Next Article in Journal
Mangiferin as New Potential Anti-Cancer Agent and Mangiferin-Integrated Polymer Systems—A Novel Research Direction
Next Article in Special Issue
Nradd Acts as a Negative Feedback Regulator of Wnt/β-Catenin Signaling and Promotes Apoptosis
Previous Article in Journal
Jasmonates, Ethylene and Brassinosteroids Control Adventitious and Lateral Rooting as Stress Avoidance Responses to Heavy Metals and Metalloids
Previous Article in Special Issue
Micromolar Valproic Acid Doses Preserve Survival and Induce Molecular Alterations in Neurodevelopmental Genes in Two Strains of Zebrafish Larvae
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Zebrafish Models of Photoreceptor Dysfunction and Degeneration

by
Nicole C. L. Noel
1,*,
Ian M. MacDonald
1,2 and
W. Ted Allison
1,3,4
1
Department of Medical Genetics, University of Alberta, Edmonton, AB T6G 2H7, Canada
2
Department of Ophthalmology and Visual Sciences, University of Alberta, Edmonton, AB T6G 2R7, Canada
3
Department of Biological Sciences, University of Alberta, Edmonton, AB T6G 2E9, Canada
4
Centre for Prions and Protein Folding Diseases, University of Alberta, Edmonton, AB T6G 2M8, Canada
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(1), 78; https://doi.org/10.3390/biom11010078
Submission received: 19 December 2020 / Revised: 2 January 2021 / Accepted: 4 January 2021 / Published: 9 January 2021
(This article belongs to the Special Issue Zebrafish: A Model for the Study of Human Diseases)

Abstract

:
Zebrafish are an instrumental system for the generation of photoreceptor degeneration models, which can be utilized to determine underlying causes of photoreceptor dysfunction and death, and for the analysis of potential therapeutic compounds, as well as the characterization of regenerative responses. We review the wealth of information from existing zebrafish models of photoreceptor disease, specifically as they relate to currently accepted taxonomic classes of human rod and cone disease. We also highlight that rich, detailed information can be derived from studying photoreceptor development, structure, and function, including behavioural assessments and in vivo imaging of zebrafish. Zebrafish models are available for a diversity of photoreceptor diseases, including cone dystrophies, which are challenging to recapitulate in nocturnal mammalian systems. Newly discovered models of photoreceptor disease and drusenoid deposit formation may not only provide important insights into pathogenesis of disease, but also potential therapeutic approaches. Zebrafish have already shown their use in providing pre-clinical data prior to testing genetic therapies in clinical trials, such as antisense oligonucleotide therapy for Usher syndrome.

1. Introduction

Vision is possible due to specialized light-detecting cells in the eye called photoreceptors. The degeneration of photoreceptors or the retinal pigment epithelium (RPE), essential retinal support cells, results in permanent vision loss and blindness. Photoreceptor diseases are extremely diverse, which creates challenges for development of treatments to prevent or reverse vision loss [1]. The zebrafish model provides opportunity to model diverse blinding diseases and elucidate therapeutic avenues. This review highlights the utility of zebrafish for modelling inherited photoreceptor disease by: (i) introducing photoreceptors and photoreceptor conditions with a genetic component, (ii) outlining zebrafish photoreceptor organization and tools to assess photoreceptor health, (iii) summarizing zebrafish models of photoreceptor disease, (iv) discussing how these models are informative for human disease, and (v) presenting the therapeutic contributions of these models. Ultimately, zebrafish provide a valuable platform for defining disease mechanisms and testing therapeutics that can be used in the clinic.

2. Photoreceptors and Their Diseases

Photoreceptors are sensory neurons responsible for converting light into signals that are transmitted to the brain for interpretation, allowing the perception of visual information. There are two types of retinal photoreceptors, named after their distinct morphologies: cones, which are maximally sensitive to particular wavelengths of light; and rods, which are extremely sensitive to low levels of light and can detect single photons. Cones allow for high-acuity daytime and colour vision, while rods facilitate vision under dim-light and night conditions. Rods and cones have an outer segment (OS), the light-sensitive part of the photoreceptor; an inner segment (IS), a specialized compartment packed with mitochondria that is connected to the OS by a structure called the connecting cilium (CC); a cell body; and a synaptic terminal, where downstream cells innervate the photoreceptors (Figure 1A). The OS is a modified cilium that is responsible for the photoreceptor names—cones have a cone-shaped OS, while rods have a rod-shaped OS. The OS contains many membranous discs that are packed with opsin, the light-sensitive protein that triggers a signalling cascade, which results in signal transmission to downstream neurons. In cones, OS discs are contiguous with the plasma membrane, whereas rod OS discs are internalized within the plasma membrane.
The human retina has three cone photoreceptor subtypes: blue light-sensitive (OPN1SW1), green light-sensitive (OPN1MW1), and red light-sensitive (OPN1LW1) cones [2,3]. The photoreceptors are organized such that the peripheral retina is rod-dense with cones interspersed throughout, and the central retina is cone-dense (Figure 1B) [4,5]. This cone-dense central region is called the macula, in the middle of which is a pit that contains exclusively cone photoreceptors, known as the fovea [6]. The fovea is important for high-acuity central vision. Fovea-like structures are rarely seen outside of primate species.
Photoreceptors are supported by the RPE. The RPE plays many essential roles in the retina, including maintenance of the blood–retina barrier, absorption of stray photons, transport of materials to and from the underlying choroid vessels, phagocytosis of shed photoreceptor OS fragments, and recycling of chromophores necessary for the visual cycle [7]. The RPE also secretes factors and signalling molecules important for photoreceptor function and neuroprotection, including pigment epithelium-derived factor (PEDF), which aids in photoreceptor survival [8,9,10].
Photoreceptor dysfunction and degeneration can be caused by mutations in many aspects of the photoreceptor cells, including structural factors, cilium components, phototransduction machinery, and ion channels [1]. In addition, RPE dysfunction leads to photoreceptor disease as the RPE has essential roles in photoreceptor maintenance and function. The complexities of photoreceptor development, function, and maintenances can result in extreme genetic and phenotypic heterogeneity for photoreceptor conditions. Moreover, photoreceptor disease can manifest as part of multi-system conditions, such as ciliopathies [11].
In the following sections, photoreceptor diseases with a genetic component are described. Inherited photoreceptor diseases cumulatively affect one in 2000 to one in 3000 individuals, and age-related macular degeneration, a multifactorial condition with a genetic component, is the leading cause of vision loss in the ageing population [12,13,14]. Vision loss has an immense impact on quality of life, and unfortunately these conditions have few effective therapies. This necessitates animal models of photoreceptor disease, as targeted therapeutics cannot be developed without an understanding of disease pathology and underlying mechanisms.

2.1. Retinitis Pigmentosa

Retinitis pigmentosa (RP) is a group of rod degenerative diseases, characterized by night blindness and peripheral vision loss. RP is the most common inherited photoreceptor condition, affecting one in 3500 to one in 5000 individuals. During RP progression, the RPE becomes degenerate and pigment granules can translocate into the retina, resulting in a feature called bone spicule. Cones are spared initially in RP, but eventually degenerate, leading to total blindness. Over 70 genes have been associated with RP, and RP can be inherited in an autosomal recessive, dominant, X-linked, digenic, or mitochondrial manner [1]. Additionally, RP can be caused by mutations that directly affect the rod photoreceptors or the RPE. RP can manifest alone or in combination with extraocular phenotypes in syndromes.

2.2. Leber Congenital Amaurosis

Leber congenital amaurosis (LCA) is a severe, early-onset disease of the RPE and photoreceptors. LCA is caused by mutations in essential RPE and photoreceptor genes. The genes most commonly mutated in LCA patients are RPE65, GUCY2D, and CEP290 [15,16,17,18]. The RPE65 gene produces an RPE-specific retinoid isomerohydrolase, which is essential for recycling of components required for the photortransduction cascade [19]. GUCY2D produces the protein retinal guanylate cyclase-1 (RETGC1), which is also needed for phototransduction [20,21]. CEP290 encodes a centrosomal protein important for development of the centrosome and cilium [22].

2.3. Choroideremia

Choroideremia is an X-linked condition characterized by degeneration of the photoreceptors, RPE, and underlying choroid vasculature. Choroideremia is caused predominantly by mutations in the gene CHM, which encodes the protein REP1 [23]. REP1 posttranslationally modifies Rab proteins, which are essential for Rab protein function. Defects in Rab protein function would impact processes in many retinal cell types, but there is evidence from animal models that the RPE defects lead to photoreceptor degeneration when REP1 is lost [24,25].

2.4. Cone-Rod Dystrophy

Cone-rod dystrophies (CORD) are conditions where both the cones and rods are affected. CORD patients can present with visual acuity loss, light sensitivity, colour vision defects, central vision loss, and partial peripheral vision loss.

2.5. Enhanced S-cone Syndrome

Enhanced S-cone syndrome is a unique condition, as it presents with increased photoreceptor function. Enhanced S-cone syndrome is caused by mutations in the transcription factors NR2E3 or NRL, which are required for rod development [26,27,28,29]. Patients with enhanced S-cone syndrome therefore have cone-only retinas with enhanced blue cone sensitivity and experience night blindness and photoreceptor degeneration [26].

2.6. Cone Dystrophy

Cone dystrophies are conditions that affect the cone photoreceptors, leading to their dysfunction and degeneration. Cone dystrophies typically present with visual acuity defects, colour vision abnormalities, and reduced cone responses, measured by electroretinography (ERG). Although extremely rare, there are reported cases of peripheral cone dystrophies, whereby the cones in the peripheral retina are more affected than those in the central macula [30].
Macular degenerations involve loss of cones in the macula and leads to visual acuity decline, colour vision disturbances, and central vision loss. Cone loss can result in macular atrophy, which is visible on routine fundoscopy imaging. There are many different types of macular degeneration, the most common of which is age-related macular degeneration (AMD).
AMD is the leading cause of vision loss in the ageing population [12]. AMD is a multifactorial condition, and a combination of genetic and environmental factors influence a person’s likelihood of developing AMD. There are two forms of AMD: dry and wet. Dry AMD is the most common, characterized by geographic atrophy that results from the degeneration of RPE, photoreceptors, and choroid vessels. Dry AMD can progress to wet AMD, which is characterized by choroidal neovascularization. The aberrant vessels are fragile and prone to leaking or bursting, which can lead to fluid buildup and retinal scarring. Lipid deposits, called drusen and subretinal drusenoid deposits, are hallmark features of AMD. Drusen develop between the RPE and underlying choroid vasculature, while subretinal drusenoid deposits occur between the photoreceptor OSs and RPE. Although characteristic features of AMD, the impact that drusen and subretinal drusenoid deposits have on the retina and photoreceptor biology is unclear.

2.7. Achromatopsia

Achromatopsia is a condition characterized by lack of colour vision. Achromatopsia can either be complete or incomplete; individuals with complete achromatopisa have no functioning cone photoreceptors, while those with incomplete achromatopsia have partial cone function. Achromatopsia is generally caused by mutations in cone-specific phototransduction machinery or ion channels that prevent the cones from relaying messages to downstream neurons [1,31].

3. Zebrafish as a Model for Photoreceptor Disease: Photoreceptor Features, Function and Morphological Assessment, and Model Generation

Zebrafish are a powerful model for photoreceptor development and disease. The zebrafish retina is remarkably conserved to the mammalian retina in terms of development, structure, and function. Researchers can study cone photoreceptor development and disease in zebrafish, which is challenging in nocturnal rodent models. There are many tools for assessing zebrafish visual responses, photoreceptor function, and morphology in live animals. Additionally, there are well-established technologies for generating transgenic and mutant animals. These features are outlined below. For reviews that compare retinal features of different animal species used to model disease, see [32,33].

3.1. The Zebrafish Visual System

Zebrafish have a visual system that develops rapidly and larval zebrafish can be assessed for visual function at early ages. Zebrafish embryos develop externally and embryos are transparent, meaning that visual system development can be easily observed. Eye field patterning begins at 28 h post fertilization (hpf), and the developing retina expresses opsin genes as early as 51 hpf [34,35]. At 5 days post fertilization (dpf), zebrafish can carry out complex visually-mediated behaviours such as prey capture [36], meaning that larvae can be assessed for visual defects very early in life.
The zebrafish retina is cone-rich, similar to the human macula. This is a huge benefit since other commonly used models, such as mice and rats, are nocturnal with few cones and no macula-like region. The abundance of cone photoreceptors allows for modeling of cone diseases in zebrafish. However, there are nevertheless notable differences between zebrafish and humans. Zebrafish have double cones, which are not observed in the mammalian retina. Additionally, the zebrafish photoreceptor mosaic is organized differently than the human mosaic, and while it can model general cone dystrophies, cannot accurately recapitulate macular degeneration. Although zebrafish have uneven distributions of cones at some stages, their retina does not have a macular region with increased cones and a central cone-only foveal pit [37]. It is important to note that although zebrafish do not have a retinal region that is structurally similar to the fovea, larvae do have a region that is important for high-acuity vision and is therefore functionally similar to the human fovea [38].
The zebrafish photoreceptor mosaic is differentially organized between larval and adult stages. Zebrafish have four cone subtypes: ultraviolet (sws1, homolog of mammalian OPN1SW), blue (sws2), green (rh2), and red (lws, homolog of mammalian OPN1MW and OPN1LW). The green and red cones exist as physically fused double cones. Larval zebrafish have a cone-dominant retina with few rods [37,39]. As the larval zebrafish age, they develop more rods, and in adulthood there are roughly equal proportions of cones and rods [37]. Adult zebrafish photoreceptors are organized into a highly structured a row mosaic, with UV and blue cones alternating in their rows, and red and green cones alternating in neighbouring rows [37,40] (Figure 2A,B). Some opsins have multiple copies that are differentially expressed according to retinal location and age. For example, zebrafish have four green opsin genes (rh2-1, rh2-2, rh2-3, and rh2-4), two red opsin genes (lws1 and lws2), and two rhodopsin genes (rho and rhol), which have different spatial localizations depending on developmental stage [41,42,43]. There are many tools available for visualizing zebrafish photoreceptors, including transgenic lines expressing fluorescent proteins in specific cell types and antibodies for immunolabelling (Figure 2B,C). A subset of commonly used transgenic lines (Table 1) and antibodies (Table 2) are highlighted here; available zebrafish tools are catalogued on the Zebrafish Information Network (ZFIN) [44]. The well-characterized promoters in these transgenes can also be utilized to drive expression of other genes, such as disease or cell death genes.
Table 1. Examples of common zebrafish transgenic lines with labelled photoreceptors.
Table 1. Examples of common zebrafish transgenic lines with labelled photoreceptors.
TransgeneDescriptionReference
Tg(3.2gnat2:eGFP)Enhanced GFP expressed in all cone photoreceptors[45]
Tg(rho:eGFP)Enhanced GFP expressed in rod photoreceptors[46]
Tg(sws1:GFP)GFP expressed in UV cones[47]
Tg(sws2:mCherry)mCherry expressed in blue cones[48]
Table 2. Examples of commonly used antibodies that label zebrafish photoreceptors.
Table 2. Examples of commonly used antibodies that label zebrafish photoreceptors.
AntibodyAntigenLabelsReference
4c12N/ARods[49]
10c9.1UV opsinUV cone OSs[50]
1D4Red opsinRed cone OSs[51]
zpr1Arrestin3aRed/green double cones[52,53]
zpr3N/ARod and double cone OSs[54]

3.2. Visually-Mediated Responses as Diagnostics in Zebrafish

Visual function can be investigated in zebrafish larvae by several behavioural tests. Visually-mediated responses in zebrafish larvae include the optokinetic response (OKR), optomotor response (OMR), and visual motor response (VMR). OKR tracks the eye movements of the larvae in response to visual stimuli. OKR involves immobilizing a larvae and placing the animal inside of a drum with moving black and white stripes [55,56,57,58]. As the drum moves, the zebrafish’s eyes will track the movement then saccade back to the origin [55,56,57,58]. OKR can be detected as early as 3 dpf [57,58]. OMR uses a visual stimulus to elicit a swim response. Larvae are placed inside narrow tracks on top of a flat screen [58,59,60], a stimulus of moving lines plays on the screen, and the fish swim in the direction of the moving stimulus [58,59]. OMR is detectable as early as 4 dpf, but is more robust at 6–7 dpf [59]. VMR is a visually-mediated startle response triggered by rapid changes in light. Individual larvae are placed into wells of a 96-well plate and then the plate is inserted into a recording chamber with lights that can be turned on and off [61]. Larvae will become active when the lights come on, followed by returned to baseline; similarly, when the lights are turned off, there is an increase in activity followed by a return to baseline. VMR can be elicited from larvae as early as 4 dpf.

3.3. Retinal Imaging and Functional Assessement for Zebrafish

Zebrafish have many tools available for the assessment of photoreceptor health in live animals. These techniques include optical coherence tomography (OCT), ERG, calcium imaging, and fundoscopy.
OCT is a live imaging technique that uses the differential reflectivity of the various retinal cell types to capture images of the retinal layers and photoreceptor mosaic. Photoreceptor ISs are highly reflective due to the abundance of densely packed mitochondria, thereby facilitating visualization of the photoreceptor layer [62]. Refractive properties of the zebrafish lens cause zebrafish OCT to be very high resolution, and individual photoreceptor ISs can be observed [63,64]. En face projections provide a view of the photoreceptor mosaic.
Electroretinography is an electrophysiological method for assessing retinal function. ERG involves placing an electrode on the subject and using light stimuli to elicit a response. The testing conditions and light stimuli can be modified to assess specific photoreceptor types. ERG has been utilized to assess both larval and adult zebrafish [65,66,67,68,69]. Calcium imaging techniques can also be used to assess photoreceptor activity in transgenic zebrafish larvae expressing a Ca2+ sensor fused to a fluorescent protein [70,71]. When Ca2+ is present in the cell, such as during neuronal activity, the protein undergoes a conformational change that increases fluorescence [72,73]. This is a useful tool to assess whether photoreceptor cells are responding to stimuli, although it cannot easily be used in live adult zebrafish due to eye size and significant pigmentation.
Fundoscopy involves shining a light into the eye to observe the back of the eye, termed the fundus. Zebrafish fundoscopy enables visualization of adult zebrafish retinal cells in transgenic animals expressing fluorescent proteins [48]. This can be used to observe progressive photoreceptor degeneration and regeneration.

3.4. Generation of Transgenics and Mutants

External fertilization and large clutch size mean that zebrafish embryos are very accessible. This makes generation of transgenic or mutant animals straightforward, as constructs can be injected directly into the single-cell embryo [74]. The generation of transgenic zebrafish can be accomplished using transposase and Tol2 transposable element sites or I-SceI meganuclease [75,76,77]. However, some transgene sequences are susceptible to silencing in zebrafish. The Gal4/UAS transcriptional activation system is a gene/enhancer trap system; the zebrafish-adapted version is KalTA4/UAS [78]. This involves the tissue-specific activation of the KalTA4 transcription factor, which then recognizes the UAS promoter and drives expression of the downstream transgene—for example, a fluorescent protein. Unfortunately, the UAS promoter becomes silenced throughout subsequent generations, and silencing can even be observed in individual fish; larvae may have high levels of expression, but expression can decrease drastically in adulthood due to silencing [79,80]. However, silencing seems to be a relatively uncommon occurrence when transgenes are generated using endogenous zebrafish promoter sequences. Promoter silencing has also created challenges for generation of conditional knockout lines in zebrafish. Cre-lox recombination utilizes Cre recombinase and lox target sites to create conditional loss-of-function alleles. Cre can be expressed ubiquitously or under tissue-specific promoters. Cre-lox has been successfully utilized in zebrafish and there are increasing numbers of tools and zebrafish Cre-driver lines [81,82,83,84,85,86]. However, it has been difficult to find tissue-specific promoters that are not silenced over time, as well as issues activating inducible Cre in zebrafish. In addition, it is challenging to insert lox sites around target genes. As a result, conditional knockout zebrafish are difficult to generate and rare.
There are many means by which mutations have been introduced to zebrafish, including in forward genetic screens using mutagens such as ethylnitrosourea, or targeted gene mutations with zinc finger nucleases (ZFNs), transcription activator-like effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR)/Cas [87,88,89,90]. However, mutants isolated from forward genetics screens require extensive mapping in order to locate the affected gene(s), and some lines never had a mutated gene identified. ZFN and TALEN editing strategies require fusing a nuclease to a sequence-specific DNA binding domain to target genomic regions; this process can therefore be strenuous, but resulting mutants are unlikely to have off-target mutations introduced. CRISPR/Cas systems are easily adaptable to different target genes, but off-target cutting is difficult to predict, though there are methodologies that help limit off-target mutagenesis [91,92,93,94]. There are nevertheless challenges associated with the zebrafish genome. The teleost lineage underwent a genome duplication event, meaning that many genes have two copies. This creates complexities when making mutants, and functional redundancy between the paralogs can necessitate generation of double mutants. Conversely, there can be sub-specialization of the paralogs. It is not uncommon for one gene to be specific to cones while the other is either specific to rods, have pan-photoreceptor expression, or be expressed elsewhere in the retina or nervous system entirely [95,96,97,98].

3.5. Regenerative Response

The mature human retina has very limited capacity to replace lost neurons, meaning that loss of any retinal neurons is irreversible and can lead to vision impairment. Conversely, zebrafish retinas have a robust regenerative capacity [99]. This is both a benefit and challenge, as it permits investigation into mechanisms of neuronal regeneration, but also means that the zebrafish retina responds differently to acute retinal lesions than the human system. Assessment of disease progression can be complicated by regeneration of lost photoreceptors; there may be cases where degeneration is underestimated due to regenerated photoreceptors restoring visual function. It is important for researchers to be aware of how regenerative responses may affect their assessment of retinal damage longitudinally. However, factors necessary for regenerative responses can be pharmacologically blocked or knocked down to allow for disease progression more comparable to humans [100,101,102], although blocking regeneration may cause additional issues with retinal biology. Potential regenerative responses are important to be aware of when investigating models for potential degeneration. Mechanisms underlying the zebrafish retinal regenerative response will be discussed further in Section 5.3.

4. Zebrafish Models of Photoreceptor Disease

There are many zebrafish models of photoreceptor disease, including models where degeneration is induced by exposure to light or toxins, transgenic lines expressing deleterious components, and mutants, which are reviewed here.

4.1. Light Ablation

Light ablation allows for photoreceptor ablation while the remaining neural retina stays relatively intact [99]. There are also differences in susceptibility of photoreceptor types depending on retina location, potentially due to the amount of light that reaches specific areas [103,104,105,106,107]. Light ablation typically causes generalized destruction of the photoreceptors, though targeted laser ablation can be performed to target specific photoreceptor types [108]. A robust regenerative response occurs post light ablation [109].

4.2. Toxic Lesions

4.2.1. Ouabain

Retinal lesion can be achieved with injection of ouabain, which inhibits Na+/K+-ATPase and thereby reduces available ATP for cellular processes [110]. Dosage determines the retinal layers that are affected. High doses can penetrate far into the retina, damaging many retinal cell types including photoreceptors, while lower doses only reach inner retinal layers [110,111,112]. This method cannot easily be limited to only one cell type, although ouabain could be injected subretinally to cause more targeted photoreceptor and RPE damage. In zebrafish, ouabain has been used successfully to ablate retinal neurons, including photoreceptors, and the resultant damage stimulates a regenerative response [111].

4.2.2. N-Methyl-N-nitrosourea

N-methyl-N-nitrosourea (MNU) is an alkylating agent that transfers its methyl group to nucleic acids, resulting in DNA mutations and damage. It is a carcinogen, typically used to induce tumours in cancer models [113,114]. MNU treatment in animals induces photoreceptor cell death, presumably through DNA damage [115,116,117]. MNU treatment primarily affects the rod photoreceptors and therefore produces rod degeneration akin to RP [115]. The acute loss of rods results in a regenerative response in zebrafish [115]. Why MNU treatment is more damaging to rods than cones is unknown.

4.3. Hypoxia-Induced Neovascularization

Exposing adult zebrafish to a hypoxic environment results in retinal neovascularization [118]. Angiogenic sprouts being to grow around day 2 and continue to grow throughout hypoxia treatment. Similar to wet AMD neovascularization, the aberrant vessel growth can be minimalized with anti-VEGF treatment. How the neovascularization impacts retina function in adult zebrafish is poorly investigated.

4.4. Gene Knockdown (Morpholinos)

Morpholinos are antisense oligonucleotide analogs, usually 20–25 bases in length, which complement and bind mRNA to prevent the production of protein products. Morpholinos can either block mRNA splicing or translation [119,120]. They can be easily introduced to zebrafish embryos by microinjection at the single-cell stage [119]. However, morpholinos can have off-target effects and morpholino experiments need to be carefully controlled. It has been reported that morpholino phenotypes are often not recapitulated in genetic mutants [121]; the reason for this is unclear, but may be due lack of genetic compensation. When there are genomic mutations, there is genetic compensation in response to the genetic lesions. However, this compensation does not occur in a gene knockdown scenario [122]. Despite their limitations, morpholinos provide the opportunity for preliminary assessment of potential gene function when properly controlled, which is extremely valuable for expeditious assessment of putative disease-causing genes.

4.5. Transgenic Models

4.5.1. XOPS:mCFP

The Tg(XOPS:mCFP) transgenic zebrafish line expresses membrane-targeted cyan fluorescent protein (mCFP) under the Xenopus rhodopsin promoter (XOPS) [123] (Table 3). The mCFP is overexpressed, resulting in toxicity and rod degeneration. Interestingly, there is no sign of cone death, despite rod loss. This differs from what is observed in humans with RP, where cone degeneration follows rod death. The reason for cone persistence despite rod degeneration in the Tg(XOPS:mCFP) zebrafish is unclear, though could be due to the cone-rich environment of the zebrafish retina, compared to the rod-rich human retina. The phenomenon of cone survival in zebrafish provides opportunities to investigate how cone photoreceptors respond to rod loss, and may provide insight into factors contributing to cone death in human RP.

4.5.2. Nitroreductase Ablation

Nitroreductase (NTR, nfsB gene) is a bacterial enzyme that can reduce otherwise inert prodrugs, such as metronidazole (MTZ) into DNA cross-linking agents, allowing for the targeted ablation of cells expressing NTR [124]. Importantly, there appears to be no toxic bystander effects as the result of NTR ablation, and neighbouring (non-NTR expressing) cells are not damaged by the treatment [60,124,125,126]. The NTR mechanism of ablation is thus one of the few methods that allows for ablation of targeted cell subtypes. Additionally, as NTR-fluorescent protein fusions have been made, the degeneration of target cells can be followed, and, where possible, regeneration of these cells can also be observed. It is also possible to delay, or prevent, the regeneration of the target cell type by continued treatment with MTZ, since regenerating cells will undergo apoptosis once they begin to express NTR.
The NTR method has been used to ablate UV, blue, and red cones as well as rods in zebrafish [60,125,126,127,128]. Intriguingly, ablation of specific photoreceptor subtypes has revealed thresholds for triggering regenerative responses. In larval zebrafish, NTR-mediated blue cone ablation did not stimulate appreciable amounts of regeneration [127]. However, simultaneously ablating both blue and red cones using the NTR mechanism successfully induced robust regeneration. Conversely, ablating solely UV or red cones successfully stimulated a regenerative response [125,127,129]. Regenerated cones can be of multiple subtypes, and are not restricted to the ablated cone subtype [129]. How NTR ablation of specific cone subtypes triggers regenerative responses in the adult zebrafish retina has not been investigated. The ablation of rods in the adult zebrafish retina demonstrated that there is similarly a threshold of damage that needs to be reached to trigger regenerative responses. Ablating the vast majority of rod photoreceptors stimulated a regenerative response, but incomplete ablation of rods did not [126].
NTR ablation of RPE resulted in RPE degeneration as well as subsequent photoreceptor degeneration [130]. RPE loss by this method triggered a regenerative response, and animals were able to re-establish an RPE monolayer. While regenerative capacity for retinal neurons in zebrafish is well established, this was the first study to show RPE regeneration post degeneration.

4.5.3. Disease Genes

Transgenesis technology allows for insertion of a disease-causing gene from humans or other species into the zebrafish genome. This can be used to generate models of photoreceptor disease, as well as investigate how the mutated protein behaves within and impacts photoreceptor cells.
There are several transgenic zebrafish models of RP-like photoreceptor disease (Table 3). Zebrafish with transgenically introduced mouse rhodopsin with a common RP mutation [131,132,133], Tg(rho:msRho-P23H-flag), have notable rod loss by early adulthood [134]. Similarly, a line expressing human rhodopsin with the RP-causing Q344X mutation in rod photoreceptors, Tg(rho:hsaRHO-Q344X), has rapid rod degeneration [135,136]. Adcy is an enzyme normally found in the IS that is related to mechanisms of photoreceptor cell death; antagonists of Adcy increase photoreceptor survival in degeneration models [136]. An RP-like transgenic line that expresses Adcy2b with a C-terminal rhodopsin tail, targeting the protein to the OS, undergoes rod degeneration [136]. PRPF31 is involved in splicing of pre-mRNA and has been associated with RP [137]. Tg(rho:prpf3-AD5-mCherry) zebrafish express mutant prpf31 in rods and have abnormal splicing for a subset of important photoreceptor genes and increased cell death in the photoreceptor layer [138]. A single nucleotide polymorphism in the HTRA1 promoter region increases an individual’s likelihood of developing AMD [139,140]. Overexpression of HTRA1 in mouse RPE has also been reported to induce choroidal neovascularization, similar to what is observed in AMD [141]. Zebrafish larvae transgenically expressing human HTRA1 in rod photoreceptors have rod cell death [142]. Of interest, the zebrafish model of RP expressing human mutant rhodopsin have increased htra1a expression [142].
There is a transgenic cone-rod dystrophy model with mutant human GUCY2D, which produces the protein retinal guanylate cyclase-1 (RETGC1) [143] (Table 3). Cones degenerate in larval zebrafish, followed by rod degeneration. In early adulthood, rods look dysmorphic, but whether rods continue to degenerate in aged adult retina has not been reported.
Table 3. Transgenic zebrafish models of photoreceptor disease.
Table 3. Transgenic zebrafish models of photoreceptor disease.
Photoreceptor DiseaseTransgenePhotoreceptor FeaturesReference
Cone-rod dystrophyTg(3.2gnat2:hsa.GUCY2D-E837D R838S)Dysmorphic cones at 5 dpf. 3-month-old animals have a thin photoreceptor layer in the central retina, dysmorphic cones, and less cone and rod staining[143]
Retinitis pigmentosaTg(rho:hsaRHO-Q344X)Rod degeneration detectable at 5 dpf, rapid progression.[136]
Retinitis pigmentosaTg(rho:msRho-P23H-flag)Rod degeneration at 3 months.[134]
Retinitis pigmentosaTg(rho:prpf3-AD5-mCherry)Increased photoreceptor cell death at 5 months.[138]
Retinitis pigmentosaTg(rho:adcy2b-rho-tail)Rod degeneration noted at 14 dpf.[136]
Retinitis pigmentosaTg(rho:hsaHTRA1)Rod degeneration at 5 dpf.[142]
Retinitis pigmentosaTg(XOPS:mCFP)Rapid rod degeneration by 5 dpf. Adults do not have a rod response on ERG.[123]

4.6. Genetic Mutants

Genetic screens and genome editing strategies have fostered the discovery and generation of many zebrafish mutants with photoreceptor phenotypes. Of note, there are instances where the disease phenotype in zebrafish differs from what is observed in human patients with mutations in the same gene. This phenomenon may be due to disparities in the type of genetic lesions present. Loss-of-function mutations are most commonly engineered in animal models, but disease-causing mutations observed in humans may instead be missense, splice site mutations, or compound. Alternatively, these divergences could be related to gene duplication and subspecialization/neospecialization in zebrafish, or differences between the human and zebrafish photoreceptor mosaic.
Mutations in genes affecting cilia or other proteins expressed in many tissue and cell types may have extraretinal phenotypes and early lethality. Only the photoreceptor phenotypes will be focused on here. For previous reviews on zebrafish models of ocular disease, please see [144,145,146,147,148].
Please note that the zebrafish mutants are categorized based on the condition that their phenotype is most similar to, and this may not necessarily align with the phenotypes observed in human patients with mutations in the same gene. In order to be included here, the zebrafish mutant lines had to have an identified mutated gene and present with photoreceptor degeneration, photoreceptor dysfunction, or otherwise fit into a photoreceptor condition category based on retinal findings.

4.6.1. Retinitis Pigmentosa

Mutations in the rod-specific opsin gene, rhodopsin (RHO), are a frequent cause of RP [131,132,133,149]. Recently, several rho mutant zebrafish lines have been generated which model dominant or recessive RP [150,151] (Table 4; see for list of zebrafish mutant RP models). These animals have rod degeneration beginning soon after rod development and continuing into adulthood [150,151]. Cones are unaffected in rho mutants.
RP2 is a cilium-associated protein and its mutation causes X-linked RP [152,153,154,155]. Knockout of rp2 in zebrafish results in photoreceptor functional defects in larval animals, as well as progressive rod OS degeneration, followed by cone OS degeneration [156]. Mutations in CERKL, a gene involved in metabolism of components important for neuron survival, can cause RP [157]. Zebrafish cerkl knockout larvae have photoreceptor defects detected by electrophysiological assessment [158]. Young adults have OS defects, primarily affecting the rods, which progress to rod degeneration and eventual cone OS defects. KIF3B encodes a kinesin motor involved in transport through the cilium and mutations in KIF3B have recently been associated with a ciliopathy that presents with RP [159]. kif3b mutant zebrafish have rapid rod degeneration and delayed OS genesis, but cones appear normal [160,161]. RPGRIP1 localizes to the CC and OS and mutations in RPGRIP1 cause several different photoreceptor diseases, including LCA, RP, and CORD [162,163,164,165]. Zebrafish rpgrip1 mutants do not have proper rod OS development, and none are observed at 5 dpf [166]. Rod cells were almost entirely absent by 3 months of age. Cones appear normal at 7 dpf, but have functional defects and cone degeneration is clear by 6 months. By 23 months, almost all photoreceptors are lost. Interestingly, double cones degenerate before single cones. RP1L1 is a photoreceptor cilium-associated protein, mutations in which lead to RP and macular degeneration [167,168,169]. Mutant rp1l1 zebrafish have rod functional defects at 6 months, which worsen over time [170]. By 12 months, animals have disorganized rod OSs, rod loss, and subretinal drusenoid deposits [170].
MYO7A encodes a myosin protein expressed in human photoreceptors and RPE [171]. MYO7A mutations lead to a subtype of Usher syndrome, a condition characterized by hearing and vision loss [172,173]. myo7aa mutant larvae have decreased photoreceptor function and rod photoreceptor loss [174]. Mutations in USH2A also cause a subtype of Usher syndrome [175]. USH2A encodes the transmembrane protein usherin, the exact function of which is unknown [175,176]. ush2a mutant zebrafish have photoreceptor functional defects in larvae, and progressive photoreceptor degeneration in adulthood [177,178]. Rods begin to degenerate prior to cone degeneration [177].
her9 is a transcriptional regulator that has increased expression in Tg(XOPS:mCFP) retinas undergoing rod degeneration [179]. her9 mutant larvae have a lower abundance of rods and there is evidence that red cones do not develop appropriately or rapidly degenerate [180]. After rods begin to decrease in number, green cones are reduced in number and appear to have small OSs, while blue and UV cones appear unaffected. Animals do not survive past 13 dpf. her9 orthologs have not been associated with human photoreceptor disease.
Table 4. Zebrafish mutant models of retinitis pigmentosa-like photoreceptor disease.
Table 4. Zebrafish mutant models of retinitis pigmentosa-like photoreceptor disease.
GenePhotoreceptor FeaturesReference
cerklPhotoreceptor functional defects at 7 dpf. Rod OS defects at 3 months, cone OS defects at 7 months. Notable thinning of the photoreceptor layer and cell death by 12 months.[158]
her9Decrease in rod photoreceptors at 5 dpf. Few double cones with short OSs at 12 dpf.[180]
kif3bDelayed OS development. Rapid rod degeneration by 5 dpf. [160,161]
myo7aaDecreased photoreceptor function at 5dpf. Reduced rods at 8 dpf.[174]
rhoRod loss observed at 6 dpf. Degeneration continues into adulthood.[150,151]
rp1l1Rod dysfunction at 6 months. Subretinal drusenoid deposits at 11 months. Photoreceptor loss observed at 12 months.[170]
rp2Photoreceptor functional defects at 7 dpf. Short rod OSs at 2 months; cone OS defects at 4 months; significant rod OS loss and decreased cone OSs by 7 months.[156]
rpgrip1No rod OSs at 5 dpf. Cone dysfunction at 7 dpf. Severe rod degeneration by 3 months, followed by cone degeneration. By 23 months, majority of photoreceptors have degenerated.[166]
ush2aDecreased photoreceptor function at 5–7 dpf and increased photoreceptor apoptosis at 8 dpf. Notable rod OS degeneration at 12 months, cone OS degeneration observed at 20 months.[177,178]

4.6.2. Leber Congenital Amaurosis

Mutations in genes involved in ciliogenesis initiation, cilia elongation, transport of cilium components, or physiological processes can result in LCA or an LCA-like phenotype in animal models. Intraflagellar transport (IFT) proteins play crucial roles in movement of cargo in the cilium, which can be facilitated by kinesin motors. ift88 and ift172 zebrafish mutants have no photoreceptor OS development and the photoreceptors degenerate [181,182,183,184] (see Table 5 for list of zebrafish mutant LCA models). Sensory cilia in general do not develop in these animals [183]. Similarly, cluap1 (also known as ift28) mutant zebrafish have ciliogenesis defects that result in no OS development and rapid photoreceptor degeneration [185]. Fish with mutations in ift57 develop photoreceptor OSs, although they are shorter than normal, and photoreceptors degenerate soon after developing [184,186]. ift122 mutant zebrafish have normal photoreceptor development, but similarly undergo degeneration [187]. Zebrafish kif3a (kinesin family 3a) mutants fail to develop OSs and photoreceptors rapidly degenerate, and larvae have an extinguished ERG [161,188,189]. napbb encodes a protein important for synaptic vesicle fusion [190] and napbb mutant photoreceptors undergo cell death almost immediately after specification [191,192]. No photoreceptor layer is detectable at early larval stages, with only a few photoreceptors observed at the ciliary margin [191,192]. IFT88, CLUAP1, IFT57, IFT122, KIF3A, and NAPB, mutations have not been associated with human photoreceptor disease, likely because of essential roles in development. IFT172 mutations have been associated with RP [193].
KIAA0586 mutations underlie a ciliopathy called Joubert syndrome, which presents with brain abnormalities and photoreceptor degeneration [194]. KIAA0586 is required for ciliogenesis. Developing photoreceptors in kiaa0586 mutant zebrafish have decreased OS development and degenerate rapidly after development [195]. Mutant larvae have significantly reduced ERG responses. TMEM216 is required for cilia assembly and TMEM216 mutations also cause Joubert syndrome in humans [196,197]. Knockout of tmem216 resulted in decreased cone OS development, short OSs, disorganized OS discs, and degeneration [198]. GDF6 encodes a morphogen and its mutation causes LCA and juvenile RP in humans [199]. gdf6a mutant zebrafish larvae are functionally blind at 7 dpf as determined by OMR, and have short, dysmorphic cone OSs and ISs, as well as overgrown, disorganized rod OSs and short ISs [199,200]. Photoreceptors do not undergo degeneration in these animals [199].
Table 5. Zebrafish mutant models of Leber congenital amaurosis-like photoreceptor disease.
Table 5. Zebrafish mutant models of Leber congenital amaurosis-like photoreceptor disease.
GenePhotoreceptor FeaturesReference
cluap1No OS development, rapid photoreceptor degeneration.[185]
gdf6aShort, dysmorphic cones and expanded, disorganized rods at 7 dpf.[199,200]
ift57Short OSs with normal disc stacking at 4 dpf. Central retina photoreceptor degeneration at 5 dpf.[184,186]
ift88No OS development, rapid photoreceptor degeneration.[181,182,183,184]
ift122Normal photoreceptor OS development. Degeneration starting at 7 dpf, severe degeneration by 10 dpf.[187]
ift172No OS development, rapid photoreceptor degeneration.[181,182,184]
kiaa0586Fewer OSs observed at 3 dpf, photoreceptor degeneration observed at 4 dpf. Decreased photoreceptor function detected at 6 dpf.[195]
kif3aNo OS development and rapid rod photoreceptor degeneration by 5 dpf, subsequent cone degeneration. Extinguished photoreceptor response at 7 dpf.[161,188,189]
napbbImmediate, severe photoreceptor degeneration; cell death observed at 3 dpf, no distinct photoreceptor layer at 6 dpf, few photoreceptors in ciliary margin.[191,192]
tmem216Short cilia and disorganized OS discs by 7 dpf, rapid degeneration. Few photoreceptors remaining by 14 dpf.[198]

4.6.3. Choroideremia

Zebrafish rep1 knockout larvae have RPE that lacks uniformity with hypertrophic and de-pigmented regions [24] (Table 6). Additionally, the RPE has large vacuoles and accumulation of undigested OS fragments. The photoreceptor layer is disorganized with dysmorphic, functionally defective photoreceptors. The photoreceptor abnormalities appear to result from RPE dysfunction in these animals, as wild-type photoreceptors transplanted into mutant retina still become dysmorphic and degenerate, while rep1-negative mutant photoreceptors transplanted into wild-type retinas are morphologically normal. rep1 mutant zebrafish do not survive to adulthood.

4.6.4. Cone-Rod Dystrophy

CC2D2A, AHI1, and ARL13B encode basal body or cilia-related proteins and their mutation causes forms of Joubert syndrome [201,202,203]. cc2d2a knockout zebrafish larvae have disorganized photoreceptor OSs and functional defects [204] (see Table 6 for zebrafish CORD mutants). Of interest, cilia assembly occurs normally in mutants, but trafficking of cilia components is impaired. Zebrafish larvae with ahi1 mutations have abnormal cone OSs but normal visual function, as assessed by OKR [205]. Cone OS morphology recovers by early adulthood, but cone morphology defects and degeneration are seen in five-month-old animals. At this age, rod photoreceptors have mislocalization of rhodopsin. Zebrafish arl13b mutant photoreceptors have short OSs [206]. Knockout arl13b larvae do not survive past 9 dpf, so transplantation experiments were performed to assess photoreceptor degeneration. The arl13b mutant photoreceptors had noticeable degeneration by 30 dpf. LCA5 is also a ciliary protein and mutations in the LCA5 gene cause LCA [207]. However, knockout of zebrafish lca5 results in a cone-rod dystrophy phenotype rather than LCA [208]. lca5 mutant larvae have decreased photoreceptor responses. Cone OS abnormalities are obvious at 1 month of age, while rod OS defects are observed at seven months. Mutations in PCARE, a regulator of cilium actin, cause RP [209]. Zebrafish larvae with mutations in pcare1 have abnormal OS morphology, photoreceptor functional defects, and decreased OKR [210]. Adult pcare1 mutants have dysmorphic photoreceptor OSs and thin photoreceptor layers, indicative of degeneration. EYS mutations typically cause RP in human patients, although EYS mutations have also been associated with a cone-rod dystrophy [211,212,213]. The function of EYS is unknown, but a recent study found evidence that EYS localizes to the photoreceptor cilium and may be involved in cilium stability [214]. Zebrafish eys mutants have CORD, whereby cones degenerate in six-month-old animals and rod degeneration is observed at 8–14 months [215,216,217]. BBS2 is a basal body-associated protein necessary for proper ciliogenesis, and mutations in BBS2 lead to RP and a ciliopathy called Bardet-Biedl syndrome [218,219]. Larval bbs2 mutants have photoreceptor deficits observed by OKR and short, disorganized photoreceptor OSs [220]. In adulthood, photoreceptor degeneration is observed, along with regeneration of rods but not cones.
KCNJ13 encodes a channel protein found in the RPE and mutations in KCNJ13 lead to LCA [221]. Adult zebrafish with mutated kcnj13 have photoreceptor loss and RPE abnormalities [222]. Specifically, the RPE has increased phagosomes, enlarged mitochondria, and changes in melanosome localization, while cone photoreceptors have abnormal mitochondria and degenerate in aged animals. POMGNT1 is an enzyme that posttranslationally modifies other proteins, and POMGNT1 mutations cause RP [223]. Zebrafish with pomgnt1 have cone and rod degeneration at six months of age [224].

4.6.5. Enhanced S-cone Syndrome

A similar phenotype to human patients with enhanced S-cone syndrome is observed in nr2e3 null zebrafish. Animals lack rod photoreceptors and undergo progressive cone degeneration [225] (Table 6). Interestingly, green and red cones degenerate while UV and blue cones persist. nrl mutant larvae also do not develop rods, but intriguingly have increased UV cone photoreceptors [226]. Adult nrl mutants unexpectedly have rods, but the rods have abnormal synapses. Whether these rods are functional is unknown.
Table 6. Zebrafish mutants of choroideremia, cone-rod dystrophy, and enhanced S-cone syndrome.
Table 6. Zebrafish mutants of choroideremia, cone-rod dystrophy, and enhanced S-cone syndrome.
Photoreceptor DiseaseGenePhotoreceptor + RPE FeaturesReference
Choroideremiarep1RPE irregularity and pigment loss. Disorganized photoreceptor layers, dysmorphic photoreceptors, and significantly reduced photoreceptor responses at 5 dpf.[24]
Cone-rod dystrophyahi1Disorganized, short OSs, but normal visual function at 5 dpf. Thin photoreceptor nuclear layer, cone degeneration, and dysmorphic rods at 5 months.[205]
Cone-rod dystrophyarl13bShort OSs at 4dpf. Cone degeneration observed at 30 dpf in mosaic animals.[206]
Cone-rod dystrophybbs2Short, disorganized photoreceptor OSs and visual deficits at 5 dpf. Photoreceptor degeneration observed in adulthood.[220]
Cone-rod dystrophycc2d2aDysmorphic photoreceptor OSs and functional defects at 5 dpf.[204]
Cone-rod dystrophyeysProgressive photoreceptor loss; cone degeneration observed at 6 months, rod degeneration observed at 14 months[215,216,217]
Cone-rod dystrophykcnj13Increased phagosomes in the RPE at 3 months, enlarged RPE mitochondria at 6 months, and abnormal melanosome localization under dark adaptation at 12 months. Cone mitochondria abnormalities at 6 months and photoreceptor loss at 12 months.[222]
Cone-rod dystrophylca5Photoreceptor functional defects at 7 dpf. Cone OS defects at 1 month, rod OS defects at 7 months, and progressive photoreceptor loss.[208]
Cone-rod dystrophypcare1Dysmorphic OSs and dysfunctional photoreceptors at 5 dpf. Abnormal OS morphology and thinner photoreceptor layer at 6 months.[210]
Cone-rod dystrophypomgnt1Reduction in cones and rods at 6 months.[224]
Enhanced S-cone syndromenr2e3No rods. Green and red cone degeneration started at 1 month.[225]
Enhanced S-cone syndromenrlRods fail to develop and UV cones are over-abundant in larvae. Adult zebrafish surprisingly have rods, but with synaptic defects.[226]

4.6.6. Cone Dystrophy

PROM1 plays a critical role in outer segment disc morphogenesis and PROM1 mutations cause a variety of photoreceptor diseases, such as RP, CORD, and macular degeneration [227,228,229]. Zebrafish with prom1b mutations have a cone degeneration, starting with red and green cones and affecting blue and UV cones soon after [230] (Table 7; see for list of zebrafish mutant models of cone dystrophy). While rods do not appear to degenerate in prom1b mutants, they have overgrown OSs. CEP290 encodes a centrosomal protein and mutations are the most frequent cause of LCA [17]. Interestingly, cep290 mutant zebrafish have cone OS disorganization and degeneration in adulthood that is severe by 12 months, but normal rod OS morphology [231]. Why cep290 results in an adult onset, cone-specific phenotype in zebrafish is unknown, but may be due to alternative photoreceptor centrosomal proteins in zebrafish.
PDE6C is an essential component of the cone phototransduction pathway and results in cone dystrophy when mutated [232]. pde6c-deficient zebrafish have rapid cone degeneration as well as dysmorphic, degenerate rods in larval stages [49,192]. In adulthood, mutants have severe cone degeneration but normal rod photoreceptors. AIPL1 is required for maintenance of photoreceptor phototransduction machinery and AIPL1 mutations can lead to LCA, RP, and CORD [233,234]. Zebrafish have two copies of this gene, aipl1a, which is expressed in rods and potentially UV cones, and aipl1b, which is expressed in all cones [235]. Zebrafish with mutations in aipl1b have progressive cone degeneration starting in larval stages, and by early adulthood have only a few single cones remaining [235]. Rod photoreceptors are unaffected.
Mutations in the retinal voltage-gated calcium channel subunit gene, CACNA2D4, lead to cone dystrophy [236]. Zebrafish have two paralogs of CACNA2D4: cacna2d4a and cacna2d4b [96]. Knocking out each paralog individually does not impact photoreceptor function in early development, but double knockout animals had reduced cone function [96]. ATP6V0E1 is a subunit of a proton pump and atp6v0e1 mutant zebrafish larvae have RPE pigment loss, outer retinal holes, visual defects determined by OKR, photoreceptor functional deficiencies assessed by ERG, and short OSs [237]. Rods were not assessed in these animals.
WRB is involved in recruitment of proteins to the endoplasmic reticulum, although its exact role in photoreceptors is unclear [238]. wrb mutant zebrafish larvae have holes in their photoreceptor layer, decreased or absent OMR, photoreceptor function deficits, and short cone ribbon synapses in early life [239,240]. Rods have not been investigated in these animals. Mutations in WRB have not been associated with human disease.

4.6.7. Achromatopsia

GNAT2 is an essential component of the cone phototransducation pathway and pathogenic variants in GNAT2 lead to achromatopsia [241]. gnat2 mutant zebrafish have cone dysfunction but the photoreceptors do not have overt morphological defects [242] (see Table 7 for mutant zebrafish achromatopsia models). Electrophysiological assessment suggested that gnat2 mutant cones are ~1000 times less sensitive than wild-type cones. The rods are unaffected in these animals.
Pathogenic mutations in CACNA1F cause CORD and congenital stationary night blindness, a non-progressive condition characterized by rod dysfunction [243,244,245]. CACNA1F protein is a subunit of a voltage-gated calcium channel required at the photoreceptor synapse [246]. Zebrafish cacna1fa is cone-specific and mutant retinas have thin photoreceptor synaptic layers and decreased photoreceptor responses [97]. cacna1fa mutant cones do not develop synaptic ribbons. Similarly, zebrafish synj1 mutants have a thin photoreceptor synaptic layer and abnormal cone pedicles with unanchored synaptic ribbons [247,248,249]. The cone IS and OS appear normal [247,248,249]. Mutant zebrafish do not survive long enough to develop appreciable numbers of functional rods, but a transplantation study demonstrated that synj1 is required for cone synapse formation but not rod synapses [249]. A related phenotype is seen in per2 knockout zebrafish; per2 mutant larvae have an abnormal VMR, untethered ribbon synapses, and decreases in cone opsin expression which may indicate cone degeneration [250]. Animals have not been aged for further investigation of cone and rod viability.

4.6.8. Age-Related Macular Degeneration

AMD models are challenging to generate as AMD is a multifactorial condition. However, there are zebrafish models with features of AMD. Recently, rp1l1 mutant zebrafish were reported to have progressive photoreceptor dysfunction and subretinal drusenoid deposits [170] (Table 4). This is the first report of subretinal drusenoid deposits in zebrafish, and may provide insight into how drusenoid deposits develop and impact photoreceptor function. In addition, models of neovascularization are relevant to wet AMD. Zebrafish larvae with mutations in vhl present with retinal neovascularization, leaky vessels, and develop oedema and retinal detachment [251] (Table 7).
Table 7. Zebrafish mutant models of cone dystrophy, achromatopsia, and retinal neovascularization.
Table 7. Zebrafish mutant models of cone dystrophy, achromatopsia, and retinal neovascularization.
Photoreceptor DiseaseGenePhotoreceptor, RPE, or Vascular FeaturesReference
Achromatopsiacacna1faThin photoreceptor synaptic layer. Photoreceptor functional defects at 5–6 dpf. Cone synapses do not develop synaptic ribbons.[97]
Achromatopsiagnat2Cone functional defects under dim to moderate light intensities.[242]
Achromatopsiaper2Visually-mediated behaviour deficiencies at 5–6 dpf. Unanchored cone ribbon synapses and decreased cone opsin expression.[250]
Achromatopsiasynj1Thin photoreceptor synaptic layer and abnormal, unanchored cone ribbon synapses at 6 dpf.[247,248,249]
Cone dystrophyaipl1Thin photoreceptor layer and OS defects at 7 dpf. Severe cone degeneration by 3 months.[235]
Cone dystrophyatp6v0e1Hypopigmented RPE, reduced or absent visual response, decreased photoreceptor function, short OSs, and outer retinal holes at 5 dpf.[237]
Cone dystrophycacna2d4a + cacna2d4bMild cone dysfunction at 20 dpf.[96]
Cone dystrophycep290Cone OS abnormalities and degeneration at 3 months, severe degeneration at 12 months. Normal rods.[231]
Cone dystrophypde6cCone degeneration at 4 dpf and rod subsequently become dysmorphic and degenerate in the larval retina. At 3 months, cones are degenerate and largely missing, rods are normal.[49,192]
Cone dystrophyprom1bProgressive cone degeneration. Reduced red and green cones at 7 dpf; reduced UV and blue cones at 1 month. Overgrown rod OSs.[230]
Cone dystrophywrbHoles in the photoreceptor layer at 4 dpf. Vision defects at 5 dpf, photoreceptor function defects at 4–5 dpf. Small cone ribbon synapses[239,240]
Retinal neovascularizationvhlRetinal neovascularization, leaky blood vessels, retinal oedema, and retinal detachment at 7–8 dpf.[251]

4.7. Relevance to Human Disease

In order to understand and prevent photoreceptor disease, a keen knowledge of how photoreceptor development, function, and maintenance is required. Zebrafish have provided additional information about factors involved in these aspects, as lines with mutations in genes not associated with human photoreceptor disease, potentially due to lethality, have impaired or degenerate photoreceptors. Additionally, the cellular abnormalities and cause of photoreceptor death can be investigated in these models. For example, genetically mosaic zebrafish generated by transplantation experiments with zebrafish rep1 mutants determined that rep1 loss in photoreceptors does not lead to photoreceptor degeneration, but rather that rep1 is required for normal RPE function, and photoreceptors die due to RPE dysfunction in rep1 mutant retina [24].
Models of cone disease are desperately needed, as preservation of the high-acuity central vision is a priority for patients and physicians. Nocturnal rodent models are not an ideal system for these conditions, as they have a low density of cones and their cones have some properties that are distinct from human cones, such as expression of multiple opsins in a single cell. Indeed, some mouse models of cone diseases do not have the expected cone degeneration phenotype [252]. The zebrafish retina allows for the investigation of cone diseases, and the numerous genetic models provide the opportunity to determine disease mechanisms based on the mutated gene.
In addition, the cone-rich zebrafish retina enables the assessment of how rod loss impacts cone photoreceptors. Cone persistence post rod death in adult zebrafish RP models permits determination of how cones respond to rod loss and the factors that influence their health when neighbouring photoreceptors degenerate. Additionally, while cones appear viable post rod loss in RP zebrafish models, pde6c mutant larvae have rod degeneration post cone loss despite the cone-specific genetic lesion [49,192]. This suggests that the minority photoreceptor type is inherently susceptible to death, as the cone-dominant larval zebrafish retina undergoes rod degeneration post cone loss while the rod-dominant human retina experiences cone degeneration post rod loss. The pde6c mutants can therefore be utilized to determine why the minority photoreceptor type is sensitive to degeneration and provide avenues to preserve the cells, which is relevant to efforts aimed at protecting cones in late-stage RP.
Interestingly, several zebrafish models of photoreceptor disease have red and green cone degeneration prior to UV and blue cone degeneration [166,180,225,230,235]. Whether this is due to structural distinctions between double and single cones, or whether there are specific features of longer-wavelength sensitive cones that make them more susceptible to cell death than short wavelength cones, is unknown. Further assessment is required and may inform macular degeneration pathologies. The central macula is comprised of primarily red and green cones, and understanding what makes these cones vulnerable to disease is important for preventing AMD.
Drusen and subretinal drusenoid deposits are characteristic features of AMD but how they develop and their effect on photoreceptor health is unknown. Animal models of drusen and drusenoid deposits are difficult to develop and there a few photoreceptor disease models that acquire them. Subretinal drusenoid deposits in particular are poorly investigated, as they are relatively newly characterized. A zebrafish model of photoreceptor disease with accompanying subretinal drusenoid deposits provides the opportunity to investigate how these lipid accumulations develop in vivo [170].

5. Contribution to Therapeutics

Zebrafish provide a model to test therapeutic strategies and pharmaceuticals. Numerous disease models combined with the small size of zebrafish larvae and visual tools, such as VMR, OMR, and OKR, allow for screening of compounds. Below, the contributions from zebrafish models are outlined.

5.1. Pharmaceuticals

5.1.1. Anti-Angiogenic Compounds

Zebrafish models of retinal neovascularization, such as the hypoxia-induced model and vhl mutants, provide the opportunity to screen for anti-angiogenic compounds that could be employed in patients with wet AMD. Indeed, Van Rooijen and colleagues found that treating vhl mutants with vascular endothelial growth factor receptor (VEGFR) tyrosine kinase inhibitors typically used in cancer treatment successfully blocked pathogenic angiogenesis [251], which could be further tested in mammalian models of retinal neovascularization.

5.1.2. Histone Deacetylase Inhibition

Cellular stress markers can activate cell death pathways, resulting in apoptosis, or trigger elimination by immune cells. This can result in loss of cells that were stressed, but otherwise functional, and has been observed in the retina [253]. Overactivation of histone deacetylases (HDACs) has been detected in models of photoreceptor degeneration [254]. Indeed, HDAC inhibitors have been effective for treating neuronal pathologies in vitro and in animal models of CNS disease, including photoreceptor degeneration models [255,256,257]. However, it is important to note that there are also HDACs that inhibit photoreceptor death after light injury [258]; it is therefore important to target specific HDACs.
HDAC6 inhibition has been reported to prevent neurodegeneration in peripheral neuropathy models [259]. In the atp6v0e mutant zebrafish model of cone photoreceptor disease, HDAC6 inhibitors successfully restored OKR, decreased the number of apoptotic cells in the photoreceptor layer, and increased OS persistence [237,260,261]. Similar results are observed in a mouse model of photoreceptor disease [260,261], further supporting HDAC6 as a potential therapeutic target for treating photoreceptor degenerations.

5.1.3. Novel Compound Screening

Zebrafish models of photoreceptor disease provide a platform for discovery of novel neuroprotective drugs. Zebrafish have been employed for numerous large drug screens, such as screens to identify compounds that impact neuronal activity [262,263]. Zebrafish can therefore be utilized to identify novel compounds in large-scale screens. Through behavioural screens and histological approaches, several compounds have been identified as protective against photoreceptor dysfunction and degeneration in zebrafish.
Ganzen and colleagues utilized Tg(rho:hsaRHO-Q344X) and transgenics with NTR-ablatable rods to screen a library of potential drugs for potential RP treatment [264] (bioRxiv preprint). VMR assessment identified carvedilol as beneficial, and carvedilol-treated Tg(rho:hsaRHO-Q344X) larvae had increased rod survival compared to control animals. As RP is the most common inherited photoreceptor disease, a pharmacological treatment may have applicability to many patients experiencing progression vision loss.
Cone degeneration in pde6c mutant zebrafish larvae occurs through nectroptosis, while the subsequent rod degeneration is apoptosis-mediated [265]. Blocking necroptosis factors using necrostatin-1s or necrosulfonamide at 10 hpf until 7 dpf results in a reduction in cone death [265]. However, whether vision is preserved due to necrostatin-1s or necrosulfonamide treatment is unknown. Further investigation is required, but pharmacological repression of necroptosis has applicability to many different photoreceptor diseases that involve necroptosis pathways.
Studies using zebrafish models of photoreceptor disease have also identified schisandrin B and gypenoside, plant-derived compounds with antioxidant and anti-inflammatory properties, as potential neuroprotective compounds [266,267,268,269,270,271]. These examples highlight the benefits of utilizing zebrafish models of photoreceptor disease to screen novel therapeutics and specific drugs.

5.1.4. Oculotoxicity

Many drugs have retinal toxicity, meaning that prolonged exposure can result in retinal damage and vision loss. Zebrafish have provided a relatively high-throughput tool to assess retinal toxicity of pharmaceuticals. Deeti and colleagues used zebrafish larvae to develop a methodology for assessing the impact of pharmaceuticals on the zebrafish visual system [272]. Thereby, 3 dpf, zebrafish larvae were treated with compounds for 2 days in 48-well plates, and then assessed for visual impairment using VMR and OKR. 5/6 known oculotoxic drugs reduced the visual responses of zebrafish. However, all of the oculotoxic drugs resulted in a reduction in zebrafish touch response, suggesting defects in other sensory neurons. Therefore, zebrafish provide a viable tool for screening drugs for retinal toxicity.

5.2. Antisense Oligonucleotide Therapy

Antisense oligonucleotide therapy can block aberrant splicing in genes with mutations in splice donor or acceptor sites. A frequent Usher syndrome-causing USH2A mutation results in incorporation of a psuedoexon and truncation of the USH2A protein [273]. Morpholino treatment successfully prevented aberrant splicing in mosaic zebrafish with ush2a containing a CRISPR/Cas-introduced region of human mutant USH2A [274]. In addition, this group recently investigated morpholino-induced skipping of a mutated exon in ush2a mutant zebrafish [275] (preprint); specifically, they skipped mutant exon 13, which is commonly mutated in Usher syndrome [178,276]. Exon skipping was successful and partially restored protein expression [275] (preprint). This therapeutic approach of using antisense oligonucleotides to partially restore USH2A function by skipping the problematic exon 13 is currently in clinical trial (ID: NCT03780257).

5.3. Photoreceptor Regenerative Pathways

Zebrafish can regenerate lost retinal neurons, including photoreceptors [277]. Investigating the zebrafish regenerative response provides information on activation of endogenous retinal stem cells as well as factors that promote photoreceptor differentiation and integration of newly generated cells, which is invaluable for effective development and deployment of stem cell therapies.
Zebrafish have two distinct stem cell pools: the ciliary marginal zone (CMZ) and Müller glia.

5.3.1. Ciliary Marginal Zone

The CMZ is a pool of retinal stem cells located at the peripheral edges of the retina. In many species, this pool becomes inactive once the retina has matured. However, in many vertebrates, such as zebrafish and other species that undergo indeterminate growth, the CMZ self-renews and constantly adds new retinal neurons at the edges of the retina. The CMZ contains both retinal stem cells, which can divide indeterminately, and retinal progenitor cells, which can only divide a certain number of times and lose contact with the CMZ as differentiation occurs [104,277,278,279]. Retinal stem cells sit in the edges of the CMZ, and typically undergo asymmetrical division to produce a retinal stem cell and a retina progenitor cell [104,277,278,279].
Mammals have a CMZ-like region that gives rise to some retinal neurons during development [280], although whether it has the capacity to produce photoreceptors is unclear. In the mature retina, these retinal margin cells seem to lose neurogenesis capacity or become quiescent. However, there are cells with stem cell-like properties in the mammalian/human peripheral retinal margin net to the pars plana; specifically, Müller glia and ciliary epithelium cells that express neural progenitor markers [281,282,283]. Whether these cell populations could be stimulated to enter a stem cell-like fate in the damaged retina requires further investigation.

5.3.2. Müller Glia

Müller glia are support cells that play many essential roles in the retina, including ion regulation, neurotransmitter uptake and degradation, cellular debris removal, and neuronal insulation. In zebrafish and other teleost fishes, Müller glia also act as endogenous stem cells by undergoing de-differentiation, proliferation, and specification to produce many different retinal neuron types, including photoreceptors [281,282,283]. Zebrafish Müller glia division is asymmetrical, allowing for replenishment of lost retinal neurons while maintaining the stem cell pool [99,104,107,279,284,285,286,287]. As mentioned above, there is a threshold of damage that must be met to trigger the Müller glial regenerative response, which has been observed in NTR ablation studies [60,125,126,127,128,129]. Interestingly, some mutant zebrafish have cone degeneration but Müller glia do not undergo proliferation to regenerate the lost photoreceptors [220,231]. Further investigation into one of these lines, the bbs2 mutants, found that the endogenous Müller glia do have the capacity to regenerate retinal neurons in bbs2 mutant retina, as evident by light ablation experiments [220]. Why photoreceptor degeneration triggers a regenerative response in some mutant models but not others is unclear, but may be related to the speed of degeneration or an inflammatory requirement for Müller glia activation.
Zebrafish Müller glial behaviour must be carefully controlled in order to prevent aberrant proliferation and production of unneeded neurons. In the uninjured retina, the reprograming pathways of Müller glia are inhibited [288]. After retinal injury, Müller glia respond to factors released by damaged cells—including TNFα [289]—and undergo reprogramming, followed by proliferation and neurogenesis. Studying the regenerative pathways in zebrafish has provided insight into stem cell and fate-determining factors that can be utilized to generate photoreceptor precursor cells for transplantation.
As Müller glia are interspersed within the human retina, how the stem cell behaviour of Müller glia is activated and whether mature mammalian Müller glia can be re-programmed to a stem cell fate to treat retinal degenerations is an area of active interest. Human Müller glia respond to damage and have stem cell-like characteristics, but do not naturally undergo neurogenesis in vivo [281,290]. Müller glia proliferation is rare in the mammalian retina, usually associated with scar formation [291]. In culture, human Müller glia treated with growth and differentiation factors can take on stem cell characteristics and differentiate into rod photoreceptors [281,290,292]. Rod photoreceptors derived from human Müller glia cultures have successfully been introduced into a rat model of RP. These cells are able to migrate into the ONL and restore some visual function [293]. Further investigations into Müller glial response to photoreceptor loss, photoreceptor cell differentiation, and integration will be essential for application of stem cell therapies in human disease, and zebrafish provide this unique avenue. For in-depth reviews of photoreceptor regeneration in zebrafish and how these mechanisms relate to mammalian systems, please see [145,294,295,296,297,298].

6. Conclusions

Zebrafish provide the unique opportunity to model photoreceptor disease in a cone-rich retina, similar to the human macula. There are inducible, transgenic, and mutant zebrafish models that encompass the entire spectrum of photoreceptor disease. These models can be further utilized to elucidate disease mechanism, test therapeutics, and screen novel compounds. In addition, the robust regenerative capacity of the zebrafish retina allows for the characterization of endogenous stem cells.

Funding

This work was funded by the Canadian Institutes of Health Research through a Rare Disease Models and Mechanisms grant to W.T.A. and I.M.M. and a Rare Disease Foundation microgrant to I.M.M. and N.C.L.N. N.C.L.N. was supported by a Doctoral Recruitment Scholarship from University of Alberta, Alberta Innovates Doctoral Graduate Student Scholarship, Frederick Banting and Charles Best Canada Graduate Scholarship from the Canadian Institutes of Health Research, a President’s Doctoral Prize of Distinction from University of Alberta, and a Queen Elizabeth II PhD Scholarship from the Government of Alberta.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Daiger, S.P.; Rossiter, B.J.F.; Greenberg, J.; Christoffels, A.; Hide, W. RetNet. Available online: https://sph.uth.edu/RetNet/ (accessed on 10 December 2020).
  2. Nathans, J.; Thomas, D.; Hogness, D.S. Molecular genetics of human color vision: The genes encoding blue, green, and red pigments. Science 1985, 232, 193–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Bowmaker, J.K.; Dartnall, H.J.A. Visual pigments of rods and cones in a human retina. J. Physiol. 1980, 298, 501–511. [Google Scholar] [CrossRef] [PubMed]
  4. Hofer, H.; Carroll, J.; Neitz, J.; Neitz, M.; Williams, D.R. Organization of the human trichromatic cone mosaic. J. Neurosci. 2005, 25, 9669–9679. [Google Scholar] [CrossRef] [PubMed]
  5. Curcio, C.A.; Sloan, K.R.; Packer, O.; Hendrickson, A.E.; Kalina, R.E. Distribution of cones in human and monkey retina: Individual variability and radial asymmetry. Science 1987, 236, 579–582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Schultze, M. Zur Anatomie and Physiologie der Retina. Ach Mikr Anat. 1866, 2, 175–286. [Google Scholar] [CrossRef] [Green Version]
  7. Strauss, O. The retinal pigment epithelium in visual function. Physiol. Rev. 2005, 85, 845–881. [Google Scholar] [CrossRef] [Green Version]
  8. Cayouette, M.; Smith, S.B.; Becerra, S.P.; Gravel, C. Pigment epithelium-derived factor delays the death of photoreceptors in mouse models of inherited retinal degenerations. Neurobiol. Dis. 1999, 6, 523–532. [Google Scholar] [CrossRef] [Green Version]
  9. Becerra, S.P.; Fariss, R.N.; Wu, Y.Q.; Montuenga, L.M.; Wong, P.; Pfeffer, B.A. Pigment epithelium-derived factor in the monkey retinal pigment epithelium and interphotoreceptor matrix: Apical secretion and distribution. Exp. Eye Res. 2004, 78, 223–234. [Google Scholar] [CrossRef]
  10. Ogata, N.; Wang, L.; Jo, N.; Tombran-Tink, J.; Takahashi, K.; Mrazek, D.; Matsumura, M. Pigment epithelium derived factor as a neuroprotective agent against ischemic retinal injury. Curr. Eye Res. 2001, 22, 245–252. [Google Scholar] [CrossRef]
  11. Reiter, J.F.; Leroux, M.R. Genes and molecular pathways underpinning ciliopathies. Nat. Rev. Mol. Cell Biol. 2017, 18, 533–547. [Google Scholar] [CrossRef]
  12. National Eye Institute (National Institutes of Health) AMD Data and Statistics. Available online: https://www.nei.nih.gov/learn-about-eye-health/resources-for-health-educators/eye-health-data-and-statistics/age-related-macular-degeneration-amd-data-and-statistics (accessed on 10 December 2020).
  13. Hamel, C. Retinitis pigmentosa. Orphanetj. Rare Dis. 2006, 1, 40. [Google Scholar] [CrossRef]
  14. Hanany, M.; Rivolta, C.; Sharon, D. Worldwide carrier frequency and genetic prevalence of autosomal recessive inherited retinal diseases. Proc. Natl. Acad. Sci. USA 2020, 117, 2710–2716. [Google Scholar] [CrossRef] [PubMed]
  15. Santer, R.; Schneppenheim, R.; Dombrowski, A.; Gotze, H.; Steinmann, B.; Schaub, J. Mutations in RPE65 cause Leber’s congenital amaurosis. Nat. Genet. 1997, 15, 57–61. [Google Scholar]
  16. Perrault, I.; Rozet, J.M.; Calvas, P.; Gerber, S.; Camuzat, A.; Dollfus, H.; Châtelin, S.; Souied, E.; Ghazi, I.; Leowski, C.; et al. Retinal-specific guanylate cyclase gene mutations in Leber’s congenital amaurosis. Nat. Genet. 1996, 14, 461–464. [Google Scholar] [CrossRef] [PubMed]
  17. den Hollander, A.I.; Koenekoop, R.K.; Yzer, S.; Lopez, I.; Arends, M.L.; Voesenek, K.E.; Zonneveld, M.N.; Strom, T.M.; Meitinger, T.; Brunner, H.G.; et al. Mutations in the CEP290 (NPHP6) gene are a frequent cause of Leber congenital amaurosis. Am. J. Hum. Genet. 2006, 79, 556–561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Chacon-Camacho, O.F. Review and update on the molecular basis of Leber congenital amaurosis. Worldj. Clin. Cases 2015, 3, 112. [Google Scholar] [CrossRef]
  19. Xue, L.; Gollapalli, D.R.; Maiti, P.; Jahng, W.J.; Rando, R.R. A palmitoylation switch mechanism in the regulation of the visual cycle. Cell 2004, 117, 761–771. [Google Scholar] [CrossRef] [Green Version]
  20. Wimberg, H.; Lev, D.; Yosovich, K.; Namburi, P.; Banin, E.; Sharon, D.; Koch, K.W. Photoreceptor guanylate cyclase (GUCY2D) mutations cause retinal dystrophies by severe malfunction of Ca2+-dependent cyclic GMP synthesis. Front. Mol. Neurosci. 2018, 11, 1–12. [Google Scholar] [CrossRef]
  21. Azadi, S.; Molday, L.L.; Molday, R.S. RD3, the protein associated with Leber congenital amaurosis type 12, is required for guanylate cyclase trafficking in photoreceptor cells. Proc. Natl. Acad. Sci. USA 2010, 107, 21158–21163. [Google Scholar] [CrossRef] [Green Version]
  22. Chang, B.; Khanna, H.; Hawes, N.; Jimeno, D.; He, S.; Lillo, C.; Parapuram, S.K.; Cheng, H.; Scott, A.; Hurd, R.E.; et al. In-frame deletion in a novel centrosomal/ciliary protein CEP290/NPHP6 perturbs its interaction with RPGR and results in early-onset retinal degeneration in the rd16 mouse. Hum. Mol. Genet. 2006, 15, 1847–1857. [Google Scholar] [CrossRef]
  23. Seabra, M.C.; Brown, M.S.; Goldstein, J.L. Retinal degeneration in choroideremia: Deficiency of Rab geranylgeranyl transferase. Science 1993, 259, 377–381. [Google Scholar] [CrossRef]
  24. Krock, B.L.; Bilotta, J.; Perkins, B.D. Noncell-autonomous photoreceptor degeneration in a zebrafish model of choroideremia. Proc. Natl. Acad. Sci. USA 2007, 104, 4600–4605. [Google Scholar] [CrossRef] [Green Version]
  25. Wavre-Shapton, S.T.; Tolmachova, T.; da Silva, M.L.; Futter, C.E.; Seabra, M.C. Conditional ablation of the choroideremia gene causes age-related changes in mouse retinal pigment epithelium. PLoS ONE 2013, 8, 1–11. [Google Scholar] [CrossRef]
  26. Haider, N.B.; Jacobson, S.G.; Cideciyan, A.V.; Swiderski, R.; Streb, L.M.; Searby, C.; Beck, G.; Hockey, R.; Hanna, D.B.; Gorman, S.; et al. Mutation of a nuclear receptor gene, NR2E3, causes enhanced S cone syndrome, a disorder of retinal cell fate. Nat. Genet. 2000, 24, 127–131. [Google Scholar] [CrossRef]
  27. Nishiguchi, K.M.; Friedman, J.S.; Sandberg, M.A.; Swaroop, A.; Berson, E.L.; Dryja, T.P. Recessive NRL mutations in patients with clumped pigmentary retinal degeneration and relative preservation of blue cone function. Proc. Natl. Acad. Sci. USA 2004, 101, 17819–17824. [Google Scholar] [CrossRef] [Green Version]
  28. Newman, H.; Blumen, S.C.; Braverman, I.; Hanna, R.; Tiosano, B.; Perlman, I.; Ben-Yosef, T. Homozygosity for a recessive loss-of-function mutation of the NRL gene is associated with a variant of enhanced S-cone syndrome. Investig. Ophthalmol. Vis. Sci. 2016, 57, 5361–5371. [Google Scholar] [CrossRef] [Green Version]
  29. Littink, K.W.; Stappers, P.T.Y.; Riemslag, F.C.C.; Talsma, H.E.; van Genderen, M.M.; Cremers, F.P.M.; Collin, R.W.J.; van den Born, L.I. Autosomal recessive NRL mutations in patients with enhanced S-cone syndrome. Genes 2018, 9, 68. [Google Scholar] [CrossRef] [Green Version]
  30. Sisk, R.A.; Hufnagel, R.B.; Laham, A.; Wohler, E.S.; Sobreira, N.; Ahmed, Z.M. Peripheral cone dystrophy: Expanded clinical spectrum, multimodal and ultrawide-field imaging, and genomic analysis. J. Ophthalmol. 2018, 2018. [Google Scholar] [CrossRef] [Green Version]
  31. Remmer, M.H.; Rastogi, N.; Ranka, M.P.; Ceisler, E.J. Achromatopsia: A review. Curr. Opin. Ophthalmol. 2015, 26, 333–340. [Google Scholar] [CrossRef]
  32. Viets, K.; Eldred, K.C.; Johnston, R.J. Mechanisms of photoreceptor patterning in vertebrates and invertebrates. Trends Genet. 2016, 32, 638–659. [Google Scholar] [CrossRef] [Green Version]
  33. Kostic, C.; Arsenijevic, Y. Animal modelling for inherited central vision loss. J. Pathol. 2016, 238, 300–310. [Google Scholar] [CrossRef] [Green Version]
  34. Schmitt, E.A.; Hyatt, G.A.; Dowling, J.E. Erratum: Temporal and spatial patterns of opsin gene expression in the zebrafish (Danio rerio): Corrections with additions. Vis. Neurosci. 1999, 16, 601–605. [Google Scholar] [CrossRef]
  35. Robinson, J.; Schmitt, E.A.; Dowling, J.E. Temporal and spatial patterns of opsin gene expression in zebrafish (Danio rerio). Vis. Neurosci. 1995, 12, 895–906. [Google Scholar] [CrossRef]
  36. Neuhauss, S.C.F. Behavioral genetic approaches to visual system development and function in zebrafish. J. Neurobiol. 2003, 54, 148–160. [Google Scholar] [CrossRef] [Green Version]
  37. Allison, W.T.; Barthel, L.K.; Skebo, K.M.; Takechi, M.; Kawamura, S.; Raymond, P.A. Ontogeny of cone photoreceptor mosaics in zebrafish. J. Comp. Neurol. 2010, 518, 4182–4195. [Google Scholar] [CrossRef] [Green Version]
  38. Yoshimatsu, T.; Schröder, C.; Nevala, N.E.; Berens, P.; Baden, T. Fovea-like photoreceptor specializations underlie single UV cone driven prey-capture behavior in zebrafish. Neuron 2020, 107, 320–337.e6. [Google Scholar] [CrossRef]
  39. Fadool, J.M. Development of a rod photoreceptor mosaic revealed in transgenic zebrafish. Dev. Biol. 2003, 258, 277–290. [Google Scholar] [CrossRef] [Green Version]
  40. Noel, N.C.L.; Allison, W.T. Connectivity of cone photoreceptor telodendria in the zebrafish retina. J. Comp. Neurol. 2018, 526, 609–625. [Google Scholar] [CrossRef]
  41. Morrow, J.M.; Lazic, S.; Fox, M.D.; Kuo, C.; Schott, R.K.; De Gutierrez, E.A.; Santini, F.; Tropepe, V.; Chang, B.S.W. A second visual rhodopsin gene, rh1-2, is expressed in zebrafish photoreceptors and found in other ray-finned fishes. J. Exp. Biol. 2017, 220, 294–303. [Google Scholar] [CrossRef] [Green Version]
  42. Takechi, M.; Kawamura, S. Temporal and spatial changes in the expression pattern of multiple red and green subtype opsin genes during zebrafish development. J. Exp. Biol. 2005, 208, 1337–1345. [Google Scholar] [CrossRef] [Green Version]
  43. Morrow, J.M.; Lazic, S.; Chang, B.S.W. A novel rhodopsin-like gene expressed in zebrafish retina. Vis. Neurosci. 2011, 28, 325–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ruzicka, L.; Howe, D.G.; Ramachandran, S.; Toro, S.; Van Slyke, C.E.; Bradford, Y.M.; Eagle, A.; Fashena, D.; Frazer, K.; Kalita, P.; et al. The Zebrafish Information Network: New support for non-coding genes, richer Gene Ontology annotations and the Alliance of Genome Resources. Nucleic Acids Res. 2019, 47, D867–D873. [Google Scholar] [CrossRef]
  45. Kennedy, B.N.; Alvarez, Y.; Brockerhoff, S.E.; Stearns, G.W.; Sapetto-Rebow, B.; Taylor, M.R.; Hurley, J.B. Identification of a zebrafish cone photoreceptor-specific promoter and genetic rescue of achromatopsia in the nof mutant. Investig. Ophthalmol. Vis. Sci. 2007, 48, 522–529. [Google Scholar] [CrossRef] [Green Version]
  46. Hamaoka, T.; Takechi, M.; Chinen, A.; Nishiwaki, Y.; Kawamura, S. Visualization of rod photoreceptor development using GFP-transgenic zebrafish. Genesis 2002, 34, 215–220. [Google Scholar] [CrossRef]
  47. Takechi, M.; Hamaoka, T.; Kawamura, S. Fluorescence visualization of ultraviolet-sensitive cone photoreceptor development in living zebrafish. Febs Lett. 2003, 553, 90–94. [Google Scholar] [CrossRef] [Green Version]
  48. DuVal, M.G.; Chung, H.; Lehmann, O.J.; Allison, W.T. Longitudinal fluorescent observation of retinal degeneration and regeneration in zebrafish using fundus lens imaging. Mol. Vis. 2013, 19, 1082–1095. [Google Scholar]
  49. Stearns, G.; Evangelista, M.; Fadool, J.M.; Brockerhoff, S.E. A mutation in the cone-specific pde6 gene causes rapid cone photoreceptor degeneration in zebrafish. J. Neurosci. 2007, 27, 13866–13874. [Google Scholar] [CrossRef] [Green Version]
  50. DuVal, M.G.; Oel, A.P.; Allison, W.T. gdf6a is required for cone photoreceptor subtype differentiation and for the actions of tbx2b in determining rod versus cone photoreceptor fate. PLoS ONE 2014, 9, e92991. [Google Scholar] [CrossRef] [Green Version]
  51. Yin, J.; Brocher, J.; Linder, B.; Hirmer, A.; Sundaramurthi, H.; Fischer, U.; Winkler, C. The 1D4 antibody labels outer segments of long double cone but not rod photoreceptors in zebrafish. Investig. Ophthalmol. Vis. Sci. 2012, 53, 4943–4951. [Google Scholar] [CrossRef] [Green Version]
  52. Larison, K.D.; Bremiller, R. Early onset of phenotype and cell patterning in the embryonic zebrafish retina. Development 1990, 109, 567–576. [Google Scholar]
  53. Ile, K.E.; Kassen, S.; Cao, C.; Vihtehlic, T.; Shah, S.D.; Mousley, C.J.; Alb, J.G., Jr.; Huijbregts, R.P.; Stearns, G.W.; Brockerhoff, S.E.; et al. Zebrafish class 1 phosphatidylinositol transfer proteins: PITPbeta and double cone cell outer segment integrity in retina. Traffic 2010, 11, 1151–1167. [Google Scholar] [CrossRef] [Green Version]
  54. Schmitt, E.A.; Dowling, J.E. Comparison of topographical patterns of ganglion and photoreceptor cell differentiation in the retina of the zebrafish, Danio rerio. J. Comp. Neurol. 1996, 371, 222–234. [Google Scholar] [CrossRef]
  55. Brockerhoff, S.E.; Hurley, J.B.; Janssen-Bienhold, U.; Neuhauss, S.C.F.; Driever, W.; Dowling, J.E. A behavioral screen for isolating zebrafish mutants with visual system defects. Proc. Natl. Acad. Sci. USA 1995, 92, 10545–10549. [Google Scholar] [CrossRef] [Green Version]
  56. Brockerhoff, S.E. Measuring the optokinetic response of zebrafish larvae. Nat. Protoc. 2006, 1, 2448–2451. [Google Scholar] [CrossRef]
  57. Easter, S.S.; Nicola, G.N. The development of eye movements in the zebrafish (Danio rerio). Dev. Psychobiol. 1997, 31, 267–276. [Google Scholar] [CrossRef]
  58. Orger, M.B.; Gahtan, E.; Muto, A.; Page-McCaw, P.; Smear, M.C.; Baier, H. Behavioral screening assays in zebrafish. Methods Cell Biol. 2004, 77, 53–68. [Google Scholar] [CrossRef]
  59. Orger, M.B.; Smear, M.C.; Anstis, S.M.; Baier, H. Perception of Fourier and non-Fourier motion by larval zebrafish. Nat. Neurosci. 2000, 3, 1128–1133. [Google Scholar] [CrossRef]
  60. Hagerman, G.F.; Noel, N.C.L.; Cao, S.; DuVal, M.G.; Oel, A.P.; Allison, W.T. Rapid recovery of visual function associated with blue cone ablation in zebrafish. PLoS ONE 2016, 11. [Google Scholar] [CrossRef]
  61. Emran, F.; Rihel, J.; Dowling, J.E. A behavioral assay to measure responsiveness of zebrafish to changes in light intensities. J. Vis. Exp. 2008, 20, 1–6. [Google Scholar] [CrossRef]
  62. Fernández, E.J.; Hermann, B.; Považay, B.; Unterhuber, A.; Sattmann, H.; Hofer, B.; Ahnelt, P.K.; Drexler, W. Ultrahigh resolution optical coherence tomography and pancorrection for cellular imaging of the living human retina. Opt. Express 2008, 16, 11083. [Google Scholar] [CrossRef] [Green Version]
  63. Huckenpahler, A.L.; Wilk, M.A.; Cooper, R.F.; Moehring, F.; Link, B.A.; Carroll, J.; Collery, R.F. Imaging the adult zebrafish cone mosaic using optical coherence tomography. Vis. Neurosci. 2016, 33, E011. [Google Scholar] [CrossRef] [Green Version]
  64. Collery, R.F.; Veth, K.N.; Dubis, A.M.; Carroll, J.; Link, B.A. Rapid, accurate, and non-invasive measurement of zebrafish axial length and other eye dimensions using SD-OCT allows longitudinal analysis of myopia and emmetropization. PLoS ONE 2014, 9. [Google Scholar] [CrossRef] [Green Version]
  65. Nadolski, N.J.; Wong, C.X.L.; Hocking, J.C. Electroretinogram analysis of zebrafish retinal function across development. Doc. Ophthalmol. 2020. [Google Scholar] [CrossRef]
  66. Van Epps, H.A.; Yim, C.M.; Hurley, J.B.; Brockerhoff, S.E. Investigations of photoreceptor synaptic transmission and light adaptation in the zebrafish visual mutant nrc. Investig. Ophthalmol. Vis. Sci. 2001, 42, 868–874. [Google Scholar]
  67. Makhankov, Y.V.; Rinner, O.; Neuhauss, S.C.F. An inexpensive device for non-invasive electroretinography in small aquatic vertebrates. J. Neurosci. Methods 2004, 135, 205–210. [Google Scholar] [CrossRef] [Green Version]
  68. Fleisch, V.C.; Jametti, T.; Neuhauss, S.C.F. Electroretinogram (ERG) measurements in larval zebrafish. Cold Spring Harb. Protoc. 2008, 3, 1–6. [Google Scholar] [CrossRef] [Green Version]
  69. Brockerhoff, S.E.; Hurley, J.B.; Niemi, G.A.; Dowling, J.E. A new form of inherited red-blindness identified in zebrafish. J. Neurosci. 1997, 17, 4236–4242. [Google Scholar] [CrossRef]
  70. Giarmarco, M.M.; Cleghorn, W.M.; Sloat, S.R.; Hurley, J.B.; Brockerhoff, S.E. Mitochondria maintain distinct Ca2+ pools in cone photoreceptors. J. Neurosci. 2017, 37, 2061–2072. [Google Scholar] [CrossRef] [Green Version]
  71. Ma, E.Y.; Lewis, A.; Barabas, P.; Stearns, G.; Suzuki, S.; Krizaj, D.; Brockerhoff, S.E. Loss of Pde6 reduces cell body Ca2+ transients within photoreceptors. Cell Death Dis. 2013, 4, 1–10. [Google Scholar] [CrossRef] [Green Version]
  72. Bisbach, C.M.; Hutto, R.A.; Poria, D.; Cleghorn, W.M.; Abbas, F.; Vinberg, F.; Kefalov, V.J.; Hurley, J.B.; Brockerhoff, S.E. Mitochondrial Calcium Uniporter (MCU) deficiency reveals an alternate path for Ca2+ uptake in photoreceptor mitochondria. Sci. Rep. 2020, 10, 1–19. [Google Scholar] [CrossRef]
  73. Hutto, R.A.; Bisbach, C.M.; Abbas, F.; Brock, D.C.; Cleghorn, W.M.; Parker, E.D.; Bauer, B.H.; Ge, W.; Vinberg, F.; Hurley, J.B.; et al. Increasing Ca2+ in photoreceptor mitochondria alters metabolites, accelerates photoresponse recovery, and reveals adaptations to mitochondrial stress. Cell Death Differ. 2020, 27, 1067–1085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Rosen, J.N.; Sweeney, M.F.; Mably, J.D. Microinjection of zebrafish embryos to analyze gene function. J. Vis. Exp. 2009, 1–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Kawakami, K.; Shima, A.; Kawakami, N. Identification of a functional transposase of the Tol2 element, an Ac-like element from the Japanese medaka fish, and its transposition in the zebrafish germ lineage. Proc. Natl. Acad. Sci. USA 2000, 97, 11403–11408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kawakami, K. Tol2: A versatile gene transfer vector in vertebrates. Genome Biol. 2007, 8, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Thermes, V.; Grabher, C.; Ristoratore, F.; Bourrat, F.; Choulika, A.; Wittbrodt, J.; Joly, J.S. I-SceI meganuclease mediates highly efficient transgenesis in fish. Mech. Dev. 2002, 118, 91–98. [Google Scholar] [CrossRef]
  78. Distel, M.; Wullimann, M.F.; Köster, R.W. Optimized Gal4 genetics for permanent gene expression mapping in zebrafish. Proc. Natl. Acad. Sci. USA 2009, 106, 13365–13370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Akitake, C.M.; Macurak, M.; Halpern, M.E.; Goll, M.G. Transgenerational analysis of transcriptional silencing in zebrafish. Dev. Biol. 2011, 352, 191–201. [Google Scholar] [CrossRef] [Green Version]
  80. Goll, M.G.; Anderson, R.; Stainier, D.Y.R.; Spradling, A.C.; Halpern, M.E. Transcriptional silencing and reactivation in transgenic zebrafish. Genetics 2009, 182, 747–755. [Google Scholar] [CrossRef] [Green Version]
  81. Jungke, P.; Hammer, J.; Hans, S.; Brand, M. Isolation of novel CreERT2-driver lines in zebrafish using an unbiased gene trap approach. PLoS ONE 2015, 10, 1–24. [Google Scholar] [CrossRef] [Green Version]
  82. Kirchgeorg, L.; Felker, A.; van Oostrom, M.; Chiavacci, E.; Mosimann, C. Cre/lox-controlled spatiotemporal perturbation of FGF signaling in zebrafish. Dev. Dyn. 2018, 247, 1146–1159. [Google Scholar] [CrossRef] [Green Version]
  83. Lin, H.J.; Lee, S.H.; Wu, J.L.; Duann, Y.F.; Chen, J.Y. Development of Cre-loxP technology in zebrafish to study the regulation of fish reproduction. Fish Physiol. Biochem. 2013, 39, 1525–1539. [Google Scholar] [CrossRef] [PubMed]
  84. Mukherjee, K.; Liao, E.C. Generation and characterization of a zebrafish muscle specific inducible Cre line. Transgenic Res. 2018, 27, 559–569. [Google Scholar] [CrossRef] [PubMed]
  85. Le, X.; Langenau, D.M.; Keefe, M.D.; Kutok, J.L.; Neuberg, D.S.; Zon, L.I. Heat shock-inducible Cre/Lox approaches to induce diverse types of tumors and hyperplasia in transgenic zebrafish. Proc. Natl. Acad. Sci. USA 2007, 104, 9410–9415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Tabor, K.M.; Marquart, G.D.; Hurt, C.; Smith, T.S.; Geoca, A.K.; Bhandiwad, A.A.; Subedi, A.; Sinclair, J.L.; Rose, H.M.; Polys, N.F.; et al. Brain-wide cellular resolution imaging of Cre transgenic zebrafish lines for functional circuit-mapping. eLife 2019, 8, 1–21. [Google Scholar] [CrossRef] [PubMed]
  87. Muto, A.; Orger, M.B.; Wehman, A.M.; Smear, M.C.; Kay, J.N.; Page-McCaw, P.S.; Gahtan, E.; Xiao, T.; Nevin, L.M.; Gosse, N.J.; et al. Forward genetic analysis of visual behavior in zebrafish. PLoS Genet. 2005, 1, e66. [Google Scholar] [CrossRef] [PubMed]
  88. Doyon, Y.; McCammon, J.M.; Miller, J.C.; Faraji, F.; Ngo, C.; Katibah, G.E.; Amora, R.; Hocking, T.D.; Zhang, L.; Rebar, E.J.; et al. Heritable targeted gene disruption in zebrafish using designed zinc-finger nucleases. Nat. Biotechnol. 2008, 26, 702–708. [Google Scholar] [CrossRef] [Green Version]
  89. Zu, Y.; Tong, X.; Wang, Z.; Liu, D.; Pan, R.; Li, Z.; Hu, Y.; Luo, Z.; Huang, P.; Wu, Q.; et al. TALEN-mediated precise genome modification by homologous recombination in zebrafish. Nat. Methods 2013, 10, 329–331. [Google Scholar] [CrossRef]
  90. Hwang, W.Y.; Fu, Y.; Reyon, D.; Maeder, M.L.; Tsai, S.Q.; Sander, J.D.; Peterson, R.T.; Yeh, J.R.J.; Joung, J.K. Efficient genome editing in zebrafish using a CRISPR-Cas system. Nat. Biotechnol. 2013, 31, 227–229. [Google Scholar] [CrossRef]
  91. Doench, J.G.; Fusi, N.; Sullender, M.; Hegde, M.; Vaimberg, E.W.; Donovan, K.F.; Smith, I.; Tothova, Z.; Wilen, C.; Orchard, R.; et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9. Nat. Biotechnol. 2016, 34, 184–191. [Google Scholar] [CrossRef] [Green Version]
  92. Ran, F.A.; Hsu, P.D.; Lin, C.Y.; Gootenberg, J.S.; Konermann, S.; Trevino, A.E.; Scott, D.A.; Inoue, A.; Matoba, S.; Zhang, Y.; et al. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 2013, 154, 1380–1389. [Google Scholar] [CrossRef] [Green Version]
  93. Maruyama, T.; Dougan, S.K.; Truttmann, M.C.; Bilate, A.M.; Ingram, J.R.; Ploegh, H.L. Increasing the efficiency of precise genome editing with CRISPR-Cas9 by inhibition of nonhomologous end joining. Nat. Biotechnol. 2015, 33, 538–542. [Google Scholar] [CrossRef] [PubMed]
  94. Kocak, D.D.; Josephs, E.A.; Bhandarkar, V.; Adkar, S.S.; Kwon, J.B.; Gersbach, C.A. Increasing the specificity of CRISPR systems with engineered RNA secondary structures. Nat. Biotechnol. 2019, 37, 657–666. [Google Scholar] [CrossRef] [PubMed]
  95. Kassahn, K.S.; Dang, V.T.; Wilkins, S.J.; Perkins, A.C.; Ragan, M.A. Evolution of gene function and regulatory control after whole-genome duplication: Comparative analyses in vertebrates. Genome Res. 2009, 19, 1404–1418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Schlegel, D.K.; Glasauer, S.M.K.; Mateos, J.M.; Barmettler, G.; Ziegler, U.; Neuhauss, S.C.F. A new zebrafish model for CACNA2D4-dysfunction. Investig. Ophthalmol. Vis. Sci. 2019, 60, 5124–5135. [Google Scholar] [CrossRef] [Green Version]
  97. Jia, S.; Muto, A.; Orisme, W.; Henson, H.E.; Parupalli, C.; Ju, B.; Baier, H.; Taylor, M.R. Zebrafish cacna1fa is required for cone photoreceptor function and synaptic ribbon formation. Hum. Mol. Genet. 2014, 23, 2981–2994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Seiler, C.; Finger-Baier, K.C.; Rinner, O.; Makhankov, Y.V.; Schwarz, H.; Neuhauss, S.C.F.; Nicolson, T. Duplicated genes with split functions: Independent roles of protocadherin 15 orthologues in zebrafish hearing and vision. Development 2005, 132, 615–623. [Google Scholar] [CrossRef] [Green Version]
  99. Hitchcock, P.F.; Raymond, P.A. The teleost retina as a model for developmental and regeneration biology. Zebrafish 2004, 1, 257–271. [Google Scholar] [CrossRef]
  100. Wan, J.; Ramachandran, R.; Goldman, D. HB-EGF is necessary and sufficient for Müller glia dedifferentiation and retina regeneration. Dev. Cell 2012, 22, 334–347. [Google Scholar] [CrossRef] [Green Version]
  101. Ramachandran, R.; Zhao, X.F.; Goldman, D. Insm1a-mediated gene repression is essential for the formation and differentiation of Müller glia-derived progenitors in the injured retina. Nat. Cell Biol. 2012, 14, 1013–1023. [Google Scholar] [CrossRef] [Green Version]
  102. Meyers, J.R.; Hu, L.; Moses, A.; Kaboli, K.; Papandrea, A.; Raymond, P.A. β-catenin/Wnt signaling controls progenitor fate in the developing and regenerating zebrafish retina. Neural Dev. 2012, 7, 1–17. [Google Scholar] [CrossRef] [Green Version]
  103. Qin, Z.; Barthel, L.K.; Raymond, P.A. Genetic evidence for shared mechanisms of epimorphic regeneration in zebrafish. Proc. Natl. Acad. Sci. USA 2009, 106, 9310–9315. [Google Scholar] [CrossRef] [Green Version]
  104. Bernardos, R.L.; Barthel, L.K.; Meyers, J.R.; Raymond, P.A. Late-stage neuronal progenitors in the retina are radial Müller glia that function as retinal stem cells. J. Neurosci. 2007, 27, 7028–7040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Penn, J.S. Effects of continuous light on the retina of a fish, Notemigonus crysoleucas. J. Comp. Neurol. 1985, 238, 121–127. [Google Scholar] [CrossRef] [PubMed]
  106. Shahinfar, S.; Edward, D.P.; Tso, M.O. A pathological study of photoreceptor cell death in retinal photic injury. Curr. Eye Res. 1991, 10, 47–59. [Google Scholar] [CrossRef] [PubMed]
  107. Thummel, R.; Kassen, S.C.; Enright, J.M.; Nelson, C.M.; Montgomery, J.E.; Hyde, D.R. Characterization of Müller glia and neuronal progenitors during adult zebrafish retinal regeneration. Exp. Eye Res. 2008, 87, 433–444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Vihtelic, T.S.; Soverly, J.E.; Kassen, S.C.; Hyde, D.R. Retinal regional differences in photoreceptor cell death and regeneration in light-lesioned albino zebrafish. Exp. Eye Res. 2006, 82, 558–575. [Google Scholar] [CrossRef]
  109. Dunn, F.A. Photoreceptor ablation initiates the immediate loss of glutamate receptors in postsynaptic bipolar cells in retina. J. Neurosci. 2015, 35, 2423–2431. [Google Scholar] [CrossRef] [Green Version]
  110. Winkler, B.S. Relative inhibitory effects of ATP depletion, ouabain and calcium on retinal photoreceptors. Exp. Eye Res. 1983, 36, 581–594. [Google Scholar] [CrossRef]
  111. Fimbel, S.M.; Montgomery, J.E.; Burket, C.T.; Hyde, D.R. Regeneration of inner retinal neurons after intravitreal injection of ouabain in zebrafish. J. Neurosci. 2007, 27, 1712–1724. [Google Scholar] [CrossRef] [Green Version]
  112. Maier, W.; Wolburg, H. Regeneration of the goldfish retina after exposure to different doses of ouabain. Cell Tissue Res. 1979, 202, 99–118. [Google Scholar] [CrossRef]
  113. Terracjni, B.; Testa, M.C. Carcinogenicity of a single administration of N-nitrosomethylurea: A comparison between newborn and 5-week-old mice and rats. Br. J. Cancer 1970, 24, 588–598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Wurdeman, R.L.; Church, K.M.; Gold, B. DNA methylation by N-methyl-N-nitrosourea, N-methyl-N’-nitro-N-nitrosoguanidine, N-nitroso(1-acetoxyethyl)methylamine and diazomethane: Mechanism for the formation of N7-methylguanine in sequence-characterized 5′-32P-end-labeled DNA. J. Am. Chem. Soc. 1989, 111, 6408–6412. [Google Scholar] [CrossRef]
  115. Tappeiner, C.; Balmer, J.; Iglicki, M.; Schuerch, K.; Jazwinska, A.; Enzmann, V.; Tschopp, M. Characteristics of rod regeneration in a novel zebrafish retinal degeneration model using N-methyl-N-nitrosourea (MNU). PLoS ONE 2013, 8, 4–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Nagar, S.; Krishnamoorthy, V.; Cherukuri, P.; Jain, V.; Dhingra, N.K. Early remodeling in an inducible animal model of retinal degeneration. Neuroscience 2009, 160, 517–529. [Google Scholar] [CrossRef] [PubMed]
  117. Zulliger, R.; Lecaudé, S.; Eigeldinger-Berthou, S.; Wolf-Schnurrbusch, U.E.K.; Enzmann, V. Caspase-3-independent photoreceptor degeneration by N-methyl-N-nitrosourea (MNU) induces morphological and functional changes in the mouse retina. Graefe’s Arch. Clin. Exp. Ophthalmol. 2011, 249, 859–869. [Google Scholar] [CrossRef] [Green Version]
  118. Cao, R.; Jensen, L.D.E.; Söll, I.; Hauptmann, G.; Cao, Y. Hypoxia-induced retinal angiogenesis in zebrafish as a model to study retinopathy. PLoS ONE 2008, 3, 1–9. [Google Scholar] [CrossRef] [Green Version]
  119. Nasevicius, A.; Ekker, S.C. Effective targeted gene “knockdown” in zebrafish. Nat. Genet. 2000, 26, 216–220. [Google Scholar] [CrossRef]
  120. Draper, B.W.; Morcos, P.A.; Kimmel, C.B. Inhibition of zebrafish fgf8 pre-mRNA splicing with morpholino oligos: A quantifiable method for gene knockdown. Genesis 2001, 30, 154–156. [Google Scholar] [CrossRef]
  121. Kok, F.O.; Shin, M.; Ni, C.W.; Gupta, A.; Grosse, A.S.; vanImpel, A.; Kirchmaier, B.C.; Peterson-Maduro, J.; Kourkoulis, G.; Male, I.; et al. Reverse genetic screening reveals poor correlation between morpholino-induced and mutant phenotypes in zebrafish. Dev. Cell 2015, 32, 97–108. [Google Scholar] [CrossRef] [Green Version]
  122. Rossi, A.; Kontarakis, Z.; Gerri, C.; Nolte, H.; Hölper, S.; Krüger, M.; Stainier, D.Y.R. Genetic compensation induced by deleterious mutations but not gene knockdowns. Nature 2015, 524, 230–233. [Google Scholar] [CrossRef]
  123. Morris, A.C.; Schroeter, E.H.; Bilotta, J.; Wong, R.O.; Fadool, J.M. Cone survival despite rod degeneration in XOPS-mCFP transgenic zebrafish. Investig. Ophthalmol. Vis. Sci. 2005, 46, 4762–4771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Curado, S.; Stainier, D.Y.; Anderson, R.M. Nitroreductase-mediated cell/tissue ablation in zebrafish: A spatially and temporally controlled ablation method with applications in developmental and regeneration studies. Nat. Protoc. 2008, 3, 948–954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Fraser, B.; DuVal, M.G.; Wang, H.; Allison, W.T. Regeneration of cone photoreceptors when cell ablation is primarily restricted to a particular cone subtype. PLoS ONE 2013, 8, e55410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Montgomery, J.E.; Parsons, M.J.; Hyde, D.R. A novel model of retinal ablation demonstrate that the extent of rod cell death regulates the origin of the regenerated zebrafish rod photoreceptors. J. Comp. Neurol. 2010, 518, 800–814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Yoshimatsu, T.; D’Orazi, F.D.; Gamlin, C.R.; Suzuki, S.C.; Suli, A.; Kimelman, D.; Raible, D.W.; Wong, R.O. Presynaptic partner selection during retinal circuit reassembly varies with timing of neuronal regeneration in vivo. Nat. Commun. 2016, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  128. White, D.T.; Sengupta, S.; Saxena, M.T.; Xu, Q.; Hanes, J.; Ding, D.; Ji, H.; Mumm, J.S. Immunomodulation-accelerated neuronal regeneration following selective rod photoreceptor cell ablation in the zebrafish retina. Proc. Natl. Acad. Sci. USA 2017, 114, E3719–E3728. [Google Scholar] [CrossRef] [Green Version]
  129. D’Orazi, F.D.; Suzuki, S.C.; Darling, N.; Wong, R.O.; Yoshimatsu, T. Conditional and biased regeneration of cone photoreceptor types in the zebrafish retina. J. Comp. Neurol. 2020, 528, 2816–2830. [Google Scholar] [CrossRef]
  130. Hanovice, N.J.; Leach, L.L.; Slater, K.; Gabriel, A.E.; Romanovicz, D.; Shao, E.; Collery, R.; Burton, E.A.; Lathrop, K.L.; Link, B.A.; et al. Regeneration of the zebrafish retinal pigment epithelium after widespread genetic ablation. PLoS Genet. 2019, 15, 1–35. [Google Scholar] [CrossRef] [Green Version]
  131. Dryja, T.P.; McGee, T.L.; Reichel, E.; Hahn, L.; Cowley, G.S.; Yandell, D.W.; Sandberg, M.A.; Berson, E.L. A point mutation of the rhodopsin gene in one form of retinitis pigmentosa. Nature 1990, 343, 364–366. [Google Scholar] [CrossRef]
  132. Sullivan, L.S.; Bowne, S.J.; Birch, D.G.; Hughbanks-Wheaton, D.; Heckenlively, J.R.; Lewis, R.A.; Garcia, C.A.; Ruiz, R.S.; Blanton, S.H.; Northrup, H.; et al. Prevalence of disease-causing mutations in families with autosomal dominant retinitis pigmentosa: A screen of known genes in 200 families. Investig. Ophthalmol. Vis. Sci. 2006, 47, 3052–3064. [Google Scholar] [CrossRef]
  133. Parmeggiani, F.S.; Sorrentino, F.; Ponzin, D.; Barbaro, V.; Ferrari, S.; Di Iorio, E. Retinitis pigmentosa: Genes and disease mechanisms. Curr. Genom. 2011, 12, 238–249. [Google Scholar] [CrossRef] [PubMed]
  134. Santhanam, A.; Shihabeddin, E.; Atkinson, J.A.; Nguyen, D.; Lin, Y.-P.; O’Brien, J. A zebrafish model of retinitis pigmentosa shows continuous degeneration and regeneration of rod photoreceptors. Cells 2020, 9, 2242. [Google Scholar] [CrossRef]
  135. Sung, C.H.; Makino, C.; Baylor, D.; Nathans, J. A rhodopsin gene mutation responsible for autosomal dominant retinitis pigmentosa results in a protein that is defective in localization to the photoreceptor outer segment. J. Neurosci. 1994, 14, 5818–5833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Nakao, T.; Tsujikawa, M.; Notomi, S.; Ikeda, Y.; Nishida, K. The role of mislocalized phototransduction in photoreceptor cell death of retinitis pigmentosa. PLoS ONE 2012, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Vithana, E.N.; Abu-Safieh, L.; Allen, M.J.; Carey, A.; Papaioannou, M.; Chakarova, C.; Al-Maghtheh, M.; Ebenezer, N.D.; Willis, C.; Moore, A.T.; et al. A human homolog of yeast pre-mRNA splicing gene, PRP31, underlies autosomal dominant retinitis pigmentosa on chromosome 19q13.4 (RP11). Mol. Cell 2001, 8, 375–381. [Google Scholar] [CrossRef]
  138. Yin, J.; Brocher, J.; Fischer, U.; Winkler, C. Mutant Prpf31 causes pre-mRNA splicing defects and rod photoreceptor cell degeneration in a zebrafish model for retinitis pigmentosa. Mol. Neurodegener. 2011, 6, 56. [Google Scholar] [CrossRef] [Green Version]
  139. Yang, Z.; Camp, N.J.; Sun, H.; Tong, Z.; Gibbs, D.; Cameron, D.J.; Chen, H.; Zhao, Y.; Pearson, E.; Li, X.; et al. A variant of the HTRA1 gene increases susceptibility to age-related macular degeneration. Science 2006, 314, 992–993. [Google Scholar] [CrossRef] [PubMed]
  140. DeWan, A.; Liu, M.; Hartman, S.; Zhang, S.S.-M.; Liu, D.T.L.; Zhao, C.; Tam, P.O.S.T.; Chan, W.M.; Lam, D.S.C.; Snyder, M.; et al. HTRA1 promoter polymorphism in wet age-related macular degeneration. Science 2006, 314, 989–993. [Google Scholar] [CrossRef]
  141. Jones, A.; Kumar, S.; Zhang, N.; Tong, Z.; Yang, J.H.; Watt, C.; Anderson, J.; Amrita; Fillerup, H.; McCloskey, M.; et al. Increased expression of multifunctional serine protease, HTRA1, in retinal pigment epithelium induces polypoidal choroidal vasculopathy in mice. Proc. Natl. Acad. Sci. USA 2011, 108, 14578–14583. [Google Scholar] [CrossRef] [Green Version]
  142. Oura, Y.; Nakamura, M.; Takigawa, T.; Fukushima, Y.; Wakabayashi, T.; Tsujikawa, M.; Nishida, K. High-temperature requirement A 1 causes photoreceptor cell death in zebrafish disease models. Am. J. Pathol. 2018, 188, 2729–2744. [Google Scholar] [CrossRef] [Green Version]
  143. Collery, R.F.; Cederlund, M.L.; Kennedy, B.N. Transgenic zebrafish expressing mutant human RETGC-1 exhibit aberrant cone and rod morphology. Exp. Eye Res. 2013, 108, 120–128. [Google Scholar] [CrossRef] [PubMed]
  144. Link, B.A.; Collery, R.F. Zebrafish models of retinal disease. Annu. Rev. Vis. Sci. 2015, 1, 125–153. [Google Scholar] [CrossRef] [PubMed]
  145. Angueyra, J.M.; Kindt, K.S. Leveraging zebrafish to study retinal degenerations. Front. Cell Dev. Biol. 2018, 6, 1–19. [Google Scholar] [CrossRef] [PubMed]
  146. Chhetri, J.; Jacobson, G.; Gueven, N. Zebrafish–on the move towards ophthalmological research. Eye 2014, 28, 367–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Gross, J.M.; Perkins, B.D. Zebrafish mutants as models for congenital ocular disorders in humans. Mol. Reprod. Dev. 2008, 75, 547–555. [Google Scholar] [CrossRef] [PubMed]
  148. Morris, A.C. The genetics of ocular disorders: Insights from the zebrafish. Birth Defects Res. Part. C Embryo Today Rev. 2011, 93, 215–228. [Google Scholar] [CrossRef]
  149. Athanasiou, D.; Aguila, M.; Bellingham, J.; Li, W.; McCulley, C.; Reeves, P.J.; Cheetham, M.E. The molecular and cellular basis of rhodopsin retinitis pigmentosa reveals potential strategies for therapy. Prog. Retin. Eye Res. 2018, 62, 1–23. [Google Scholar] [CrossRef] [Green Version]
  150. Zelinka, C.P.; Sotolongo-Lopez, M.; Fadool, J.M. Targeted disruption of the endogenous zebrafish rhodopsin locus as models of rapid rod photoreceptor degeneration. Mol. Vis. 2018, 24, 587–602. [Google Scholar]
  151. Eroglu, A.U.; Mulligan, T.S.; Zhang, L.; White, D.T.; Sengupta, S.; Nie, C.; Lu, N.Y.; Qian, J.; Xu, L.; Pei, W.; et al. Multiplexed CRISPR/Cas9 targeting of genes implicated in retinal regeneration and degeneration. Front. Cell Dev. Biol. 2018, 6. [Google Scholar] [CrossRef] [Green Version]
  152. Schwahn, U.; Lenzner, S.; Dong, J.; Feil, S.; Hinzmann, B.; Van Duijnhoven, G.; Kirschner, R.; Hemberger, M.; Bergen, A.A.B.; Rosenberg, T.; et al. Positional cloning of the gene for X-linked retinitis pigmentosa 2. Nat. Genet. 1998, 19, 327–332. [Google Scholar] [CrossRef]
  153. Hurd, T.W.; Fan, S.; Margolis, B.L. Localization of retinitis pigmentosa 2 to cilia is regulated by importin β2. J. Cell Sci. 2011, 124, 718–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Schwarz, N.; Hardcastle, A.J.; Cheetham, M.E. Arl3 and RP2 mediated assembly and traffic of membrane associated cilia proteins. Vis. Res. 2012, 75, 2–4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Evans, R.J.; Schwarz, N.; Nagel-Wolfrum, K.; Wolfrum, U.; Hardcastle, A.J.; Cheetham, M.E. The retinitis pigmentosa protein RP2 links pericentriolar vesicle transport between the Golgi and the primary cilium. Hum. Mol. Genet. 2010, 19, 1358–1367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Liu, F.; Chen, J.; Yu, S.; Raghupathy, R.K.; Liu, X.; Qin, Y.; Li, C.; Huang, M.; Liao, S.; Wang, J.; et al. Knockout of RP2 decreases GRK1 and rod transducin subunits and leads to photoreceptor degeneration in zebrafish. Hum. Mol. Genet. 2015, 24, 4648–4659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Tuson, M.; Marfany, G.; Gonzàlez-Duarte, R. Mutation of CERKL, a novel human ceramide kinase gene, causes autosomal recessive retinitis pigmentosa (RP26). Am. J. Hum. Genet. 2004, 74, 128–138. [Google Scholar] [CrossRef] [Green Version]
  158. Yu, S.; Li, C.; Biswas, L.; Hu, X.; Liu, F.; Reilly, J.; Liu, X.; Liu, Y.; Huang, Y.; Lu, Z.; et al. CERKL gene knockout disturbs photoreceptor outer segment phagocytosis and causes rod-cone dystrophy in zebrafish. Hum. Mol. Genet. 2017, 26, 2335–2345. [Google Scholar] [CrossRef]
  159. Cogné, B.; Latypova, X.; Senaratne, L.D.S.; Martin, L.; Koboldt, D.C.; Kellaris, G.; Fievet, L.; Le Meur, G.; Caldari, D.; Debray, D. Mutations in the kinesin-2 motor KIF3B cause an autosomal-dominant ciliopathy. Am. J. Hum. Genet. 2020, 106, 893–904. [Google Scholar] [CrossRef]
  160. Zhao, C.; Omori, Y.; Brodowska, K.; Kovach, P.; Malicki, J. Kinesin-2 family in vertebrate ciliogenesis. Proc. Natl. Acad. Sci. USA 2012, 109, 2388–2393. [Google Scholar] [CrossRef] [Green Version]
  161. Feng, D.; Chen, Z.; Yang, K.; Miao, S.; Xu, B.; Kang, Y.; Xie, H.; Zhao, C. The cytoplasmic tail of rhodopsin triggers rapid rod degeneration in kinesin-2 mutants. J. Biol. Chem. 2017, 292, 1735–17386. [Google Scholar] [CrossRef] [Green Version]
  162. Dryja, T.P.; Adams, S.M.; Grimsby, J.L.; McGee, T.L.; Hong, D.H.; Li, T.; Andréasson, S.; Berson, E.L. Null RPGRIP1 alleles in patients with Leber congenital amaurosis. Am. J. Hum. Genet. 2001, 68, 1295–1298. [Google Scholar] [CrossRef] [Green Version]
  163. Hameed, A.; Abid, A.; Aziz, A.; Ismail, M.; Mehdi, S.Q.; Khaliq, S. Evidence of RPGRIP1 gene mutations associated with recessive cone-rod dystrophy. J. Med. Genet. 2003, 40, 616–619. [Google Scholar] [CrossRef] [PubMed]
  164. Booij, J.C.; Florijn, R.J.; ten Brink, J.B.; Loves, W.; Meire, F.; van Schooneveld, M.J.; de Jong, P.T.; Bergen, A.A. Identification of mutations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. J. Med. Genet. 2005, 42, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Mavlyutov, T.A.; Zhao, H.; Ferreira, P.A. Species-specific subcellular localization of RPGR and RPGRIP isoforms: Implications for the phenotypic variability of congenital retinopathies among species. Hum. Mol. Genet. 2002, 11, 1899–1907. [Google Scholar] [CrossRef] [PubMed]
  166. Raghupathy, R.K.; Zhang, X.; Liu, F.; Alhasani, R.H.; Biswas, L.; Akhtar, S.; Pan, L.; Moens, C.B.; Li, W.; Liu, M.; et al. Rpgrip1 is required for rod outer segment development and ciliary protein trafficking in zebrafish. Sci. Rep. 2017, 7, 1–14. [Google Scholar] [CrossRef] [Green Version]
  167. Davidson, A.E.; Sergouniotis, P.I.; Mackay, D.S.; Wright, G.A.; Waseem, N.H.; Michaelides, M.; Holder, G.E.; Robson, A.G.; Moore, A.T.; Plagnol, V.; et al. RP1L1 variants are associated with a spectrum of inherited retinal diseases including retinitis pigmentosa and occult macular dystrophy. Hum. Mutat. 2013, 34, 506–514. [Google Scholar] [CrossRef]
  168. Noel, N.C.L.; MacDonald, I.M. RP1L1 and inherited photoreceptor disease: A review. Surv. Ophthalmol. 2020, 65, 725–739. [Google Scholar] [CrossRef]
  169. Akahori, M.; Tsunoda, K.; Miyake, Y.; Fukuda, Y.; Ishiura, H.; Tsuji, S.; Usui, T.; Hatase, T.; Nakamura, M.; Ohde, H.; et al. Dominant mutations in RP1L1 are responsible for occult macular dystrophy. Am. J. Hum. Genet. 2010, 87, 424–429. [Google Scholar] [CrossRef] [Green Version]
  170. Noel, N.C.L.; Nadolski, N.J.; Hocking, J.C.; Macdonald, I.M.; Allison, W.T. Progressive photoreceptor dysfunction and age-related macular degeneration-like features in rp1l1 mutant zebrafish. Cells 2020, 9, 2214. [Google Scholar] [CrossRef]
  171. El-Amraoui, A.; Sahly, I.; Picaud, S.; Sahel, J.; Abitbol, M.; Petit, C. Human Usher 1B/mouse shaker-1: The retinal phenotype discrepancy explained by the presence/absence of myosin VIIA in the photoreceptor cells. Hum. Mol. Genet. 1996, 5, 1171–1178. [Google Scholar] [CrossRef] [Green Version]
  172. Zina, Z.B.; Masmoudi, S.; Ayadi, H.; Chaker, F.; Ghorbel, A.M.; Drira, M.; Petit, C. From DFNB2 to usher syndrome: Variable expressivity of the same disease. Am. J. Med. Genet. 2001, 101, 181–183. [Google Scholar] [CrossRef]
  173. Weil, D.; Küssel, P.; Blanchard, S.; Lévy, G.; Levi-Acobas, F.; Drira, M.; Ayadi, H.; Petit, C. The autosomal recessive isolated deafness, DFNB2, and the Usher 1B syndrome are allelic defects of the myosin-VIIA gene. Nat. Genet. 1997, 16, 191–193. [Google Scholar] [CrossRef] [PubMed]
  174. Wasfy, M.M.; Matsui, J.I.; Miller, J.; Dowling, J.E.; Perkins, B.D. Myosin 7aa-/- mutant zebrafish show mild photoreceptor degeneration and reduced electroretinographic responses. Exp. Eye Res. 2014, 122, 65–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Eudy, J.D.; Weston, M.D.; Yao, S.F.; Hoover, D.M.; Rehm, H.L.; Ma-Edmonds, M.; Yan, D.; Ahmad, I.; Cheng, J.J.; Ayuso, C.; et al. Mutation of a gene encoding a protein with extracellular matrix motifs in Usher syndrome type IIa. Science 1998, 280, 1753–1757. [Google Scholar] [CrossRef] [PubMed]
  176. van Wijk, E.; Pennings, R.J.E.; te Brinke, H.; Claassen, A.; Yntema, H.G.; Hoefsloot, L.H.; Cremers, F.P.M.; Cremers, W.R.J.; Kremer, H. Identification of 51 Novel Exons of the Usher Syndrome Type 2A (USH2A) Gene That Encode Multiple Conserved Functional Domains and That Are Mutated in Patients with Usher Syndrome Type II. Am. J. Hum. Genet. 2004, 74, 738–744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Han, S.; Liu, X.; Xie, S.; Gao, M.; Liu, F.; Yu, S.; Sun, P.; Wang, C.; Archacki, S.; Lu, Z.; et al. Knockout of ush2a gene in zebrafish causes hearing impairment and late onset rod-cone dystrophy. Hum. Genet. 2018, 137, 779–794. [Google Scholar] [CrossRef]
  178. Dona, M.; Slijkerman, R.; Lerner, K.; Broekman, S.; Wegner, J.; Howat, T.; Peters, T.; Hetterschijt, L.; Boon, N.; de Vrieze, E.; et al. Usherin defects lead to early-onset retinal dysfunction in zebrafish. Exp. Eye Res. 2018, 173, 148–159. [Google Scholar] [CrossRef]
  179. Morris, A.C.; Forbes-Osborne, M.A.; Pillai, L.S.; Fadool, J.M. Microarray analysis of XOPS-mCFP zebrafish retina identifies genes associated with rod photoreceptor degeneration and regeneration. Investig. Ophthalmol. Vis. Sci. 2011, 52, 2255–2266. [Google Scholar] [CrossRef] [Green Version]
  180. Coomer, C.E.; Wilson, S.G.; Titialii-Torres, K.F.; Bills, J.D.; Krueger, L.A.; Petersen, R.A.; Turnbaugh, E.M.; Janesch, E.L.; Morris, A.C. Her9/Hes4 is required for retinal photoreceptor development, maintenance, and survival. Sci. Rep. 2020, 10, 1–20. [Google Scholar] [CrossRef]
  181. Sukumaran, S.; Perkins, B.D. Early defects in photoreceptor outer segment morphogenesis in zebrafish ift57, ift88 and ift172 intraflagellar transport mutants. Vis. Res. 2009, 49, 479–489. [Google Scholar] [CrossRef] [Green Version]
  182. Doerre, G.; Malicki, J. Genetic analysis of photoreceptor cell development in the zebrafish retina. Mech. Dev. 2002, 110, 125–138. [Google Scholar] [CrossRef]
  183. Tsujikawa, M.; Malicki, J. Intraflagellar transport genes are essential for differentiation and survival of vertebrate sensory neurons. Neuron 2004, 42, 703–716. [Google Scholar] [CrossRef] [Green Version]
  184. Krock, B.L.; Perkins, B.D. The intraflagellar transport protein IFT57 is required for cilia maintenance and regulates IFT-particle-kinesin-II dissociation in vertebrate photoreceptors. J. Cell Sci. 2008, 121, 1907–1915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Lee, C.; Wallingford, J.B.; Gross, J.M. Cluap1 is essential for ciliogenesis and photoreceptor maintenance in the vertebrate eye. Investig. Ophthalmol. Vis. Sci. 2014, 55, 4585–4592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Gross, J.M.; Perkins, B.D.; Amsterdam, A.; Egaña, A.; Darland, T.; Matsui, J.I.; Sciascia, S.; Hopkins, N.; Dowling, J.E. Identification of zebrafish insertional mutants with defects in visual system development and function. Genetics 2005, 170, 245–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Boubakri, M.; Chaya, T.; Hirata, H.; Kajimura, N.; Kuwahara, R.; Ueno, A.; Malicki, J.; Furukawa, T.; Omori, Y. Loss of ift122, a retrograde intraflagellar transport (IFT) complex component, leads to slow, progressive photoreceptor degeneration due to inefficient opsin transport. J. Biol. Chem. 2016, 291, 24465–24474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Pooranachandran, N.; Malicki, J.J. Unexpected roles for ciliary kinesins and intraflagellar transport proteins. Genetics 2016, 203, 771–785. [Google Scholar] [CrossRef] [PubMed]
  189. Raghupathy, R.K.; Zhang, X.; Alhasani, R.H.; Zhou, X.; Mullin, M.; Reilly, J.; Li, W.; Liu, M.; Shu, X. Abnormal photoreceptor outer segment development and early retinal degeneration in kif3a mutant zebrafish. Cell Biochem. Funct. 2016, 34, 429–440. [Google Scholar] [CrossRef] [Green Version]
  190. Burgalossi, A.; Jung, S.; Meyer, G.; Jockusch, W.J.; Jahn, O.; Taschenberger, H.; O’Connor, V.M.; Nishiki, T.-i.; Takahashi, M.; Brose, N.; et al. SNARE protein recycling by αSNAP and βSNAP supports synaptic vesicle priming. Neuron 2010, 68, 473–487. [Google Scholar] [CrossRef] [Green Version]
  191. Nishiwaki, Y.; Yoshizawa, A.; Kojima, Y.; Oguri, E.; Nakamura, S.; Suzuki, S.; Yuasa-Kawada, J.; Kinoshita-Kawada, M.; Mochizuki, T.; Masai, I. The BH3-only SNARE BNip1 mediates photoreceptor apoptosis in response to vesicular fusion defects. Dev. Cell 2013, 25, 374–387. [Google Scholar] [CrossRef] [Green Version]
  192. Nishiwaki, Y.; Komori, A.; Sagara, H.; Suzuki, E.; Manabe, T.; Hosoya, T.; Nojima, Y.; Wada, H.; Tanaka, H.; Okamoto, H.; et al. Mutation of cGMP phosphodiesterase 6α′-subunit gene causes progressive degeneration of cone photoreceptors in zebrafish. Mech. Dev. 2008, 125, 932–946. [Google Scholar] [CrossRef]
  193. Bujakowska, K.M.; Zhang, Q.; Siemiatkowska, A.M.; Liu, Q.; Place, E.; Falk, M.J.; Consugar, M.; Lancelot, M.E.; Antonio, A.; Lonjou, C.; et al. Mutations in IFT172 cause isolated retinal degeneration and Bardet-Biedl syndrome. Hum. Mol. Genet. 2015, 24, 230–242. [Google Scholar] [CrossRef] [PubMed]
  194. Bachmann-Gagescu, R.; Phelps, I.G.; Dempsey, J.C.; Sharma, V.A.; Ishak, G.E.; Boyle, E.A.; Wilson, M.; Marques lourenço, C.; Arslan, M.; Shendure, J.; et al. KIAA0586 is mutated in Joubert syndrome. Hum. Mutat. 2015, 36, 831–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Naharros, I.O.; Cristian, F.B.; Zang, J.; Gesemann, M.; Ingham, P.W.; Neuhauss, S.C.F.; Bachmann-Gagescu, R. The ciliopathy protein TALPID3/KIAA0586 acts upstream of Rab8 activation in zebrafish photoreceptor outer segment formation and maintenance. Sci. Rep. 2018, 8, 1–13. [Google Scholar] [CrossRef] [Green Version]
  196. Edvardson, S.; Shaag, A.; Zenvirt, S.; Erlich, Y.; Hannon, G.J.; Shanske, A.L.; Gomori, J.M.; Ekstein, J.; Elpeleg, O. Joubert syndrome 2 (JBTS2) in Ashkenazi Jews is associated with a TMEM216 mutation. Am. J. Hum. Genet. 2010, 86, 93–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Valente, E.M.; Logan, C.V.; Mougou-Zerelli, S.; Lee, J.H.; Silhavy, J.L.; Brancati, F.; Iannicelli, M.; Travaglini, L.; Romani, S.; Illi, B.; et al. Mutations in TMEM216 perturb ciliogenesis and cause Joubert, Meckel and related syndromes. Nat. Genet. 2010, 42, 619–625. [Google Scholar] [CrossRef] [PubMed]
  198. Liu, Y.; Cao, S.; Yu, M.; Hu, H. TMEM216 deletion causes mislocalization of cone opsin and rhodopsin and photoreceptor degeneration in zebrafish. Investig. Ophthalmol. Vis. Sci. 2020, 61, 1–12. [Google Scholar] [CrossRef] [PubMed]
  199. Asai-Coakwell, M.; March, L.; Dai, X.H.; DuVal, M.; Lopez, I.; French, C.R.; Famulski, J.; De Baere, E.; Francis, P.J.; Sundaresan, P.; et al. Contribution of growth differentiation factor 6-dependent cell survival to early-onset retinal dystrophies. Hum. Mol. Genet. 2013, 22, 1432–1442. [Google Scholar] [CrossRef] [Green Version]
  200. Nadolski, N.J.; Balay, S.D.; Wong, C.X.L.; Waskiewicz, A.J.; Hocking, J.C. Abnormal cone and rod photoreceptor morphogenesis in gdf6a mutant zebrafish. Investig. Ophthalmol. Vis. Sci. 2020, 61. [Google Scholar] [CrossRef]
  201. Noor, A.; Windpassinger, C.; Patel, M.; Stachowiak, B.; Mikhailov, A.; Azam, M.; Irfan, M.; Siddiqui, Z.K.; Naeem, F.; Paterson, A.D.; et al. CC2D2A, encoding a coiled-coil and C2 domain protein, causes autosomal-recessive mental retardation with retinitis pigmentosa. Am. J. Hum. Genet. 2008, 82, 1011–1018. [Google Scholar] [CrossRef] [Green Version]
  202. Ferland, R.J.; Eyaid, W.; Collura, R.V.; Tully, L.D.; Hill, R.S.; Al-Nouri, D.; Al-Rumayyan, A.; Topcu, M.; Gascon, G.; Bodell, A.; et al. Abnormal cerebellar development and axonal decussation due to mutations in AHI1 in Joubert syndrome. Nat. Genet. 2004, 36, 1008–1013. [Google Scholar] [CrossRef] [Green Version]
  203. Cantagrel, V.; Silhavy, J.L.; Bielas, S.L.; Swistun, D.; Marsh, S.E.; Bertrand, J.Y.; Audollent, S.; Attié-Bitach, T.; Holden, K.R.; Dobyns, W.B.; et al. Mutations in the cilia gene ARL13B lead to the classical form of Joubert syndrome. Am. J. Hum. Genet. 2008, 83, 170–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Bachmann-Gagescu, R.; Phelps, I.G.; Stearns, G.; Link, B.A.; Brockerhoff, S.E.; Moens, C.B.; Doherty, D. The ciliopathy gene cc2d2a controls zebrafish photoreceptor outer segment development through a role in Rab8-dependent vesicle trafficking. Hum. Mol. Genet. 2011, 20, 4041–4055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Lessieur, E.M.; Fogerty, J.; Gaivin, R.J.; Song, P.; Perkins, B.D. The ciliopathy gene ahi1 is required for zebrafish cone photoreceptor outer segment morphogenesis and survival. Investig. Ophthalmol. Vis. Sci. 2017, 58, 448–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Song, P.; Dudinsky, L.; Fogerty, J.; Gaivin, R.; Perkins, B.D. Arl13b interacts with vangl2 to regulate cilia and photoreceptor outer segment length in zebrafish. Investig. Ophthalmol. Vis. Sci. 2016, 57, 4517–4526. [Google Scholar] [CrossRef] [PubMed]
  207. Den Hollander, A.I.; Koenekoop, R.K.; Mohamed, M.D.; Arts, H.H.; Boldt, K.; Towns, K.V.; Sedmak, T.; Beer, M.; Nagel-Wolfrum, K.; McKibbin, M.; et al. Mutations in LCA5, encoding the ciliary protein lebercilin, cause Leber congenital amaurosis. Nat. Genet. 2007, 39, 889–895. [Google Scholar] [CrossRef]
  208. Qu, Z.; Yimer, T.A.; Xie, S.; Wong, F.; Yu, S.; Liu, X.; Han, S.; Ma, J.; Lu, Z.; Hu, X.; et al. Knocking out lca5 in zebrafish causes cone-rod dystrophy due to impaired outer segment protein trafficking. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 2694–2705. [Google Scholar] [CrossRef]
  209. Nishimura, D.Y.; Baye, L.M.; Perveen, R.; Searby, C.C.; Avila-Fernandez, A.; Pereiro, I.; Ayuso, C.; Valverde, D.; Bishop, P.N.; Manson, F.D.C.; et al. Discovery and functional analysis of a retinitis pigmentosa gene, C2ORF71. Am. J. Hum. Genet. 2010, 86, 686–695. [Google Scholar] [CrossRef] [Green Version]
  210. Corral-Serrano, J.C.; Messchaert, M.; Dona, M.; Peters, T.A.; Kamminga, L.M.; Van Wijk, E.; Collin, R.W.J. C2orf71a/pcare1 is important for photoreceptor outer segment morphogenesis and visual function in zebrafish. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef]
  211. Abd El-Aziz, M.M.; Barragan, I.; O’Driscoll, C.A.; Goodstadt, L.; Prigmore, E.; Borrego, S.; Mena, M.; Pieras, J.I.; El-Ashry, M.F.; Safieh, L.A.; et al. EYS, encoding an ortholog of Drosophila spacemaker, is mutated in autosomal recessive retinitis pigmentosa. Nat. Genet. 2008, 40, 1285–1287. [Google Scholar] [CrossRef] [Green Version]
  212. Collin, R.W.J.; Littink, K.W.; Klevering, B.J.; van den Born, L.I.; Koenekoop, R.K.; Zonneveld, M.N.; Blokland, E.A.W.; Strom, T.M.; Hoyng, C.B.; den Hollander, A.I.; et al. Identification of a 2 Mb human ortholog of Drosophila eyes shut/spacemaker that Is mutated in patients with retinitis pigmentosa. Am. J. Hum. Genet. 2008, 83, 594–603. [Google Scholar] [CrossRef] [Green Version]
  213. Katagiri, S.; Akahori, M.; Hayashi, T.; Yoshitake, K.; Gekka, T.; Ikeo, K.; Tsuneoka, H.; Iwata, T. Autosomal recessive cone-rod dystrophy associated with compound heterozygous mutations in the EYS gene. Doc. Ophthalmol. 2014, 128, 211–217. [Google Scholar] [CrossRef] [PubMed]
  214. Alfano, G.; Kruczek, P.M.; Shah, A.Z.; Kramarz, B.; Jeffery, G.; Zelhof, A.C.; Bhattacharya, S.S. EYS is a protein associated with the ciliary axoneme in rods and cones. PLoS ONE 2016, 11, 1–20. [Google Scholar] [CrossRef] [PubMed]
  215. Yu, M.; Liu, Y.; Li, J.; Natale, B.N.; Cao, S.; Wang, D.; Amack, J.D.; Hu, H. Eyes shut homolog is required for maintaining the ciliary pocket and survival of photoreceptors in zebrafish. Biol. Open 2016, 5, 1662–1673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Messchaert, M.; Dona, M.; Broekman, S.; Peters, T.A.; Corral-Serrano, J.C.; Slijkerman, R.W.N.; van Wijk, E.; Collin, R.W.J. Eyes shut homolog is important for the maintenance of photoreceptor morphology and visual function in zebrafish. PLoS ONE 2018, 13, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Lu, Z.; Hu, X.; Liu, F.; Soares, D.C.; Liu, X.; Yu, S.; Gao, M.; Han, S.; Qin, Y.; Li, C.; et al. Ablation of EYS in zebrafish causes mislocalisation of outer segment proteins, F-actin disruption and cone-rod dystrophy. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef]
  218. Nishimura, D.Y.; Searby, C.C.; Carmi, R.; Elbedour, K.; Van Maldergem, L.; Fulton, A.B.; Lam, B.L.; Powell, B.R.; Swiderski, R.E.; Bugge, K.E.; et al. Positional cloning of a novel gene on chromosome 16q causing Bardet-Biedl syndrome (BBS2). Hum. Mol. Genet. 2001, 10, 865–874. [Google Scholar] [CrossRef] [Green Version]
  219. Shevach, E.; Ali, M.; Mizrahi-Meissonnier, L.; McKibbin, M.; El-Asrag, M.; Watson, C.M.; Inglehearn, C.F.; Ben-Yosef, T.; Blumenfeld, A.; Jalas, C.; et al. Association between missense mutations in the BBS2 gene and nonsyndromic retinitis pigmentosa. Jama Ophthalmol. 2015, 133, 312–318. [Google Scholar] [CrossRef] [Green Version]
  220. Song, P.; Fogerty, J.; Cianciolo, L.T.; Stupay, R.; Perkins, B.D. Cone photoreceptor degeneration and neuroinflammation in the zebrafish Bardet-Biedl syndrome 2 (bbs2) mutant does not lead to retinal regeneration. Front. Cell Dev. Biol. 2020, 8, 1–13. [Google Scholar] [CrossRef]
  221. Sergouniotis, P.I.; Davidson, A.E.; Mackay, D.S.; Li, Z.; Yang, X.; Plagnol, V.; Moore, A.T.; Webster, A.R. Recessive mutations in KCNJ13, encoding an inwardly rectifying potassium channel subunit, cause leber congenital amaurosis. Am. J. Hum. Genet. 2011, 89, 183–190. [Google Scholar] [CrossRef] [Green Version]
  222. Toms, M.; Burgoyne, T.; Tracey-White, D.; Richardson, R.; Dubis, A.M.; Webster, A.R.; Futter, C.; Moosajee, M. Phagosomal and mitochondrial alterations in RPE may contribute to KCNJ13 retinopathy. Sci. Rep. 2019, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  223. Xu, M.; Yamada, T.; Sun, Z.; Eblimit, A.; Lopez, I.; Wang, F.; Manya, H.; Xu, S.; Zhao, L.; Li, Y.; et al. Mutations in POMGNT1 cause non-syndromic retinitis pigmentosa. Hum. Mol. Genet. 2016, 25, 1479–1488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Liu, Y.; Yu, M.; Shang, X.; Nguyen, M.H.H.; Balakrishnan, S.; Sager, R.; Hu, H. Eyes shut homolog (EYS) interacts with matriglycan of O-mannosyl glycans whose deficiency results in EYS mislocalization and degeneration of photoreceptors. Sci. Rep. 2020, 10, 1–14. [Google Scholar] [CrossRef] [PubMed]
  225. Xie, S.; Han, S.; Qu, Z.; Liu, F.; Li, J.; Yu, S.; Reilly, J.; Tu, J.; Liu, X.; Lu, Z.; et al. Knockout of Nr2e3 prevents rod photoreceptor differentiation and leads to selective L-/M-cone photoreceptor degeneration in zebrafish. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 1273–1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Oel, A.P.; Neil, G.N.; Dong, E.M.; Balay, S.D.; Collett, K.; Allison, W.T. Nrl is dispensable for specification of rod photoreceptors in adult zebrafish contrasting a deeply conserved requirement earlier in ontogeny. iScience 2020, 23. [Google Scholar] [CrossRef]
  227. Maw, M.A.; Corbeil, D.; Koch, J.; Hellwig, A.; Wilson-Wheeler, J.C.; Bridges, R.J.; Kumaramanickavel, G.; John, S.; Nancarrow, D.; Röper, K.; et al. A frameshift mutation in prominin (mouse)-like 1 causes human retinal degeneration. Hum. Mol. Genet. 2000, 9, 27–34. [Google Scholar] [CrossRef]
  228. Zhang, Q.; Zulfiqar, F.; Xiao, X.; Riazuddin, S.A.; Ahmad, Z.; Caruso, R.; MacDonald, I.; Sieving, P.; Riazuddin, S.A.; Hejtmancik, J.F. Severe retinitis pigmentosa mapped to 4p15 and associated with a novel mutation in the PROM1 gene. Hum. Genet. 2007, 122, 293–299. [Google Scholar] [CrossRef]
  229. Yang, Z.; Chen, Y.; Lillo, C.; Chien, J.; Yu, Z.; Michaelides, M.; Klein, M.; Howes, K.A.; Li, Y.; Kaminoh, Y.; et al. Mutant prominin 1 found in patients with macular degeneration disrupts photoreceptor disk morphogenesis in mice. J. Clin. Investig. 2008, 118, 2908–2916. [Google Scholar] [CrossRef] [Green Version]
  230. Lu, Z.; Hu, X.; Reilly, J.; Jia, D.; Liu, F.; Yu, S.; Liu, X.; Xie, S.; Qu, Z.; Qin, Y.; et al. Deletion of the transmembrane protein Prom1b in zebrafish disrupts outer-segment morphogenesis and causes photoreceptor degeneration. J. Biol. Chem. 2019, 294, 13953–13963. [Google Scholar] [CrossRef]
  231. Lessieur, E.M.; Song, P.; Nivar, G.C.; Piccillo, E.M.; Fogerty, J.; Rozic, R.; Perkins, B.D. Ciliary genes arl13b, ahi1 and cc2d2a differentially modify expression of visual acuity phenotypes but do not enhance retinal degeneration due to mutation of cep290 in zebrafish. PLoS ONE 2019, 14, 1–26. [Google Scholar] [CrossRef] [Green Version]
  232. Thiadens, A.A.H.J.; den Hollander, A.I.; Roosing, S.; Nabuurs, S.B.; Zekveld-Vroon, R.C.; Collin, R.W.J.; De Baere, E.; Koenekoop, R.K.; van Schooneveld, M.J.; Strom, T.M.; et al. Homozygosity mapping reveals PDE6C mutations in patients with early-onset cone photoreceptor disorders. Am. J. Hum. Genet. 2009, 85, 240–247. [Google Scholar] [CrossRef] [Green Version]
  233. Sohocki, M.M.; Bowne, S.J.; Sullivan, L.S.; Blackshaw, S.; Cepko, C.L.; Payne, A.M.; Bhattacharya, S.S.; Khaliq, S.; Mehdi, S.Q.; Birch, D.G.; et al. Mutations in a new photoreceptor-pineal gene on 17p cause Leber congenital amaurosis. Nat. Genet. 2000, 24, 79–83. [Google Scholar] [CrossRef] [PubMed]
  234. Sohocki, M.M.; Perrault, I.; Leroy, B.P.; Payne, A.M.; Dharmaraj, S.; Bhattacharya, S.S.; Kaplan, J.; Maumenee, I.H.; Koenekoop, R.; Meire, F.M.; et al. Prevalence of AIPL1 mutations in inherited retinal degenerative disease. Mol. Genet. Metab. 2000, 70, 142–150. [Google Scholar] [CrossRef] [PubMed]
  235. Iribarne, M.; Nishiwaki, Y.; Nakamura, S.; Araragi, M.; Oguri, E.; Masai, I. Aipl1 is required for cone photoreceptor function and survival through the stability of Pde6c and Gc3 in zebrafish. Sci. Rep. 2017, 7, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Wycisk, K.A.; Zeitz, C.; Feil, S.; Wittmer, M.; Forster, U.; Neidhardt, J.; Wissinger, B.; Zrenner, E.; Wilke, R.; Kohl, S.; et al. Mutation in the auxiliary calcium-channel subunit CACNA2D4 causes autosomal recessive cone dystrophy. Am. J. Hum. Genet. 2006, 79, 973–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Daly, C.; Shine, L.; Heffernan, T.; Deeti, S.; Reynolds, A.L.; O’Connor, J.J.; Dillon, E.T.; Duffy, D.J.; Kolch, W.; Cagney, G.; et al. A brain-derived neurotrophic factor mimetic is sufficient to restore cone photoreceptor visual function in an inherited blindness model. Sci. Rep. 2017, 7, 1–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Vilardi, F.; Stephan, M.; Clancy, A.; Janshoff, A.; Schwappach, B. WRB and CAML are necessary and sufficient to mediate tail-anchored protein targeting to the ER membrane. PLoS ONE 2014, 9. [Google Scholar] [CrossRef]
  239. Daniele, L.L.; Emran, F.; Lobo, G.P.; Gaivin, R.J.; Perkins, B.D. Mutation of wrb, a component of the guided entry of Tail-Anchored protein pathway, disrupts photoreceptor synapse structure and function. Investig. Ophthalmol. Vis. Sci. 2016, 57, 2942–2954. [Google Scholar] [CrossRef] [Green Version]
  240. Lin, S.Y.; Vollrath, M.A.; Mangosing, S.; Shen, J.; Cardenas, E.; Corey, D.P. The zebrafish pinball wizard gene encodes WRB, a tail-anchored-protein receptor essential for inner-ear hair cells and retinal photoreceptors. J. Physiol. 2016, 594, 895–914. [Google Scholar] [CrossRef] [Green Version]
  241. Kohl, S.; Baumann, B.; Rosenberg, T.; Kellner, U.; Lorenz, B.; Vadalà, M.; Jacobson, S.G.; Wissinger, B. Mutations in the cone photoreceptor G-protein α-subunit gene GNAT2 in patients with achromatopsia. Am. J. Hum. Genet. 2002, 71, 422–425. [Google Scholar] [CrossRef] [Green Version]
  242. Brockerhoff, S.E.; Rieke, F.; Matthews, H.R.; Taylor, M.R.; Kennedy, B.; Ankoudinova, I.; Niemi, G.A.; Tucker, C.L.; Xiao, M.; Cilluffo, M.C.; et al. Light stimulates a transducin-independent increase of cytoplasmic Ca2+ and suppression of current in cones from the zebrafish mutant nof. J. Neurosci. 2003, 23, 470–480. [Google Scholar] [CrossRef] [Green Version]
  243. Bech-Hansen, N.T.; Naylor, M.J.; Maybaum, T.A.; Pearce, W.G.; Koop, B.; Fishman, G.A.; Mets, M.; Musarella, M.A.; Boycott, K.M. Loss-of-function mutations in a calcium-channel α1-subunit gene in Xp11.23 cause incomplete X-linked congenital stationary night blindness. Nat. Genet. 1998, 19, 264–267. [Google Scholar] [CrossRef]
  244. Strom, T.M.; Nyakatura, G.; Apfelstedt-Sylla, E.; Hellebrand, H.; Lorenz, B.; Weber, B.H.F.; Wutz, K.; Gutwillinger, N.; Rüther, K.; Drescher, B.; et al. An L-type calcium-channel gene mutated in incomplete X-linked congenital stationary night blindness. Nat. Genet. 1998, 19, 260–263. [Google Scholar] [CrossRef] [PubMed]
  245. Jalkanen, R.; Mänlyjärvi, M.; Tobias, R.; Isosomppi, J.; Sankila, E.M.; Alitalo, T.; Bech-Hansen, N.T. X linked cone-rod dystrophy, CORDX3, is caused by a mutation in the CACNA1F gene. J. Med. Genet. 2006, 43, 699–704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Baumann, L.; Gerstner, A.; Zong, X.; Biel, M.; Wahl-Schott, C. Functional characterization of the L-type Ca2+ channel Cav1.4α1 from mouse retina. Investig. Ophthalmol. Vis. Sci. 2004, 45, 708–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Van Epps, H.A.; Hayashi, M.; Lucast, L.; Stearns, G.W.; Hurley, J.B.; De Camilli, P.; Brockerhoff, S.E. The zebrafish nrc mutant reveals a role for the polyphosphoinositide phosphatase synaptojanin 1 in cone photoreceptor ribbon anchoring. J. Neurosci. 2004, 24, 8641–8650. [Google Scholar] [CrossRef] [Green Version]
  248. Allwardt, B.A.; Lall, A.B.; Brockerhoff, S.E.; Dowling, J.E. Synapse formation is arrested in retinal photoreceptors of the zebrafish nrc mutant. J. Neurosci. 2001, 21, 2330–2342. [Google Scholar] [CrossRef] [Green Version]
  249. Holzhausen, L.C.; Lewis, A.A.; Cheong, K.K.; Brockerhoff, S.E. Differential role for synaptojanin 1 in rod and cone photoreceptors. J. Comp. Neurol. 2009, 517, 633–644. [Google Scholar] [CrossRef] [Green Version]
  250. Huang, D.F.; Wang, M.Y.; Yin, W.U.; Ma, Y.Q.; Wang, H.A.N.; Xue, T.; Ren, D.L.; Hu, B. Zebrafish lacking circadian gene per2 exhibit visual function deficiency. Front. Behav. Neurosci. 2018, 12, 1–10. [Google Scholar] [CrossRef] [Green Version]
  251. Van Rooijen, E.; Voest, E.E.; Logister, I.; Bussmann, J.; Korving, J.; Van Eeden, F.J.; Giles, R.H.; Schulte-Merker, S. Von Hippel-Lindau tumor suppressor mutants faithfully model pathological hypoxia-driven angiogenesis and vascular retinopathies in zebrafish. Dmm Dis. Model. Mech. 2010, 3, 343–353. [Google Scholar] [CrossRef] [Green Version]
  252. Schori, C.; Agbaga, M.P.; Brush, R.S.; Ayyagari, R.; Grimm, C.; Samardzija, M. Elovl4 5-bp deletion does not accelerate cone photoreceptor degeneration in an all-cone mouse. PLoS ONE 2018, 13, 1–16. [Google Scholar] [CrossRef] [Green Version]
  253. Zhao, L.; Zabel, M.K.; Wang, X.; Ma, W.; Shah, P.; Fariss, R.N.; Qian, H.; Parkhurst, C.N.; Gan, W.; Wong, W.T. Microglial phagocytosis of living photoreceptors contributes to inherited retinal degeneration. Embo Mol. Med. 2015, 7, 1179–1197. [Google Scholar] [CrossRef] [PubMed]
  254. Sancho-Pelluz, J.; Alavi, M.V.; Sahaboglu, A.; Kustermann, S.; Farinelli, P.; Azadi, S.; Van Veen, T.; Romero, F.J.; Paquet-Durand, F.; Ekström, P. Excessive HDAC activation is critical for neurodegeneration in the rd1 mouse. Cell Death Dis. 2010, 1, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Trifunović, D.; Arango-Gonzalez, B.; Comitato, A.; Barth, M.; del Amo, E.M.; Kulkarni, M.; Sahaboglu, A.; Hauck, S.M.; Urtti, A.; Arsenijevic, Y.; et al. HDAC inhibition in the cpfl1 mouse protects degenerating cone photoreceptors in vivo. Hum. Mol. Genet. 2016, 25, 4462–4472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Lai, J.I.; Leman, L.J.; Ku, S.; Vickers, C.J.; Olsen, C.A.; Montero, A.; Ghadiri, M.R.; Gottesfeld, J.M. Cyclic tetrapeptide HDAC inhibitors as potential therapeutics for spinal muscular atrophy: Screening with iPSC-derived neuronal cells. Bioorganic Med. Chem. Lett. 2017, 27, 3289–3293. [Google Scholar] [CrossRef]
  257. Janczura, K.J.; Volmar, C.H.; Sartor, G.C.; Rao, S.J.; Ricciardi, N.R.; Lambert, G.; Brothers, S.P.; Wahlestedt, C. Inhibition of HDAC3 reverses Alzheimer’s disease-related pathologies in vitro and in the 3xTg-AD mouse model. Proc. Natl. Acad. Sci. USA 2018, 115, E11148–E11157. [Google Scholar] [CrossRef] [Green Version]
  258. Kawase, R.; Nishimura, Y.; Ashikawa, Y.; Sasagawa, S.; Murakami, S.; Yuge, M.; Okabe, S.; Kawaguchi, K.; Yamamoto, H.; Moriyuki, K.; et al. EP300 protects from light-induced retinopathy in zebrafish. Front. Pharm. 2016, 7, 1–13. [Google Scholar] [CrossRef] [Green Version]
  259. D’Ydewalle, C.; Krishnan, J.; Chiheb, D.M.; Van Damme, P.; Irobi, J.; Kozikowski, A.P.; Berghe, P.V.; Timmerman, V.; Robberecht, W.; Van Den Bosch, L. HDAC6 inhibitors reverse axonal loss in a mouse model of mutant HSPB1-induced Charcot-Marie-Tooth disease. Nat. Med. 2011, 17, 968–974. [Google Scholar] [CrossRef]
  260. Sundaramurthi, H.; Roche, S.L.; Grice, G.L.; Moran, A.; Dillion, E.T.; Campiani, G.; Nathan, J.A.; Kennedy, B.N. Selective histone deacetylase 6 inhibitors restore cone photoreceptor vision or outer segment morphology in zebrafish and mouse models of retinal blindness. Front. Cell Dev. Biol. 2020, 8. [Google Scholar] [CrossRef]
  261. Leyk, J.; Daly, C.; Janssen-Bienhold, U.; Kennedy, B.N.; Richter-Landsberg, C. HDAC6 inhibition by tubastatin A is protective against oxidative stress in a photoreceptor cell line and restores visual function in a zebrafish model of inherited blindness. Cell Death Dis. 2017, 8, e3028. [Google Scholar] [CrossRef] [Green Version]
  262. Kokel, D.; Bryan, J.; Laggner, C.; White, R.; Cheung, C.Y.J.; Mateus, R.; Healey, D.; Kim, S.; Werdich, A.A.; Haggarty, S.J.; et al. Rapid behavior-based identification of neuroactive small molecules in the zebrafish. Nat. Chem. Biol. 2010, 6, 231–237. [Google Scholar] [CrossRef] [Green Version]
  263. Rihel, J.; Prober, D.A.; Arvanites, A.; Lam, K.; Zimmerman, S.; Jang, S.; Haggarty, S.J.; Kokel, D.; Rubin, L.L.; Peterson, R.T.; et al. Zebrafish behavioral profiling links drugs to biological targets and rest/wake regulation. Science 2010, 327, 348–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Ganzen, L.; Ko, M.; Zhang, M.; Xie, R.; Chen, Y.; Zhang, L.; James, R.; Mumm, J.; van Rijn, R.; Zhong, W.; et al. Drug screening with zebrafish visual behavior identifies carvedilol as a potential treatment for retinitis pigmentosa. bioRxiv 2020. [Google Scholar] [CrossRef]
  265. Viringipurampeer, I.A.; Shan, X.; Gregory-Evans, K.; Zhang, J.P.; Mohammadi, Z.; Gregory-Evans, C.Y. Rip3 knockdown rescues photoreceptor cell death in blind pde6c zebrafish. Cell Death Differ. 2014, 21, 665–675. [Google Scholar] [CrossRef] [Green Version]
  266. Giridharan, V.V.; Thandavarayan, R.A.; Sato, S.; Ko, K.M.; Konishi, T. Prevention of scopolamine-induced memory deficits by schisandrin B, an antioxidant lignan from Schisandra chinensis in mice. Free Radic. Res. 2011, 45, 950–958. [Google Scholar] [CrossRef] [PubMed]
  267. Ko, K.M.; Lam, B.Y.H. Schisandrin B protects against tert-butylhydroperoxide induced cerebral toxicity by enhancing glutathione antioxidant status in mouse brain. Mol. Cell. Biochem. 2002, 238, 181–186. [Google Scholar] [CrossRef] [PubMed]
  268. Zhang, L.; Xiang, L.; Liu, Y.; Venkatraman, P.; Chong, L.; Cho, J.; Bonilla, S.; Jin, Z.B.; Pang, C.P.; Ko, K.M.; et al. A naturally-derived compound schisandrin B enhanced light sensation in the pde6c zebrafish model of retinal degeneration. PLoS ONE 2016, 11, 1–19. [Google Scholar] [CrossRef] [Green Version]
  269. Chen, N.; Chiu, P.Y.; Ko, K.M. Schisandrin B enhances cerebral mitochondrial antioxidant status and structural integrity, and protects against cerebral ischemia/reperfusion injury in rats. Biol. Pharm. Bull. 2008, 31, 1387–1391. [Google Scholar] [CrossRef] [Green Version]
  270. Ji, X.; Shen, Y.; Guo, X. Isolation, structures, and bioactivities of the polysaccharides from Gynostemma pentaphyllum (Thunb.) Makino: A review. Biomed. Res. Int. 2018, 2018. [Google Scholar] [CrossRef] [Green Version]
  271. Alhasani, R.H.; Zhou, X.; Biswas, L.; Li, X.; Reilly, J.; Zeng, Z.; Shu, X. Gypenosides attenuate retinal degeneration in a zebrafish retinitis pigmentosa model. Exp. Eye Res. 2020, 201, 108291. [Google Scholar] [CrossRef]
  272. Deeti, S.; O’Farrell, S.; Kennedy, B.N. Early safety assessment of human oculotoxic drugs using the zebrafish visualmotor response. J. Pharm. Toxicol. Methods 2014, 69, 1–8. [Google Scholar] [CrossRef]
  273. Vaché, C.; Besnard, T.; le Berre, P.; García-García, G.; Baux, D.; Larrieu, L.; Abadie, C.; Blanchet, C.; Bolz, H.J.; Millan, J.; et al. Usher syndrome type 2 caused by activation of an USH2A pseudoexon: Implications for diagnosis and therapy. Hum. Mutat. 2012, 33, 104–108. [Google Scholar] [CrossRef] [PubMed]
  274. Slijkerman, R.; Goloborodko, A.; Broekman, S.; de Vrieze, E.; Hetterschijt, L.; Peters, T.; Gerits, M.; Kremer, H.; van Wijk, E. Poor splice-site recognition in a humanized zebrafish knockin model for the recurrent deep-intronic c.7595-2144A>G mutation in USH2A. Zebrafish 2018, 15, 597–609. [Google Scholar] [CrossRef] [PubMed]
  275. Slijkerman, R.; van Diepen, H.; Albert, S.; Dona, M.; Venselaar, H.; Zang, J.; Neuhauss, S.; Peters, T.; Broekman, S.; Pennings, R.; et al. Antisense oligonucleotide-based treatment of retinitis pigmentosa caused by mutations in USH2A exon 13. bioRxiv 2020. [Google Scholar] [CrossRef]
  276. Baux, D.; Blanchet, C.; Hamel, C.; Meunier, I.; Larrieu, L.; Faugère, V.; Vaché, C.; Castorina, P.; Puech, B.; Bonneau, D.; et al. Enrichment of LOVD-USHbases with 152 USH2A genotypes defines an extensive mutational spectrum and highlights missense hotspots. Hum. Mutat. 2014, 35, 1179–1186. [Google Scholar] [CrossRef]
  277. Cameron, D.A. Cellular proliferation and neurogenesis in the injured retina of adult zebrafish. Vis. Neurosci. 2000, 17, 789–797. [Google Scholar] [CrossRef]
  278. Wan, Y.; Almeida, A.D.; Rulands, S.; Chalour, N.; Muresan, L.; Wu, Y.; Simons, B.D.; He, J.; Harris, W.A. The ciliary marginal zone of the zebrafish retina: Clonal and time-lapse analysis of a continuously growing tissue. Development 2016, 143, 1099–1107. [Google Scholar] [CrossRef] [Green Version]
  279. Yurco, P.; Cameron, D.A. Responses of Muller glia to retinal injury in adult zebrafish. Vis. Res. 2005, 45, 991–1002. [Google Scholar] [CrossRef] [Green Version]
  280. Marcucci, F.; Murcia-Belmonte, V.; Wang, Q.; Coca, Y.; Ferreiro-Galve, S.; Kuwajima, T.; Khalid, S.; Ross, M.E.; Mason, C.; Herrera, E. The ciliary margin zone of the mammalian retina generates retinal ganglion cells. Cell Rep. 2016, 17, 3153–3164. [Google Scholar] [CrossRef]
  281. Bhatia, B.; Jayaram, H.; Singhal, S.; Jones, M.F.; Limb, G.A. Differences between the neurogenic and proliferative abilities of Müller glia with stem cell characteristics and the ciliary epithelium from the adult human eye. Exp. Eye Res. 2011, 93, 852–861. [Google Scholar] [CrossRef] [Green Version]
  282. Ahmad, I.; Tang, L.; Pham, H. Identification of neural progenitors in the adult mammalian eye. Biochem. Biophys. Res. Commun. 2000, 270, 517–521. [Google Scholar] [CrossRef]
  283. Bhatia, B.; Singhal, S.; Lawrence, J.M.; Khaw, P.T.; Limb, G.A. Distribution of Müller stem cells within the neural retina: Evidence for the existence of a ciliary margin-like zone in the adult human eye. Exp. Eye Res. 2009, 89, 373–382. [Google Scholar] [CrossRef] [PubMed]
  284. Fleisch, V.C.; Fraser, B.; Allison, W.T. Investigating regeneration and functional integration of CNS neurons: Lessons from zebrafish genetics and other fish species. Biochim. Biophys. Acta Mol. Basis Dis. 2011, 1812, 364–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Nagashima, M.; Barthel, L.K.; Raymond, P.A. A self-renewing division of zebrafish Muller glial cells generates neuronal progenitors that require N-cadherin to regenerate retinal neurons. Development 2013, 140, 4510–4521. [Google Scholar] [CrossRef] [Green Version]
  286. Fausett, B.V.; Goldman, D. A role for alpha1 tubulin-expressing Muller glia in regeneration of the injured zebrafish retina. J. Neurosci. 2006, 26, 6303–6313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Otteson, D.C.; Hitchcock, P.F. Stem cells in the teleost retina: Persistent neurogenesis and injury-induced regeneration. Vis. Res. 2003, 43, 927–936. [Google Scholar] [CrossRef] [Green Version]
  288. Ramachandran, R.; Fausett, B.V.; Goldman, D. Ascl1a regulates Müller glia dedifferentiation and retinal regeneration through a Lin-28-dependent, let-7 microRNA signalling pathway. Nat. Cell Biol. 2010, 12, 1101–1107. [Google Scholar] [CrossRef]
  289. Nelson, C.M.; Ackerman, K.M.; O’Hayer, P.; Bailey, T.J.; Gorsuch, R.A.; Hyde, D.R. Tumor necrosis factor-alpha is produced by dying retinal neurons and is required for Müller glia proliferation during zebrafish retinal regeneration. J. Neurosci. 2013, 33, 6524–6539. [Google Scholar] [CrossRef] [Green Version]
  290. Lawrence, J.M.; Singhal, S.; Bhatia, B.; Keegan, D.J.; Reh, T.A.; Luthert, P.J.; Khaw, P.T.; Limb, G.A. MIO-M1 cells and similar muller glial cell lines derived from adult human retina exhibit neural stem cell characteristics. Stem Cells 2007, 25, 2033–2043. [Google Scholar] [CrossRef]
  291. Bringmann, A.; Pannicke, T.; Grosche, J.; Francke, M.; Wiedemann, P.; Skatchkov, S.N.; Osborne, N.N.; Reichenbach, A. Müller cells in the healthy and diseased retina. Prog. Retin. Eye Res. 2006, 25, 397–424. [Google Scholar] [CrossRef]
  292. Giannelli, S.G.; Demontis, G.C.; Pertile, G.; Rama, P.; Broccoli, V. Adult human Muller glia cells are a highly efficient source of rod photoreceptors. Stem Cells 2011, 29, 344–356. [Google Scholar] [CrossRef]
  293. Jayaram, H.; Jones, M.F.; Eastlake, K.; Cottrill, P.B.; Becker, S.; Wiseman, J.; Khaw, P.T.; Limb, G.A. Transplantation of photoreceptors derived from human Muller glia restore rod function in the P23H rat. Stem Cells Transl Med. 2014, 3, 323–333. [Google Scholar] [CrossRef] [PubMed]
  294. Lenkowski, J.R.; Raymond, P.A. Müller glia: Stem cells for generation and regeneration of retinal neurons in teleost fish. Prog. Retin. Eye Res. 2014, 40, 94–123. [Google Scholar] [CrossRef] [PubMed]
  295. Qian, C.; Dong, B.; Wang, X.-Y.; Zhou, F.-Q. In vivo glial trans-differentiation for neuronal replacement and functional recovery in central nervous system. Febs J. 2020. [Google Scholar] [CrossRef] [PubMed]
  296. Goldman, D. Müller glial cell reprogramming and retina regeneration. Nat. Rev. Neurosci. 2014, 15, 431–442. [Google Scholar] [CrossRef] [Green Version]
  297. Wan, J.; Goldman, D. Retina regeneration in zebrafish. Curr. Opin. Genet. Dev. 2016, 40, 41–47. [Google Scholar] [CrossRef] [Green Version]
  298. Madelaine, R.; Mourrain, P. Endogenous retinal neural stem cell reprogramming for neuronal regeneration. Neural Regen. Res. 2017, 12, 1765–1767. [Google Scholar] [CrossRef]
Figure 1. Anatomy of rod and cone photoreceptors and their organization in the human retina. (A) Cartoons of a rod (grey) and cone (blue) photoreceptor. Photoreceptors have an outer segment (OS) that is packed with light-sensitive opsin proteins, a connecting cilium (CC) that connects the OS with the mitochondria-rich inner segment (IS), a cell body (CB), and a synapse. (B) Cartoon of the human photoreceptor mosaic. Humans have three types of cones: red, green, and blue, depicted in those respective colours. The peripheral retina is rod-dense with cones interspersed throughout, while the central retina is cone-dense.
Figure 1. Anatomy of rod and cone photoreceptors and their organization in the human retina. (A) Cartoons of a rod (grey) and cone (blue) photoreceptor. Photoreceptors have an outer segment (OS) that is packed with light-sensitive opsin proteins, a connecting cilium (CC) that connects the OS with the mitochondria-rich inner segment (IS), a cell body (CB), and a synapse. (B) Cartoon of the human photoreceptor mosaic. Humans have three types of cones: red, green, and blue, depicted in those respective colours. The peripheral retina is rod-dense with cones interspersed throughout, while the central retina is cone-dense.
Biomolecules 11 00078 g001
Figure 2. Zebrafish photoreceptor organization. (A) Cartoon depiction of the adult zebrafish photoreceptor mosaic. UV and blue cones, depicted in purple and blue respectively, alternate in their rows while red and green double cones alternate in neighbouring rows. Rods are studded throughout. (B) Fluorescent image of a flat-mounted adult transgenic zebrafish retina, with GFP expressed in UV cones (magenta) and mCherry expressed in blue cones (cyan). The alternation of UV and blue cones in their rows is apparent. (C) Immunofluorescent image of a cryosectioned adult Tg(sws1:GFP) zebrafish retina with GFP in UV cones (green). UV opsin (red) and rhodopsin (magenta) are labelled, as well as nuclei (cyan).
Figure 2. Zebrafish photoreceptor organization. (A) Cartoon depiction of the adult zebrafish photoreceptor mosaic. UV and blue cones, depicted in purple and blue respectively, alternate in their rows while red and green double cones alternate in neighbouring rows. Rods are studded throughout. (B) Fluorescent image of a flat-mounted adult transgenic zebrafish retina, with GFP expressed in UV cones (magenta) and mCherry expressed in blue cones (cyan). The alternation of UV and blue cones in their rows is apparent. (C) Immunofluorescent image of a cryosectioned adult Tg(sws1:GFP) zebrafish retina with GFP in UV cones (green). UV opsin (red) and rhodopsin (magenta) are labelled, as well as nuclei (cyan).
Biomolecules 11 00078 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Noel, N.C.L.; MacDonald, I.M.; Allison, W.T. Zebrafish Models of Photoreceptor Dysfunction and Degeneration. Biomolecules 2021, 11, 78. https://doi.org/10.3390/biom11010078

AMA Style

Noel NCL, MacDonald IM, Allison WT. Zebrafish Models of Photoreceptor Dysfunction and Degeneration. Biomolecules. 2021; 11(1):78. https://doi.org/10.3390/biom11010078

Chicago/Turabian Style

Noel, Nicole C. L., Ian M. MacDonald, and W. Ted Allison. 2021. "Zebrafish Models of Photoreceptor Dysfunction and Degeneration" Biomolecules 11, no. 1: 78. https://doi.org/10.3390/biom11010078

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