Next Article in Journal
Two-Time Multiplexed Targeted Next-Generation Sequencing Might Help the Implementation of Germline Screening Tools for Myelodysplastic Syndromes/Hematologic Neoplasms
Next Article in Special Issue
Skin Bioimpedance Analysis to Determine Cellular Integrity by Phase Angle in Women with Fibromyalgia: A Cross-Sectional Study
Previous Article in Journal
Derazantinib Inhibits the Bioactivity of Keloid Fibroblasts via FGFR Signaling
Previous Article in Special Issue
Impact of CPAP Therapy on the Autonomic Nervous System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Developments in Gene Therapy for Neovascular Age-Related Macular Degeneration: A Review

by
Lucia Finocchio
1,†,
Marco Zeppieri
1,*,†,
Andrea Gabai
1,
Giacomo Toneatto
1,
Leopoldo Spadea
2 and
Carlo Salati
1
1
Department of Ophthalmology, University Hospital of Udine, 33100 Udine, Italy
2
Eye Clinic, Policlinico Umberto I, “Sapienza” University of Rome, 00142 Rome, Italy
*
Author to whom correspondence should be addressed.
Shared first authorship.
Biomedicines 2023, 11(12), 3221; https://doi.org/10.3390/biomedicines11123221
Submission received: 1 November 2023 / Revised: 28 November 2023 / Accepted: 1 December 2023 / Published: 5 December 2023
(This article belongs to the Special Issue 10th Anniversary of Biomedicines—Ophthalmology Disorders)

Abstract

:
Age-related macular degeneration (AMD) is a complex and multifactorial disease and a leading cause of irreversible blindness in the elderly population. The anti-vascular endothelial growth factor (anti-VEGF) therapy has revolutionized the management and prognosis of neovascular AMD (nAMD) and is currently the standard of care for this disease. However, patients are required to receive repeated injections, imposing substantial social and economic burdens. The implementation of gene therapy methods to achieve sustained delivery of various therapeutic proteins holds the promise of a single treatment that could ameliorate the treatment challenges associated with chronic intravitreal therapy, and potentially improve visual outcomes. Several early-phase trials are currently underway, evaluating the safety and efficacy of gene therapy for nAMD; however, areas of controversy persist, including the therapeutic target, route of administration, and potential safety issues. In this review, we assess the evolution of gene therapy for nAMD and summarize several preclinical and early-stage clinical trials, exploring challenges and future directions.

1. Introduction

1.1. Definition of Age-Related Macular Degeneration

Age-related macular degeneration (AMD) is a chronic inflammatory eye disease that involves the macular region with a strong hereditary component, typically affecting people over 60 years of age. Pathologic changes involve the deeper retinal layers of the macula and surrounding vasculature, resulting in central vision loss. The disease has a prevalence of 8.7% worldwide [1] and is the most common cause of severe visual impairment in developed countries [2,3,4,5,6]. Its prevalence is likely expected to rise due to the exponential aging of the population. Principal risk factors for developing AMD include age, smoking history, hyperlipidemia, family history, and ethnicity [7]. There are two main types of AMD: non-neovascular and neovascular. Non-neovascular AMD (“dry” AMD) accounts for almost 80–85% of all cases and is usually related to a more favorable visual prognosis. The accumulation of retinal deposits, called drusen, is a distinctive clinical finding and may be the first sign of the “dry” form of the disease. Retinal pigment epithelial (RPE) changes, pigment clumping, or autofluorescence abnormalities may also be early clinical signs [8]. The type and quantity of drusen define the early, intermediate, and late stages of initial dry AMD but are not necessarily associated with vision impairment [9]. Progression to geographic atrophy (GA) and neovascular AMD (“wet” AMD or nAMD) is the key reason for severe vision loss as a result of AMD [10]. Neovascular AMD affects the remaining 15–20% of eyes with AMD and is characterized by the formation of macular neovascularization (MNV). These new blood vessels may cause an accumulation of subretinal, intraretinal, and sub-RPE fluid and bleeding, with metamorphopsia and central scotoma, respectively. Vascular endothelial growth factor (VEGF) is the main factor responsible for this abnormal vascular proliferation.

1.2. Pathophysiology of nAMD

Although there has been extensive research conducted to understand AMD pathogenesis, it remains not completely clear due to its multifactorial character. AMD is the consequence of the interaction between metabolism, genetics, and the environment.
The RPE promotes a vascular background along its basal surface and an avascular background along its apical surface. This creates an essential environment to maintain retinal photoreceptor cells in healthy conditions. Vascular endothelial cells require survival factors such as vascular endothelial growth factor (VEGF), basic fibroblast growth factor (bFGF), and angiopoietin-1 [11] that can originate from the extracellular matrix, surrounding cells, or plasma; VEGF represents an important mediator of paracrine and endocrine trophic support [12,13]. With aging, several changes happen in the RPE/Bruch’s membrane complex, modifying its capacity to remove residual substances such as lipofuscin [14], including the thickening of Bruch’s membrane, reduction of metabolic activity, loss of mitochondria and reduced choroidal blood supply [15]. Moreover, aging increases the risk of retinal and choroidal hypoxia. With age, the accumulation of lipofuscin, reactive oxygen species (ROS) and other factors cause a thickening of Bruch’s membrane, that not only decreases the removal of debris by the choriocapillaris but also acts as a barrier to the diffusion of oxygen and nutrients from the choroid to the photoreceptors and the RPE [16,17]. Hypoxia in turn causes an upregulation of a heterodimer made up of HIF-1α and HIF-1β, which is the Hypoxia-inducible factor-1 (HIF-1), [18]. HIF-1 regulates the transcription of genes of VEGF and its receptor VEGFR, platelet-derived growth factor-B (PDGF-B) and its receptor PDGFRβ, angiopoietin-2 (Ang2), stromal-derived factor-1 (SDF-1) and its receptor CXCR4, and vascular endothelial-protein tyrosine phosphatase (VE-PTP) [19]. HIF-1, VEGF, and VEGFR family are the principal intermediaries of angiogenesis regulation [20]. VEGF is the most studied factor in the context of ocular neovascularization. Encoded by the VEGF gene, this glycoprotein family, including VEGF-A, VEGF-B, VEGF-C, VEGF-D, VEGF-E, and VEGF-F and placental growth factor (PLGF), primarily activates cellular signaling pathways to facilitate the development of new blood vessels, either de novo or from existing ones. VEGF-A plays a crucial role in vascular proliferation and the migration of endothelial cells for both physiological and pathological angiogenesis [21,22]. However, VEGF expression by RPE cells seems to not be sufficient to cause MNV. In addition to secreted factors, which are mainly governed by HIF-1, signals contributing to MNV formation also encompass cues from the extracellular matrix and neighboring cells [12].
Inflammation is believed to play a central role in the pathogenesis of both dry and wet AMD. In the literature, there is strong evidence to suggest that an abnormal complement activation is significantly involved in the pathogenesis of the disease.
Even though complement activation products were detected in drusen nearly thirty years ago [23,24], it was not until the early 21st century that a major heritable determinant of AMD was identified as a single nucleotide polymorphism in the complement factor H (CFH) gene [25,26,27,28,29]. This discovery motivated genotyping initiatives in AMD, leading to the identification of further risk variants in other complement gene loci, such as C3, C2, CFB, C9, CFI, and CFHR4 [26,27,28,29,30,31,32,33,34]. C3a and C5a have been found in soft drusen and induce the upregulation of VEGF in RPE, increasing the risk of MNV associated with soft drusen [35].
The membrane attack complex (C5b-9, MAC) is also found in drusen and in impaired RPE cells of AMD-affected eyes [36]. Various factors are suggested to trigger complement activation in the retina, including oxidative stress and the buildup of pro-inflammatory derivatives from the visual cycle, such as lipofuscin components, apolipoproteins, and amyloid beta [37,38]. These elements are believed to interact with and amplify other established pathological mechanisms, like RPE dysfunction stemming from choroidal vascular insufficiency. Oxidative stress has the potential to disrupt the regulation of the complement system by RPE cells. This disruption includes a decrease in the surface expression of complement inhibitors, decay accelerating factor (CD55), and CD59, as well an impairment of complement regulation at the cell surface by CFH [39]. Furthermore, oxidative stress may impede interferon-γ from effectively enhancing CFH expression in RPE cells. In vitro, products resulting from the photo-oxidation of N-retinylidene-N-retinylethanolamine (A2E) in RPE can activate the complement system [40]. Choroidal dendritic cells and retinal microglia cells, both fundamental to retinal structure and metabolism, are “activated and recruited” by locally injured and/or sublethal damaged RPE cells, related to RPE blebs, fragments, and debris. They can maintain and enhance the local inflammation, not only activating the complement, but also forming an immune complex and recruiting choroidal T-cells or phagocytic cells, collectively contributing to the development of AMD [41,42]. Under the stimulation of inflammatory mediators, RPE cells produce cytokines and chemokines, including IL-4, -5, -6, -8, -10, -13, -17, IFN-β, IFN-γ, TGF-β, MCP-1, and VEGF. Inflammatory cytokines can also enhance the secretion of VEGF [43,44,45].
As discussed above, genetics plays a role in the pathogenesis of nAMD. Genetic variants implicated in nAMD encompass CFH, CFH-related genes, complement proteins C3 and C9, the age-related maculopathy susceptibility (ARMS)2 gene, and the VEGF and VEGFR axis [46,47,48,49,50,51,52,53]. Furthermore, inhibitor metalloproteinase (TIMP) 3, fibrillin, collagen 4A3, and metalloproteinase 19 and 9 appear to play a role in the development of nAMD [49]. (Figure 1).

1.3. Angiogenesis and VEGF Pathway

VEGF induces a vigorous angiogenic response in a variety of in vivo models [54,55]. VEGFR 1 and 2 are expressed in the cell surface of most blood endothelial cells. VEGFR-3, instead, is largely restricted to lymphatic endothelial cells. The primary contributor to angiogenesis is suggested to be VEGF-A, as it interacts with both VEGFR-1 and VEGFR-2 [56]. In contrast, PLGF and VEGF-B interact only with VEGFR-1; VEGF-E (ort-virus-derived) is a selective VEGFR-2 agonist; VEGF-C and VEGF-D bind VEGFR-2 and VEGFR-3. There is much evidence that VEGFR-2 is the major mediator of endothelial cell mitogenesis and survival, as well as angiogenesis and microvascular permeability [21]. In contrast, VEGFR-1 does not mediate an effective mitogenic signal in endothelial cells, and it may perform an inhibitory role by sequestering and preventing VEGF interaction with VEGFR-2 [57]. However, VEGFR-1 has an established signaling role in mediating monocyte chemotaxis [58].

2. Current Treatment Landscape in nAMD

2.1. Anti-VEGF Therapy: The Gold Standard

Intravitreal anti-VEGF therapies that target vascular permeability, angiogenesis, and inflammatory responses by inhibiting VEGF signaling are the current gold standard treatments for patients with nAMD [59]. Intravitreal anti-VEGF agents include aflibercept (Eylea®; Regeneron Pharmaceutical, Inc., Tarrytown, USA and Bayer Healthcare, Berlin, Germany) [60], ranibizumab (Lucentis®, Genentech, South San Francisco, CA, USA/Roche, Basel, Switzerland) [57,61] and bevacizumab (Avastin; Genentech, off-label use) [62]. These intravitreal anti-angiogenic agents used in clinical settings are based on limited activity against factors belonging to the VEGF family. This entails the inhibition of the VEGF-A activity observed in ranibizumab and bevacizumab, as well as aflibercept which has a broader spectrum of action, neutralizing, in addition to VEGF-A, other VEGF family ligands, such as VEGF-B, the placental growth factor-1 (PlGF-1), and placental growth factor-2 (PlGF-2). Pegaptanib sodium (Macugen®; Bausch + Lomb; Bridgewater, NJ, US)), the first VEGF-targeting agent approved by the U.S. Food and Drug Administration (FDA), is no longer used. Recently, brolucizumab (Beovu®; Novartis, Basel, Switzerland) [63] and faricimab (Vabysmo; Genentech/Roche, South San Francisco, CA, USA) have been approved for ophthalmic use. Brolucizumab targets the major VEGF-A isoforms [64], but, after an initial widespread adoption, its use has since been significantly limited in certain countries following several cases of severe occlusive retinal vasculitis [65]; faricimab is the latest antibody to receive approval, and it simultaneously targets VEGF-A and angiopoietin-2 (Ang II) [66]. Biosimilars have now started to enter the market in certain healthcare systems, in an attempt to make better use of resources. In terms of treatment regimens, non-monthly treatment schedules, such as the pro re nata (PRN) strategy and the treat and extend (T&E) protocol, have garnered considerable interest. In the PRN approach, the disease status is evaluated monthly, and treatment is provided if deemed necessary. In the T&E approach, treatment intervals are fixed based on the disease status observed during each visit [67]. At present, the T&E regimen has emerged as the prevailing treatment approach on a global scale, and future research directions are concentrating on even longer intervals between injections. This highlights the need for more persistent therapeutic agents and the exploration of alternative strategies to achieve prolonged efficacy.
Among the surgical methods for administering gene therapy products targeting the inner retina, intravitreal (IVT) injections are the least invasive.
The majority of currently administered intravitreal AAVs face challenges in adequately reaching the outer retina, RPE, and choroid because of the inner limiting membrane (ILM), which acts as a physical barrier between the vitreous and the retina [68]. Although the delivery of gene therapy products to the outer retina and choroid has been enhanced with the development of recombinant vectors like AAV2.GL and AAV2.NN [69,70], the utilization of ADVM-022, an intravitreal AAV-based aflibercept gene therapy, was discontinued due to dose-limiting toxicity [71]. Nevertheless, the fact that IVT injection is more immunogenic than subretinal delivery has been highlighted [72,73] and may be useful in future studies.

2.2. Limitations of Anti-VEGF Therapy

Although positive results are achieved in the majority of patients, approximately 25–35% of individuals with nAMD either show suboptimal responses to existing anti-VEGF treatments, experience delayed treatment failure, or require intensive and frequent IVT therapy [74,75]. Among the 35% who do not respond optimally to therapy, more than 10% experience deterioration despite treatment, and an additional 25% exhibit no signs of improvement [76,77]. For patients who achieve disease stability, discontinuing therapy could potentially have negative consequences, necessitating the continuation of treatment at consistent intervals to maintain vision. Furthermore, in certain individuals with aggressive nAMD, the continued use of anti-VEGF therapy, even after achieving stability, does not sufficiently prevent the recurrence of the disease [78,79]. The consequences of a suboptimal response and limited effectiveness over time, leading to poor vision, significantly affect the outcomes reported by the patients. Moreover, the need for repeated treatments for nAMD places a substantial burden on healthcare systems, patients, and their caregivers. Moreover, present anti-VEGF treatments are associated with certain adverse events which, although infrequent, can considerably affect eyesight. Endophthalmitis is a severe complication that arises in approximately 1 in 3500 injections [80]. Another notable potential complication of anti-VEGF therapy is the risk of intraocular inflammation, which, if severe, may lead to irreversible vision loss [81]. Furthermore, a temporary increase in intraocular pressure is frequently noticed shortly after IVT injection of all anti-VEGF agents [82]. The repeated use of anti-VEGF treatments can also be associated with unfavorable effects. For example, macular atrophy, which represents an advanced phenotype of nAMD and may result in permanent vision loss, has been documented as being potentially linked to long-term anti-VEGF use [83]. The causative relationship is still not well-defined, and it is possible that macular atrophy is simply part of the natural history of some forms of treated MNV [84,85]; subretinal fibrosis represents another advanced manifestation of nAMD linked to permanent vision impairment and can be the result of untreated nAMD itself, which complicates any potential association between sub-retinal fibrosis and long-term anti-VEGF therapy. Exploring alternative agents that provide comparable or enhanced efficacy while requiring fewer injections and maintaining a longer duration of action could potentially address many of these limitations.

2.3. Emerging Therapies and Need for Alternative Treatment Options

Conbercept (Lumitin; Chengdu Kang Hong Biotech, Chengdu, China) is a recently tested agent designed to bind VEGFA, VEGF-B and PlGF [86], but the PANDA-1 and PANDA-2 phase III trials for nAMD were concluded in 2021. They were halted because the desired primary outcome, specifically the non-inferiority of conbercept compared to aflibercept, was not attained (NCT03577899 and NCT03630952). OPT-302, a novel “trap” molecule, binds to VEGF-C and VEGF-D, in turn inhibiting their activation of VEGFR2 and VEGFR3 [87]. This new agent is currently being assessed in phase III trials with and without either ranibizumab (ShORe trial, NCT04757610) or aflibercept (COAST trial, NCT04757636). Researchers have also attempted to develop eye drops either as potential standalone treatments [88,89] or as an adjunct to optimize IVT therapy, but they have yet to be proven effective. Port delivery systems (PDS) are innovative devices utilizing surgically implantable reservoirs that are required to be refilled periodically and have been developed to enable the continuous administration of anti-VEGF agents directly into the eye through passive diffusion [90]. The FDA initially granted approval for a PDS containing ranibizumab as a treatment option for nAMD but the product has since been recalled and is no longer available commercially. Other treatments for nAMD, focusing on the VEGF system and alternative pathways, including heparin-binding variants of VEGF receptor 1 [91] and cells derived from induced pluripotent stem cells (iPSCs), have been under assessment [92]. These multi-targeted therapies and alternative treatment options, such as retinal gene therapy, may be the answer to the unmet needs in the treatment of nAMD [93,94]. An attractive alternative approach, in fact, involves using a single intraocular injection of a gene therapy vector that would continuously express an anti-angiogenic protein to block the pathological neovascularization in AMD. The aim of this review is to summarize the rationale and progress of preclinical and clinical trials using gene delivery strategies for the treatment of nAMD.

3. Gene Therapy Strategies for nAMD

3.1. Overview of Genes Targeted

As stated above, AMD is known to be a multifactorial disease, the development and progression of which is governed by the complex interaction of various environmental and genetic elements; aging is the primary factor, and drives the overexpression of VEGF-A in the macular microenvironment among elderly patients. Advancements in technologies, such as single-cell sequencing and genome-wide association studies (GWASs), have revealed mutations and factors that contribute to the progression of AMD. Through GWASs, specific genes, including CFH on chromosome 1 and ARMS2 and HTRA1, both residing on chromosome 10, have emerged as significant loci closely linked to advanced AMD [26,95]. The CHF variant is primarily connected to the presence of drusen, whereas the ARMS2-HTRA1 variant is correlated with the occurrence of subretinal or sub-RPE hemorrhages [96]. Although these genes are involved in the development of nAMD and may be useful predictors of treatment response, they have yet to be shown to have a significant role in its treatment. Other genes including MMP9, CETP, and TIMP3 have been linked to nAMD due to their roles in regulating the extracellular matrix remodeling [97], and the FGD6, HTRA1, and CFH genes play pivotal roles in governing oxidative stress and inflammation, which in turn regulate the advancement of angiogenesis, thereby contributing to the progression of nAMD [98].
However, the RPE hypoxia previously described promotes an over-expression of the hypoxia-inducible factor alpha (HIF-α) and VEGF-A by RPE cells, with the consequent degeneration of the RPE cells themselves and of Bruch’s membrane [99]. Anti-VEGF treatments have really shown that VEGFA/HIF-α-related genes (VEGF, VEGFR, PDGF, PEGF) can be useful treatments; this makes VEGF, VEGFR, PDGF, and PEGF the primary targets for the current gene therapy [100].
Gene therapy for nAMD faces challenges due to the complexity of the genes associated with the condition. Unlike monogenic disorders with a small gene that can fit into an AAV for the standard gene augmentation therapy, nAMD involves multiple genetic factors; the already mentioned genes contribute to disease susceptibility, making it challenging to devise a one-size-fits-all gene therapy. The diverse genetic landscape of nAMD adds a layer of complexity, requiring a nuanced approach in developing gene therapies tailored to the specific genetic factors.

3.2. Gene Silencing and Inhibition of VEGF Expression

Exploring gene silencing through small interfering RNA (siRNA) or microRNA (miRNA) targeting VEGF is considered as a potential approach for AMD treatment [101]. Numerous clinical trials are currently underway, focusing on the utilization of precise gene silencing methods [102,103,104]. After being introduced into cells, siRNA binds and activates the RNA-induced silencing complex, which in turn targets and degrades any cells complementary to the siRNA sequence, thereby preventing protein synthesis.
Bevasiranib, a modified naked RNA, results in the downregulation of VEGF-A by means of its intracellular transcriptional inhibitor action and possibly its TLR3-mediated activity, and may be the treatment of nAMD. A phase III human trial, which involved the intravitreal administration of siRNA bevasiranib (NCT00499590), was halted, as it was deemed unlikely to achieve its primary objective [105]. As bevasiranib may only inhibit new VEGF synthesis, without impacting existing VEGF levels, a phase III trial (NCT00499590) was also performed to assess the efficacy of the combined bevasiranib and ranibizumab therapy for nAMD treatment, but this too was unlikely to meet its primary endpoint and was terminated.
AGN211745 (formerly Sirna-027) is a chemically modified naked siRNA that has VEGFR-1 as the target gene, inducing gene silencing by binding the complementary target RNA with the lytic cytoplasmic protein complexes known as RNA-induced silencing complexes, thereby reducing the level of VEGFR-mRNA and significantly inhibiting MNV development, with the potential to treat nAMD. However, despite positive findings in the phase I/II study, a phase II trial administering Sirna-027 (NCT00395057) did not meet crucial efficacy endpoints. (NCT00363714) [106].
Despite many efforts in multiple trials exploring gene silencing, studies have never advanced beyond phase III, as gene silencing methods encounter several obstacles, including RNA instability, limited bioavailability, and the potential for non-specific targeting. These challenges, common to most drug delivery systems, significantly hamper the successful application of siRNA therapeutics in the treatment of nAMD. Additionally, while siRNA-based therapies have demonstrated theoretical advances for patients with nAMD, this approach has not shown any superiority compared to conventional anti-VEGF treatments. This is primarily because even with siRNA therapies, the requirement for repeated injections persists, as their effect is temporary (3–7 days) due to their degradation by tissue nucleases. Nonetheless, the possibility of extending these effects exists through chemical alterations or the use of viral vectors, which could help maintain the efficacy of therapies based on RNA interference.
An alternative to siRNAs involves the use of microRNAs (miRNAs) which are small (18–22 nucleotide), single-stranded, noncoding RNAs that down-regulate the gene expression post-transcriptionally [107]. Various research studies have shown that the dysregulation of miRNAs is relevant both in experimental AMD models and in AMD subjects, and may therefore potentially be associated with an increased risk of developing AMD [108,109,110]. MicroRNA mimics or anti-miRNA have the potential to be biomarkers, diagnostic tools, or targets for the control and treatment of this disease, by modulating retinal cellular function [111]. Unfortunately, the miRNAs evaluated in animal models of AMD behave differently compared to AMD patients; thus, their role in the disease remains unclear [112].

3.3. Gene Delivery Approaches: Viral Vector-Based and Non-Viral Delivery

To achieve successful results in gene therapy, it is essential to use a vector that ensures prolonged gene expression levels while minimizing the risks of toxicity and immune reactions. Different types of vectors have been used.

3.3.1. Viral Vector-Based Delivery

Viral vectors are modified viruses commonly used in gene therapy approaches to deliver therapeutic genes or RNA-based molecules to target disease cells. They have been used as delivery vehicles to precisely transport therapeutic genetic material into the target cells within the eye and achieve a sustained therapeutic effect. In gene therapy for nAMD, vector selection is of paramount importance. For retinal gene supplementation, the optimal selection is the recombinant adeno-associated viral vector (AAV) [113,114]. Its small, single-stranded DNA genome of approximately 4.6 kilobases (kb) with organized capsid structure makes it conducive to genetic modifications [115]. AAVs are currently the most commonly used vector for retinal gene transfer in both preclinical studies and clinical trials [116]. They provide advantages like extended transgene expression, minimal risk of insertional mutagenesis, only slight inflammatory responses induced, and a low chance of germline transmission [117,118]. The most extensive AAV serotypes studied in ocular gene therapies are AAV2, AAV5 and AAV8 [119,120,121]. Gene therapy products utilizing AAV vector systems, including Glybera (alipogene tiparvovec to treat hereditary lipoprotein lipase deficiency) [122], Luxturna (voretigene neparvovec-rzyl), Zolgensma (onasemnogene abeparvovec to treat spinal muscular atrophy type 1) [123] and Hemgenix (Etranacogene dezaparvovec for the treatment of hemophilia B) [124], have received notable approvals. Among these, Luxturna, the first approved gene therapy for a genetic disease, is a recombinant AAV 2 vector containing human RPE65 complementary DNA that enables RPE cells to produce the retinoid isomerohydrolase RPE65. After its efficacy and safety were ultimately confirmed in an open-label, randomized and controlled phase 3 trial conducted at two centers in the United States, Luxturna was authorized for gene augmentation therapy in RPE65-associated retinal dystrophy [125] and stands out as a retinal gene therapy designed to treat Leber congenital amaurosis (LCA) [126]. However, a subset of patients undergoing subretinal Luxturna injection developed progressive perifoveal chorioretinal atrophy following surgery. Despite that, most patients did well on visual function measures. Although the mechanism for chorioretinal atrophy is not known at this time, there are several potential factors that must be considered, alone or in combination, namely: direct toxicity of the AAV2 vector to the photoreceptors and RPE, inflammation or immune response to the vector, surgical delivery and ocular factors [127]. Further studies are necessary to determine what potential factors predispose patients to this complication and to clarify what the implications are for gene therapy in nAMD, especially in terms of a immune response.
Moreover, retroviruses and lentiviruses have been employed in various gene therapy products, such as RetinoStat® (Oxford BioMedica, Oxford, UK, OXB-201) targeted for nAMD (NCT01301443) and stem cell therapy. Notably, subretinal administration of RetinoStat, a lentiviral vector expressing endostatin and angiostatin, demonstrated safety and good tolerance. Patients with severe nAMD exhibited signs of clinical improvement, including visual acuity stabilization and reduction in vascular leakage [128]. Nonetheless, retroviruses and lentiviruses carry risks such as the potential for insertional mutagenesis and germline transmission. Additionally, they might trigger more pronounced inflammatory responses compared to AAVs. An important aspect of gene therapy is the possible immune reaction towards the AAV capsid: in humans, administering AAV vectors, unlike in many animal models, triggers antigen-specific T-cell activation, posing an increased risk during the initial postoperative phase. A brief period of immunosuppression around the surgery can help regulate immune responses until the capsid antigens are eliminated from the infected cells [129]. The route of vector delivery significantly influences immunogenicity. Subretinal delivery is a favorable option for disorders primarily affecting the RPE and/or photoreceptors. Given that the majority of inherited retinal disorders (IRD) involve either or both of these cell types, the subretinal delivery emerges as the prevailing administration route in gene therapy trials targeting monogenic conditions. This method involves the creation of a retinotomy near the temporal vascular arcades, allowing the bleb to slowly spread toward the foveal region, creating a shallow elevation [68]. Despite this type of delivery method involving a temporary detachment of the retina, the existing trial data indicates that it is generally safe and has the potential to offer effective therapeutic outcomes [69,70,71].

3.3.2. Non-Viral Delivery

Among the non-viral delivery techniques, the most straightforward approach is physical delivery, which involves injecting naked plasmid DNA, siRNA, mRNA or miRNA. However, this method has a limited efficacy due to the rapid degradation and minimal uptake [130]. Non-viral gene delivery through chemical techniques is attractive due to its lower potential to trigger immune responses, straightforward scalability and cost savings in production [131].
DNA nanoparticles consisting of a single molecule, compacted using polyethylene glycol (PEG)-substituted lysine peptides (CK30-PEG), have been utilized for transporting payloads of up to 20 kb in size [132,133]. These nanoparticles have demonstrated safety in diverse mouse models of retinal degeneration [134,135].
Lipid-based transfection systems have proven to be effective in delivering target genes to retinal cells in various studies. Numerous lipid-based drugs designed for eye diseases are accessible for transporting CRISPR or ribonucleoproteins for base editing [136,137]. Niosomes, consisting of cholesterol and uncharged single-chain surfactant [138,139], exhibit potential as non-viral carriers for gene delivery [140,141,142]. In ocular gene therapy, polymer-based platforms like chitosan, hyaluronic acid, polyethyleneimine (PEI), poly(amidoamine) (PAMAM), PEG, poly (lactic-glycolic acid) (PLGA) and poly(L-lysine) (PLL) have been under investigation [101,130].

3.4. Gene Editing Technologies and CRISPR/Cas9

Gene editing technology involves the manipulation of the target gene at the DNA or genomic level. The most common gene editing system to date uses clustered, regularly interspaced, short palindromic repeat (CRISPR) endonucleases such as Cas9, which can cut the DNA at a precise, targeted location, to either ablate or repair a destructive mutation [143]. The CRISPR/Cas9 system comprises a guide RNA targeting the gene of interest and an endonuclease that creates a site-specific double-stranded DNA “cut”, enabling precise genetic modification [144]. This allows for the lasting and accurate modification or removal of a mutation associated with a specific disease [145]. However, when addressing mutations in a single gene, CRISPR may not be effective for patients without a recognized genetic diagnosis. The CRISPR-Cas9 system has several potential advantages over other editing tools such as its simplicity of target design, ease of generating large-scale libraries and relatively low cost [146,147]. Moreover, the genome editing with CRISPR-Cas9 enables multiple editing through the engagement of multiple guide RNAs (gRNAs) [148,149]. Treatment for AMD patients can involve the use of the adeno-associated viral vector (AAV)-CRISPR tool, utilizing CjCas9 (Campylobacter jejuni) [150,151], and type-V CRISPR-Cas systems with LbCpf1 nucleases. AAV-delivered CjCas9 can accurately target and modify specific sites in the human or mouse genome, inducing mutations in RPE cells. In this context, CjCas9 can target the VEGFA or Hif1a gene in RPE cells, potentially reducing the size of laser-induced neovascularization. This approach may evolve into an in vivo genome editing therapy for nAMD [152]. Progress in CRISPR/Cas9 technology, including base and prime editing, holds the promise of improving the efficiency and cost-effectiveness of using CRISPR/Cas9 to treat retinal diseases such as nAMD.
However, a key challenge in the application of CRISPR/Cas9 technology remains the manufacturing and production for in vivo editing [153], and all CRISPR applications in retinal diseases including nAMD have been largely experimental; clinical trials of CRISPR for nAMD are lacking, as the field is still exploring safety and efficacy concerns.
The genomic impacts of transduction using AAV vectors encoding CRISPR-Cas nucleases are still under investigation; high levels of AAV integration (up to 47%) into Cas9-induced double-strand breaks (DSBs) are in therapeutically relevant genes in cultured murine neurons, mouse brain, muscle, and cochlea, and this should be recognized as a common outcome for applications that utilize AAV for genome editing [113]. Moreover, efficient gene delivery and editing can be achieved through the ocular delivery of mRNA packaged in lipid nanoparticles (LNPs). Subretinal injections of LNPa containing Cre mRNA in the mouse show a tdTomato signal in the RPE, enabling genome editing in the retina; in the future, this can be used to correct genetic mutations that lead to blindness [114].

4. Clinical Trials and Promising Gene Therapy Approaches

Clinical trials investigating gene therapy for nAMD currently adopt two strategies: the intraocular administration of modified viral vectors expressing antiangiogenic proteins, and RNA interference molecules to contrast the VEGF overexpression.
To this purpose, PEDF, endostatin, angiostatin, secreted extracellular domain of VEGFR1 and sFLT-1 have been targeted by gene therapy [154].

4.1. PEDF

A phase I clinical trial (ClinicalTrials.gov: NCT00109499) explored the safety of AdGVPEDF.11D in patients affected by advanced nAMD. The investigators delivered the PEDF gene via an adenoviral vector with deficient replication (by deletion of E1, E3, and E4). PEDF is an important endogenous antiangiogenic factor, and its levels are low in the presence of nAMD. Adenovirus, a double-strand DNA virus, can carry up to 37 kb for transgene delivery [155,156,157,158]. The participants received an intravitreal injection of AdGVPEDF.11D with dosages ranging from 1E6 and 1E9 particle units (PU). In 25% of cases, there were reports of mild and temporary intraocular inflammation, with no severe adverse events. Although the study was not designed to assess the therapeutic efficacy, neovascularization was observed to be stable or reduced in patients receiving 1E8 or 1E9 PU, compared to those receiving lower doses.

4.2. Anti-VEGF

Intravitreal and subretinal injection of FLT-1 (also known as VEGFR-1) or FLT-1 derivates have been tested on nAMD patients after encouraging results on animal models [159]. FLT-1 expression is normally upregulated by hypoxia, neutralizing VEGF-A, and thereby preventing its dimerization with membrane receptor VEGFR-2 and the consequent pro-angiogenic pathway. The intravitreal injection of AAV2-sFLT01, encoding for a fusion protein composed by sFLT-1 domain 2 and the Fc domain of IgG1, was tested in a phase I trial (ClinicalTrials.gov: NCT01024998, Sanofi Genzyme, Paris, France), whereas the subretinal administration of recombinant AAV (rAAV).sFLT-1, encoding the natural soluble FLT-1, was experimented on in a phase I/IIa trial (ClinicalTrials.gov: NCT01494805, Avalanche Biotechnologies).
In the first trial, the viral vector was demonstrated to be safe, not detectable systemically and not eliciting immunogenic activity. Moreover, the encoded protein was detectable within 52 weeks in 5 of the 10 patients treated with the highest dosage (2E10 vector genomes). In general, the expression was dose-related, but variable among the subjects, with 80% of non-expressers showing, at baseline, anti-AAV2 antibody titers of 1:400 or greater, indicating a considerable impact of individual characteristics in determining the response to treatment. Although the treatment was well tolerated at all dosages, it did not produce any significant anatomical (retinal thickness) and functional (BCVA) improvement [160].
The phase I/IIa trial NCT01494805 confirmed the safety and effectiveness of the subretinal injection of the rAAV.sFlt-1 vector, resulting in an increase in retinal sFLT-1 levels. Forty patients suffering from nAMD were assigned to low-dose, high-dose or control arms. A regular intravitreal injection of ranibizumab was administered when patients showed a BCVA reduction or intraretinal/subretinal fluid increase on OCT or augmented leakage on fluorescein angiography during the 36-month follow-up. The number of intravitreal treatments and changes in BCVA and retinal thickness were recorded during the 36-month follow-up. This gene therapy demonstrated safety and good tolerance; however, no notable changes were observed in the examined endpoints [161,162,163]. The induced endogenous expression of anti-VEGF has also been explored in humans, after encouraging results on animal models.
In a phase I clinical trial (ClinicalTrials.gov: NCT03748784, Adverum Biotechnologies, Redwood City, CA, USA), an AAV2-derived vector, 7m8 (AAV.7m8-aflibercept), named ADVM-22, was administered via intravitreal injection in 18 nAMD aflibercept-responder patients. Aflibercept expression, BCVA, OCT changes and the need for a rescue treatment with standard intravitreal injection of aflibercept were assessed. The BCVA maintenance and retinal thickness reduction on OCT were observed in 12 patients who received 2E11 or 6E11 doses of ADVM-22, with 10 of them (83%) not requiring rescue treatment for about 11 months [164]. The ADVM-22 was assessed in the INFINITY trial for diabetic macular edema (DME) and in the OPTIC clinical trial in patients with nAMD. The data from the studies show marked differences in the safety profile between the two patient populations, with rapid, clinically relevant decreases in intraocular pressure refractory to steroids, requiring subsequent additional treatment in the treated eye of some of the patients with DME. Although no similar clinically relevant events were observed in the OPTIC trial patients, this unexpected occurrence has disrupted further evaluation of the intervention [165].
Positive results were also obtained with the delivery of a gene encoding a soluble monoclonal portion of an anti-VEGF antibody structurally similar to ranibizumab. The safety and tolerability of this gene treatment, called RGX-314 and administered via subretinal injection, was tested in a phase I/IIa trial (ClinicalTrials.gov: NCT03066258, REGENXBIO). The 42 enrolled nAMD patients had previously been treated with anti-VEGF intravitreal injections. They were divided into 5 cohorts receiving the adeno-associated viral vector (NAV AAV8) at different doses (3E9, 1E10, 6E10, 1.6E11 and 2.5E11 genome copies [GC] per eye). The rescue treatment consisted of intravitreal anti-VEGF in the case of vision loss of 5 or more ETDRS letters; persistent, increased or new intra/subretinal fluid on OCT; or the appearance of new macular hemorrhage. The aqueous levels of the encoded protein were observed to increase in a dose-dependent manner in the five subgroups, with the RGX-314 protein reaching 260.5 ng/mL in 1 year in the 6E10 cohort (six patients). In the same sub-group at 24 months, BCVA was improved by 14 ETDRS letters, and the central retinal thickness remained stable at the baseline. BCVA remained stable at 2 years, with changes within one ETDRS line in cohort 2, 4, and 5. The rescue treatment after 2 years was necessary in all cohorts, with a lower mean number in the higher dose cohorts (2.8, 4.4 and 2 ranibizumab injections in cohorts 3, 4 and 5, respectively), whereas the first two cohorts received a higher number of rescue injections (10.3 and 9.3 in cohort 1 and 2, respectively). RGX-314 showed a good tolerability; overall, the intervention demonstrated no severe adverse events at the lower doses; 2 participants in the highest dose group developed retinal pigmentary changes that resulted in vision loss. As a consequence, the protocol was amended [166]. These results encouraged several more studies on RGX-314 safety and efficacy on nAMD patients: a phase II trial (ClinicalTrials.gov: NCT04832724) comparing the effects of two different doses in two subretinal formulations, the clinical and the eventual commercial formulations, a phase II trial (ClinicalTrials.gov: NCT04514653) comparing 3 different doses of RGX-314 with ranibizumab, a randomized combination of a fixed dose of RGX-314 with either topical or local steroid formulations post-treatment, and a 5-year follow-up trial with a sub-study on the affected fellow-eye (ClinicalTrials.gov: NCT03999801). Unfortunately, no conclusive data from these trials are currently available [167].

4.3. Endostatins and Angiostatins

The subretinal injection of viral vectors encoding endostatin and angiostatin, which are endogenous inhibitors of angiogenesis, showed good preclinical results on mice with laser-induced neovascularization [168,169]. These results prompted a phase I clinical study (ClinicalTrials.gov: NCT01301443, Oxford Biomedica) on subretinal treatment with the non-replicating bicistronic EIAV vector encoding both endostatin and angiostatin (RetinoStat) on humans with advanced nAMD. The trial enrolled 21 patients that were divided into three cohorts receiving a different treatment dose (4E4, 2.4E5 and 8E5 transduction units [TU]). The gene therapy was safe, well tolerated and generated a sustained expression of angiostatin and endostatin, which was detected in aqueous humor samples of eight patients for up to 2.5 years and in 2 patients for more than 4 years. Unfortunately, despite a documented reduction of fluorescein leakage, the treatment produced no functional improvement [170].

4.4. Complement Cascade Inhibition

The complement cascade activation with membrane attack complex (MAC) accumulation has been observed to be upregulated in AMD patients, with consequent RPE cell damage. This process is thought to play an important pathogenetic role in both atrophic and nAMD [171]. CD59 is a membrane that prevents MAC formation on the cell membrane in the final phases of the complement cascade leading to cell lysis. Therefore, a soluble form of this molecule has been studied for gene therapy applications in dry AMD and nAMD. For the latter, a phase I trial (NCT03585556) adopting the intravitreal injection of AAVCAGsCD59, a viral vector encoding for soluble CD59 was initiated, but the results have yet to be made available.

4.5. RNA Interference

Another gene-based therapeutic strategy to reduce the expression of VEGF and its receptors is gene silencing with siRNAs. As previously mentioned, these artificial RNA strands are capable of forming complexes with complementary mRNA, selectively silencing their expression after transcription. Preclinical successes on nAMD models led to clinical trials adopting bevasiranib, a 21-nt RNA silencing the VEGF encoding mRNA, and ANG 211745, a 21-nt RNA that silences the mRNA encoding for FLT-1, also known as VEGFR1.
Bevasiranib was the first siRNA approved for IVT use in clinical trials on nAMD patients. This treatment was proved safe in a phase I trial (ClinicalTrials.gov: NCT00722384, OPKO Health) adopting five dosing regimens (0.1, 0.33, 1, 1.5 and 3 mg). Since bevasiranib inhibits VEGF synthesis, but does not affect the preexisting VEGF levels, and its efficacy as a monotherapy was demonstrated insufficient in a phase II trial (ClinicalTrials.gov: NCT00259753), resulting in BCVA loss and neovascular lesion enlargement, this gene therapy was subsequently associated with intravitreal ranibizumab in patients affected by nAMD in a phase III trial (ClinicalTrials.gov: NCT00499590). Despite its promising rationale, the trial did not meet its primary endpoint and was terminated.
Another phase III trial testing the efficacy and safety of the combined bevasirinab-ranibizumab therapy was aborted even before the enrollment started due to concerns regarding its Toll-like receptor (TLR) action, which was detected in murine models, and observed to induce RPE cell apoptosis [105].
The first phase I trial (ClinicalTrials.gov: NCT00363714, Allergan, Dublin, Ireland) assessing another intravitreal si RNA (ANG 211745) safety in nAMD patients showed good results, but the following phase II trial failed to reach its therapeutic targets [106]. Further concerns emerged on the TLR3 pathway activation; thus, more specific gene treatments were developed to overcome this limitation. In a phase I trial, the intravitreal injection of PF-045236 (ClinicalTrials.gov: NCT00725685), a 19-nt siRNA silencing the hypoxia-induced gene RTP801, was tested on patients with MNV or DME and was demonstrated to be safe and well tolerated [172]. In the subsequent phase II MONET trial (ClinicalTrials.gov: NCT00713518), this gene therapy showed no superiority in improving BCVA when compared with ranibizumab, but the two treatments combined showed synergetic efficacy [173].
The need to frequently combine treatment regimes in order to obtain the best outcome for the patient highlights the complexity of nAMD pathogenesis and, consequently, the need for a multifactorial therapeutic approach. To this purpose, a single gene therapy that regulates the expression of different proangiogenic molecules simultaneously would represent the ideal solution; preclinical studies by Askou et al. evaluated multigenic lentiviral vectors in human cells and in mouse retina that encode for both PEDF and anti-VEGF miRNA [174].
Several other clinical trials are currently recruiting, but the results are still awaited. These trials are included in Table 1 but have not been discussed in this review, as conclusive data have yet to be published.

5. Conclusions and Future Perspective

AMD is a complex disease characterized by a variety of genetic and molecular factors contributing to its pathogenesis and development. The approach to date for nAMD management is VEGF-A-based antiangiogenics; currently, the developed therapeutics are anti-VEGF antibodies or recombinant fusion proteins. However, the monthly repetitive intravitreal injection of these agents only achieves a limited control of nAMD, and its progression.
Gene therapy is a rapidly evolving field and is radically different from previously available forms of treatment. It has both advantages and potential disadvantages, but it could reduce the treatment burden by providing sustained and long-lasting therapeutic effects; when a therapeutic gene is successfully integrated into patient cells, it can continuously produce the desired protein, which in the case of nAMD most likely results in the down-regulation of VEGF. The need for a reduced frequency of treatment clearly has advantages for the patient, with significant improvement in the quality of life and preservation of vision and a reduction in the socio-economic burden associated with sight loss. Viral vector gene therapy appears to be the most promising option in nAMD, but the multifactorial character of the disease implies that gene therapy will require considerable effort to realize desirable therapeutic outcomes.
Furthermore, gene therapy raises several concerns. The potential off-target effects have already been mentioned; however, the main safety concern is ensuring precise control over gene expression; once a gene is introduced or modified in a cell, it cannot be easily switched off. This is of particular concern if the therapeutic gene has the potential to produce a protein that may be harmful if overexpressed. Furthermore, once a gene is modified, there is limited control over its expression; this is important, as the level of function may not be adjusted according to changing patient needs [175,176].
Moreover, the selection of promoters plays a crucial role in gene therapy for retinal diseases. Promoters are DNA sequences that control the initiation of gene expression, determining when and where a therapeutic gene is activated. In retinal gene therapy, the choice of promoter influences the specificity, strength and duration of gene expression within the target cells. Different retinal diseases may require distinct promoter characteristics. Precision in promoter selection helps avoid off-target effects and enhances the therapeutic gene’s therapeutic efficacy. Additionally, the durability of gene expression is a critical consideration. Some diseases may benefit from sustained expression over an extended period, necessitating the use of promoters that support long-term gene activity. Overall, the strategic choice of promoters in nAMD gene therapy is pivotal for optimizing treatment outcomes, tailoring expression patterns to specific cell types, and achieving the desired therapeutic effects while minimizing unintended consequences [177].
The number of studies reported in the literature is vast; this review does provide an exhaustive analysis but represents the issues the authors considered pertinent regarding gene therapy for nAMD. While gene therapy approaches, including those involving AAV vectors and CRISPR-Cas nucleases, are actively being explored, their widespread clinical application for nAMD is still in the investigational stage.
The results of the clinical trials described in this review may lay the foundations to revolutionize treatment plans for nAMD in the future. Challenges and areas of controversy persist, but progress has been made to optimize the dosage of these drugs, the routes of administration, post-injection management and long-term benefits. Future gene therapy studies need to live up to several key expectations to advance the field effectively. Future studies should focus on enhancing the precision and specificity of gene therapies. This involves developing technologies that can accurately target specific cells or tissues, minimizing off-target effects. The sustainability of gene therapies is crucial; ensuring that the therapeutic effects persist over the long term is essential for the success of these treatments, particularly in chronic conditions like nAMD. Continued efforts must be made to improve the safety profiles of gene therapies. This includes minimizing adverse reactions, immune responses, and other potential risks associated with the delivery of therapeutic genes. Moreover, gene therapy approaches can be combined with other treatment modalities, such as anti-VEGF therapies, to potentially enhance the overall effectiveness of treatment. While gene therapy for nAMD is still in its early stages, ongoing research and advancements in gene delivery methods, vector development, and gene editing technologies hold promise for improving its efficacy and safety profile and advancing the potential of clinical applications and patient benefit.

Author Contributions

Conceptualization, L.F., A.G., G.T., M.Z., C.S. and L.S.; methodology, A.G., G.T., M.Z. and L.F.; validation, A.G., G.T., M.Z., L.F. and C.S.; formal analysis, A.G., G.T. and L.F.; investigation, A.G., G.T. and L.F.; resources, C.S.; writing—original draft preparation, A.G, G.T. and L.F.; writing—review and editing A.G., G.T., M.Z. and L.F.; visualization, A.G.,G.T., M.Z., L.F., L.S. and C.S.; supervision, M.Z., L.S. and C.S.; project administration, M.Z. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wong, W.L.; Su, X.; Li, X.; Cheung, C.M.G.; 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] [PubMed]
  2. Klein, R.; Klein, B.E.K.; Cruickshanks, K.J. The Prevalence of Age-Related Maculopathy by Geographic Region and Ethnicity1This Paper Has Been Edited by Neville, N. Osborne, PhD, DSc, Nuffield Laboratory of Ophthalmology, University of Oxford, Walton Street, Oxford, UK; and Gerald, J. Chader, The Foundation Fighting Blindness, Hunt Valley, MS.1. Prog. Retin. Eye Res. 1999, 18, 371–389. [Google Scholar] [CrossRef] [PubMed]
  3. Mitchell, P.; Smith, W.; Attebo, K.; Wang, J.J. Prevalence of Age-Related Maculopathy in Australia. Ophthalmology 1995, 102, 1450–1460. [Google Scholar] [CrossRef] [PubMed]
  4. Klaver, C.C.; Assink, J.J.; van Leeuwen, R.; Wolfs, R.C.; Vingerling, J.R.; Stijnen, T.; Hofman, A.; de Jong, P.T. Incidence and Progression Rates of Age-Related Maculopathy: The Rotterdam Study. Investig. Ophthalmol. Vis. Sci. 2001, 42, 2237–2241. [Google Scholar]
  5. 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. Ophthalmology 2010, 117, 921–927. [Google Scholar] [CrossRef]
  6. Wong, T.; Chakravarthy, U.; Klein, R.; Mitchell, P.; Zlateva, G.; Buggage, R.; Fahrbach, K.; Probst, C.; Sledge, I. The Natural History and Prognosis of Neovascular Age-Related Macular Degeneration. Ophthalmology 2008, 115, 116–126.e1. [Google Scholar] [CrossRef] [PubMed]
  7. Thomas, C.J.; Mirza, R.G.; Gill, M.K. Age-Related Macular Degeneration. Med. Clin. N. Am. 2021, 105, 473–491. [Google Scholar] [CrossRef]
  8. Nusinowitz, S.; Wang, Y.; Kim, P.; Habib, S.; Baron, R.; Conley, Y.; Gorin, M. Retinal Structure in Pre-Clinical Age-Related Macular Degeneration. Curr. Eye Res. 2018, 43, 376–382. [Google Scholar] [CrossRef]
  9. Pinelli, R.; Bertelli, M.; Scaffidi, E.; Fulceri, F.; Busceti, C.L.; Biagioni, F.; Fornai, F. Measurement of Drusen and Their Correlation with Visual Symptoms in Patients Affected by Age-Related Macular Degeneration. Arch. Ital. Biol. 2021, 3, 82–104. [Google Scholar] [CrossRef]
  10. Ferris, F.L.; Fine, S.L.; Hyman, L. Age-Related Macular Degeneration and Blindness Due to Neovascular Maculopathy. Arch. Ophthalmol. 1984, 102, 1640–1642. [Google Scholar] [CrossRef]
  11. Chavakis, E.; Dimmeler, S. Regulation of endothelial cell survival and apoptosis during angiogenesis. Arter. Thromb. Vasc. Biol. 2002, 22, 887–893. [Google Scholar] [CrossRef]
  12. Campochiaro, P.A. Molecular Pathogenesis of Retinal and Choroidal Vascular Diseases. Prog. Retin. Eye Res. 2015, 49, 67–81. [Google Scholar] [CrossRef]
  13. Alon, T.; Hemo, I.; Itin, A.; Pe’er, J.; Stone, J.; Keshet, E. Vascular Endothelial Growth Factor Acts as a Survival Factor for Newly Formed Retinal Vessels and Has Implications for Retinopathy of Prematurity. Nat. Med. 1995, 1, 1024–1028. [Google Scholar] [CrossRef] [PubMed]
  14. Okubo, A.; Rosa, R.H.; Bunce, C.V.; Alexander, R.A.; Fan, J.T.; Bird, A.C.; Luthert, P.J. The Relationships of Age Changes in Retinal Pigment Epithelium and Bruch’s Membrane. Investig. Ophthalmol. Vis. Sci. 1999, 40, 443–449. [Google Scholar]
  15. Bonilha, V.L. Age and disease-related structural changes in the retinal pigment epithelium. Clin. Ophthalmol. 2008, 2, 413–424. [Google Scholar] [CrossRef]
  16. Gehrs, K.M.; Anderson, D.H.; Johnson, L.V.; Hageman, G.S. Age-related macular degeneration—Emerging pathogenetic and therapeutic concepts. Ann. Med. 2009, 38, 450–471. [Google Scholar] [CrossRef] [PubMed]
  17. Algvere, P.V.; Kvanta, A.; Seregard, S. Drusen maculopathy: A risk factor for visual deterioration. Acta Ophthalmol. 2016, 94, 427–433. [Google Scholar] [CrossRef] [PubMed]
  18. Semenza, G.L. HIF-1: Mediator of Physiological and Pathophysiological Responses to Hypoxia. J. Appl. Physiol. 2000, 88, 1474–1480. [Google Scholar] [CrossRef]
  19. Kelly, B.D.; Hackett, S.F.; Hirota, K.; Oshima, Y.; Cai, Z.; Berg-Dixon, S.; Rowan, A.; Yan, Z.; Campochiaro, P.A.; Semenza, G.L. Cell Type–Specific Regulation of Angiogenic Growth Factor Gene Expression and Induction of Angiogenesis in Nonischemic Tissue by a Constitutively Active Form of Hypoxia-Inducible Factor 1. Circ. Res. 2003, 93, 1074–1081. [Google Scholar] [CrossRef]
  20. Metzger, C.S.; Koutsimpelas, D.; Brieger, J. Transcriptional Regulation of the VEGF Gene in Dependence of Individual Genomic Variations. Cytokine 2015, 76, 519–526. [Google Scholar] [CrossRef]
  21. Ferrara, N.; Gerber, H.-P.; LeCouter, J. The Biology of VEGF and Its Receptors. Nat. Med. 2003, 9, 669–676. [Google Scholar] [CrossRef]
  22. Ferrara, N. Vascular Endothelial Growth Factor: Basic Science and Clinical Progress. Endocr. Rev. 2004, 25, 581–611. [Google Scholar] [CrossRef] [PubMed]
  23. Baudouin, C.; Peyman, G.A.; Fredj-Reygrobellet, D.; Gordon, W.C.; Lapalus, P.; Gastaud, P.; Bazan, N.G. Immunohistological Study of Subretinal Membranes in Age-Related Macular Degeneration. Jpn. J. Ophthalmol. 1992, 36, 443–451. [Google Scholar] [PubMed]
  24. 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] [PubMed]
  25. Toomey, C.B.; Johnson, L.V.; Bowes Rickman, C. Complement factor H in AMD: Bridging genetic associations and pathobiology. Prog. Retin. Eye Res. 2018, 62, 38–57. [Google Scholar] [CrossRef]
  26. 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]
  27. 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]
  28. 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]
  29. Zareparsi, S.; Branham, K.E.H.; 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]
  30. The AMD Genetics Clinical Study Group; Gold, B.; Merriam, J.E.; Zernant, J.; Hancox, L.S.; Taiber, A.J.; Gehrs, K.; Cramer, K.; Neel, J.; Bergeron, J.; 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]
  31. Lorés-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. Ophthalmology 2018, 125, 1064–1074. [Google Scholar] [CrossRef]
  32. 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]
  33. 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] [PubMed]
  34. 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] [PubMed]
  35. 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]
  36. 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] [PubMed]
  37. Bradt, B.M.; Kolb, W.P.; Cooper, N.R. Complement-Dependent Proinflammatory Properties of the Alzheimer’s Disease β-Peptide. J. Exp. Med. 1998, 188, 431–438. [Google Scholar] [CrossRef]
  38. Johnson, L.V.; Leitner, W.P.; Rivest, A.J.; Staples, M.K.; Radeke, M.J.; Anderson, D.H. The Alzheimer’s Aβ-Peptide Is Deposited at Sites of Complement Activation in Pathologic Deposits Associated with Aging and Age-Related Macular Degeneration. Proc. Natl. Acad. Sci. USA 2002, 99, 11830–11835. [Google Scholar] [CrossRef]
  39. Thurman, J.M.; Renner, B.; Kunchithapautham, K.; Ferreira, V.P.; Pangburn, M.K.; Ablonczy, Z.; Tomlinson, S.; Holers, V.M.; Rohrer, B. Oxidative Stress Renders Retinal Pigment Epithelial Cells Susceptible to Complement-Mediated Injury. J. Biol. Chem. 2009, 284, 16939–16947. [Google Scholar] [CrossRef]
  40. Wu, Z.; Lauer, T.W.; Sick, A.; Hackett, S.F.; Campochiaro, P.A. Oxidative Stress Modulates Complement Factor H Expression in Retinal Pigmented Epithelial Cells by Acetylation of FOXO3. J. Biol. Chem. 2007, 282, 22414–22425. [Google Scholar] [CrossRef]
  41. Hageman, G.; 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]
  42. Johnson, L.V.; Leitner, W.P.; Staples, M.K.; Anderson, D.H. Complement Activation and Inflammatory Processes in Drusen Formation and Age Related Macular Degeneration. Exp. Eye Res. 2001, 73, 887–896. [Google Scholar] [CrossRef]
  43. Leung, K.W.; Barnstable, C.J.; Tombran-Tink, J. Bacterial Endotoxin Activates Retinal Pigment Epithelial Cells and Induces Their Degeneration through IL-6 and IL-8 Autocrine Signaling. Mol. Immunol. 2009, 46, 1374–1386. [Google Scholar] [CrossRef]
  44. Parmeggiani, F.; Romano, M.R.; Costagliola, C.; Semeraro, F.; Incorvaia, C.; D’Angelo, S.; Perri, P.; De Palma, P.; De Nadai, K.; Sebastiani, A. Mechanism of Inflammation in Age-Related Macular Degeneration. Mediat. Inflamm. 2012, 2012, 546786. [Google Scholar] [CrossRef] [PubMed]
  45. Ricci, F.; Bandello, F.; Navarra, P.; Staurenghi, G.; Stumpp, M.; Zarbin, M. Neovascular Age-Related Macular Degeneration: Therapeutic Management and New-Upcoming Approaches. Int. J. Mol. Sci. 2020, 21, 8242. [Google Scholar] [CrossRef]
  46. Skerka, C.; Chen, Q.; Fremeaux-Bacchi, V.; Roumenina, L.T. Complement Factor H Related Proteins (CFHRs). Mol. Immunol. 2013, 56, 170–180. [Google Scholar] [CrossRef] [PubMed]
  47. Mattapallil, M.J.; Caspi, R.R. Compliments of Factor H: What’s in It for AMD? Immunity 2017, 46, 167–169. [Google Scholar] [CrossRef] [PubMed]
  48. 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. [Google Scholar] [CrossRef]
  49. Natoli, R.; Fernando, N.; Jiao, H.; Racic, T.; Madigan, M.; Barnett, N.L.; Chu-Tan, J.A.; Valter, K.; Provis, J.; Rutar, M. Retinal Macrophages Synthesize C3 and Activate Complement in AMD and in Models of Focal Retinal Degeneration. Investig. Ophthalmol. Vis. Sci. 2017, 58, 2977. [Google Scholar] [CrossRef]
  50. Grassmann, F.; Heid, I.M.; Weber, B.H.F.; International AMD Genomics Consortium (IAMDGC). Recombinant Haplotypes Narrow the ARMS2/HTRA1 Association Signal for Age-Related Macular Degeneration. Genetics 2017, 205, 919–924. [Google Scholar] [CrossRef]
  51. Lazzeri, S.; Orlandi, P.; Figus, M.; Fioravanti, A.; Cascio, E.; Di Desidero, T.; Agosta, E.; Canu, B.; Sartini, M.S.; Danesi, R.; et al. The Rs2071559 AA VEGFR-2 Genotype Frequency Is Significantly Lower in Neovascular Age-Related Macular Degeneration Patients. Sci. World J. 2012, 2012, 420190. [Google Scholar] [CrossRef] [PubMed]
  52. Barchitta, M.; Maugeri, A. Association between Vascular Endothelial Growth Factor Polymorphisms and Age-Related Macular Degeneration: An Updated Meta-Analysis. Dis. Markers 2016, 2016, 8486406. [Google Scholar] [CrossRef]
  53. Mammadzada, P.; Corredoira, P.M.; André, H. The Role of Hypoxia-Inducible Factors in Neovascular Age-Related Macular Degeneration: A Gene Therapy Perspective. Cell Mol. Life Sci. 2020, 77, 819–833. [Google Scholar] [CrossRef]
  54. Leung, D.W.; Cachianes, G.; Kuang, W.-J.; Goeddel, D.V.; Ferrara, N. Vascular Endothelial Growth Factor Is a Secreted Angiogenic Mitogen. Science 1989, 246, 1306–1309. [Google Scholar] [CrossRef] [PubMed]
  55. Plouët, J.; Schilling, J.; Gospodarowicz, D. Isolation and Characterization of a Newly Identified Endothelial Cell Mitogen Produced by AtT-20 Cells. EMBO J. 1989, 8, 3801–3806. [Google Scholar] [CrossRef] [PubMed]
  56. Michels, S.; Schmidt-Erfurth, U.; Rosenfeld, P.J. Promising New Treatments for Neovascular Age-Related Macular Degeneration. Expert Opin. Investig. Drugs 2006, 15, 779–793. [Google Scholar] [CrossRef]
  57. Park, J.E.; Chen, H.H.; Winer, J.; Houck, K.A.; Ferrara, N. Placenta Growth Factor. Potentiation of Vascular Endothelial Growth Factor Bioactivity, in Vitro and in Vivo, and High Affinity Binding to Flt-1 but Not to Flk-1/KDR. J. Biol. Chem. 1994, 269, 25646–25654. [Google Scholar] [CrossRef]
  58. Barleon, B.; Sozzani, S.; Zhou, D.; Weich, H.A.; Mantovani, A.; Marmé, D. Migration of Human Monocytes in Response to Vascular Endothelial Growth Factor (VEGF) Is Mediated via the VEGF Receptor Flt-1. Blood 1996, 87, 3336–3343. [Google Scholar] [CrossRef]
  59. Kaiser, S.M.; Arepalli, S.; Ehlers, J.P. Current and future anti-VEGF agents for neovascular age-related macular degeneration. J. Exper. Pharmacol. 2021, 13, 905–912. [Google Scholar] [CrossRef]
  60. Korobelnik, J.F.; Do, D.V.; Schmidt-Erfurth, U.; Boyer, D.S.; Holz, F.G.; Heier, J.S.; Midena, E.; Kaiser, P.K.; Terasaki, H.; Marcus, D.M.; et al. Intravitreal aflibercept for DME. Ophthalmology 2014, 121, 2247–2254. [Google Scholar] [CrossRef]
  61. Folk, J.C.; Stone, E.M. Ranibizumab therapy for neovascular age-related macular degeneration. N. Engl. J. Med. 2010, 363, 1648–1655. [Google Scholar] [CrossRef] [PubMed]
  62. Avery, R.L.; Pieramici, D.J.; Rabena, M.D.; Castellarin, A.A.; Nasir, M.A.; Giust, M.J. Intravitreal bevacizumab (Avastin) for neovascular age-related macular degeneration. Ophthalmology 2006, 113, 363–372.e5. [Google Scholar] [CrossRef] [PubMed]
  63. Dugel, P.U.; Koh, A.; Ogura, Y.; Jaffe, G.J.; Schmidt-Erfurth, U.; Brown, D.M.; Gomes, A.V.; Warburton, J.; Weichselberger, A.; Holz, F.G.; et al. HAWK and HARRIER: Phase 3, multicenter, randomized, double-masked trials of brolucizumab for neovascular age-related macular degeneration. Ophthalmology 2020, 127, 72–84. [Google Scholar] [CrossRef] [PubMed]
  64. Garweg, J.G.; Blum, C.A.; Copt, R.P.; Eandi, C.M.; Hatz, K.; Prünte, C.F.; Seelig, E.; Somfai, G.M. Brolucizumab in Neovascular Age-Related Macular Degeneration and Diabetic Macular Edema: Ophthalmology and Diabetology Treatment Aspects. Ophthalmol. Ther. 2023, 12, 639–655. [Google Scholar] [CrossRef] [PubMed]
  65. Monés, J.; Srivastava, S.K.; Jaffe, G.J.; Tadayoni, R.; Albini, T.A.; Kaiser, P.K.; Holz, F.G.; Korobelnik, J.F.; Kim, I.K.; Pruente, C.; et al. Risk of Inflammation, Retinal Vasculitis, and Retinal Occlusion-Related Events with Brolucizumab: Post Hoc Review of HAWK and HARRIER. Ophthalmology 2021, 128, 1050–1059. [Google Scholar] [CrossRef] [PubMed]
  66. Nair, A.A.; Finn, A.P.; Sternberg, P., Jr. Spotlight on Faricimab in the Treatment of Wet Age-Related Macular Degeneration: Design, Development and Place in Therapy. Drug Des. Dev. Ther. 2022, 16, 3395–3400. [Google Scholar] [CrossRef] [PubMed]
  67. Wykoff, C.C.; Ou, W.C.; Brown, D.M.; Croft, D.E.; Wang, R.; Payne, J.F.; Clark, W.L.; Abdelfattah, N.S.; Sadda, S.R.; TREX-AMD Study Group. TREX-AMD Study Group. Randomized trial of treat-and-extend versus monthly dosing for neovascular age-related macular degeneration: 2-year results of the TREX-AMD study. Ophthalmol. Retina 2017, 1, 314–321. [Google Scholar] [CrossRef]
  68. Xue, K.; Groppe, M.; Salvetti, A.P.; MacLaren, R.E. Technique of retinal gene therapy: Delivery of viral vector into the subretinal space. Eye 2017, 31, 1308–1316. [Google Scholar] [CrossRef]
  69. Bainbridge, J.W.; Mehat, M.S.; Sundaram, V.; Robbie, S.J.; Barker, S.E.; Ripamonti, C.; Georgiadis, A.; Mowat, F.M.; Beattie, S.G.; Gardner, P.J.; et al. Long-Term effect of gene therapy on Leber’s congenital amaurosis. N. Engl. J. Med. 2015, 372, 1887–1897. [Google Scholar] [CrossRef]
  70. Weleber, R.G.; Pennesi, M.E.; Wilson, D.J.; Kaushal, S.; Erker, L.R.; Jensen, L.; McBride, M.T.; Flotte, T.R.; Humphries, M.; Calcedo, R.; et al. Results at 2 years after gene therapy for Rpe65-deficient Leber congenital amaurosis and severe Early-Childhood-Onset retinal dystrophy. Ophthalmology 2016, 123, 1606–1620. [Google Scholar] [CrossRef]
  71. MacLaren, R.E.; Groppe, M.; Barnard, A.R.; Cottriall, C.L.; Tolmachova, T.; Seymour, L.; Clark, K.R.; During, M.J.; Cremers, F.P.; Black, G.C.; et al. Retinal gene therapy in patients with choroideremia: Initial findings from a phase 1/2 clinical trial. Lancet 2014, 383, 1129–1137. [Google Scholar] [CrossRef] [PubMed]
  72. Tan, T.E.; Fenner, B.J.; Barathi, V.A.; Tun, S.B.B.; Wey, Y.S.; Tsai, A.S.H.; Su, X.; Lee, S.Y.; Cheung, C.M.G.; Wong, T.Y.; et al. Gene-based therapeutics for acquired retinal disease: Opportunities and progress. Front. Genet. 2021, 12, 795010. [Google Scholar] [CrossRef] [PubMed]
  73. Planul, A.; Dalkara, D. Vectors and gene delivery to the retina. Annu. Rev. Vis. Sci. 2017, 3, 121–140. [Google Scholar] [CrossRef] [PubMed]
  74. Spooner, K.; Hong, T.; Wijeyakumar, W.; Chang, A.A. Switching to aflibercept among patients with treatment-resistant neovascular age-related macular degeneration: A systematic review with meta-analysis. Clin. Ophthalmol. 2017, 11, 161–177. [Google Scholar] [CrossRef] [PubMed]
  75. Broadhead, G.K.; Hong, T.; Chang, A.A. Treating the untreatable patient: Current options for the management of treatment-resistant neovascular age-related macular degeneration. Acta Ophthalmol. 2014, 92, 713–723. [Google Scholar] [CrossRef] [PubMed]
  76. Comparison of Age-related Macular Degeneration Treatments Trials (CATT) Research Group; Martin, D.F.; Maguire, M.G.; Fine, S.L.; Ying, G.S.; Jaffe, G.J.; Grunwald, J.E.; Toth, C.; Redford, M.; Ferris, F.L., 3rd. Comparison of Age-related Macular Degeneration Treatments Trials Research Group. Ranibizumab and bevacizumab for treatment of neovascular age-related macular degeneration: Two-year results. Ophthalmology 2012, 119, 1388–1398. [Google Scholar] [CrossRef]
  77. Heier, J.S.; Brown, D.M.; Chong, V.; Korobelnik, J.F.; Kaiser, P.K.; Nguyen, Q.D.; Kirchhof, B.; Ho, A.; Ogura, Y.; Yancopoulos, G.D.; et al. Intravitreal aflibercept (VEGF trap-eye) in wet age-related macular degeneration. Ophthalmology 2012, 119, 2537–2548. [Google Scholar] [CrossRef]
  78. Garweg, J.G.; Traine, P.G.; Garweg, R.A.; Wons, J.; Gerhardt, C.; Pfister, I.B. Continued anti-VEGF treatment does not prevent recurrences in eyes with stable neovascular age-related macular degeneration using a treat-and-extend regimen: A retrospective case series. Eye 2022, 36, 862–868. [Google Scholar] [CrossRef]
  79. Khurana, R.N. Long-term management of neovascular age-related macular degeneration: To suspend or not to suspend? Ophthalmol. Retina 2019, 3, 621–622. [Google Scholar] [CrossRef]
  80. Torres-Costa, S.; Ramos, D.; Brandão, E.; Carneiro, Â.; Rosas, V.; Rocha-Sousa, A.; Falcão-Reis, F.; Falcão, M. Incidence of endophthalmitis after intravitreal injection with and without topical antibiotic prophylaxis. Eur. J. Ophthalmol. 2021, 31, 600–606. [Google Scholar] [CrossRef]
  81. Patil, N.S.; Dhoot, A.S.; Popovic, M.M.; Kertes, P.J.; Muni, R.H. Risk Of Intraocular Inflammation After Injection Of Antivascular Endothelial Growth Factor Agents: A Meta-Analysis. Retina 2022, 42, 2134–2142. [Google Scholar] [CrossRef]
  82. Levin, A.M.; Chaya, C.J.; Kahook, M.Y.; Wirostko, B.M. Intraocular Pressure Elevation Following Intravitreal Anti-VEGF Injections: Short- and Long-term Considerations. J. Glaucoma 2021, 30, 1019–1026. [Google Scholar] [CrossRef]
  83. Daien, V.; Nguyen, V.; Essex, R.W.; Guymer, R.; Arnold, J.J.; Munk, M.; Ceklic, L.; Gillies, M.C.; Barthelmes, D.; Fight Retinal Blindness! investigators. Prevalence and characteristics of macular atrophy in eyes with neovascular age-related macular degeneration. A study from a long-term observational dataset: The Fight Retinal Blindness project. Br. J. Ophthalmol. 2020, 104, 1064–1069. [Google Scholar] [CrossRef]
  84. Sadda, S.R.; Tuomi, L.L.; Ding, B.; Fung, A.E.; Hopkins, J.J. Macular Atrophy in the HARBOR Study for Neovascular Age-Related Macular Degeneration. Ophthalmology 2018, 125, 878–886. [Google Scholar] [CrossRef]
  85. Rofagha, S.; Bhisitkul, R.B.; Boyer, D.S.; Sadda, S.R.; Zhang, K. Group S-US: Seven-year outcomes in ranibizumab-treated patients in ANCHOR, MARINA, and HORIZON: A multicenter cohort study (SEVEN-UP). Ophthalmology 2013, 120, 2292–2299. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, Q.; Li, T.; Wu, Z.; Wu, Q.; Ke, X.; Luo, D.; Wang, H. Novel VEGF decoy receptor fusion protein conbercept targeting multiple VEGF isoforms provide remarkable anti-angiogenesis effect in vivo. PLoS ONE 2013, 8, e70544. [Google Scholar] [CrossRef] [PubMed]
  87. Dugel, P.U.; Boyer, D.S.; Antoszyk, A.N.; Steinle, N.C.; Varenhorst, M.P.; Pearlman, J.A.; Gillies, M.C.; Finger, R.P.; Baldwin, M.E.; Leitch, I.M. Phase 1 study of OPT-302 inhibition of vascular endothelial growth factors C and D for neovascular age-related macular degeneration. Ophthalmol. Retina 2020, 4, 250–263. [Google Scholar] [CrossRef] [PubMed]
  88. Samanta, A.; Aziz, A.A.; Jhingan, M.; Singh, S.R.; Khanani, A.M.; Chhablani, J. Emerging therapies in neovascular age-related macular degeneration in 2020. Asia Pac. J. Ophthalmol. 2020, 9, 250–259. [Google Scholar] [CrossRef]
  89. Chandrasekaran, P.R.; Madanagopalan, V.G. KSI-301: Antibody biopolymer conjugate in retinal disorders. Ther. Adv. Ophthalmol. 2021, 13, 25158414211027708. [Google Scholar] [CrossRef]
  90. Chen, E.R.; Kaiser, P.K. Therapeutic potential of the ranibizumab port delivery system in the treatment of AMD: Evidence to date. Clin. Ophthalmol. 2020, 14, 1349–1355. [Google Scholar] [CrossRef]
  91. Xin, H.; Biswas, N.; Li, P.; Zhong, C.; Chan, T.C.; Nudleman, E.; Ferrara, N. Heparin-binding VEGFR1 variants as long-acting VEGF inhibitors for treatment of intraocular neovascular disorders. Proc. Natl. Acad. Sci. USA 2021, 118, e1921252118. [Google Scholar] [CrossRef]
  92. Mandai, M.; Watanabe, A.; Kurimoto, Y.; Hirami, Y.; Morinaga, C.; Daimon, T.; Fujihara, M.; Akimaru, H.; Sakai, N.; Shibata, Y.; et al. Autologous induced stem-cell-derived retinal cells for macular degeneration. N. Engl. J. Med. 2017, 376, 1038–1046. [Google Scholar] [CrossRef] [PubMed]
  93. Parravano, M.; Costanzo, E.; Scondotto, G.; Trifirò, G.; Virgili, G. Anti-VEGF and Other Novel Therapies for Neovascular Age-Related Macular Degeneration: An Update. BioDrugs 2021, 35, 673–692. [Google Scholar] [CrossRef]
  94. Khachigian, L.M.; Liew, G.; Teo, K.Y.C.; Wong, T.Y.; Mitchell, P. Emerging therapeutic strategies for unmet need in neovascular age-related macular degeneration. J. Transl. Med. 2023, 21, 133. [Google Scholar] [CrossRef] [PubMed]
  95. 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]
  96. van Asten, F.; Simmons, M.; Singhal, A.; Keenan, T.D.; Ratnapriya, R.; Agrón, E.; Clemons, T.E.; Swaroop, A.; Lu, Z.; Chew, E.Y. Age-Related Eye Disease Study 2 Research Group. A deep phenotype association study reveals specific phenotype associations with genetic variants in age-related macular degeneration: Age-Related Eye Disease Study 2 (AREDS2) report no. 14. Ophthalmology 2018, 125, 559–568. [Google Scholar] [CrossRef]
  97. 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]
  98. Huang, L.; Zhang, H.; Cheng, C.Y.; Wen, F.; Tam, P.O.; Zhao, P.; Chen, H.; Li, Z.; Chen, L.; Tai, Z.; et al. A missense variant in FGD6 confers increased risk of polypoidal choroidal vasculopathy. Nat. Genet. 2016, 48, 640–647. [Google Scholar] [CrossRef]
  99. Botto, C.; Rucli, M.; Tekinsoy, M.D.; Pulman, J.; Sahel, J.-A.; Dalkara, D. Early and late stage gene therapy interventions for inherited retinal degenerations. Prog. Retin Eye Res. 2021, 86, 100975. [Google Scholar] [CrossRef]
  100. De Guimaraes, T.A.C.; Georgiou, M.; Bainbridge, J.W.; 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]
  101. Jiang, J.; Zhang, X.; Tang, Y.; Li, S.; Chen, J. Progress on ocular siRNA gene-silencing therapy and drug delivery systems. Fundam. Clin. Pharmacol. 2021, 35, 4–24. [Google Scholar] [CrossRef]
  102. Salminen, A.; Kauppinen, A.; Hyttinen, J.M.; Toropainen, E.; Kaarniranta, K. Endoplasmic reticulum stress in age-related macular degeneration: Trigger for neovascularization. Mol. Med. 2010, 16, 535–542. [Google Scholar] [CrossRef] [PubMed]
  103. Hoy, S.M. Patisiran: First global approval. Drugs 2018, 78, 1625–1631. [Google Scholar] [CrossRef]
  104. Padda, I.S.; Mahtani, A.U.; Parmar, M. Small Interfering RNA (siRNA) Based Therapy; StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  105. Garba, A.O.; Mousa, S.A. Bevasiranib for the treatment of wet, age-related macular degeneration. Ophthalmol. Eye Dis. 2010, 2, 75–83. [Google Scholar] [CrossRef] [PubMed]
  106. Kaiser, P.K.; Symons, R.C.; Shah, S.M.; Quinlan, E.J.; Tabandeh, H.; Do, D.V.; Reisen, G.; Lockridge, J.A.; Short, B.; Guerciolini, R.; et al. RNAi-Based Treatment for Neovascular Age-Related Macular Degeneration by Sirna-027. Am. J. Ophthalmol. 2010, 150, 33–39.e2. [Google Scholar] [CrossRef] [PubMed]
  107. Winter, J.; Jung, S.; Keller, S.; Gregory, R.I.; Diederichs, S. Many roads to maturity: MicroRNA biogenesis pathways and their regulation. Nat. Cell Biol. 2009, 11, 228–234. [Google Scholar] [CrossRef]
  108. Pogue, A.I.; Lukiw, W.J. Up–regulated pro–inflammatory MicroRNAs (miRNAs) in alzheimer’s disease (AD) and age–related macular degeneration (AMD). Cell Mol. Neurobiol. 2018, 38, 1021–1031. [Google Scholar] [CrossRef]
  109. Zhou, Q.; Anderson, C.; Hanus, J.; Zhao, F.; Ma, J.; Yoshimura, A.; Wang, S. Strand and cell type–specific function of microRNA–126 in angiogenesis. Mol. Ther. 2016, 24, 1823–1835. [Google Scholar] [CrossRef]
  110. Martinez, B.; Peplow, P. MicroRNAs as diagnostic and prognostic biomarkers of age–related macular degeneration: Advances and limitations. Neural Regen. Res. 2021, 16, 440–447. [Google Scholar] [CrossRef]
  111. Szemraj, M.; Bielecka-Kowalska, A.; Oszajca, K.; Krajewska, M.; Goś, R.; Jurowski, P.; Kowalski, M.; Szemraj, J. Serum micrornas as potential biomarkers of AMD. Med. Sci. Monitor. 2015, 21, 2734–2742. [Google Scholar] [CrossRef]
  112. Cruz-Aguilar, M.; Groman-Lupa, S.; Jimenez-Martınez, M.C. MicroRNAs as potential biomarkers and therapeutic targets in age-related macular degeneration. Front. Ophthalmol. 2023, 3, 1023782. [Google Scholar] [CrossRef]
  113. Hanlon, K.S.; Kleinstiver, B.P.; Garcia, S.P.; Zaborowski, M.P.; Volak, A.; Spirig, S.E.; Muller, A.; Sousa, A.A.; Tsai, S.Q.; Bengtsson, N.E.; et al. High levels of AAV vector integration into CRISPR-induced DNA breaks. Nat. Commun. 2019, 10, 4439. [Google Scholar] [CrossRef] [PubMed]
  114. Gautam, M.; Jozic, A.; Su, G.L.; Herrera-Barrera, M.; Curtis, A.; Arrizabalaga, S.; Tschetter, W.; Ryals, R.C.; Sahay, G. Lipid nanoparticles with PEG-variant surface modifications mediate genome editing in the mouse retina. Nat. Commun. 2023, 14, 6468. [Google Scholar] [CrossRef] [PubMed]
  115. Rabinowitz, J.E.; Rolling, F.; Li, C.; Conrath, H.; Xiao, W.; Xiao, X.; Samulski, R.J. Cross-packaging of a single adeno-associated virus (AAV) type 2 vector genome into multiple AAV serotypes enables transduction with broad specificity. J. Virol. 2002, 76, 791–801. [Google Scholar] [CrossRef] [PubMed]
  116. Day, T.P.; Byrne, L.C.; Schaffer, D.V.; Flannery, J.G. Advances in AAV vector development for gene therapy in the retina. Adv. Exp. Med. Biol. 2014, 801, 687–693. [Google Scholar] [PubMed]
  117. Grimm, D.; Büning, H. Small but increasingly mighty: Latest advances in AAV vector research, design, and evolution. Hum. Gene Ther. 2017, 28, 1075–1086. [Google Scholar] [CrossRef]
  118. Dismuke, D.J.; Tenenbaum, L.; Samulski, R.J. Biosafety of recombinant adeno-associated virus vectors. Curr. Gene Ther. 2013, 13, 434–452. [Google Scholar] [CrossRef]
  119. Bulcha, J.T.; Wang, Y.; Ma, H.; Tai, P.W.L.; Gao, G. Viral vector platforms within the gene therapy landscape. Signal Transduct. Target. Ther. 2021, 6, 53. [Google Scholar] [CrossRef]
  120. Han, I.C.; Burnight, E.R.; Ulferts, M.J.; Worthington, K.S.; Russell, S.R.; Sohn, E.H.; Mullins, R.F.; Stone, E.M.; Tucker, B.A.; Wiley, L.A. Helper-dependent adenovirus transduces the human and rat retina but elicits an inflammatory reaction when delivered subretinally in rats. Hum. Gene Ther. 2019, 30, 1371–1384. [Google Scholar] [CrossRef]
  121. Arbabi, A.; Liu, A.; Ameri, H. Gene therapy for inherited retinal degeneration. J. Ocul. Pharmacol. Ther. 2019, 35, 79–97. [Google Scholar] [CrossRef]
  122. Allocca, M.; Mussolino, C.; Garcia-Hoyos, M.; Sanges, D.; Iodice, C.; Petrillo, M.; Vandenberghe, L.H.; Wilson, J.M.; Marigo, V.; Surace, E.M.; et al. Novel adeno-associated virus serotypes efficiently transduce murine photoreceptors. J. Virol. 2007, 81, 11372–11380. [Google Scholar] [CrossRef] [PubMed]
  123. Petrs-Silva, H.; Dinculescu, A.; Li, Q.; Deng, W.T.; Pang, J.J.; Min, S.H.; Chiodo, V.; Neeley, A.W.; Govindasamy, L.; Bennett, A.; et al. Novel properties of tyrosine-mutant AAV2 vectors in the mouse retina. Mol. Ther. 2011, 19, 293–301. [Google Scholar] [CrossRef]
  124. Ylä-Herttuala, S. Endgame: Glybera finally recommended for approval as the first gene therapy drug in the European union. Mol. Ther. 2012, 20, 1831–1832. [Google Scholar] [CrossRef]
  125. Russell, S.; Bennett, J.; Wellman, J.A.; Chung, D.C.; Yu, Z.F.; Tillman, A.; Wittes, J.; Pappas, J.; Elci, O.; McCague, S.; et al. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: A randomised, controlled, open-label, phase 3 trial. Lancet 2017, 390, 849–860. [Google Scholar] [CrossRef]
  126. Gao, J.; Hussain, R.M.; Weng, C.Y. Voretigene Neparvovec in Retinal Diseases: A Review of the Current Clinical Evidence. Clin. Ophthalmol. 2020, 14, 3855–3869. [Google Scholar] [CrossRef] [PubMed]
  127. Gange, W.S.; Sisk, R.A.; Besirli, C.G.; Lee, T.C.; Havunjian, M.; Schwartz, H.; Borchert, M.; Sengillo, J.D.; Mendoza, C.; Berrocal, A.M.; et al. Perifoveal Chorioretinal Atrophy after Subretinal Voretigene Neparvovec-rzyl for RPE65-Mediated Leber Congenital Amaurosis. Ophthalmol. Retina 2022, 6, 58–64. [Google Scholar] [CrossRef] [PubMed]
  128. Lauer, A.K.; Campochiaro, P.A.; Sohn, E.H.; Kelleher, M.; Harrop, R.; Loader, J.; Ellis, S.; Mitrophanous, K. Phase I Safety and Tolerability results for RetinoStat®, a Lentiviral Vector Expressing Endostatin and Angiostatin, in Patients with Advanced Neovascular Age-Related Macular Degeneration. Investig. Ophthalmol. Vis. Sci. 2016, 57. [Google Scholar]
  129. Mingozzi, F.; High, K.A. Immune responses to AAV in clinical trials. Curr. Gene Ther. 2011, 11, 321–330. [Google Scholar] [CrossRef]
  130. Oliveira, A.V.; Rosa da Costa, A.M.; Silva, G.A. Non-viral strategies for ocular gene delivery. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 77, 1275–1289. [Google Scholar] [CrossRef]
  131. Patel, S.; Ryals, R.C.; Weller, K.K.; Pennesi, M.E.; Sahay, G. Lipid nanoparticles for delivery of messenger RNA to the back of the eye. J. Control. Release 2019, 303, 91–100. [Google Scholar] [CrossRef]
  132. Cai, X.; Nash, Z.; Conley, S.M.; Fliesler, S.J.; Cooper, M.J.; Naash, M.I. A partial structural and functional rescue of a retinitis pigmentosa model with compacted DNA nanoparticles. PLoS ONE 2009, 4, e5290. [Google Scholar] [CrossRef]
  133. Farjo, R.; Skaggs, J.; Quiambao, A.B.; Cooper, M.J.; Naash, M.I. Efficient non-viral ocular gene transfer with compacted DNA nanoparticles. PLoS ONE 2006, 1, e38. [Google Scholar] [CrossRef] [PubMed]
  134. Han, Z.; Banworth, M.J.; Makkia, R.; Conley, S.M.; Al-Ubaidi, M.R.; Cooper, M.J.; Naash, M.I. Genomic DNA nanoparticles rescue rhodopsin-associated retinitis pigmentosa phenotype. FASEB J. 2015, 29, 2535–2544. [Google Scholar] [CrossRef] [PubMed]
  135. Han, Z.; Conley, S.M.; Makkia, R.S.; Cooper, M.J.; Naash, M.I. DNA nanoparticle-mediated ABCA4 delivery rescues Stargardt dystrophy in mice. J. Clin. Investig. 2012, 122, 3221–3226. [Google Scholar] [CrossRef] [PubMed]
  136. Jang, H.K.; Jo, D.H.; Lee, S.N.; Cho, C.S.; Jeong, Y.K.; Jung, Y.; Yu, J.; Kim, J.H.; Woo, J.S.; Bae, S. High-purity production and precise editing of DNA base editing ribonucleoproteins. Sci. Adv. 2021, 7, eabg2661. [Google Scholar] [CrossRef] [PubMed]
  137. Zuris, J.A.; Thompson, D.B.; Shu, Y.; Guilinger, J.P.; Bessen, J.L.; Hu, J.H.; Maeder, M.L.; Joung, J.K.; Chen, Z.-K.; Liu, D.R. Cationic lipidmediated delivery of proteins enables efficient protein based genome editing in vitro and in vivo. Nat. Biotechnol. 2015, 33, 73–80. [Google Scholar] [CrossRef] [PubMed]
  138. Kazi, K.M.; Mandal, A.S.; Biswas, N.; Guha, A.; Chatterjee, S.; Behera, M.; Kuotsu, K. Niosome: A future of targeted drug delivery systems. J. Adv. Pharm. Technol. Res. 2010, 1, 374–380. [Google Scholar]
  139. Chen, S.; Hanning, S.; Falconer, J.; Locke, M.; Wen, J. Recent advances in non-ionic surfactant vesicles (niosomes): Fabrication, characterization, pharmaceutical and cosmetic applications. Eur. J. Pharm. Biopharm. 2019, 144, 18–39. [Google Scholar] [CrossRef]
  140. Al Qtaish, N.; Gallego, I.; Villate-Beitia, I.; Sainz-Ramos, M.; López-Méndez, T.B.; Grijalvo, S.; Eritja, R.; Soto-Sánchez, C.; Martínez-Navarrete, G.; Fernández, E.; et al. Niosomebased approach for in situ gene delivery to retina and brain cortex as immune-privileged tissues. Pharmaceutics 2020, 12, 198. [Google Scholar] [CrossRef]
  141. Durak, S.; Esmaeili Rad, M.; Alp Yetisgin, A.; Eda Sutova, H.; Kutlu, O.; Cetinel, S.; Zarrabi, A. Niosomal drug delivery systems for ocular disease-recent advances and future prospects. Nanomaterials 2020, 10, 1191. [Google Scholar] [CrossRef]
  142. Villate-Beitia, I.; Gallego, I.; Martínez-Navarrete, G.; Zárate, J.; López-Méndez, T.; Soto-Sánchez, C.; Santos-Vizcaíno, E.; Puras, G.; Fernández, E.; Pedraz, J.L. Polysorbate 20 non-ionic surfactant enhances retinal gene delivery efficiency of cationic niosomes after intravitreal and subretinal administration. Int. J. Pharm. 2018, 550, 388–397. [Google Scholar] [CrossRef]
  143. Yiu, G. Genome editing in retinal diseases using CRISPR technology. Ophthalmol. Retina 2018, 2, 1–3. [Google Scholar] [CrossRef]
  144. Redman, M.; King, A.; Watson, C.; King, D. What is CRISPR/Cas9? Arch. Dis. Child Educ. Pract. Ed. 2016, 101, 213–215. [Google Scholar] [CrossRef]
  145. Da Costa, B.L.; Levi, S.R.; Eulau, E.; Tsai, Y.T.; Quinn, P.M.J. Prime editing for inherited retinal diseases. Front. Genome Ed. 2021, 3, 775330. [Google Scholar] [CrossRef] [PubMed]
  146. Guha, T.K.; Wai, A.; Hausner, G. Programmable genome editing tools and their regulation for efficient genome engineering. Comput. Struct. Biotechnol. J. 2017, 15, 146–160. [Google Scholar] [CrossRef]
  147. Sander, J.D.; Joung, J.K. CRISPR-Cas systems for editing, regulating and targeting genomes. Nat. Biotechnol. 2014, 32, 347–355. [Google Scholar] [CrossRef] [PubMed]
  148. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex genome engineering using CRISPR/Cas systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef] [PubMed]
  149. Sakuma, T.; Nishikawa, A.; Kume, S.; Chayama, K.; Yamamoto, T. Multiplex genome engineering in human cells using all-in-one CRISPR/Cas9 vector system. Sci. Rep. 2014, 4, 5400. [Google Scholar] [CrossRef] [PubMed]
  150. Kuo, J.Z.; Wong, T.Y.; Ong, F.S. Genetic risk, ethnic variations and pharmacogenetic biomarkers in AMD and polypoidal choroidal vasculopathy. Expert Rev. Ophthalmol. 2013, 8, 127–140. [Google Scholar] [CrossRef]
  151. Imamura, Y.; Engelbert, M.; Iida, T.; Freund, K.B.; Yannuzzi, L.A. Polypoidal choroidal vasculopathy: A review. Surv. Ophthalmol. 2010, 55, 501–515. [Google Scholar] [CrossRef] [PubMed]
  152. Kim, K.; Park, S.W.; Kim, J.H.; Lee, S.H.; Kim, D.; Koo, T.; Kim, K.-e.; Kim, J.H.; Kim, J.-S. Genome surgery using Cas9 ribonucleoproteins for the treatment of age-related macular degeneration. Genome Res. 2017, 27, 419–426. [Google Scholar] [CrossRef]
  153. Ahmad, I. CRISPR/Cas9—A Promising Therapeutic Tool to Cure Blindness: Current Scenario and Future Prospects. Int. J. Mol. Sci. 2022, 23, 11482. [Google Scholar] [CrossRef]
  154. Lin, F.L.; Wang, P.Y.; Chuang, Y.F.; Wang, J.H.; Wong, V.H.Y.; Bui, B.V.; Liu, G.S. Gene Therapy Intervention in Neovascular Eye Disease: A Recent Update. Mol. Ther. 2020, 28, 2120–2138. [Google Scholar] [CrossRef]
  155. Bennett, J.; Wilson, J.; Sun, D.; Forbes, B.; Maguire, A. Adenovirus vector-mediated in vivo gene transfer into adult murine retina. Invest. Ophthalmol. Vis. Sci. 1994, 35, 2535–2542. [Google Scholar]
  156. Parks, R.J.; Chen, L.; Anton, M.; Sankar, U.; Rudnicki, M.A.; Graham, F.L. A helper-dependent adenovirus vector system: Removal of helper virus by Cre-mediated excision of the viral packaging signal. Proc. Natl. Acad. Sci. USA 1996, 93, 13565–13570. [Google Scholar] [CrossRef]
  157. 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 angio-genesis. Science 1999, 285, 245–248. [Google Scholar] [CrossRef] [PubMed]
  158. Holekamp, N.M.; Bouck, N.; Volpert, O. Pigment epithelium-derived factor is deficient in the vitreous of patients with choroidal neovascularization due to age-related macular degeneration. Am. J. Ophthalmol. 2002, 134, 220–227. [Google Scholar] [CrossRef]
  159. Shibuya, M. Vascular endothelial growth factor receptor-1 (VEGFR-1/Flt-1): A dual regulator for angiogenesis. Angiogenesis 2006, 9, 225–230. [Google Scholar] [CrossRef] [PubMed]
  160. Heier, J.S.; Kherani, S.; Desai, S.; Dugel, P.; Kaushal, S.; Cheng, S.H.; Delacono, C.; Purvis, A.; Richards, S.; Le-Halpere, A.; et al. Intravitreous injection of AAV2-sFLT01 in patients with advanced neovascular age-related macular degeneration: A phase 1, open-label trial. Lancet 2017, 390, 50–61. [Google Scholar] [CrossRef] [PubMed]
  161. RRakoczy, E.P.; Lai, C.M.; Magno, A.L.; Wikstrom, M.E.; French, M.A.; Pierce, C.M.; Schwartz, S.D.; Blumenkranz, M.S.; Chalberg, T.W.; Degli-Esposti, M.A.; et al. Gene therapy with recombinant adeno-associated vectors for neovascular age-related macular degeneration: 1 year follow-up of a phase 1 randomised clinical trial. Lancet 2015, 386, 2395–2403. [Google Scholar] [CrossRef]
  162. Rakoczy, E.P.; Magno, A.L.; Lai, C.M.; Pierce, C.M.; Degli-Esposti, M.A.; Blumenkranz, M.S.; Constable, I.J. Three-year follow-up of phase 1 and 2a rAAV.sFLT-1 subretinal gene therapy trials for exudative age-related macular degeneration. Am. J. Ophthalmol. 2019, 204, 113–123. [Google Scholar] [CrossRef]
  163. Constable, I.J.; Lai, C.M.; Magno, A.L.; French, M.A.; Barone, S.B.; Schwartz, S.D.; Blumenkranz, M.S.; Degli-Esposti, M.A.; Rakoczy, E.P. Gene therapy in neovascular age-related macular degeneration: Three-year follow-up of a phase 1 randomized dose-escalation trial. Am. J. Ophthalmol. 2017, 177, 150–158. [Google Scholar] [CrossRef]
  164. Adverum Biotechnologies. Transforming Gene Therapy. 2020. Available online: http://investors.adverum.com/static-files/c8256955-641c-45a3-bdda-5b99c2336a14 (accessed on 20 August 2023).
  165. Available online: https://investors.adverum.com/news/news-details/2021/Adverum-Provides-Update-on-ADVM-022-and-the-INFINITY-Trial-in-Patients-with-Diabetic-Macular-Edema/default.aspx (accessed on 20 August 2023).
  166. Available online: https://www.regenxbio.com/getmedia/d311cd48-532b-49ac-bf41-59d3d8907c8d/RGX-314_BobAvery-AAO-2021_11Nov21_FINAL.pdf?ext=.pdf (accessed on 20 August 2023).
  167. REGENXBIO Inc. Key Takeaways from the RGX-314 phase I/IIa Clinical Trial for Wet AMD (Cohorts 1–5). 2019. Available online: https://regenxbio.com/wp-content/uploads/2019/10/Key-Takeaways-From-The-RGX-314-Phase-I-IIa-Clinical-Trial-For-Wet-AMD-Cohorts-1-5.pdf (accessed on 20 August 2023).
  168. Kachi, S.; Binley, K.; Yokoi, K.; Umeda, N.; Akiyama, H.; Muramatu, D.; Iqball, S.; Kan, O.; Naylor, S.; Campochiaro, P.A. Equine infectious anemia viral vector-mediated codelivery of endostatin and angiostatin driven by retinal pigmented epithelium-specific VMD2 promoter inhibits choroidal neovascularization. Hum. Gene Ther. 2009, 20, 31–39. [Google Scholar] [CrossRef]
  169. Balaggan, K.S.; Binley, K.; Esapa, M.; MacLaren, R.E.; Iqball, S.; Duran, Y.; Pearson, R.A.; Kan, O.; Barker, S.E.; Smith, A.J.; et al. EIAV vector-mediated delivery of endostatin or angiostatin inhibits angiogenesis and vascular hyperpermeability in experimental CNV. Gene Ther. 2006, 13, 1153–1165. [Google Scholar] [CrossRef]
  170. Campochiaro, P.A.; Lauer, A.K.; Sohn, E.H.; Mir, T.A.; Naylor, S.; Anderton, M.C.; Kelleher, M.; Harrop, R.; Ellis, S.; Mitrophanous, K.A. Lentiviral vector gene transfer of endostatin/angiostatin for macular degeneration (GEM) study. Hum. Gene Ther. 2017, 28, 99–111. [Google Scholar] [CrossRef] [PubMed]
  171. Kumar-Singh, R. The role of complement membrane attack complex in dry and wet AMD—From hypothesis to clinical trials. Exp. Eye Res. 2019, 184, 266–277. [Google Scholar] [CrossRef] [PubMed]
  172. Nguyen, Q.D.; Schachar, R.A.; Nduaka, C.I.; Sperling, M.; Basile, A.S.; Klamerus, K.J.; Chi-Burris, K.; Yan, E.; Paggiarino, D.A.; Rosenblatt, I.; et al. PF-04523655 Study Group Phase 1 dose-escalation study of a siRNA targeting the RTP801 gene in age-related macular degeneration patients. Eye 2012, 26, 1099–1105. [Google Scholar] [CrossRef] [PubMed]
  173. Nguyen, Q.D.; Schachar, R.A.; Nduaka, C.I.; Sperling, M.; Klamerus, K.J.; Chi-Burris, K.; Yan, E.; Paggiarino, D.A.; Rosenblatt, I.; Aitchison, R.; et al. Evaluation of the siRNA PF-04523655 versus ranibizumab for the treatment of neovascular age-related macular degeneration (MONET Study). Ophthalmology 2012, 119, 1867–1873. [Google Scholar] [CrossRef]
  174. Askou, A.L.; Alsing, S.; Benckendorff, J.N.E.; Holmgaard, A.; Mikkelsen, J.G.; Aagaard, L.; Bek, T.; Corydon, T.J. Suppression of choroidal neovascularization by AAV-based dual-acting antiangiogenic gene therapy. Mol. Ther. Nucleic Acids 2019, 16, 38–50. [Google Scholar] [CrossRef]
  175. Khanani, A.M.; Thomas, M.J.; Aziz, A.A.; Weng, C.Y.; Danzig, C.J.; Yiu, G.; Kiss, S.; Waheed, N.K.; Kaiser, P.K. Review of gene therapies for age-related macular degeneration. Eye 2022, 36, 303–311. [Google Scholar] [CrossRef]
  176. Drag, S.; Dotiwala, F.; Upadhyay, A.K. Gene therapy for retinal degenerative diseases: Progress, challenges, and future directions. Investig. Ophthalmol. Vis. Sci. 2023, 64, 39. [Google Scholar] [CrossRef] [PubMed]
  177. Buck, T.M.; Wijnholds, J. Recombinant Adeno-Associated Viral Vectors (rAAV)-Vector Elements in Ocular Gene Therapy Clinical Trials and Transgene Expression and Bioactivity Assays. Int. J. Mol. Sci. 2020, 21, 4197. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of the pathophysiological process generating macular neovascularization (MNV). Adapted from Campochiaro, Anderson et al., Ricci et al. [12,36,45].
Figure 1. Overview of the pathophysiological process generating macular neovascularization (MNV). Adapted from Campochiaro, Anderson et al., Ricci et al. [12,36,45].
Biomedicines 11 03221 g001
Table 1. Clinical trials investigating gene therapy for nAMD.
Table 1. Clinical trials investigating gene therapy for nAMD.
Trial IDDevelopmentTested DrugRoute of AdministrationMechanismResults
NCT00109499Phase IAdGVPEDF.11DIntravitreal injectionInduction of PEDF expressionSafe, MNV size stable or reduced with dose 1E8 or 1E9 PU
NCT01024998Phase IAAV2-sFLT01Intravitreal injectionInduction of gene AAV2-sFLT01 encoding for an anti-angiogenic fusion protein formed by FLT-1 and IgG1 Fc domain that neutralizes VEGF-A before it binds its receptorSafe, good protein expression levels, but without significant anatomo-functional results
NCT01494805Phase I/IIaAAV (rAAV).sFLT-1Subretinal injectionInduction of gene encoding the natural anti-angiogenic protein FLT-1 that neutralizes VEGF-A before it binds its receptorSafe, no significant anatomo-functional results
NCT03748784Phase IAAV.7m8-afliberceptIntravitreal injectionInduction of endogenous aflibercept expression in confirmed exogenous aflibercept-responding patientsBCVA and retinal thickness maintenance in 12 patients out of 18 (10 of them not requiring rescue treatment for about 11 months)
NCT03066258Phase I/IIaRGX-314 (5 cohorts with different doses)Subretinal injectionInduction of endogenous anti-VEGF protein (similar to ranibizumab) expression in confirmed ranibizumab-responding patients
Induction of endogenous anti-VEGF protein (similar to ranibizumab) expression in confirmed ranibizumab-responding patients
Safe, good efficacy with functional and anatomical stabilization or improvement and less rescue treatments in patients treated with higher doses
NCT04832724Phase IIRGX-314 (clinical vs. eventual commercial formulation)Recruiting, data not available
NCT04514653Phase IIRGX-314 vs. ranibizumab
NCT03999801Phase IIRGX-314 vs. ranibizumab
and
RGX-314 + local vs. RGX-314 + topical steroids
NCT01301443Phase IRetinoStatSubretinal injectionInduction of supplemental endogenous endostatin and angiostatin expressionSafe.
Non-significant effectiveness
NCT03585556Phase IAAVCAGsCD59Intravitreal injectionInduction of soluble CD59 expression to prevent MAC formation and cellular damage and apoptosisData not available
NCT00722384Phase IBevasiranibIntravitreal injectionPost-transcription silencing of VEGF mRNASafe
NCT00259753Phase IIBevasiranibVision loss and MNV extension
NCT00499590Phase IIIBevasiranib
combined with intravitreal ranibizumab
Terminated due to missed primary endpoints
NCT00363714Phase IAGN 211745Intravitreal injectionPost-transcription silencing of FLT-1 (VEGFR-2) mRNASafe
NCT00725685Phase IPF-04523655Intravitreal injectionPost-transcription silencing of hypoxia-induced gene RTP801Safe
NCT00713518Phase IIPF-04523655 versus RanibizumabIntravitreal injectionA 19-nucleotide methylated double stranded siRNA targeting the RTP801 geneNot significantly more effective than ranbizumab, but synergetic with it in improving BCVA
NCT05657301Phase IKH631Subretinal injectionAdeno-associated virus 8 vector that encodes a human VEGF receptor fusion proteinRecruiting, no results posted
NCT05672121Phase I/IIKH631Subretinal injectionAdeno-associated virus 8 vector that encodes a human VEGF receptor fusion proteinRecruiting, no results posted
NCT05536973Phase IIADVM-022 (AAV.7m8-aflibercept)Intravitreal injectionInduction of endogenous aflibercept expression in confirmed exogenous aflibercept-responding patientsRecruiting, no results posted
NCT05197270Phase I/II4D-150Intravitreal injectionDual transgene payload, expressing aflibercept and an anti-VEGF-C RNAiRecruiting, no results posted
NCT06031727Phase IHG202Not specifiedKnockdown of Vascular Endothelial Growth Factor ARecruiting, no results posted
NCT05903794Phase IEXG102-031Not specifiedExpressing a fusion protein that is able to bind all subtypes of VEGF as well as the angiopoietin 2Recruiting, no results posted
NCT05099094Phase IIDLVIntravitreal/intracameral/subretinalIDLV vector is engineered to carry the VEGFA antibody geneRecruiting, no results posted
NCT05407636Phase IIIRGX-314Subretinal/suprachoroidalInduction of endogenous anti-VEGF proteinRecruiting, no results posted
NCT04704921Phase IIb/IIIRGX-314Subretinal/suprachoroidalInduction of endogenous anti-VEGF proteinRecruiting, no results posted
PEDF = pigment epithelium derived factor; MNV = macular neovascularization; VEGF = vascular endothelial growth factor; PU = particle units; VEGFR = vascular endothelial growth factor receptor.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Finocchio, L.; Zeppieri, M.; Gabai, A.; Toneatto, G.; Spadea, L.; Salati, C. Recent Developments in Gene Therapy for Neovascular Age-Related Macular Degeneration: A Review. Biomedicines 2023, 11, 3221. https://doi.org/10.3390/biomedicines11123221

AMA Style

Finocchio L, Zeppieri M, Gabai A, Toneatto G, Spadea L, Salati C. Recent Developments in Gene Therapy for Neovascular Age-Related Macular Degeneration: A Review. Biomedicines. 2023; 11(12):3221. https://doi.org/10.3390/biomedicines11123221

Chicago/Turabian Style

Finocchio, Lucia, Marco Zeppieri, Andrea Gabai, Giacomo Toneatto, Leopoldo Spadea, and Carlo Salati. 2023. "Recent Developments in Gene Therapy for Neovascular Age-Related Macular Degeneration: A Review" Biomedicines 11, no. 12: 3221. https://doi.org/10.3390/biomedicines11123221

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