Next Article in Journal
Distribution of Polyphenolic and Isoprenoid Compounds and Biological Activity Differences between in the Fruit Skin + Pulp, Seeds, and Leaves of New Biotypes of Elaeagnusmultiflora Thunb
Previous Article in Journal
Nrf2 Regulates Anti-Inflammatory A20 Deubiquitinase Induction by LPS in Macrophages in Contextual Manner
Previous Article in Special Issue
Expression of Pro-Angiogenic Markers Is Enhanced by Blue Light in Human RPE Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms Related to Oxidative Stress in Retinitis Pigmentosa

1
Department of Medical Sciences, University of Ferrara, 44121 Ferrara, Italy
2
Department of Specialized Surgery, Section of Ophthalmology, Sant’Anna University Hospital, 44121 Ferrara, Italy
3
Department of Head and Neck, Section of Ophthalmology, San Paolo Hospital, 17100 Savona, Italy
4
Department of Neuroscience and Rehabilitation, Section of Ophthalmology, University of Ferrara, 44121 Ferrara, Italy
*
Author to whom correspondence should be addressed.
Authors equally contributed.
Antioxidants 2021, 10(6), 848; https://doi.org/10.3390/antiox10060848
Submission received: 9 February 2021 / Revised: 13 May 2021 / Accepted: 20 May 2021 / Published: 26 May 2021

Abstract

:
Retinitis pigmentosa (RP) is an inherited retinopathy. Nevertheless, non-genetic biological factors play a central role in its pathogenesis and progression, including inflammation, autophagy and oxidative stress. The retina is particularly affected by oxidative stress due to its high metabolic rate and oxygen consumption as well as photosensitizer molecules inside the photoreceptors being constantly subjected to light/oxidative stress, which induces accumulation of ROS in RPE, caused by damaged photoreceptor’s daily recycling. Oxidative DNA damage is a key regulator of microglial activation and photoreceptor degeneration in RP, as well as mutations in endogenous antioxidant pathways involved in DNA repair, oxidative stress protection and activation of antioxidant enzymes (MUTYH, CERKL and GLO1 genes, respectively). Moreover, exposure to oxidative stress alters the expression of micro-RNA (miRNAs) and of long non-codingRNA (lncRNAs), which might be implicated in RP etiopathogenesis and progression, modifying gene expression and cellular response to oxidative stress. The upregulation of the P2X7 receptor (P2X7R) also seems to be involved, causing pro-inflammatory cytokines and ROS release by macrophages and microglia, contributing to neuroinflammatory and neurodegenerative progression in RP. The multiple pathways analysed demonstrate that oxidative microglial activation may trigger the vicious cycle of non-resolved neuroinflammation and degeneration, suggesting that microglia may be a key therapy target of oxidative stress in RP.

1. Introduction

Retinitis pigmentosa (RP) is an inherited retinal degeneration caused by a collection of different genetic mutations, most of which are related to the progressive loss of photoreceptors (both rod and cone cells) and retinal pigmented epithelium (RPE) dysfunction [1]. This significant genetic burden is also associated with an important phenotypic heterogeneity, due to different penetrance, expressivity [2] and interaction with oxidative stress factors.
Cellular homeostasis requires a balance between oxidative species and antioxidant defence mechanisms. An excessive production of reactive oxygen species (ROS) and free radicals leads to oxidative stress, causing cellular dysfunction, necrosis, apoptosis or autophagic cell death [3,4].
The retina is particularly affected by oxidative stress due to different factors: high oxygen consumption related to a high metabolic demand, the presence of photoreceptors, which contain photosensitizer molecules such as polyunsaturated fatty acids, and the accumulation of ROS in RPE, as an effect of damaged photoreceptor’s daily recycling [5].
Due to high metabolic activity, the retinal tissue has developed multiple defence mechanisms against oxidative damage. Oxidative stress (OS) is related to the activation of specific molecular pathways, such as PERK (PKR-like endoplasmic reticulum kinase) and IRE1 (inositol-requiring enzyme 1), that promote the transcription of genes encoding antioxidant enzymes. However, the chronic activation of some of these pathways causes mitochondrial damage in the long term, with intracellular accumulation of ROS resulting in retinal damage [6].
On the other side, inactivating mutations of genes involved in endogenous antioxidant defences leads to a rapid progression of the disease. In fact, in this review, we also analysed the role of GLO1 (Glyoxalase 1) and CERKL (ceramide-kinase like): mutations in these genes increase the sensitivity of retinal tissue to oxidative damage, resulting in cellular apoptosis and retinal neuro-degeneration due to the lack of resilience of the retinal tissue exposed to oxidative stress [7].
Among the defence mechanisms, the role of autophagy was also highlighted, prevalent in the retinal ganglion cells (RGC), which have a dual effect: reduction of intracellular ROS levels and mitochondrial support [8,9].
In a study conducted by Rodriguez-Muela et al. using a rd10 mouse model, it was shown that the calcium overload, the activation of calpain, the increase in cathepsin B activity, the reduced colocalization of cathepsin B with lysosomal markers and the reduction in the autophagosomal marker LC3-II (lipidated form of LC3-microtubule-associated protein 1A/1B-light chain 3) expression lead to an increase in permeability of the lysosomal membrane. All these changes in cellular activity, which occur before the death of photoreceptors, are markers of lysosomal dysfunction and down-regulation of autophagic activity [10].
Indeed, recent studies have discovered that the exposure to oxidative stress determines altered expression of micro-RNA (miRNAs) and of long non-coding RNA (lncRNAs) that might be implicated in the etiopathogenesis and progression of RP, since they alter gene expression along with the cellular response to oxidative stress [11].
This review focuses on the molecular mechanisms related to oxidative stress occurring in RP. These mechanisms play a central role in RP pathogenesis and progression.

Association between RP-Causative Mutations and Activation of UPR, PERK and IRE1 as a Response to Oxidative Stress

In response to environmental stress, such as OS, the translation of stress-inducible transcripts encoding heat shock proteins (HSP) is enhanced [3]. HSP perform chaperone functions by ensuring the correct folding of new proteins or refolding proteins that are damaged by OS.
A pathway that is responsible for cell translation reprogramming upon stress is PERK, a protein that is active in response to the accumulation of misfolded proteins [4].
Genetic mutations that were discovered to be causative of RP are involved in misfolding of transmembrane proteins implicated in photoreception and phototransduction. The accumulation of misfolded or unfolded proteins causes an increase in ROS, enhancing oxidative stress and the activation of an unfolded protein response (UPR), PERK and IRE1 pathways in photoreceptors cells. These pathways, involved in endogenous antioxidant defence, if chronically stimulated, determine the activation of pro-apoptotic programs associated with oxidative stress, pro-inflammatory signalling, dysfunctional autophagy, free cytosolic Ca2+ overload and an altered protein synthesis rate in the retina [4], leading to retinal degeneration [5] (Table 1).

2. Mutations in Endogenous Antioxidant Pathways: MUTYH, CERKL and GLO1

2.1. Role of 8-Oxoguanine and MUTYH in RP

One of the most prevalent genotoxic lesions is 8-oxoguanine (8-oxoG), and it is generated in DNA attacked by ROS [12]. MUTYH (mutY DNA glycosylase) plays an important role in the maintenance of genomic integrity through the activation of pathways involved in DNA repair. It removes adenine (A) from 8-oxoG:A mispairs, through the base excision repair (BER) pathway, preventing mutations in the genome [13].
MUTYH deficiency prevents single-strand breaks (SSBs) formation and cell death under oxidative stress [14,15].
However, under severe oxidative DNA damage, the excessive activation of MUTYH leads to formation of SSBs of DNA, causing disturbed homeostasis and cell death [16].
MUTYH-mediated BER is critical to promote retinal degeneration and inflammation in RP. As demonstrated by Oka et al. in a rd 10 mice model, it occurs through two different pathways, (i) mitochondrial SSBs mediate calpain activation and (ii) nuclear SSBs induce poly (ADP-ribose) polymerase (PARP) activation [17].
The accumulation of 8-oxoG in the rd10 mouse retina is attenuated by hMTH1 over-expression and the activation of MUTYH, which leads to a reduction of rod and cone photoreceptor cell death [18], as well as microgliosis in rd10 mice [17]. Oxidized nucleic acids are increased in the photoreceptor layer but also in immune cells, such as microglial cells and macrophages, which infiltrate outer retinal regions. These findings suggest that oxidative DNA damage in the outer nuclear layer (ONL) of rd10 mice is derived from the oxidized nucleotide pool that occurs in photoreceptor cells and in other non-neuronal proliferating cells in the ONL, such as microglia, that may incorporate oxidized nucleic acids into their nuclear DNA during retinal degeneration, suggesting that these inflammatory cells could be an alternative source of ROS in RP [19].
In microglia, nuclear accumulation of 8-oxoG is associated with PARP activation occurring before the peak of photoreceptor degeneration; thereafter, it expands to the photoreceptor nuclei along with microglial activation [17]. Therefore, oxidative microglial activation may trigger the vicious cycle of non-resolved neuroinflammation and degeneration in RP as it happens in the brain [20], suggesting that the microglia, and especially the MUTYH-SSBs-PARP pathway, may be a key target of oxidative stress in RP (Table 1).

2.2. Oxidative Stress in RP: The Role of CERKL

CERKL (ceramide-kinase like) is a gene involved in the oxidative stress protection [21]. The precise function of CERKL is yet to be determined, but many studies show that it is implicated in the cellular response to oxidative stress and may play a role in protecting cells against stress injury [22,23]. The name of the gene stems from the diacylglycerol kinase domain, which shares homology with ceramide kinases.
Ceramide is a chore sphingolipid (SL), a precursor of other bioactive and complex SLs lipid secondary messengers that control cell status [24,25] and plays a key role in stress-induced apoptosis [24].
To avoid entering apoptosis, induced by the increase of ceramide, cells activate enzymatic pathways involved in its clearance. In this context, the phosphorylation of ceramide, by ceramide kinase, produces a protective effect against apoptosis. Overexpression of CERK-L isoforms protects cells from apoptosis induced by oxidative stress [21].
Mutations in CERKL, that have been reported to cause distinct RP [26,27], with characteristic macular and peripheral lesions and other cone-rod dystrophy (CRD), support the concept that failure in the endogenous mechanisms to overcome oxidative stress leads to an accelerated progression of retinal neurodegeneration (Table 1).

2.3. Glyoxalase 1 (GLO1) Related Genes and Pathways

Glyoxalase 1 (GLO1) is a ubiquitous cellular enzyme involved in detoxification of cytotoxic products of glycolysis, such as α-oxoaldehydes, methylglyoxal (MG), glyoxal (GO) and 3-deoxyglucosone (3-DG). In detail, GLO1 metabolizes MG and prevents MG-induced damage. An excess of MG inactivates antioxidant enzymes, such as glutathione peroxidase and superoxide dismutase (SOD) enzymes, impairing the degradation of MG, determining a positive feedback loop [28].
These cytotoxic products’ levels are increased in cells undergoing hyperglycemic metabolism, such as the RPE cells and photoreceptors, representing the principal source of intra- and extracellular advanced glycation end products (AGEs) [29]. High levels of AGEs, in addition to ROS, determine hyperinflammation and permanent tissue damage [30]. Intracellular AGE precursors, such as MG and GO, can also modify and inhibit the function of important enzymes, such as glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and GLO1 [31].
AGEs exert their injuring effects by direct glycation of intracellular proteins and lipids, by the activation of cell signalling pathways through their binding to cellular receptors and the modulation of gene expression [32].
In the retina, AGEs could alter intra- and extracellular protein structure, increase inflammation and oxidative stress and, therefore, promote vascular dysfunction [33]. The retinal AGEs deposition could cause an upregulation of vascular endothelial growth factor (VEGF), a downregulation of pigment epithelium-derived factor (PEDF) and, eventually, a significant disruption of the inner blood–retinal barrier (iBRB) [34]. Additionally, increased advanced lipoxidation end-products (ALEs) accumulation was also detected in the outer retina. This portion contains photoreceptors, mainly rich in polyunsaturated fatty acids and therefore highly susceptible to lipid peroxidation [35].
In the retina, RPE cells exert the activity of protection by oxidative stress [36,37]. Numerous studies have validated the presence of high levels of ROS and AGEs in RPE, which are able to alter transduction pathways and gene expression [38].
GLO1 mutations, which are part of the endogenous detoxification system regulating ROS and AGE levels, contribute to the accumulation of AGEs in the retina, playing a role in RP pathogenesis [7,39] (Table 1).
Recent studies have identified 22 GLO1-related genes, with their related pathways, to be involved in a complex network of intracellular biochemical mechanisms that might be associated with RP onset and progression. Such pathways include microtubules and actin assembly, ubiquitin-proteasome activity, RE and Golgi integrity, vesicular trafficking, transcriptional and translational control, glycolytic metabolism regulation and glycosylation modifications [40].
In fact, the global down-expression of these genes, excluding the upregulation of AUTS2 (cytoplasmatic activator of transcription and developmental regulator) and ANKH (progressive ankylosis protein homolog), could mostly lead to impairment in cell polarity and adhesion, through the alteration of actin filament structure and activity, with the final result being an increase in RPE apoptosis. Some of these genes are also involved in cell death through the dysregulation of energy metabolism or the translation machinery (SIK3, IPO3 and MRPS33).
The main mutations involving genes that compromise cellular metabolism, leading to RPE dysfunction and photoreceptor damage, which could accelerate retinal degeneration in RP, are explained below and summarized in Table 2.

2.3.1. AUTS2

Cytoplasmic activator of transcription and developmental regulator, AUTS2, is involved in the activation of the Rho family small GTPase Rac1. This pathway aims to control neuronal migration and neurite extension through the coordination of actin polymerization and microtubule dynamics. In the nucleus, AUTS2 functions as a transcriptional activator of many target genes, together with the polycomb complex 1 (PRC1) [41].

2.3.2. ARHGAP21 and PTPN13

Rho GTPase activating protein 21 (ARHGAP21) is involved in many pathways, both extra and intracellular, such as the inhibition of cell migration and proliferation, cell polarity, cell adhesion, Golgi regulation and positioning, intracellular trafficking and glucose homeostasis [42].
Protein tyrosine phosphatase non-receptor type 13 (PTPN13) is a tyrosine phosphatase which mediates phosphoinositide 3-kinase (PI3K) signalling through dephosphorylation of phosphoinositide-3-kinase regulatory subunit 2 (PIK3R2) [43]. In addition, PTPN13 plays a relevant role in the same extracellular mechanism as ARHGAP21. Therefore, it controls negative apoptotic signalling [44] as the antagonist of Rho GTPase A, SLIT-ROBO Rho GTPase activating protein 1 (SRGAP1) [45].

2.3.3. FMNL2

Formin-like protein 2-FMNL2 is involved in the modulation of actin polymerization and organization of the cytoskeleton [46] through the regulation of contractility during epithelial junction maturation [47].

2.3.4. UBC, MYO18A, EPS15, ANKH

Ubiquitin C (UBC), Myosin XVIIIA (MYO18A), epidermal growth factor receptor pathway substrate 15 (EPS15) and ANKH inorganic pyrophosphate transport regulator are also involved in intracellular transport processes, and the dysregulation of these genes could influence vesicular trafficking of RPE cells, essential for photoreceptor outer segment (POS) renewal, visual cycle intermediate regeneration and avoiding AGE accumulation.
Specifically, MYO18A is involved in actin’s retrograde treadmilling and its transport from focal adhesions to the leading edge [48]. ANKH regulates trans-Golgi network trafficking and endocytosis [49], and finally, EPS15, with its encoded protein, is involved in clathrin-coated pit maturation, including invagination or budding and cell growth regulation [50,51].

2.3.5. RFFL, FBXW2, CAND1

These three genes are involved in the ubiquitin-proteasome system activity, and their down-regulation could lead to ER stress, inducing the accumulation of misfolded proteins and AGEs.
RFFL (ring finger and FYVE-like domain containing E3 ubiquitin protein ligase /rififylin) is an E3 ubiquitin-protein ligase which plays an important role in the extrinsic pathway of apoptosis, modulating cellular death domain receptors [52,53].
FBXW2 (F-box/WD repeat-containing protein 2) enhances the ubiquitylation and degradation of ß-catenin, overexpressed in the WNT/ß-catenin pathway during inflammatory processes.
CAND1 (Cullin-associated NEDD8-dissociated protein 1) controls the tubules’ elongation and retraction, thereby regulating the tubular endoplasmic reticulum network [54].
The down-expression of FBXW2 and CAND1 also plays a role in ROS production and cell death, altering cellular respiration processes through the impairment of glycolytic metabolism in RPE cells.

2.3.6. SIK3

SIK family kinase 3 (SIK3) gene encodes for a specific kinase that up-regulates mTOR, CREB signalling and cholesterol biosynthesis, correlating with retinoid metabolism and melanogenesis [55].
The down-expression of SIK3 could cause a defect in energy metabolism, decreasing mitochondrial respiration and up-regulating autophagy in order to remove dysfunctional cellular components, leading to cellular antioxidant mechanisms’ impairment [56,57].

2.3.7. IPO7

Importin 7 (IPO7) functions as a receptor for nuclear localization signals (NLS) and promotes translocation of import substrates through the nuclear pore complex (NPC). Its down-regulation could trigger p53-dependent growth arrest, ribosomal biogenesis stress and nucleolar morphology changes [58].

2.3.8. MRPS33, MORC4, MCPH1, NFIA, CTIF and LMBRD1

Down-expression of mitochondrial ribosomal protein S33 (MRPS33) could damage mitochondrial protein synthesis [59]. MORC family CW-type zinc finger 4 (MORC4) and Microcephalin 1 (MCPH1) dysregulation arrests DNA damage repair and causes apoptosis [60,61]. Reduced levels of nuclear Factor I A (NFIA) impair mitotic exit and cell differentiation [62]. Cap-binding complex-dependent translation initiation factor (CTIF) down-expression alters the intermodulation between translation and the aggresome-autophagy pathway, compromising mRNA and protein’s quality control [63]. Lastly, LMBR1 domain containing 1 (LMBRD1) down-regulation could affect the transport and metabolism of cobalamin, decreasing distribution of cyanocobalamin (vitamin B12) from the blood to the retina [64,65].

3. miRNA Altered Expression and Oxidative Stress:

miRNAs represent a group of short non-coding RNAs that are involved in transcript degradation or translational inhibition of their target mRNAs, whose function is to regulate post-transcriptional gene expression [66]. They are key regulators of many important biological processes [67], controlling specific pathways by targeting networks of functionally correlated genes.
Alterations of miRNA expression, due to mutations in either the miRNA itself or its target genes, could lead to several pathological conditions.
There is much evidence to support the role of miRNAs in normal retinal development and functions [68]: deletion of specific retina-enriched miRNAs has relevant effects on the development of retinal diseases, such as RP [69].
Luigi Donato et al. investigated the complexity of human retina miRNome (murine miRNA transcriptome), analysing data from human RPE cell transcriptomes.
Due to its specific proteins, RPE has many functions, including (i) the regeneration of outer segments of photoreceptors by phagocytizing the spent discs, (ii) regulating the trafficking of nutrients and waste products to and from the retina, (iii) protecting the outer retina from excessive high-energy light and the subsequent light-generated reactive oxygen species and (iv) maintaining retinal homeostasis thanks to the release of diffusible factors.
As a result of all this metabolic activity, RPE cells are very susceptible to oxidative stress [36,70]. Furthermore, RPE cells contain a significant number of mitochondria that are the principal cause of ROS production and removal inside the cell [70]. OS plays a critical role in the etiopathogenesis of RP [71] and leads to pathobiological changes in RPE cells [72] determining outer blood–retina barrier dysfunction [73], inhibition of processing of photoreceptor outer segments by RPE [74], expression of transforming growth factor-β2 [75] and synthesis alterations of extracellular matrix components [76]. All these changes lead to increased RPE apoptosis [77] and senescence changes [72,78].
In a recent paper [11], authors compared changes in the expression of miRNAs obtained from whole transcriptome analyses between two groups of RPE cells, one untreated and the other exposed to the oxidant agent oxidized low-density lipoprotein (oxLDL). In the treated samples, 23 miRNAs revealed altered expression, targeting genes involved in several biochemical pathways, many of which were associated with RP. Moreover, five RP causative genes (KLHL7, RDH11, CERKL, AIPL1 and USH1G) emerged as already confirmed targets of five altered miRNAs (hsa-miR-1307, hsa-miR-3064, hsa-miR-4709, hsa-miR-3615 and hsa-miR-637), suggesting a connection between induced oxidative stress and RP development and progression.
The finding of new regulative functions of miRNAs, and especially their altered expression induced by OS in RPE, should lead to the discovery of alternative mechanism responsible of the etiopathogenesis and progression of RP.

4. Role of Long Non-Codingrna

Long non-coding RNAs (lncRNAs) are untranslated transcripts that regulate many biological processes through epigenetic modifications, RNA splicing, mRNA decay and mRNA translation [79], acting as scaffolds for chromatin-modifying complexes [80].
Recent studies have highlighted the close connection between OS, the biochemical pathway involved in RP pathogenesis, and lncRNAs differential expression [81,82] in RPE metabolism.
In cells such as RPE, the high metabolic demand determines the up-regulation of DNA metabolic processes that cause an increased rDNA silencing, due to chromatin-associated lncRNAs, that leads to cellular growth alterations and RPE cell death [83].
In this process, lncRNAs have a key role, as it is well-known that they are involved in DNA metabolic processes: lncRNAs up- or down-regulation could alter gene expression, along with cellular responses to OS, which is one of the most important pathways involved in RP.
Considering DNA damage, two different lncRNAs are likely induced and overexpressed: MNX-AS1 and MIR31HG.
They interact with Cyclin D1 mRNA, whose encoding gene is already known to transcribe a specific lncRNA involved in DNA damage condition, acting as transcription repressors [84].
Oxidative stress also determines changes in glucose metabolism. Specifically, the dysregulation of two lnc-RNAs that are implicated in bioenergetic reactions related to glucose, BDNF-AS and TUG1 [85], were discovered to induce RPE apoptosis [86,87].
Moreover, high glucose levels influence the synthesis of IGF-1, PEDF, AGEs and their receptors (RAGE) that determine OS and inflammatory reactions, leading to retinal degeneration [88].
Additionally, other deregulated lncRNAs connected with insulin-related pathways, such as ARF, AKT1 [89], CRNDE, CYTOR CAP1 and ACACA [90,91], were discovered to induce RPE cell apoptosis [37].
Considering lipid metabolism and homeostasis, RPE cells present intracellular signalling pathways whose gene transcription is regulated by the peroxisome proliferator-activated receptor (PPAR) [92].
In the study by Donato et al. [37], three lncRNAs (AC007283.1, AC012442.2 and AC089983) were identified as being involved in fatty acids biosynthesis and metabolism. In particular, the down-expression of AC007283.1 and AC012442.2, along with the over-expression AC089983.1, could alter the gene expression and lipid metabolism regulation by PPAR-alfa.
Additionally, several down-regulated lncRNAs, such as AC004943.2 and AC007036.3, were found to interact with various miRNAs involved in fatty acid metabolism and biosynthesis, leading to the impairment of the integrity and functionality of lipidic retinal structures [37].
Lastly, the involvement of numerous lncRNAs was also identified in protein metabolism. It is well-known that protein’s misfolding, including those related to retinal survival and vision process, like rhodopsin, determine the disruptions of cellular protein homeostasis [93] and could lead to cell death.
Two clusters made of dysregulated lncRNAs and their interactors/host genes were detected to be involved in cellular amide metabolism [37]. Among them, the up-regulation of PTEN-induced putative kinase protein 1 (PINK1) antisense RNA and the downregulation of FMRP translational regulator 1 (FMR1)-IT1 sense intronic and vimentin (VIM) antisense 1 RNA were found to be particularly interesting.
PINK1 regulates mitochondrial damage, promotes mitophagy and protects cells from death and apoptosis, especially during high glucose-mediated regulation of RPE [91], reflecting the apoptosis status of RPE.
FMR1-IT1 and VIM-AS1 are related to synaptogenesis, intracellular trafficking and cellular stability [94,95], representing the attempt of RPE cells to boost the production of vital proteins.

5. P2X7 Receptor and Inflammation in RP

The P2X7 receptor (P2X7R) is an ATP-gated ion channel expressed by immune and inflammatory cells and is over expressed during inflammation and by stressed or dying cells involved in innate and adaptive immune responses. It is recognized as a potent trigger of ROS production, and its over-stimulation leads to the impairment of mitochondrial metabolism, caspase activation as well as apoptosis induction [96,97]. Its stimulation also induces ATP release by means of a membrane pore formation or in association with pannexin hemichannels, activating the NLRP3 (NLR family pyrin domain-containing 3) inflammasome that induces the maturation and release of pro-inflammatory cytokines (IL-1β and IL-18) and the production of ROS, released by macrophages and microglia, contributing to the progression of neuroinflammatory and neurodegenerative diseases [98,99].
The expression of P2X7R was showed in several components of the retinal layers: not only on photoreceptor cells, but also on retinal ganglion cell, amacrine and horizontal cells, microglia and Müller glial cells, astrocytes and pericytes as well as RPE cells [100].
An upregulation of P2x7R mRNA was demonstrated within the retina of an RP mouse model when photoreceptor degeneration occurred. Furthermore, intravitreal administration of ATP in this animal model caused photoreceptor cell death and loss of function due to the activation of P2x7R. As a demonstration, this process can be delayed by intravitreal injection of a P2x7 receptor antagonist in rd1 mouse models [101].

6. The Dual Role of Microglia in RP: Between Neurotoxicity and Neuroprotection

We have already analysed two mechanisms that determine an activation of microglia, involving MUTYH and P2X7 receptors. The excessive activation of MUTYH determines nuclear accumulation of 8-oxoG, which is the most prevalent genotoxic lesion, and the PARP pathway as a consequence of accumulation of oxidized nucleic acids, while the upregulation of P2X7R is induced by stressed retinal cells that release extracellular ATP, leading to microglial chemotaxis and activation [100].
Herein, we report in detail the molecular mechanism underlying the activation of microglia and the role of these cells in the development of RP.
The microglia is composed of immune cells implicated in neuronal homeostasis and innate immune defences. Microglia’s cells are activated in response to the altered physiology of mutation-bearing photoreceptors [102], inducing the production of proinflammatory cytokines and chemokines [103,104,105].
A study conducted on a rd1 mouse line demonstrates that in the early stages of disease, there is a persistent upregulation of microglial markers, such as Tmem119, C1qa, TNF, Il1a and Il1b, that act as inflammatory factors, determining neurotoxicity [106,107,108,109]. A rd1 mouse line treated with PLX5622, a potent colony-stimulator factor 1 receptor (CSF1R) inhibitor that eliminates microglial population, did not show an increase of these factors.
However, it is known that the role of microglia in the development of RP is dictated by a balance between neurotoxic and neuroprotective/neurotrophic influences.
In fact, the activation of retinal microglia induces the expression of neurotrophic factors in Muller cells, exerting neuroprotection in degenerating photoreceptors [110]. Furthermore, it has been demonstrated that overexpression of TGF-beta, an anti-inflammatory cytokine that has a modulatory effect on the action of microglia [111], exercises a protective effect on the degeneration of the cones.
More specifically, through RNA-seq, it has been highlighted that the effects are explicit through post-translational modifications of the proteome not detectable using RNA-seq.
Inhibition of this neuroprotective pathway, induced by the expression of TGF-beta, determines microglial activation, resulting in degenerative changes in retinal tissue given by the expression of proinflammatory cytokines [112].
Finally, the last mechanism concerns the interaction between microglia and complement activation, specifically between the central complement component (C3) and the microglia-expressed receptor (CR3) [113].
An increased C3 expression was found in microglia translocated in ONL after rod degeneration, with a concomitant opsonization of the degenerating photoreceptor by iC3b, a product of C3 activation. This complement activation is therefore a microglial response aimed at phagocytosis of the apoptotic photoreceptor and the restoration of homeostasis. When both C3 and CR3 are deficient, it causes an increased proinflammatory cytokine expression, accumulation of apoptotic cells and neurodegeneration [114].
These findings could lead to other possible immunomodulatory therapy opportunities.

7. Conclusions

The molecular mechanisms related to oxidative stress occurring in retinitis pigmentosa (Figure 1) play a central role in its pathogenesis and progression. The burden of the analysed pathways demonstrate that oxidative microglial activation may trigger the vicious cycle of non-resolved neuroinflammation and degeneration in RP, suggesting that the microglia may be a key target of oxidative stress in RP. Failure of the endogenous mechanisms to overcome oxidative stress leads to an accelerated progression of retinal neurodegeneration.
This analytic review aimed to highlight possible therapeutic targets inside the different pathogenetic mechanisms that induce the formation of ROS and potentially slow down the evolution of the disease over time.

Funding

The works of the authors cited herein have been funded by AIRC grants to FDV and MT, and University of Ferrara, FAR projects to FDV, MT and PP.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

3-DG3-DeoxyGlucosone
8-oxoG8oxoGuanine
ACACA Acetyl-CoA Carboxylase Alpha
AGEsAdvanced Glycation End products
AIPL1Aryl Hydrocarbon Receptor Interacting Protein Like 1
AKT1AKT Serine/Threonine Kinase 1
ALEsAdvanced lipoxidation End-products
ANKHProgressive ankylosis protein homolog
ARFAlternate Reading Frame
ARHGAP21Rho GTPase Activating Protein 21
AUTS2Cytoplasmatic activator of transcription and developmental regulator
BDNF-ASBrain-derived neurotrophic factor Antisense
BERBase Exicision Repair
C3Complement Component 3
CAND1Cullin-associated NEDD8-dissociated protein 1
CAP1Cyclase Associated Actin Cytoskeleton Regulatory Protein 1
CERKLCeramide Kinase Like
CR3Complement Receptor 3
CRDCone-Rode Dystrophy
CREBcAMP response element-binding protein
CRNDEColorectal Neoplasia Differentially Expressed
CSF1RColony Stimulator Factor 1 Receptor (PLX5622)
CTIFCap Binding Complex Dependent Translation Initiation Factor
CYTORCytoskeleton Regulator RNA
EPS15Epidermal Growth Factor Receptor Pathway Substrate 15
FBXW2F-box/WD repeat-containing protein 2
FMNL2Formin Like Protein 2
FMR1FMRP Translational Regulator 1
GAPDHGlyceraldehyde 3-phosphate Dehydrogenase
GLO1 Glyoxalase 1
GOGlyoxal
hMTH1Human MutT Homologue
HSPHeat Shock Protein
iBRBInner Blood-Retinal Barrier
IGF1 insulin-like growth factor-1
IL-18Interleukin-18
IL-1βInterleukin-1β
IPO3Transportin-3
IPO7Importin 7
IRE1 Inositol-Requiring Enzyme 1
KLHL7Kelch Like Family Member 7
LC3-IILipidated form of LC3 (Microtubule-associated protein 1A/1B-light chain 3)
LMBRD1Limb Development Membrane Protein Domain 1
lnc-RNAlong non-coding-RNA
MCPH Microcephalin
MGMethylGlyoxal
MIR31HGMicroRNA 31 Host Gene
miRNA microRNA
miRNome Murine miRNA
MNXMotor Neuron And Pancreas Homeobox 1
MNX-AS1 antisense transcript of MNX1
MORC 4 MORC family CW-type zinc finger protein 4
MRPS3328S ribosomal protein S33
mTORMammalian target of Rapamycin
MUTYHmutY DNA glycosylase
MYO18AMyosin XVIIIA
NFIANuclear Factor I A
NLSNuclear Localization Signals
NPCNuclear Pore Complex
ONLOuter Nuclear Layer
OS Oxidative Stress
P2X7RP2X 7 receptor
PARPPoly(ADP-ribose) polymerase
PEDFPigment Epithelium-derived Factor
PERKPKR-like Endoplasmic Reticulum Kinase
PI3KPhosphoinositide 3 Kinase
PINK1PTEN-induced kinase 1
POSPhotoreceptor Outer Segment
PPAR peroxisome proliferator-activated receptor
PRC1 Polycomb Complex 1
PTEN Phosphatase and tensin homolog
PTPN13Protein Tyrosine Phosphatase Non-Receptor type 13
RAC1Rac Family Small GTPase 1
RDH11Retinol Dehydrogenase 11
RERough Endoplasmic reticulum
RFFL E3 ubiquitin-protein ligase rififylin
ROSReactive Oxygen Species
RPERetinal Pigment Epithelium
SIK3SIK Family Kinase 3
SL SphingoLipid
SODSuperoxide Dismutase
SRGAP1Slit-Robo Rho GTPase Activating Protein 1
SSBSingle Strand Breaks
TUG1 Taurine Up-Regulated 1
UBC Ubiquitin C
UPRUnfolded Protein Response
USH1GUsher syndrome type-1G protein
VEGFVascular Endothelial Growth Factor
VIM Vimentin

References

  1. Campa, C.; Gallenga, C.E.; Bolletta, E.; Perri, P. The Role of Gene Therapy in the Treatment of Retinal Diseases: A Review. Curr. Gene Ther. 2017, 17, 194–213. [Google Scholar] [CrossRef] [PubMed]
  2. Sorrentino, F.S.; Gallenga, C.E.; Bonifazzi, C.; Perri, P. A challenge to the striking genotypic heterogeneity of retinitis pigmentosa: A better understanding of the pathophysiology using the newest genetic strategies. Eye 2016, 30, 1542–1548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Sies, H.; Jones, D.P. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 2020, 21, 363–383. [Google Scholar] [CrossRef] [PubMed]
  4. Gorbatyuk, M.S.; Starra, C.R.; Gorbatyuk, O.S. Endoplasmic reticulum stress: New insights into the pathogenesis and treatment of retinal degenerative diseases. Prog. Retin. Eye Res. 2020, 79, 100860. [Google Scholar] [CrossRef]
  5. Domènech, E.B.; Marfany, G. The Relevance of Oxidative Stress in the Pathogenesis and Therapy of Retinal Dystrophies. Antioxidants 2020, 9, 347. [Google Scholar] [CrossRef] [Green Version]
  6. Kroeger, H.; Chiang, W.C.; Felden, J.; Nguyen, A.; Lin, J.H. ER stress and unfolded protein response in ocular health and disease. FEBS J. 2019, 286, 399–412. [Google Scholar] [CrossRef]
  7. Donato, L.; Scimone, C.; Nicocia, G.; Denaro, L.; Robledo, R.; Sidoti, A.; D’Angelo, R. GLO1 gene polymorphisms and their association with retinitis pigmentosa: A case–control study in a Sicilian population. Mol. Biol. Rep. 2018, 45, 1349–1355. [Google Scholar] [CrossRef]
  8. Boya, P. Why autophagy is good for retinal ganglion cells? Eye 2017, 31, 185–190. [Google Scholar] [CrossRef] [Green Version]
  9. Boya, P.; Esteban-Martínez, L.; Serrano-Puebla, A.; Gómez-Sintes, R.; Villarejo-Zori, B. Autophagy in the eye: Development, degeneration, and aging. Prog. Retin. Eye Res. 2016, 55, 206–245. [Google Scholar] [CrossRef] [PubMed]
  10. Rodríguez-Muela, N.; Hernández-Pinto, A.M.; Serrano-Puebla, A.; García-Ledo, L.; Latorre, S.H.; de la Rosa, E.J.; Boya, P. Lysosomal membrane permeabilization and autophagy blockade contribute to photoreceptor cell death in a mouse model of retinitis pigmentosa. Cell Death Differ. 2015, 22, 476–487. [Google Scholar] [CrossRef] [Green Version]
  11. Donato, L.; Bramanti, P.; Scimone, C.; Rinaldi, C.; Giorgianni, F.; Beranova-Giorgianni, S.; Koirala, D.; D’Angelo, R.; Sidoti, A. miRNAexpression profile of retinal pigment epithelial cells under oxidative stress conditions. FEBS Open Bio 2018, 8, 219–233. [Google Scholar] [CrossRef] [Green Version]
  12. Jingwen, C.; Zhenqian, H.; Xin, W.; Jaqi, K.; Yan, R.; Wei, G.; Xiang, L.; Jingmei, W.; Weidong, D.; Yusaku, N.; et al. Oxidative stress induces different tissue dependent effects on Mutyh-deficient mice. Free Radic. Biol. Med. 2019, 143, 482–493. [Google Scholar]
  13. Douglas, M.; Nunez, N.N.; Burnside, M.A.; Bradshaw, K.M.; David, S.S. Repair of 8-oxoG: A Mismatches by the MUTYH Glycosylase: Mechanism, Metals and Medicine. Free Radic. Biol. Med. 2017, 107, 202–215. [Google Scholar]
  14. Foti, J.J.; Devadoss, B.; Winkler, J.A.; Collins, J.J.; Walker, G.C. Oxidation of the guanine nucleotide pool underlies cell death by bactericidal antibiotics. Science 2012, 336, 315–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Nakatake, S.; Murakami, Y.; Ikeda, Y.; Morioka, N.; Tachibana, T.; Fujiwara, K.; Yoshida, N.; Notomi, S.; Hisatomi, T.; Yoshida, S.; et al. MUTYH promotes oxidative microglial activation and inherited retinal degeneration. JCI Insight 2016, 1, e87781. [Google Scholar] [CrossRef] [Green Version]
  16. Oka, S.; Nakabeppu, Y. DNA glycosylase encoded byMUTYH functions as a molecular switch for programmed cell death under oxidative stress to suppress tumorigenesis. Cancer Sci. 2011, 102, 677–682. [Google Scholar] [CrossRef] [PubMed]
  17. Oka, S.; Ohno, M.; Tsuchimoto, D.; Sakumi, K.; Furuichi, M.; Nakabeppu, Y. Two distinct pathways of cell death triggered by oxidative damage to nuclear and mitochondrial DNAs. EMBO J. 2008, 27, 421–432. [Google Scholar] [CrossRef] [Green Version]
  18. Murakami, Y.; Ikeda, Y.; Yoshida, N.; Notomi, S.; Hisatomi, T.; Oka, S.; De Luca, G.; Yonemitsu, Y.; Bignami, M.; Nakabeppu, Y.; et al. MutT homolog-1 attenuates oxidative DNA damage and delays photoreceptor cell death in inherited retinal degeneration. Am. J. Pathol. 2012, 181, 1378–1386. [Google Scholar] [CrossRef] [PubMed]
  19. Usui, S.; Oveson, B.C.; Lee, S.Y.; Jo, Y.J.; Yoshida, T.; Miki, A.; Miki, K.; Iwase, T.; Lu, L.; Campochiaro, P.A. NADPH oxidase plays a central role in cone cell death in retinitis pigmentosa. J. Neurochem. 2009, 110, 1028–1037. [Google Scholar] [CrossRef] [Green Version]
  20. Conti, P.; Lauritano, D.; Caraffa, A.; Gallenga, C.E.; Kritas, S.K.; Ronconi, G.; Martinotti, S. Microglia and mast cells generate proinflammatory cytokines in the brain and worsen inflammatory state: Suppressor effect of IL-37. Eur. J. Pharmacol. 2020, 15, 875. [Google Scholar] [CrossRef]
  21. Tuson, M.; Garanto, A.; Gonzàlez-Duarte, R.; Marfany, G. Overexpression of CERKL, a gene responsible for retinitis pigmentosa in humans, protects cells from apoptosis induced by oxidative stress. Mol. Vis. 2009, 15, 168–180. [Google Scholar]
  22. Fathinajafabadi, A.; Pérez-Jiménez, E.; Riera, M.; Knecht, E.; Gonzàez-Duarte, R. CERKL, a retinal disease gene, encodes an mRNA-binding protein that localizes in compact and untranslated mRNPs associated with microtubules. PLoS ONE 2014, 9, e87898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Li, C.; Wang, L.; Zhang, J.; Huang, M.; Wong, F.; Liu, X.; Liu, F.; Cui, X.; Yang, G.; Chen, J.; et al. CERKL interacts with mitochondrial TRX2 and protects retinal cells from oxidative stress-induced apoptosis. Biochim. Biophys. Acta Mol. Basis Dis. 2014, 1842, 1121–1129. [Google Scholar] [CrossRef] [Green Version]
  24. Hannun, Y.A.; Obeid, L.M. The ceramide-centric universe of lipid- mediated cell regulation: Stress encounters of the lipid kind. J. Biol. Chem. 2002, 277, 25847–25850. [Google Scholar] [CrossRef] [Green Version]
  25. Hannun, Y.A.; Obeid, L.M. Principles of bioactive lipid signalling: Lessons from sphingolipids. Nat. Rev. Mol. Cell Biol. 2008, 9, 139–150. [Google Scholar] [CrossRef] [PubMed]
  26. Tuson, M.; Marfany, G.; Gonzalez-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] [PubMed] [Green Version]
  27. Auslender, N.; Sharon, D.; Abbasi, A.H.; Garzozi, H.J.; Banin, E.; Ben- Yosef, T. A common founder mutation of CERKL underlies autosomal recessive retinal degeneration with early macular involvement among Yemenite Jews. Investig. Ophthalmol. Vis. Sci. 2007, 48, 5431–5438. [Google Scholar] [CrossRef] [Green Version]
  28. Khan, M.A.; Anwar, S.; Aljarbou, A.N.; Al-Orainy, M.; Aldebasi, Y.H.; Islam, S.; Younus, H. Protective effect of thymoquinone on glucose or methylglyoxal-induced glycation of superoxide dismutase. Int. J. Biol. Macromol. 2014, 65, 16–20. [Google Scholar] [CrossRef]
  29. Groener, J.B.; Oikonomou, D.; Cheko, R.; Kender, Z.; Zemva, J.; Kihm, L.; Muckenthaler, M.; Peters, V.; Fleming, T.; Kopf, S.; et al. Methylglyoxal and Advanced Glycation End Products in Patients with Diabetes—What We Know so Far and the Missing Links. Exp. Clin. Endocrinol. Diabetes 2019, 127, 497–504. [Google Scholar] [CrossRef] [PubMed]
  30. Nedic, O.; Rattan, S.I.; Grune, T.; Trougakos, I.P. Molecular effects of advanced glycation end products on cell signalling pathways, ageing and pathophysiology. Free Radic. Res. 2013, 47, 28–38. [Google Scholar] [CrossRef]
  31. Lee, H.J.; Howell, S.K.; Sanford, R.J.; Beisswenger, P.J. Methylglyoxal can modify GAPDH activity and structure. Ann. N. Y. Acad. Sci. 2005, 1043, 135–145. [Google Scholar] [CrossRef]
  32. Wautier, M.P.; Guillausseau, P.J.; Wautier, J.L. Activation of the receptor for advanced glycation end products and consequences on health. Diabetes Metab. Syndr. 2017, 11, 305–309. [Google Scholar] [CrossRef]
  33. Xu, J.; Chen, L.J.; Yu, J.; Wang, H.J.; Zhang, F.; Liu, Q.; Wu, J. Involvement of Advanced Glycation End Products in the Pathogenesis of Diabetic Retinopathy. Cell Physiol. Biochem. 2018, 48, 705–717. [Google Scholar] [CrossRef]
  34. Grigsby, J.G.; Allen, D.M.; Ferrigno, A.S.; Vellanki, S.; Pouw, C.E.; Hejny, W.A.; Tsin, A.T.C. Autocrine and Paracrine Secretion of Vascular Endothelial Growth Factor in the Pre-Hypoxic Diabetic Retina. Curr. Diabetes Rev. 2017, 13, 161–174. [Google Scholar] [CrossRef]
  35. Chen, M.; Curtis, T.M.; Stitt, A.W. Advanced glycation end products and diabetic retinopathy. Curr. Med. Chem. 2013, 20, 3234–3240. [Google Scholar] [CrossRef]
  36. Datta, S.; Cano, M.; Ebrahimi, K.; Wang, L.; Handa, J.T. The impact of oxidative stress and inflammation on RPE degeneration in non-neovascular AMD. Prog. Retin. Eye Res. 2017, 60, 201–218. [Google Scholar] [CrossRef] [PubMed]
  37. Donato, L.; Scimone, C.; Alibrandi, S.; Rinaldi, C.; Sidoti, A.; D’Angelo, R. Transcriptome Analyses of lncRNAs in A2E-Stressed Retinal Epithelial Cells Unveil Advanced Links between Metabolic Impairments Related to Oxidative Stress and Retinitis Pigmentosa. Antioxidants 2020, 9, 318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Kaarniranta, K.; Koskela, A.; Felszeghy, S.; Kivinen, N.; Salminen, A.; Kauppinen, A. Fatty acids and oxidized lipoproteins contribute to autophagy and innate immunity responses upon the degeneration of retinal pigment epithelium and development of age-related macular degeneration. Biochimie 2019, 159, 49–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Peculis, R.; Konrade, I.; Skapare, E.; Fridmanis, D.; NikitinaZake, L.; Lejnieks, A.; Pirags, V.; Dambrova, M.; Klovins, J. Identification of glyoxalase 1 polymorphisms associatedwith enzyme activity. Gene 2013, 515, 140–143. [Google Scholar] [CrossRef] [PubMed]
  40. Donato, L.; Scimone, C.; Alibrandi, S.; Nicocia, G.; Rinaldi, C.; Sidoti, A.; D’Angelo, R. Discovery of GLO1 New Related Genes and Pathways by RNA-Seq on A2E-Stressed Retinal Epithelial Cells Could Improve Knowledge on Retinitis Pigmentosa. Antioxidants 2020, 9, 416. [Google Scholar] [CrossRef] [PubMed]
  41. Hori, K.; Hoshino, M. Neuronal Migration and AUTS2 Syndrome. Brain Sci. 2017, 7, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Rosa, L.R.O.; Soares, G.M.; Silveira, L.R.; Boschero, A.C.; Barbosa-Sampaio, H.C.L. ARHGAP21 as a master regulator of multiple cellular processes. J. Cell Physiol. 2018, 233, 8477–8481. [Google Scholar] [CrossRef] [PubMed]
  43. Kuchay, S.; Duan, S.; Schenkein, E.; Peschiaroli, A.; Saraf, A.; Florens, L.; Washburn, M.P.; Pagano, M. FBXL2- and PTPL1-mediated degradation of p110-free p85beta regulatory subunit controls the PI(3)K signalling cascade. Nat. Cell Biol. 2013, 15, 472–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Villa, F.; Deak, M.; Bloomberg, G.B.; Alessi, D.R.; van Aalten, D.M. Crystal structure of the PTPL1/FAP-1 human tyrosine phosphatase mutated in colorectal cancer: Evidence for a second phosphotyrosine substrate recognition pocket. J. Biol. Chem. 2005, 280, 8180–8187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kaur, H.; Xu, N.; Doycheva, D.M.; Malaguit, J.; Tang, J.; Zhang, J.H. Recombinant Slit2 attenuates neuronal apoptosis via the Robo1-srGAP1 pathway in a rat model of neonatal HIE. Neuropharmacology 2019, 158, 107727. [Google Scholar] [CrossRef]
  46. Kage, F.; Steffen, A.; Ellinger, A.; Ranftler, C.; Gehre, C.; Brakebusch, C.; Pavelka, M.; Stradal, T.; Rottner, K. FMNL2 and -3 regulate Golgi architecture and anterograde transport downstream of Cdc42. Sci. Rep. 2017, 7, 9791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Liang, X.; Kiru, S.; Gomez, G.A.; Yap, A.S. Regulated recruitment of SRGAP1 modulates RhoA signaling for contractility during epithelial junction maturation. Cytoskeleton 2018, 75, 61–69. [Google Scholar] [CrossRef]
  48. Buschman, M.D.; Field, S.J. MYO18A: An unusual myosin. Adv. Biol. Regul. 2018, 67, 84–92. [Google Scholar] [CrossRef]
  49. Van Gils, M.; Nollet, L.; Verly, E.; Deianova, N.; Vanakker, O.M. Cellular signaling in pseudoxanthoma elasticum: An update. Cell Signal. 2019, 55, 119–129. [Google Scholar] [CrossRef]
  50. Khanobdee, K.; Kolberg, J.B.; Dunlevy, J.R. Nuclear and plasma membrane localization of SH3BP4 in retinal pigment epithelial cells. Mol. Vis. 2004, 10, 933–942. [Google Scholar]
  51. Majumdar, A.; Ramagiri, S.; Rikhy, R. Drosophila homologue of Eps15 is essential for synaptic vesicle recycling. Exp. Cell Res. 2006, 312, 2288–2298. [Google Scholar] [CrossRef]
  52. Gan, X.; Wang, C.; Patel, M.; Kreutz, B.; Zhou, M.; Kozasa, T.; Wu, D. Different Raf protein kinases mediate different signaling pathways to stimulate E3 ligase RFFL gene expression in cell migration regulation. J. Biol. Chem. 2013, 288, 33978–33984. [Google Scholar] [CrossRef] [Green Version]
  53. Sakai, R.; Fukuda, R.; Unida, S.; Aki, M.; Ono, Y.; Endo, A.; Kusumi, S.; Koga, D.; Fukushima, T.; Komada, M.; et al. The integral function of the endocytic recycling compartment is regulated by RFFL-mediated ubiquitylation of Rab11 effectors. J. Cell Sci. 2019, 132. [Google Scholar] [CrossRef] [Green Version]
  54. Kajiho, H.; Yamamoto, Y.; Sakisaka, T. CAND1 regulates lunapark for the proper tubular network of the endoplasmic reticulum. Sci. Rep. 2019, 9, 13152. [Google Scholar] [CrossRef] [Green Version]
  55. Uebi, T.; Itoh, Y.; Hatano, O.; Kumagai, A.; Sanosaka, M.; Sasaki, T.; Sasagawa, S.; Doi, J.; Tatsumi, K.; Mitamura, K.; et al. Involvement of SIK3 in glucose and lipid homeostasis in mice. PLoS ONE 2012, 7, e37803. [Google Scholar] [CrossRef]
  56. Chong, C.M.; Zheng, W. Artemisinin protects human retinal pigment epithelial cells from hydrogen peroxide-induced oxidative damage through activation of ERK/CREB signaling. Redox Biol. 2016, 9, 50–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Kriete, A.; Bosl, W.J.; Booker, G. Rule-based cell systems model of aging using feedback loop motifs mediated by stress responses. PLoS Comput. Biol. 2010, 6, e1000820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Golomb, L.; Bublik, D.R.; Wilder, S.; Nevo, R.; Kiss, V.; Grabusic, K.; Volarevic, S.; Oren, M. Importin 7 and exportin 1 link c-Myc and p53 to regulation of ribosomal biogenesis. Mol. Cell. 2012, 45, 222–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Cavdar Koc, E.; Burkhart, W.; Blackburn, K.; Moseley, A.; Spremulli, L.L. The small subunit of the mammalian mitochondrial ribosome. Identification of the full complement of ribosomal proteins present. J. Biol. Chem. 2001, 276, 19363–19374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Yang, Z.; Zhuang, Q.; Hu, G.; Geng, S. MORC4 is a novel breast cancer oncogene regulated by miR-193b-3p. J. Cell Biochem. 2019, 120, 4634–4643. [Google Scholar] [CrossRef] [PubMed]
  61. Liu, X.; Zong, W.; Li, T.; Wang, Y.; Xu, X.; Zhou, Z.W.; Wang, Z.Q. The E3 ubiquitin ligase APC/C(C)(dh1) degrades MCPH1 after MCPH1-betaTrCP2-Cdc25A-mediated mitotic entry to ensure neurogenesis. EMBO J. 2017, 36, 3666–3681. [Google Scholar] [CrossRef] [PubMed]
  62. Clark, B.S.; Stein-O’Brien, G.L.; Shiau, F.; Cannon, G.H.; Davis-Marcisak, E.; Sherman, T.; Santiago, C.P.; Hoang, T.V.; Rajaii, F.; James-Esposito, R.E.; et al. Single-Cell RNA-Seq Analysis of Retinal Development Identifies NFI Factors as Regulating Mitotic Exit and Late-Born Cell Specification. Neuron 2019, 102, 1111–1126. [Google Scholar] [CrossRef] [PubMed]
  63. Park, Y.; Park, J.; Kim, Y.K. Crosstalk between translation and the aggresome-autophagy pathway. Autophagy 2018, 14, 1079–1081. [Google Scholar] [CrossRef]
  64. Kawaguchi, K.; Okamoto, T.; Morita, M.; Imanaka, T. Translocation of the ABC transporter ABCD4 from the endoplasmic reticulum to lysosomes requires the escort protein LMBD1. Sci. Rep. 2016, 6, 30183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Kinoshita, Y.; Nogami, K.; Jomura, R.; Akanuma, S.I.; Abe, H.; Inouye, M.; Kubo, Y.; Hosoya, K.I. Investigation of Receptor-Mediated Cyanocobalamin (Vitamin B12) Transport across the Inner Blood-Retinal Barrier Using Fluorescence-Labeled Cyanocobalamin. Mol. Pharm. 2018, 15, 3583–3594. [Google Scholar] [CrossRef]
  66. He, L.; Hannon, G.J. MicroRNAs: Small RNAs with a big role in gene regulation. Nat. Rev. Genet. 2004, 5, 522–531. [Google Scholar] [CrossRef]
  67. Zamore, P.D.; Haley, B. Ribo-gnome: The big world of small RNAs. Science 2005, 309, 1519–1524. [Google Scholar] [CrossRef] [Green Version]
  68. Ohana, R.; Weiman-Kelman, B.; Raviv, S.; Tamm, E.R.; Pasmanik-Chor, M.; Rinon, A.; Netanely, D.; Shamir, R.; Solomon, A.S.; Ashery-Padan, R. MicroRNAs are essential for differentiation of the retinal pigmented epithelium and maturation of adjacent photoreceptors. Development 2015, 142, 2487–2498. [Google Scholar]
  69. Romano, G.L.; Platania, C.B.M.; Drago, F.; Salomone, S.; Ragusa, M.; Barbagallo, C.; Di Pietro, C.; Purrello, M.; Reibaldi, M.; Avitabile, T.; et al. Retinal and circulating miRNAs in age-related macular degeneration: An in vivo animal and human study. Front. Pharmacol. 2017, 8, 168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Tian, B.; Maidana, D.E.; Dib, B.; Miller, J.B.; Bouzika, P.; Miller, J.W.; Vavvas, D.G.; Lin, H. MiR-17-3p Exacerbates Oxidative Damage in Human Retinal Pigment Epithelial Cells. PLoS ONE 2016, 11, e0160887. [Google Scholar] [CrossRef] [Green Version]
  71. Kruk, J.; Kubasik-Kladna, K.; Aboul-Enein, H.Y. The role oxidative stress in the pathogenesis of eye diseases: Current status and a dual role of physical activity. Mini Rev. Med. Chem. 2015, 16, 241–257. [Google Scholar] [CrossRef] [PubMed]
  72. Mao, K.; Shu, W.; Qiu, Q.; Gu, Q.; Wu, X. Salvianolic acid A protects retinal pigment epithelium from OX-LDL-induced inflammation in an age-related macular degeneration model. Discov. Med. 2017, 23, 129–147. [Google Scholar] [PubMed]
  73. Kim, J.H.; Lee, S.J.; Kim, K.W.; Yu, Y.S.; Kim, J.H. Oxidized low density lipoprotein-induced senescence of retinal pigment epithelial cells is followed by outer blood-retinal barrier dysfunction. Int. J. Biochem. Cell Biol. 2012, 44, 808–814. [Google Scholar] [CrossRef] [PubMed]
  74. Hoppe, G.; Marmorstein, A.D.; Pennock, E.A.; Hoff, H.F. Oxidized low density lipoprotein-induced inhibition of processing of photoreceptor outer segments by RPE. Investig. Ophthalmol. Vis. Sci. 2001, 42, 2714–2720. [Google Scholar]
  75. Yu, A.L.; Lorenz, R.L.; Haritoglou, C.; Kampik, A.; Welge-Lussen, U. Biological effects of native and oxidized low-density lipoproteins in cultured human retinal pigment epithelial cells. Exp. Eye Res. 2009, 88, 495–503. [Google Scholar] [CrossRef] [PubMed]
  76. Saneipour, M.; Ghatreh-Samani, K.; Heydarian, E.; Farrokhi, E.; Abdian, N. Adiponectin inhibits oxidized low density lipoprotein-induced increase in matrix metalloproteinase 9 expression in vascular smooth muscle cells. ARYA Atheroscler. 2015, 11, 191–195. [Google Scholar]
  77. Yating, Q.; Yuan, Y.; Wei, Z.; Qing, G.; Xingwei, W.; Qiu, Q.; Lili, Y. Oxidized LDL induces apoptosis of human retinal pigment epithelium through activation of ERK-Bax/Bcl-2 signaling pathways. Curr. Eye Res. 2015, 40, 415–422. [Google Scholar] [CrossRef] [PubMed]
  78. Bailey, T.A.; Kanuga, N.; Romero, I.A.; Greenwood, J.; Luthert, P.J.; Cheetham, M.E. Oxidative stress affects the junctional integrity of retinal pigment epithelial cells. Investig. Ophthalmol. Vis. Sci. 2004, 45, 675–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Kopp, F.; Mendell, J.T. Functional Classification and Experimental Dissection of Long Noncoding RNAs. Cell 2018, 172, 393–407. [Google Scholar] [CrossRef] [Green Version]
  80. Kondo, Y.; Shinjo, K.; Katsushima, K. Long non-coding RNAs as an epigenetic regulator in human cancers. Cancer Sci. 2017, 108, 1927–1933. [Google Scholar] [CrossRef] [Green Version]
  81. Wawrzyniak, O.; Zarebska, Z.; Rolle, K.; Gotz-Wieckowska, A. Circular and long non-coding RNAs and their role in ophthalmologic diseases. Acta Biochim. Pol. 2018, 65, 497–508. [Google Scholar] [CrossRef] [PubMed]
  82. Donato, L.; Scimone, C.; Rinaldi, C.; D’Angelo, R.; Sidoti, A. Non-coding RNAome of RPE cells under oxidative stress suggests unknown regulative aspects of Retinitis pigmentosa etiopathogenesis. Sci. Rep. 2018, 8, 16638. [Google Scholar] [CrossRef]
  83. Manelyte, L.; Strohner, R.; Gross, T.; Langst, G. Chromatin targeting signals, nucleosome positioning mechanism and non-coding RNA-mediated regulation of the chromatin remodeling complex NoRC. PLoS Genet. 2014, 10, e1004157. [Google Scholar] [CrossRef] [Green Version]
  84. Thapar, R. Regulation of DNA Double-Strand Break Repair by Non-Coding RNAs. Molecules 2018, 23, 2789. [Google Scholar] [CrossRef] [Green Version]
  85. Han, X.; Yang, Y.; Sun, Y.; Qin, L.; Yang, Y. LncRNA TUG1 affects cell viability by regulating glycolysis in osteosarcoma cells. Gene 2018, 674, 87–92. [Google Scholar] [CrossRef] [PubMed]
  86. Chen, S.; Wang, M.; Yang, H.; Mao, L.; He, Q.; Jin, H.; Ye, Z.M.; Luo, X.Y.; Xia, Y.P.; Hu, B. LncRNA TUG1 sponges microRNA-9 to promote neurons apoptosis by up-regulated Bcl2l11 under ischemia. Biochem. Biophys. Res. Commun. 2017, 485, 167–173. [Google Scholar] [CrossRef]
  87. Li, Y.; Xu, F.; Xiao, H.; Han, F. Long noncoding RNA BDNF-AS inversely regulated BDNF and modulated high-glucose induced apoptosis in human retinal pigment epithelial cells. J. Cell Biochem. 2018, 119, 817–823. [Google Scholar] [CrossRef] [PubMed]
  88. Kang, M.K.; Lee, E.J.; Kim, Y.H.; Kim, D.Y.; Oh, H.; Kim, S.I.; Kang, Y.H. Chrysin Ameliorates Malfunction of Retinoid Visual Cycle through Blocking Activation of AGE-RAGE-ER Stress in Glucose-Stimulated Retinal Pigment Epithelial Cells and Diabetic Eyes. Nutrients 2018, 10, 1046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Millar, C.A.; Powell, K.A.; Hickson, G.R.; Bader, M.F.; Gould, G.W. Evidence for a role for ADP-ribosylation factor 6 in insulin-stimulated glucose transporter-4 (GLUT4) trafficking in 3T3-L1 adipocytes. J. Biol. Chem. 1999, 274, 17619–17625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Ellis, B.C.; Graham, L.D.; Molloy, P.L. CRNDE, a long non-coding RNA responsive to insulin/IGF signaling, regulates genes involved in central metabolism. Biochim. Biophys. Acta 2014, 1843, 372–386. [Google Scholar] [CrossRef] [Green Version]
  91. Zhang, Y.; Xi, X.; Mei, Y.; Zhao, X.; Zhou, L.; Ma, M.; Liu, S.; Zha, X.; Yang, Y. High-glucose induces retinal pigment epithelium mitochondrial pathways of apoptosis and inhibits mitophagy by regulating ROS/PINK1/Parkin signal pathway. Biomed. Pharmacother. 2019, 111, 1315–1325. [Google Scholar] [CrossRef] [PubMed]
  92. Yanagi, Y. Role of Peoxisome Proliferator Activator Receptor gamma on Blood Retinal Barrier Breakdown. PPAR Res. 2008, 2008, doi. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Lin, J.H.; Lavail, M.M. Misfolded proteins and retinal dystrophies. Adv. Exp. Med. Biol. 2010, 664, 115–121. [Google Scholar] [PubMed] [Green Version]
  94. Lundkvist, A.; Reichenbach, A.; Betsholtz, C.; Carmeliet, P.; Wolburg, H.; Pekny, M. Under stress, the absence of intermediate filaments from Muller cells in the retina has structural and functional consequences. J. Cell Sci. 2004, 117, 3481–3488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Rossignol, R.; Ranchon-Cole, I.; Paris, A.; Herzine, A.; Perche, A.; Laurenceau, D.; Bertrand, P.; Cercy, C.; Pichon, J.; Mortaud, S.; et al. Visual sensorial impairments in neurodevelopmental disorders: Evidence for a retinal phenotype in Fragile X Syndrome. PLoS ONE 2014, 9, e105996. [Google Scholar] [CrossRef] [Green Version]
  96. Adinolfi, E.; Giuliani, A.L.; De Marchi, E.; Pegoraro, A.; Orioli, E.; Di Virgilio, F. The P2X7 receptor: A main player in inflammation. Biochem. Pharmacol. 2018, 151, 234–244. [Google Scholar] [CrossRef] [PubMed]
  97. Di Virgilio, F.; Tang, Y.; Sarti, A.C.; Rossato, M. A rationale for targeting the P2X7 receptor in Coronavirus disease 19. Br. J. Pharmacol. 2020, 177, 4990–4994. [Google Scholar] [CrossRef] [PubMed]
  98. Di Virgilio, F.; Dal Ben, D.; Sarti, A.C.; Giuliani, A.L.; Falzoni, S. The P2X7 Receptor in Infection and Inflammation. Immunity 2017, 47, 15–31. [Google Scholar] [CrossRef] [Green Version]
  99. Savio, L.E.B.; De Andrade Mello, P.; Da Silva, C.G.; Coutinho-Silva, R. The P2X7 Receptor in Inflammatory Diseases: Angel or Demon? Front. Pharmacol. 2018, 9, 52. [Google Scholar] [PubMed] [Green Version]
  100. Reichenbach, A.; Bringmann, A. Purinergic signaling in retinal degeneration and regeneration. Neuropharmacology 2016, 104, 194–211. [Google Scholar] [CrossRef] [PubMed]
  101. Calzaferri, F.; Ruiz-Ruiz, C.; Diego, A.M.G.; Pascual, R.; Méndez-López, I.; Cano-Abad, M.F.; García, A.G. The purinergic P2X7 receptor as a potential drug target to combat neuroinflammation in neurodegenerative diseases. Med. Res. Rev. 2020, 40, 2427–2465. [Google Scholar] [CrossRef] [PubMed]
  102. Appelbaum, T.; Santana, E.; Aguirre, G.D. Strong Upregulation of Inflammatory Genes Accompanies Photoreceptor demise in Canine Models of Retinal Degeneration. PLoS ONE 2017, 12, e0177224. [Google Scholar]
  103. Block, M.L.; Zecca, L.; Hong, J.S. Microglia-mediated neurotoxicity: Uncovering the molecular mechanism. Nat. Rev. Neurosci. 2007, 8, 57–69. [Google Scholar] [CrossRef] [PubMed]
  104. Subhramanyam, C.S.; Wang, C.; Hu, Q.; Dheen, S.T. Microglia-mediated neuroinflammation in neurodegenerative diseases. Semin. Cell Dev. Biol. 2019, 94, 112–120. [Google Scholar] [CrossRef]
  105. Peng, B.; Xiao, J.; Wang, K.; So, K.F.; Tipoe, G.L.; Lin, B. Suppression of microglial activation is neuroprotective in a mouse model of human retinitis pigmentosa. J. Neurosci. 2014, 34, 8139–8150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wang, S.K.; Xue, Y.; Cepko, C.L. Microglia modulation by TGF-Beta1 protects cones in mouse models of retinal degeneration. J. Clin. Investig. 2020, 130, 4360–4369. [Google Scholar] [PubMed]
  107. Smith, J.A.; Das, A.; Ray, S.K.; Banik, N.L. Role of pro-inflammatory cytokines released from microglia in neurodegenerative diseases. Brain Res. Bull. 2012, 87, 10–20. [Google Scholar] [CrossRef]
  108. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.-S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  109. Chitnis, T.; Weiner, H.L. CNS inflammation and neurodegeneration. J. Clin. Investig. 2017, 127, 3577–3587. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, M.; Ma, W.; Zhao, L.; Fariss, R.N.; Wong, W.T. Adaptive Muller Cell Responses to Microglial Activation Mediate Neuroprotection and Coordinates Inflammation in The Retina. J. Neuroinflamm. 2011, 8, 173. [Google Scholar] [CrossRef] [Green Version]
  111. Taylor, R.A.; Chang, C.F.; Goods, B.A.; Hammond, M.D.; Mac Grory, B.; Ai, Y.; Steinschneider, A.F.; Renfroe, S.C.; Askenase, M.H.; McCullough, L.D.; et al. TGF-beta1 modulates microglial phenotype and promotes recovery after intracerebral hemorrhage. J. Clin. Investig. 2017, 127, 280–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. O’Koren, E.G.; Yu, C.; Klingeborn, M.; Wong, A.Y.; Prigge, C.L.; Mathew, R.; Kalnitsky, J.; Msallam, R.A.; Silvin, A.; Kay, J.N.; et al. Microglial function is distinct in different anatomical locations during retinal homeostasis and degenration. Immunity 2019, 50, 723–737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Rutar, M.; Valter, K.; Natoli, R.; Provis, J.M. Synthesis and propagation of complement C3by microglia/monocytes in the aging retina. PLoS ONE. 2014, 9, e93343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Silverman, S.M.; Ma, W.; Wang, X.; Zhao, L.; Wong, W.T. C3- and CR3- dependent microglial clearence protects photoreceptors in retinitis pigmentosa. J. Exp. Med. 2019, 216, 1925–1943. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Main molecular mechanisms occurring during oxidative stress in RP.
Figure 1. Main molecular mechanisms occurring during oxidative stress in RP.
Antioxidants 10 00848 g001
Table 1. Altered pathway involved in endogenous antioxidant defence and their effects on retinal tissue.
Table 1. Altered pathway involved in endogenous antioxidant defence and their effects on retinal tissue.
Pathway InvolvedEffectsAssociation with RPReferences
CHRONIC ACTIVATION OF PERK AND IRE1Misfolded proteins: UPR activationActivation of pro-apoptotic programs
Pro-inflammatory signalling
Dysfunctional autophagy, free cytosolic Ca2+
Altered protein synthesis rate in the retina and retinal degeneration[4,5]
MUTYH MUTATIONDNA repairFormation of single-strand breaks (SSBs) of DNA
Disturbed homeostasis and cell death
Oxidative microglial activation
Retinal degeneration and neuroinflammation in RP [16,18,20]
CERKL MUTATIONOxidative stress protectionActivation of pro-apoptotic programsAccelerated progression of retinal neurodegeneration. [21,26,27]
GLO1 MUTATIONDetoxification of cytotoxic products of glycolysisInactivation of antioxidant enzymes (glutathione peroxidase and SOD enzymes)Hyperinflammation and permanent tissue damage Vascular dysfunction
Altered transduction pathways and genetic expression in EPR
[28,30,33,38]
Table 2. GLO1-related genes mutations: RPE dysfunction and photoreceptors damage.
Table 2. GLO1-related genes mutations: RPE dysfunction and photoreceptors damage.
GeneEffects of MutationConsequencesReferences
AUTS2, ANKHAlteration of actin filament structure and activityRPE apoptosis[41]
SIK3, IPO3, MRPS33Dysregulation of energy metabolism and translation machineryCell death [56,58,59]
ARHGAP2, PTPN13Alteration of cell migration and proliferation
Modification of cell polarity, cell adhesion, Golgi regulation
Impairment of intracellular trafficking and glucose homeostasis
RPE apoptosis[42,43,44,45]
FMNL2Alteration of actin polymerization and organization of the cytoskeleton Alteration of vesicular trafficking of RPE cells[46,47]
UBC, MYO18A, EPS15, ANKHImpairment of intracellular transport processesInfluence of vesicular trafficking of RPE cells (essential for POS renewal and visual cycle intermediate regeneration)
AGE accumulation
[48,49,50,51]
RFFL, FBXW2, CAND1ER stress
Accumulation of misfolded proteins
AGEs and ROS production
Cell death
Alteration of cellular respiration process
Impairment of glycolytic metabolism in RPE cells.
[52,53,54]
SIK3Defect in energy metabolism
Decrease of mitochondrial respiration
Up-regulation of autophagy
Cellular antioxidant mechanisms impairment
Alteration of cellular respiration process
[56,57]
IPO7Ribosomal biogenesis stress
Nucleolar morphology changes
p53-dependent growth arrest[58]
MRPS33, MORC4, MCPH1, NFIA, CTIF, LMBRD1Damage of mitochondrial protein synthesis
Arrests in DNA damage repair
Impairment of mitotic exit and cell differentiation
Impairment of mRNA and protein quality control
Alteration of transport and metabolism of cobalamin
RPE apoptosis and retinal degeneration
Decreased
distribution of cyanocobalamin (vitamin B12) from blood to retina
[59,62,63,64]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gallenga, C.E.; Lonardi, M.; Pacetti, S.; Violanti, S.S.; Tassinari, P.; Di Virgilio, F.; Tognon, M.; Perri, P. Molecular Mechanisms Related to Oxidative Stress in Retinitis Pigmentosa. Antioxidants 2021, 10, 848. https://doi.org/10.3390/antiox10060848

AMA Style

Gallenga CE, Lonardi M, Pacetti S, Violanti SS, Tassinari P, Di Virgilio F, Tognon M, Perri P. Molecular Mechanisms Related to Oxidative Stress in Retinitis Pigmentosa. Antioxidants. 2021; 10(6):848. https://doi.org/10.3390/antiox10060848

Chicago/Turabian Style

Gallenga, Carla Enrica, Maria Lonardi, Sofia Pacetti, Sara Silvia Violanti, Paolo Tassinari, Francesco Di Virgilio, Mauro Tognon, and Paolo Perri. 2021. "Molecular Mechanisms Related to Oxidative Stress in Retinitis Pigmentosa" Antioxidants 10, no. 6: 848. https://doi.org/10.3390/antiox10060848

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