Next Article in Journal
Management Strategies in Arrhythmogenic Cardiomyopathy across the Spectrum of Ventricular Involvement
Next Article in Special Issue
Natural Fatty Acid Guards against Brain Endothelial Cell Death and Microvascular Pathology following Ischemic Insult in the Presence of Acute Hyperglycemia
Previous Article in Journal
Role of Telomere Length in Survival of Patients with Idiopathic Pulmonary Fibrosis and Other Interstitial Lung Diseases
Previous Article in Special Issue
How Many Alzheimer–Perusini’s Atypical Forms Do We Still Have to Discover?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bridging Retinal and Cerebral Neurodegeneration: A Focus on Crosslinks between Alzheimer–Perusini’s Disease and Retinal Dystrophies

by
Luigi Donato
1,2,†,
Domenico Mordà
2,3,†,
Concetta Scimone
1,2,
Simona Alibrandi
1,2,*,
Rosalia D’Angelo
1 and
Antonina Sidoti
1
1
Department of Biomedical and Dental Sciences and Morphofunctional Imaging, Division of Medical Biotechnologies and Preventive Medicine, University of Messina, 98122 Messina, Italy
2
Department of Biomolecular Strategies, Genetics, Cutting-Edge Therapies, Euro-Mediterranean Institute of Science and Technology (I.E.ME.S.T.), 90139 Palermo, Italy
3
Department of Veterinary Sciences, University of Messina, 98122 Messina, Italy
*
Author to whom correspondence should be addressed.
These authors have contributed equally to this work and share first authorship.
Biomedicines 2023, 11(12), 3258; https://doi.org/10.3390/biomedicines11123258
Submission received: 1 November 2023 / Revised: 2 December 2023 / Accepted: 7 December 2023 / Published: 8 December 2023

Abstract

:
In the early stages of Alzheimer–Perusini’s disease (AD), individuals often experience vision-related issues such as color vision impairment, reduced contrast sensitivity, and visual acuity problems. As the disease progresses, there is a connection with glaucoma and age-related macular degeneration (AMD) leading to retinal cell death. The retina’s involvement suggests a link with the hippocampus, where most AD forms start. A thinning of the retinal nerve fiber layer (RNFL) due to the loss of retinal ganglion cells (RGCs) is seen as a potential AD diagnostic marker using electroretinography (ERG) and optical coherence tomography (OCT). Amyloid beta fragments (Aβ), found in the eye’s vitreous and aqueous humor, are also present in the cerebrospinal fluid (CSF) and accumulate in the retina. Aβ is known to cause tau hyperphosphorylation, leading to its buildup in various retinal layers. However, diseases like AD are now seen as mixed proteinopathies, with deposits of the prion protein (PrP) and α-synuclein found in affected brains and retinas. Glial cells, especially microglial cells, play a crucial role in these diseases, maintaining immunoproteostasis. Studies have shown similarities between retinal and brain microglia in terms of transcription factor expression and morphotypes. All these findings constitute a good start to achieving better comprehension of neurodegeneration in both the eye and the brain. New insights will be able to bring the scientific community closer to specific disease-modifying therapies.

1. Retinal Degeneration in Alzheimer’s Disease (AD)

Numerous studies have highlighted visual impairments in individuals with AD, often manifesting earlier than other dementia-related symptoms [1,2,3]. Ashok et al. [4] elucidated that these visual changes arise not solely from pathological developments in the visual cortex but also from the loss of Retinal Ganglion Cells (RGC) and degeneration associated with age-related macular degeneration (AMD) [5,6]. Intermediate hard drusen were commonly found in the temporal region of AD eyes compared to older normal eyes. A recent study has shown that, starting from postmortem brain and eye tissues samples from AD donors, there was a significant relationship between cerebral amyloid angiopathy (CAA) level and number of temporal intermediate hard drusen [7]. Other retinal manifestations in AD encompass diminished macular blood flow and astrogliosis [8]. Interestingly, certain eye diseases exhibit hallmark AD histopathological changes, indicating some overlap. For instance, amyloid-beta (Aβ) and iron accumulate in drusen [9,10,11], a defining sign of AMD, while phosphorylated tau (p-tau) and Aβ are found in RGCs, which are primarily affected in glaucoma.
However, in sporadic AD (SAD), which accounts for 85% of AD cases [12], the accumulation of Aβ and p-tau in RGCs is not consistently observed [13,14,15]. Vision-related abnormalities are prevalent in AD [16]. A decrease in the retinal nerve fiber layer (RNFL) thickness, especially in the inferior and superior quadrants [17,18,19,20], due to selective RGC loss, is now recognized as a potential AD diagnostic marker [21,22].
This marker can be non-invasively detected using optical coherence tomography (OCT) and electroretinography (ERG). However, some authors argue that these tests lack the specificity and sensitivity needed for broad clinical applications [19,23,24,25].
The processing of amyloid precursor protein (APP) in the eye, and its impairment in the brain remains a topic of debate, as does the efficiency of Aβ clearance mechanisms in both the brain and the eye. A significant correlation has been observed between Aβ and degeneration in the RGC layer, photoreceptors, and the retinal pigmented epithelium (RPE) [4,26]. Conversely, p-Tau has been identified from the ganglion cell to the outer plexiform layer [27,28]. While the presence of Aβ in drusen is established, the primary factor causing APP metabolism imbalance in RPE cells remains unidentified. Zhao et al. [29] noted that during normal aging, the production and secretion of Aβ1−42 increase in RPE cells. This leads to its deposition at the interface of RPE cells and the outer segments of photoreceptors and in the subretinal space, where it should be cleared by microglia. However, excessive Aβ expression in RPE cells results in AMD-like pathology [30], causing microglia to accumulate Aβ and other cellular debris, leading to inflammation and typical AMD drusen deposition. Rong et al.’s meta-analysis [31] supports the role of Aβ in AMD pathogenesis, highlighting a significant AD–AMD link [32].
Glaucomatous degeneration of RGCs is linked with p-tau, potentially overexpressed due to increased shear stress and other cytoskeletal proteins via the ROCK kinase pathway. AD also affects other eye regions beyond the pars nervosa. The cornea exhibits increased sensitivity [33] and decreased thickness [34,35]. Pupillary abnormalities in AD patients include slower responses to light and target detection tasks [36], exaggerated reactions to dilute tropicamide [37,38], and a reduced baseline size [39]. Additionally, during cognitive tasks, the pupil diameter appears enlarged [40]. Some AD patients develop equatorial supranuclear cataracts [41,42], but the link between lens opacity and AD is debated. While Sun et al. [43] found no optic nerve axonal damage in AD patients, other studies describe axonal degeneration [44,45] and lamina cribrosa thinning [35]. Bayer et al. and Wostyn et al. [46,47] reported elevated intraocular pressure in AD patients. Expectedly, these changes lead to various visual dysfunctions, including impaired color vision, contrast sensitivity, visual acuity, and visual integration, along with visuospatial impairments, reduced macular thickness, visual field loss, and visuomotor deficits [48].

2. Histopathological Alterations in Different Eye Cytotypes Due to β-Amyloid Production

During the progression of AD, multiple eye cell types, both neural and non-neural, exhibit distinct metabolic changes, which subsequently influence gene expression. While these changes have been documented (Figure 1), a comprehensive understanding remains elusive. It is widely believed that a chronic elevation of intraocular pressure (IOP) may instigate the accumulation of not only p-tau in RGCs, but also Aβ, leading to cell death. The processing of APP appears to be consistent in the retina and other eye cell types. This is further supported by the presence of soluble APPα, APPβ, and pathogenic Aβ in the vitreous humor (VH) and aqueous humor (AH), as well as in cerebrospinal fluid (CSF) [49]. Aβ deposits span across all retina layers, from the nerve fiber layer to the photoreceptor layer. Analogous to the brain, retinal Aβ fosters tau protein hyperphosphorylation, leading to its aggregation in structures like neurofibrillary tangles (NFT) [15,50,51]. Elevated IOP has been linked to increased tau oligomerization, resulting in RGC death. The suppression of tau using short interfering RNA (shRNA) has been shown to rescue RGCs, further emphasizing the relationship between p-tau accumulation and cell death [52].
Tau’s role in stabilizing microtubules means that its hyperphosphorylation and aggregation can disrupt anterograde axonal transport, affecting several neuronal functions. For instance, it can hinder mitochondrial transport, leading to energy depletion and the generation of reactive oxygen species (ROS) [53]. In the cornea, both epithelial cells and fibroblasts seem to be involved. Fibroblasts show increased expression of ADAM-10 (α-secretase, the Amyloid Precursor Protein processing enzyme which starts non-amyloidogenic pathway cleaving at residue 697) and BACE-1 (β-secretase, the Amyloid Precursor Protein processing enzyme which starts amyloidogenic pathway cleaving at residue 11) [54], while epithelial cells exhibit heightened APP expression, Aβ accumulation, morphological changes, and an increased rate of apoptosis [34]. Both cell types express APP and produce Aβ [55].
In the retina, notable histopathological markers include a shift towards anaerobic metabolism [56,57], the presence of Aβ plaques [14,58], p-tau deposits [13,52], vascular changes [59,60], and blood–retinal barrier disruptions [61]. Aβ also accumulates in the retinal microvasculature and pericytes [61]. Significant gene expression changes have been observed, such as increased levels of retinal vascular β40 and β42 Aβ fragments and decreased levels of vascular PDGFR-β and LDL-1 [61]. The lens’s involvement in AD is debated; while some researchers have observed increased Aβ aggregation [42,62] and presenilin expression [63], others have not [54,64]. The aqueous humor has shown elevated Aβ levels [65,66], whereas the vitreous humor is rich in AD-associated proteins [49,66]. The choroid appears thinner [67,68], and tau deposition has been observed in the optic nerve [69].

3. Retinal Histopathological Abnormalities in AD Mouse Models

Various AD animal models have exhibited retinal Aβ deposits, often accompanied by apoptotic RGCs and axonal degeneration [70,71]. For example, in 3xTG-AD, APP-PS1ΔE9, and APPswe/PS1ΔE9 mouse models, which are known to develop Aβ deposits in the brain, there have been observations of retinal Aβ oligomers and thinning of the retinal nerve fiber layer (RNFL) [6,72,73]. In a glaucoma rat model, the retinal threshold for Aβ1−42 increased in response to elevated intraocular pressure (IOP), aging, and light exposure, leading to RGC apoptosis [52]. Walsh et al. [74] replicated this condition by administering intravitreal injections of Aβ1−42. Their research demonstrated that using agents that either reduce toxic Aβ fragments or induce anti-amyloidogenic mutations can protect RGCs. Lastly, in tauopathy mouse models (where mutations are introduced in MAPT, the tau gene), a direct relationship was established between p-tau, Aβ deposits, and RGC death, highlighting the three primary contributors to the retinal pathology associated with AD [75].

4. The Role of the Prion Protein in Retinal Allostasis

In the progression of Alzheimer-Perusini’s disease, the prion protein (PrP) plays a multifaceted role. As described by H. H. Jarosz-Griffiths et al. [76], its function can be likened to the “Ugly” character in Ennio Morricone’s renowned film. PrP can act as a conduit for the cytotoxic effects of Amyloid-beta (Aβ), serving as a scavenger receptor. Conversely, it can also block Aβ-derived fragments, preventing their aggregation in the extracellular space, thus acting as a decoy [12]. Currently, two Aβ binding sites on the PrPC structure have been identified: one spanning amino acid residues 95–105 and the other 23–27 [77]. However, only blocking the larger site (residues 95–105) protects neurons from Aβ’s harmful effects. When Aβ oligomers bind to PrP, the downstream tyrosine kinase Fyn pathway is activated, leading to increasingly severe synaptic alterations until their eventual destruction [78].
Like APP, PrPC undergoes various cleavage patterns, termed α-, β-, or γ-cleavage. α-cleavage is facilitated by ADAM17 [79], which belongs to the α-secretase family (also known as TNFα-cleaving enzyme, TACE). This cleavage results in the release of the N-terminal soluble fragment N1 into the extracellular space, while the C-terminal fragment remains attached as C1 via a GPI anchor. Given that C1 lacks the primary Aβ-binding site, this post-translational modification is believed to protect against Aβ-induced toxicity. Conversely, β-cleavage produces a membrane-anchored C-terminal fragment, C2, which contains the entire primary Aβ-binding region, and a released N-terminal fragment N2, which contains the secondary Aβ-binding site spanning residues 23−27. Thus, C2 fragments may facilitate Aβ neurotoxic transmission. γ-cleavage results in a soluble form of PrPC. In neurons, the majority (65–80%) of PrPC cleavage occurs at position 111/112, releasing the N1 fragment (α-cleavage). However, in the retina, β-cleavage predominates, primarily due to oxidative stress. This cleavage pattern is believed to protect cells from oxidative damage. The same trend is observed in other eye cell types. This distinction between the eye, especially the retina, and the brain might be attributed to the eye’s constant exposure to light radiation. Another potential factor is the downregulation of ADAM17, the enzyme initiating Aβ’s non-amyloidogenic processing [80]. PrPC itself is believed to regulate ADAM17/TACE levels, indicating a mutual and intricate control mechanism [81,82,83,84,85]. This proteolytic molecule is activated in immune cells, where TNF-α initiates several pathways crucial for their activity [86]. Additionally, PrPC stabilizes the extracellular matrix by interacting with β1 integrin [87,88]. When PrPC expression is reduced, the RhoA-associated coiled-coil-containing kinase (ROCK) pathway is activated. Overactivation of ROCK, through the LIMK-cofilin pathway [89,90], results in a shift in the cytoskeleton status. These changes lead to increased resistance to aqueous outflow, elevated IOP, and ultimately, RGC loss.
In the brain, PrPC acts as a radical scavenger. PrPC-knockdown mice exhibit increased susceptibility to intracellular ROS [91]. The N-terminal octapeptide repeat of PrPC coordinates redox-active metals (e.g., Cu, Zn, and Fe) [92,93,94,95], enabling it to act as both a scavenger and a metal transporter [81,96,97]. Given these findings, further research is needed to fully understand PrP’s multifaceted role in the eye [4]. Striebel et al. [32] and others have shown that prion-induced degeneration of photoreceptor cells in mice and humans resembles the pathology of human retinitis pigmentosa caused by retinal protein gene mutations. The Scrapie isoform of the prion protein (PrPSc) [98] has been found to be associated with the base of cilia and swollen cone inner segments. These findings suggest that ciliopathy might be the underlying pathogenic mechanism. PrPSc has also been detected on the dendrites of cone and rod bipolar cells, extending into ribbon synapses. It is plausible that a similar mechanism might be activated in prion-like diseases, such as AD, as well as in human retinitis pigmentosa (RP) [32].
Not only AD but also other prion and “prion-like” neurodegenerative progressions lead to retinal pathological changes in humans and other species. While there may be some overlap in histopathological abnormalities, each disease seems to have a unique impact on the retina [99]. For instance, in patients with Parkinson’s disease (PD), abnormal α-synuclein oligomers [100] have been observed in the retina, particularly in the nerve fiber (ganglion cell axons), inner plexiform, and ganglion cell body layers [99].

5. Different Microglial Phenotypes and Aβ Clearance

Microglia constitute 0.2% of the total retinal cells and 5–10% of the entire CNS. These cells share lineage with monocytes, macrophages, and dendritic cells. They originate from the yolk sac, which is the embryonic location for human hematopoiesis [101], and develop from primitive progenitors [102,103]. Their migration to the CNS occurs post-differentiation [104]. They have been identified in development from embryonic day (E) 8.5–9.5. In mice with compromised circulation (due to sodium-calcium exchanger1, Ncx-1 deficiency), microglia do not reach the brain, suggesting this as their primary migration route [105]. While microglial presence in the human retina is noted by the 10th week of gestation and in mice by E11.5, it is plausible that they enter the brain even earlier [106,107]. This timing is consistent in rats at E12 [108] and in quails at E7 [109].
It is hypothesized that retinal microglial infiltration happens in two sequential migratory waves, based on the spatiotemporal positioning of these cells, both pre- and post-vascularization. Initially, microglia might enter the retina by traversing the vitreal retina surface or migrating from non-neural ciliary regions in the periphery [106,107,110,111]. Later, the invasion likely begins from the optic disc or through blood vessels [112]. Regardless, the arrival of microglia aligns with the differentiation of retinal neurons from retinal progenitor cells (RPCs). A study by Li et al. [104] reported that the outer plexiform layer (OPL) houses 47% of the microglial population, with the remainder primarily in the inner plexiform layer (IPL). The authors speculate that this distribution might mirror the spatiotemporal distribution of synapses in the developing retina. Supporting this theory, synapse formation starts in the nascent IPL around E17, coinciding with the presence of 99% of microglial cells [113,114]. By postnatal day (P)3, approximately 80% remain localized between the developing IPL and the ganglion cell layer (GCL). By P9, microglia are observed in the developing OPL. In adulthood, this pattern persists, with microglia and their processes predominantly found in the inner retina and OPL. In contrast, the outer nuclear layer (ONL) largely lacks them [107,115]. This might be because phagocytosis and general allostasis in this region are overseen by the retinal pigmented epithelium (RPE).
Microglia exhibit a diverse range of phenotypes in response to changes in the CNS microenvironments, including those in the retina. They play a pivotal role in monitoring allostasis within these environments. Historically, two primary microglial phenotypes have been identified. From a dichotomous perspective, microglial activation can be categorized into M1 or M2 phenotypes, similar to macrophages [99]. The M1 phenotype is associated with neuroinflammatory responses, while the M2 phenotype is linked to anti-inflammatory, neuroprotective, and restorative functions. Recent studies have identified at least three subpopulations of the M2 phenotype, labeled M2a-c [116]. Specifically, M2a is involved in the production of IL-10 and IGF-1, facilitating cell debris clearance and neuroprotection [99,116,117]. M2b cells, on the other hand, appear to be activated by inflammatory agents such as IL-1β and LPS and respond by upregulating IL-10 expression [116]. They also exhibit phagocytic activity in AD model brains and have increased levels of CD64, a typical M1 Fc receptor [118]. M2c cells produce significant amounts of TGFβ but are inhibited by IL-10 and glucocorticoids (as illustrated in Figure 2). However, the spectrum of microglial phenotypes is vast. For a more comprehensive overview of microglial phenotypes observed in AD brains, such as damage-associated microglia (DAM), readers are referred to Donato et al. [12].
Our current focus is on specific morphotypes, which are molecular shifts resulting in evident morphological changes. A recent theory has identified two primary microglial states: ramified and amoeboid, with three transitional forms: hyper-ramified, activated, and rod morphotypes [119,120]. Ramified microglia, representing the inactive state, have small, round somas with thin, elongated processes that continuously extend and retract, facilitating their monitoring function [104,119,121]. These cells are distributed in a mosaic pattern [122]. Upon detecting specific microenvironmental stimuli, they transition to the “hyper-ramified” state, characterized by denser, longer, and thicker dendrites and a larger, irregularly shaped soma [121,123]. Persistent harmful stimuli can cause these primed microglia to adopt the “amoeboid” state, which is marked by even larger, more uniformly shaped cell bodies with minimal to no extensions [121,123]. This state is inherently phagocytic, enabling the cells to migrate to inflammation sites and engulf dying neurons and cellular debris [119,124]. “Rod” microglia resemble sausage-shaped cells with few dendrites, sometimes shorter than those of ramified microglia [125,126]. While their exact functions remain elusive, evidence suggests that rod microglia position themselves near neurons, aligning with nerve fibers [125,126,127,128]. Studies have shown that ramified microglia express high levels of P2RY12, associated with surveillance functions, whereas amoeboid microglia predominantly express the phagocytosis marker CD68 [119,129].
Recognizing variations in microglial morphology, especially through retinal neuroimaging, could pave the way for early diagnosis of AD or other neurodegenerative conditions. Fernandez-Arjona et al. [130] defined 15 parameters to facilitate an objective morphometric analysis using automated software (Fraclac, v.2.5; ImageJ v.1.53). During brain development, the amoeboid form of microglia is predominant, especially when synapse remodeling is at its peak. Following this period, within the first two postnatal weeks, they transition to the ramified form [131,132]. A similar transition is observed in the retina. At birth, these immune cells in the retina exhibit an amoeboid structure, extending their dendrites towards the retinal basal side. As the retina matures, they evolve into the ramified morphotype [107]. However, in the event of CNS injury or infection, microglia revert to the amoeboid state [133,134]. Koso et al. highlighted the significance of the zinc finger transcription factor Sall1 in this transition, noting its specific expression in amoeboid retinal microglia. A knockout of this factor can induce a shift from the ramified to the amoeboid form [135]. Both brain and retinal microglia express several common transcription factors [104]. While numerous studies [104,119,127,136,137,138] have identified each morphotype in both the retina and the broader brain, it remains unclear if these morphotypes exhibit identical characteristics across both regions [104]. O’Koren et al. [139] suggested that the cytokine IL-34 might delineate the spatial distribution of distinct microglial subsets in the retina. In healthy conditions, most IL-34- microglial cells are found in the OPL, while IL-34+ cells are predominantly in the IPL. As neurodegeneration progresses, both subsets migrate to the RPE. Interestingly, microglia near damaged retinal neurons show elevated levels of CD11c [140]. In the future, we might harness the potential to reprogram microglial phenotypes to address early pathological developments. Distinct microglial shapes have been observed across various CNS regions, such as the striatum, frontal cortex, and hippocampus, where the soma is larger, and processes are more abundant. Additionally, as one transitions through cerebral cortex layers, the dimensions of microglial cells change [141]. Typically, aging microglia shed their dendrites, resembling a leafless tree, adopting a “dystrophic” appearance [142]. This transformation complicates myelin turnover, further exacerbated by alterations in the myelin itself. Data regarding this phenomenon in the retina is limited, possibly due to its relative lack of this lipid sheath [143].

6. Brain–Retina Microglia Axes and Alzheimer–Perusini’s Continuum

Numerous studies [125,136,144,145,146,147,148] have highlighted the association between microglial activation and AD pathology, not only in the brain but also in the retina. The age-induced shift in the epigenetic profile of microglia might influence the onset of specific pathologies, such as AMD or even AMD-like retinal degeneration in AD patients. For instance, Ma et al. [149] observed an age-related upregulation of the AMD-associated gene, C3. Several research groups noted primed microglia in the retinas of 5-week-old 3× transgenic AD mice, suggesting an early microglial response in AD, potentially preceding cerebral involvement [146].
Rod microglia, in particular, warrant special attention due to their potential region-specific involvement in AD [125,136]. In a study by Bachstetter et al. [136], rod microglia were exclusively identified in the parietal cortex of AD patient autopsies, with none found in the hippocampus or temporal cortex when compared to healthy controls. This observation aligns with another study where no significant presence of this microglial morphotype was noted in the hippocampal and cortical brain regions of patients with a history of traumatic brain injury [136]. Furthermore, a correlation between rod microglia and patients with mutated C9orf72 was observed, given its prevalence in this patient group and relative absence in age-matched controls. Interestingly, a higher density of rod microglia in the grey matter of Down’s syndrome AD patients (with three copies of the APP gene) was linked to a more severe pathological progression compared to those with only AD and no chromosomal aberration [125]. Grimaldi et al. [150] examined postmortem AD retinas and identified classic AD histopathological markers, such as Aß plaques and neurofibrillary tangles. Additionally, they reported an increased expression of caspase 3, an AD-associated neurodegeneration marker, and a higher density of iba-1 positive microglia, which also exhibited elevated levels of IL-1ß. However, these microglia maintained normal TREM2 levels, even though AD retinas showed higher TREM2 mRNA levels compared to healthy patients. If these observations are validated by further research, they could serve as effective early diagnostic markers. Understanding the underlying molecular mechanisms could not only facilitate early diagnosis but also pave the way for innovative therapeutic approaches. Moreover, new insights into retinal DAM marker expression, combined with enhanced morphological parameter assessments, could offer a novel diagnostic approach for AD retinopathies [143].
In conclusion, another noteworthy microglial phenotype is the perivascular microglia (PM). PM microglia appear to oversee exchanges across the blood–brain barrier (BBB), monitoring compounds entering the CNS from the bloodstream [151]. A study found that drug-induced depletion of PM microglia in AD mice led to the formation of corticovascular amyloid plaques [152]. Figure 3 summarizes the discussed microglial morphotypes.

7. The Retina as a Gateway to Early AD Diagnosis

As previously mentioned, AD patients often exhibit visual symptoms related to retinal morphological and functional changes before the manifestation of typical AD symptoms [153]. Many of these changes could serve as early indicators of the disease. For instance, the loss of amacrine and ganglion cells, leading to dysfunction in the RNFL, is frequently reported in the literature and is associated with Aβ oligomerization throughout the retina [75,154]. However, the pattern of Aβ deposition in the retina differs from that in the brain. As discussed in Chiquita et al. [155], advanced studies on neural networks across multiple space–time dimensions [156,157] may provide insights into the parallels between the progression of the disease in the retina and the brain. Additionally, the AD retina exhibits changes related to the immune system (e.g., astrocytes and microglia), vasculature, and electroretinogram (ERG) responses [158]. Events such as cell death and the disrupted allostasis of transition metals (e.g., iron, copper, and zinc) [159] can induce astrocyte and microglia activation [146,160]. This is further complicated by the co-localization with Tau oligomers [161,162] and Aβ plaques. Such a disrupted microenvironment might lead to rhodopsin instability, further exacerbating cell death events [159]. Movio et al. [163] provided a concise summary of the current limitations in analyzing AD in the retina. The data describing the simultaneous structural changes in the retina and AD brain tissue, as reported in the literature, seem inconsistent [155,164]. However, obtaining accurate electrophysiological tracings of RGC in AD remains a challenge, even though it could be pivotal for early diagnosis. In the subsequent sections, we will delve into recent efforts to overcome these challenges.

8. Retinal Organoids (RO) and Other Culture Systems

To date, our understanding of neurodegenerative disease-related retinal changes primarily stems from imaging analyses of patients, post-mortem tissue studies, and animal experimental models [165,166,167]. However, a more personalized approach is achievable through the use of induced pluripotent stem cells (hiPSC) and organoid technologies. These can be derived from patients who may be at risk of future neurodegenerative progression, identified either due to genetic predispositions or prolonged exposure to known pollutants, long before symptom onset [163]. Both sporadic AD (SAD) and familial AD (FAD) derived hiPSC cells exhibit similar brain tissue characteristics in vitro [168], suggesting similar characteristics in ROs [169,170]. Key indicators to monitor in these models include rhodopsin stability, neuroinflammatory responses, microvascular changes, and oxidative stress [171]. Given the importance of microglial cells in AD, as discussed previously, a reliable RO model should incorporate these cells, as described in Movio et al. [163]. Considering the significant alterations in AD-associated microglial cells, such as cytokine secretion changes and impaired phagocytosis [172], the use of hiPSC-derived microglial-like cells might be beneficial [173]. Park (2018) proposed a microfluidic system-based approach for an AD 3D model, co-culturing human neurons, astrocytes, and microglia, which effectively maintained astrocyte and microglia activation [159]. Given the significant cell loss observed in the inner retina, a comprehensive 3D organoid model might not be essential for early diagnosis or addressing specific queries [165]. A simpler 2D culture of ganglion and amacrine cells might suffice. However, the inclusion of astrocytes and microglia, combined with microfluidic devices, could enhance the model’s relevance [174]. Utilizing CRISPR/Cas genome editing in hiPSC-derived diseased cells offers the potential to deeply explore the genetic contributions, illuminating the underlying molecular mechanisms [163]. Lastly, when developing neurodegenerative disease models, the intrinsic association with aging should be acknowledged. Strategies such as progerin overexpression, anti-aging gene knockouts, and the induction of reactive oxygen species (ROS) production can make these models more reflective of the actual disease state [175,176].

9. Rhodopsin Quantification

Rhodopsin, a combination of opsin and 11-cis-retinal, is a G-protein coupled receptor found in the rod cells of the retina, playing a crucial role in light signal transduction. It serves as a marker for retinal thinning [177,178], which is increasingly recognized as an early indicator of neurodegenerative diseases. Notably, rhodopsin instability and depletion appear to precede the loss of retinal neural cells [159,179]. A study by Ni et al. in 2017 highlighted the role of drosophila rhodopsin receptors as circadian pacemakers in neural cells, suggesting a connection between circadian rhythm disruptions and the etiopathogenesis of neurodegenerative diseases [180], with the discovery of the glymphatic system being particularly noteworthy in this context [181].
Stojanovic et al. proposed a theory in 2004, suggesting that rhodopsin misfolding could be the underlying cause of retinal degeneration observed in conditions like AD and RP [182]. They posited that specific mutations in a high-affinity transmembrane site of rhodopsin, which coordinates with zinc, trigger this misfolding. Interestingly, low zinc levels have been linked to retinal neurodegeneration and night blindness, a result of rod cell death [183]. Elevated zinc concentrations, on the other hand, diminish rhodopsin’s affinity for 11-cis-retinal [184]. The dysregulation of transition metals is closely associated with the onset of neurodegenerative diseases [185].
Recent clinical studies suggest the potential of optical coherent tomography (OCT) as a screening tool for early signs of neurodegenerative progression, given its ability to detect volumetric changes in the retina [186]. As rhodopsin instability is believed to precede cell loss, its quantification could further advance early diagnosis. Liu et al. in 2015 introduced a rhodopsin imaging technique known as nanosecond pulsed scanning laser ophthalmoscopy (SLO), later improved to eSLO (confocal laser ophthalmoscope). This technique offers high-resolution mapping of rhodopsin and pigment epithelium distribution by analyzing lipofuscin autofluorescence in approximately 50 μm resolution pixel clusters. eSLO is designed for compatibility with commonly used clinical imaging instruments [187]. Combining rhodopsin quantification with retinal thickness assessment could pave the way for early diagnosis of brain neurodegeneration [159]. Considering that rhodopsin kinetics slow down with age [188], it is essential to differentiate between pathological and natural reductions in this pigment and establish appropriate benchmarks. While rhodopsin levels are most closely associated with Alzheimer’s disease, they cannot be used for making a differential diagnosis with other neurodegenerative disorders, e.g., Parkinson’s disease. However, when combined with other techniques or early clinical signs, rhodopsin quantification holds significant potential for early diagnosis.

10. Conclusions and Future Scenarios

This review offers a comprehensive look at various promising discoveries and theories, some of which are still distant from practical clinical application. To bring these findings to fruition, several challenges must be addressed:
  • Enhancing the sensitivity and specificity of certain techniques (as ones mentioned in the previous paragraphs);
  • Establishing definitive cut-offs and thresholds (which indicate when a certain accumulation of non-functioning rhodopsin or when a specific modification of the molecule is unequivocally indicative of an emerging neurodegenerative progression);
  • Developing a comprehensive diagnostic framework that encompasses various stages and provides refined guidelines for individual evaluations;
  • Ensuring global accessibility to advanced equipment and technologies.
The age-old adage suggests that the eyes are windows to the soul. Given their transparency, could the eyes serve as a lens through which early signs of neurodegeneration are detected? Based on our discussions, the potential is promising. However, a deeper understanding of retinal pathophysiology, molecular mechanisms, and gene expression is essential. In many neurodegenerative diseases, specific organs or systems show early signs of affliction. For example, the onset of Parkinson’s disease (PD) is often linked to the olfactory nerve. Such patterns, although unique to each individual, can be grouped into discernible clusters. In the future, advanced bioinformatics tools powered by artificial intelligence (AI) could map these associations on a grand scale, correlating gene expression with structural data. This would bring the scientific community a step closer to ensuring early diagnoses for countless patients globally. Presently, by integrating cutting-edge eye imaging techniques with valuable genetic data, we can potentially screen individuals at medium to high risk of developing dementia. Achieving this goal requires a collaborative effort among neurologists, ophthalmologists, bioinformaticians, and biotechnologists, all working together to elevate scientific knowledge.

Author Contributions

All authors (L.D., D.M., C.S., S.A., R.D. and A.S.) reviewed the literature. L.D. and D.M. wrote the initial draft and drew the figures. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Moons, L.; De Groef, L. Multimodal retinal imaging to detect and understand Alzheimer’s and Parkinson’s disease. Curr. Opin. Neurobiol. 2022, 72, 1–7. [Google Scholar] [CrossRef] [PubMed]
  2. Shi, H.; Koronyo, Y.; Rentsendorj, A.; Fuchs, D.T.; Sheyn, J.; Black, K.L.; Mirzaei, N.; Koronyo-Hamaoui, M. Retinal Vasculopathy in Alzheimer’s Disease. Front. Neurosci. 2021, 15, 731614. [Google Scholar] [CrossRef] [PubMed]
  3. Yoon, S.P.; Grewal, D.S.; Thompson, A.C.; Polascik, B.W.; Dunn, C.; Burke, J.R.; Fekrat, S. Retinal Microvascular and Neurodegenerative Changes in Alzheimer’s Disease and Mild Cognitive Impairment Compared with Control Participants. Ophthalmol. Retin. 2019, 3, 489–499. [Google Scholar] [CrossRef]
  4. Ashok, A.; Singh, N.; Chaudhary, S.; Bellamkonda, V.; Kritikos, A.E.; Wise, A.S.; Rana, N.; McDonald, D.; Ayyagari, R. Retinal Degeneration and Alzheimer’s Disease: An Evolving Link. Int. J. Mol. Sci. 2020, 21, 7290. [Google Scholar] [CrossRef] [PubMed]
  5. Asanad, S.; Ross-Cisneros, F.N.; Nassisi, M.; Barron, E.; Karanjia, R.; Sadun, A.A. The Retina in Alzheimer’s Disease: Histomorphometric Analysis of an Ophthalmologic Biomarker. Investig. Ophthalmol. Vis. Sci. 2019, 60, 1491–1500. [Google Scholar] [CrossRef]
  6. Gupta, V.; Gupta, V.B.; Chitranshi, N.; Gangoda, S.; Vander Wall, R.; Abbasi, M.; Golzan, M.; Dheer, Y.; Shah, T.; Avolio, A.; et al. One protein, multiple pathologies: Multifaceted involvement of amyloid beta in neurodegenerative disorders of the brain and retina. Cell Mol. Life Sci. 2016, 73, 4279–4297. [Google Scholar] [CrossRef]
  7. Ukalovic, K.; Cao, S.; Lee, S.; Tang, Q.; Beg, M.F.; Sarunic, M.V.; Hsiung, G.R.; Mackenzie, I.R.; Hirsch-Reinshagen, V.; Cui, J.Z.; et al. Drusen in the Peripheral Retina of the Alzheimer’s Eye. Curr. Alzheimer Res. 2018, 15, 743–750. [Google Scholar] [CrossRef]
  8. Blanks, J.C.; Schmidt, S.Y.; Torigoe, Y.; Porrello, K.V.; Hinton, D.R.; Blanks, R.H. Retinal pathology in Alzheimer’s disease. II. Regional neuron loss and glial changes in GCL. Neurobiol. Aging 1996, 17, 385–395. [Google Scholar] [CrossRef]
  9. Guo, L.Y.; Alekseev, O.; Li, Y.; Song, Y.; Dunaief, J.L. Iron increases APP translation and amyloid-beta production in the retina. Exp. Eye Res. 2014, 129, 31–37. [Google Scholar] [CrossRef]
  10. Ohno-Matsui, K. Parallel findings in age-related macular degeneration and Alzheimer’s disease. Prog. Retin. Eye Res. 2011, 30, 217–238. [Google Scholar] [CrossRef] [PubMed]
  11. Ong, S.S.; Proia, A.D.; Whitson, H.E.; Farsiu, S.; Doraiswamy, P.M.; Lad, E.M. Ocular amyloid imaging at the crossroad of Alzheimer’s disease and age-related macular degeneration: Implications for diagnosis and therapy. J. Neurol. 2019, 266, 1566–1577. [Google Scholar] [CrossRef] [PubMed]
  12. Donato, L.; Morda, D.; Scimone, C.; Alibrandi, S.; D’Angelo, R.; Sidoti, A. How Many Alzheimer-Perusini’s Atypical Forms Do We Still Have to Discover? Biomedicines 2023, 11, 2035. [Google Scholar] [CrossRef] [PubMed]
  13. den Haan, J.; Morrema, T.H.J.; Verbraak, F.D.; de Boer, J.F.; Scheltens, P.; Rozemuller, A.J.; Bergen, A.A.B.; Bouwman, F.H.; Hoozemans, J.J. Amyloid-beta and phosphorylated tau in post-mortem Alzheimer’s disease retinas. Acta Neuropathol. Commun. 2018, 6, 147. [Google Scholar] [CrossRef]
  14. Koronyo, Y.; Biggs, D.; Barron, E.; Boyer, D.S.; Pearlman, J.A.; Au, W.J.; Kile, S.J.; Blanco, A.; Fuchs, D.T.; Ashfaq, A.; et al. Retinal amyloid pathology and proof-of-concept imaging trial in Alzheimer’s disease. JCI Insight 2017, 2, e93621. [Google Scholar] [CrossRef] [PubMed]
  15. Shah, T.M.; Gupta, S.M.; Chatterjee, P.; Campbell, M.; Martins, R.N. Beta-amyloid sequelae in the eye: A critical review on its diagnostic significance and clinical relevance in Alzheimer’s disease. Mol. Psychiatry 2017, 22, 353–363. [Google Scholar] [CrossRef]
  16. Cerquera-Jaramillo, M.A.; Nava-Mesa, M.O.; Gonzalez-Reyes, R.E.; Tellez-Conti, C.; de-la-Torre, A. Visual Features in Alzheimer’s Disease: From Basic Mechanisms to Clinical Overview. Neural Plast. 2018, 2018, 2941783. [Google Scholar] [CrossRef]
  17. Berisha, F.; Feke, G.T.; Trempe, C.L.; McMeel, J.W.; Schepens, C.L. Retinal abnormalities in early Alzheimer’s disease. Investig. Ophthalmol. Vis. Sci. 2007, 48, 2285–2289. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, D.; Zhang, L.; Li, Z.; Zhang, X.; Wu, Y.; Yang, H.; Min, B.; Zhang, X.; Ma, D.; Lu, Y. Thinner changes of the retinal nerve fiber layer in patients with mild cognitive impairment and Alzheimer’s disease. BMC Neurol. 2015, 15, 14. [Google Scholar] [CrossRef] [PubMed]
  19. Marziani, E.; Pomati, S.; Ramolfo, P.; Cigada, M.; Giani, A.; Mariani, C.; Staurenghi, G. Evaluation of retinal nerve fiber layer and ganglion cell layer thickness in Alzheimer’s disease using spectral-domain optical coherence tomography. Investig. Ophthalmol. Vis. Sci. 2013, 54, 5953–5958. [Google Scholar] [CrossRef]
  20. Polo, V.; Rodrigo, M.J.; Garcia-Martin, E.; Otin, S.; Larrosa, J.M.; Fuertes, M.I.; Bambo, M.P.; Pablo, L.E.; Satue, M. Visual dysfunction and its correlation with retinal changes in patients with Alzheimer’s disease. Eye 2017, 31, 1034–1041. [Google Scholar] [CrossRef]
  21. Jones-Odeh, E.; Hammond, C.J. How strong is the relationship between glaucoma, the retinal nerve fibre layer, and neurodegenerative diseases such as Alzheimer’s disease and multiple sclerosis? Eye 2015, 29, 1270–1284. [Google Scholar] [CrossRef]
  22. Sivak, J.M. The aging eye: Common degenerative mechanisms between the Alzheimer’s brain and retinal disease. Investig. Ophthalmol. Vis. Sci. 2013, 54, 871–880. [Google Scholar] [CrossRef] [PubMed]
  23. Liao, H.; Zhu, Z.; Peng, Y. Potential Utility of Retinal Imaging for Alzheimer’s Disease: A Review. Front. Aging Neurosci. 2018, 10, 188. [Google Scholar] [CrossRef] [PubMed]
  24. Ngolab, J.; Honma, P.; Rissman, R.A. Reflections on the Utility of the Retina as a Biomarker for Alzheimer’s Disease: A Literature Review. Neurol. Ther. 2019, 8, 57–72. [Google Scholar] [CrossRef] [PubMed]
  25. Tsokolas, G.; Tsaousis, K.T.; Diakonis, V.F.; Matsou, A.; Tyradellis, S. Optical Coherence Tomography Angiography in Neurodegenerative Diseases: A Review. Eye Brain 2020, 12, 73–87. [Google Scholar] [CrossRef]
  26. Ratnayaka, J.A.; Serpell, L.C.; Lotery, A.J. Dementia of the eye: The role of amyloid beta in retinal degeneration. Eye 2015, 29, 1013–1026. [Google Scholar] [CrossRef]
  27. Liu, B.; Rasool, S.; Yang, Z.; Glabe, C.G.; Schreiber, S.S.; Ge, J.; Tan, Z. Amyloid-peptide vaccinations reduce beta-amyloid plaques but exacerbate vascular deposition and inflammation in the retina of Alzheimer’s transgenic mice. Am. J. Pathol. 2009, 175, 2099–2110. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, C.; Zhang, C.W.; Zhou, Y.; Wong, W.Q.; Lee, L.C.; Ong, W.Y.; Yoon, S.O.; Hong, W.; Fu, X.Y.; Soong, T.W.; et al. APP upregulation contributes to retinal ganglion cell degeneration via JNK3. Cell Death Differ. 2018, 25, 663–678. [Google Scholar] [CrossRef]
  29. Zhao, Y.; Bhattacharjee, S.; Jones, B.M.; Hill, J.M.; Clement, C.; Sambamurti, K.; Dua, P.; Lukiw, W.J. Beta-Amyloid Precursor Protein (betaAPP) Processing in Alzheimer’s Disease (AD) and Age-Related Macular Degeneration (AMD). Mol. Neurobiol. 2015, 52, 533–544. [Google Scholar] [CrossRef]
  30. Prasad, T.; Zhu, P.; Verma, A.; Chakrabarty, P.; Rosario, A.M.; Golde, T.E.; Li, Q. Amyloid beta peptides overexpression in retinal pigment epithelial cells via AAV-mediated gene transfer mimics AMD-like pathology in mice. Sci. Rep. 2017, 7, 3222. [Google Scholar] [CrossRef]
  31. Rong, S.S.; Lee, B.Y.; Kuk, A.K.; Yu, X.T.; Li, S.S.; Li, J.; Guo, Y.; Yin, Y.; Osterbur, D.L.; Yam, J.C.S.; et al. Comorbidity of dementia and age-related macular degeneration calls for clinical awareness: A meta-analysis. Br. J. Ophthalmol. 2019, 103, 1777–1783. [Google Scholar] [CrossRef] [PubMed]
  32. Striebel, J.F.; Race, B.; Leung, J.M.; Schwartz, C.; Chesebro, B. Prion-induced photoreceptor degeneration begins with misfolded prion protein accumulation in cones at two distinct sites: Cilia and ribbon synapses. Acta Neuropathol. Commun. 2021, 9, 17. [Google Scholar] [CrossRef] [PubMed]
  33. Ornek, N.; Dag, E.; Ornek, K. Corneal sensitivity and tear function in neurodegenerative diseases. Curr. Eye Res. 2015, 40, 423–428. [Google Scholar] [CrossRef] [PubMed]
  34. Dong, Z.; Luo, A.; Gan, Y.; Li, J. Amyloid Beta Deposition Could Cause Corneal Epithelial Cell Degeneration Associated with Increasing Apoptosis in APPswePS1 Transgenic Mice. Curr. Eye Res. 2018, 43, 1326–1333. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, E.J.; Kim, T.W.; Lee, D.S.; Kim, H.; Park, Y.H.; Kim, J.; Lee, J.W.; Kim, S. Increased CSF tau level is correlated with decreased lamina cribrosa thickness. Alzheimers Res. Ther. 2016, 8, 6. [Google Scholar] [CrossRef]
  36. Dragan, M.C.; Leonard, T.K.; Lozano, A.M.; McAndrews, M.P.; Ng, K.; Ryan, J.D.; Tang-Wai, D.F.; Wynn, J.S.; Hoffman, K.L. Pupillary responses and memory-guided visual search reveal age-related and Alzheimer’s-related memory decline. Behav. Brain Res. 2017, 322, 351–361. [Google Scholar] [CrossRef]
  37. Iijima, A.; Haida, M.; Ishikawa, N.; Ueno, A.; Minamitani, H.; Shinohara, Y. Re-evaluation of tropicamide in the pupillary response test for Alzheimer’s disease. Neurobiol. Aging 2003, 24, 789–796. [Google Scholar] [CrossRef] [PubMed]
  38. Scinto, L.F. ApoE allelic variability influences pupil response to cholinergic challenge and cognitive impairment. Genes. Brain Behav. 2007, 6, 209–215. [Google Scholar] [CrossRef]
  39. Kawasaki, A.; Ouanes, S.; Crippa, S.V.; Popp, J. Early-Stage Alzheimer’s Disease Does Not Alter Pupil Responses to Colored Light Stimuli. J. Alzheimers Dis. 2020, 75, 1273–1282. [Google Scholar] [CrossRef]
  40. Kremen, W.S.; Panizzon, M.S.; Elman, J.A.; Granholm, E.L.; Andreassen, O.A.; Dale, A.M.; Gillespie, N.A.; Gustavson, D.E.; Logue, M.W.; Lyons, M.J.; et al. Pupillary dilation responses as a midlife indicator of risk for Alzheimer’s disease: Association with Alzheimer’s disease polygenic risk. Neurobiol. Aging 2019, 83, 114–121. [Google Scholar] [CrossRef]
  41. Kerbage, C.; Sadowsky, C.H.; Jennings, D.; Cagle, G.D.; Hartung, P.D. Alzheimer’s disease diagnosis by detecting exogenous fluorescent signal of ligand bound to Beta amyloid in the lens of human eye: An exploratory study. Front. Neurol. 2013, 4, 62. [Google Scholar] [CrossRef] [PubMed]
  42. Melov, S.; Wolf, N.; Strozyk, D.; Doctrow, S.R.; Bush, A.I. Mice transgenic for Alzheimer disease beta-amyloid develop lens cataracts that are rescued by antioxidant treatment. Free Radic. Biol. Med. 2005, 38, 258–261. [Google Scholar] [CrossRef]
  43. Sun, S.W.; Nishioka, C.; Labib, W.; Liang, H.F. Axonal Terminals Exposed to Amyloid-beta May Not Lead to Pre-Synaptic Axonal Damage. J. Alzheimers Dis. 2015, 45, 1139–1148. [Google Scholar] [CrossRef] [PubMed]
  44. Nishioka, C.; Liang, H.F.; Barsamian, B.; Sun, S.W. Amyloid-beta induced retrograde axonal degeneration in a mouse tauopathy model. Neuroimage 2019, 189, 180–191. [Google Scholar] [CrossRef]
  45. Sadun, A.A.; Bassi, C.J. Optic nerve damage in Alzheimer’s disease. Ophthalmology 1990, 97, 9–17. [Google Scholar] [CrossRef] [PubMed]
  46. Bayer, A.U.; Keller, O.N.; Ferrari, F.; Maag, K.P. Association of glaucoma with neurodegenerative diseases with apoptotic cell death: Alzheimer’s disease and Parkinson’s disease. Am. J. Ophthalmol. 2002, 133, 135–137. [Google Scholar] [CrossRef] [PubMed]
  47. Wostyn, P.; Audenaert, K.; De Deyn, P.P. Alzheimer’s disease-related changes in diseases characterized by elevation of intracranial or intraocular pressure. Clin. Neurol. Neurosurg. 2008, 110, 101–109. [Google Scholar] [CrossRef] [PubMed]
  48. Lenoir, H.; Sieroff, E. Visual perceptual disorders in Alzheimer’s disease. Geriatr. Psychol. Neuropsychiatr. Vieil. 2019, 17, 307–316. [Google Scholar] [CrossRef]
  49. Wright, L.M.; Stein, T.D.; Jun, G.; Chung, J.; McConnell, K.; Fiorello, M.; Siegel, N.; Ness, S.; Xia, W.; Turner, K.L.; et al. Association of Cognitive Function with Amyloid-beta and Tau Proteins in the Vitreous Humor. J. Alzheimers Dis. 2019, 68, 1429–1438. [Google Scholar] [CrossRef]
  50. Hart, N.J.; Koronyo, Y.; Black, K.L.; Koronyo-Hamaoui, M. Ocular indicators of Alzheimer’s: Exploring disease in the retina. Acta Neuropathol. 2016, 132, 767–787. [Google Scholar] [CrossRef]
  51. Ho, W.L.; Leung, Y.; Tsang, A.W.; So, K.F.; Chiu, K.; Chang, R.C. Review: Tauopathy in the retina and optic nerve: Does it shadow pathological changes in the brain? Mol. Vis. 2012, 18, 2700–2710. [Google Scholar] [PubMed]
  52. Chiasseu, M.; Cueva Vargas, J.L.; Destroismaisons, L.; Vande Velde, C.; Leclerc, N.; Di Polo, A. Tau Accumulation, Altered Phosphorylation, and Missorting Promote Neurodegeneration in Glaucoma. J. Neurosci. 2016, 36, 5785–5798. [Google Scholar] [CrossRef] [PubMed]
  53. Rhein, V.; Song, X.; Wiesner, A.; Ittner, L.M.; Baysang, G.; Meier, F.; Ozmen, L.; Bluethmann, H.; Drose, S.; Brandt, U.; et al. Amyloid-beta and tau synergistically impair the oxidative phosphorylation system in triple transgenic Alzheimer’s disease mice. Proc. Natl. Acad. Sci. USA 2009, 106, 20057–20062. [Google Scholar] [CrossRef] [PubMed]
  54. Michael, R.; Otto, C.; Lenferink, A.; Gelpi, E.; Montenegro, G.A.; Rosandic, J.; Tresserra, F.; Barraquer, R.I.; Vrensen, G.F. Absence of amyloid-beta in lenses of Alzheimer patients: A confocal Raman microspectroscopic study. Exp. Eye Res. 2014, 119, 44–53. [Google Scholar] [CrossRef] [PubMed]
  55. Choi, S.I.; Lee, B.; Woo, J.H.; Jeong, J.B.; Jun, I.; Kim, E.K. APP processing and metabolism in corneal fibroblasts and epithelium as a potential biomarker for Alzheimer’s disease. Exp. Eye Res. 2019, 182, 167–174. [Google Scholar] [CrossRef] [PubMed]
  56. Einarsdottir, A.B.; Hardarson, S.H.; Kristjansdottir, J.V.; Bragason, D.T.; Snaedal, J.; Stefansson, E. Retinal oximetry imaging in Alzheimer’s disease. J. Alzheimers Dis. 2016, 49, 79–83. [Google Scholar] [CrossRef]
  57. Olafsdottir, O.B.; Saevarsdottir, H.S.; Hardarson, S.H.; Hannesdottir, K.H.; Traustadottir, V.D.; Karlsson, R.A.; Einarsdottir, A.B.; Jonsdottir, K.D.; Stefansson, E.; Snaedal, J. Retinal oxygen metabolism in patients with mild cognitive impairment. Alzheimers Dement 2018, 10, 340–345. [Google Scholar] [CrossRef]
  58. Koronyo, Y.; Salumbides, B.C.; Black, K.L.; Koronyo-Hamaoui, M. Alzheimer’s disease in the retina: Imaging retinal abeta plaques for early diagnosis and therapy assessment. Neurodegener. Dis. 2012, 10, 285–293. [Google Scholar] [CrossRef] [PubMed]
  59. Cheung, C.Y.; Ong, Y.T.; Ikram, M.K.; Ong, S.Y.; Li, X.; Hilal, S.; Catindig, J.A.; Venketasubramanian, N.; Yap, P.; Seow, D.; et al. Microvascular network alterations in the retina of patients with Alzheimer’s disease. Alzheimers Dement. 2014, 10, 135–142. [Google Scholar] [CrossRef]
  60. Frost, S.; Kanagasingam, Y.; Sohrabi, H.; Vignarajan, J.; Bourgeat, P.; Salvado, O.; Villemagne, V.; Rowe, C.C.; Macaulay, S.L.; Szoeke, C.; et al. Retinal vascular biomarkers for early detection and monitoring of Alzheimer’s disease. Transl. Psychiatry 2013, 3, e233. [Google Scholar] [CrossRef]
  61. Shi, H.; Koronyo, Y.; Rentsendorj, A.; Regis, G.C.; Sheyn, J.; Fuchs, D.T.; Kramerov, A.A.; Ljubimov, A.V.; Dumitrascu, O.M.; Rodriguez, A.R.; et al. Identification of early pericyte loss and vascular amyloidosis in Alzheimer’s disease retina. Acta Neuropathol. 2020, 139, 813–836. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, S.S.; Zhu, S.Q. [Correlation between Alzheimer disease and cataract]. Zhonghua Yan Ke Za Zhi 2017, 53, 314–316. [Google Scholar] [CrossRef] [PubMed]
  63. Frederikse, P.H.; Zigler, J.S., Jr. Presenilin expression in the ocular lens. Curr. Eye Res. 1998, 17, 947–952. [Google Scholar] [CrossRef]
  64. Ho, C.Y.; Troncoso, J.C.; Knox, D.; Stark, W.; Eberhart, C.G. Beta-amyloid, phospho-tau and alpha-synuclein deposits similar to those in the brain are not identified in the eyes of Alzheimer’s and Parkinson’s disease patients. Brain Pathol. 2014, 24, 25–32. [Google Scholar] [CrossRef] [PubMed]
  65. Kwak, D.E.; Ko, T.; Koh, H.S.; Ji, Y.W.; Shin, J.; Kim, K.; Kim, H.Y.; Lee, H.K.; Kim, Y. Alterations of aqueous humor Abeta levels in Abeta-infused and transgenic mouse models of Alzheimer disease. PLoS ONE 2020, 15, e0227618. [Google Scholar] [CrossRef] [PubMed]
  66. Prakasam, A.; Muthuswamy, A.; Ablonczy, Z.; Greig, N.H.; Fauq, A.; Rao, K.J.; Pappolla, M.A.; Sambamurti, K. Differential accumulation of secreted AbetaPP metabolites in ocular fluids. J. Alzheimers Dis. 2010, 20, 1243–1253. [Google Scholar] [CrossRef] [PubMed]
  67. Cunha, J.P.; Proenca, R.; Dias-Santos, A.; Melancia, D.; Almeida, R.; Aguas, H.; Santos, B.O.; Alves, M.; Ferreira, J.; Papoila, A.L.; et al. Choroidal thinning: Alzheimer’s disease and aging. Alzheimers Dement 2017, 8, 11–17. [Google Scholar] [CrossRef]
  68. Trebbastoni, A.; Marcelli, M.; Mallone, F.; D’Antonio, F.; Imbriano, L.; Campanelli, A.; de Lena, C.; Gharbiya, M. Attenuation of Choroidal Thickness in Patients with Alzheimer Disease: Evidence From an Italian Prospective Study. Alzheimer Dis. Assoc. Disord. 2017, 31, 128–134. [Google Scholar] [CrossRef]
  69. Rizzo, M.; Anderson, S.W.; Dawson, J.; Nawrot, M. Vision and cognition in Alzheimer’s disease. Neuropsychologia 2000, 38, 1157–1169. [Google Scholar] [CrossRef] [PubMed]
  70. Chiu, K.; Chan, T.F.; Wu, A.; Leung, I.Y.; So, K.F.; Chang, R.C. Neurodegeneration of the retina in mouse models of Alzheimer’s disease: What can we learn from the retina? Age 2012, 34, 633–649. [Google Scholar] [CrossRef] [PubMed]
  71. Guo, L.; Salt, T.E.; Luong, V.; Wood, N.; Cheung, W.; Maass, A.; Ferrari, G.; Russo-Marie, F.; Sillito, A.M.; Cheetham, M.E.; et al. Targeting amyloid-beta in glaucoma treatment. Proc. Natl. Acad. Sci. USA 2007, 104, 13444–13449. [Google Scholar] [CrossRef]
  72. Edwards, M.M.; Rodriguez, J.J.; Gutierrez-Lanza, R.; Yates, J.; Verkhratsky, A.; Lutty, G.A. Retinal macroglia changes in a triple transgenic mouse model of Alzheimer’s disease. Exp. Eye Res. 2014, 127, 252–260. [Google Scholar] [CrossRef]
  73. Joly, S.; Lamoureux, S.; Pernet, V. Nonamyloidogenic processing of amyloid beta precursor protein is associated with retinal function improvement in aging male APP(swe)/PS1DeltaE9 mice. Neurobiol. Aging 2017, 53, 181–191. [Google Scholar] [CrossRef] [PubMed]
  74. Walsh, D.T.; Bresciani, L.; Saunders, D.; Manca, M.F.; Jen, A.; Gentleman, S.M.; Jen, L.S. Amyloid beta peptide causes chronic glial cell activation and neuro-degeneration after intravitreal injection. Neuropathol. Appl. Neurobiol. 2005, 31, 491–502. [Google Scholar] [CrossRef] [PubMed]
  75. Chiasseu, M.; Alarcon-Martinez, L.; Belforte, N.; Quintero, H.; Dotigny, F.; Destroismaisons, L.; Vande Velde, C.; Panayi, F.; Louis, C.; Di Polo, A. Tau accumulation in the retina promotes early neuronal dysfunction and precedes brain pathology in a mouse model of Alzheimer’s disease. Mol. Neurodegener. 2017, 12, 58. [Google Scholar] [CrossRef] [PubMed]
  76. Jarosz-Griffiths, H.H.; Noble, E.; Rushworth, J.V.; Hooper, N.M. Amyloid-beta Receptors: The Good, the Bad, and the Prion Protein. J. Biol. Chem. 2016, 291, 3174–3183. [Google Scholar] [CrossRef] [PubMed]
  77. Lauren, J.; Gimbel, D.A.; Nygaard, H.B.; Gilbert, J.W.; Strittmatter, S.M. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-beta oligomers. Nature 2009, 457, 1128–1132. [Google Scholar] [CrossRef]
  78. Nygaard, H.B.; van Dyck, C.H.; Strittmatter, S.M. Fyn kinase inhibition as a novel therapy for Alzheimer’s disease. Alzheimers Res. Ther. 2014, 6, 8. [Google Scholar] [CrossRef] [PubMed]
  79. Altmeppen, H.C.; Puig, B.; Dohler, F.; Thurm, D.K.; Falker, C.; Krasemann, S.; Glatzel, M. Proteolytic processing of the prion protein in health and disease. Am. J. Neurodegener. Dis. 2012, 1, 15–31. [Google Scholar]
  80. Asai, M.; Hattori, C.; Szabo, B.; Sasagawa, N.; Maruyama, K.; Tanuma, S.; Ishiura, S. Putative function of ADAM9, ADAM10, and ADAM17 as APP alpha-secretase. Biochem. Biophys. Res. Commun. 2003, 301, 231–235. [Google Scholar] [CrossRef]
  81. Asthana, A.; Baksi, S.; Ashok, A.; Karmakar, S.; Mammadova, N.; Kokemuller, R.; Greenlee, M.H.; Kong, Q.; Singh, N. Prion protein facilitates retinal iron uptake and is cleaved at the beta-site: Implications for retinal iron homeostasis in prion disorders. Sci. Rep. 2017, 7, 9600. [Google Scholar] [CrossRef]
  82. Castle, A.R.; Gill, A.C. Physiological Functions of the Cellular Prion Protein. Front. Mol. Biosci. 2017, 4, 19. [Google Scholar] [CrossRef]
  83. Ezpeleta, J.; Boudet-Devaud, F.; Pietri, M.; Baudry, A.; Baudouin, V.; Alleaume-Butaux, A.; Dagoneau, N.; Kellermann, O.; Launay, J.M.; Schneider, B. Protective role of cellular prion protein against TNFalpha-mediated inflammation through TACE alpha-secretase. Sci. Rep. 2017, 7, 7671. [Google Scholar] [CrossRef]
  84. Singh, N. The role of iron in prion disease and other neurodegenerative diseases. PLoS Pathog. 2014, 10, e1004335. [Google Scholar] [CrossRef]
  85. Singh, N.; Singh, A.; Das, D.; Mohan, M.L. Redox control of prion and disease pathogenesis. Antioxid. Redox Signal 2010, 12, 1271–1294. [Google Scholar] [CrossRef] [PubMed]
  86. Ramana, K.V. Tumor necrosis factor-alpha converting enzyme: Implications for ocular inflammatory diseases. Int. J. Biochem. Cell Biol. 2010, 42, 1076–1079. [Google Scholar] [CrossRef] [PubMed]
  87. Frigg, R.; Wenzel, A.; Samardzija, M.; Oesch, B.; Wariwoda, H.; Navarini, A.A.; Seeliger, M.W.; Tanimoto, N.; Reme, C.; Grimm, C. The prion protein is neuroprotective against retinal degeneration in vivo. Exp. Eye Res. 2006, 83, 1350–1358. [Google Scholar] [CrossRef] [PubMed]
  88. Loubet, D.; Dakowski, C.; Pietri, M.; Pradines, E.; Bernard, S.; Callebert, J.; Ardila-Osorio, H.; Mouillet-Richard, S.; Launay, J.M.; Kellermann, O.; et al. Neuritogenesis: The prion protein controls beta1 integrin signaling activity. FASEB J. 2012, 26, 678–690. [Google Scholar] [CrossRef] [PubMed]
  89. Alleaume-Butaux, A.; Nicot, S.; Pietri, M.; Baudry, A.; Dakowski, C.; Tixador, P.; Ardila-Osorio, H.; Haeberle, A.M.; Bailly, Y.; Peyrin, J.M.; et al. Double-Edge Sword of Sustained ROCK Activation in Prion Diseases through Neuritogenesis Defects and Prion Accumulation. PLoS Pathog. 2015, 11, e1005073. [Google Scholar] [CrossRef] [PubMed]
  90. Kim, H.J.; Choi, H.S.; Park, J.H.; Kim, M.J.; Lee, H.G.; Petersen, R.B.; Kim, Y.S.; Park, J.B.; Choi, E.K. Regulation of RhoA activity by the cellular prion protein. Cell Death Dis. 2017, 8, e2668. [Google Scholar] [CrossRef] [PubMed]
  91. Bertuchi, F.R.; Bourgeon, D.M.; Landemberger, M.C.; Martins, V.R.; Cerchiaro, G. PrPC displays an essential protective role from oxidative stress in an astrocyte cell line derived from PrPC knockout mice. Biochem. Biophys. Res. Commun. 2012, 418, 27–32. [Google Scholar] [CrossRef] [PubMed]
  92. Dupiereux, I.; Falisse-Poirrier, N.; Zorzi, W.; Watt, N.T.; Thellin, O.; Zorzi, D.; Pierard, O.; Hooper, N.M.; Heinen, E.; Elmoualij, B. Protective effect of prion protein via the N-terminal region in mediating a protective effect on paraquat-induced oxidative injury in neuronal cells. J. Neurosci. Res. 2008, 86, 653–659. [Google Scholar] [CrossRef] [PubMed]
  93. Haigh, C.L.; Collins, S.J. Endoproteolytic cleavage as a molecular switch regulating and diversifying prion protein function. Neural Regen. Res. 2016, 11, 238–239. [Google Scholar] [CrossRef] [PubMed]
  94. Zeng, F.; Watt, N.T.; Walmsley, A.R.; Hooper, N.M. Tethering the N-terminus of the prion protein compromises the cellular response to oxidative stress. J. Neurochem. 2003, 84, 480–490. [Google Scholar] [CrossRef] [PubMed]
  95. Zeng, L.; Zou, W.; Wang, G. Cellular prion protein (PrP(C)) and its role in stress responses. Int. J. Clin. Exp. Med. 2015, 8, 8042–8050. [Google Scholar] [PubMed]
  96. Brown, D.R.; Clive, C.; Haswell, S.J. Antioxidant activity related to copper binding of native prion protein. J. Neurochem. 2001, 76, 69–76. [Google Scholar] [CrossRef] [PubMed]
  97. Singh, A.; Haldar, S.; Horback, K.; Tom, C.; Zhou, L.; Meyerson, H.; Singh, N. Prion protein regulates iron transport by functioning as a ferrireductase. J. Alzheimers Dis. 2013, 35, 541–552. [Google Scholar] [CrossRef]
  98. Wille, H.; Requena, J.R. The Structure of PrP(Sc) Prions. Pathogens 2018, 7, 20. [Google Scholar] [CrossRef] [PubMed]
  99. Ramirez, A.I.; de Hoz, R.; Salobrar-Garcia, E.; Salazar, J.J.; Rojas, B.; Ajoy, D.; Lopez-Cuenca, I.; Rojas, P.; Trivino, A.; Ramirez, J.M. The Role of Microglia in Retinal Neurodegeneration: Alzheimer’s Disease, Parkinson, and Glaucoma. Front. Aging Neurosci. 2017, 9, 214. [Google Scholar] [CrossRef]
  100. Surguchov, A.; McMahan, B.; Masliah, E.; Surgucheva, I. Synucleins in ocular tissues. J. Neurosci. Res. 2001, 65, 68–77. [Google Scholar] [CrossRef]
  101. Moore, K.L.; Persaud, T.V.; Torchia, M.G. The Developing Human: Clinically Oriented Embryology, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  102. Alliot, F.; Godin, I.; Pessac, B. Microglia derive from progenitors, originating from the yolk sac, and which proliferate in the brain. Brain Res. Dev. Brain Res. 1999, 117, 145–152. [Google Scholar] [CrossRef]
  103. Ashwell, K. Microglia and cell death in the developing mouse cerebellum. Brain Res. Dev. Brain Res. 1990, 55, 219–230. [Google Scholar] [CrossRef]
  104. Li, F.; Jiang, D.; Samuel, M.A. Microglia in the developing retina. Neural Dev. 2019, 14, 12. [Google Scholar] [CrossRef] [PubMed]
  105. Ginhoux, F.; Greter, M.; Leboeuf, M.; Nandi, S.; See, P.; Gokhan, S.; Mehler, M.F.; Conway, S.J.; Ng, L.G.; Stanley, E.R.; et al. Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science 2010, 330, 841–845. [Google Scholar] [CrossRef] [PubMed]
  106. Diaz-Araya, C.M.; Provis, J.M.; Penfold, P.L.; Billson, F.A. Development of microglial topography in human retina. J. Comp. Neurol. 1995, 363, 53–68. [Google Scholar] [CrossRef] [PubMed]
  107. Santos, A.M.; Calvente, R.; Tassi, M.; Carrasco, M.C.; Martin-Oliva, D.; Marin-Teva, J.L.; Navascues, J.; Cuadros, M.A. Embryonic and postnatal development of microglial cells in the mouse retina. J. Comp. Neurol. 2008, 506, 224–239. [Google Scholar] [CrossRef] [PubMed]
  108. Ashwell, K.W.; Hollander, H.; Streit, W.; Stone, J. The appearance and distribution of microglia in the developing retina of the rat. Vis. Neurosci. 1989, 2, 437–448. [Google Scholar] [CrossRef]
  109. Martin-Estebane, M.; Navascues, J.; Sierra-Martin, A.; Martin-Guerrero, S.M.; Cuadros, M.A.; Carrasco, M.C.; Marin-Teva, J.L. Onset of microglial entry into developing quail retina coincides with increased expression of active caspase-3 and is mediated by extracellular ATP and UDP. PLoS ONE 2017, 12, e0182450. [Google Scholar] [CrossRef]
  110. Marin-Teva, J.L.; Calvente, R.; Cuadros, M.A.; Almendros, A.; Navascues, J. Circumferential migration of ameboid microglia in the margin of the developing quail retina. Glia 1999, 27, 226–238. [Google Scholar] [CrossRef]
  111. Provis, J.M.; Diaz, C.M.; Penfold, P.L. Microglia in human retina: A heterogeneous population with distinct ontogenies. Perspect. Dev. Neurobiol. 1996, 3, 213–222. [Google Scholar]
  112. Chen, L.; Yang, P.; Kijlstra, A. Distribution, markers, and functions of retinal microglia. Ocul. Immunol. Inflamm. 2002, 10, 27–39. [Google Scholar] [CrossRef] [PubMed]
  113. Pei, Y.F.; Rhodin, J.A. The prenatal development of the mouse eye. Anat. Rec. 1970, 168, 105–125. [Google Scholar] [CrossRef] [PubMed]
  114. Sernagor, E.; Eglen, S.J.; Wong, R.O. Development of retinal ganglion cell structure and function. Prog. Retin. Eye Res. 2001, 20, 139–174. [Google Scholar] [CrossRef] [PubMed]
  115. Silverman, S.M.; Wong, W.T. Microglia in the Retina: Roles in Development, Maturity, and Disease. Annu. Rev. Vis. Sci. 2018, 4, 45–77. [Google Scholar] [CrossRef] [PubMed]
  116. Walker, D.G.; Lue, L.F. Immune phenotypes of microglia in human neurodegenerative disease: Challenges to detecting microglial polarization in human brains. Alzheimers Res. Ther. 2015, 7, 56. [Google Scholar] [CrossRef]
  117. Luo, C.; Jian, C.; Liao, Y.; Huang, Q.; Wu, Y.; Liu, X.; Zou, D.; Wu, Y. The role of microglia in multiple sclerosis. Neuropsychiatr. Dis. Treat. 2017, 13, 1661–1667. [Google Scholar] [CrossRef] [PubMed]
  118. Akinrinmade, O.A.; Chetty, S.; Daramola, A.K.; Islam, M.U.; Thepen, T.; Barth, S. CD64: An Attractive Immunotherapeutic Target for M1-type Macrophage Mediated Chronic Inflammatory Diseases. Biomedicines 2017, 5, 56. [Google Scholar] [CrossRef] [PubMed]
  119. Holloway, O.G.; Canty, A.J.; King, A.E.; Ziebell, J.M. Rod microglia and their role in neurological diseases. Semin. Cell Dev. Biol. 2019, 94, 96–103. [Google Scholar] [CrossRef]
  120. Walker, F.R.; Beynon, S.B.; Jones, K.A.; Zhao, Z.; Kongsui, R.; Cairns, M.; Nilsson, M. Dynamic structural remodelling of microglia in health and disease: A review of the models, the signals and the mechanisms. Brain Behav. Immun. 2014, 37, 1–14. [Google Scholar] [CrossRef] [PubMed]
  121. Torres-Platas, S.G.; Comeau, S.; Rachalski, A.; Bo, G.D.; Cruceanu, C.; Turecki, G.; Giros, B.; Mechawar, N. Morphometric characterization of microglial phenotypes in human cerebral cortex. J. Neuroinflammation 2014, 11, 12. [Google Scholar] [CrossRef]
  122. Nimmerjahn, A.; Kirchhoff, F.; Helmchen, F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 2005, 308, 1314–1318. [Google Scholar] [CrossRef] [PubMed]
  123. Karperien, A.; Ahammer, H.; Jelinek, H.F. Quantitating the subtleties of microglial morphology with fractal analysis. Front. Cell Neurosci. 2013, 7, 3. [Google Scholar] [CrossRef] [PubMed]
  124. Davis, B.M.; Salinas-Navarro, M.; Cordeiro, M.F.; Moons, L.; De Groef, L. Characterizing microglia activation: A spatial statistics approach to maximize information extraction. Sci. Rep. 2017, 7, 1576. [Google Scholar] [CrossRef] [PubMed]
  125. Giordano, K.R.; Denman, C.R.; Dubisch, P.S.; Akhter, M.; Lifshitz, J. An update on the rod microglia variant in experimental and clinical brain injury and disease. Brain Commun. 2021, 3, fcaa227. [Google Scholar] [CrossRef] [PubMed]
  126. Taylor, S.E.; Morganti-Kossmann, C.; Lifshitz, J.; Ziebell, J.M. Rod microglia: A morphological definition. PLoS ONE 2014, 9, e97096. [Google Scholar] [CrossRef] [PubMed]
  127. de Hoz, R.; Gallego, B.I.; Ramirez, A.I.; Rojas, B.; Salazar, J.J.; Valiente-Soriano, F.J.; Aviles-Trigueros, M.; Villegas-Perez, M.P.; Vidal-Sanz, M.; Trivino, A.; et al. Rod-like microglia are restricted to eyes with laser-induced ocular hypertension but absent from the microglial changes in the contralateral untreated eye. PLoS ONE 2013, 8, e83733. [Google Scholar] [CrossRef]
  128. Yuan, T.F.; Liang, Y.X.; Peng, B.; Lin, B.; So, K.F. Local proliferation is the main source of rod microglia after optic nerve transection. Sci. Rep. 2015, 5, 10788. [Google Scholar] [CrossRef] [PubMed]
  129. Dubbelaar, M.L.; Kracht, L.; Eggen, B.J.L.; Boddeke, E. The Kaleidoscope of Microglial Phenotypes. Front. Immunol. 2018, 9, 1753. [Google Scholar] [CrossRef]
  130. Fernandez-Arjona, M.D.M.; Grondona, J.M.; Granados-Duran, P.; Fernandez-Llebrez, P.; Lopez-Avalos, M.D. Microglia Morphological Categorization in a Rat Model of Neuroinflammation by Hierarchical Cluster and Principal Components Analysis. Front. Cell Neurosci. 2017, 11, 235. [Google Scholar] [CrossRef]
  131. Okajima, T.; Tsuruta, F. Microglial dynamics during brain development. Neural Regen. Res. 2018, 13, 222–223. [Google Scholar] [CrossRef]
  132. Perez-Pouchoulen, M.; VanRyzin, J.W.; McCarthy, M.M. Morphological and Phagocytic Profile of Microglia in the Developing Rat Cerebellum. eNeuro 2015, 2, ENEURO.0036-15.2015. [Google Scholar] [CrossRef]
  133. Davis, E.J.; Foster, T.D.; Thomas, W.E. Cellular forms and functions of brain microglia. Brain Res. Bull. 1994, 34, 73–78. [Google Scholar] [CrossRef]
  134. Kreutzberg, G.W. Microglia: A sensor for pathological events in the CNS. Trends Neurosci. 1996, 19, 312–318. [Google Scholar] [CrossRef]
  135. Koso, H.; Tsuhako, A.; Lai, C.Y.; Baba, Y.; Otsu, M.; Ueno, K.; Nagasaki, M.; Suzuki, Y.; Watanabe, S. Conditional rod photoreceptor ablation reveals Sall1 as a microglial marker and regulator of microglial morphology in the retina. Glia 2016, 64, 2005–2024. [Google Scholar] [CrossRef] [PubMed]
  136. Bachstetter, A.D.; Ighodaro, E.T.; Hassoun, Y.; Aldeiri, D.; Neltner, J.H.; Patel, E.; Abner, E.L.; Nelson, P.T. Rod-shaped microglia morphology is associated with aging in 2 human autopsy series. Neurobiol. Aging 2017, 52, 98–105. [Google Scholar] [CrossRef] [PubMed]
  137. Zeng, H.Y.; Green, W.R.; Tso, M.O. Microglial activation in human diabetic retinopathy. Arch. Ophthalmol. 2008, 126, 227–232. [Google Scholar] [CrossRef] [PubMed]
  138. Zhao, L.; Zabel, M.K.; Wang, X.; Ma, W.; Shah, P.; Fariss, R.N.; Qian, H.; Parkhurst, C.N.; Gan, W.B.; Wong, W.T. Microglial phagocytosis of living photoreceptors contributes to inherited retinal degeneration. EMBO Mol. Med. 2015, 7, 1179–1197. [Google Scholar] [CrossRef]
  139. O’Koren, E.G.; Yu, C.; Klingeborn, M.; Wong, A.Y.W.; Prigge, C.L.; Mathew, R.; Kalnitsky, J.; Msallam, R.A.; Silvin, A.; Kay, J.N.; et al. Microglial Function Is Distinct in Different Anatomical Locations during Retinal Homeostasis and Degeneration. Immunity 2019, 50, 723–737.e727. [Google Scholar] [CrossRef]
  140. Tang, P.H.; Pierson, M.J.; Heuss, N.D.; Gregerson, D.S. A subpopulation of activated retinal macrophages selectively migrated to regions of cone photoreceptor stress, but had limited effect on cone death in a mouse model for type 2 Leber congenital amaurosis. Mol. Cell Neurosci. 2017, 85, 70–81. [Google Scholar] [CrossRef] [PubMed]
  141. Tan, Y.L.; Yuan, Y.; Tian, L. Microglial regional heterogeneity and its role in the brain. Mol. Psychiatry 2020, 25, 351–367. [Google Scholar] [CrossRef]
  142. Hendrickx, D.A.; Schuurman, K.G.; van Draanen, M.; Hamann, J.; Huitinga, I. Enhanced uptake of multiple sclerosis-derived myelin by THP-1 macrophages and primary human microglia. J. Neuroinflamm. 2014, 11, 64. [Google Scholar] [CrossRef] [PubMed]
  143. Choi, S.; Guo, L.; Cordeiro, M.F. Retinal and Brain Microglia in Multiple Sclerosis and Neurodegeneration. Cells 2021, 10, 1507. [Google Scholar] [CrossRef]
  144. Baik, S.H.; Kang, S.; Son, S.M.; Mook-Jung, I. Microglia contributes to plaque growth by cell death due to uptake of amyloid beta in the brain of Alzheimer’s disease mouse model. Glia 2016, 64, 2274–2290. [Google Scholar] [CrossRef]
  145. Cao, S.; Fisher, D.W.; Rodriguez, G.; Yu, T.; Dong, H. Comparisons of neuroinflammation, microglial activation, and degeneration of the locus coeruleus-norepinephrine system in APP/PS1 and aging mice. J. Neuroinflamm. 2021, 18, 10. [Google Scholar] [CrossRef] [PubMed]
  146. Grimaldi, A.; Brighi, C.; Peruzzi, G.; Ragozzino, D.; Bonanni, V.; Limatola, C.; Ruocco, G.; Di Angelantonio, S. Inflammation, neurodegeneration and protein aggregation in the retina as ocular biomarkers for Alzheimer’s disease in the 3xTg-AD mouse model. Cell Death Dis. 2018, 9, 685. [Google Scholar] [CrossRef]
  147. Vandenabeele, M.; Veys, L.; Lemmens, S.; Hadoux, X.; Gelders, G.; Masin, L.; Serneels, L.; Theunis, J.; Saito, T.; Saido, T.C.; et al. The App(NL-G-F) mouse retina is a site for preclinical Alzheimer’s disease diagnosis and research. Acta Neuropathol. Commun. 2021, 9, 6. [Google Scholar] [CrossRef] [PubMed]
  148. Yao, K.; Zu, H.B. Microglial polarization: Novel therapeutic mechanism against Alzheimer’s disease. Inflammopharmacology 2020, 28, 95–110. [Google Scholar] [CrossRef]
  149. Ma, W.; Cojocaru, R.; Gotoh, N.; Gieser, L.; Villasmil, R.; Cogliati, T.; Swaroop, A.; Wong, W.T. Gene expression changes in aging retinal microglia: Relationship to microglial support functions and regulation of activation. Neurobiol. Aging 2013, 34, 2310–2321. [Google Scholar] [CrossRef]
  150. Grimaldi, A.; Pediconi, N.; Oieni, F.; Pizzarelli, R.; Rosito, M.; Giubettini, M.; Santini, T.; Limatola, C.; Ruocco, G.; Ragozzino, D.; et al. Neuroinflammatory Processes, A1 Astrocyte Activation and Protein Aggregation in the Retina of Alzheimer’s Disease Patients, Possible Biomarkers for Early Diagnosis. Front. Neurosci. 2019, 13, 925. [Google Scholar] [CrossRef]
  151. Okunuki, Y.; Mukai, R.; Nakao, T.; Tabor, S.J.; Butovsky, O.; Dana, R.; Ksander, B.R.; Connor, K.M. Retinal microglia initiate neuroinflammation in ocular autoimmunity. Proc. Natl. Acad. Sci. USA 2019, 116, 9989–9998. [Google Scholar] [CrossRef]
  152. Koizumi, T.; Kerkhofs, D.; Mizuno, T.; Steinbusch, H.W.M.; Foulquier, S. Vessel-Associated Immune Cells in Cerebrovascular Diseases: From Perivascular Macrophages to Vessel-Associated Microglia. Front. Neurosci. 2019, 13, 1291. [Google Scholar] [CrossRef]
  153. Lopez-de-Eguileta, A.; Cervero, A.; Ruiz de Sabando, A.; Sanchez-Juan, P.; Casado, A. Ganglion Cell Layer Thinning in Alzheimer’s Disease. Medicina 2020, 56, 553. [Google Scholar] [CrossRef] [PubMed]
  154. Zhao, H.; Chang, R.; Che, H.; Wang, J.; Yang, L.; Fang, W.; Xia, Y.; Li, N.; Ma, Q.; Wang, X. Hyperphosphorylation of tau protein by calpain regulation in retina of Alzheimer’s disease transgenic mouse. Neurosci. Lett. 2013, 551, 12–16. [Google Scholar] [CrossRef]
  155. Chiquita, S.; Rodrigues-Neves, A.C.; Baptista, F.I.; Carecho, R.; Moreira, P.I.; Castelo-Branco, M.; Ambrosio, A.F. The Retina as a Window or Mirror of the Brain Changes Detected in Alzheimer’s Disease: Critical Aspects to Unravel. Mol. Neurobiol. 2019, 56, 5416–5435. [Google Scholar] [CrossRef] [PubMed]
  156. Markram, H. The blue brain project. Nat. Rev. Neurosci. 2006, 7, 153–160. [Google Scholar] [CrossRef]
  157. Sharpee, T.O.; Destexhe, A.; Kawato, M.; Sekulic, V.; Skinner, F.K.; Wojcik, D.K.; Chintaluri, C.; Cserpan, D.; Somogyvari, Z.; Kim, J.K.; et al. 25th Annual Computational Neuroscience Meeting: CNS-2016. BMC Neurosci. 2016, 17 (Suppl. S1), 54. [Google Scholar] [CrossRef] [PubMed]
  158. Ge, Y.J.; Xu, W.; Ou, Y.N.; Qu, Y.; Ma, Y.H.; Huang, Y.Y.; Shen, X.N.; Chen, S.D.; Tan, L.; Zhao, Q.H.; et al. Retinal biomarkers in Alzheimer’s disease and mild cognitive impairment: A systematic review and meta-analysis. Ageing Res. Rev. 2021, 69, 101361. [Google Scholar] [CrossRef] [PubMed]
  159. Lenahan, C.; Sanghavi, R.; Huang, L.; Zhang, J.H. Rhodopsin: A Potential Biomarker for Neurodegenerative Diseases. Front. Neurosci. 2020, 14, 326. [Google Scholar] [CrossRef] [PubMed]
  160. Cheignon, C.; Tomas, M.; Bonnefont-Rousselot, D.; Faller, P.; Hureau, C.; Collin, F. Oxidative stress and the amyloid beta peptide in Alzheimer’s disease. Redox Biol. 2018, 14, 450–464. [Google Scholar] [CrossRef] [PubMed]
  161. Fernandez-Albarral, J.A.; Salobrar-Garcia, E.; Martinez-Paramo, R.; Ramirez, A.I.; de Hoz, R.; Ramirez, J.M.; Salazar, J.J. Retinal glial changes in Alzheimer’s disease—A review. J. Optom. 2019, 12, 198–207. [Google Scholar] [CrossRef]
  162. Nilson, A.N.; English, K.C.; Gerson, J.E.; Barton Whittle, T.; Nicolas Crain, C.; Xue, J.; Sengupta, U.; Castillo-Carranza, D.L.; Zhang, W.; Gupta, P.; et al. Tau Oligomers Associate with Inflammation in the Brain and Retina of Tauopathy Mice and in Neurodegenerative Diseases. J. Alzheimers Dis. 2017, 55, 1083–1099. [Google Scholar] [CrossRef]
  163. Movio, M.I.; de Lima-Vasconcellos, T.H.; Dos Santos, G.B.; Echeverry, M.B.; Colombo, E.; Mattos, L.S.; Resende, R.R.; Kihara, A.H. Retinal organoids from human-induced pluripotent stem cells: From studying retinal dystrophies to early diagnosis of Alzheimer’s and Parkinson’s disease. Semin. Cell Dev. Biol. 2023, 144, 77–86. [Google Scholar] [CrossRef]
  164. Golzan, S.M.; Goozee, K.; Georgevsky, D.; Avolio, A.; Chatterjee, P.; Shen, K.; Gupta, V.; Chung, R.; Savage, G.; Orr, C.F.; et al. Retinal vascular and structural changes are associated with amyloid burden in the elderly: Ophthalmic biomarkers of preclinical Alzheimer’s disease. Alzheimers Res. Ther. 2017, 9, 13. [Google Scholar] [CrossRef] [PubMed]
  165. Gupta, V.B.; Chitranshi, N.; den Haan, J.; Mirzaei, M.; You, Y.; Lim, J.K.; Basavarajappa, D.; Godinez, A.; Di Angelantonio, S.; Sachdev, P.; et al. Retinal changes in Alzheimer’s disease- integrated prospects of imaging, functional and molecular advances. Prog. Retin. Eye Res. 2021, 82, 100899. [Google Scholar] [CrossRef]
  166. Johnson, E.C.B.; Dammer, E.B.; Duong, D.M.; Ping, L.; Zhou, M.; Yin, L.; Higginbotham, L.A.; Guajardo, A.; White, B.; Troncoso, J.C.; et al. Large-scale proteomic analysis of Alzheimer’s disease brain and cerebrospinal fluid reveals early changes in energy metabolism associated with microglia and astrocyte activation. Nat. Med. 2020, 26, 769–780. [Google Scholar] [CrossRef] [PubMed]
  167. Mirzaei, M.; Pushpitha, K.; Deng, L.; Chitranshi, N.; Gupta, V.; Rajput, R.; Mangani, A.B.; Dheer, Y.; Godinez, A.; McKay, M.J.; et al. Upregulation of Proteolytic Pathways and Altered Protein Biosynthesis Underlie Retinal Pathology in a Mouse Model of Alzheimer’s Disease. Mol. Neurobiol. 2019, 56, 6017–6034. [Google Scholar] [CrossRef] [PubMed]
  168. Penney, J.; Ralvenius, W.T.; Tsai, L.H. Modeling Alzheimer’s disease with iPSC-derived brain cells. Mol. Psychiatry 2020, 25, 148–167. [Google Scholar] [CrossRef] [PubMed]
  169. Artero Castro, A.; Rodriguez Jimenez, F.J.; Jendelova, P.; Erceg, S. Deciphering retinal diseases through the generation of three dimensional stem cell-derived organoids: Concise Review. Stem Cells 2019, 37, 1496–1504. [Google Scholar] [CrossRef] [PubMed]
  170. Brighi, C.; Cordella, F.; Chiriatti, L.; Soloperto, A.; Di Angelantonio, S. Retinal and Brain Organoids: Bridging the Gap Between in vivo Physiology and in vitro Micro-Physiology for the Study of Alzheimer’s Diseases. Front. Neurosci. 2020, 14, 655. [Google Scholar] [CrossRef]
  171. Jiang, H.; Wang, J.; Levin, B.E.; Baumel, B.S.; Camargo, C.J.; Signorile, J.F.; Rundek, T. Retinal Microvascular Alterations as the Biomarkers for Alzheimer Disease: Are We There Yet? J. Neuroophthalmol. 2021, 41, 251–260. [Google Scholar] [CrossRef] [PubMed]
  172. Xu, M.; Zhang, L.; Liu, G.; Jiang, N.; Zhou, W.; Zhang, Y. Pathological Changes in Alzheimer’s Disease Analyzed Using Induced Pluripotent Stem Cell-Derived Human Microglia-Like Cells. J. Alzheimers Dis. 2019, 67, 357–368. [Google Scholar] [CrossRef]
  173. Abud, E.M.; Ramirez, R.N.; Martinez, E.S.; Healy, L.M.; Nguyen, C.H.H.; Newman, S.A.; Yeromin, A.V.; Scarfone, V.M.; Marsh, S.E.; Fimbres, C.; et al. iPSC-Derived Human Microglia-like Cells to Study Neurological Diseases. Neuron 2017, 94, 278–293.e9. [Google Scholar] [CrossRef] [PubMed]
  174. Henstridge, C.M.; Spires-Jones, T.L. Modeling Alzheimer’s disease brains in vitro. Nat. Neurosci. 2018, 21, 899–900. [Google Scholar] [CrossRef] [PubMed]
  175. Campos, P.B.; Paulsen, B.S.; Rehen, S.K. Accelerating neuronal aging in in vitro model brain disorders: A focus on reactive oxygen species. Front. Aging Neurosci. 2014, 6, 292. [Google Scholar] [CrossRef] [PubMed]
  176. Grenier, K.; Kao, J.; Diamandis, P. Three-dimensional modeling of human neurodegeneration: Brain organoids coming of age. Mol. Psychiatry 2020, 25, 254–274. [Google Scholar] [CrossRef] [PubMed]
  177. Cideciyan, A.V.; Jacobson, S.G.; Aleman, T.S.; Gu, D.; Pearce-Kelling, S.E.; Sumaroka, A.; Acland, G.M.; Aguirre, G.D. In vivo dynamics of retinal injury and repair in the rhodopsin mutant dog model of human retinitis pigmentosa. Proc. Natl. Acad. Sci. USA 2005, 102, 5233–5238. [Google Scholar] [CrossRef] [PubMed]
  178. Xiong, B.; Bellen, H.J. Rhodopsin homeostasis and retinal degeneration: Lessons from the fly. Trends Neurosci. 2013, 36, 652–660. [Google Scholar] [CrossRef]
  179. Di Pierdomenico, J.; Garcia-Ayuso, D.; Pinilla, I.; Cuenca, N.; Vidal-Sanz, M.; Agudo-Barriuso, M.; Villegas-Perez, M.P. Early Events in Retinal Degeneration Caused by Rhodopsin Mutation or Pigment Epithelium Malfunction: Differences and Similarities. Front. Neuroanat. 2017, 11, 14. [Google Scholar] [CrossRef]
  180. Ni, J.D.; Baik, L.S.; Holmes, T.C.; Montell, C. A rhodopsin in the brain functions in circadian photoentrainment in Drosophila. Nature 2017, 545, 340–344. [Google Scholar] [CrossRef] [PubMed]
  181. Reeves, B.C.; Karimy, J.K.; Kundishora, A.J.; Mestre, H.; Cerci, H.M.; Matouk, C.; Alper, S.L.; Lundgaard, I.; Nedergaard, M.; Kahle, K.T. Glymphatic System Impairment in Alzheimer’s Disease and Idiopathic Normal Pressure Hydrocephalus. Trends Mol. Med. 2020, 26, 285–295. [Google Scholar] [CrossRef]
  182. Stojanovic, A.; Stitham, J.; Hwa, J. Critical role of transmembrane segment zinc binding in the structure and function of rhodopsin. J. Biol. Chem. 2004, 279, 35932–35941. [Google Scholar] [CrossRef]
  183. Ugarte, M.; Osborne, N.N. Zinc in the retina. Prog. Neurobiol. 2001, 64, 219–249. [Google Scholar] [CrossRef] [PubMed]
  184. del Valle, L.J.; Ramon, E.; Canavate, X.; Dias, P.; Garriga, P. Zinc-induced decrease of the thermal stability and regeneration of rhodopsin. J. Biol. Chem. 2003, 278, 4719–4724. [Google Scholar] [CrossRef] [PubMed]
  185. Das, N.; Raymick, J.; Sarkar, S. Role of metals in Alzheimer’s disease. Metab. Brain Dis. 2021, 36, 1627–1639. [Google Scholar] [CrossRef] [PubMed]
  186. Doustar, J.; Torbati, T.; Black, K.L.; Koronyo, Y.; Koronyo-Hamaoui, M. Optical Coherence Tomography in Alzheimer’s Disease and Other Neurodegenerative Diseases. Front. Neurol. 2017, 8, 701. [Google Scholar] [CrossRef] [PubMed]
  187. Ehler, M.; Dobrosotskaya, J.; Cunningham, D.; Wong, W.T.; Chew, E.Y.; Czaja, W.; Bonner, R.F. Modeling Photo-Bleaching Kinetics to Create High Resolution Maps of Rod Rhodopsin in the Human Retina. PLoS ONE 2015, 10, e0131881. [Google Scholar] [CrossRef]
  188. Jackson, G.R.; Owsley, C.; McGwin, G., Jr. Aging and dark adaptation. Vision. Res. 1999, 39, 3975–3982. [Google Scholar] [CrossRef]
Figure 1. Where Aβ and p-tau were found in eye tissues.
Figure 1. Where Aβ and p-tau were found in eye tissues.
Biomedicines 11 03258 g001
Figure 2. Microglia polarization.
Figure 2. Microglia polarization.
Biomedicines 11 03258 g002
Figure 3. Microglial morphotypes.
Figure 3. Microglial morphotypes.
Biomedicines 11 03258 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Donato, L.; Mordà, D.; Scimone, C.; Alibrandi, S.; D’Angelo, R.; Sidoti, A. Bridging Retinal and Cerebral Neurodegeneration: A Focus on Crosslinks between Alzheimer–Perusini’s Disease and Retinal Dystrophies. Biomedicines 2023, 11, 3258. https://doi.org/10.3390/biomedicines11123258

AMA Style

Donato L, Mordà D, Scimone C, Alibrandi S, D’Angelo R, Sidoti A. Bridging Retinal and Cerebral Neurodegeneration: A Focus on Crosslinks between Alzheimer–Perusini’s Disease and Retinal Dystrophies. Biomedicines. 2023; 11(12):3258. https://doi.org/10.3390/biomedicines11123258

Chicago/Turabian Style

Donato, Luigi, Domenico Mordà, Concetta Scimone, Simona Alibrandi, Rosalia D’Angelo, and Antonina Sidoti. 2023. "Bridging Retinal and Cerebral Neurodegeneration: A Focus on Crosslinks between Alzheimer–Perusini’s Disease and Retinal Dystrophies" Biomedicines 11, no. 12: 3258. https://doi.org/10.3390/biomedicines11123258

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

Article Metrics

Back to TopTop