Next Article in Journal
Leukocyte Telomere Length as a Molecular Biomarker of Coronary Heart Disease
Next Article in Special Issue
The Genetic Variants in the Renin-Angiotensin System and the Risk of Heart Failure in Polish Patients
Previous Article in Journal
Identification of Runs of Homozygosity Islands and Genomic Estimated Inbreeding Values in Caqueteño Creole Cattle (Colombia)
Previous Article in Special Issue
Genetics and Sport Injuries: New Perspectives for Athletic Excellence in an Italian Court of Rugby Union Players
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Genetic Mechanisms in Age-Related Macular Degeneration

by
Aumer Shughoury
1,
Duriye Damla Sevgi
1 and
Thomas A. Ciulla
1,2,3,*
1
Department of Ophthalmology, Eugene and Marilyn Glick Eye Institute, Indiana University School of Medicine, Indianapolis, IN 46202, USA
2
Clearside Biomedical, Inc., Alpharetta, GA 30005, USA
3
Midwest Eye Institute, Indianapolis, IN 46290, USA
*
Author to whom correspondence should be addressed.
Genes 2022, 13(7), 1233; https://doi.org/10.3390/genes13071233
Submission received: 20 May 2022 / Revised: 30 June 2022 / Accepted: 5 July 2022 / Published: 12 July 2022
(This article belongs to the Special Issue Feature Papers: Molecular Genetics and Genomics)

Abstract

:
Age-related macular degeneration (AMD) is among the leading causes of irreversible blindness worldwide. In addition to environmental risk factors, such as tobacco use and diet, genetic background has long been established as a major risk factor for the development of AMD. However, our ability to predict disease risk and personalize treatment remains limited by our nascent understanding of the molecular mechanisms underlying AMD pathogenesis. Research into the molecular genetics of AMD over the past two decades has uncovered 52 independent gene variants and 34 independent loci that are implicated in the development of AMD, accounting for over half of the genetic risk. This research has helped delineate at least five major pathways that may be disrupted in the pathogenesis of AMD: the complement system, extracellular matrix remodeling, lipid metabolism, angiogenesis, and oxidative stress response. This review surveys our current understanding of each of these disease mechanisms, in turn, along with their associated pathogenic gene variants. Continued research into the molecular genetics of AMD holds great promise for the development of precision-targeted, personalized therapies that bring us closer to a cure for this debilitating disease.

1. Introduction

Age-related macular degeneration (AMD) is one of the most common causes of irreversible blindness, affecting up to 10% of individuals over the age of 45 worldwide [1]. Progressive central vision loss due to AMD has a significant impact on patient quality of life [2,3] and is associated with substantial psychosocial [2,3,4,5] and economic [6] burden. As the global population ages over the next two decades, AMD is projected to affect up to 288 million individuals [1] and prevalence of AMD in the USA is expected to double [7].
AMD is characterized by the macular accumulation of sub-retinal pigment epithelial (RPE) and/or subretinal deposits known as drusen or reticular pseudodrusen, respectively, typically beginning after the seventh decade of life (Figure 1). This leads to progressive degeneration of the RPE, which plays critical metabolic and regulatory roles in supporting the health of overlying photoreceptors [8,9]. Local photoreceptor degeneration ensues, with consequent progressive loss of central visual acuity.
Several pathophysiologic mechanisms have been implicated in the development of AMD, all of which converge on RPE dysfunction and degeneration. Oxidative stress appears to play a strong role [10]. Cigarette smoking—which increases oxidant load and impairs antioxidant defense mechanisms [11,12]—has been therefore identified as the most significant modifiable risk factor influencing the development of AMD [13]. Conversely, increased dietary and supplemental intake of antioxidant vitamins and minerals is one of the only interventions proven to slow the progression to advanced forms of disease [14]. Cumulative oxidative damage with age may cause structural degeneration of the choriocapillaris, resulting in decreased blood flow to the RPE and photoreceptors [15]. Impaired circulation reduces the clearance of lipids and cellular byproducts, which accumulate as drusen [16]. In turn, lipid accumulation can induce remodeling of the extracellular matrix (ECM) and stimulate an inflammatory response. Complex interplay between these pathogenic processes [17] ultimately causes disease progression to either tissue atrophy or macular neovascularization (MNV, formerly known as choroidal neovascularization [18]). These two endpoints represent the two major forms of advanced AMD: geographic atrophy (GA, traditionally known as “dry” AMD) and neovascular AMD (traditionally known as “wet” AMD). However, the advent of optical coherence tomography (OCT) angiography has allowed cases of nonexudative neovascularization to be more readily identified, blurring the distinction between these traditional categories [19].
Age is the strongest demographic risk factor for the development of AMD [13,14,20]. Environmental and lifestyle factors, such as cigarette smoking [12,13,21,22,23] and diet [24,25] can modify this risk. Over the past two decades, the contribution of specific genetic factors to AMD risk has been increasingly recognized.
The genetic influence on AMD risk has long been suspected [26]. Historically, evidence for the genetic contribution to AMD came from epidemiological studies of racial disparities [27,28] and twin–twin concordance studies [29,30,31,32,33]. Over the course of history, millions of mutations in the human genome have accumulated, resulting in genetic variations such as single nucleotide polymorphisms (SNPs). These single base changes in the DNA sequence occur relatively frequently (>1 percent) and have been the subject of much study [34,35,36,37,38,39,40], especially prior to the wide availability of whole genome sequencing. In AMD, familial-based linkage [41,42,43,44,45] and genome-wide association studies (GWAS) [46,47,48,49,50,51,52] have rapidly uncovered a large number of common and rare disease-associated gene variants [53]. Such genetic research has increased our understanding of this complex disease [17,54,55] and spurred the development of novel gene-based therapies for the treatment of AMD [56,57].
This review surveys the literature to date implicating genetic factors in the pathogenesis of AMD. Genes identified as contributing to AMD risk are largely associated with the major pathophysiological mechanisms underlying the development of AMD: immune dysregulation and inflammation, ECM disruption, lipid accumulation, angiogenesis, and cellular apoptosis. In what follows, we discuss the major gene variants to date that have been associated with each of these pathways.

2. Immune Dysregulation and the Complement System

The innate immune system appears to play a key role in the development and progression of AMD. For example, immune complex deposition has been strongly implicated in the formation and biomolecular makeup of drusen [58]. It has been hypothesized that drusen may be biomarkers of immune-mediated processes occurring at the RPE–Bruch’s Membrane (BrM) interface in the aging retina [59]. Moreover, localized inflammation and microglial cell recruitment appear to be key mediators of AMD pathogenesis [54].
Dysregulation of the complement system (Figure 2) has been strongly associated with the development of AMD [58,60] and the most commonly identified high-risk genetic variants involve genes coding for key components of the complement system [51,54]. The complement cascade consists of an array of specialized plasma proteins and enzymes (complement factors) that react with one another in complex patterns to target foreign pathogens and tag them for destruction by immunologic phagocytosis [61]. Three distinct pathways have been characterized: an antibody-dependent ‘classical’ pathway, an antibody-independent ‘alternative’ pathway, and a ‘lectin’ pathway, which involves binding to specific sugars on the surface of microorganisms. All of these pathways trigger a localized inflammatory response when activated via convergence on complement factor 3 (C3), which ultimately leads to cleavage of complement factor 5 (C5) to form key terminal fragments (C5a and C5b). C5b is involved in the formation of a membrane attack complex (MAC), along with C6, C7, C8, and C9 complement factors. MAC, in turn, disrupts the lipid bilayer that forms the extracellular membrane of invading pathogens, resulting in cell lysis [61]. At sub-lytic concentrations, MAC can also induce several localized inflammatory reactions [62,63,64]. Importantly, the regulatory factors involved in these complement pathways limit complement activation specifically to the pathogenic surface [61].

2.1. Complement Factor H (CFH)

The complement factor H (CFH) gene on chromosome 1 was the earliest gene found to play a role in the development of AMD [66,67,68,69,70]. CFH has since been recognized as the major susceptibility locus for the development of AMD [71] in Caucasians and Asians [72,73], and CFH is deposited in significant concentrations in drusen [67,74,75]. While the exact mechanism by which CFH dysfunction confers AMD risk has not yet been established, the highest-risk CFH variants tend to affect the functional domains or serum expression of the enzyme [76]. CFH is a complement regulatory protein expressed in the BrM and sub-retinal space [77]. In the complement cascade, it functions to inhibit the alternative pathway by acting as a cofactor to complement factor I (CFI) in the inactivation of complement component 3b (C3b) [78]. CFH additionally acts to promote the decay of C3 convertase [79]. Consequently, compromise of CFH function leads to dysregulation of the complement cascade [80]. CFH also downregulates proinflammatory activity by binding to C-reactive protein (CRP) [81,82] and the immune checkpoint molecule CD47 [83]. Finally, recent functional studies have suggested that in vitro knockdown of CFH results in increased MAC deposition in choroidal endothelial cells (CECs), while overexpression of CFH protects against MAC deposition [84]. Thus, CFH dysfunction may lead to the development of localized chronic inflammation in the retina [85]. Beyond its immunomodulatory roles, it has also been suggested that CFH may independently act as a protective antioxidant in the RPE [86,87] and play key roles in lipid metabolism [88,89]. CFH involvement in multiple pathogenic pathways may help explain its strong association with AMD.
CFH gene variants have been found to largely affect the functional activity of CFH, rather than serum levels [90]. Two common loss-of-function CFH variants—rs1061170 (Y402H) [66,67,68,69] and rs1410996 [91]—account for a significant portion of AMD genetic risk across populations [42,92,93,94,95,96,97,98,99]. These two variants are notably found at rates nearly seven-fold higher in European compared to Asian populations [100], potentially accounting for some of the difference in AMD prevalence between these two populations. The mechanism by which these variants increase AMD risk has not yet been established, although the Y402H variant has been associated with uncontrolled complement system activation [80,101]. Interestingly, although the Y402H variant is associated with a poor response to antioxidant therapy in non-neovascular AMD [98], it is associated with a strong positive response to anti-VEGF therapy in neovascular AMD [102].
A number of other rare CFH variants have been identified [42,95,101,103,104,105], many of which are highly penetrant and correlate with more severe or extensive disease phenotypes [45,106,107,108,109]. For example, the rs121913059 (p.Arg1210Cys) mutation has been found to confer a more than 20-fold increase in AMD risk [45].
In addition to significantly increasing the risk of AMD in general [110], CFH mutations have also been associated with a higher risk of particular AMD phenotypes compared to other gene variants. For example, some CFH mutations have been associated with a slightly stronger risk of progression to GA as opposed to MNV [54]. Conversely, CFH mutations may also increase the genetic risk of progression to MNV after zinc and antioxidant supplementation, as found in AREDS vitamin formulations [44,111,112,113]. Certain CFH variants, such as the Y402H and rs1410966 variants, have been associated with a higher risk of peripheral retinal involvement [114]. Additionally, compared to the common rs1061170 variant, rare CFH variants rs800292, rs1410996, and rs1329428 are associated with a relatively poor response to anti-VEGF therapy [102]. As with Y402H, many of these rare variants exhibit differences in ethnic distribution and may account for differences in AMD phenotypes across populations [115].

2.2. Complement Factor I (CFI)

The CFI gene on chromosome 4 has also been implicated in AMD. CFI functions in the alternative complement pathway to inactivate C3b in conjunction with CFH. Rare gene variants affecting CFI expression [116,117,118,119,120] and/or disrupting its functional domain [44,117,119,121] impair its ability to modulate complement activation and have been highly associated with the development of AMD [67,117]. Such variants include the p.Gly119Arg [117] and p.Leu131Arg [121] missense mutations, both of which have been associated with reduced CFI concentration and activity. An additional variant, rs915370426 (p.Pro553Ser), appears to confer high risk despite normal serum CFI concentration [121]. An investigational gene therapy, GT005 (Gyroscope Therapeutics) [122], has recently been developed to increase intraocular CFI expression in the treatment of GA. Three Phase I/II multicenter, randomized, controlled trials (“FOCUS”, NCT03846193; “EXPLORE”, NCT04437368; and “HORIZON”, NCT04566445) are currently evaluating the safety of subretinal administration of GT005 in patients with GA. Preliminary data have suggested a positive safety profile and promising evidence of a sustained increase in intravitreal CFI levels following treatment.

2.3. Complement Component 3 (C3)

Several rare variants in the complement C3 gene on chromosome 19 have also been identified with an increased risk of AMD. The most common variants include rs2230199 (p.Arg102Gly) [123,124,125] and the associated polymorphism rs1047286 (p.Pro292Leu) [124,125]. These variants are particularly common in Caucasian populations and rare in Asian and African populations [54,124,125]. The rs2230199 variant is associated with impaired CFH binding, conferring resistance to CFH-mediated inactivation [126]. It has also been shown to more efficiently activate the alternative complement pathway [126]. An additional rare variant, rs147859257 (p.Lys155Gln), has been found to impair the inactivation of C3b by the CFI/CFH complex, resulting in constitutive activation of the alternative cascade pathway and a significantly increased risk of AMD [43,44,45]. Two further C3 variants, rs117793540 (p.Arg735Trp) and rs2230210 (p.Ser1619Arg), have been inconsistently associated with AMD risk, with differing results across cohort studies [44,127]. Finally, a recent meta-analysis found no association between the two other known major C3 polymorphisms—rs2230205, rs2250656—and AMD [125]. In fact, the rs2250656 variant may have a protective effect against the development of neovascular AMD in the Chinese population [128]. A randomized phase 2 trial investigating intravitreal injection of the C3 inhibitor Pegcetacoplan for the treatment of GA has recently published positive results, with data suggesting statistically significant effects on the progression of GA [129]; two phase 3 trials are currently underway, with GA area growth at 12 months showing a statistically significant reduction in one study (“Oaks” NCT03525613) and a trend of reduction in the other (“Derby”, NCT03525600).

2.4. Complement Component 5 (C5)

Few studies have evaluated the association between AMD and complement component 5 (C5). C5 is cleaved into its bioactive fragment in one of the final steps of the complement cascade (Figure 2). The role of C5 in AMD has been suggested by its presence in drusen [130,131] as well as the observation of elevated serum C5a levels in AMD [132]. It has also been suggested that C5a can induce VEGF expression, leading to the development of MNV [133]. MNV mouse models have also demonstrated that RPE and choroidal C5a levels are elevated in laser-induced MNV, while genetic knockout or pharmacologic blockade of C5a receptors reduces VEGF expression and MNV formation after laser injury [133].
To date, most human population studies have not demonstrated a significant association between known C5 SNPs and AMD [123,124,134,135]. However, the prominent role of C5 in the complement cascade makes it a prime target for pharmacologic downregulation of the complement system in the treatment of AMD. A recent phase 3 study (“Gather 1” NCT02686658) demonstrated that intravitreal injection of avacincaptad pegol, a C5 inhibitor, reduced GA enlargement versus sham over a 12-month period [136]. Another confirmatory phase 3 study is underway (“Gather 2” NCT04435366). Further research is required to clarify the exact role of C5 in AMD and to uncover genetic variants that may impact risk.

2.5. Complement Component 9 (C9)

Recently, variants in the gene coding complement component 9 (C9) have been associated with a significantly increased risk of AMD [44,51,121,137] and progression to more advanced stages of the disease [137,138]. C9 is a key protein in the terminal complement pathway, joining with C5–C8 to form MAC, as discussed above [61]. Patients with advanced AMD exhibit higher serum MAC concentrations [139]. The rare variant p.Pro167Ser (rs34882957) has been significantly associated with AMD [44,139] and appears to be associated with an increased serum concentration of C9 [121], as well as with increased polymerization rates [139,140] and hemolytic function [139]. The C9 p.Pro167Ser variant may therefore represent a gain-of-function mutation, leading to increased MAC formation and lytic activity in the terminal complement pathway. Other rare C9 variants, such as p.Met45Leu, p.Phe62Ser, and p.Ala529Thr, have been associated with increased C9 expression in AMD patients, though without an increase in lytic activity [140], while other variants such as p.Arg118Trp and p.Thr170Ile have been found to confer risk without elevated C9 levels [140]. Conversely, the nonsense C9 mutation p.Arg95* found commonly in Japanese populations [141] confers a strong (nearly 5-fold) protective effect against the development of AMD [142] and has also been correlated with decreased VEGF levels and reduced risk of progression to MNV [142].

2.6. Complement Component 2 (C2) and Complement Factor B (CFB)

Rare variants in the complement component 2 (C2) and complement factor B (CFB) genes on chromosome 6 have also been associated with a strong protective effect against the development of AMD [143,144,145,146,147]. C2 and CFB are expressed in the neural retina, Bruch’s membrane, and choroid [143], and function as activators of the classical and alternative complement cascades, respectively [148]. Variants reducing the function of these enzymes therefore dampen the activity of the complement cascade [149], potentially explaining their protective effects against the development of AMD. Such variants include C2 rs9332739, rs547154, and rs429608, as well as CFB rs9332739, rs547154, rs4151667, and rs641153 [93,145,146,147,150]. Two 2012 meta-analyses found that these variants may reduce the risk of AMD by nearly half [146,147]. In the Japanese population, C2 rs547154 and CFB rs541862 additionally protect against the development of MNV [151]. Interestingly, antioxidant supplementation has also been found to be more effective in retarding the progression of non-exudative AMD in patients with these protective C2 and CFB variants [98].

2.7. Complement Factor D (CFD)

Finally, complement factor D (CFD) has also been implicated in the development of AMD. CFD functions as a rate-limiting enzyme in the activation of the alternative complement pathway by cleaving and activating CFB [152,153]. Six CFD variants have been identified (rs1683564, rs35186399, rs1683563, rs3826945, rs34337649, and rs1651896). These variants have largely not been significantly associated with AMD [154]. A small case-control series did find rs3826945 to be correlated with increased AMD risk [153], however this association could not be demonstrated in a separate population [93]. Interestingly, intravitreal injection of lampalizumab, a selective CFD inhibitor, was not found to reduce GA enlargement versus sham during 48 weeks of treatment in two Phase 3 trials (“Chroma” and “Spectri,” NCT02247479 and NCT02247531) [155].

3. Extracellular Matrix (ECM) Remodeling

The ECM is a supportive framework consisting of the stroma and basement membrane between the epithelial and endothelial tissue layers. The retinal ECM consists of a five-layered BrM, bridging the space between the choroid and RPE [156]. It plays a crucial role in the physical support and remodeling of the RPE, as well as the exchange of biomolecules, oxygen, nutrients, and waste products between the RPE and choriocapillaris [157]. The structure of BrM is controlled by a balance between local proteolytic enzymes known as matrix metalloproteinases (MMPs) and their tissue inhibitors (TIMPs) [158]. Specifically, MMP-1, MMP-2, MMP-3, MMP-9, TIMP-1, TIMP-2, and TIMP-3 are found at high concentrations in BrM and are critical to maintaining its structure and function [159].
Changes in the structure and function of BrM have been strongly implicated in the pathogenesis of AMD [160]. Impaired MMP-mediated ECM degradation has been associated with both aging and macular degeneration, resulting in thickening of BrM [161]. As BrM increases in thickness, its filtration capacity declines, leading to the focal accumulation of waste products and the formation of drusen [157]. Additionally, reduced permeability of BrM impairs the transport of critical metabolites and nutrients between the RPE and choroid, ultimately leading to RPE and photoreceptor degeneration [162]. It is therefore hypothesized that an imbalance between MMP and TIMP activity may play a role in AMD pathogenesis [163]. Several gene variants affecting local MMP and TIMP expression have been associated with AMD risk [160].

3.1. Tissue Inhibitor of Metalloproteinases (TIMPs)

Variants in TIMP-3 have been particularly strongly linked to the development of AMD [46,51,164,165]. TIMP-3 is expressed in the RPE adjacent to BrM [166] and localizes directly to the ECM [167]. In the aging retina, TIMP-3 accumulates in BrM [162,168] and particularly elevated concentrations of TIMP-3 are found in AMD [162]. As an MMP inhibitor, TIMP-3 excess results in impaired ECM turnover [169] and pathologic thickening of BrM, leading to RPE and photoreceptor atrophy [162,170]. Additionally, TIMP-3 physiologically acts as a potent local inhibitor of angiogenesis, competitively impairing VEGF binding to its receptor [171]. Decreased TIMP-3 activity thus increased retinal VEGF activity and angiogenesis. Notably, genetic variants that reduce TIMP-3 expression are the main etiology of Sorsby’s fundus dystrophy, a rare autosomal dominant macular dystrophy presenting with features of neovascular AMD at a young age [172].
Variants in TIMP-3 contribute significantly to the genetic burden in patients with AMD [51]. In their 2016 GWAS, Fritsche et al. found nine rare variants in TIMP-3 to be cumulatively associated with a >30-fold increased risk of AMD [51]. In addition to contributing to the overall risk of AMD, it is thought that TIMP-3 dysfunction or suppression may contribute to the development of MNV [173]. The rs5754227 [51], rs713685 [174], rs743751 [174], and rs5749482 [175] TIMP-3 intron variants have been particularly associated with an increased risk of AMD. Conversely, the rs9621532 TIMP-3 variant has been found to have a slight to moderate protective effect against AMD in some studies [47,49,165,176], as well as a protective effect against the development of MNV [177]—although other studies have found no association with AMD risk [178,179]. Finally, TIMP-3 variants rs6518799, rs756481, rs5749498, rs12170368, and rs1427385 have not been found to be associated with AMD [174,180]. Beyond TIMP-3, the TIMP-2 gene (specifically polymorphism rs8179090) has also been identified and found to be inconsistently associated with decreased AMD risk [181,182].

3.2. Matrix Metalloproteinases (MMPs)

Several studies have also demonstrated an association between MMP gene mutations and AMD. MMP-2 is currently the most widely studied gene in this context. Cheng et al. demonstrated that the T allele (TT and CT genotypes) of the rs243865 MMP-2 polymorphism is protective against AMD [173], while Liutkeviciene et al. found an association between the homozygous CC genotype and hard drusen development in AMD [183]. Conversely, several studies [181,184,185,186] including a recent meta-analysis [187] have demonstrated no statistically significant association between the MMP-2 rs243865 variant and AMD. Beyond rs243865, the MMP-2 rs2287074 variant has been found to have a protective effect against AMD in one study [186], while the rs243866 and rs2285053 variants have not demonstrated an association with AMD [182,184,188].
The MMP-9 rs142450006 [51,189,190], rs3918241 [182], rs3918242 [182,184], rs4810482 [190], rs17576 [190], and rs17577 [190] variants are associated with an increased risk of AMD and progression to MNV. Similarly, the MMP-9 CA (13–27) microsatellite expansion variant has also been associated with progression to MNV, with risk directly proportional to the length of microsatellite expansion [191].
No association has been found between the known variants in MMP-1 (rs1799756) [192], MMP-3 (rs3025058) [184,193], or MMP-7 (rs11568818) [182,192] and AMD.

3.3. Other Extracellular Matrix Components

Other genes coding for ECM components have been studied in association with AMD. Collagen Type 8, α 1 (COL8A1), is a short-chain component of type VIII collagen found in the basement membrane of many components of the human eye [194], including BrM [195]. A common gene variant, rs140647181, near the COLA8A1 gene has been associated with increased AMD risk [51]. A wide array of additional rare, protein-altering variants in the COL8A1 gene itself has also been independently associated with AMD [195,196]. Similarly, the rs1999930 gene variant near the collagen matrix protein-coding genes COL10A1 (coding for α chain of type X collagen) and FRK (coding for fyn-related kinase) have also been associated with increased AMD risk [48], as they have copy number variants in the EFEMP1 gene coding for fibulin 3, a matrix glycoprotein [197]. The mechanism by which many of these gene alterations influence AMD risk remains unclear, although it is suspected that these variants may alter the integrity of BrM [195] or cause ECM protein accumulation in BrM as drusen [198].

4. Lipid Metabolism

Lipid accumulation in BrM is strongly implicated in AMD pathogenesis [199,200]. Much like the intima of atherosclerotic arterial walls, lipids, and cholesterol accumulate significantly in human BrM with age [201], causing pathologic thickening and dysfunction of BrM [202]. Additionally, the RPE basolaterally secretes large lipoproteins (Figure 3) containing apolipoproteins B and E into BrM as a byproduct of photoreception [200]. In AMD, the reduced clearance of lipids from BrM may be a key mechanism in the formation of drusen [200,203]. Lipids are among the most significant components of drusen [204,205,206], accounting for over 40% of the drusen volume [205]. Large lipoproteins containing apolipoproteins B and E are secreted basolaterally by the RPE into BrM; these accumulate in both the sub-RPE space as soft drusen, as well as in the subretinal space as drusenoid deposits (also known as “pseudodrusen”) [207]. Pathologic accumulation of lipids in these spaces leads directly to RPE and photoreceptor loss in AMD [208] as in the critical flow of biomolecules across the BrM between the RPE-photoreceptor complex and choriocapillaris is impaired [203]. Drusen expansion additionally disrupts the local cellular architecture by driving RPE cells into the retina, resulting in RPE degeneration and atrophy [203]. Moreover, as in atherosclerotic disease [209], oxidation of lipoproteins in BrM and the local inflammatory response they induce [210] likely also play a key role in the pathogenesis and progression of AMD [59].
Variants in several genes coding for proteins involved in lipid metabolism and cholesterol transport (Figure 4) are associated with the risk of AMD. Genes involved in the structure and function of high-density lipoprotein (HDL) are particularly implicated. HDL functions in the removal and transport of excess cholesterol from peripheral tissues to the liver [211]. HDL also exerts localized anti-inflammatory effects by inhibiting monocyte activity [212] and is a major transporter of lutein and zeaxanthin in the retina [213].

4.1. Apolipoprotein E (ApoE)

Lipoproteins function in the human body as transport vehicles for water-insoluble cholesterol. In particular, HDL-mediated transport of cholesterol to the liver is the main mechanism by which peripheral tissues recycle excess cholesterol [211]. The formation and metabolism of such plasma lipoproteins is regulated by associated apolipoproteins [214]. Apolipoprotein E (ApoE) is significantly involved in regulating cholesterol metabolism and lipid transport in nervous tissue [215]. ApoE dysfunction has therefore notably been found to play a key role in the pathogenesis of Alzheimer’s disease and other neurodegenerative disorders [Huang 2014]. It is also a major component of drusen [215]. ApoE has consequently been a particular focus of AMD research.
The APOE gene is among the earliest genes associated with AMD [41]. Three major allelic variants, ApoE2, ApoE3, and ApoE4, have historically been identified as giving rise to structurally and functionally distinct protein products [216]. The ApoE3 allele is the most common, occurring in more than 75% of chromosomes worldwide [217], and it is considered the wild-type, “neutral” allele. The ApoE4 allele, with an estimated frequency of 15% worldwide [218], is associated with a strong protective effect against AMD [41,100,219,220,221,222,223,224,225,226]. A 2006 meta-analysis estimated that the ApoE4 allele confers up to a 40% reduction in the risk of developing AMD [100]. On the other hand, ApoE2, with an estimated frequency of 8% worldwide [217], has been associated with a slightly increased risk of AMD in several independent studies [222,223,227,228,229], although a recent large meta-analysis did not find any association between ApoE2 and AMD [224]. A large study of a Brazilian population failed to find an association between ApoE2 and AMD [225]. However, ApoE2 may play a role in the pathogenesis of MNV by upregulating angiogenic factors [230].

4.2. Hepatic Lipase (LIPC)

Plasma HDL levels are strongly regulated by hepatic lipase (LIPC) [231], which is expressed in the subretinal space and involved in intra-retinal lipid transport [232]. Several studies have demonstrated a protective effect of LIPC variants (rs493258, rs10468017, rs9621532, rs11755724, rs493258, rs509859, rs12637095) against the development of AMD [47,48,233,234,235,236,237]. Two of these LIPC variants—rs493258 and rs10468017—have been specifically associated with increased systemic HDL levels [238]. However, the mechanism by which these variants reduce AMD risk is unclear, as elevated HDL levels alone do not appear to account for the observed risk reduction [47,233]. In fact, elevated HDL levels may instead be associated with an increased risk of AMD [60]. It has been suggested that these variants may increase the HDL-mediated efficiency of carotenoid delivery to the retina [47]. The rare LIPC rs13095226 and rs3748391 variants have also been associated with a slightly increased AMD risk [47]. Finally, LIPC variants may be associated with a poorer response to anti-VEGF therapy [239].

4.3. Cholesteryl Ester Transfer Protein (CETP)

Cholesteryl ester transfer protein (CETP) has been strongly associated with AMD. CETP plays a key function in the transport of cholesterol from peripheral tissue to the liver by transferring cholesterol esters from low-density lipoproteins (LDLs) and very low-density lipoproteins (VLDLs) to HDLs [240]. In the human retina, CETP localizes to photoreceptor outer segments and the choriocapillaris and is involved in local lipid trafficking [241]. The major CETP variant, rs3764261, has been strongly associated with an increased risk of AMD [46,47,49,50,234]. Interestingly, Wang et al. noted this association only after adjusting for pathologic CFH variants, suggesting the presence of major interplay between the complement and lipid metabolism pathways [234]. Two additional intronic variants of CETP have been associated with AMD: rs5817082 is associated with a slightly reduced risk of AMD, while rs17231506 is associated with a slightly increased risk [51]. Curiously, CETP rs17231506 is associated with elevated HDL levels [60,242] similarly to LIPC rs493258, yet the two gene variants appear to have opposite effects on AMD pathogenesis.

4.4. ATP-Binding Cassette Transporter A1 (ABCA1)

The ATP-binding cassette transporter A1 (ABCA1) protein plays an important role in the elimination of excess tissue cholesterol by initiating the formation of HDL [243] and mediating the efflux of cellular cholesterol into the extracellular space, where HDL may bind and transport it back to the liver [181]. The complete absence of ABCA1 in a knockout mouse model results in retinal lipid accumulation and RPE degeneration [244], whereas increased ABCA1 expression leads to reduced lipid accumulation in in-vitro RPE cell models [245]. Population-based studies of the rs1883025 ABCA1 variant have demonstrated opposite effects of its two alleles; the C allele is generally associated with increased plasma HDL levels and increased risk of AMD, while the T allele is associated with decreased HDL and decreased risk of AMD [49,246]. The protective effect of the rs1883025 T allele has also been demonstrated in a 2016 meta-analysis [247]. Similarly, the A allele of the rs2740488 variant has been associated with an increased risk of AMD, whereas the C allele is associated with a slightly decreased risk [51,245].

4.5. Lipoprotein Lipase (LPL)

Finally, lipoprotein lipase (LPL) is an enzyme ubiquitously expressed throughout the human body with key roles in the metabolism of lipoprotein triglycerides into free fatty acids in the blood stream [248]. A single intergenic LPL variant, rs12678919, has been analyzed in the context of AMD, with inconsistent findings. Several studies have demonstrated no statistically significant association between rs12678919 and AMD [47,176,249,250,251]. However, two studies have demonstrated an association with increased risk [46,252]. As with CETP, a meta-analysis by Wang et al. discovered a strong association between LPL rs12678919 and AMD risk only after adjusting for CFH gene mutations [234], which may account for the variable findings of prior studies.

5. Angiogenesis

Angiogenic pathways are involved in the development of MNV in the advanced stages of neovascular AMD. MNV involves the proliferation of new abnormal blood vessels growing from the choroid into the sub-RPE or subretinal space, or both [253]. Various hypotheses have been proposed to account for this phenomenon. It is predominantly thought that subretinal angiogenesis is controlled by the RPE [254], which secretes VEGF in states of ischemia to stimulate endothelial cell proliferation [255]. VEGF is therefore recognized as the major driver of neovascularization, and currently represents the main target for treatment of neovascular disease [256]. The RPE also secretes pigment epithelial-derived factor (PEDF), a potent inhibitor of angiogenesis [257], allowing the modulation of neovascularization via PEDF activity. It has also been suggested that certain ECM components can stimulate angiogenesis in states of dysregulation [258].

5.1. Vascular Endothelial Growth Factor (VEGF)

VEGF consists of seven biologically active isoforms (Figure 5), with VEGF-A playing the most significant role in angiogenesis [259]. VEGF-A is a glycoprotein that mainly targets endothelial cells to stimulate proliferation, migration, and vessel formation while inhibiting cellular apoptosis [260]. Elevated levels of VEGF-A have consequently been associated with a range of ocular neovascular diseases [261]. Multiple VEGF-A variants have been identified, most of which surprisingly have not been strongly or consistently associated with AMD [38,262]. However, the VEGF-A rs3025033 variant and haplotype rs1570360A-rs699947A-rs3025033G-rs2146323A have recently been associated with a lower risk of neovascular AMD [263]. Conversely, the VEGF-A rs3025039 C-allele may increase the risk of AMD, although apparently only in the context of the rs1048661 lysine oxidase 1 (LOXL1) gene variant [264]. The T-allele of the VEGF-A rs3025000 variant increases the likelihood of clinical response to anti-VEGF therapy in MNV [265], an effect that has not been demonstrated with other VEGFA alleles [266] and may be relatively minor compared to phenotypic predictors of response to VEGF therapy [267].

5.2. Fibulin 5

Fibulin 5 (FBLN5) is an ECM protein that localizes to sub-RPE deposits and drusen [268,269] and plays a role in modulating angiogenesis in part by antagonizing VEGF [270]. AMD patients have been found to secrete FBLN5 at reduced rates [271]. Missense FBLN5 and other fibulin gene mutations can be found in up to 2% of AMD patients [272,273], suggesting that fibulin dysfunction may play a role in AMD pathogenesis [272].

6. Oxidative Stress Response and Photoreceptor Survival

The shared endpoint of the pathophysiologic processes underlying macular degeneration is RPE degeneration, with consequent photoreceptor cell death and dysfunction [274]. This process may be expedited by direct oxidative damage to cellular DNA in the retina directly causing cellular apoptosis [275]. Genetic mutations that increase susceptibility to oxidative damage and impair photoreceptor survival may therefore play a role in AMD pathogenesis [10]. For example, the RAD51 family of genes plays a critical role in DNA repair and protection against oxidative damage [276]. Several rare mutations in RAD51B have been associated with a significantly increased risk of AMD [223,277,278], and abnormally decreased serum concentrations of RAD51B have been noted in AMD patients [279].
Similarly, the tumor necrosis factor receptor superfamily, member 10a (TNFRSF10A), also known as Death Receptor 4 (DR4), plays key roles in promoting cellular apoptosis [280] and can be localized to the RPE [281]. Reduced expression of TNFRSF10A is associated with decreased RPE cell viability and increased apoptotic susceptibility in mice [281]. TNFRSF10A mutations have been associated with an increased risk of AMD, particularly in Asian populations [50,51,151,282,283].
Finally, excision repair cross complexes (ERCC) have been associated with AMD susceptibility [284]. ERCC6 functions in the transcription-coupled excision repair of DNA mutations [285]. Patients with AMD have been found to have decreased retinal ERCC6 expression [286]. Mutations in the ERCC6 gene have been inconsistently associated with increased AMD risk [179,284,286,287,288] and may have a synergic effect with CFH mutations [284].

7. Genes Implicated in Multiple Pathways

Age-Related Maculopathy Susceptibility 2 (ARMS2) and High-Temperature Requirement Factor A Serine Peptidase 1 (HTRA1)

One of the earliest discovered and most significant of the susceptibility loci associated with AMD is the region of chromosome 10q26 spanning the age-related maculopathy susceptibility 2 (ARMS2) gene coding region and the high-temperature requirement factor A serine peptidase 1 (HTRA1) gene promoter [44,289,290,291,292,293,294]. The ARMS2-HTRA1 region is a major susceptibility locus among Caucasians and East Asians [72,73], together with CFH accounting for over half of the genetic risk associated with AMD [51,153,229,295,296,297]. Notably, patients with mutations in the ARMS2-HTRA1 locus are more than twice as likely to progress to late-stage disease compared to patients harboring CFH variants, and ARMS2-HTRA1 variants have also been associated with more rapid progression of disease [17].
Multiple high-risk gene variants have been discovered in the ARMS2-HTRA1 locus [95,223,294,296,298,299,300,301,302,303,304,305]. These variants are in high linkage disequilibrium (LD), and most studies to date have been unable to statistically distinguish between them. It is therefore unclear to date which of ARMS2 or HTRA1 is responsible for increased AMD risk [306]. Moreover, the exact structure and function of ARMS2 and HTRA1 remain unclear.
Local ARMS2 dysfunction may play a role in oxidative stress and damage to the retina [306,307]. Early studies suggested that ARMS2 localizes to the outer mitochondrial membrane in photoreceptors and RPE cells [302,307,308], although this finding has been disputed [309,310]. It has also been suggested that ARMS2 localizes to the perinuclear cytoplasm [309] or endoplasmic reticulum [310], suggesting a non-mitochondrial mechanism of increasing AMD risk. Finally, ARMS2 has recently been implicated in the local inflammatory response secreted by macrophages to activate the complement cascade and clear cellular debris [311]. HTRA1, on the other hand, has been implicated in a broad range of physiologic processes, including ECM remodeling and TGF-β cytokine signaling [306]. Its most prominent roles appear to be in ECM deposition, angiogenesis, and regulation of local subretinal inflammation [294,312,313]. It has also been hypothesized that HTRA1 plays a key role in maintaining the RPE–BrM–choroid interface in the aging retina [314]. High-risk variants that reduce HTRA1 expression at the RPE–BrM interface may therefore give rise to AMD phenotypes due to loss of its protective function against the effects of advancing age [314]. Some have recently argued that the weight of the evidence to date supports HTRA1, rather than ARMS2, as the main causal genetic factor of the two associated with AMD [306].
The HTRA1 gene variant rs11200638 is found in the HTRA1 promoter region in strong LD with the ARMS2 missense variant rs10490924 [294,300,309]. Both have been highly associated with increased AMD risk [95,223,305,308,309,315,316,317] and a younger onset of the disease [304]. ARMS2 rs10490924 is correlated with elevated C-reactive protein levels [318], suggesting an inflammatory mechanism at play. Interestingly, while the ARMS2 rs10490924 G allele is associated with an increased risk of progression to advanced stages of disease [296], the G allele is also associated with a higher likelihood of MNV response to anti-VEGF therapy in patients who have progressed to neovascular disease [98,317,319,320,321,322], particularly in East Asian populations [321]. A similar correlation has been demonstrated between HTRA1 rs11200638 and increased likelihood of response to anti-VEGF therapy [317]. The ARMS2 rs10490924 variant also appears to play a more prominent role in AMD development in Asian populations than in European populations, with risk allele frequencies of 40% vs. 20%, respectively [54]. As ARMS2 dysfunction is slightly associated with the progression to neovascular disease, these differences in allele distribution may in part explain the higher prevalence of MNV in Asian populations compared to Europeans [323]. [
An additional unstable ARMS2 insertion-deletion variant c.*372_815del443ins54 in complete LD with ARMS2 rs10490924 and HTRA1 rs11200638 has also been strongly associated with increased AMD risk and progression to MNV [95,308,324]. However, its effects on HTRA1 and ARMS2 expression and function are unclear, and the risk it confers is difficult to distinguish from that associated with rs10490924 and rs11200638 [306]. Another ARMS-HTRA1 variant, rs2284665, has also been found to correlate significantly with the development of MNV [325]. Finally, the nonsense ARMS2 variant rs10490924 [308] appears to have no effect on AMD risk despite decreasing ARMS2 expression [326].
In addition to HTRA1 rs11200638, other variants, such as rs1049331, rs2293870, and rs2284665, have also been identified as strongly associated with AMD progression [327,328,329]. The mechanism by which theses HTRA1 mutations confer risk is unclear. Impaired HTRA1-mediated inhibition of cellular apoptosis [330] and insulin-like growth factor 1 (IGF1) [331] may play a role. Additionally, HTRA1 rs2284665 may increase HTRA1 expression in lymphocytes and elsewhere [294], although this has not been consistently demonstrated [294,332,333,334,335].

8. Conclusions

AMD is a complex and multifactorial disease that we have yet to fully understand. Several pathogenic pathways involving the complement system, extracellular matrix remodeling, lipid metabolism, angiogenesis, and oxidative stress response are currently thought to intertwine to give rise to AMD in its various forms. Age plays a significant role in the risk of disease development, and environmental factors such as smoking and diet represent important modifiable risk factors to prevent AMD. Despite its late onset, AMD has also been shown to have a strong genetic component, and a broad array of gene variants affecting these key pathways have been identified as significantly affecting the risk of disease. With the advent of GWAS, at least 52 independent gene variants and 34 genetic loci have been identified to date, accounting for over 50% of the genetic risk [51]. There are now commercially available tests (e.g., Macula Risk and Vita Risk® from Arctic Medical Laboratories, Grand Rapids, Michigan, USA) that assess for the presence of gene variants affecting these pathways. Such testing may allow an estimation of personalized risk and expected response to treatment using algorithms that integrate genetic information with other disease-associated factors such as age, smoking status, fellow-eye status, and body mass index [336]. While the utility of genetic testing remains limited at this time due to the paucity of approved gene-targeted treatment modalities [337], gene therapies currently in development [57,338] hold great promise for the future implementation of personalized interventions that significantly alter the course of AMD. Future research on the molecular and pharmacogenetics of AMD may thus hold the key to curing what is currently among the most common incurable causes of blindness worldwide.

Author Contributions

All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

Shughoury and Sevgi declare no conflict of interest. Ciulla has an employment relation with, and equity ownership in, Clearside Bio. This work was undertaken in his role as Volunteer Clinical Professor at Indiana University School of Medicine.

References

  1. Wong, W.L.; Su, X.; Li, X.; Cheung, C.M.; Klein, R.; Cheng, C.Y.; Wong, T.Y. Global prevalence of age-related macular degeneration and disease burden projection for 2020 and 2040: A systematic review and meta-analysis. Lancet Glob. Health 2014, 2, e106–e116. [Google Scholar] [CrossRef] [Green Version]
  2. Brown, M.M.; Brown, G.C.; Sharma, S.; Kistler, J.; Brown, H. Utility values associated with blindness in an adult population. Br. J. Ophthalmol. 2001, 85, 327–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hernandez Trillo, A.; Dickinson, C.M. The impact of visual and nonvisual factors on quality of life and adaptation in adults with visual impairment. Investig. Ophthalmol. Vis. Sci. 2012, 53, 4234–4241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Williams, R.A.; Brody, B.L.; Thomas, R.G.; Kaplan, R.M.; Brown, S.I. The psychosocial impact of macular degeneration. Arch. Ophthalmol. 1998, 116, 514–520. [Google Scholar] [CrossRef]
  5. Spooner, K.L.; Mhlanga, C.T.; Hong, T.H.; Broadhead, G.K.; Chang, A.A. The burden of neovascular age-related macular degeneration: A patient’s perspective. Clin. Ophthalmol. 2018, 12, 2483–2491. [Google Scholar] [CrossRef] [Green Version]
  6. Brown, M.M.; Brown, G.C.; Stein, J.D.; Roth, Z.; Campanella, J.; Beauchamp, G.R. Age-related macular degeneration: Economic burden and value-based medicine analysis. Can. J. Ophthalmol. 2005, 40, 277–287. [Google Scholar] [CrossRef]
  7. Rein, D.B.; Wittenborn, J.S.; Zhang, X.; Honeycutt, A.A.; Lesesne, S.B.; Saaddine, J.; Vision Health Cost-Effectiveness Study, G. Forecasting age-related macular degeneration through the year 2050: The potential impact of new treatments. Arch. Ophthalmol. 2009, 127, 533–540. [Google Scholar] [CrossRef] [Green Version]
  8. Sparrow, J.R. Bisretinoids of RPE lipofuscin: Trigger for complement activation in age-related macular degeneration. Adv. Exp. Med. Biol. 2010, 703, 63–74. [Google Scholar] [CrossRef]
  9. Strauss, O. The retinal pigment epithelium in visual function. Physiol. Rev. 2005, 85, 845–881. [Google Scholar] [CrossRef] [Green Version]
  10. Beatty, S.; Koh, H.; Phil, M.; Henson, D.; Boulton, M. The role of oxidative stress in the pathogenesis of age-related macular degeneration. Surv. Ophthalmol. 2000, 45, 115–134. [Google Scholar] [CrossRef] [Green Version]
  11. Van der Vaart, H.; Postma, D.S.; Timens, W.; ten Hacken, N.H. Acute effects of cigarette smoke on inflammation and oxidative stress: A review. Thorax 2004, 59, 713–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Thornton, J.; Edwards, R.; Mitchell, P.; Harrison, R.A.; Buchan, I.; Kelly, S.P. Smoking and age-related macular degeneration: A review of association. Eye 2005, 19, 935–944. [Google Scholar] [CrossRef] [PubMed]
  13. Smith, W.; Assink, J.; Klein, R.; Mitchell, P.; Klaver, C.C.; Klein, B.E.; Hofman, A.; Jensen, S.; Wang, J.J.; de Jong, P.T. Risk factors for age-related macular degeneration: Pooled findings from three continents. Ophthalmology 2001, 108, 697–704. [Google Scholar] [CrossRef]
  14. Age-Related Eye Disease Study Research Group. Risk factors associated with age-related macular degeneration. A case-control study in the age-related eye disease study: Age-Related Eye Disease Study Report Number 3. Ophthalmology 2000, 107, 2224–2232. [Google Scholar] [CrossRef]
  15. Friedman, E. A hemodynamic model of the pathogenesis of age-related macular degeneration. Am. J. Ophthalmol. 1997, 124, 677–682. [Google Scholar] [CrossRef]
  16. Curcio, C.A. Complementing apolipoprotein secretion by cultured retinal pigment epithelium. Proc. Natl. Acad. Sci. USA 2011, 108, 18569–18570. [Google Scholar] [CrossRef] [Green Version]
  17. Fleckenstein, M.; Keenan, T.D.L.; Guymer, R.H.; Chakravarthy, U.; Schmitz-Valckenberg, S.; Klaver, C.C.; Wong, W.T.; Chew, E.Y. Age-related macular degeneration. Nat. Rev. Dis. Primers 2021, 7, 31. [Google Scholar] [CrossRef]
  18. Spaide, R.F.; Jaffe, G.J.; Sarraf, D.; Freund, K.B.; Sadda, S.R.; Staurenghi, G.; Waheed, N.K.; Chakravarthy, U.; Rosenfeld, P.J.; Holz, F.G.; et al. Consensus Nomenclature for Reporting Neovascular Age-Related Macular Degeneration Data: Consensus on Neovascular Age-Related Macular Degeneration Nomenclature Study Group. Ophthalmology 2020, 127, 616–636. [Google Scholar] [CrossRef]
  19. Palejwala, N.V.; Jia, Y.; Gao, S.S.; Liu, L.; Flaxel, C.J.; Hwang, T.S.; Lauer, A.K.; Wilson, D.J.; Huang, D.; Bailey, S.T. Detection of Nonexudative Choroidal Neovascularization in Age-Related Macular Degeneration with Optical Coherence Tomography Angiography. Retina 2015, 35, 2204–2211. [Google Scholar] [CrossRef] [Green Version]
  20. Choudhury, F.; Varma, R.; McKean-Cowdin, R.; Klein, R.; Azen, S.P.; Los Angeles Latino Eye Study, G. Risk factors for four-year incidence and progression of age-related macular degeneration: The los angeles latino eye study. Am. J. Ophthalmol. 2011, 152, 385–395. [Google Scholar] [CrossRef] [Green Version]
  21. Seddon, J.M.; Willett, W.C.; Speizer, F.E.; Hankinson, S.E. A prospective study of cigarette smoking and age-related macular degeneration in women. JAMA 1996, 276, 1141–1146. [Google Scholar] [CrossRef] [PubMed]
  22. Christen, W.G.; Glynn, R.J.; Manson, J.E.; Ajani, U.A.; Buring, J.E. A prospective study of cigarette smoking and risk of age-related macular degeneration in men. JAMA 1996, 276, 1147–1151. [Google Scholar] [CrossRef]
  23. Chakravarthy, U.; Augood, C.; Bentham, G.C.; de Jong, P.T.; Rahu, M.; Seland, J.; Soubrane, G.; Tomazzoli, L.; Topouzis, F.; Vingerling, J.R.; et al. Cigarette smoking and age-related macular degeneration in the EUREYE Study. Ophthalmology 2007, 114, 1157–1163. [Google Scholar] [CrossRef] [PubMed]
  24. Seddon, J.M.; Rosner, B.; Sperduto, R.D.; Yannuzzi, L.; Haller, J.A.; Blair, N.P.; Willett, W. Dietary fat and risk for advanced age-related macular degeneration. Arch. Ophthalmol. 2001, 119, 1191–1199. [Google Scholar] [CrossRef] [Green Version]
  25. Van Leeuwen, R.; Boekhoorn, S.; Vingerling, J.R.; Witteman, J.C.; Klaver, C.C.; Hofman, A.; de Jong, P.T. Dietary intake of antioxidants and risk of age-related macular degeneration. JAMA 2005, 294, 3101–3107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Hyman, L.G.; Lilienfeld, A.M.; Ferris, F.L., 3rd; Fine, S.L. Senile macular degeneration: A case-control study. Am. J. Epidemiol. 1983, 118, 213–227. [Google Scholar] [CrossRef]
  27. Klein, R.; Klein, B.E.; Cruickshanks, K.J. The prevalence of age-related maculopathy by geographic region and ethnicity. Prog. Retin. Eye Res. 1999, 18, 371–389. [Google Scholar] [CrossRef]
  28. Klein, R.; Klein, B.E.; Knudtson, M.D.; Wong, T.Y.; Cotch, M.F.; Liu, K.; Burke, G.; Saad, M.F.; Jacobs, D.R., Jr. Prevalence of age-related macular degeneration in 4 racial/ethnic groups in the multi-ethnic study of atherosclerosis. Ophthalmology 2006, 113, 373–380. [Google Scholar] [CrossRef]
  29. Klein, M.L.; Mauldin, W.M.; Stoumbos, V.D. Heredity and age-related macular degeneration. Observations in monozygotic twins. Arch. Ophthalmol. 1994, 112, 932–937. [Google Scholar] [CrossRef]
  30. Meyers, S.M.; Greene, T.; Gutman, F.A. A twin study of age-related macular degeneration. Am. J. Ophthalmol. 1995, 120, 757–766. [Google Scholar] [CrossRef] [Green Version]
  31. Seddon, J.M.; Ajani, U.A.; Mitchell, B.D. Familial aggregation of age-related maculopathy. Am. J. Ophthalmol. 1997, 123, 199–206. [Google Scholar] [CrossRef]
  32. Klein, B.E.; Klein, R.; Lee, K.E.; Moore, E.L.; Danforth, L. Risk of incident age-related eye diseases in people with an affected sibling: The Beaver Dam Eye Study. Am. J. Epidemiol. 2001, 154, 207–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Hammond, C.J.; Webster, A.R.; Snieder, H.; Bird, A.C.; Gilbert, C.E.; Spector, T.D. Genetic influence on early age-related maculopathy: A twin study. Ophthalmology 2002, 109, 730–736. [Google Scholar] [CrossRef]
  34. Bojanowski, C.M.; Tuo, J.; Chew, E.Y.; Csaky, K.G.; Chan, C.C. Analysis of Hemicentin-1, hOgg1, and E-selectin single nucleotide polymorphisms in age-related macular degeneration. Trans. Am. Ophthalmol. Soc. 2005, 103, 37–44; discussion 44–35. [Google Scholar]
  35. Maguire, M.G.; Ying, G.S.; Jaffe, G.J.; Toth, C.A.; Daniel, E.; Grunwald, J.; Martin, D.F.; Hagstrom, S.A.; Group, C.R. Single-Nucleotide Polymorphisms Associated With Age-Related Macular Degeneration and Lesion Phenotypes in the Comparison of Age-Related Macular Degeneration Treatments Trials. JAMA Ophthalmol. 2016, 134, 674–681. [Google Scholar] [CrossRef] [Green Version]
  36. Popp, N.A.; Agron, E.; Hageman, G.S.; Tuo, J.; Chew, E.Y.; Chan, C.C. No Sex Differences in the Frequencies of Common Single Nucleotide Polymorphisms Associated with Age-Related Macular Degeneration. Curr. Eye Res. 2017, 42, 470–475. [Google Scholar] [CrossRef]
  37. Lin, L.Y.; Zhou, Q.; Hagstrom, S.; Maguire, M.G.; Daniel, E.; Grunwald, J.E.; Martin, D.F.; Ying, G.S.; Group, C.R. Association of Single-Nucleotide Polymorphisms in Age-Related Macular Degeneration With Pseudodrusen: Secondary Analysis of Data From the Comparison of AMD Treatments Trials. JAMA Ophthalmol. 2018, 136, 682–688. [Google Scholar] [CrossRef]
  38. Richardson, A.J.; Islam, F.M.; Guymer, R.H.; Cain, M.; Baird, P.N. A tag-single nucleotide polymorphisms approach to the vascular endothelial growth factor-A gene in age-related macular degeneration. Mol. Vis. 2007, 13, 2148–2152. [Google Scholar]
  39. Zhou, T.Q.; Guan, H.J.; Hu, J.Y. Genome-wide analysis of single nucleotide polymorphisms in patients with atrophic age-related macular degeneration in oldest old Han Chinese. Genet. Mol. Res. 2015, 14, 17432–17438. [Google Scholar] [CrossRef]
  40. Shin, H.T.; Yoon, B.W.; Seo, J.H. Comparison of risk allele frequencies of single nucleotide polymorphisms associated with age-related macular degeneration in different ethnic groups. BMC Ophthalmol. 2021, 21, 97. [Google Scholar] [CrossRef]
  41. Klaver, C.C.; Kliffen, M.; van Duijn, C.M.; Hofman, A.; Cruts, M.; Grobbee, D.E.; van Broeckhoven, C.; de Jong, P.T. Genetic association of apolipoprotein E with age-related macular degeneration. Am. J. Hum. Genet. 1998, 63, 200–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Raychaudhuri, S.; Iartchouk, O.; Chin, K.; Tan, P.L.; Tai, A.K.; Ripke, S.; Gowrisankar, S.; Vemuri, S.; Montgomery, K.; Yu, Y.; et al. A rare penetrant mutation in CFH confers high risk of age-related macular degeneration. Nat. Genet. 2011, 43, 1232–1236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Helgason, H.; Sulem, P.; Duvvari, M.R.; Luo, H.; Thorleifsson, G.; Stefansson, H.; Jonsdottir, I.; Masson, G.; Gudbjartsson, D.F.; Walters, G.B.; et al. A rare nonsynonymous sequence variant in C3 is associated with high risk of age-related macular degeneration. Nat. Genet. 2013, 45, 1371–1374. [Google Scholar] [CrossRef] [PubMed]
  44. Seddon, J.M.; Yu, Y.; Miller, E.C.; Reynolds, R.; Tan, P.L.; Gowrisankar, S.; Goldstein, J.I.; Triebwasser, M.; Anderson, H.E.; Zerbib, J.; et al. Rare variants in CFI, C3 and C9 are associated with high risk of advanced age-related macular degeneration. Nat. Genet. 2013, 45, 1366–1370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhan, X.; Larson, D.E.; Wang, C.; Koboldt, D.C.; Sergeev, Y.V.; Fulton, R.S.; Fulton, L.L.; Fronick, C.C.; Branham, K.E.; Bragg-Gresham, J.; et al. Identification of a rare coding variant in complement 3 associated with age-related macular degeneration. Nat. Genet. 2013, 45, 1375–1379. [Google Scholar] [CrossRef]
  46. Chen, W.; Stambolian, D.; Edwards, A.O.; Branham, K.E.; Othman, M.; Jakobsdottir, J.; Tosakulwong, N.; Pericak-Vance, M.A.; Campochiaro, P.A.; Klein, M.L.; et al. Genetic variants near TIMP3 and high-density lipoprotein-associated loci influence susceptibility to age-related macular degeneration. Proc. Natl. Acad. Sci. USA 2010, 107, 7401–7406. [Google Scholar] [CrossRef] [Green Version]
  47. Neale, B.M.; Fagerness, J.; Reynolds, R.; Sobrin, L.; Parker, M.; Raychaudhuri, S.; Tan, P.L.; Oh, E.C.; Merriam, J.E.; Souied, E.; et al. Genome-wide association study of advanced age-related macular degeneration identifies a role of the hepatic lipase gene (LIPC). Proc. Natl. Acad. Sci. USA 2010, 107, 7395–7400. [Google Scholar] [CrossRef] [Green Version]
  48. Yu, Y.; Bhangale, T.R.; Fagerness, J.; Ripke, S.; Thorleifsson, G.; Tan, P.L.; Souied, E.H.; Richardson, A.J.; Merriam, J.E.; Buitendijk, G.H.; et al. Common variants near FRK/COL10A1 and VEGFA are associated with advanced age-related macular degeneration. Hum. Mol. Genet. 2011, 20, 3699–3709. [Google Scholar] [CrossRef] [Green Version]
  49. Yu, Y.; Reynolds, R.; Fagerness, J.; Rosner, B.; Daly, M.J.; Seddon, J.M. Association of variants in the LIPC and ABCA1 genes with intermediate and large drusen and advanced age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 2011, 52, 4663–4670. [Google Scholar] [CrossRef] [Green Version]
  50. Fritsche, L.G.; Chen, W.; Schu, M.; Yaspan, B.L.; Yu, Y.; Thorleifsson, G.; Zack, D.J.; Arakawa, S.; Cipriani, V.; Ripke, S.; et al. Seven new loci associated with age-related macular degeneration. Nat. Genet. 2013, 45, 433–439. [Google Scholar] [CrossRef]
  51. Fritsche, L.G.; Igl, W.; Bailey, J.N.; Grassmann, F.; Sengupta, S.; Bragg-Gresham, J.L.; Burdon, K.P.; Hebbring, S.J.; Wen, C.; Gorski, M.; et al. A large genome-wide association study of age-related macular degeneration highlights contributions of rare and common variants. Nat. Genet. 2016, 48, 134–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Han, X.; Gharahkhani, P.; Mitchell, P.; Liew, G.; Hewitt, A.W.; MacGregor, S. Genome-wide meta-analysis identifies novel loci associated with age-related macular degeneration. J. Hum. Genet. 2020, 65, 657–665. [Google Scholar] [CrossRef] [PubMed]
  53. Black, J.R.; Clark, S.J. Age-related macular degeneration: Genome-wide association studies to translation. Genet. Med. 2016, 18, 283–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Fritsche, L.G.; Fariss, R.N.; Stambolian, D.; Abecasis, G.R.; Curcio, C.A.; Swaroop, A. Age-related macular degeneration: Genetics and biology coming together. Annu. Rev. Genom. Hum. Genet. 2014, 15, 151–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Yamashiro, K.; Hosoda, Y.; Miyake, M.; Ooto, S.; Tsujikawa, A. Characteristics of Pachychoroid Diseases and Age-Related Macular Degeneration: Multimodal Imaging and Genetic Backgrounds. J. Clin. Med. 2020, 9, 2034. [Google Scholar] [CrossRef]
  56. Al-Khersan, H.; Hussain, R.M.; Ciulla, T.A.; Dugel, P.U. Innovative therapies for neovascular age-related macular degeneration. Expert Opin. Pharmacother. 2019, 20, 1879–1891. [Google Scholar] [CrossRef]
  57. Guimaraes, T.A.C.; Georgiou, M.; Bainbridge, J.W.B.; Michaelides, M. Gene therapy for neovascular age-related macular degeneration: Rationale, clinical trials and future directions. Br. J. Ophthalmol. 2021, 105, 151–157. [Google Scholar] [CrossRef] [Green Version]
  58. Anderson, D.H.; Radeke, M.J.; Gallo, N.B.; Chapin, E.A.; Johnson, P.T.; Curletti, C.R.; Hancox, L.S.; Hu, J.; Ebright, J.N.; Malek, G.; et al. The pivotal role of the complement system in aging and age-related macular degeneration: Hypothesis re-visited. Prog. Retin. Eye Res. 2010, 29, 95–112. [Google Scholar] [CrossRef] [Green Version]
  59. Hageman, G.S.; Luthert, P.J.; Victor Chong, N.H.; Johnson, L.V.; Anderson, D.H.; Mullins, R.F. An integrated hypothesis that considers drusen as biomarkers of immune-mediated processes at the RPE-Bruch’s membrane interface in aging and age-related macular degeneration. Prog. Retin. Eye Res. 2001, 20, 705–732. [Google Scholar] [CrossRef]
  60. Colijn, J.M.; Meester-Smoor, M.; Verzijden, T.; de Breuk, A.; Silva, R.; Merle, B.M.J.; Cougnard-Gregoire, A.; Hoyng, C.B.; Fauser, S.; Coolen, A.; et al. Genetic Risk, Lifestyle, and Age-Related Macular Degeneration in Europe: The EYE-RISK Consortium. Ophthalmology 2021, 128, 1039–1049. [Google Scholar] [CrossRef]
  61. Janeway, C.A., Jr. How the immune system protects the host from infection. Microbes Infect. 2001, 3, 1167–1171. [Google Scholar] [CrossRef]
  62. Kilgore, K.S.; Imlay, M.M.; Szaflarski, J.P.; Silverstein, F.S.; Malani, A.N.; Evans, V.M.; Warren, J.S. Neutrophils and reactive oxygen intermediates mediate glucan-induced pulmonary granuloma formation through the local induction of monocyte chemoattractant protein-1. Lab. Investig. 1997, 76, 191–201. [Google Scholar] [PubMed]
  63. Triantafilou, K.; Hughes, T.R.; Triantafilou, M.; Morgan, B.P. The complement membrane attack complex triggers intracellular Ca2+ fluxes leading to NLRP3 inflammasome activation. J. Cell Sci. 2013, 126, 2903–2913. [Google Scholar] [CrossRef] [Green Version]
  64. Vlaicu, S.I.; Tegla, C.A.; Cudrici, C.D.; Danoff, J.; Madani, H.; Sugarman, A.; Niculescu, F.; Mircea, P.A.; Rus, V.; Rus, H. Role of C5b-9 complement complex and response gene to complement-32 (RGC-32) in cancer. Immunol. Res. 2013, 56, 109–121. [Google Scholar] [CrossRef] [PubMed]
  65. Danobeitia, J.S.; Djamali, A.; Fernandez, L.A. The role of complement in the pathogenesis of renal ischemia-reperfusion injury and fibrosis. Fibrogenesis Tissue Repair 2014, 7, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Edwards, A.O.; Ritter, R., III; Abel, K.J.; Manning, A.; Panhuysen, C.; Farrer, L.A. Complement factor H polymorphism and age-related macular degeneration. Science 2005, 308, 421–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Hageman, G.S.; Anderson, D.H.; Johnson, L.V.; Hancox, L.S.; Taiber, A.J.; Hardisty, L.I.; Hageman, J.L.; Stockman, H.A.; Borchardt, J.D.; Gehrs, K.M.; et al. A common haplotype in the complement regulatory gene factor H (HF1/CFH) predisposes individuals to age-related macular degeneration. Proc. Natl. Acad. Sci. USA 2005, 102, 7227–7232. [Google Scholar] [CrossRef] [Green Version]
  68. Haines, J.L.; Hauser, M.A.; Schmidt, S.; Scott, W.K.; Olson, L.M.; Gallins, P.; Spencer, K.L.; Kwan, S.Y.; Noureddine, M.; Gilbert, J.R.; et al. Complement factor H variant increases the risk of age-related macular degeneration. Science 2005, 308, 419–421. [Google Scholar] [CrossRef] [Green Version]
  69. Klein, R.J.; Zeiss, C.; Chew, E.Y.; Tsai, J.Y.; Sackler, R.S.; Haynes, C.; Henning, A.K.; SanGiovanni, J.P.; Mane, S.M.; Mayne, S.T.; et al. Complement factor H polymorphism in age-related macular degeneration. Science 2005, 308, 385–389. [Google Scholar] [CrossRef]
  70. Zareparsi, S.; Branham, K.E.; Li, M.; Shah, S.; Klein, R.J.; Ott, J.; Hoh, J.; Abecasis, G.R.; Swaroop, A. Strong association of the Y402H variant in complement factor H at 1q32 with susceptibility to age-related macular degeneration. Am. J. Hum. Genet. 2005, 77, 149–153. [Google Scholar] [CrossRef] [Green Version]
  71. Conley, Y.P.; Thalamuthu, A.; Jakobsdottir, J.; Weeks, D.E.; Mah, T.; Ferrell, R.E.; Gorin, M.B. Candidate gene analysis suggests a role for fatty acid biosynthesis and regulation of the complement system in the etiology of age-related maculopathy. Hum. Mol. Genet. 2005, 14, 1991–2002. [Google Scholar] [CrossRef] [PubMed]
  72. Cheng, C.Y.; Yamashiro, K.; Chen, L.J.; Ahn, J.; Huang, L.; Huang, L.; Cheung, C.M.; Miyake, M.; Cackett, P.D.; Yeo, I.Y.; et al. New loci and coding variants confer risk for age-related macular degeneration in East Asians. Nat. Commun. 2015, 6, 6063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Miyake, M.; Saito, M.; Yamashiro, K.; Sekiryu, T.; Yoshimura, N. Complement factor H R1210C among Japanese patients with age-related macular degeneration. Jpn J. Ophthalmol. 2015, 59, 273–278. [Google Scholar] [CrossRef] [PubMed]
  74. Johnson, P.T.; Betts, K.E.; Radeke, M.J.; Hageman, G.S.; Anderson, D.H.; Johnson, L.V. Individuals homozygous for the age-related macular degeneration risk-conferring variant of complement factor H have elevated levels of CRP in the choroid. Proc. Natl. Acad. Sci. USA 2006, 103, 17456–17461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Laine, M.; Jarva, H.; Seitsonen, S.; Haapasalo, K.; Lehtinen, M.J.; Lindeman, N.; Anderson, D.H.; Johnson, P.T.; Jarvela, I.; Jokiranta, T.S.; et al. Y402H polymorphism of complement factor H affects binding affinity to C-reactive protein. J. Immunol. 2007, 178, 3831–3836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Triebwasser, M.P.; Roberson, E.D.; Yu, Y.; Schramm, E.C.; Wagner, E.K.; Raychaudhuri, S.; Seddon, J.M.; Atkinson, J.P. Rare Variants in the Functional Domains of Complement Factor H Are Associated With Age-Related Macular Degeneration. Investig. Ophthalmol. Vis. Sci. 2015, 56, 6873–6878. [Google Scholar] [CrossRef] [Green Version]
  77. Rodriguez de Cordoba, S.; Esparza-Gordillo, J.; Goicoechea de Jorge, E.; Lopez-Trascasa, M.; Sanchez-Corral, P. The human complement factor H: Functional roles, genetic variations and disease associations. Mol. Immunol. 2004, 41, 355–367. [Google Scholar] [CrossRef]
  78. DiScipio, R.G. Formation and structure of the C5b-7 complex of the lytic pathway of complement. J. Biol. Chem. 1992, 267, 17087–17094. [Google Scholar] [CrossRef]
  79. Sharma, A.K.; Pangburn, M.K. Identification of three physically and functionally distinct binding sites for C3b in human complement factor H by deletion mutagenesis. Proc. Natl. Acad. Sci. USA 1996, 93, 10996–11001. [Google Scholar] [CrossRef] [Green Version]
  80. Skerka, C.; Lauer, N.; Weinberger, A.A.; Keilhauer, C.N.; Suhnel, J.; Smith, R.; Schlotzer-Schrehardt, U.; Fritsche, L.; Heinen, S.; Hartmann, A.; et al. Defective complement control of factor H (Y402H) and FHL-1 in age-related macular degeneration. Mol. Immunol. 2007, 44, 3398–3406. [Google Scholar] [CrossRef]
  81. Makou, E.; Herbert, A.P.; Barlow, P.N. Functional anatomy of complement factor H. Biochemistry 2013, 52, 3949–3962. [Google Scholar] [CrossRef] [PubMed]
  82. Molins, B.; Fuentes-Prior, P.; Adan, A.; Anton, R.; Arostegui, J.I.; Yague, J.; Dick, A.D. Complement factor H binding of monomeric C-reactive protein downregulates proinflammatory activity and is impaired with at risk polymorphic CFH variants. Sci. Rep. 2016, 6, 22889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Calippe, B.; Augustin, S.; Beguier, F.; Charles-Messance, H.; Poupel, L.; Conart, J.B.; Hu, S.J.; Lavalette, S.; Fauvet, A.; Rayes, J.; et al. Complement Factor H Inhibits CD47-Mediated Resolution of Inflammation. Immunity 2017, 46, 261–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Mulfaul, K.; Mullin, N.K.; Giacalone, J.C.; Voigt, A.P.; DeVore, M.R.; Stone, E.M.; Tucker, B.A.; Mullins, R.F. Local factor H production by human choroidal endothelial cells mitigates complement deposition: Implications for macular degeneration. J. Pathol. 2022, 257, 29–38. [Google Scholar] [CrossRef] [PubMed]
  85. Clark, S.J.; Perveen, R.; Hakobyan, S.; Morgan, B.P.; Sim, R.B.; Bishop, P.N.; Day, A.J. Impaired binding of the age-related macular degeneration-associated complement factor H 402H allotype to Bruch’s membrane in human retina. J. Biol. Chem. 2010, 285, 30192–30202. [Google Scholar] [CrossRef] [Green Version]
  86. Borras, C.; Canonica, J.; Jorieux, S.; Abache, T.; El Sanharawi, M.; Klein, C.; Delaunay, K.; Jonet, L.; Salvodelli, M.; Naud, M.C.; et al. CFH exerts anti-oxidant effects on retinal pigment epithelial cells independently from protecting against membrane attack complex. Sci. Rep. 2019, 9, 13873. [Google Scholar] [CrossRef] [Green Version]
  87. Armento, A.; Honisch, S.; Panagiotakopoulou, V.; Sonntag, I.; Jacob, A.; Bolz, S.; Kilger, E.; Deleidi, M.; Clark, S.; Ueffing, M. Loss of Complement Factor H impairs antioxidant capacity and energy metabolism of human RPE cells. Sci. Rep. 2020, 10, 10320. [Google Scholar] [CrossRef]
  88. Weismann, D.; Hartvigsen, K.; Lauer, N.; Bennett, K.L.; Scholl, H.P.; Charbel Issa, P.; Cano, M.; Brandstatter, H.; Tsimikas, S.; Skerka, C.; et al. Complement factor H binds malondialdehyde epitopes and protects from oxidative stress. Nature 2011, 478, 76–81. [Google Scholar] [CrossRef] [Green Version]
  89. Landowski, M.; Kelly, U.; Klingeborn, M.; Groelle, M.; Ding, J.D.; Grigsby, D.; Bowes Rickman, C. Human complement factor H Y402H polymorphism causes an age-related macular degeneration phenotype and lipoprotein dysregulation in mice. Proc. Natl. Acad. Sci. USA 2019, 116, 3703–3711. [Google Scholar] [CrossRef] [Green Version]
  90. Geerlings, M.J.; Kremlitzka, M.; Bakker, B.; Nilsson, S.C.; Saksens, N.T.; Lechanteur, Y.T.; Pauper, M.; Corominas, J.; Fauser, S.; Hoyng, C.B.; et al. The Functional Effect of Rare Variants in Complement Genes on C3b Degradation in Patients With Age-Related Macular Degeneration. JAMA Ophthalmol. 2017, 135, 39–46. [Google Scholar] [CrossRef]
  91. Maller, J.; George, S.; Purcell, S.; Fagerness, J.; Altshuler, D.; Daly, M.J.; Seddon, J.M. Common variation in three genes, including a noncoding variant in CFH, strongly influences risk of age-related macular degeneration. Nat. Genet. 2006, 38, 1055–1059. [Google Scholar] [CrossRef] [PubMed]
  92. Maugeri, A.; Barchitta, M.; Mazzone, M.G.; Giuliano, F.; Agodi, A. Complement System and Age-Related Macular Degeneration: Implications of Gene-Environment Interaction for Preventive and Personalized Medicine. Biomed. Res. Int. 2018, 2018, 7532507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Rajendran, A.; Dhoble, P.; Sundaresan, P.; Saravanan, V.; Vashist, P.; Nitsch, D.; Smeeth, L.; Chakravarthy, U.; Ravindran, R.D.; Fletcher, A.E. Genetic risk factors for late age-related macular degeneration in India. Br. J. Ophthalmol. 2018, 102, 1213–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Connolly, E.; Rhatigan, M.; O’Halloran, A.M.; Muldrew, K.A.; Chakravarthy, U.; Cahill, M.; Kenny, R.A.; Doyle, S.L. Prevalence of age-related macular degeneration associated genetic risk factors and 4-year progression data in the Irish population. Br. J. Ophthalmol. 2018, 102, 1691–1695. [Google Scholar] [CrossRef]
  95. Supanji, S.; Romdhoniyyah, D.F.; Sasongko, M.B.; Agni, A.N.; Wardhana, F.S.; Widayanti, T.W.; Prayogo, M.E.; Perdamaian, A.B.I.; Dianratri, A.; Kawaichi, M.; et al. Associations of ARMS2 and CFH Gene Polymorphisms with Neovascular Age-Related Macular Degeneration. Clin. Ophthalmol. 2021, 15, 1101–1108. [Google Scholar] [CrossRef] [PubMed]
  96. Matuskova, V.; Zeman, T.; Ewerlingova, L.; Hlinomazova, Z.; Soucek, J.; Vlkova, E.; Goswami, N.; Balcar, V.J.; Sery, O. An association of neovascular age-related macular degeneration with polymorphisms of CFH, ARMS2, HTRA1 and C3 genes in Czech population. Acta Ophthalmol. 2020, 98, e691–e699. [Google Scholar] [CrossRef] [PubMed]
  97. Gili, P.; Lloreda Martin, L.; Martin-Rodrigo, J.C.; Kim-Yeon, N.; Modamio-Gardeta, L.; Fernandez-Garcia, J.L.; Rebolledo-Poves, A.B.; Gomez-Blazquez, E.; Pazos-Rodriguez, R.; Perez-Fernandez, E.; et al. Gene polymorphisms associated with an increased risk of exudative age-related macular degeneration in a Spanish population. Eur. J. Ophthalmol. 2022, 32, 651–657. [Google Scholar] [CrossRef]
  98. Gourgouli, K.; Gourgouli, I.; Tsaousis, G.; Spai, S.; Niskopoulou, M.; Efthimiopoulos, S.; Lamnissou, K. Investigation of genetic base in the treatment of age-related macular degeneration. Int. Ophthalmol. 2020, 40, 985–997. [Google Scholar] [CrossRef]
  99. Neto, J.M.; Viturino, M.G.; Ananina, G.; Bajano, F.F.; Costa, S.; Roque, A.B.; Borges, G.F.; Franchi, R.; Rim, P.H.; Medina, F.M.; et al. Association of genetic variants rs641153 (CFB), rs2230199 (C3), and rs1410996 (CFH) with age-related macular degeneration in a Brazilian population. Exp. Biol. Med. 2021, 246, 2290–2296. [Google Scholar] [CrossRef]
  100. Thakkinstian, A.; Bowe, S.; McEvoy, M.; Smith, W.; Attia, J. Association between apolipoprotein E polymorphisms and age-related macular degeneration: A HuGE review and meta-analysis. Am. J. Epidemiol. 2006, 164, 813–822. [Google Scholar] [CrossRef]
  101. Lores-Motta, L.; Paun, C.C.; Corominas, J.; Pauper, M.; Geerlings, M.J.; Altay, L.; Schick, T.; Daha, M.R.; Fauser, S.; Hoyng, C.B.; et al. Genome-Wide Association Study Reveals Variants in CFH and CFHR4 Associated with Systemic Complement Activation: Implications in Age-Related Macular Degeneration. Ophthalmology 2018, 125, 1064–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Wang, Z.; Zou, M.; Chen, A.; Liu, Z.; Young, C.A.; Wang, S.B.; Zheng, D.; Jin, G. Genetic associations of anti-vascular endothelial growth factor therapy response in age-related macular degeneration: A systematic review and meta-analysis. Acta Ophthalmol. 2022, 100, e669–e680. [Google Scholar] [CrossRef] [PubMed]
  103. Yu, Y.; Triebwasser, M.P.; Wong, E.K.; Schramm, E.C.; Thomas, B.; Reynolds, R.; Mardis, E.R.; Atkinson, J.P.; Daly, M.; Raychaudhuri, S.; et al. Whole-exome sequencing identifies rare, functional CFH variants in families with macular degeneration. Hum. Mol. Genet. 2014, 23, 5283–5293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Association Between Complement Factor C2/C3/CFB/CFH Polymorphisms and Age-Related Macular Degeneration: A Meta-Analysis. Genet. Test. Mol. Biomark. 2018, 22, 526–540. [CrossRef] [PubMed]
  105. Supanji, S.; Perdamaian, A.B.I.; Anindita, D.A.; Widayanti, T.W.; Wardhana, F.S.; Sasongko, M.B.; Prayogo, M.E.; Agni, A.N.; Oka, C. rs3753394 Complement Factor H (CFH) Gene Polymorphism in Patients with Age-Related Macular Degeneration (AMD) in Indonesian Population. BIO Web Conf. 2021, 41, 06001. [Google Scholar] [CrossRef]
  106. Duvvari, M.R.; Saksens, N.T.; van de Ven, J.P.; de Jong-Hesse, Y.; Schick, T.; Nillesen, W.M.; Fauser, S.; Hoefsloot, L.H.; Hoyng, C.B.; de Jong, E.K.; et al. Analysis of rare variants in the CFH gene in patients with the cuticular drusen subtype of age-related macular degeneration. Mol. Vis. 2015, 21, 285–292. [Google Scholar] [PubMed]
  107. Ferrara, D.; Seddon, J.M. Phenotypic Characterization of Complement Factor H R1210C Rare Genetic Variant in Age-Related Macular Degeneration. JAMA Ophthalmol. 2015, 133, 785–791. [Google Scholar] [CrossRef] [Green Version]
  108. Kersten, E.; Paun, C.C.; Schellevis, R.L.; Hoyng, C.B.; Delcourt, C.; Lengyel, I.; Peto, T.; Ueffing, M.; Klaver, C.C.W.; Dammeier, S.; et al. Systemic and ocular fluid compounds as potential biomarkers in age-related macular degeneration. Surv. Ophthalmol. 2018, 63, 9–39. [Google Scholar] [CrossRef] [Green Version]
  109. Yu, Y.; Wagner, E.K.; Souied, E.H.; Seitsonen, S.; Immonen, I.J.; Happola, P.; Raychaudhuri, S.; Daly, M.J.; Seddon, J.M. Protective coding variants in CFH and PELI3 and a variant near CTRB1 are associated with age-related macular degenerationdagger. Hum. Mol. Genet. 2016, 25, 5276–5285. [Google Scholar] [CrossRef] [Green Version]
  110. De Breuk, A.; Lechanteur, Y.T.E.; Heesterbeek, T.J.; Fauser, S.; Klaver, C.C.W.; Hoyng, C.B.; den Hollander, A.I. Genetic Risk in Families with Age-Related Macular Degeneration. Ophthalmol. Sci. 2021, 1, 100087. [Google Scholar] [CrossRef]
  111. Klein, M.L.; Francis, P.J.; Rosner, B.; Reynolds, R.; Hamon, S.C.; Schultz, D.W.; Ott, J.; Seddon, J.M. CFH and LOC387715/ARMS2 genotypes and treatment with antioxidants and zinc for age-related macular degeneration. Ophthalmology 2008, 115, 1019–1025. [Google Scholar] [CrossRef] [PubMed]
  112. Awh, C.C.; Hawken, S.; Zanke, B.W. Treatment response to antioxidants and zinc based on CFH and ARMS2 genetic risk allele number in the Age-Related Eye Disease Study. Ophthalmology 2015, 122, 162–169. [Google Scholar] [CrossRef] [PubMed]
  113. Seddon, J.M.; Silver, R.E.; Rosner, B. Response to AREDS supplements according to genetic factors: Survival analysis approach using the eye as the unit of analysis. Br. J. Ophthalmol. 2016, 100, 1731–1737. [Google Scholar] [CrossRef] [PubMed]
  114. Seddon, J.M.; Reynolds, R.; Rosner, B. Peripheral retinal drusen and reticular pigment: Association with CFHY402H and CFHrs1410996 genotypes in family and twin studies. Investig. Ophthalmol. Vis. Sci. 2009, 50, 586–591. [Google Scholar] [CrossRef] [Green Version]
  115. Geerlings, M.J.; Kersten, E.; Groenewoud, J.M.M.; Fritsche, L.G.; Hoyng, C.B.; de Jong, E.K.; den Hollander, A.I. Geographic distribution of rare variants associated with age-related macular degeneration. Mol. Vis. 2018, 24, 75–82. [Google Scholar]
  116. Fagerness, J.A.; Maller, J.B.; Neale, B.M.; Reynolds, R.C.; Daly, M.J.; Seddon, J.M. Variation near complement factor I is associated with risk of advanced AMD. Eur. J. Hum. Genet. 2009, 17, 100–104. [Google Scholar] [CrossRef]
  117. Van de Ven, J.P.; Nilsson, S.C.; Tan, P.L.; Buitendijk, G.H.; Ristau, T.; Mohlin, F.C.; Nabuurs, S.B.; Schoenmaker-Koller, F.E.; Smailhodzic, D.; Campochiaro, P.A.; et al. A functional variant in the CFI gene confers a high risk of age-related macular degeneration. Nat. Genet. 2013, 45, 813–817. [Google Scholar] [CrossRef]
  118. Hallam, T.M.; Marchbank, K.J.; Harris, C.L.; Osmond, C.; Shuttleworth, V.G.; Griffiths, H.; Cree, A.J.; Kavanagh, D.; Lotery, A.J. Rare Genetic Variants in Complement Factor I Lead to Low FI Plasma Levels Resulting in Increased Risk of Age-Related Macular Degeneration. Investig. Ophthalmol. Vis. Sci. 2020, 61, 18. [Google Scholar] [CrossRef]
  119. Java, A.; Baciu, P.; Widjajahakim, R.; Sung, Y.J.; Yang, J.; Kavanagh, D.; Atkinson, J.; Seddon, J. Functional Analysis of Rare Genetic Variants in Complement Factor I (CFI) using a Serum-Based Assay in Advanced Age-related Macular Degeneration. Transl. Vis. Sci. Technol. 2020, 9, 37. [Google Scholar] [CrossRef]
  120. De Jong, S.; Volokhina, E.B.; de Breuk, A.; Nilsson, S.C.; de Jong, E.K.; van der Kar, N.; Bakker, B.; Hoyng, C.B.; van den Heuvel, L.P.; Blom, A.M.; et al. Effect of rare coding variants in the CFI gene on Factor I expression levels. Hum. Mol. Genet. 2020, 29, 2313–2324. [Google Scholar] [CrossRef]
  121. Geerlings, M.J.; de Jong, E.K.; den Hollander, A.I. The complement system in age-related macular degeneration: A review of rare genetic variants and implications for personalized treatment. Mol. Immunol. 2017, 84, 65–76. [Google Scholar] [CrossRef] [PubMed]
  122. Ellis, S.; Buchberger, A.; Holder, J.; Orhan, e.; Hughes, J. GT005, a gene therapy for the treatment of dry age-related macular degeneration (AMD). Investig. Ophthalmol. Vis. Sci. 2020, 61, 2295. [Google Scholar]
  123. Maller, J.B.; Fagerness, J.A.; Reynolds, R.C.; Neale, B.M.; Daly, M.J.; Seddon, J.M. Variation in complement factor 3 is associated with risk of age-related macular degeneration. Nat. Genet. 2007, 39, 1200–1201. [Google Scholar] [CrossRef]
  124. Yates, J.R.; Sepp, T.; Matharu, B.K.; Khan, J.C.; Thurlby, D.A.; Shahid, H.; Clayton, D.G.; Hayward, C.; Morgan, J.; Wright, A.F.; et al. Complement C3 variant and the risk of age-related macular degeneration. N. Engl. J. Med. 2007, 357, 553–561. [Google Scholar] [CrossRef] [Green Version]
  125. Zhang, J.; Li, S.; Hu, S.; Yu, J.; Xiang, Y. Association between genetic variation of complement C3 and the susceptibility to advanced age-related macular degeneration: A meta-analysis. BMC Ophthalmol. 2018, 18, 274. [Google Scholar] [CrossRef] [PubMed]
  126. Heurich, M.; Martinez-Barricarte, R.; Francis, N.J.; Roberts, D.L.; Rodriguez de Cordoba, S.; Morgan, B.P.; Harris, C.L. Common polymorphisms in C3, factor B, and factor H collaborate to determine systemic complement activity and disease risk. Proc. Natl. Acad. Sci. USA 2011, 108, 8761–8766. [Google Scholar] [CrossRef] [Green Version]
  127. Duvvari, M.R.; Paun, C.C.; Buitendijk, G.H.; Saksens, N.T.; Volokhina, E.B.; Ristau, T.; Schoenmaker-Koller, F.E.; van de Ven, J.P.; Groenewoud, J.M.; van den Heuvel, L.P.; et al. Analysis of rare variants in the C3 gene in patients with age-related macular degeneration. PLoS ONE 2014, 9, e94165. [Google Scholar] [CrossRef] [Green Version]
  128. Pei, X.T.; Li, X.X.; Bao, Y.Z.; Yu, W.Z.; Yan, Z.; Qi, H.J.; Qian, T.; Xiao, H.X. Association of c3 gene polymorphisms with neovascular age-related macular degeneration in a chinese population. Curr. Eye Res. 2009, 34, 615–622. [Google Scholar] [CrossRef]
  129. Liao, D.S.; Grossi, F.V.; El Mehdi, D.; Gerber, M.R.; Brown, D.M.; Heier, J.S.; Wykoff, C.C.; Singerman, L.J.; Abraham, P.; Grassmann, F.; et al. Complement C3 Inhibitor Pegcetacoplan for Geographic Atrophy Secondary to Age-Related Macular Degeneration: A Randomized Phase 2 Trial. Ophthalmology 2020, 127, 186–195. [Google Scholar] [CrossRef] [Green Version]
  130. Anderson, D.H.; Mullins, R.F.; Hageman, G.S.; Johnson, L.V. A role for local inflammation in the formation of drusen in the aging eye. Am. J. Ophthalmol. 2002, 134, 411–431. [Google Scholar] [CrossRef]
  131. Johnson, L.V.; Ozaki, S.; Staples, M.K.; Erickson, P.A.; Anderson, D.H. A potential role for immune complex pathogenesis in drusen formation. Exp. Eye Res. 2000, 70, 441–449. [Google Scholar] [CrossRef] [PubMed]
  132. Scholl, H.P.; Charbel Issa, P.; Walier, M.; Janzer, S.; Pollok-Kopp, B.; Borncke, F.; Fritsche, L.G.; Chong, N.V.; Fimmers, R.; Wienker, T.; et al. Systemic complement activation in age-related macular degeneration. PLoS ONE 2008, 3, e2593. [Google Scholar] [CrossRef]
  133. Nozaki, M.; Raisler, B.J.; Sakurai, E.; Sarma, J.V.; Barnum, S.R.; Lambris, J.D.; Chen, Y.; Zhang, K.; Ambati, B.K.; Baffi, J.Z.; et al. Drusen complement components C3a and C5a promote choroidal neovascularization. Proc. Natl. Acad. Sci. USA 2006, 103, 2328–2333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Baas, D.C.; Ho, L.; Ennis, S.; Merriam, J.E.; Tanck, M.W.; Uitterlinden, A.G.; de Jong, P.T.; Cree, A.J.; Griffiths, H.L.; Rivadeneira, F.; et al. The complement component 5 gene and age-related macular degeneration. Ophthalmology 2010, 117, 500–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Liu, K.; Ma, L.; Lai, T.Y.Y.; Brelen, M.E.; Tam, P.O.S.; Tham, C.C.; Pang, C.P.; Chen, L.J. Evaluation of the association of C5 with neovascular age-related macular degeneration and polypoidal choroidal vasculopathy. Eye Vis. 2019, 6, 34. [Google Scholar] [CrossRef]
  136. Jaffe, G.J.; Westby, K.; Csaky, K.G.; Mones, J.; Pearlman, J.A.; Patel, S.S.; Joondeph, B.C.; Randolph, J.; Masonson, H.; Rezaei, K.A. C5 Inhibitor Avacincaptad Pegol for Geographic Atrophy Due to Age-Related Macular Degeneration: A Randomized Pivotal Phase 2/3 Trial. Ophthalmology 2021, 128, 576–586. [Google Scholar] [CrossRef]
  137. Saksens, N.T.; Geerlings, M.J.; Bakker, B.; Schick, T.; Daha, M.R.; Fauser, S.; Boon, C.J.; de Jong, E.K.; Hoyng, C.B.; den Hollander, A.I. Rare Genetic Variants Associated With Development of Age-Related Macular Degeneration. JAMA Ophthalmol. 2016, 134, 287–293. [Google Scholar] [CrossRef]
  138. Ratnapriya, R.; Acar, I.E.; Geerlings, M.J.; Branham, K.; Kwong, A.; Saksens, N.T.M.; Pauper, M.; Corominas, J.; Kwicklis, M.; Zipprer, D.; et al. Family-based exome sequencing identifies rare coding variants in age-related macular degeneration. Hum Mol. Genet. 2020, 29, 2022–2034. [Google Scholar] [CrossRef]
  139. McMahon, O.; Hallam, T.M.; Patel, S.; Harris, C.L.; Menny, A.; Zelek, W.M.; Widjajahakim, R.; Java, A.; Cox, T.E.; Tzoumas, N.; et al. The rare C9 P167S risk variant for age-related macular degeneration increases polymerization of the terminal component of the complement cascade. Hum. Mol. Genet. 2021, 30, 1188–1199. [Google Scholar] [CrossRef]
  140. Kremlitzka, M.; Geerlings, M.J.; de Jong, S.; Bakker, B.; Nilsson, S.C.; Fauser, S.; Hoyng, C.B.; de Jong, E.K.; den Hollander, A.I.; Blom, A.M. Functional analyses of rare genetic variants in complement component C9 identified in patients with age-related macular degeneration. Hum. Mol. Genet. 2018, 27, 2678–2688. [Google Scholar] [CrossRef]
  141. Horiuchi, T.; Nishizaka, H.; Kojima, T.; Sawabe, T.; Niho, Y.; Schneider, P.M.; Inaba, S.; Sakai, K.; Hayashi, K.; Hashimura, C.; et al. A non-sense mutation at Arg95 is predominant in complement 9 deficiency in Japanese. J. Immunol. 1998, 160, 1509–1513. [Google Scholar] [PubMed]
  142. Nishiguchi, K.M.; Yasuma, T.R.; Tomida, D.; Nakamura, M.; Ishikawa, K.; Kikuchi, M.; Ohmi, Y.; Niwa, T.; Hamajima, N.; Furukawa, K.; et al. C9-R95X polymorphism in patients with neovascular age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 2012, 53, 508–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Gold, B.; Merriam, J.E.; Zernant, J.; Hancox, L.S.; Taiber, A.J.; Gehrs, K.; Cramer, K.; Neel, J.; Bergeron, J.; Barile, G.R.; et al. Variation in factor B (BF) and complement component 2 (C2) genes is associated with age-related macular degeneration. Nat. Genet. 2006, 38, 458–462. [Google Scholar] [CrossRef] [Green Version]
  144. Mantel, I.; Ambresin, A.; Moetteli, L.; Droz, I.; Roduit, R.; Munier, F.L.; Schorderet, D.F. Complement factor B polymorphism and the phenotype of early age-related macular degeneration. Ophthalmic. Genet. 2014, 35, 12–17. [Google Scholar] [CrossRef] [PubMed]
  145. Spencer, K.L.; Hauser, M.A.; Olson, L.M.; Schmidt, S.; Scott, W.K.; Gallins, P.; Agarwal, A.; Postel, E.A.; Pericak-Vance, M.A.; Haines, J.L. Protective effect of complement factor B and complement component 2 variants in age-related macular degeneration. Hum. Mol. Genet. 2007, 16, 1986–1992. [Google Scholar] [CrossRef]
  146. Sun, C.; Zhao, M.; Li, X. CFB/C2 gene polymorphisms and risk of age-related macular degeneration: A systematic review and meta-analysis. Curr. Eye Res. 2012, 37, 259–271. [Google Scholar] [CrossRef]
  147. Thakkinstian, A.; McEvoy, M.; Chakravarthy, U.; Chakrabarti, S.; McKay, G.J.; Ryu, E.; Silvestri, G.; Kaur, I.; Francis, P.; Iwata, T.; et al. The association between complement component 2/complement factor B polymorphisms and age-related macular degeneration: A HuGE review and meta-analysis. Am. J. Epidemiol. 2012, 176, 361–372. [Google Scholar] [CrossRef]
  148. Gehrs, K.M.; Jackson, J.R.; Brown, E.N.; Allikmets, R.; Hageman, G.S. Complement, age-related macular degeneration and a vision of the future. Arch. Ophthalmol. 2010, 128, 349–358. [Google Scholar] [CrossRef] [Green Version]
  149. Lokki, M.L.; Koskimies, S.A. Allelic differences in hemolytic activity and protein concentration of BF molecules are found in association with particular HLA haplotypes. Immunogenetics 1991, 34, 242–246. [Google Scholar] [CrossRef]
  150. Yoneyama, S.; Sakurada, Y.; Kikushima, W.; Sugiyama, A.; Matsubara, M.; Fukuda, Y.; Tanabe, N.; Parikh, R.; Mabuchi, F.; Kashiwagi, K.; et al. Genetic factors associated with response to as-needed aflibercept therapy for typical neovascular age-related macular degeneration and polypoidal choroidal vasculopathy. Sci. Rep. 2020, 10, 7188. [Google Scholar] [CrossRef]
  151. Nakata, I.; Yamashiro, K.; Yamada, R.; Gotoh, N.; Nakanishi, H.; Hayashi, H.; Akagi-Kurashige, Y.; Tsujikawa, A.; Otani, A.; Saito, M.; et al. Significance of C2/CFB variants in age-related macular degeneration and polypoidal choroidal vasculopathy in a Japanese population. Investig. Ophthalmol. Vis. Sci. 2012, 53, 794–798. [Google Scholar] [CrossRef] [Green Version]
  152. Forneris, F.; Ricklin, D.; Wu, J.; Tzekou, A.; Wallace, R.S.; Lambris, J.D.; Gros, P. Structures of C3b in complex with factors B and D give insight into complement convertase formation. Science 2010, 330, 1816–1820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Stanton, C.M.; Yates, J.R.; den Hollander, A.I.; Seddon, J.M.; Swaroop, A.; Stambolian, D.; Fauser, S.; Hoyng, C.; Yu, Y.; Atsuhiro, K.; et al. Complement factor D in age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 2011, 52, 8828–8834. [Google Scholar] [CrossRef] [PubMed]
  154. Zeng, J.; Chen, Y.; Tong, Z.; Zhou, X.; Zhao, C.; Wang, K.; Hughes, G.; Kasuga, D.; Bedell, M.; Lee, C.; et al. Lack of association of CFD polymorphisms with advanced age-related macular degeneration. Mol. Vis. 2010, 16, 2273–2278. [Google Scholar]
  155. Holz, F.G.; Sadda, S.R.; Busbee, B.; Chew, E.Y.; Mitchell, P.; Tufail, A.; Brittain, C.; Ferrara, D.; Gray, S.; Honigberg, L.; et al. Efficacy and Safety of Lampalizumab for Geographic Atrophy Due to Age-Related Macular Degeneration: Chroma and Spectri Phase 3 Randomized Clinical Trials. JAMA Ophthalmol. 2018, 136, 666–677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Hogan, M.J. Ultrastructure of the choroid. Its role in the pathogenesis of chorioretinal disease. Trans. Pac. Coast Otoophthalmol. Soc. Annu. Meet 1961, 42, 61–87. [Google Scholar] [PubMed]
  157. Booij, J.C.; Baas, D.C.; Beisekeeva, J.; Gorgels, T.G.; Bergen, A.A. The dynamic nature of Bruch’s membrane. Prog. Retin. Eye Res. 2010, 29, 1–18. [Google Scholar] [CrossRef]
  158. Massova, I.; Kotra, L.P.; Fridman, R.; Mobashery, S. Matrix metalloproteinases: Structures, evolution, and diversification. FASEB J. 1998, 12, 1075–1095. [Google Scholar] [CrossRef] [Green Version]
  159. Hussain, A.A.; Lee, Y.; Zhang, J.J.; Marshall, J. Disturbed matrix metalloproteinase activity of Bruch’s membrane in age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 2011, 52, 4459–4466. [Google Scholar] [CrossRef]
  160. Garcia-Onrubia, L.; Valentin-Bravo, F.J.; Coco-Martin, R.M.; Gonzalez-Sarmiento, R.; Pastor, J.C.; Usategui-Martin, R.; Pastor-Idoate, S. Matrix Metalloproteinases in Age-Related Macular Degeneration (AMD). Int. J. Mol. Sci. 2020, 21, 5934. [Google Scholar] [CrossRef]
  161. Guo, L.; Hussain, A.A.; Limb, G.A.; Marshall, J. Age-dependent variation in metalloproteinase activity of isolated human Bruch’s membrane and choroid. Investig. Ophthalmol. Vis. Sci. 1999, 40, 2676–2682. [Google Scholar]
  162. Kamei, M.; Hollyfield, J.G. TIMP-3 in Bruch’s membrane: Changes during aging and in age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 1999, 40, 2367–2375. [Google Scholar]
  163. Krogh Nielsen, M.; Subhi, Y.; Rue Molbech, C.; Nilsson, L.L.; Nissen, M.H.; Sorensen, T.L. Imbalances in tissue inhibitors of metalloproteinases differentiate choroidal neovascularization from geographic atrophy. Acta Ophthalmol. 2019, 97, 84–90. [Google Scholar] [CrossRef] [Green Version]
  164. Abecasis, G.R.; Yashar, B.M.; Zhao, Y.; Ghiasvand, N.M.; Zareparsi, S.; Branham, K.E.; Reddick, A.C.; Trager, E.H.; Yoshida, S.; Bahling, J.; et al. Age-related macular degeneration: A high-resolution genome scan for susceptibility loci in a population enriched for late-stage disease. Am. J. Hum. Genet. 2004, 74, 482–494. [Google Scholar] [CrossRef] [Green Version]
  165. Ardeljan, D.; Chan, C.C. Aging is not a disease: Distinguishing age-related macular degeneration from aging. Prog. Retin. Eye Res. 2013, 37, 68–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Della, N.G.; Campochiaro, P.A.; Zack, D.J. Localization of TIMP-3 mRNA expression to the retinal pigment epithelium. Investig. Ophthalmol. Vis. Sci. 1996, 37, 1921–1924. [Google Scholar]
  167. Yu, W.H.; Yu, S.; Meng, Q.; Brew, K.; Woessner, J.F., Jr. TIMP-3 binds to sulfated glycosaminoglycans of the extracellular matrix. J. Biol. Chem. 2000, 275, 31226–31232. [Google Scholar] [CrossRef] [Green Version]
  168. Fariss, R.N.; Apte, S.S.; Olsen, B.R.; Iwata, K.; Milam, A.H. Tissue inhibitor of metalloproteinases-3 is a component of Bruch’s membrane of the eye. Am. J. Pathol. 1997, 150, 323–328. [Google Scholar]
  169. Arpino, V.; Brock, M.; Gill, S.E. The role of TIMPs in regulation of extracellular matrix proteolysis. Matrix Biol. 2015, 44–46, 247–254. [Google Scholar] [CrossRef]
  170. Bailey, T.A.; Alexander, R.A.; Dubovy, S.R.; Luthert, P.J.; Chong, N.H. Measurement of TIMP-3 expression and Bruch’s membrane thickness in human macula. Exp. Eye Res. 2001, 73, 851–858. [Google Scholar] [CrossRef]
  171. Qi, J.H.; Ebrahem, Q.; Moore, N.; Murphy, G.; Claesson-Welsh, L.; Bond, M.; Baker, A.; Anand-Apte, B. A novel function for tissue inhibitor of metalloproteinases-3 (TIMP3): Inhibition of angiogenesis by blockage of VEGF binding to VEGF receptor-2. Nat. Med. 2003, 9, 407–415. [Google Scholar] [CrossRef] [PubMed]
  172. Weber, B.H.; Vogt, G.; Wolz, W.; Ives, E.J.; Ewing, C.C. Sorsby’s fundus dystrophy is genetically linked to chromosome 22q13-qter. Nat. Genet. 1994, 7, 158–161. [Google Scholar] [CrossRef] [PubMed]
  173. Cheng, Y.; Cheng, T.; Qu, Y. TIMP-3 suppression induces choroidal neovascularization by moderating the polarization of macrophages in age-related macular degeneration. Mol. Immunol. 2019, 106, 119–126. [Google Scholar] [CrossRef] [PubMed]
  174. Kauer, W.K.; Stein, H.J. Emerging concepts of bile reflux in the constellation of gastroesophageal reflux disease. J. Gastrointest. Surg. 2010, 14, S9–S16. [Google Scholar] [CrossRef]
  175. Cascella, R.; Strafella, C.; Longo, G.; Ragazzo, M.; Manzo, L.; De Felici, C.; Errichiello, V.; Caputo, V.; Viola, F.; Eandi, C.M.; et al. Uncovering genetic and non-genetic biomarkers specific for exudative age-related macular degeneration: Significant association of twelve variants. Oncotarget 2018, 9, 7812–7821. [Google Scholar] [CrossRef] [Green Version]
  176. Fauser, S.; Smailhodzic, D.; Caramoy, A.; van de Ven, J.P.; Kirchhof, B.; Hoyng, C.B.; Klevering, B.J.; Liakopoulos, S.; den Hollander, A.I. Evaluation of serum lipid concentrations and genetic variants at high-density lipoprotein metabolism loci and TIMP3 in age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 2011, 52, 5525–5528. [Google Scholar] [CrossRef] [Green Version]
  177. Tuo, J.; Grob, S.; Zhang, K.; Chan, C.C. Genetics of immunological and inflammatory components in age-related macular degeneration. Ocul. Immunol. Inflamm. 2012, 20, 27–36. [Google Scholar] [CrossRef] [Green Version]
  178. Zeng, R.; Wen, F.; Zhang, X.; Zuo, C.; Li, M.; Chen, H.; Wu, K. An rs9621532 variant near the TIMP3 gene is not associated with neovascular age-related macular degeneration and polypoidal choroidal vasculopathy in a Chinese Han population. Ophthalmic Genet. 2012, 33, 139–143. [Google Scholar] [CrossRef]
  179. Liutkeviciene, R.; Vilkeviciute, A.; Gedvilaite, G.; Kaikaryte, K.; Kriauciuniene, L. Haplotypes of HTRA1 rs1120638, TIMP3 rs9621532, VEGFA rs833068, CFI rs10033900, ERCC6 rs3793784, and KCTD10 rs56209061 Gene Polymorphisms in Age-Related Macular Degeneration. Dis. Markers 2019, 2019, 9602949. [Google Scholar] [CrossRef] [Green Version]
  180. Ardeljan, D.; Meyerle, C.B.; Agron, E.; Wang, J.J.; Mitchell, P.; Chew, E.Y.; Zhao, J.; Maminishkis, A.; Chan, C.C.; Tuo, J. Influence of TIMP3/SYN3 polymorphisms on the phenotypic presentation of age-related macular degeneration. Eur. J. Hum. Genet. 2013, 21, 1152–1157. [Google Scholar] [CrossRef]
  181. Ortak, H.; Demir, S.; Ates, O.; Benli, I.; Sogut, E.; Sahin, M. The role of MMP2 (-1306C>T) and TIMP2 (-418 G>C) promoter variants in age-related macular degeneration. Ophthalmic Genet 2013, 34, 217–222. [Google Scholar] [CrossRef] [PubMed]
  182. Oszajca, K.; Szemraj, M.; Szemraj, J.; Jurowski, P. Association analysis of genetic polymorphisms and expression levels of selected genes involved in extracellular matrix turnover and angiogenesis with the risk of age-related macular degeneration. Ophthalmic Genet. 2018, 39, 684–698. [Google Scholar] [CrossRef] [PubMed]
  183. Liutkeviciene, R.; Lesauskaite, V.; Sinkunaite-Marsalkiene, G.; Simonyte, S.; Zemaitiene, R.; Kriauciuniene, L.; Zaliuniene, D. MMP-2 Rs24386 (C-->T) gene polymorphism and the phenotype of age-related macular degeneration. Int. J. Ophthalmol. 2017, 10, 1349–1353. [Google Scholar] [CrossRef] [PubMed]
  184. Liutkeviciene, R.; Lesauskaite, V.; Sinkunaite-Marsalkiene, G.; Zaliuniene, D.; Zaliaduonyte-Peksiene, D.; Mizariene, V.; Gustiene, O.; Jasinskas, V.; Jariene, G.; Tamosiunas, A. The Role of Matrix Metalloproteinases Polymorphisms in Age-Related Macular Degeneration. Ophthalmic Genet. 2015, 36, 149–155. [Google Scholar] [CrossRef]
  185. Liutkeviciene, R.; Vilkeviciute, A.; Borisovaite, D.; Miniauskiene, G. Association of exudative age-related macular degeneration with matrix metalloproteinases-2 (-1306 C/T) rs243865 gene polymorphism. Indian J. Ophthalmol. 2018, 66, 551–557. [Google Scholar] [CrossRef]
  186. Seitzman, R.L.; Mahajan, V.B.; Mangione, C.; Cauley, J.A.; Ensrud, K.E.; Stone, K.L.; Cummings, S.R.; Hochberg, M.C.; Hillier, T.A.; Sinsheimer, J.S.; et al. Estrogen receptor α and matrix metalloproteinase 2 polymorphisms and age-related maculopathy in older women. Am. J. Epidemiol. 2008, 167, 1217–1225. [Google Scholar] [CrossRef] [Green Version]
  187. Usategui-Martin, R.; Pastor-Idoate, S.; Chamorro, A.J.; Fernandez, I.; Fernandez-Bueno, I.; Marcos-Martin, M.; Gonzalez-Sarmiento, R.; Carlos Pastor, J. Meta-analysis of the rs243865 MMP-2 polymorphism and age-related macular degeneration risk. PLoS ONE 2019, 14, e0213624. [Google Scholar] [CrossRef] [Green Version]
  188. Cheng, J.; Hao, X.; Zhang, Z. Risk of macular degeneration affected by polymorphisms in Matrix metalloproteinase-2: A case-control study in Chinese Han population. Medicine 2017, 96, e8190. [Google Scholar] [CrossRef]
  189. Yan, Q.; Ding, Y.; Liu, Y.; Sun, T.; Fritsche, L.G.; Clemons, T.; Ratnapriya, R.; Klein, M.L.; Cook, R.J.; Liu, Y.; et al. Genome-wide analysis of disease progression in age-related macular degeneration. Hum. Mol. Genet. 2018, 27, 929–940. [Google Scholar] [CrossRef] [Green Version]
  190. Sohn, E.H.; Han, I.C.; Roos, B.R.; Faga, B.; Luse, M.A.; Binkley, E.M.; Boldt, H.C.; Folk, J.C.; Russell, S.R.; Mullins, R.F.; et al. Genetic Association between MMP9 and Choroidal Neovascularization in Age-Related Macular Degeneration. Ophthalmol. Sci. 2021, 1, 100002. [Google Scholar] [CrossRef]
  191. Fiotti, N.; Pedio, M.; Battaglia Parodi, M.; Altamura, N.; Uxa, L.; Guarnieri, G.; Giansante, C.; Ravalico, G. MMP-9 microsatellite polymorphism and susceptibility to exudative form of age-related macular degeneration. Genet. Med. 2005, 7, 272–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Budiene, B.; Liutkeviciene, R.; Gustiene, O.; Ugenskiene, R.; Laukaitiene, D.; Savukaityte, A.; Vilkeviciute, A.; Steponaviciute, R.; Rocyte, A.; Zaliuniene, D. The association of matrix metalloproteinases polymorphisms and interleukins in advanced age-related macular degeneration. Ophthalmic Genet. 2018, 39, 463–472. [Google Scholar] [CrossRef] [PubMed]
  193. Liutkeviciene, R.; Zaliaduonyte-Peksiene, D.; Zaliuniene, D.; Gustiene, O.; Jasinskas, V.; Lesauskaite, V.; Tamosiunas, A.; Zaliunas, R. Does matrix metalloproteinase-3 polymorphism play a role in age-related macular degeneration in patients with myocardial infarction? Medicina 2012, 48, 404–409. [Google Scholar] [CrossRef] [PubMed]
  194. Tamura, Y.; Konomi, H.; Sawada, H.; Takashima, S.; Nakajima, A. Tissue distribution of type VIII collagen in human adult and fetal eyes. Investig. Ophthalmol. Vis. Sci. 1991, 32, 2636–2644. [Google Scholar]
  195. Corominas, J.; Colijn, J.M.; Geerlings, M.J.; Pauper, M.; Bakker, B.; Amin, N.; Lores Motta, L.; Kersten, E.; Garanto, A.; Verlouw, J.A.M.; et al. Whole-Exome Sequencing in Age-Related Macular Degeneration Identifies Rare Variants in COL8A1, a Component of Bruch’s Membrane. Ophthalmology 2018, 125, 1433–1443. [Google Scholar] [CrossRef] [Green Version]
  196. Den Hollander, A.I.; Hoyng, C.B.; Boon, C.J.F. Complement Factor H Gene Mutations: Implications for Genetic Testing and Precision Medicine in Macular Degeneration. Ophthalmology 2019, 126, 1422–1423. [Google Scholar] [CrossRef] [Green Version]
  197. Meyer, K.J.; Davis, L.K.; Schindler, E.I.; Beck, J.S.; Rudd, D.S.; Grundstad, A.J.; Scheetz, T.E.; Braun, T.A.; Fingert, J.H.; Alward, W.L.; et al. Genome-wide analysis of copy number variants in age-related macular degeneration. Hum. Genet. 2011, 129, 91–100. [Google Scholar] [CrossRef] [Green Version]
  198. Marmorstein, A.D.; Marmorstein, L.Y.; Sakaguchi, H.; Hollyfield, J.G. Spectral profiling of autofluorescence associated with lipofuscin, Bruch’s Membrane, and sub-RPE deposits in normal and AMD eyes. Investig. Ophthalmol. Vis. Sci. 2002, 43, 2435–2441. [Google Scholar]
  199. Van Leeuwen, E.M.; Emri, E.; Merle, B.M.J.; Colijn, J.M.; Kersten, E.; Cougnard-Gregoire, A.; Dammeier, S.; Meester-Smoor, M.; Pool, F.M.; de Jong, E.K.; et al. A new perspective on lipid research in age-related macular degeneration. Prog. Retin. Eye Res. 2018, 67, 56–86. [Google Scholar] [CrossRef]
  200. Curcio, C.A. Antecedents of Soft Drusen, the Specific Deposits of Age-Related Macular Degeneration, in the Biology of Human Macula. Investig. Ophthalmol. Vis. Sci. 2018, 59, AMD182–AMD194. [Google Scholar] [CrossRef] [Green Version]
  201. Curcio, C.A. Photoreceptor topography in ageing and age-related maculopathy. Eye 2001, 15, 376–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Curcio, C.A.; Millican, C.L.; Bailey, T.; Kruth, H.S. Accumulation of cholesterol with age in human Bruch’s membrane. Investig. Ophthalmol. Vis. Sci. 2001, 42, 265–274. [Google Scholar]
  203. Curcio, C.A. Soft Drusen in Age-Related Macular Degeneration: Biology and Targeting Via the Oil Spill Strategies. Investig. Ophthalmol. Vis. Sci. 2018, 59, AMD160–AMD181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Crabb, J.W.; Miyagi, M.; Gu, X.; Shadrach, K.; West, K.A.; Sakaguchi, H.; Kamei, M.; Hasan, A.; Yan, L.; Rayborn, M.E.; et al. Drusen proteome analysis: An approach to the etiology of age-related macular degeneration. Proc. Natl. Acad. Sci. USA 2002, 99, 14682–14687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Curcio, C.A.; Johnson, M.; Huang, J.D.; Rudolf, M. Aging, age-related macular degeneration, and the response-to-retention of apolipoprotein B-containing lipoproteins. Prog. Retin. Eye Res. 2009, 28, 393–422. [Google Scholar] [CrossRef] [Green Version]
  206. Mullins, R.F.; Russell, S.R.; Anderson, D.H.; Hageman, G.S. Drusen associated with aging and age-related macular degeneration contain proteins common to extracellular deposits associated with atherosclerosis, elastosis, amyloidosis, and dense deposit disease. FASEB J. 2000, 14, 835–846. [Google Scholar] [CrossRef]
  207. Zweifel, S.A.; Spaide, R.F.; Curcio, C.A.; Malek, G.; Imamura, Y. Reticular pseudodrusen are subretinal drusenoid deposits. Ophthalmology 2010, 117, 303–312.e301. [Google Scholar] [CrossRef]
  208. Zweifel, S.A.; Imamura, Y.; Spaide, T.C.; Fujiwara, T.; Spaide, R.F. Prevalence and significance of subretinal drusenoid deposits (reticular pseudodrusen) in age-related macular degeneration. Ophthalmology 2010, 117, 1775–1781. [Google Scholar] [CrossRef]
  209. Ross, R. Atherosclerosis—an inflammatory disease. N. Engl. J. Med. 1999, 340, 115–126. [Google Scholar] [CrossRef]
  210. Tall, A.R.; Yvan-Charvet, L. Cholesterol, inflammation and innate immunity. Nat. Rev. Immunol. 2015, 15, 104–116. [Google Scholar] [CrossRef] [Green Version]
  211. Lusis, A.J. Atherosclerosis. Nature 2000, 407, 233–241. [Google Scholar] [CrossRef] [PubMed]
  212. Murphy, A.J.; Woollard, K.J.; Hoang, A.; Mukhamedova, N.; Stirzaker, R.A.; McCormick, S.P.; Remaley, A.T.; Sviridov, D.; Chin-Dusting, J. High-density lipoprotein reduces the human monocyte inflammatory response. Arter. Thromb. Vasc. Biol. 2008, 28, 2071–2077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Wang, W.; Connor, S.L.; Johnson, E.J.; Klein, M.L.; Hughes, S.; Connor, W.E. Effect of dietary lutein and zeaxanthin on plasma carotenoids and their transport in lipoproteins in age-related macular degeneration. Am. J. Clin. Nutr. 2007, 85, 762–769. [Google Scholar] [CrossRef]
  214. Mahley, R.W.; Innerarity, T.L.; Rall, S.C., Jr.; Weisgraber, K.H. Plasma lipoproteins: Apolipoprotein structure and function. J. Lipid Res. 1984, 25, 1277–1294. [Google Scholar] [CrossRef]
  215. Mahley, R.W. Apolipoprotein E: Cholesterol transport protein with expanding role in cell biology. Science 1988, 240, 622–630. [Google Scholar] [CrossRef] [PubMed]
  216. Zannis, V.I.; Just, P.W.; Breslow, J.L. Human apolipoprotein E isoprotein subclasses are genetically determined. Am. J. Hum. Genet. 1981, 33, 11–24. [Google Scholar] [PubMed]
  217. Farrer, L.A.; Cupples, L.A.; Haines, J.L.; Hyman, B.; Kukull, W.A.; Mayeux, R.; Myers, R.H.; Pericak-Vance, M.A.; Risch, N.; van Duijn, C.M. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease. A meta-analysis. APOE and Alzheimer Disease Meta Analysis Consortium. JAMA 1997, 278, 1349–1356. [Google Scholar] [CrossRef]
  218. Zannis, V.I.; Kardassis, D.; Zanni, E.E. Genetic mutations affecting human lipoproteins, their receptors, and their enzymes. Adv. Hum. Genet. 1993, 21, 145–319. [Google Scholar] [CrossRef]
  219. Souied, E.H.; Benlian, P.; Amouyel, P.; Feingold, J.; Lagarde, J.P.; Munnich, A.; Kaplan, J.; Coscas, G.; Soubrane, G. The epsilon4 allele of the apolipoprotein E gene as a potential protective factor for exudative age-related macular degeneration. Am. J. Ophthalmol. 1998, 125, 353–359. [Google Scholar] [CrossRef]
  220. Schmidt, S.; Klaver, C.; Saunders, A.; Postel, E.; De La Paz, M.; Agarwal, A.; Small, K.; Udar, N.; Ong, J.; Chalukya, M.; et al. A pooled case-control study of the apolipoprotein E (APOE) gene in age-related maculopathy. Ophthalmic Genet. 2002, 23, 209–223. [Google Scholar] [CrossRef]
  221. Zareparsi, S.; Reddick, A.C.; Branham, K.E.; Moore, K.B.; Jessup, L.; Thoms, S.; Smith-Wheelock, M.; Yashar, B.M.; Swaroop, A. Association of apolipoprotein E alleles with susceptibility to age-related macular degeneration in a large cohort from a single center. Investig. Ophthalmol. Vis. Sci. 2004, 45, 1306–1310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. McKay, G.J.; Patterson, C.C.; Chakravarthy, U.; Dasari, S.; Klaver, C.C.; Vingerling, J.R.; Ho, L.; de Jong, P.T.; Fletcher, A.E.; Young, I.S.; et al. Evidence of association of APOE with age-related macular degeneration: A pooled analysis of 15 studies. Hum. Mutat. 2011, 32, 1407–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Shen, L.; Hoffmann, T.J.; Melles, R.B.; Sakoda, L.C.; Kvale, M.N.; Banda, Y.; Schaefer, C.; Risch, N.; Jorgenson, E. Differences in the Genetic Susceptibility to Age-Related Macular Degeneration Clinical Subtypes. Investig. Ophthalmol. Vis. Sci. 2015, 56, 4290–4299. [Google Scholar] [CrossRef] [Green Version]
  224. Xiying, M.; Wenbo, W.; Wangyi, F.; Qinghuai, L. Association of Apolipoprotein E Polymorphisms with Age-related Macular Degeneration Subtypes: An Updated Systematic Review and Meta-analysis. Arch. Med. Res. 2017, 48, 370–377. [Google Scholar] [CrossRef] [PubMed]
  225. Viturino, M.G.; Neto, J.M.; Bajano, F.F.; Costa, S.M.; Roque, A.B.; Borges, G.F.; Ananina, G.; Rim, P.H.; Medina, F.M.; Costa, F.F.; et al. Evaluation of APOE polymorphisms and the risk for age-related macular degeneration in a Southeastern Brazilian population. Exp. Biol. Med. 2021, 246, 1148–1155. [Google Scholar] [CrossRef] [PubMed]
  226. Liutkeviciene, R.; Vilkeviciute, A.; Smalinskiene, A.; Tamosiunas, A.; Petkeviciene, J.; Zaliuniene, D.; Lesauskaite, V. The role of apolipoprotein E (rs7412 and rs429358) in age-related macular degeneration. Ophthalmic Genet. 2018, 39, 457–462. [Google Scholar] [CrossRef]
  227. Baird, P.N.; Richardson, A.J.; Robman, L.D.; Dimitrov, P.N.; Tikellis, G.; McCarty, C.A.; Guymer, R.H. Apolipoprotein (APOE) gene is associated with progression of age-related macular degeneration (AMD). Hum. Mutat. 2006, 27, 337–342. [Google Scholar] [CrossRef] [PubMed]
  228. Adams, M.K.; Simpson, J.A.; Richardson, A.J.; English, D.R.; Aung, K.Z.; Makeyeva, G.A.; Guymer, R.H.; Giles, G.G.; Hopper, J.; Robman, L.D.; et al. Apolipoprotein E gene associations in age-related macular degeneration: The Melbourne Collaborative Cohort Study. Am. J. Epidemiol. 2012, 175, 511–518. [Google Scholar] [CrossRef] [Green Version]
  229. Holliday, E.G.; Smith, A.V.; Cornes, B.K.; Buitendijk, G.H.; Jensen, R.A.; Sim, X.; Aspelund, T.; Aung, T.; Baird, P.N.; Boerwinkle, E.; et al. Insights into the genetic architecture of early stage age-related macular degeneration: A genome-wide association study meta-analysis. PLoS ONE 2013, 8, e53830. [Google Scholar] [CrossRef] [Green Version]
  230. Sun, Y.; Song, R.; Ai, Y.; Zhu, J.; He, J.; Dang, M.; Li, H. APOE2 promotes the development and progression of subretinal neovascularization in age-related macular degeneration via MAPKs signaling pathway. Saudi J. Biol. Sci. 2020, 27, 2770–2777. [Google Scholar] [CrossRef]
  231. Guerra, R.; Wang, J.; Grundy, S.M.; Cohen, J.C. A hepatic lipase (LIPC) allele associated with high plasma concentrations of high density lipoprotein cholesterol. Proc. Natl. Acad. Sci. USA 1997, 94, 4532–4537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Chavali, V.R.M.; Haider, N.; Bell, B.A.; Song, Y.; Dunaief, J.L. Hepatic Lipase C knockout mouse has diminished ERGs and Bruch’s lipid deposits. Investig. Ophthalmol. Vis. Sci. 2019, 60, 1343. [Google Scholar]
  233. Reynolds, R.; Rosner, B.; Seddon, J.M. Serum lipid biomarkers and hepatic lipase gene associations with age-related macular degeneration. Ophthalmology 2010, 117, 1989–1995. [Google Scholar] [CrossRef] [Green Version]
  234. Wang, Y.F.; Han, Y.; Zhang, R.; Qin, L.; Wang, M.X.; Ma, L. CETP/LPL/LIPC gene polymorphisms and susceptibility to age-related macular degeneration. Sci. Rep. 2015, 5, 15711. [Google Scholar] [CrossRef]
  235. Liutkeviciene, R.; Vilkeviciute, A.; Kriauciuniene, L.; Deltuva, V.P. SIRT1 rs12778366, FGFR2 rs2981582, STAT3 rs744166, LIPC rs10468017, rs493258 and LPL rs12678919 genotypes and haplotype evaluation in patients with age-related macular degeneration. Gene 2019, 686, 8–15. [Google Scholar] [CrossRef]
  236. Lee, J.; Zeng, J.; Hughes, G.; Chen, Y.; Grob, S.; Zhao, L.; Lee, C.; Krupa, M.; Quach, J.; Luo, J.; et al. Association of LIPC and advanced age-related macular degeneration. Eye 2013, 27, 265–270; quiz 271. [Google Scholar] [CrossRef] [Green Version]
  237. Fernández-Vega, B.; García, M.; Olivares, L.; Álvarez, L.; González-Fernández, A.; Artime, E.; Fernández-Vega Cueto, A.; Cobo, T.; Coca-Prados, M.; Vega, J.A.; et al. The association study of lipid metabolism gene polymorphisms with AMD identifies a protective role for APOE-E2 allele in the wet form in a Northern Spanish population. Acta Ophthalmol. 2020, 98, e282–e291. [Google Scholar] [CrossRef] [PubMed]
  238. Kathiresan, S.; Willer, C.J.; Peloso, G.M.; Demissie, S.; Musunuru, K.; Schadt, E.E.; Kaplan, L.; Bennett, D.; Li, Y.; Tanaka, T.; et al. Common variants at 30 loci contribute to polygenic dyslipidemia. Nat. Genet. 2009, 41, 56–65. [Google Scholar] [CrossRef] [Green Version]
  239. Sharma, K.; Battu, P.; Singh, R.; Sharma, S.K.; Anand, A. Modulated anti-VEGF therapy under the influence of lipid metabolizing proteins in Age related macular degeneration: A pilot study. Sci. Rep. 2022, 12, 714. [Google Scholar] [CrossRef]
  240. Barter, P.; Kastelein, J.; Nunn, A.; Hobbs, R.; Future Forum Editorial, B. High density lipoproteins (HDLs) and atherosclerosis; the unanswered questions. Atherosclerosis 2003, 168, 195–211. [Google Scholar] [CrossRef]
  241. Tserentsoodol, N.; Sztein, J.; Campos, M.; Gordiyenko, N.V.; Fariss, R.N.; Lee, J.W.; Fliesler, S.J.; Rodriguez, I.R. Uptake of cholesterol by the retina occurs primarily via a low density lipoprotein receptor-mediated process. Mol. Vis. 2006, 12, 1306–1318. [Google Scholar]
  242. Van Leeuwen, R.; Klaver, C.C.; Vingerling, J.R.; Hofman, A.; van Duijn, C.M.; Stricker, B.H.; de Jong, P.T. Cholesterol and age-related macular degeneration: Is there a link? Am. J. Ophthalmol. 2004, 137, 750–752. [Google Scholar] [CrossRef] [PubMed]
  243. Yokoyama, S. ABCA1 and biogenesis of HDL. J. Atheroscler. Thromb. 2006, 13, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Storti, F.; Klee, K.; Todorova, V.; Steiner, R.; Othman, A.; van der Velde-Visser, S.; Samardzija, M.; Meneau, I.; Barben, M.; Karademir, D.; et al. Impaired ABCA1/ABCG1-mediated lipid efflux in the mouse retinal pigment epithelium (RPE) leads to retinal degeneration. eLife 2019, 8, e45100. [Google Scholar] [CrossRef] [PubMed]
  245. Peters, F.; Ebner, L.J.A.; Atac, D.; Maggi, J.; Berger, W.; den Hollander, A.I.; Grimm, C. Regulation of ABCA1 by AMD-Associated Genetic Variants and Hypoxia in iPSC-RPE. Int. J. Mol. Sci. 2022, 23, 3194. [Google Scholar] [CrossRef]
  246. Pikuleva, I.A.; Curcio, C.A. Cholesterol in the retina: The best is yet to come. Prog. Retin. Eye Res. 2014, 41, 64–89. [Google Scholar] [CrossRef] [Green Version]
  247. Wang, Y.; Wang, M.; Han, Y.; Zhang, R.; Ma, L. ABCA1 rs1883025 polymorphism and risk of age-related macular degeneration. Graefes Arch. Clin. Exp. Ophthalmol. 2016, 254, 323–332. [Google Scholar] [CrossRef]
  248. Mead, J.R.; Irvine, S.A.; Ramji, D.P. Lipoprotein lipase: Structure, function, regulation, and role in disease. J. Mol. Med. 2002, 80, 753–769. [Google Scholar] [CrossRef]
  249. Peter, I.; Huggins, G.S.; Ordovas, J.M.; Haan, M.; Seddon, J.M. Evaluation of new and established age-related macular degeneration susceptibility genes in the Women’s Health Initiative Sight Exam (WHI-SE) Study. Am. J. Ophthalmol. 2011, 152, 1005–1013.e1001. [Google Scholar] [CrossRef] [Green Version]
  250. Tian, J.; Yu, W.; Qin, X.; Fang, K.; Chen, Q.; Hou, J.; Li, J.; Chen, D.; Hu, Y.; Li, X. Association of genetic polymorphisms and age-related macular degeneration in Chinese population. Investig. Ophthalmol. Vis. Sci. 2012, 53, 4262–4269. [Google Scholar] [CrossRef] [Green Version]
  251. Zhang, X.; Li, M.; Wen, F.; Zuo, C.; Chen, H.; Wu, K.; Zeng, R. Different impact of high-density lipoprotein-related genetic variants on polypoidal choroidal vasculopathy and neovascular age-related macular degeneration in a Chinese Han population. Exp. Eye Res. 2013, 108, 16–22. [Google Scholar] [CrossRef] [PubMed]
  252. Merle, B.M.; Maubaret, C.; Korobelnik, J.F.; Delyfer, M.N.; Rougier, M.B.; Lambert, J.C.; Amouyel, P.; Malet, F.; Le Goff, M.; Dartigues, J.F.; et al. Association of HDL-related loci with age-related macular degeneration and plasma lutein and zeaxanthin: The Alienor study. PLoS ONE 2013, 8, e79848. [Google Scholar] [CrossRef]
  253. Grossniklaus, H.E.; Green, W.R. Choroidal neovascularization. Am. J. Ophthalmol. 2004, 137, 496–503. [Google Scholar] [CrossRef] [PubMed]
  254. Kaplan, H.J.; Leibole, M.A.; Tezel, T.; Ferguson, T.A. Fas ligand (CD95 ligand) controls angiogenesis beneath the retina. Nat. Med. 1999, 5, 292–297. [Google Scholar] [CrossRef]
  255. Kim, I.; Ryan, A.M.; Rohan, R.; Amano, S.; Agular, S.; Miller, J.W.; Adamis, A.P. Constitutive expression of VEGF, VEGFR-1, and VEGFR-2 in normal eyes. Investig. Ophthalmol. Vis. Sci. 1999, 40, 2115–2121. [Google Scholar]
  256. Ferrara, N. VEGF and Intraocular Neovascularization: From Discovery to Therapy. Transl. Vis. Sci. Technol. 2016, 5, 10. [Google Scholar] [CrossRef] [Green Version]
  257. Dawson, D.W.; Volpert, O.V.; Gillis, P.; Crawford, S.E.; Xu, H.; Benedict, W.; Bouck, N.P. Pigment epithelium-derived factor: A potent inhibitor of angiogenesis. Science 1999, 285, 245–248. [Google Scholar] [CrossRef]
  258. Zarbin, M.A. Current concepts in the pathogenesis of age-related macular degeneration. Arch. Ophthalmol. 2004, 122, 598–614. [Google Scholar] [CrossRef] [Green Version]
  259. Ferrara, N.; Adamis, A.P. Ten years of anti-vascular endothelial growth factor therapy. Nat. Rev. Drug Discov. 2016, 15, 385–403. [Google Scholar] [CrossRef] [Green Version]
  260. Apte, R.S.; Chen, D.S.; Ferrara, N. VEGF in Signaling and Disease: Beyond Discovery and Development. Cell 2019, 176, 1248–1264. [Google Scholar] [CrossRef] [Green Version]
  261. Miller, J.W.; Le Couter, J.; Strauss, E.C.; Ferrara, N. Vascular endothelial growth factor a in intraocular vascular disease. Ophthalmology 2013, 120, 106–114. [Google Scholar] [CrossRef] [PubMed]
  262. Fang, A.M.; Lee, A.Y.; Kulkarni, M.; Osborn, M.P.; Brantley, M.A., Jr. Polymorphisms in the VEGFA and VEGFR-2 genes and neovascular age-related macular degeneration. Mol. Vis. 2009, 15, 2710–2719. [Google Scholar] [PubMed]
  263. Vilkeviciute, A.; Cebatoriene, D.; Kriauciuniene, L.; Liutkeviciene, R. VEGFA Haplotype and VEGF-A and VEGF-R2 Protein Associations with Exudative Age-Related Macular Degeneration. Cells 2022, 11, 996. [Google Scholar] [CrossRef] [PubMed]
  264. Chen, H.; Mo, M.; Liu, G.Y.; Gong, Y.M.; Yu, K.D.; Xu, G.Z. Interaction of two functional genetic variants LOXL1 rs1048661 and VEGFA rs3025039 on the risk of age-related macular degeneration in Chinese women. Ann. Transl. Med. 2020, 8, 818. [Google Scholar] [CrossRef] [PubMed]
  265. Abedi, F.; Wickremasinghe, S.; Richardson, A.J.; Islam, A.F.; Guymer, R.H.; Baird, P.N. Genetic influences on the outcome of anti-vascular endothelial growth factor treatment in neovascular age-related macular degeneration. Ophthalmology 2013, 120, 1641–1648. [Google Scholar] [CrossRef]
  266. Hagstrom, S.A.; Ying, G.S.; Pauer, G.J.; Sturgill-Short, G.M.; Huang, J.; Maguire, M.G.; Martin, D.F.; Comparison of Age-Related Macular Degeneration Treatments Trials (CATT) Research Group. VEGFA and VEGFR2 gene polymorphisms and response to anti-vascular endothelial growth factor therapy: Comparison of age-related macular degeneration treatments trials (CATT). JAMA Ophthalmol. 2014, 132, 521–527. [Google Scholar] [CrossRef] [Green Version]
  267. Finger, R.P.; Wickremasinghe, S.S.; Baird, P.N.; Guymer, R.H. Predictors of anti-VEGF treatment response in neovascular age-related macular degeneration. Surv. Ophthalmol. 2014, 59, 1–18. [Google Scholar] [CrossRef]
  268. Kucukevcilioglu, M.; Patel, C.B.; Stone, E.M.; Russell, S.R. Clinically detectable drusen domains in fibulin-5-associated age-related macular degeneration (AMD): Drusen subdomains in fibulin-5 AMD. Int. Ophthalmol. 2016, 36, 569–575. [Google Scholar] [CrossRef]
  269. Mullins, R.F.; Olvera, M.A.; Clark, A.F.; Stone, E.M. Fibulin-5 distribution in human eyes: Relevance to age-related macular degeneration. Exp. Eye Res. 2007, 84, 378–380. [Google Scholar] [CrossRef] [Green Version]
  270. Albig, A.R.; Schiemann, W.P. Fibulin-5 antagonizes vascular endothelial growth factor (VEGF) signaling and angiogenic sprouting by endothelial cells. DNA Cell Biol. 2004, 23, 367–379. [Google Scholar] [CrossRef]
  271. Lotery, A.J.; Baas, D.; Ridley, C.; Jones, R.P.; Klaver, C.C.; Stone, E.; Nakamura, T.; Luff, A.; Griffiths, H.; Wang, T.; et al. Reduced secretion of fibulin 5 in age-related macular degeneration and cutis laxa. Hum. Mutat. 2006, 27, 568–574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Stone, E.M.; Braun, T.A.; Russell, S.R.; Kuehn, M.H.; Lotery, A.J.; Moore, P.A.; Eastman, C.G.; Casavant, T.L.; Sheffield, V.C. Missense variations in the fibulin 5 gene and age-related macular degeneration. N. Engl. J. Med. 2004, 351, 346–353. [Google Scholar] [CrossRef] [PubMed]
  273. Auer-Grumbach, M.; Weger, M.; Fink-Puches, R.; Papic, L.; Frohlich, E.; Auer-Grumbach, P.; El Shabrawi-Caelen, L.; Schabhuttl, M.; Windpassinger, C.; Senderek, J.; et al. Fibulin-5 mutations link inherited neuropathies, age-related macular degeneration and hyperelastic skin. Brain 2011, 134, 1839–1852. [Google Scholar] [CrossRef] [PubMed]
  274. Bhutto, I.; Lutty, G. Understanding age-related macular degeneration (AMD): Relationships between the photoreceptor/retinal pigment epithelium/Bruch’s membrane/choriocapillaris complex. Mol. Asp. Med. 2012, 33, 295–317. [Google Scholar] [CrossRef] [Green Version]
  275. Winkler, B.S.; Kapousta-Bruneau, N.; Arnold, M.J.; Green, D.G. Effects of inhibiting glutamine synthetase and blocking glutamate uptake on b-wave generation in the isolated rat retina. Vis. Neurosci. 1999, 16, 345–353. [Google Scholar] [CrossRef] [Green Version]
  276. Bonilla, B.; Hengel, S.R.; Grundy, M.K.; Bernstein, K.A. RAD51 Gene Family Structure and Function. Annu. Rev. Genet. 2020, 54, 25–46. [Google Scholar] [CrossRef]
  277. Zhou, J.; Wang, D.; Zhang, J.; Zhang, M.; Lu, F.; Qiu, G.; Zhao, L.; Nguyen, D.H.; Luo, H.; Cao, G.; et al. RAD51 gene is associated with advanced age-related macular degeneration in Chinese population. Clin. Biochem. 2013, 46, 1689–1693. [Google Scholar] [CrossRef]
  278. Chu, X.K.; Meyerle, C.B.; Liang, X.; Chew, E.Y.; Chan, C.C.; Tuo, J. In-depth analyses unveil the association and possible functional involvement of novel RAD51B polymorphisms in age-related macular degeneration. Age 2014, 36, 9627. [Google Scholar] [CrossRef] [Green Version]
  279. Battu, P.; Sharma, K.; Rain, M.; Singh, R.; Anand, A. Serum Levels of ARMS2, COL8A1, RAD51B, and VEGF and their Correlations in Age-related Macular Degeneration. Curr. Neurovasc. Res. 2021, 18, 181–188. [Google Scholar] [CrossRef]
  280. Sartorius, U.; Schmitz, I.; Krammer, P.H. Molecular mechanisms of death-receptor-mediated apoptosis. Chembiochem 2001, 2, 20–29. [Google Scholar] [CrossRef]
  281. Mori, K.; Ishikawa, K.; Fukuda, Y.; Ji, R.; Wada, I.; Kubo, Y.; Akiyama, M.; Notomi, S.; Murakami, Y.; Nakao, S.; et al. TNFRSF10A downregulation induces retinal pigment epithelium degeneration during the pathogenesis of age-related macular degeneration and central serous chorioretinopathy. Hum. Mol. Genet. 2022, ddac020. [Google Scholar] [CrossRef] [PubMed]
  282. Arakawa, S.; Takahashi, A.; Ashikawa, K.; Hosono, N.; Aoi, T.; Yasuda, M.; Oshima, Y.; Yoshida, S.; Enaida, H.; Tsuchihashi, T.; et al. Genome-wide association study identifies two susceptibility loci for exudative age-related macular degeneration in the Japanese population. Nat. Genet. 2011, 43, 1001–1004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Sun, Y.; Li, S.; Li, H.; Yang, F.; Bai, Y.; Zhao, M.; Guo, J.; Zhao, M.; Zhou, P.; Khor, C.C.; et al. TNFRSF10A-LOC389641 rs13278062 but not REST-C4orf14-POLR2B-IGFBP7 rs1713985 was found associated with age-related macular degeneration in a Chinese population. Investig. Ophthalmol. Vis. Sci. 2013, 54, 8199–8203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Tuo, J.; Ning, B.; Bojanowski, C.M.; Lin, Z.N.; Ross, R.J.; Reed, G.F.; Shen, D.; Jiao, X.; Zhou, M.; Chew, E.Y.; et al. Synergic effect of polymorphisms in ERCC6 5’ flanking region and complement factor H on age-related macular degeneration predisposition. Proc. Natl. Acad. Sci. USA 2006, 103, 9256–9261. [Google Scholar] [CrossRef] [Green Version]
  285. Troelstra, C.; van Gool, A.; de Wit, J.; Vermeulen, W.; Bootsma, D.; Hoeijmakers, J.H. ERCC6, a member of a subfamily of putative helicases, is involved in Cockayne’s syndrome and preferential repair of active genes. Cell 1992, 71, 939–953. [Google Scholar] [CrossRef] [Green Version]
  286. Baas, D.C.; Despriet, D.D.; Gorgels, T.G.; Bergeron-Sawitzke, J.; Uitterlinden, A.G.; Hofman, A.; van Duijn, C.M.; Merriam, J.E.; Smith, R.T.; Barile, G.R.; et al. The ERCC6 gene and age-related macular degeneration. PLoS ONE 2010, 5, e13786. [Google Scholar] [CrossRef] [Green Version]
  287. Blasiak, J.; Synowiec, E.; Salminen, A.; Kaarniranta, K. Genetic variability in DNA repair proteins in age-related macular degeneration. Int. J. Mol. Sci. 2012, 13, 13378–13397. [Google Scholar] [CrossRef]
  288. Liu, M.M.; Agron, E.; Chew, E.; Meyerle, C.; Ferris, F.L., 3rd; Chan, C.C.; Tuo, J. Copy number variations in candidate genes in neovascular age-related macular degeneration. Investig. Ophthalmol. Vis. Sci. 2011, 52, 3129–3135. [Google Scholar] [CrossRef]
  289. Weeks, D.E.; Conley, Y.P.; Tsai, H.J.; Mah, T.S.; Rosenfeld, P.J.; Paul, T.O.; Eller, A.W.; Morse, L.S.; Dailey, J.P.; Ferrell, R.E.; et al. Age-related maculopathy: An expanded genome-wide scan with evidence of susceptibility loci within the 1q31 and 17q25 regions. Am. J. Ophthalmol. 2001, 132, 682–692. [Google Scholar] [CrossRef]
  290. Weeks, D.E.; Conley, Y.P.; Tsai, H.J.; Mah, T.S.; Schmidt, S.; Postel, E.A.; Agarwal, A.; Haines, J.L.; Pericak-Vance, M.A.; Rosenfeld, P.J.; et al. Age-related maculopathy: A genomewide scan with continued evidence of susceptibility loci within the 1q31, 10q26, and 17q25 regions. Am. J. Hum. Genet. 2004, 75, 174–189. [Google Scholar] [CrossRef] [Green Version]
  291. Iyengar, S.K.; Song, D.; Klein, B.E.; Klein, R.; Schick, J.H.; Humphrey, J.; Millard, C.; Liptak, R.; Russo, K.; Jun, G.; et al. Dissection of genomewide-scan data in extended families reveals a major locus and oligogenic susceptibility for age-related macular degeneration. Am. J. Hum. Genet. 2004, 74, 20–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Kenealy, S.J.; Schmidt, S.; Agarwal, A.; Postel, E.A.; De La Paz, M.A.; Pericak-Vance, M.A.; Haines, J.L. Linkage analysis for age-related macular degeneration supports a gene on chromosome 10q26. Mol. Vis. 2004, 10, 57–61. [Google Scholar] [CrossRef]
  293. Fisher, S.A.; Abecasis, G.R.; Yashar, B.M.; Zareparsi, S.; Swaroop, A.; Iyengar, S.K.; Klein, B.E.; Klein, R.; Lee, K.E.; Majewski, J.; et al. Meta-analysis of genome scans of age-related macular degeneration. Hum. Mol. Genet. 2005, 14, 2257–2264. [Google Scholar] [CrossRef] [Green Version]
  294. 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]
  295. Marmorstein, A.D.; Marmorstein, L.Y. The challenge of modeling macular degeneration in mice. Trends Genet. 2007, 23, 225–231. [Google Scholar] [CrossRef] [PubMed]
  296. Klein, R.; Myers, C.E.; Meuer, S.M.; Gangnon, R.E.; Sivakumaran, T.A.; Iyengar, S.K.; Lee, K.E.; Klein, B.E. Risk alleles in CFH and ARMS2 and the long-term natural history of age-related macular degeneration: The Beaver Dam Eye Study. JAMA Ophthalmol. 2013, 131, 383–392. [Google Scholar] [CrossRef] [Green Version]
  297. Wyatt, M.K.; Tsai, J.Y.; Mishra, S.; Campos, M.; Jaworski, C.; Fariss, R.N.; Bernstein, S.L.; Wistow, G. Interaction of complement factor h and fibulin3 in age-related macular degeneration. PLoS ONE 2013, 8, e68088. [Google Scholar] [CrossRef]
  298. Jakobsdottir, J.; Conley, Y.P.; Weeks, D.E.; Mah, T.S.; Ferrell, R.E.; Gorin, M.B. Susceptibility genes for age-related maculopathy on chromosome 10q26. Am. J. Hum. Genet. 2005, 77, 389–407. [Google Scholar] [CrossRef] [Green Version]
  299. Rivera, A.; Fisher, S.A.; Fritsche, L.G.; Keilhauer, C.N.; Lichtner, P.; Meitinger, T.; Weber, B.H. Hypothetical LOC387715 is a second major susceptibility gene for age-related macular degeneration, contributing independently of complement factor H to disease risk. Hum. Mol. Genet. 2005, 14, 3227–3236. [Google Scholar] [CrossRef] [Green Version]
  300. Dewan, A.; Liu, M.; Hartman, S.; Zhang, S.S.; Liu, D.T.; Zhao, C.; Tam, P.O.; Chan, W.M.; Lam, D.S.; Snyder, M.; et al. HTRA1 promoter polymorphism in wet age-related macular degeneration. Science 2006, 314, 989–992. [Google Scholar] [CrossRef]
  301. Conley, Y.P.; Jakobsdottir, J.; Mah, T.; Weeks, D.E.; Klein, R.; Kuller, L.; Ferrell, R.E.; Gorin, M.B. CFH, ELOVL4, PLEKHA1 and LOC387715 genes and susceptibility to age-related maculopathy: AREDS and CHS cohorts and meta-analyses. Hum. Mol. Genet. 2006, 15, 3206–3218. [Google Scholar] [CrossRef] [PubMed]
  302. Kanda, A.; Chen, W.; Othman, M.; Branham, K.E.; Brooks, M.; Khanna, R.; He, S.; Lyons, R.; Abecasis, G.R.; Swaroop, A. A variant of mitochondrial protein LOC387715/ARMS2, not HTRA1, is strongly associated with age-related macular degeneration. Proc. Natl. Acad. Sci. USA 2007, 104, 16227–16232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Francis, P.J.; Klein, M.L. Update on the role of genetics in the onset of age-related macular degeneration. Clin. Ophthalmol. 2011, 5, 1127–1133. [Google Scholar] [CrossRef] [Green Version]
  304. Hautamaki, A.; Seitsonen, S.; Holopainen, J.M.; Moilanen, J.A.; Kivioja, J.; Onkamo, P.; Jarvela, I.; Immonen, I. The genetic variant rs4073 AT of the Interleukin-8 promoter region is associated with the earlier onset of exudative age-related macular degeneration. Acta Ophthalmol. 2015, 93, 726–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Woo, S.J.; Ahn, J.; Morrison, M.A.; Ahn, S.Y.; Lee, J.; Kim, K.W.; DeAngelis, M.M.; Park, K.H. Analysis of Genetic and Environmental Risk Factors and Their Interactions in Korean Patients with Age-Related Macular Degeneration. PLoS ONE 2015, 10, e0132771. [Google Scholar] [CrossRef] [Green Version]
  306. Lu, Z.G.; May, A.; Dinh, B.; Lin, V.; Su, F.; Tran, C.; Adivikolanu, H.; Ehlen, R.; Che, B.; Wang, Z.H.; et al. The interplay of oxidative stress and ARMS2-HTRA1 genetic risk in neovascular AMD. Vessel. Plus 2021, 5, 4. [Google Scholar] [CrossRef]
  307. Smailhodzic, D.; Klaver, C.C.; Klevering, B.J.; Boon, C.J.; Groenewoud, J.M.; Kirchhof, B.; Daha, M.R.; den Hollander, A.I.; Hoyng, C.B. Risk alleles in CFH and ARMS2 are independently associated with systemic complement activation in age-related macular degeneration. Ophthalmology 2012, 119, 339–346. [Google Scholar] [CrossRef]
  308. Fritsche, L.G.; Loenhardt, T.; Janssen, A.; Fisher, S.A.; Rivera, A.; Keilhauer, C.N.; Weber, B.H. Age-related macular degeneration is associated with an unstable ARMS2 (LOC387715) mRNA. Nat. Genet. 2008, 40, 892–896. [Google Scholar] [CrossRef]
  309. Wang, G.; Spencer, K.L.; Court, B.L.; Olson, L.M.; Scott, W.K.; Haines, J.L.; Pericak-Vance, M.A. Localization of age-related macular degeneration-associated ARMS2 in cytosol, not mitochondria. Investig. Ophthalmol. Vis. Sci. 2009, 50, 3084–3090. [Google Scholar] [CrossRef]
  310. Kortvely, E.; Hauck, S.M.; Duetsch, G.; Gloeckner, C.J.; Kremmer, E.; Alge-Priglinger, C.S.; Deeg, C.A.; Ueffing, M. ARMS2 is a constituent of the extracellular matrix providing a link between familial and sporadic age-related macular degenerations. Investig. Ophthalmol. Vis. Sci. 2010, 51, 79–88. [Google Scholar] [CrossRef] [Green Version]
  311. Micklisch, S.; Lin, Y.; Jacob, S.; Karlstetter, M.; Dannhausen, K.; Dasari, P.; von der Heide, M.; Dahse, H.M.; Schmolz, L.; Grassmann, F.; et al. Age-related macular degeneration associated polymorphism rs10490924 in ARMS2 results in deficiency of a complement activator. J. Neuroinflam. 2017, 14, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. Canfield, A.E.; Hadfield, K.D.; Rock, C.F.; Wylie, E.C.; Wilkinson, F.L. HtrA1: A novel regulator of physiological and pathological matrix mineralization? Biochem. Soc. Trans. 2007, 35, 669–671. [Google Scholar] [CrossRef] [PubMed]
  313. Beguier, F.; Housset, M.; Roubeix, C.; Augustin, S.; Zagar, Y.; Nous, C.; Mathis, T.; Eandi, C.; Benchaboune, M.; Drame-Maigne, A.; et al. The 10q26 Risk Haplotype of Age-Related Macular Degeneration Aggravates Subretinal Inflammation by Impairing Monocyte Elimination. Immunity 2020, 53, 429–441.e428. [Google Scholar] [CrossRef] [PubMed]
  314. Williams, B.L.; Seager, N.A.; Gardiner, J.D.; Pappas, C.M.; Cronin, M.C.; Amat di San Filippo, C.; Anstadt, R.A.; Liu, J.; Toso, M.A.; Nichols, L.; et al. Chromosome 10q26-driven age-related macular degeneration is associated with reduced levels of HTRA1 in human retinal pigment epithelium. Proc. Natl. Acad. Sci. USA 2021, 118, e2103617118. [Google Scholar] [CrossRef] [PubMed]
  315. Ross, R.J.; Verma, V.; Rosenberg, K.I.; Chan, C.C.; Tuo, J. Genetic markers and biomarkers for age-related macular degeneration. Expert Rev. Ophthalmol. 2007, 2, 443–457. [Google Scholar] [CrossRef] [Green Version]
  316. Bergeron-Sawitzke, J.; Gold, B.; Olsh, A.; Schlotterbeck, S.; Lemon, K.; Visvanathan, K.; Allikmets, R.; Dean, M. Multilocus analysis of age-related macular degeneration. Eur. J. Hum. Genet. 2009, 17, 1190–1199. [Google Scholar] [CrossRef]
  317. Mohamad, N.A.; Ramachandran, V.; Mohd Isa, H.; Chan, Y.M.; Ngah, N.F.; Ching, S.M.; Hoo, F.K.; Wan Sulaiman, W.A.; Inche Mat, L.N.; Mohamed, M.H. Association of HTRA1 and ARMS2 gene polymorphisms with response to intravitreal ranibizumab among neovascular age-related macular degenerative subjects. Hum. Genom. 2019, 13, 13. [Google Scholar] [CrossRef]
  318. Shijo, T.; Sakurada, Y.; Fukuda, Y.; Yoneyama, S.; Sugiyama, A.; Matsubara, M.; Kikushima, W.; Tanabe, N.; Parikh, R.; Kashiwagi, K. Association of CRP levels with ARMS2 and CFH variants in age-related macular degeneration. Int. Ophthalmol. 2020, 40, 2735–2742. [Google Scholar] [CrossRef]
  319. Kitchens, J.W.; Kassem, N.; Wood, W.; Stone, T.W.; Isernhagen, R.; Wood, E.; Hancock, B.A.; Radovich, M.; Waymire, J.; Li, L.; et al. A pharmacogenetics study to predict outcome in patients receiving anti-VEGF therapy in age related macular degeneration. Clin. Ophthalmol. 2013, 7, 1987–1993. [Google Scholar] [CrossRef] [Green Version]
  320. Nakai, S.; Matsumiya, W.; Miki, A.; Honda, S.; Nakamura, M. Association of an age-related maculopathy susceptibility 2 gene variant with the 12-month outcomes of intravitreal aflibercept combined with photodynamic therapy for polypoidal choroidal vasculopathy. Jpn J. Ophthalmol. 2019, 63, 389–395. [Google Scholar] [CrossRef]
  321. Zhang, J.; Liu, Z.; Hu, S.; Qi, J. Meta-Analysis of the Pharmacogenetics of ARMS2 A69S Polymorphism and the Response to Advanced Age-Related Macular Degeneration. Ophthalmic Res. 2021, 64, 192–204. [Google Scholar] [CrossRef]
  322. Fang, K.; Gao, P.; Tian, J.; Qin, X.; Yu, W.; Li, J.; Chen, Q.; Huang, L.; Chen, D.; Hu, Y.; et al. Joint Effect of CFH and ARMS2/HTRA1 Polymorphisms on Neovascular Age-Related Macular Degeneration in Chinese Population. J. Ophthalmol. 2015, 2015, 821918. [Google Scholar] [CrossRef] [PubMed]
  323. Kawasaki, R.; Yasuda, M.; Song, S.J.; Chen, S.J.; Jonas, J.B.; Wang, J.J.; Mitchell, P.; Wong, T.Y. The prevalence of age-related macular degeneration in Asians: A systematic review and meta-analysis. Ophthalmology 2010, 117, 921–927. [Google Scholar] [CrossRef] [PubMed]
  324. Yasuma, T.; Nakamura, M.; Kikuchi, M.; Ishikawa, K.; Nishihara, H.; Kaneko, H.; Nishiguchi, K.M.; Terasaki, H. Association of c.*372_815del443ins54 Polymorphism in the ARMS2 (LOC387715) Gene With Neovascular Age-Related Macular Degeneration (AMD) and Polypoidal Choroidal Vasculopathy (PCV) in a Japanese Population. Investig. Ophthalmol. Vis. Sci. 2009, 50, 3436. [Google Scholar]
  325. Akagi-Kurashige, Y.; Yamashiro, K.; Gotoh, N.; Miyake, M.; Morooka, S.; Yoshikawa, M.; Nakata, I.; Kumagai, K.; Tsujikawa, A.; Yamada, R.; et al. MMP20 and ARMS2/HTRA1 Are Associated with Neovascular Lesion Size in Age-Related Macular Degeneration. Ophthalmology 2015, 122, 2295–2302.e2292. [Google Scholar] [CrossRef]
  326. Yang, Z.; Tong, Z.; Chen, Y.; Zeng, J.; Lu, F.; Sun, X.; Zhao, C.; Wang, K.; Davey, L.; Chen, H.; et al. Genetic and functional dissection of HTRA1 and LOC387715 in age-related macular degeneration. PLoS Genet. 2010, 6, e1000836. [Google Scholar] [CrossRef]
  327. Andreoli, M.T.; Morrison, M.A.; Kim, B.J.; Chen, L.; Adams, S.M.; Miller, J.W.; DeAngelis, M.M.; Kim, I.K. Comprehensive analysis of complement factor H and LOC387715/ARMS2/HTRA1 variants with respect to phenotype in advanced age-related macular degeneration. Am. J. Ophthalmol. 2009, 148, 869–874. [Google Scholar] [CrossRef] [Green Version]
  328. Deangelis, M.M.; Ji, F.; Adams, S.; Morrison, M.A.; Harring, A.J.; Sweeney, M.O.; Capone, A., Jr.; Miller, J.W.; Dryja, T.P.; Ott, J.; et al. Alleles in the HtrA serine peptidase 1 gene alter the risk of neovascular age-related macular degeneration. Ophthalmology 2008, 115, 1209–1215.e1207. [Google Scholar] [CrossRef] [Green Version]
  329. Tam, P.O.; Ng, T.K.; Liu, D.T.; Chan, W.M.; Chiang, S.W.; Chen, L.J.; DeWan, A.; Hoh, J.; Lam, D.S.; Pang, C.P. HTRA1 variants in exudative age-related macular degeneration and interactions with smoking and CFH. Investig. Ophthalmol. Vis. Sci. 2008, 49, 2357–2365. [Google Scholar] [CrossRef] [Green Version]
  330. Gerhardt, M.J.; Marsh, J.A.; Morrison, M.; Kazlauskas, A.; Khadka, A.; Rosenkranz, S.; DeAngelis, M.M.; Saint-Geniez, M.; Jacobo, S.M.P. ER stress-induced aggresome trafficking of HtrA1 protects against proteotoxicity. J. Mol. Cell Biol. 2017, 9, 516–532. [Google Scholar] [CrossRef]
  331. Jacobo, S.M.; Deangelis, M.M.; Kim, I.K.; Kazlauskas, A. Age-related macular degeneration-associated silent polymorphisms in HtrA1 impair its ability to antagonize insulin-like growth factor 1. Mol. Cell Biol. 2013, 33, 1976–1990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Kanda, H.; Nakano, Y.; Terasawa, Y.; Morimoto, T.; Fujikado, T. The relationship between retinal damage and current intensity in a pre-clinical suprachoroidal-transretinal stimulation model using a laser-formed microporous electrode. J. Neural. Eng. 2017, 14, 056013. [Google Scholar] [CrossRef] [Green Version]
  333. Wang, W.; Dean, D.C.; Kaplan, H.J. Age-related macular degeneration. Discov. Med. 2010, 9, 13–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  334. Wang, G.; Dubovy, S.R.; Kovach, J.L.; Schwartz, S.G.; Agarwal, A.; Scott, W.K.; Haines, J.L.; Pericak-Vance, M.A. Variants at chromosome 10q26 locus and the expression of HTRA1 in the retina. Exp. Eye Res. 2013, 112, 102–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Friedrich, U.; Myers, C.A.; Fritsche, L.G.; Milenkovich, A.; Wolf, A.; Corbo, J.C.; Weber, B.H.F. Risk- and non-risk-associated variants at the 10q26 AMD locus influence ARMS2 mRNA expression but exclude pathogenic effects due to protein deficiency. Hum. Mol. Genet. 2011, 20, 1387–1399. [Google Scholar] [CrossRef]
  336. Ajana, S.; Cougnard-Gregoire, A.; Colijn, J.M.; Merle, B.M.J.; Verzijden, T.; de Jong, P.; Hofman, A.; Vingerling, J.R.; Hejblum, B.P.; Korobelnik, J.F.; et al. Predicting Progression to Advanced Age-Related Macular Degeneration from Clinical, Genetic, and Lifestyle Factors Using Machine Learning. Ophthalmology 2021, 128, 587–597. [Google Scholar] [CrossRef]
  337. Csaky, K.G.; Schachat, A.P.; Kaiser, P.K.; Small, K.W.; Heier, J.S. The Use of Genetic Testing in the Management of Patients With Age-Related Macular Degeneration:American Society of Retina Specialists Genetics Task Force Special Report. J. Vitr. Dis. 2017, 1, 75–78. [Google Scholar] [CrossRef] [Green Version]
  338. Cabral de Guimaraes, T.A.; Daich Varela, M.; Georgiou, M.; Michaelides, M. Treatments for dry age-related macular degeneration: Therapeutic avenues, clinical trials and future directions. Br. J. Ophthalmol. 2022, 106, 297–304. [Google Scholar] [CrossRef]
Figure 1. (A) Color fundus photo of macular drusen in the right eye of a 70-year-old male. (B) Optical coherence tomography image of macular drusen demonstrating sub-retinal pigment epithelium deposits. [Adapted with permission from Ipoliker, CC BY-SA 3.0 https://creativecommons.org/licenses/by-sa/3.0, accessed on 15 May 2022, via Wikimedia Commons].
Figure 1. (A) Color fundus photo of macular drusen in the right eye of a 70-year-old male. (B) Optical coherence tomography image of macular drusen demonstrating sub-retinal pigment epithelium deposits. [Adapted with permission from Ipoliker, CC BY-SA 3.0 https://creativecommons.org/licenses/by-sa/3.0, accessed on 15 May 2022, via Wikimedia Commons].
Genes 13 01233 g001
Figure 2. Overview of the complement system. The complement system may be activated by three distinct biomolecular pathways: classical, lectin, and alternative pathways. These three pathways converge on the cleavage of C3 to C3b, which in turn acts to cleave C5 to C5b. C5b subsequently combines with C6-C9 to form the membrane attack complex (MAC), which forms pores in the target cell membranes, resulting in lysis of the target cells. Image adapted with permission from Danobeitia et al., 2014 [65].
Figure 2. Overview of the complement system. The complement system may be activated by three distinct biomolecular pathways: classical, lectin, and alternative pathways. These three pathways converge on the cleavage of C3 to C3b, which in turn acts to cleave C5 to C5b. C5b subsequently combines with C6-C9 to form the membrane attack complex (MAC), which forms pores in the target cell membranes, resulting in lysis of the target cells. Image adapted with permission from Danobeitia et al., 2014 [65].
Genes 13 01233 g002
Figure 3. Structure of a lipoprotein. Lipoproteins are biochemical structures that function in the transport of water-insoluble lipids in the bloodstream. They consist of a hydrophilic outer shell (surface coat) composed of apolipoproteins, phospholipids, and unesterified cholesterol, and a lipid core consisting of cholesteryl esters and triglycerides. (Image adapted with permission from: AntiSense, CC BY-SA 3.0, https://creativecommons.org/licenses/by-sa/3.0, accessed on 15 May 2022, via Wikimedia Commons).
Figure 3. Structure of a lipoprotein. Lipoproteins are biochemical structures that function in the transport of water-insoluble lipids in the bloodstream. They consist of a hydrophilic outer shell (surface coat) composed of apolipoproteins, phospholipids, and unesterified cholesterol, and a lipid core consisting of cholesteryl esters and triglycerides. (Image adapted with permission from: AntiSense, CC BY-SA 3.0, https://creativecommons.org/licenses/by-sa/3.0, accessed on 15 May 2022, via Wikimedia Commons).
Genes 13 01233 g003
Figure 4. Lipid transport and metabolism in the human body. Apolipoproteins (e.g., A1, A2, B-48, B-100, C2, D, and E) bind water-insoluble lipids to form water-soluble lipoproteins such as chylomicrons, very low-density lipoproteins (VLDL), low-density lipoprotein (LDL), and high-density lipoprotein (HDL) to allow the transport of lipids (e.g., cholesterol, triglycerides) through the bloodstream. The process of lipoprotein formation is mediated by enzymes such as lecithin-cholesterol acyltransferase (LCAT). Upon arrival in their target tissues, lipids carried by lipoproteins are broken down by enzymes such as lipoprotein lipase (LPL) in peripheral tissue and hepatic lipase (LIPC) in the liver. (Image adapted with permission from: Vtosha, CC BY-SA 3.0, https://creativecommons.org/licenses/by-sa/3.0, accessed on 15 May 2022, via Wikimedia Commons).
Figure 4. Lipid transport and metabolism in the human body. Apolipoproteins (e.g., A1, A2, B-48, B-100, C2, D, and E) bind water-insoluble lipids to form water-soluble lipoproteins such as chylomicrons, very low-density lipoproteins (VLDL), low-density lipoprotein (LDL), and high-density lipoprotein (HDL) to allow the transport of lipids (e.g., cholesterol, triglycerides) through the bloodstream. The process of lipoprotein formation is mediated by enzymes such as lecithin-cholesterol acyltransferase (LCAT). Upon arrival in their target tissues, lipids carried by lipoproteins are broken down by enzymes such as lipoprotein lipase (LPL) in peripheral tissue and hepatic lipase (LIPC) in the liver. (Image adapted with permission from: Vtosha, CC BY-SA 3.0, https://creativecommons.org/licenses/by-sa/3.0, accessed on 15 May 2022, via Wikimedia Commons).
Genes 13 01233 g004
Figure 5. Vascular endothelial growth factor (VEGF) isoforms and their receptors (VEGFR). (Image adapted with permission from: Mjeltsch, CC BY-SA 4.0, https://creativecommons.org/licenses/by-sa/4.0, accessed on 15 May 2022, via Wikimedia Commons).
Figure 5. Vascular endothelial growth factor (VEGF) isoforms and their receptors (VEGFR). (Image adapted with permission from: Mjeltsch, CC BY-SA 4.0, https://creativecommons.org/licenses/by-sa/4.0, accessed on 15 May 2022, via Wikimedia Commons).
Genes 13 01233 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shughoury, A.; Sevgi, D.D.; Ciulla, T.A. Molecular Genetic Mechanisms in Age-Related Macular Degeneration. Genes 2022, 13, 1233. https://doi.org/10.3390/genes13071233

AMA Style

Shughoury A, Sevgi DD, Ciulla TA. Molecular Genetic Mechanisms in Age-Related Macular Degeneration. Genes. 2022; 13(7):1233. https://doi.org/10.3390/genes13071233

Chicago/Turabian Style

Shughoury, Aumer, Duriye Damla Sevgi, and Thomas A. Ciulla. 2022. "Molecular Genetic Mechanisms in Age-Related Macular Degeneration" Genes 13, no. 7: 1233. https://doi.org/10.3390/genes13071233

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