Next Article in Journal
Polymorphism of Carbamazepine Pharmaceutical Cocrystal: Structural Analysis and Solubility Performance
Previous Article in Journal
Magnetic Nanoparticle-Based High-Performance Positive and Negative Magnetic Resonance Imaging Contrast Agents
Previous Article in Special Issue
Retinitis Pigmentosa: Novel Therapeutic Targets and Drug Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances of Ocular Drug Delivery Systems: Prominence of Ocular Implants for Chronic Eye Diseases

1
Department of Pharmaceutics, Faculty of Pharmacy, Minia University, Minya 61519, Egypt
2
Department of Pharmaceutics, College of Pharmacy, King Khalid University, Abha 62223, Saudi Arabia
3
School of Pharmacy, Kingston University London, Kingston Upon Tames KT1 2EE, UK
4
School of Pharmacy, The University of Auckland, Auckland 1010, New Zealand
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(6), 1746; https://doi.org/10.3390/pharmaceutics15061746
Submission received: 16 May 2023 / Revised: 2 June 2023 / Accepted: 13 June 2023 / Published: 15 June 2023
(This article belongs to the Special Issue Novel Therapeutic Targets and Drug Development in Retinal Diseases)

Abstract

:
Chronic ocular diseases can seriously impact the eyes and could potentially result in blindness or serious vision loss. According to the most recent data from the WHO, there are more than 2 billion visually impaired people in the world. Therefore, it is pivotal to develop more sophisticated, long-acting drug delivery systems/devices to treat chronic eye conditions. This review covers several drug delivery nanocarriers that can control chronic eye disorders non-invasively. However, most of the developed nanocarriers are still in preclinical or clinical stages. Long-acting drug delivery systems, such as inserts and implants, constitute the majority of the clinically used methods for the treatment of chronic eye diseases due to their steady state release, persistent therapeutic activity, and ability to bypass most ocular barriers. However, implants are considered invasive drug delivery technologies, especially those that are nonbiodegradable. Furthermore, in vitro characterization approaches, although useful, are limited in mimicking or truly representing the in vivo environment. This review focuses on long-acting drug delivery systems (LADDS), particularly implantable drug delivery systems (IDDS), their formulation, methods of characterization, and clinical application for the treatment of eye diseases.

Graphical Abstract

1. Introduction

Over the last three decades, the prevalence of sight-related diseases has received increased attention; this is mainly due to the increasing life expectancy of the global population. There were around 188 million people who had minor vision impairment, 216 million people who had moderate-to-severe sight impairment, and approximately 40 million people who were legally blind [1]. These numbers are only expected to increase over time. The eye has a complicated vital structure with several anatomical and physiological constraints. The anterior part of the eye, which is implicated in refraction and vision, is made up of several ocular tissues, including the cornea, conjunctiva, aqueous humor, iris, ciliary body, and the lens, whereas the back segment of the eye is mostly made up of the vitreous humor, choroid, retina, and posterior sclera. The posterior segment recognizes and transmits light signals though the optic nerve so that the eye can view the outside world. Many chronic eye diseases can affect these specialized ocular tissues.
Common conditions that affect the front of the eye include glaucoma, anterior uveitis, cataracts, and dry eye diseases [2,3], while the conditions that most often affect the back of the eye include AMD, diabetic retinopathy (DR), CMV, vitreoretinopathy, and posterior uveitis [3]. Topical eye drops supply drugs to most of anterior segment tissues, whilst eye injections (most notably, intravitreal) are the standard drug administration option for posterior segment diseases. Poor bioavailability (less than 5%) represents a major issue with topically administered ocular medications, while invasiveness (typically repeated monthly intravitreal injections) and non-compliance issues are the main hurdles to treating the diseases of the posterior segment.
The barriers to treating diseases of the anterior segment include a tight corneal–epithelial junction, reflex blinking and tearing, ocular tissue/s metabolism, tear turnover, nasolacrimal drainage, efflux transporter pumps, and the blood–aqueous barrier [4]. These anatomical and physiological constraints have been discussed in detail elsewhere [3,5]. The main barrier to medication absorption following topical application is the corneal epithelium. Tight intercellular connections surround cells that are on the surface serve as barriers to prevent drug molecules from entering the cells through the paracellular route [6]. The typical drop size of topically instilled eye drops, which is delivered to the eye, is in the volume range of 25–56 µL. Although the human eye can temporarily accommodate up to 30 µL, any excess amount is quickly wasted due to reflex blinking, greatly reducing the amount of medication that is ultimately accessible for a therapeutic effect [7,8]. P-glycoprotein and multidrug resistant proteins primarily cause drug efflux. P-glycoprotein, which is located in the blood–aqueous and blood–retinal barriers [9,10], eliminates amphipathic substances, while multidrug resistant proteins, which are found in the ciliary body and blood–aqueous barrier [11], are known to export organic anions. Endothelial cells from the blood vessels in the iris and cilia form the blood–aqueous barrier together with the non-pigmented ciliary epithelium. This prevents the bulk of medications from reaching deeper ocular tissues and controls the diffusion of soluble molecules between the front and back of the eye by building tight connections at the cellular level [12,13].
For drugs targeted at the back of the eye, the retinal pigmented epithelium, ciliary body, and ocular metabolic enzymes reduce how much of the drug remains [13,14]. The posterior segment barriers include the inner limiting membrane, vitreous diffusion, tight retinal–pigmented epithelium junctions, and the blood–retinal barrier [15]. The inner limiting membrane is a substantial physical barrier that inhibits drugs from being delivered to the posterior portion of the eye [16]. The vitreous body represents a second major barrier for drug delivery for the posterior segment. In the human eye, the vitreous body, a transparent, gel-like substance, accounts for around 80% of the total volume. The vitreous body consists of extraordinarily high water content (>97%) and collagen fibers [17,18]. The collagen fibers make up the network that fabricate the gel structure in the 3D shape and make it flexible and strong against mechanical pressures. The vitreous body can act as a barrier either physiologically or anatomically. The physiological barrier action is represented by the slowing down of drug diffusion and the anatomical barrier is represented by the 3D gel-like structure [19]. The retina and retinal pigment epithelium’s limiting structure prevents the free flow or diffusion of therapeutic drugs, which is what gives the retina its tight junctions [20]. Another significant challenge to drug delivery to the posterior portion is the blood–retinal barrier. The outer and inner blood–retinal barriers make up the blood–retinal barrier. The inner blood–retinal barrier is made up of retinal capillary endothelial cells, whereas the outside blood–retinal barrier is composed of tightly connected retinal pigment epithelial cells. Similar to the blood–brain barrier, the absence of wide gaps in the retinal pigment epithelium and retinal endothelial cells prevents passive drug transport. Only very small molecules from the choroid, including carbon dioxide, oxygen, and lipophilic compounds, can diffuse to the inner retinal tissues [21].
In this review, the most common ocular chronic disorders will be discussed. These conditions necessitate longer treatment intervals with drugs, and the most effective drug delivery systems should ideally improve the activity, stability, and distribution of drug molecules to target the ocular tissues. The utilization of long-acting drug delivery systems (LADDS), particularly implantable drug delivery systems (IDDS), and their formulation and methods of characterization, assessment, and their clinical application are covered.

2. Chronic Eye Diseases That Require Long-Acting Therapy

2.1. Dry Eye Diseases

Dry eye diseases, which are characterized by symptoms such as ocular surface irritation and vision impairment, are brought about by insufficient tear production or tear hyperosmolarity [22]. Accordingly, dry eye diseases could be classified into deficient or evaporative diseases [23]. The tear film’s osmolality rises as a result, and the ocular surface becomes inflamed [24]. According to estimates, 5–30% of adults over 50 are at risk for developing dry eye diseases [25,26,27]. Increasing evidence suggests that ocular inflammation is a major contributor to the pathophysiology of dry eye because it has demonstrated that regardless of the origin of dry eye condition, proinflammatory cytokines and T helper cells are present on the ocular surface [28].
There are various pathophysiological factors that might trigger dry eye diseases (Figure 1). The major etiological causes are ocular surface injury, meibomian gland dysfunction, and tear film hyperosmolarity and instability [29,30]. Thus, for dry eye to be defined with the greatest degree of precision and to be distinguished from other ocular surface disorders, the etiological factors are essential. In addition, the symptoms dry eye syndrome are associated with malfunction in particular brain regions [31]. In addition, gut microbiome disturbance or dysbiosis was identified to be associated with the development of dry eye, particularly primary Sjogren’s disease (Figure 1) [32].
To provide comfort to the ocular surface, tears are replenished with a variety of lubricants. These lubricants, which are called artificial tears, include several polymer solutions, such as hyaluronic acid, carboxymethyl cellulose, polyvinyl alcohol, polyethylene glycol, and polyvinyl pyrrolidone [33]. Products made of these polymers can be supplemented with additional additives to increase lubrication and prolong their duration in the eye because they do not include any physiologically bioactive molecules such as those found in real tears [34,35]. Diquafosol sodium and other aqueous secretagogues are useful for treating dry eye conditions and promoting mucin and tear production [36]. Punctal plugs, which are microscopic implants made of silicone or collagen, were initially developed to treat dry eyes by occluding the punctal duct, causing tear fluid to accumulate [37]. Topical glucocorticoid formulations have gained widespread acceptance as a temporary therapeutic option for dry eye diseases due to their well-recognized anti-inflammatory effects. Topical steroids have been demonstrated to have anti-inflammatory effects on a number of targets associated with the symptoms and signs of dry eye, such as lowering cytokine expression, maintaining the integrity of the corneal epithelium [38,39,40], and increasing tear production in animal models [41]. Topical glucocorticoid drops have been demonstrated to ameliorate symptoms and clinical indicators after a month of usage in various trials, and provided prominent lowering in the level of pro-inflammatory cytokines [40,41,42,43]. Nonsteroidal anti-inflammatory drugs were also employed topically to treat dry eye syndrome. Topical diclofenac and topical ketorolac have demonstrated enhanced effectiveness against dry eye syndrome [44,45]. Topical cyclosporine A, a typical immunomodulatory, reduces the number of T cells that are activated and the level of inflammatory markers in dry eye syndrome, as well; it controls inflammation, as well as the death of conjunctival epithelial cells [46,47,48]. Tacrolimus, which is a 10–100 times more potent immunosuppressant than cyclosporine A, is routinely used to treat dry eye diseases [49].
The continuous precorneal clearing caused by the dynamic nature of the ocular surface, together with blinking, nasolacrimal discharge and response, and basal tearing, all help to quickly remove foreign particles from the eye. Less than 20% of the applied dosage remains on the ocular surface after a single blink, providing a brief window for drug absorption (5–7 min) [50]. This is especially true when the quick turnover of tear fluid is taken into consideration. When two or more eye drops are applied at once, there is greater competition for space in the precorneal cavity, which can further reduce precorneal retention time and ocular bioavailability when treating dry eye diseases [51]. As a result, the development of better drug delivery systems might increase the efficacy of drugs used topically to treat dry eye disorders. Punctal plugs have been proven to increase the action of loaded medications in the treatment of dry eye diseases and to move beyond the ocular barriers [52,53,54]. The incorporation of mucin secretagogue rebamipide into nanocarriers significantly increased the activity and penetration into ocular tissues. The optimum use of cyclosporin A to treat dry eye disorders is hampered by its very hydrophobic nature and very poor water solubility [55,56]. Nanocarriers have been widely utilized to increase the activity, effectiveness, penetration, and duration of cyclosporin A [57]. The FDA had approved the use of several cyclosporine A drug delivery systems, such Restasis®, Ikervis®, and Cequa®, for the treatment of individuals with moderate-to-severe dry eye diseases [57,58,59]. Corticosteroids and non-steroidal anti-inflammatory drugs showed improved activity and bioavailability, reduced toxicity, and extended release upon incorporation into nanocarriers for ocular applications [60,61]. The drug delivery systems used to develop treatments for dry eye disease are summarized in Table 1.

2.2. Glaucoma

The progressive loss of retinal ganglion cells is a hallmark of the ocular neuropathy known as glaucoma [62,63]. Ganglion cell degeneration is currently untreatable, leading to a focus on slowing the disease’s development as the aim of glaucoma treatment [64]. Glaucoma is therefore seen as a chronic condition that needs ongoing management [65]. Glaucoma, a major contributor to irreversible blindness, affects over 80 million people globally today. More than 100 million people are anticipated to be afflicted by this blinding condition by the year 2040 [66,67]. The primary disease-related risk factor that can be modified is increased intraocular pressure [68]. Glaucoma can manifest in two different ways: open-angle and angle-closure. In open-angle glaucoma, the outflow channel is still accessible, but the outflow resistance is increased because of pathological alterations to the outflow tissue (Figure 2a). On the other hand, angle-closure glaucoma prevents the aqueous humor from leaving the anterior chamber of the eye because the iridocorneal angle is closed (Figure 2b).
The type of glaucoma that affects people most frequently is primary open-angle [69]. Consequently, common therapy choices include ocular drops with the goal of lowering the intraocular pressure, prostaglandin analogues, Rho-kinase inhibitors, β-adrenergic blockers, α-receptor agonists, carbonic anhydrase suppressors, and cholinergic agonists. They achieve this either by boosting the aqueous humor’s outflow through a unique method or reducing its formation [68,70].
The bioavailability of anti-glaucomatous medicines is only 1–7% inside the eye because of their short time on the corneal surface, poor corneal penetration, and quick drainage with the tear fluid [71,72]. Additionally, up to three applications each day limit the effectiveness of their therapeutic effects [72,73]. Therefore, the development of improved drug delivery systems may boost the effectiveness of medications utilized topically to successfully lower the intraocular pressure. Prostaglandin analogues, which are medications that are poorly soluble in water, are typically coupled with preservatives, such as benzalkonium chloride or polyquaternium, to make them more soluble. However, repeated instillation of these medications may irritate the ocular surface [74]. Thus, employment of nanocarriers into the formulation of prostaglandin analogues would reduce the dependence on the solubilization effect of the preservative and could lead to the development of preservative-free formulations [75]. Moreover, nanocarriers have been used effectively to enhance ocular bioavailability and therapeutic activity, as well as to reduce systemic toxicity of topically applied β blockers, α-adrenergic agonists, and carbonic anhydrase inhibitors [76,77,78]. A summary of the drug delivery systems used to enhance the therapeutic activity of anti-glaucomatous drugs is found in Table 1.

2.3. Uveitis

Uveitis is the inflammation of the uveal tract. The uveal tract, which is the middle part of the eye, is located between the retina on the inside and the sclera, conjunctiva, and anterior chamber on the outside and consists of the ciliary body, the choroid, and the iris [79]. Uveitis is considered the fourth most common reason for acquired blindness, especially for chronic uveitis, and is characterized by a high rate of related complications [80,81,82,83]. Uveitis is subcategorized according to the inflamed anatomical section into either the anterior, intermediate, or posterior, where the inflammation and accompanied leucocytes are present in the iris, vitreous humor, or choroid, respectively [79]. The concurrent presence of anterior, intermediate, and posterior uveitis is called panuveitis. Anterior uveitis, which is far more common than intermediate, posterior, or panuveitis, accounts for around 85% of all incidences of uveitis [84]. Corticosteroids, such as fluocinolone acetonide, difluprednate, fluormetholone, and triamcinolone acetonide, as well as immunomodulatory medications, including rapamycin, infliximab, and methotrexate, may be applied topically to treat anterior uveitis.
The inadequate bioavailability and the ocular tissues’ barrier properties prevent the transfer of administered medications to deeper ocular tissues, which may lead to the failure of the uveitis treatment. Consequently, the use of an efficient drug delivery system could enhance the bioavailability and improve the activity of ocularly applied corticosteroids and immunomodulatory medications [85,86]. As seen in Table 1, the use of different drug delivery systems led to the formation of more efficient treatment choices.

2.4. Endophthalmitis

The word “endophthalmitis” refers to an infection of the aqueous vitreous humors and/or the surrounding ocular tissues brought on by bacteria or fungi. Endophthalmitis is considered an uncommon eye disease; however, it may cause a severe type of inflammation and might result in irreparable vision loss. Endophthalmitis can be exogenous or endogenous depending on how the infection is transmitted to the eye. Exogenous endophthalmitis is most usually brought on by microorganisms that enter the eye through an infection in the cornea, surgery, or an eye injury. In contrast, endogenous endophthalmitis occurs when the bacteria or fungus enter the eye through the bloodstream [87]. Gram positive bacteria e.g., Staphylococcus aureus and Streptococcus species [88], as well as gram negative bacteria e.g., Klebsiella species and E. coli, are the major causes of endophthalmitis [88,89,90].
The current treatment for endophthalmitis involves repeated intravitreal injections of antimicrobial, antifungal, or antiviral agents. This procedure increases the complications and commonly results in blindness by causing irritation, ocular pain, a rise in the intraocular pressure, intraocular hemorrhage, a greater risk of retinal detachment, and retinal damage [91,92,93,94,95]. Consequently, the development of drug delivery systems for antimicrobial, antifungal, or antiviral agents could enhance ocular tissue penetration and activity noninvasively (Table 1).

2.5. Cytomegalovirus Retinitis

Cytomegalovirus (CMV) retinitis is still the most common ocular-invading virus in patients with acquired immunodeficiency syndrome (AIDS) [96,97]. Patients continue to be at risk for developing CMV retinitis predominantly as a result of either a delayed diagnosis of HIV infection or as a result of noncompliance, intolerance, or resistance to antiretroviral therapy [98]. Even though the prevalence of CMV retinitis has significantly decreased due to development of more effective treatments, CMV retinitis is still a major contributor to vision loss in AIDS patients managed with antiretroviral drugs [99]. Therefore, understanding the incidence rate and risk factors associated with the development of CMV retinitis is essential for both patients and medical professionals.
CMV retinitis could be controlled via intravitreal injection of antiviral drugs, such as ganciclovir, foscarnet, and cidofovir [100,101]. These drugs were produced in noninvasive sustained release nanocarrier formulations employing a range of drug delivery systems (Table 1).

2.6. Retinal Diseases

2.6.1. Age-Related Macular Degeneration

Age-related macular degeneration (AMD) is a condition that damages the retina’s macular area and results in a gradual loss of the central vision clarity [102,103,104]. The incidence of AMD is rising progressively with age. The percentage of AMD cases in the United States increases from 2% at age of forty to about 25% by the age of eighty [105]. Wet and dry are two different types of AMD. A persistent disorder called dry AMD often causes some degree of visual impairment and sometimes leads to complete blindness. In contrast, wet AMD only affects about 15% of AMD patients; it manifests suddenly and, if untreated, advances quickly to blindness [106,107]. When AMD is first developing in the asymptomatic early stages, the retina develops drusen, which are aggregates of insoluble extracellular lipid and protein [108]. Although AMD typically never develops without antecedent drusen development, drusen regression is connected to the progression of intermediate AMD to geographic atrophy [109,110]. Geographic atrophy, which is one late stage of dry AMD, is characterized by dispersed zones of degeneration of the overlaying light-sensitive receptors of the retina, which depend on the retinal pigment epithelium cells for alimentary maintenance [111]. Choroidal neovascularization (Figure 3, CNV), another late stage of AMD in which newly immature blood vessels sprout from the choroid toward the retina, is considered a hallmark of the wet type of AMD [111,112]. Due to lack of rigidity, fluids leak around or into the retina from these blood vessels. The late stage of AMD includes the development of neovasculature. Thus, intraocular injections of medications that target vascular endothelial growth factor (VEGF), one of the key molecules in the development of neovascularization, have been shown to be particularly effective [113,114]. Tyrosine kinase inhibitors are furthermore utilized in AMD to reduce choroidal neovascularization [115]. VEGF can activate CNV via binding to two receptors, VEGFR-1/flt-1 and VEGFR-2/KDR, both of which have intrinsic tyrosine kinase activity. Small molecule tyrosine kinase inhibitors are used to disrupt this pathway [116]. Patients with early, moderate, or atrophic AMD, however, are not eligible for any form of treatment. Additionally, there are no effective ways to stop the transition from early to advanced phases at this time [117,118].
Anti-VEGF agents are not readily able to cross the biological membranes that limit their therapeutic activity in the management of AMD [119]. Modern drug delivery systems could make the currently prescribed treatments more effective and delay this change (Table 1).

2.6.2. Diabetic Retinopathy

Diabetic retinopathy, which is a microvasculature diabetes-related problem, continues to be a predominant cause of vision loss and preventable blindness in individuals aged 20 to 74, especially in middle- and high-income nations [120]. Globally, an estimated 415 million people had diabetes in 2015, and by 2040, that figure is projected to increase to 642 million [121]. The number of those who have diabetic retinopathy and visual impairment is growing globally due to the rising incidence of diabetes and the increasing number of diabetics living longer [122]. The primary pathophysiology of diabetic retinopathy is a combination of changes brought on by hyperglycemia that leads to neovascularization (Figure 4). The neovascularization is caused by increased retinal vascular permeability, increased thickness of the retinal capillary basement membrane, inadequate blood supply to the tissues, and the release of numerous vasoactive molecules. The neovasculature is usually fragile, unstable, and leaky, causing retinal detachment and vitreous bleeding. Diabetic retinopathy combined with neovasculature is usually referred to as proliferative diabetic retinopathy, which can ultimately cause vision loss. Contrarily, the subtype of diabetic retinopathy known as non-proliferative diabetic retinopathy lacks neovascularization in the early stages. The development of microaneurysms and the minor dilatation of retinal blood vessels, which are recognized as early clinical indications of diabetic retinopathy, are common features of non-proliferative diabetic retinopathy [123,124]. The most common cause of visual loss in those with diabetic retinopathy is diabetic macular edema. In diabetic macular edema, the macula swells or thickens as a result of fluid building up sub- and intra-retinally and inside the macula as a result of the collapse of the blood–retinal barrier [125].
Antiangiogenics, steroids, anti-inflammatories, and antioxidant medications are the most popular treatments for diabetic retinopathy. However, most of these medicaments have poor ocular penetration and require implantation via surgery. Therefore, the development of more advanced drug delivery technologies may improve the potency of currently prescribed drugs, as shown in Table 1.
Table 1. Chronic eye conditions, available therapies, and drug delivery systems and their merits.
Table 1. Chronic eye conditions, available therapies, and drug delivery systems and their merits.
DiseaseTreatmentDrugDelivery System PlatformAdvantages of Delivery Systems In VivoRefs.
Dry eye syndrome Tear substitutesHypromelloseSolution [126]
Methylcellulose and derivativesSolution [127]
hyaluronic acidSolution [128]
Aqueous secretagoguesDiquafosol sodiumSolution [129]
Punctal plugsCollagen and atelocollagenIn situ hydrogelProlonged activity[49,50,130]
methacrylate-modified silk fibroinIn situ hydrogelProlonged activity[54]
Mucin secretagoguesRebamipideNanoparticlesSustained release[131]
LiposomesImproved activity[132]
MicellesImproved penetration[133]
Anti-inflammatory and immunomodulatory drugs CyclosporineMicellesImproved activity[134]
Self-nanoemulsifyingImproved efficacy[135]
LiposomesImproved activity[136]
NanoparticlesImproved activity[137]
Nano-emulsionImproved penetration[138]
Solid lipid nanoparticlesControlled release[139]
In situ hydrogelImproved activity[134]
Epigallocatechin gallateNanoparticlesExtended activity[140]
In situ gelsEnhanced efficacy[141]
LactoferrinNanoparticlesEnhanced efficacy[142]
NanocapsulesControlled release[143]
LiposomesReduced irritation[144]
Nanostructured lipid carriersControlled release[145]
Vitamin ALiposomesImproved activity[146]
TacrolimusNanoparticlesImproved penetration [147]
ProgylcosomesImproved activity[148]
MicrocrystalsImproved efficacy[149]
LiposomesImproved retention time[150]
MicellesProlonged activity[151]
NanocapsulesImproved activity[152]
CorticosteroidsDexamethasoneDendrimer Improved activity[153]
Nano-waferImproved activity[154]
Nanostructured lipid carriersImproved activity[155]
NanoparticlesImproved penetration[156]
MicellesRelease modulation[157]
NanosuspensionProlonged activity[158]
Nano emulsionImproved activity[159]
NanospongesImproved permeability[160]
FluorometholoneNanoparticlesImproved activity[161]
Triamcinolone acetonideMicellesRelease modulation[60]
NanoparticlesImproved activity[162]
HydrocortisoneNanosuspensionProlonged activity[158]
MicellesImproved targeting[163]
NanoparticlesImproved penetration[163]
NanosuspensionProlonged activity[158]
PrednisoloneNanoparticlesProlonged activity[164]
Nano capsulesReduced toxicity[165]
Lotep
rednol etabonate
NanoparticlesImproved penetration[166]
Non-steroidal anti-inflammatory drugsDiclofenac sodiumNanoparticlesImproved bioavailability[167]
NanosuspensionProlonged activity[168]
PranoprofenNanosuspensionImproved activity[169]
NanoparticlesImproved activity[169,170]
Bromfenac sodiumLiposomesExtended release[171]
NanoparticlesImproved permeation[172]
CubosomesImproved bioavailability[61]
KetorolacNanoparticlesImproved delivery[173]
lymphocyte function-associated antigen-1 antagonistsLifitegrastSolution [174,175]
Glaucoma Prostaglandin analoguesLatanoprostNanoparticlesControlled release[176]
PEGylated solid lipid Improved permeability[177]
MicellesExtended release[178]
CubosomesSustained release[179]
NanoparticlesImproved permeability[180]
TravoprostGold nanoparticlesImproved stability[181]
LiposomesSustained release[182]
SpanlasticsProlonged activity[183]
NanoemulsionImproved pharmacokinetics[75]
ImplantControlled release[184]
BimatoprostNanoparticlesImproved therapeutic activity[185]
Gold nanoparticlesControlled release[186]
Nanoparticle hydrogelControlled release[187]
MicroemulsionImproved permeability[188]
Graphene oxide-ladenControlled release[189]
ImplantsSustained release[190]
Nanovesicular systemsSustained release[191]
InsertsExtended release[192]
UnoprostoneTransscleral deviceSustained release[193]
Rho kinase inhibitorsFasudilLiposomesEnhanced bioavailability[194]
MicrospheresSustained release[195]
RipasudilSolution [196]
NetarsudilSolution [197]
β-adrenergic blockersTimololNanoparticlesExtended release[198]
MicellesExtended release[178]
CubosomesImproved bioavailability[199]
NanogelSustained release[200]
Gelatinized core liposomesImproved encapsulation[201]
MicroemulsionImproved bioavailability[202]
LevobunololNanoparticlesExtended release[203]
MicroparticlesSustained release[76]
CarteololNanocapsulesImproved activity[204]
NanoparticlesImproved activity[205]
ChitosomesImproved penetration[206]
MetipranololNanocapsulesReduced systemic side effects[207]
Betaxolol LiposomesExtended activity[208]
NanoparticlesControlled release[209]
NiosomesImproved bioavailability[210]
BilosomesImproved transcorneal permeation[211]
α-adrenergic agonistsBrimonidineNanoparticlesSustained release[212]
InsertsControlled release[213]
NiosomesSustained release[214]
MicrospheresSustained release[215]
Liposomes Improved effectiveness[216]
ImplantSustained release[217]
Gelatin-core liposomesImproved drug loading[77]
Carbonic anhydrase inhibitorsDorzolamideNanoparticlesImproved activity[218]
NanoemulsionEnhanced ocular delivery[219]
LiposomesProlonged action[78]
Microparticles Sustained release[220]
NiosomesImproved activity[221]
Implant Extended drug delivery[222]
InsertsImproved activity[223]
BrinzolamideNanoparticlesImproved therapeutic activity[224]
NanocrystalsImproved penetration[225]
LiposomesSustained release[226]
NanocapsulesImproved bioavailability[227]
Nanoemulsion Improved therapeutic efficacy[228]
NanofibersImproved patient compliance[229]
ImplantSustained release[230]
AcetazolamideCubosomesImproved therapeutic efficacy[231]
SpanlasticsEnhanced ocular delivery[232]
TransgelosomesEnhanced ocular delivery[233]
ImplantsSustained release[234]
NiosomesImproved permeability[235]
BilosomesImproved permeability[236]
MicrospongesImproved therapeutic efficacy[237]
DendrimersSustained release[238]
Cholinergic agonistsPilocarpineNanoparticlesSustained release[239]
NanocapsulesImproved bioavailability[240]
DendrimersProlonged residence time[241]
Uveitis CorticosteroidsFluocinolone acetonideImplant (Retisert®)Sustained release[242]
NanoparticlesImproved bioavailability[243]
DifluprednateMicroneedlesSustained release[244]
FluormetholoneNanoparticlesImproved penetration[245]
NanocrystalsImproved sustained activity[246]
Triamcinolone acetonideNano lipid carriersImproved penetration[247]
Immunomodulator drugsAdalimumabHydrogelImproved permeability[248]
InfliximabLiposomesProlonged activity[249]
MethotrexateImplantSustained release[250]
Sirolimus (Rapamycin)ImplantExtended release[251]
MicellesSustained release[252]
ExosomesImproved therapeutic activity[253]
LiposomesImproved therapeutic activity[86]
Endophthalmitis AntimicrobialsDaptomycinNanoparticlesNoninvasive and improved activity[254]
VancomycinNanostructured lipid carriersImproved permeability and activity[255]
NanoparticlesSustained release[256]
Thermoresponsive hydrogelsControlled release[257]
LiposomesImproved permeability[258]
ImplantControlled release[259]
NiosomesImproved permeability[260]
CeftazidimeNanoparticlesImproved activity and permeability[261]
AntifungalsAmphotericin BLiposomesImproved activity-reduce toxicity[262]
VoriconazoleThermo-sensitive in situ gelSustained release[263]
NanoparticlesImproved permeability[264]
MicroemulsionControlled release[265]
ElastosomesImproved activity and reduced toxicity[266]
MicellesImproved stability[266]
LiposomesImproved permeability[267]
AntiviralsCidofovirMicellesProlonged activity[268]
LiposomesProlonged activity[269]
FoscarnetLiposomesImproved activity and permeability[270]
GanciclovirNanoparticlesSustained release[271]
GlycerosomesSustained release[272]
MicroemulsionImproved permeability[273]
VitrasertProlonged activity[274]
MinitabletsSustained release[275]
Retinal diseases Age-related macular degenerationAnti-VEGF AgentsRanibizumab NanoparticlesImproved activity [276]
(Antibody fragment)Microparticles Improved intravitreal delivery[277]
LiposomesIncreased encapsulation-release[278]
Quantum dotsSustained release[279]
ImplantSustained release[280]
Bevacizumab NanoparticlesSustained delivery[281]
(Monoclonal antibody)Bi-layered capsuleSustained delivery[282]
NanocapsulesImproved bioavailability[283]
ImplantSustained release[284]
MicroparticlesSustained release[285]
LiposomesSustained release[286]
Aflibercept (VEGF-Trap)NanoparticlesSustained drug release[287]
MicrospheresExtended release[288]
SunitinibNanoparticlesSuperior prolonged activity[289]
MicellesExtended release[290]
AxitinibNanoparticlesSuperior activity[291]
PegaptanibPEGylated aptamerProlonged activity[113]
Gene therapyVEGF-siRNALiposomesImproved activity-stability[292]
NanoballImproved activity-targeting[293]
NanoparticlesImproved therapeutic activity[294]
Integrin antagonistsC16Y peptideNanoparticlesSustained release[295]
AntioxidantsSerine-threonine-tyrosine peptideNanoparticlesTargeting[296]
ResveratrolNanoparticlesSustained release[297]
CurcuminLiposomesImproved activity[298]
AstragalosideNanocapsulesImproved activity[299]
Diabetic retinopathyAntiangiogenicsAnti-Flt1 peptideNanoparticlesSustained release[300]
Micropump implantOn-demand targeting[301]
FenofibrateNanoparticlesControlled release[302]
PioglitazoneNanoparticlesControlled/improved activity[303]
ApatinibNanoparticlesImproved activity[304]
Silicate NanoparticlesImproved activity[305]
TacrolimusNanoparticlesImproved activity[306]
Sorafenib tosylateNanoparticlesImproved activity[307]
OctreotideNanoparticlesImproved activity-targeting[308]
Anti-inflammatory and antioxidantsp-Coumaric acidNanoparticlesImproved activity[309]
Connexin43 mimetic peptideNanoparticlesTargeting[310]
Inulin D α-tocopherol succinateNanomicellesImproved activity[311]
CiticolineLiposomesImproved permeation[312]
Melatonin NanoparticlesControlled release and enhanced tolerability[312]

3. Overview of Ocular Delivery Systems

Many disorders of the anterior region of the eye may be efficiently treated via topical administration; however, it is more challenging to target conventional therapeutic doses to the posterior of the eye in this manner. Thus, various nanocarriers have been created and investigated for the transport of drugs and genes to the anterior or the posterior portions of the eyes. The most popular nano-drug delivery systems are depicted in Figure 5, and these can be utilized to increase the activity and bioavailability, and/or lessen the toxicity of the active pharmaceutical ingredients used. Liposomes, nanoparticles, micelles, inserts, implants, hydrogel, and emulsions are some of the most frequently utilized drug delivery systems.

3.1. Liposomes

Liposomes (Figure 5) are closed vesicles made of a phospholipid bilayer that can contain both drugs that are soluble in fat [313] and those that are soluble in water [314]. Due to their biodegradability, biocompatibility, and capacity to serve as drug carriers, liposomes have been thoroughly investigated for topical ocular administration (Table 1). Liposomes are particularly useful for large molecular weight and inadequately water-soluble drugs because they promote drug permeation through ocular tissues by virtue of their superior spreading ability and rheological properties that enable prolonging the drug availability on the surface of the eye [86,315]. Liposomes’ amphiphilic lipids form a tear compound-interacting sublayer when they make contact with tear lipid components. Polar heads and tails face the polar and non-polar tear lipid components, respectively, and help distribute the medication throughout the ocular surface [316]. Extensive research has examined the merits of liposomes for ocular use, minimizing potential drug toxicity and improving their absorption and bioavailability compared to an unencapsulated drug. These medications include vancomycin, tobramycin, ganciclovir, fluconazole, brinzolamide, triamcinolone acetonide, and cyclosporine A. Drug-loaded liposomal formulations injected intravitreally have a number of benefits. Some benefits include lengthening the half-life of drugs [317], safeguarding labile compounds [318], and prolonging the time that liposomes spend in the tissues of the eye [319].

3.2. Polymeric Micelles

Amphiphilic polymers can self-assemble into different structures, known as micelles (Figure 5). The formation of micelles within the nanometer range can efficiently improve the aqueous stability and enhance cell permeability. Prior research has demonstrated that the use of a nanomicelle formulation increased medication absorption in the eye [320,321]. Nanomicelle formulations are primarily used to improve the solubility of medications with low solubility and subsequently improve their bioavailability. Hesperidin, Sirolimus, Voriconazole, and Sunitinib are only a few of the medications that were made into polymeric nanomicelles with better solubility and therapeutic efficacy (Table 1).

3.3. Polymeric Nanoparticles

Polymeric nanoparticles (Figure 5) could be produced by the use of naturally occurring or synthetically produced polymers. Chitosan, hyaluronic acid, carboxymethylcellulose sodium, albumin, and sodium alginate are some examples of natural polymers, whereas poly(lactic-co-glycolic acid), poly(-caprolactone), and poly(ethylene glycol) are examples of synthetic polymers [322]. Some retinal drugs are currently not performing as expected due to the physical and chemical properties of the medications, as well as the distinctive anatomical structure of the eye. The bioavailability of these drugs was significantly increased [323], their toxicity was reduced [324], invasive procedures could be avoided [325], and pharmacokinetic modulation was achieved [326] via incorporation into polymeric nanoparticles (Table 1). These drugs include dexamethasone, cyclosporin, latanoprost, voriconazole, and ganciclovir. Hyaluronic acid, polyethylene glycol, and chitosan are examples of mucoadhesive polymers that may be employed to alter nanoparticles to lengthen their pericorneal residence duration [327]. Moreover, mucus penetrating nanoparticles, which possess low surface tension, low viscosity, and higher hydration water content, can enhance the penetration of therapeutic medicines through the cornea, increasing their bioavailability and resulting in better pharmacologic results. Consequently, mucus penetrating nanoparticles may significantly improve the treatment of posterior ocular problems, which include posterior uveitis, CMV, and retinal disorders [328].

3.4. Solid Lipid Nanoparticles

Lipids have been used to ameliorate the limited water solubility of several lipophilic drugs and adapt them as a drug delivery system [329]. Müller and Lucks initially developed solid lipid nanoparticles (SLNs, Figure 5) in 1996, which received the attention of scientists as a popular, stable, safe, and effective nanoscale drug delivery device. A surfactant layer that surrounds a solid lipid core in SLNs stabilizes and holds the medication [330]. Drug molecules can be found mostly in the center of particles or molecularly scattered throughout the matrix, depending on the drug solubility and the drug/lipid ratio [331]. SLNs are considered an efficient system intended for ocular drug delivery. SLNs can improve corneal drug absorptivity, enhance ocular tissue penetration and bioavailability, prolong residence time, and provide extended drug release properties [332]. SLNs were efficiently used to improve the delivery of bimatoprost [185], ofloxacin [333], and dorzolamide [334], as shown in Table 1.

3.5. Hydrogels

Hydrogels (Figure 5) are produced when polymeric solutions are crosslinked to form a network. The hydrogel complexation is formed on the basis of hydrophilic interactions between the polymer tail and water molecules [335]. Hydrogels are widely employed to provide ocularly applied or injectable dosage forms to a variety of eye regions. For ocular application, there are various hydrogel formulations that have FDA approval. A hydrogel sealant called ReSure® has been authorized for use in the non-operative treatment of clear corneal incisions. Hydrogels were also used to formulate and enhance the therapeutic activity of ocularly applied drugs, such as dexamethasone [336], bevacizumab [337], and timolol [338], as shown in Table 1.

3.6. Dendrimers

Dendrimers (Figure 5) are globular, negatively, positively, or neutrally branch-like nanostructured polymers. They derive their net charge from the functional groups, which are located at the ends of their branches [339]. These molecules consist of a fundamental unit called the “core”, which comprises the major component, and side chain units called “dendrons” [340]. Drugs may be conjugated to the ligands on the dendrimer surface or may be retained in the dendrimer core. Dendrimer manufacturing, generation, surface characteristics, and conjugation technique all have an impact on the drug-loading and drug-release kinetics of dendrimers [341]. As a result of their ability to selectively target inflammatory cells while causing no harm to healthy tissue, dendrimers have proven to be a viable drug delivery vehicle for the treatment of inflammatory eye conditions. The capacity to lower medication toxicity off-target is the key advantage of dendrimers’ targeting abilities [153]. Utilizing dendrimers effectively can increase the therapeutic effectiveness of various active pharmaceutical compounds (Table 1), including pilocarpine [241], tropicamide [241], dexamethasone [342], brimonidine, and timolol [343].

3.7. Nanocrystals

Nanocrystals (Figure 5) are crystals of therapeutic drugs with particle sizes as small as a few hundred nanometers, where pure drug crystals may occasionally be stabilized by the addition of surface active agents or polymeric solutions [344]. The benefits of nanocrystals over conventional nanocarriers, such as their high drug payload and comparative ease of manufacture, make them appealing candidates for the delivery of medications that are not readily water soluble [345,346]. The preparation of therapeutic drugs in the form of nanocrystals for ocular administration has various advantages. These advantages include better tolerability, increased ocular absorption, providing intermediated and prolonged release of drugs in the eye, and improved ocular permeation [347]. They also include improved ocular safety, increased formulation retention in cul-de-sac, and enhanced ocular permeation [152]. A number of medications used ocularly have been transformed into nanocrystals (Table 1) with enhanced properties, and these include dexamethasone [348], itraconazole [349], tedizolid [350], and brinzolamide [227]. Moreover, Novartis Pharmaceutical Corporation’s formulation of nepafenac nanocrystals received approval for commercial release (FDA, 2012) under the brand name Ilevro®.

3.8. Cubosomes

Cubosomes (Figure 5) are made up of two inner aqueous pathways that are separated into two arched interpenetrating lipid bilayers, which are structured in three dimensions resembling honeycombs [351]. These pathways can be occupied by a variety of bioactive molecules, including natural bioactives, chemical pharmaceuticals, peptides, polypeptides, and proteins [352]. Cubosomes are thought to be promising delivery systems because of their special characteristics, including thermodynamic stability, bioadhesion, the capacity to encapsulate different types of drugs, and their potential to control drug release [353]. Active medicines and macromolecules can successfully be applied topically to the posterior portion of the eye using cubosomes (Table 1). These drugs include beclomethasone [352], flurbiprofen [354], timolol [199], and brimonidine [355].

3.9. Niosomes

Niosomes, which are a type of vesicular system that includes a non-ionic surfactant, are closed bilayer structures produced once the nonionic surfactants self-assemble in an aqueous media to create nanocarriers (Figure 5). Researchers have begun using niosomal systems to treat severe inflammatory diseases and conditions, such various malignancies, because of their potential to boost the bioavailability and efficiency of the encapsulated therapeutics [356]. Niosomes are being investigated more and more for improving drug delivery to both segments of the eye, anterior and posterior, as well as promoting drug penetration and retention in ocular tissues. As a consequence, niosomes showed a considerable increase in the absorption and transcorneal permeability of topically applied drugs at the ocular surface (Table 1). These drugs include cyclopentolate [357], voriconazole [358], acetazolamide [359], gentamicin [360], brinzolamide [361], pilocarpine [362], and tacrolimus [363]. Additionally, niosomes, particularly charged vesicles, have been effectively used to transfer genes by subretinal or intravitreal injection to the retinal area [364].

3.10. Emulsions

An emulsion (Figure 5) is a uniform dispersion system that is formed upon mixing two or more immiscible liquids under certain circumstances [365]. Lipid-based emulsions have become a potential vehicle for ocular medication administration. The emulsions enhance ocular delivery using one of two major strategies, either by improving ocular permeability or by lengthening the period the formulation is retained on the ocular surface [366]. Both hydrophilic and lipophilic drug types may be loaded into emulsions [367,368]. Emulsions have been successfully used to create more effective formulations for several medications used intraocularly that have increased absorption and therapeutic effectiveness. These drugs include cyclosporine A [369], coumarin-6 [370], azithromycin, and disulfiram [371].

3.11. Bilosomes

One type of vesicular drug delivery system is the bilosome (Figure 5), which is made up of non-ionic amphiphilic compounds with integrated bile salt molecules. The negatively charged bile salts serve to maintain the bilosomal structure [372]. In comparison to niosomes and liposomes, these drug carriers are more stable and can effectively increase drug absorption through biological membranes [373]. Moreover, bilosomes can enhance the permeability of polysaccharides, proteins, and polypeptides, which are poorly transported through mucosal epithelial cells [374]. Previous research studies have assessed the effectiveness of bilosomes in the administration of ocular drugs (Table 1) and found that bilosomes are well tolerated by corneal tissues [236]. These drugs include terconazole [375], acetazolamide [236], ciprofloxacin [376], ciprofloxacin [376], agomelatine [377], and betaxolol [211].

3.12. Nanocapsules

Nanocapsules (Figure 5) are a subtype of nanoparticles that are comparable to vesicular systems, in which a medicine is contained in a hollow vessel with an inner liquid core encircled by a polymeric coating [378]. Nanocapsules are well-known to be retained in the cornea for a prolonged time and to enhance penetration throughout the deep ocular tissues [152]. Thus, the development of topically applied drug-loaded nanocapsules could reduce uncomfortable intravitreal injections and systemic delivery, which have serious side effects [152]. The therapeutic action of several medications was effectively potentiated via formulation in the form of nanocapsules (Table 1). These drugs include bevacizumab [379], prednisolone [165], tacrolimus [152], brinzolamide [227], and cyclosporine [380].

3.13. Spanlastics

Elastic niosomes, also known as spanlastics (Figure 5), are a subtype of vesicular drug delivery systems that are relatively new to the market. They resemble niosomes (non-ionic surfactant vesicles), except they contain an edge activator. They were first described as systems for ocular drug delivery [381], but since then, they have been used to deliver medications to a variety of bodily organs. The spanlastics’ bilayers become more elastic and deformable when an edge activator is present, which improves drug absorption across biological membranes. Spanlastics were efficiently used to payload hydrophilic, hydrophobic, and amphiphilic therapeutic pharmaceuticals for ocular use, especially the delivery to the posterior segment (Table 1). These drugs include ketoconazole [381], cyclosporine A [382], clotrimazole [383], and vanillic acid [384].

4. Long-Acting Ocular Drug Delivery Devices

4.1. Solid Devices

Solid ocular devices are applied to the eye in a solid form and include inserts, implants, contact lenses, and films. Ocular inserts are objects that could be loaded with therapeutic drugs and inserted into the conjunctival sac for extending the duration of medicine delivery. Based on their physicochemical characteristics, the inserts are divided into three categories: bioerodible, soluble, and insoluble [385]. Soluble and erodible devices gradually dissolve while dispensing the medication and require no need for removal, while insoluble inserts can typically distribute medications at a controlled, predetermined rate through reservoir and matrix systems, but they must be removed from the eye [386]. The system prolongs drug activity, increases drug residency, improves bioavailability, and prevents crest and trough release profiles to subvert the negative effects that go along with those features [385]. Bimatoprost [387], acyclovir [388], triamcinolone acetonide [389], voriconazole [390], ketorolac [391], azithromycin [392], and dorzolamide [223] are just a few of the medications that have been delivered non-invasively to the eye using ocular inserts. A list of the commercially available long-acting drug delivery systems is shown in Table 2.
Ocular implants are solid devices that are used as medication delivery systems to slowly release molecules from polymeric matrices that are either biodegradable or not over the course of months to years. Contrary to non-biodegradable implants, which must be surgically removed after treatment, biodegradable solid implants are made utilizing biodegradable polymers, including polycaprolactones, polyglycolic acid, polylactic acid, and polylactic-co-glycolic acid, and polyanhydrides. However, these implants can have unpredictable drug release characteristics [431]. These implants can be placed at different sites in the eye, including the cameral, vitreal, scleral, episcleral, and subconjunctival areas. Implants have a number of benefits over more conventional means of administering medication to the eye, including bypassing the blood–ocular barrier and delivering a defined drug amount directly to the target site for a long period. Therefore, the danger of infection or retinal detachment may be reduced with the use of implants placed intravitreally, which may also localize therapy to the vitreous with minimal exposure to the systemic circulation [412]. In addition, implants minimize the need for repeated treatments by continuously supplying medication over a long period and are consequently suitable as a treatment of long-term eye disorders. DurasertTM is a solid polymer implant technology in which small drug molecules can be released for up to three years. Three FDA approved implants using this technology including Iluvien®, Retisert®, and Vitrasert® [432].
Drug-eluting contact lenses are solid dosage forms that have high potential to produce prolonged drug residence and close drug contact with the cornea, resulting in a major enhancement of drug bioavailability [433]. Consequently, drug-laden contact lenses provide several advantages, such as lowering the overall amount of medication required, decreasing dosing frequency, and diminishing the quantity of medication lost through systemic absorption [434]. Molecular imprinting, supercritical fluids, ion ligation, and colloidal polymeric nanoparticles are a few techniques that have been designed to payload pharmaceuticals into contact lenses [435]. Many medications have been placed into contact lenses in an effort to increase their pharmacological activity and move past the eye’s barriers, which hinder drug delivery, especially for the posterior chamber of the eye. These medications include dexamethasone [436], phomopsidione [437], latanoprost [438], a timolol–bimatoprost combination [433], and flurbiprofen [439]. Before some of these technologies can be used in clinical settings and made commercially available, a number of problems need to be resolved, including protein adhesion, diversity in the swelling ability, changes in water content, opacity, surface integrity, strength properties, ion and oxygen permeation, and drug leakage during manufacturing and storage.
Ocular films are solid sterile dosage forms that are applied topically on the eye sac to improve ocular bioavailability and remove barriers to ocular drug delivery [440]. The use of ocular films improves therapeutic efficacy, reduces systemic adverse effects, and minimizes dose frequency. In order to maximize the therapeutic response and patient compliance, ocular films could present intriguing prospects as a vehicle for the administration of therapies. They could thus replace the conventional dosage forms. However, the designation of efficient films for the ocular delivery of therapeutic medications depends on a thorough understanding of the medication, the restrictions of drug permeation to ocular tissues, and the excipients employed. The construction of ocular films should therefore take into account each of these elements. In an effort to maximize their therapeutic action, a plethora of medications were administered as ocular films, including acetazolamide [441], timolol maleate [440], ofloxacin [442], fluconazole [442], and dorzolamide hydrochloride [443].

4.2. Microneedles

Microneedles are structures of a metallic or polymeric nature that range in size from a few to 200 µm. Microneedles contain tiny protrusions, which reduces their degree of invasion. There are several microneedle subtypes with a variety of pharmaceutical purposes; however, just three microneedle subtypes play a substantial role on drug delivery to ocular tissues. These subtypes include solid coated, hollow, and microneedles of dissolving polymers, as shown in Figure 6 [444].
Microneedles with solid coatings are the type that can be used to pierce tissue and the coating instantly disintegrates. They can then be removed. The perforation will create a channel with a diameter of a few microns that will effectively distribute the drug [445]. Consequently, the main goal of solid microneedles is to increase the porosity of the cornea or sclera of the eye (Figure 6a). Metals such as stainless steel and materials such as silicon probes are employed in the manufacture of microneedles. These materials’ non-biodegradability and complexity in production make them undesirable for use in ocular delivery [446]. Coated microneedles have been used successfully to improve the effectiveness of loaded medications in a variety of eye conditions. Pilocarpine, a medication used to treat glaucoma, has shown improved absorption when loaded onto coated microneedles [447]. The anti-VEGF drug, bevacizumab, provided customized medicament delivery to the corneal stroma and a potential impact with fewer adverse effects to treat corneal neovascularization when loaded into coated microneedles [448].
Micron-sized hollow microneedles are needles with the drug formulation entirely inside the needles. The loading of microneedles with drug delivery systems could potentially improve therapeutic drug activity [449]. Microneedles were efficiently loaded with nanoparticles, nano-emulsions, liposomes, and microparticles [450]. Hollow microneedles are mainly composed of borosilicate; however, stainless steel might be used for their manufacturing. The method for administering medication involves puncturing the ocular tissue, after which the drug will leak from the microneedles’ hollow spaces (Figure 6b) [449]. Several medicaments were included in hollow microneedles with improved therapeutic activity. Triamcinolone acetonide was injected into the suprachoroidal area in hollow microneedles to efficiently manage posterior acute uveitis. The utilization of the microneedles helped to successfully alleviate posterior uveitis for up to three days with minimal invasion and without affecting the retina or raising the intraocular pressure [451].
Dissolving polymeric microneedles have been developed as a solution to the many drawbacks of hollow and solid coated microneedles, such as their manufacture, application, and reliability. They have demonstrated their compatibility with ocular tissue in comparison to their hollow and solid coated counterparts [452]. They are made of several biocompatible and biodegradable polymers that are simple to implant into the ocular tissue. After applying the polymeric microneedles to the eye tissue, the medication, which has already been loaded into the polymeric matrix, is released into the eye tissue (Figure 6c) [453]. Dissolving polymeric microneedles have significantly improved the therapeutic action of several drugs. The poorly soluble medication amphotericin B was effectively used with dissolving polymeric microneedles to increase its antifungal effectiveness by enabling rapid dissolution, excellent tissue penetration, low toxicity, and long-lasting therapeutic action [454]. By incorporating cyclosporin A into polymeric microneedles that dissolve, a high molecular weight medication with weak water solubility was effectively delivered to the ocular tissue with improved drug permeability and activity [455].

4.3. Three-Dimensional Printable Systems

In recent years, experts have predicted that 3D printing will revolutionize the pharmaceutical industry since it can generate specific doses of individualized medications with novel designs [456], drug mixtures [457], and targeted drug release properties [458]. Additionally, 3D printing could be employed for the development of highly accurate, individualized medical instruments [459]. Over the past 10 years, 3D printing has been heavily utilized in the fields of contact lens manufacturing, drug delivery to ocular tissues, implants, ocular research, and diagnostic models production [460].
Ocular prostheses, which aid ophthalmic patients in restoring the symmetry of their face, were successfully developed throughout 3D bioprinting technology with minimal cytotoxic effects. The 3D-printed prosthesis showed no negative effects on the conjunctival sac or membrane and provided the best resemblance to the look of a human eye, including iris color, sclera, and vascular structures [461]. Ocuserts made by 3D printing were also used to modulate the pharmacokinetics of ganciclovir-loaded glycerosomes, resulting in prolonged release, enhanced tissue penetration, and therapeutic potential [272].
Three-dimensional (3D) printing was additionally incorporated into the development of prosthetic corneal structures in an effort to bypass religious restrictions and drug histories. Artificial corneas created by 3D printing technology were proven a reliable, quick, convenient, and useful choice [462,463]. Gelatin, collagen, polyvinyl alcohol, and sodium alginate are the primary materials used to create the 3D-engineered corneas because they are biodegradable, translucent, permeable to oxygen and nutrients, able to endure shear stress, and sufficiently robust mechanically [464].
The development of artificial retinas is essential for the design of more efficient systems for drug delivery, research into disease causes, and the development of cutting-edge therapeutic choices. Artificial retinas with the best cytocompatibility were created via 3D bioprinting, simulating the natural structure of the human retina [465]. Moreover, human retinal progenitor cells were effectively maintained in vitro by the use of 3D-printed polymeric scaffolds. The subretinal implantation of the cell-free scaffolds into retinitis pigmentosa porcine models did not result in inflammation, infections, or cytotoxicity, supporting the possibility that they may be used in preclinical studies [466].
Dexamethasone-loaded punctal plugs created by 3D printing demonstrated sustained drug release for 1 to 3 weeks, depending on the particular polymer or blends of polymers chosen [37]. Ocular 3D-printed patches have successfully been designed to hold various pharmacological active constituents, and they may be adjusted to release varying amounts based on the patient’s demands [467]. Timolol maleate-loaded 3D-printed contact lenses were successfully utilized to treat glaucoma in patients who did not take their prescribed glaucoma medications [468]. The lenses had a smooth surface with high printing quality and released timolol maleate steadily over a period of three days [468].
Three-dimensional micro-stereolithography has been enrolled in the production of therapeutic devices for controlling intraocular pressure and, consequently, glaucoma. It combines the advantages of both digital light processing and stereolithography technologies. Over the past ten years, minimally invasive glaucoma devices have been designed to boost aqueous humor discharge in an effort to control glaucoma [469]. With the use of 3D printing techniques, a complex surgical device can be produced with significant flexibility while maintaining functionality [470].

4.4. In Situ Gelling Implants

Drug implants that are generated when certain conditions are fulfilled are known as in situ forming implants. They are currently quite popular since they do not require regular injections into the eye or insertion via surgery. In situ forming implants are administered as low viscosity solutions that solidify as depots or implants at the targeted site, controlling the administration of drugs [471]. According to the underlying phase separation process, numerous parameters might be used to influence the sol-to-gel transition [472]. In situ gelling implants are advantageous since they are simple to use, very stable, have an optimized drug release profile, and do not require complicated equipment for ocular injection [191].
In situ gelling implants were used to formulate a number of ocularly administered drugs in an effort to increase their therapeutic action, facilitate administration, extend disease management, and boost patient compliance. Moreover, in situ gelling may offer a good substitute for the currently available therapies. Triamcinolone acetonide was successfully developed as a gelling implant with extended drug release that met acceptable rheological and syringe ability standards. Triamcinolone acetonide was maintained by the formulation for a course of six weeks [473]. Bimatoprost was effectively combined into in situ gelling implant for subconjunctival injection with higher stability, cheap cost, improved solubility, and ease of processing using nano-vesicular systems. With just one injection, the newly developed formulation may maintain intraocular pressure for up to 8 weeks [191]. Additionally, peptides were loaded onto polymeric nanoparticles in an effort to lessen the burst release. In situ (light-sensitive) gelling implants with the peptide-loaded nanoparticles were used to deliver the medication to the posterior eye in a sustained and effective manner [474,475].
There are currently no in situ gelling implants available for use in the eyes. However, two formulations, including bevacizumab, are in the final stages of approval. The first is a photosensitive in situ gelling implant (OcuLiefTM), while the second is a premade photosensitive implant (EyeLiefTM). Both of these medicines were developed by the Re-Vana Therapeutics corporation [476].
A list of the commercially available long-acting drug delivery systems and devices is shown in Table 2.

5. Implantable Systems/Devices for Drug Delivery

Many chronic ocular illnesses necessitate the use of implanted drug-delivery systems or devices (IDDS) for management or therapy. IDDS are made to be implanted in order to regulate the drug efflux and, as a result, lengthen the time that the disease condition is under control. IDDS have significant benefits over conventional systemic administration. Higher medicament concentrations in the intended locations can be achieved via site-specific implantation, which can avoid oral absorption and distribution phases [477]. Additionally, IDDS increases patient compliance, minimizes parenteral treatment pain, and sustains the drug concentration in the therapeutic window by a continuous controlled release of the loaded medication [478]. As a result, IDDS were successfully used in the production of a number of authorized marketed medications to control a variety of chronic diseases, including eye chronic disorders, which include glaucoma, uveitis, endophthalmitis, dry eye diseases, AMD, and diabetic retinopathy. These products include Ozurdex® (Allergan Co., Ltd.), Retisert® (Bausch&Lomb), Vitrasert® (Bausch&Lomb), and I-vation® (Surmodics Inc.). The technologies or techniques used to generate IDDS and characterize these products are discussed in the following sections.

5.1. Polymers Used to Formulate IDDS

The choice of polymer is essential for adjusting the release profile of IDDS. Polymers used for intraocular IDDS might be biodegradable or nonbiodegradable. In the next section, we will discuss the polymers often used to formulate ocular IDDS.

5.1.1. Nonbiodegradable Polymers

The virtue of nonbiodegradable polymers, which are used to formulate nonbiodegradable IDDS, is that they may achieve very long-term release and have high biocompatibility [479]. On the other hand, the matrix polymer needs to be surgically removed after drug exhaustion. These polymers include EVA, polyimide, polyethylene terephthalate, and silicones. Several intraocular IDDS are commercially available, including Retisert®, Vitrasert®, Iluvien®, and Renexus®.

5.1.2. Biodegradable Polymers

Biodegradable polymers have the benefits of degrading once implanted into biological tissues. However, the type of polymer and degree of crosslinking greatly affect the degradation dynamics [421]. Biodegradable polymers include PLGA, polycaprolactone, and acetyl triethyl citrate. These polymers were successfully employed to develop several commercially available implants, including Ozurdex®, Posurdex®, Durysta®, and Dexycu®.
Table 2 outlines several commercially available IDDS with their polymeric composition.

5.2. Techniques for the Preparation of IDDS

5.2.1. Solvent Casting

For the production of polymeric inserts and implants, solvent casting is an efficient and scalable technique. Various experimental conditions, such as heating and lyophilization, were used to produce and cast polymeric solutions containing drug(s) and plasticizer(s). The type of drug loaded, as well as its thermal stability, play a major role in the choice of the condition. In an effort to improve the stability and therapeutic activity and prolong residency, this approach was used to formulate inserts or implants for several therapeutic medications that had received approval for ocular use. These drugs include dexamethasone [480], acetazolamide [234], bimatoprost [387], etoposide [481], and dorzolamide [482]. Table 3 displays a list of the FDA-approved polymers or copolymers used in ocular preparations. Figure 7a displays a schematic illustration showing the solvent casting process.

5.2.2. Extrusion

The hot melt extrusion method involves forcing a polymer(s) through a mold after it melts or softens at higher temperatures, often with the use of a conveyor system that leads into a tube. The actual process may be split into a number of processes, including heating the polymer mixture, loading, blending, transporting, allowing it to flow through the die, and downstream material processing (Figure 7b) [483]. Controlling each of these stages will ultimately affect the final features of the product [484]. In the hot melting extrusion method, a number of polymers were used, including aliphatic polyesters, poly (ortho esters), polyurethanes, polyvinyl lactams, ethylene-co-vinyl acetate, polyanhydrides, polyacrylics, polyethylene glycol, and polyethylene oxide [483]. In addition, several FDA-approved ocular inserts or implants were developed using the hot melt extrusion technique, including Lacrisert® and Ozurdex® [411,424].

5.2.3. Electrospinning

Electrospun inserts and implants are automatically generated utilizing a system that includes a syringe pump, collector electrode, and high voltage generator (Figure 7c). Several factors may have an impact on the manufactured inserts or implants, including the polymeric solution pump rate, the distance between the syringe tip and collection electrode, and the applied volage [485]. Electrospinning became popular due to its benefits, including simple control of the shape, diameter, surface properties, and porosity, and the simplicity of achieving nanosized inserts/implants [486]. Moreover, electrospinning enables the administration of many medicaments at once.

5.2.4. Other Techniques

Several other techniques might be employed in the development of inserts or implants, including 3D printing, hot isostatic pressing, selective laser melting, and the creation of in situ systems.
Over the past 10 years, printing throughout three dimensions has been frequently employed in the production of implants or inserts [487]. Three-dimensional printing includes the development of inserts or implants by polymer deposition in a layer-by-layer manner [488]. The pharmaceutical industry has lately boosted its usage of 3D printing due to its capacity to produce unique, individualized, and complicated dosage forms and medical equipment [489,490]. A triamcinolone acetonide sustained release implant with great clinical promise was produced as a result of a successful 3D printing application [491]. Further, 3D technology makes it possible to manage the features of the produced implants, including their form, size, and dosage, and to provide customization based on the patient’s clinical situations [491]. Hot melt extrusion coupled with 3D-printed fused deposition were effectively used to generate ciprofloxacin-loaded ocular inserts that have improved therapeutic results for treating ocular pathogenic infections and sustained antibacterial activity [492].
In the industrial process known as hot isostatic pressing, components or powders are heated to a high temperature while also being compressed in a pressurized cylinder [493]. Metal-based implants made of titanium [494] and stainless steel [495] are frequently produced via hot isostatic pressing.
Selective laser melting primarily relies on the employment of a high intensity laser beam to fuse the powder that is present in its focus zone and enable the manufacturing of items layer-by-layer from a 3D computer-assisted design [496]. The production of inserts or implants is now regarded to be a viable application for selective laser melting [496]. Selective laser melting makes it feasible to generate implants that have a crooked structure, which was previously not conceivable commercially.
In situ forming implants are solutions that go through phase separation to produce a drug depot formulation. Crosslinking, solidifying, and phase separation are some of the in situ gelling systems’ mechanisms [472]. In situ crosslinking of polymers could be initiated chemically, physically, or through photosensitization, while solidifying organogels initiate in situ gelling through solubility alteration. In situ gelling through phase separation systems could be triggered by a pH change, temperature change, or solvent exchange [472]. Drug-loaded in situ ocular inserts or implants provide benefits in terms of better therapeutic action, increased stability, simplicity of administration, and control over drug release [191,497].

5.3. Characterization and Evaluation of IDDS

In vitro testing and characterization of drug delivery systems or devices is a key element in pharmaceutical development’s quality control process for evaluating and determining the best formulation(s). One of the most crucial characterization criteria, in vitro testing for dissolution, is utilized to develop in vitro–in vivo relationships, which aid formulation marketing and reduce medication costs. However, the idea of developing in vitro–in vivo correlations becomes more difficult for IDDS, owing to the sophistication of ocular physiological conditions, ocular tissue barriers, and the insertion site of the implant. Therefore, the design of practical in vitro testing for drug release and dissolution from IDDS remains challenging. Several in vitro simulation experiments have been developed to mimic the in vivo insertion of IDDS in the eye tissue.
The static diffusion system (Figure 8a) was developed to investigate the in vitro efflux rate of drugs formulated as IDDS. In this system, the appropriate release media is chosen and directly incubated with the IDDS and kept at a standard temperature with or without mechanical agitation [498]. The amount of medication released is then measured at predetermined intervals of time. Despite the static diffusion method’s widespread use, it was restricted in its capacity to investigate IDDS because it lacked ocular flow modelling and the capacity to control diffusion layers [499].
The agar diffusion system was also adopted to assess the drug release from IDDS under very viscous conditions [500]. The procedures involved inserting IDDS into agar gel and, using the proper analytical technique, the gel was evaluated for the amount of drug dispersed at predetermined time intervals [500]. Figure 8b displays a schematic illustration of the procedures involved. However, the implants are made to be inserted in certain environments; this technique does not accurately reflect such environments. Additionally, this method only relies on a diffusion mechanism to control the drug outflow from the implant, avoiding any potential impact from the actual vitreous environment [499].
The dialysis bags system is the most straightforward method for predicting the in vitro dissolution and release of therapeutic drugs from IDDS. This technique employed a dialysis bag, which was closed on both sides once the implant was inserted (Figure 8c). Drug molecules should be able to pass through the specified molecular weight cutoff for the dialysis bag. After that, the dialysis bag is placed in the release medium solution, which is constantly agitated at standard temperature. At regular intervals, samples from the release media were obtained, analyzed, and quantified [501]. The Franz diffusion cells or modified Franz diffusion cells with a modified curved donor compartment to accommodate the curvature of the excised corneal tissues operates with the same principle of the dialysis bag, but with a more consistent and reproducible surface area for drug diffusion; they have the capacity to hold ocular tissues.
The pharmacokinetic eye model is a more complex system that simulates drug clearance via the anterior chamber, including intraocular aqueous outflow. The apparatus has two compartments that are partitioned by a dialysis membrane that simulates the posterior and anterior ocular chambers. It is hypothesized that this model may also be used to determine how much of the drug would be released from IDDS that are placed in the cavity of the vitreous [502]. The device was designed to mimic the actual insertion operation of IDDS into the eye. Both the injection and the aqueous inflow ports were positioned inside the replicating vitreous cavity, while a single output port was positioned in the simulating anterior chamber. The device proved highly effective in evaluating the in vitro release studies of various commercially available medicines, including Kenalog®, Avastin®, and Lucentis®.
The eye movement system model was created as an in vitro simulation system to imitate the vitreous body, as well as environmental stimuli that move the eyes, such as head movement [503]. IDDS were inserted inside the chamber, which imitates eye and head movements. The release medium is refreshed every 24 h, and the drug concentration is assessed using the proper analytical technique. The method may demonstrate how the vitreous body’s gelled region, together with conscious motions of the head and eyes, influence the release of produced IDDS.
Despite the wide advancements of in vitro testing, there is still no in vitro experimental design that accurately mimics the factors that determine release in an in vivo setting. This happens as a result of the drug’s distribution and penetration process in the eye being more difficult to simulate than with other routes. Furthermore, it would be unethical to repeatedly monitor drug levels in a living eye in order to demonstrate an in vitro–in vivo association relationship.

5.4. Sites for Delivery and Implantation

The selection of an implant location is contingent on the required pharmacokinetics, biocompatibility, and clinical factors. A close proximity between the insertion site and the target tissue allows for a high drug concentration in the target tissue. The most popular sites for implantation include intravitreal [419], intracameral [420], and subconjunctival injections [504]. The intravitreal injection is widely used to deliver several commercially available corticosteroids, anti-VEGF agents, antivirals, and encapsulated cells (Table 2). Extensive research was conducted on intracameral injection during the past decade, and as a consequence, the FDA granted approval to the first intracamerally injected implant, Durysta, in the year 2020 (Table 2) [505]. Subconjunctival injection is considered one of the most effective approaches to deliver several medications to the vitreous and retinal area at higher levels [506].

5.5. Regulatory Aspects of IDDS

The FDA classifies IDDS as either class II or class III medical devices, which, respectively, denote intermediate and notably higher risk levels, because of their direct and persistent contact with the living tissues [507]. In order for IDDS to be approved by the FDA and marketed in the USA, it needs to obtain the pre-market notification 510(K). A 510(K) is a premarket application submitted to the FDA to prove that the product being marketed is essentially identical to, or equally safe and effective to, a product that has previously gained FDA approval [508]. If there is no comparable product on the market, the innovative device must receive pre-market approval with sufficient reliable scientific data that must prove that it is effective and safe for the intended usage(s) [509].
For medical device manufacturers to follow guidelines while designing, producing, packing, and distributing their products, the FDA introduced “Design Control Guidance for Medical Device Manufacturers”. The FDA periodically inspects manufacturers to ensure that they adhere to the required good manufacturing practice requirements [510]. For IDDS approval, further laboratory tests, including those for sterility, biocompatibility, and material characterization, are required.
Sterilization assures patient safety during implantation procedures via the lack of live microorganisms on the device. The FDA recommends terminal sterilization using either ethylene oxide or gamma radiation. A vital component of good terminal sterilization is the packaging mechanism, which must permit gas penetration and radiation to reach the biomaterial. The FDA’s primary criteria are equipment validation, microbiological testing, and sterilization testing [511].
The chosen material must also be biocompatible and should not result in any undesirable unfavorable biological reactions when in contact with the human body. The material’s biocompatibility must be verified with tests for cytotoxicity, hemocompatibility, pyrogenicity, sensitization, genotoxicity, and carcinogenicity [511].
The physical, chemical, and mechanical characteristics should be determined for biomaterials allowed to generate IDDS. The pore size, pore size distribution, structure, and connectivity are examples of physical characteristics. The potential for toxicity, carcinogenicity, and immunogenicity are all factors of chemical characterization, and compressibility and mechanical strength are examples of mechanical properties [512].

6. Conclusions and Future Prospective

The use of IDDS is advantageous for the management of a number of ocular chronic disorders, including glaucoma, uveitis, endophthalmitis, and retinal disease, over the traditional ocular dosage forms. Less frequent administration, sustained and local action, bypassing several ocular barriers, and prolonged pharmacological impact are some of the positive characteristics of IDDS [386]. However, IDDS suffer several limitations that affect their pharmacological activity. IDDS are considered an invasive technique for ocular drug delivery. Additionally, some IDDS that are not biodegradable need to be surgically removed at the end of the treatment period, which has an impact on patient compliance. Likewise, IDDS were made to release the loaded drug over the course of treatment at a fixed value without being affected by environmental factors. A change in the medication release profile could be necessary, though, due to fluctuations in the course of the disease, how it responds to therapy, and other disorders. The key obstacle to the successful implementation of adjustable delivery ocular implantable drug delivery systems/devices is still related to their size limitation, which necessitates the use of extremely potent medications to accomplish long-term release. While this succeeds well with most steroids, it may cause issues with certain larger biomolecules. Additionally, the expense of therapy is greatly increased by injection and retrieval procedures for currently marketed IDDS.
This review has identified the most prevalent ocular chronic disorders that require longer treatment durations with their therapeutic drugs and the most advanced systems for drug delivery, which might be able to boost the activity, stability, and penetration of these pharmaceuticals throughout the ocular tissue. The enrollment of drugs into advanced systems for drug delivery may be sufficient to surmount all the impediments that stand in the way of drug activity. This approach is also more cost-effective than creating more effective drug molecules with desirable properties. This review also concentrated on the use of long-acting drug delivery systems, particularly IDDS, and their production processes, techniques for characterization, and assessment, as well as the legal and ethical issues of their clinical implication.

Author Contributions

M.M. Original draft writing; A.A.F. Revision and editing; R.G.A. Revision and editing, data curation, writing and editing; H.A. Conceptualization, revision, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research at King Khalid University through the Large Group Research Project under grant number RGP 2/55/44.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available from the corresponding author upon request.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through Large Group Research Project under grant number RGP 2/55/44.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bourne, R.R.A.; Flaxman, S.R.; Braithwaite, T.; Cicinelli, M.V.; Das, A.; Jonas, J.B.; Keeffe, J.; Kempen, J.H.; Leasher, J.; Limburg, H.; et al. Magnitude, temporal trends, and projections of the global prevalence of blindness and distance and near vision impairment: A systematic review and meta-analysis. Lancet Glob. Health 2017, 5, e888–e897. [Google Scholar] [CrossRef] [Green Version]
  2. Patel, A.; Cholkar, K.; Agrahari, V.; Mitra, A.K. Ocular drug delivery systems: An overview. World J. Pharmacol. 2013, 2, 47–64. [Google Scholar] [CrossRef]
  3. Suri, R.; Beg, S.; Kohli, K. Target strategies for drug delivery bypassing ocular barriers. J. Drug Deliv. Sci. Technol. 2020, 55, 101389. [Google Scholar] [CrossRef]
  4. Bachu, R.D.; Chowdhury, P.; Al-Saedi, Z.H.F.; Karla, P.K.; Boddu, S.H.S. Ocular Drug Delivery Barriers—Role of Nanocarriers in the Treatment of Anterior Segment Ocular Diseases. Pharmaceutics 2018, 10, 28. [Google Scholar] [CrossRef] [Green Version]
  5. Abdelkader, H.; G Alany, R. Controlled and Continuous Release Ocular Drug Delivery Systems: Pros and Cons. Curr. Drug Deliv. 2012, 9, 421–430. [Google Scholar] [CrossRef] [PubMed]
  6. Gumbiner, B. Structure, biochemistry, and assembly of epithelial tight junctions. Am. J. Physiol.-Cell Physiol. 1987, 253, C749–C758. [Google Scholar] [CrossRef]
  7. Gaudana, R.; Ananthula, H.K.; Parenky, A.; Mitra, A.K. Ocular Drug Delivery. AAPS J. 2010, 12, 348–360. [Google Scholar] [CrossRef]
  8. Hovanesian, J.; Singh, I.P.; Bauskar, A.; Vantipalli, S.; Ozden, R.G.; Goldstein, M.H. Identifying and addressing common contributors to nonadherence with ophthalmic medical therapy. Curr. Opin. Ophthalmol. 2023, 34, S1–S13. [Google Scholar] [CrossRef] [PubMed]
  9. Constable, P.A.; Lawrenson, J.G.; Dolman, D.E.M.; Arden, G.B.; Abbott, N.J. P-Glycoprotein expression in human retinal pigment epithelium cell lines. Exp. Eye Res. 2006, 83, 24–30. [Google Scholar] [CrossRef] [PubMed]
  10. Kajikawa, T.; Mishima, H.; Mishima, H.; Murakami, T.; Takano, M. Role of P-glycoprotein in distribution of rhodamine 123 into aqueous humor in rabbits. Curr. Eye Res. 1999, 18, 240–246. [Google Scholar] [CrossRef]
  11. Pelis, R.M.; Shahidullah, M.; Ghosh, S.; Coca-Prados, M.; Wright, S.H.; Delamere, N.A. Localization of Multidrug Resistance-Associated Protein 2 in the Nonpigmented Ciliary Epithelium of the Eye. J. Pharmacol. Exp. Ther. 2009, 329, 479–485. [Google Scholar] [CrossRef] [Green Version]
  12. Freddo, T.F. Shifting the Paradigm of the Blood–Aqueous Barrier. Exp. Eye Res. 2001, 73, 581–592. [Google Scholar] [CrossRef]
  13. Maulvi, F.A.; Shetty, K.H.; Desai, D.T.; Shah, D.O.; Willcox, M.D.P. Recent advances in ophthalmic preparations: Ocular barriers, dosage forms and routes of administration. Int. J. Pharm. 2021, 608, 121105. [Google Scholar] [CrossRef] [PubMed]
  14. Das, N.D.; Shichi, H. Enzymes of mercapturate synthesis and other drug-metabolizing reactions-specific localization in the eye. Exp. Eye Res. 1981, 33, 525–533. [Google Scholar] [CrossRef] [PubMed]
  15. Thrimawithana, T.R.; Young, S.; Bunt, C.R.; Green, C.; Alany, R.G. Drug delivery to the posterior segment of the eye. Drug Discov. Today 2011, 16, 270–277. [Google Scholar] [CrossRef] [PubMed]
  16. Dalkara, D.; Kolstad, K.D.; Caporale, N.; Visel, M.; Klimczak, R.R.; Schaffer, D.V.; Flannery, J.G. Inner Limiting Membrane Barriers to AAV-mediated Retinal Transduction From the Vitreous. Mol. Ther. 2009, 17, 2096–2102. [Google Scholar] [CrossRef] [Green Version]
  17. Mains, J.; Wilson, C.G. The vitreous humor as a barrier to nanoparticle distribution. J. Ocul. Pharmacol. Ther. 2013, 29, 143–150. [Google Scholar] [CrossRef]
  18. Halfter, W.; Winzen, U.; Bishop, P.N.; Eller, A. Regulation of Eye Size by the Retinal Basement Membrane and Vitreous Body. Investig. Ophthalmol. Vis. Sci. 2006, 47, 3586–3594. [Google Scholar] [CrossRef] [Green Version]
  19. Sakamoto, T.; Ueno, H.; Goto, Y.; Oshima, Y.; Ishibashi, T.; Inomata, H. A vitrectomy improves the transfection efficiency of adenoviral vector-mediated gene transfer to Müller cells. Gene Ther. 1998, 5, 1088–1097. [Google Scholar] [CrossRef] [Green Version]
  20. Jackson, T.L.; Antcliff, R.J.; Hillenkamp, J.; Marshall, J. Human Retinal Molecular Weight Exclusion Limit and Estimate of Species Variation. Investig. Ophthalmol. Vis. Sci. 2003, 44, 2141–2146. [Google Scholar] [CrossRef] [Green Version]
  21. Matet, A.; Behar-Cohen, F.; Cassoux, N.; Declèves, X.; Cisternino, S. Chapter 10-Retinal and choroidal cancers: Blood-retinal barriers considerations in ocular chemotherapy. In Drug Efflux Pumps in Cancer Resistance Pathways: From Molecular Recognition and Characterization to Possible Inhibition Strategies in Chemotherapy; Sosnik, A., Bendayan, R., Eds.; Academic Press: Cambridge, MA, USA, 2020; Volume 7, pp. 303–335. [Google Scholar]
  22. Verjee, M.A.; Brissette, A.R.; Starr, C.E. Dry Eye Disease: Early Recognition with Guidance on Management and Treatment for Primary Care Family Physicians. Ophthalmol. Ther. 2020, 9, 877–888. [Google Scholar] [CrossRef] [PubMed]
  23. The Definition and Classification of Dry Eye Disease: Report of the Definition and Classification Subcommittee of the International Dry Eye Workshop (2007). Ocul. Surf. 2007, 5, 75–92. [CrossRef] [PubMed]
  24. Navarro-Lopez, S.; Moya-Ramón, M.; Gallar, J.; Carracedo, G.; Aracil-Marco, A. Effects of physical activity/exercise on tear film characteristics and dry eye associated symptoms: A literature review. Contact Lens Anterior Eye 2023, 43, 101854. [Google Scholar] [CrossRef] [PubMed]
  25. Johnson, M.E.; Murphy, P.J. Changes in the tear film and ocular surface from dry eye syndrome. Prog. Retin. Eye Res. 2004, 23, 449–474. [Google Scholar] [CrossRef]
  26. Messmer, E.M. The pathophysiology, diagnosis, and treatment of dry eye disease. Dtsch. Arztebl. Int. 2015, 112, 71–81, quiz 82. [Google Scholar] [CrossRef] [Green Version]
  27. Vehof, J.; Snieder, H.; Jansonius, N.; Hammond, C.J. Prevalence and risk factors of dry eye in 79,866 participants of the population-based Lifelines cohort study in the Netherlands. Ocul. Surf. 2021, 19, 83–93. [Google Scholar] [CrossRef]
  28. Bron, A.J.; de Paiva, C.S.; Chauhan, S.K.; Bonini, S.; Gabison, E.E.; Jain, S.; Knop, E.; Markoulli, M.; Ogawa, Y.; Perez, V.; et al. TFOS DEWS II pathophysiology report. Ocul. Surf. 2017, 15, 438–510. [Google Scholar] [CrossRef]
  29. Craig, J.P.; Nichols, K.K.; Akpek, E.K.; Caffery, B.; Dua, H.S.; Joo, C.-K.; Liu, Z.; Nelson, J.D.; Nichols, J.J.; Tsubota, K.; et al. TFOS DEWS II Definition and Classification Report. Ocul. Surf. 2017, 15, 276–283. [Google Scholar] [CrossRef]
  30. Periman, L.M.; Perez, V.L.; Saban, D.R.; Lin, M.C.; Neri, P. The Immunological Basis of Dry Eye Disease and Current Topical Treatment Options. J. Ocul. Pharmacol. Ther. 2020, 36, 137–146. [Google Scholar] [CrossRef] [Green Version]
  31. Sun, T.; Shu, H.-Y.; Wu, J.-L.; Su, T.; Liu, Y.-J.; Zhang, L.-J.; Li, Q.-Y.; Pan, Y.-C.; Ge, Q.-M.; Shao, Y. Investigation of changes in the activity and function of dry eye-associated brain regions using the amplitude of low-frequency fluctuations method. Biosci. Rep. 2022, 42, BSR20210941. [Google Scholar] [CrossRef] [PubMed]
  32. Yao, Y.F.; Wang, M.Y.; Dou, X.Y. Gastrointestinal microbiome and primary Sjögren’s syndrome: A review of the literature and conclusions. Int. J. Ophthalmol. 2022, 15, 1864–1872. [Google Scholar] [CrossRef]
  33. Jones, L.; Downie, L.E.; Korb, D.; Benitez-Del-Castillo, J.M.; Dana, R.; Deng, S.X.; Dong, P.N.; Geerling, G.; Hida, R.Y.; Liu, Y.; et al. TFOS DEWS II Management and Therapy Report. Ocul. Surf. 2017, 15, 575–628. [Google Scholar] [CrossRef]
  34. Management and Therapy of Dry Eye Disease: Report of the Management and Therapy Subcommittee of the International Dry Eye WorkShop (2007). Ocul. Surf. 2007, 5, 163–178. [CrossRef] [PubMed]
  35. Thacker, M.; Singh, V.; Basu, S.; Singh, S. Biomaterials for dry eye disease treatment: Current overview and future perspectives. Exp. Eye Res. 2023, 226, 109339. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, Q.; Cheng, W.; Liu, C.; Jin, X.; Ming, S.; Zhao, D.; Feng, X. Evaluation of effects of 3% diquafosol ophthalmic solution on preocular tear film stability after trabeculectomy. Int. Ophthalmol. 2023, 43, 1903–1910. [Google Scholar] [CrossRef] [PubMed]
  37. Xu, X.; Awwad, S.; Diaz-Gomez, L.; Alvarez-Lorenzo, C.; Brocchini, S.; Gaisford, S.; Goyanes, A.; Basit, A.W. 3D Printed Punctal Plugs for Controlled Ocular Drug Delivery. Pharmaceutics 2021, 13, 1421. [Google Scholar] [CrossRef] [PubMed]
  38. De Paiva, C.S.; Corrales, R.M.; Villarreal, A.L.; Farley, W.J.; Li, D.-Q.; Stern, M.E.; Pflugfelder, S.C. Corticosteroid and doxycycline suppress MMP-9 and inflammatory cytokine expression, MAPK activation in the corneal epithelium in experimental dry eye. Exp. Eye Res. 2006, 83, 526–535. [Google Scholar] [CrossRef]
  39. De Paiva, C.S.; Corrales, R.M.; Villarreal, A.L.; Farley, W.; Li, D.-Q.; Stern, M.E.; Pflugfelder, S.C. Apical Corneal Barrier Disruption in Experimental Murine Dry Eye Is Abrogated by Methylprednisolone and Doxycycline. Investig. Ophthalmol. Vis. Sci. 2006, 47, 2847–2856. [Google Scholar] [CrossRef]
  40. Prinz, J.; Maffulli, N.; Fuest, M.; Walter, P.; Bell, A.; Migliorini, F. Efficacy of Topical Administration of Corticosteroids for the Management of Dry Eye Disease: Systematic Review and Meta-Analysis. Life 2022, 12, 1932. [Google Scholar] [CrossRef]
  41. Lekhanont, K.; Leyngold, I.M.; Suwan-Apichon, O.; Rangsin, R.; Chuck, R.S. Comparison of Topical Dry Eye Medications for the Treatment of Keratoconjunctivitis Sicca in a Botulinum Toxin B-Induced Mouse Model. J. Cornea Extern. Dis. 2007, 26, 84–89. [Google Scholar] [CrossRef]
  42. Avunduk, A.M.; Avunduk, M.C.; Varnell, E.D.; Kaufman, H.E. The comparison of efficacies of topical corticosteroids and nonsteroidal anti-inflammatory drops on dry eye patients: A clinical and immunocytochemical study. Am. J. Ophthalmol. 2003, 136, 593–602. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, H.K.; Ryu, I.H.; Seo, K.Y.; Hong, S.; Kim, H.C.; Kim, E.K. Topical 0.1% Prednisolone Lowers Nerve Growth Factor Expression in Keratoconjunctivitis Sicca Patients. Ophthalmology 2006, 113, 198–205. [Google Scholar] [CrossRef] [PubMed]
  44. Schechter, B.A. Ketorolac During the Induction Phase of Cyclosporin-A Therapy. J. Ocul. Pharmacol. Ther. 2006, 22, 150–154. [Google Scholar] [CrossRef] [PubMed]
  45. Bai, R.; Liu, L.-p.; Chen, Z.; Ma, Q. Cyclosporine (0.05%) Combined with Diclofenac Sodium Eye Drops for the Treatment of Dry Eye Disease. J. Ophthalmol. 2022, 2022, 2334077. [Google Scholar] [CrossRef]
  46. Kunert, K.S.; Tisdale, A.S.; Stern, M.E.; Smith, J.A.; Gipson, I.K. Analysis of Topical Cyclosporine Treatment of Patients with Dry Eye Syndrome: Effect on Conjunctival Lymphocytes. Arch. Ophthalmol. 2000, 118, 1489–1496. [Google Scholar] [CrossRef]
  47. Turner, K.; Pflugfelder, S.C.; Ji, Z.; Feuer, W.J.; Stern, M.; Reis, B.L. Interleukin-6 Levels in the Conjunctival Epithelium of Patients with Dry Eye Disease Treated with Cyclosporine Ophthalmic Emulsion. J. Cornea Extern. Dis. 2000, 19, 492–496. [Google Scholar] [CrossRef]
  48. Gao, J.; Sana, R.; Calder, V.; Calonge, M.; Lee, W.; Wheeler, L.A.; Stern, M.E. Mitochondrial Permeability Transition Pore in Inflammatory Apoptosis of Human Conjunctival Epithelial Cells and T Cells: Effect of Cyclosporin A. Investig. Ophthalmol. Vis. Sci. 2013, 54, 4717–4733. [Google Scholar] [CrossRef]
  49. Modi, D.; Nirmal, J.; Warsi, M.H.; Bhatia, M.; Hasan, N.; Kesharwani, P.; Jain, G.K. Formulation and development of tacrolimus-gellan gum nanoformulation for treatment of dry eye disease. Colloids Surf. B Biointerfaces 2022, 211, 112255. [Google Scholar] [CrossRef]
  50. Pearce, E.I.; Dorman, M.; Wilkinson, B.C.; Oliver, K.M. Effect of Blink Frequency on Tear Turnover Rate. Investig. Ophthalmol. Vis. Sci. 2011, 52, 3726. [Google Scholar]
  51. Saarinen-Savolainen, P.; Järvinen, T.; Araki-Sasaki, K.; Watanabe, H.; Urtti, A. Evaluation of Cytotoxicity of Various Ophthalmic Drugs, Eye Drop Excipients and Cyclodextrins in an Immortalized Human Corneal Epithelial Cell Line. Pharm. Res. 1998, 15, 1275–1280. [Google Scholar] [CrossRef]
  52. Kojima, T.; Matsumoto, Y.; Ibrahim, O.M.A.; Wakamatsu, T.H.; Dogru, M.; Tsubota, K. Evaluation of a Thermosensitive Atelocollagen Punctal Plug Treatment for Dry Eye Disease. Am. J. Ophthalmol. 2014, 157, 311–317.e311. [Google Scholar] [CrossRef] [PubMed]
  53. Jin, H.D.; Surbeck, J.W.; Marsh, H.R.; Ding, K.; Kingsley, R.M.; Riaz, K.M.; Park, S.S.K.; Shah, V.A. The effect of punctal plugs in reducing ocular surface irritation after povidone-iodine preparation of intravitreal injection—A randomized trial. Eye 2022, 36, 568–574. [Google Scholar] [CrossRef] [PubMed]
  54. Dai, M.; Xu, K.; Xiao, D.; Zheng, Y.; Zheng, Q.; Shen, J.; Qian, Y.; Chen, W. In Situ Forming Hydrogel as a Tracer and Degradable Lacrimal Plug for Dry Eye Treatment. Adv. Healthc. Mater. 2022, 11, 2200678. [Google Scholar] [CrossRef] [PubMed]
  55. Mithani, S.D.; Bakatselou, V.; TenHoor, C.N.; Dressman, J.B. Estimation of the Increase in Solubility of Drugs as a Function of Bile Salt Concentration. Pharm. Res. 1996, 13, 163–167. [Google Scholar] [CrossRef] [PubMed]
  56. El Tayar, N.; Mark, A.E.; Vallat, P.; Brunne, R.M.; Testa, B.; van Gunsteren, W.F. Solvent-dependent conformation and hydrogen-bonding capacity of cyclosporin A: Evidence from partition coefficients and molecular dynamics simulations. J. Med. Chem. 1993, 36, 3757–3764. [Google Scholar] [CrossRef]
  57. Mandal, A.; Gote, V.; Pal, D.; Ogundele, A.; Mitra, A.K. Ocular Pharmacokinetics of a Topical Ophthalmic Nanomicellar Solution of Cyclosporine (Cequa®) for Dry Eye Disease. Pharm. Res. 2019, 36, 36. [Google Scholar] [CrossRef]
  58. Perry, H.D.; Solomon, R.; Donnenfeld, E.D.; Perry, A.R.; Wittpenn, J.R.; Greenman, H.E.; Savage, H.E. Evaluation of Topical Cyclosporine for the Treatment of Dry Eye Disease. Arch. Ophthalmol. 2008, 126, 1046–1050. [Google Scholar] [CrossRef] [Green Version]
  59. Leonardi, A.; Labetoulle, M.; Ismail, D.; Garrigue, J.S.; Rancho, L.; Brignole-Baudouin, F.; Amrane, M.; Baudouin, C. The Effect of Ikervis® (1 mg/mL Ciclosporin cationic emulsion) on severe keratitis in patients with dry eye disease participating in a phase III study. Acta Pharmacol. 2015, 93. [Google Scholar] [CrossRef]
  60. Safwat, M.A.; Mansour, H.F.; Hussein, A.K.; Abdelwahab, S.; Soliman, G.M. Polymeric micelles for the ocular delivery of triamcinolone acetonide: Preparation and in vivo evaluation in a rabbit ocular inflammatory model. Drug Deliv. 2020, 27, 1115–1124. [Google Scholar] [CrossRef]
  61. Shoman, N.A.; Gebreel, R.M.; El-Nabarawi, M.A.; Attia, A. Optimization of hyaluronan-enriched cubosomes for bromfenac delivery enhancing corneal permeation: Characterization, ex vivo, and in vivo evaluation. Drug Deliv. 2023, 30, 2162162. [Google Scholar] [CrossRef]
  62. Weinreb, R.N.; Aung, T.; Medeiros, F.A. The Pathophysiology and Treatment of Glaucoma: A Review. JAMA 2014, 311, 1901–1911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Razeghinejad, R.; Lin, M.M.; Lee, D.; Katz, L.J.; Myers, J.S. Pathophysiology and management of glaucoma and ocular hypertension related to trauma. Surv. Ophthalmol. 2020, 65, 530–547. [Google Scholar] [CrossRef]
  64. Khatib, T.Z.; Martin, K.R. Protecting retinal ganglion cells. Eye 2017, 31, 218–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Jindal, V. Glaucoma: An Extension of Various Chronic Neurodegenerative Disorders. Mol. Neurobiol. 2013, 48, 186–189. [Google Scholar] [CrossRef] [PubMed]
  66. Tham, Y.-C.; Li, X.; Wong, T.Y.; Quigley, H.A.; Aung, T.; Cheng, C.-Y. Global Prevalence of Glaucoma and Projections of Glaucoma Burden through 2040: A Systematic Review and Meta-Analysis. Ophthalmology 2014, 121, 2081–2090. [Google Scholar] [CrossRef]
  67. Quigley, H.A.; Broman, A.T. The number of people with glaucoma worldwide in 2010 and 2020. Br. J. Ophthalmol. 2006, 90, 262–267. [Google Scholar] [CrossRef] [Green Version]
  68. O’Callaghan, J.; Cassidy, P.S.; Humphries, P. Open-angle glaucoma: Therapeutically targeting the extracellular matrix of the conventional outflow pathway. Expert Opin. Ther. Targets 2017, 21, 1037–1050. [Google Scholar] [CrossRef] [Green Version]
  69. Braunger, B.M.; Fuchshofer, R.; Tamm, E.R. The aqueous humor outflow pathways in glaucoma: A unifying concept of disease mechanisms and causative treatment. Eur. J. Pharm. Biopharm. 2015, 95, 173–181. [Google Scholar] [CrossRef]
  70. Pattabiraman, P.P.; Toris, C.B. The exit strategy: Pharmacological modulation of extracellular matrix production and deposition for better aqueous humor drainage. Eur. J. Pharmacol. 2016, 787, 32–42. [Google Scholar] [CrossRef]
  71. Janagam, D.R.; Wu, L.; Lowe, T.L. Nanoparticles for drug delivery to the anterior segment of the eye. Adv. Drug Deliv. Rev. 2017, 122, 31–64. [Google Scholar] [CrossRef]
  72. Szigiato, A.-A.; Podbielski, D.W.; Ahmed, I.I.K. Sustained drug delivery for the management of glaucoma. Expert Rev. Ophthalmol. 2017, 12, 173–186. [Google Scholar] [CrossRef]
  73. Pita-Thomas, D.W.; Goldberg, J.L. Nanotechnology and glaucoma: Little particles for a big disease. Curr. Opin. Ophthalmol. 2013, 24, 130–135. [Google Scholar] [CrossRef]
  74. Alviset, G.; Corvis, Y.; Hammad, K.; Lemut, J.; Maury, M.; Mignet, N.; Boudy, V. New Preservative-Free Formulation for the Enhanced Ocular Bioavailability of Prostaglandin Analogues in Glaucoma. Pharmaceutics 2022, 14, 453. [Google Scholar] [CrossRef] [PubMed]
  75. Ismail, A.; Nasr, M.; Sammour, O. Nanoemulsion as a feasible and biocompatible carrier for ocular delivery of travoprost: Improved pharmacokinetic/pharmacodynamic properties. Int. J. Pharm. 2020, 583, 119402. [Google Scholar] [CrossRef] [PubMed]
  76. Karataş, A.; Sonakin, Ö.; KiliÇarslan, M.; Baykara, T. Poly (ε-caprolactone) microparticles containing Levobunolol HCl prepared by a multiple emulsion (W/O/W) solvent evaporation technique: Effects of some formulation parameters on microparticle characteristics. J. Microencapsul. 2009, 26, 63–74. [Google Scholar] [CrossRef]
  77. Abdel Azim, E.A.; Elkheshen, S.A.; Hathout, R.M.; Fouly, M.A.; El Hoffy, N.M. Augmented in vitro and in vivo Profiles of Brimonidine Tartrate Using Gelatinized-Core Liposomes. Int. J. Nanomed. 2022, 17, 2753–2776. [Google Scholar] [CrossRef] [PubMed]
  78. Kouchak, M.; Malekahmadi, M.; Bavarsad, N.; Saki Malehi, A.; Andishmand, L. Dorzolamide nanoliposome as a long action ophthalmic delivery system in open angle glaucoma and ocular hypertension patients. Drug Dev. Ind. Pharm. 2018, 44, 1239–1242. [Google Scholar] [CrossRef] [PubMed]
  79. Rosenbaum, J.T.; Asquith, M. The microbiome and HLA-B27-associated acute anterior uveitis. Nat. Rev. Rheumatol. 2018, 14, 704–713. [Google Scholar] [CrossRef]
  80. Nussenblatt, R.B. The natural history of uveitis. Int. Ophthalmol. 1990, 14, 303–308. [Google Scholar] [CrossRef]
  81. Suttorp-Schulten, M.S.; Rothova, A. The possible impact of uveitis in blindness: A literature survey. Br. J. Ophthalmol. 1996, 80, 844–848. [Google Scholar] [CrossRef] [Green Version]
  82. Mérida, S.; Palacios, E.; Navea, A.; Bosch-Morell, F. New Immunosuppressive Therapies in Uveitis Treatment. Int. J. Mol. Sci. 2015, 16, 18778–18795. [Google Scholar] [CrossRef] [Green Version]
  83. Oh, B.-L.; Lee, J.S.; Lee, E.Y.; Lee, H.Y.; Yu, H.G. Incidence and Risk Factors for Blindness in Uveitis: A Nationwide Cohort Study from 2002 to 2013. Ocul. Immunol. Inflamm. 2021, 29, 1040–1044. [Google Scholar] [CrossRef]
  84. Gritz, D.C.; Wong, I.G. Incidence and prevalence of uveitis in Northern California: The Northern California Epidemiology of Uveitis Study. Ophthalmology 2004, 111, 491–500. [Google Scholar] [CrossRef]
  85. Yamaguchi, M.; Yasueda, S.-i.; Isowaki, A.; Yamamoto, M.; Kimura, M.; Inada, K.; Ohtori, A. Formulation of an ophthalmic lipid emulsion containing an anti-inflammatory steroidal drug, difluprednate. Int. J. Pharm. 2005, 301, 121–128. [Google Scholar] [CrossRef]
  86. Suri, R.; Neupane, Y.R.; Mehra, N.; Jain, G.K.; Kohli, K. Sirolimus loaded polyol modified liposomes for the treatment of Posterior Segment Eye Diseases. Med. Hypotheses 2020, 136, 109518. [Google Scholar] [CrossRef] [PubMed]
  87. Durand, M.L. Endophthalmitis. Clin. Microbiol. Infect. 2013, 19, 227–234. [Google Scholar] [CrossRef] [Green Version]
  88. Callegan, M.C.; Engelbert, M.; Parke, D.W.; Jett, B.D.; Gilmore, M.S. Bacterial Endophthalmitis: Epidemiology, Therapeutics, and Bacterium-Host Interactions. Clin. Microbiol. Rev. 2002, 15, 111–124. [Google Scholar] [CrossRef] [Green Version]
  89. Asencio-Egea, M.A.; Huertas-Vaquero, M.; Carranza-González, R.; Cells-Sánchez, J.; González-del Valle, F.; Tenías-Burillo, J.M.; Barberá-Farré, J.R. Endogenous endophthalmitis: Case report and brief review of a serious ocular disease. Rev. Chil. Infectol. 2013, 30, 516–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Wong, J.-S.; Chan, T.-K.; Lee, H.-M.; Chee, S.-P. Endogenous bacterial endophthalmitis: An East Asian experience and a reappraisal of a severe ocular affliction. Ophthalmology 2000, 107, 1483–1491. [Google Scholar] [CrossRef]
  91. Cam, D.; Saatci, A.O.; Micili, S.C.; Ergur, B.U.; Karabag, R.Y.; Durak, I.; Berk, A.T. The Effect of Intravitreal Azithromycin on the Albino Newborn Rabbit Retina. Open Ophthalmol. J. 2016, 10, 12–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Citirik, M.; Dilsiz, N.; Batman, C.; Zilelioglu, O. Comparative toxicity of 4 commonly used intravitreal corticosteroids on rat retina. Can. J. Ophthalmol. 2009, 44, e3–e8. [Google Scholar] [CrossRef]
  93. del Amo, E.M.; Urtti, A. Current and future ophthalmic drug delivery systems: A shift to the posterior segment. Drug Discov. Today 2008, 13, 135–143. [Google Scholar] [CrossRef]
  94. D’Amico, D.J.; Caspers-Velu, L.; Libert, J.; Shanks, E.; Schrooyen, M.; Hanninen, L.A.; Kenyon, K.R. Comparative Toxicity of Intravitreal Aminoglycoside Antibiotics. Am. J. Ophthalmol. 1985, 100, 264–275. [Google Scholar] [CrossRef]
  95. Yu, S.-Y.; Damico, F.M.; Viola, F.; D’Amico, D.J.; Young, L.H. Retinal Toxicity of intravitreal triamcinolone Acetonide: A Morphological Study. RETINA 2006, 26, 531–536. [Google Scholar] [CrossRef]
  96. Ye, Z.; Yang, Y.; Ke, W.; Li, Y.; Wang, K.; Chen, M. Overview and update on cytomegalovirus-associated anterior uveitis and glaucoma. Front. Public Health 2023, 11, 1117412. [Google Scholar] [CrossRef] [PubMed]
  97. Melancia, D.; Fernandes, A.; Manita, M.; Cordeiro, I.M. Cytomegalovirus optic neuropathy in a young immunocompetent patient. J. NeuroVirology 2021, 27, 364–366. [Google Scholar] [CrossRef] [PubMed]
  98. Holland, G.N.; Vaudaux, J.D.; Shiramizu, K.M.; Yu, F.; Goldenberg, D.T.; Gupta, A.; Carlson, M.; Read, R.W.; Novack, R.D.; Kuppermann, B.D. Characteristics of Untreated AIDS-related Cytomegalovirus Retinitis. II. Findings in the Era of Highly Active Antiretroviral Therapy (1997 to 2000). Am. J. Ophthalmol. 2008, 145, 12–22.e10. [Google Scholar] [CrossRef]
  99. Jabs, D.A.; Ahuja, A.; Van Natta, M.; Lyon, A.; Srivastava, S.; Gangaputra, S. Course of Cytomegalovirus Retinitis in the Era of Highly Active Antiretroviral Therapy: Five-Year Outcomes. Ophthalmology 2010, 117, 2152–2161.e2152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Port, A.D.; Orlin, A.; Kiss, S.; Patel, S.; D’Amico, D.J.; Gupta, M.P. Cytomegalovirus Retinitis: A Review. J. Ocul. Pharmacol. Ther. 2017, 33, 224–234. [Google Scholar] [CrossRef]
  101. Gupta, M.P.; Koenig, L.R.; Doubrovina, E.; Hasan, A.; Dahi, P.B.; O’Reilly, R.J.; Koehne, G.; Orlin, A.; Chan, R.V.P.; D’Amico, D.J.; et al. Ocular Outcomes after Treatment of Cytomegalovirus Retinitis Using Adoptive Immunotherapy with Cytomegalovirus-Specific Cytotoxic T Lymphocytes. Ophthalmol. Retin. 2021, 5, 838–849. [Google Scholar] [CrossRef]
  102. Coleman, H.R.; Chan, C.-C.; Ferris, F.L.; Chew, E.Y. Age-related macular degeneration. Lancet 2008, 372, 1835–1845. [Google Scholar] [CrossRef]
  103. Lim, L.S.; Mitchell, P.; Seddon, J.M.; Holz, F.G.; Wong, T.Y. Age-related macular degeneration. Lancet 2012, 379, 1728–1738. [Google Scholar] [CrossRef]
  104. de Jong, S.; Tang, J.; Clark, S.J. Age-related macular degeneration: A disease of extracellular complement amplification. Immunol. Rev. 2023, 313, 279–297. [Google Scholar] [CrossRef]
  105. Friedman, D.S.; O’Colmain, B.J.; Munoz, B.; Tomany, S.C.; McCarty, C.; De Jong, P.T.; Nemesure, B.; Mitchell, P.; Kempen, J. Prevalence of Age-Related Macular Degeneration in the United States. Arch. Ophthalmol. 2004, 122, 564–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. 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: A Systematic Review of the Literature and Meta-analysis. Ophthalmology 2008, 115, 116–126.e111. [Google Scholar] [CrossRef]
  107. Guyer, D.R.; Fine, S.L.; Maguire, M.G.; Hawkins, B.S.; Owens, S.L.; Murphy, R.P. Subfoveal Choroidal Neovascular Membranes in Age-Related Macular Degeneration: Visual Prognosis in Eyes with Relatively Good Initial Visual Acuity. Arch. Ophthalmol. 1986, 104, 702–705. [Google Scholar] [CrossRef]
  108. Hageman, G.S.; Luthert, P.J.; Victor Chong, N.H.; Johnson, L.V.; Anderson, D.H.; Mullins, R.F. An Integrated Hypothesis That Considers Drusen as Biomarkers of Immune-Mediated Processes at the RPE-Bruch’s Membrane Interface in Aging and Age-Related Macular Degeneration. Prog. Retin. Eye Res. 2001, 20, 705–732. [Google Scholar] [CrossRef] [PubMed]
  109. Ambati, J.; Atkinson, J.P.; Gelfand, B.D. Immunology of age-related macular degeneration. Nat. Rev. Immunol. 2013, 13, 438–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Klein, M.L.; Ferris, F.L.; Armstrong, J.; Hwang, T.S.; Chew, E.Y.; Bressler, S.B.; Chandra, S.R. Retinal Precursors and the Development of Geographic Atrophy in Age-Related Macular Degeneration. Ophthalmology 2008, 115, 1026–1031. [Google Scholar] [CrossRef]
  111. Ambati, J.; Fowler, B.J. Mechanisms of Age-Related Macular Degeneration. Neuron 2012, 75, 26–39. [Google Scholar] [CrossRef] [Green Version]
  112. Jensen, E.G.; Jakobsen, T.S.; Thiel, S.; Askou, A.L.; Corydon, T.J. Associations between the Complement System and Choroidal Neovascularization in Wet Age-Related Macular Degeneration. Int. J. Mol. Sci. 2020, 21, 9752. [Google Scholar] [CrossRef]
  113. Eandi, C.M.; Alovisi, C.; De Sanctis, U.; Grignolo, F.M. Treatment for neovascular age related macular degeneration: The state of the art. Eur. J. Pharmacol. 2016, 787, 78–83. [Google Scholar] [CrossRef]
  114. 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]
  115. Muñoz-Ramón, P.V.; Hernández Martínez, P.; Muñoz-Negrete, F.J. New therapeutic targets in the treatment of age-related macular degeneration. Arch. Soc. Española Oftalmol. (Engl. Ed.) 2020, 95, 75–83. [Google Scholar] [CrossRef]
  116. Yafai, Y.; Yang, X.M.; Niemeyer, M.; Nishiwaki, A.; Lange, J.; Wiedemann, P.; King, A.G.; Yasukawa, T.; Eichler, W. Anti-angiogenic effects of the receptor tyrosine kinase inhibitor, pazopanib, on choroidal neovascularization in rats. Eur. J. Pharmacol. 2011, 666, 12–18. [Google Scholar] [CrossRef]
  117. Age-Related Eye Disease Study Research Group. A Randomized, Placebo-Controlled, Clinical Trial of High-Dose Supplementation with Vitamins C and E, Beta Carotene, and Zinc for Age-Related Macular Degeneration and Vision Loss: AREDS Report No. 8. Arch. Ophthalmol. 2001, 119, 1417–1436. [Google Scholar] [CrossRef] [Green Version]
  118. Age-Related Eye Disease Study 2 (AREDS2) Research Group. Lutein + Zeaxanthin and Omega-3 Fatty Acids for Age-Related Macular Degeneration: The Age-Related Eye Disease Study 2 (AREDS2) Randomized Clinical Trial. JAMA 2013, 309, 2005–2015. [Google Scholar] [CrossRef] [Green Version]
  119. Ng, E.W.M.; Shima, D.T.; Calias, P.; Cunningham, E.T.; Guyer, D.R.; Adamis, A.P. Pegaptanib, a targeted anti-VEGF aptamer for ocular vascular disease. Nat. Rev. Drug Discov. 2006, 5, 123–132. [Google Scholar] [CrossRef] [PubMed]
  120. Teo, Z.L.; Tham, Y.-C.; Yu, M.; Chee, M.L.; Rim, T.H.; Cheung, N.; Bikbov, M.M.; Wang, Y.X.; Tang, Y.; Lu, Y.; et al. Global Prevalence of Diabetic Retinopathy and Projection of Burden through 2045: Systematic Review and Meta-analysis. Ophthalmology 2021, 128, 1580–1591. [Google Scholar] [CrossRef] [PubMed]
  121. Ogurtsova, K.; da Rocha Fernandes, J.D.; Huang, Y.; Linnenkamp, U.; Guariguata, L.; Cho, N.H.; Cavan, D.; Shaw, J.E.; Makaroff, L.E. IDF Diabetes Atlas: Global estimates for the prevalence of diabetes for 2015 and 2040. Diabetes Res. Clin. Pract. 2017, 128, 40–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Yau, J.W.Y.; Rogers, S.L.; Kawasaki, R.; Lamoureux, E.L.; Kowalski, J.W.; Bek, T.; Chen, S.-J.; Dekker, J.M.; Fletcher, A.; Grauslund, J.; et al. Global Prevalence and Major Risk Factors of Diabetic Retinopathy. Diabetes Care 2012, 35, 556–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Andersen, N.; Hjortdal, J.; Schielke, K.C.; Bek, T.; Grauslund, J.; Laugesen, C.S.; Lund-Andersen, H.; Cerqueira, C.; Andresen, J. The Danish Registry of Diabetic Retinopathy. Clin. Epidemiol. 2016, 8, 613–619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Rodríguez, M.L.; Pérez, S.; Mena-Mollá, S.; Desco, M.C.; Ortega, Á.L. Oxidative Stress and Microvascular Alterations in Diabetic Retinopathy: Future Therapies. Oxidative Med. Cell Longev. 2019, 2019, 4940825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Romero-Aroca, P.; Baget-Bernaldiz, M.; Pareja-Rios, A.; Lopez-Galvez, M.; Navarro-Gil, R.; Verges, R. Diabetic Macular Edema Pathophysiology: Vasogenic versus Inflammatory. J. Diabetes Res. 2016, 2016, 2156273. [Google Scholar] [CrossRef] [Green Version]
  126. Tauber, J. Efficacy, tolerability and comfort of a 0.3% hypromellose gel ophthalmic lubricant in the treatment of patients with moderate to severe dry eye syndrome. Curr. Med. Res. Opin. 2007, 23, 2629–2636. [Google Scholar] [CrossRef]
  127. Rajendraprasad, R.M.; Kwatra, G.; Batra, N. Carboxymethyl Cellulose versus Hydroxypropyl Methylcellulose Tear Substitutes for Dry Eye Due to Computer Vision Syndrome: Comparison of Efficacy and Safety. Int. J. Appl. Basic Med. Res. 2021, 11, 4–8. [Google Scholar] [CrossRef]
  128. Hynnekleiv, L.; Magno, M.; Vernhardsdottir, R.R.; Moschowits, E.; Tønseth, K.A.; Dartt, D.A.; Vehof, J.; Utheim, T.P. Hyaluronic acid in the treatment of dry eye disease. Acta Ophthalmol. 2022, 100, 844–860. [Google Scholar] [CrossRef]
  129. Sun, X.; Liu, L.; Liu, C. Topical diquafosol versus hyaluronic acid for the treatment of dry eye disease: A meta-analysis of randomized controlled trials. Graefe’s Arch. Clin. Exp. Ophthalmol. 2023, 1–13. [Google Scholar] [CrossRef]
  130. Chen, M.; Yung Choi, S. Preliminary Outcomes of Temporary Collagen Punctal Plugs for Patients with Dry Eye and Glaucoma. Med. Hypothesis Discov. Innov. Ophthalmol. J. 2020, 9, 56–60. [Google Scholar]
  131. Nagai, N.; Ishii, M.; Seiriki, R.; Ogata, F.; Otake, H.; Nakazawa, Y.; Okamoto, N.; Kanai, K.; Kawasaki, N. Novel Sustained-Release Drug Delivery System for Dry Eye Therapy by Rebamipide Nanoparticles. Pharmaceutics 2020, 12, 155. [Google Scholar] [CrossRef] [Green Version]
  132. Qiao, H.; Xu, Z.; Sun, M.; Fu, S.; Zhao, F.; Wang, D.; He, Z.; Zhai, Y.; Sun, J. Rebamipide liposome as an effective ocular delivery system for the management of dry eye disease. J. Drug Deliv. Sci. Technol. 2022, 75, 103654. [Google Scholar] [CrossRef]
  133. Li, Q.; Wu, X.; Xin, M. Strengthened rebamipide ocular nanoformulation to effectively treat corneal alkali burns in mice through the HMGB1 signaling pathway. Exp. Eye Res. 2021, 213, 108824. [Google Scholar] [CrossRef]
  134. Terreni, E.; Zucchetti, E.; Tampucci, S.; Burgalassi, S.; Monti, D.; Chetoni, P. Combination of Nanomicellar Technology and In Situ Gelling Polymer as Ocular Drug Delivery System (ODDS) for Cyclosporine-A. Pharmaceutics 2021, 13, 192. [Google Scholar] [CrossRef]
  135. Bang, S.P.; Yeon, C.Y.; Adhikari, N.; Neupane, S.; Kim, H.; Lee, D.C.; Son, M.J.; Lee, H.G.; Kim, J.-Y.; Jun, J.H. Cyclosporine A eyedrops with self-nanoemulsifying drug delivery systems have improved physicochemical properties and efficacy against dry eye disease in a murine dry eye model. PLoS ONE 2019, 14, e0224805. [Google Scholar] [CrossRef] [Green Version]
  136. Wong, K.-Y.; Liu, Y.; Zhou, L.; Wong, M.-S.; Liu, J. Mucin-targeting-aptamer functionalized liposomes for delivery of cyclosporin A for dry eye diseases. J. Mater. Chem. B 2023, 11. [Google Scholar] [CrossRef]
  137. Chhowala, I.; Patel, A.; Patel, R.; Bhavsar, V.; Dharamsi, A. Optimisation of PCL-HA laden biodegradable nanoparticles containing cyclosporine-A for the treatment of dry eye syndrome: In vitro-in vivo evaluation. Int. J. Nanoparticles 2021, 13, 106–120. [Google Scholar] [CrossRef]
  138. Lallemand, F.; Daull, P.; Benita, S.; Buggage, R.; Garrigue, J.-S.J.J.O.D.D. Successfully improving ocular drug delivery using the cationic nanoemulsion, novasorb. J. Drug Deliv. 2012, 2012, 604204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Gökçe, E.H.; Sandri, G.; Eğrilmez, S.; Bonferoni, M.C.; Güneri, T.; Caramella, C. Cyclosporine A-Loaded Solid Lipid Nanoparticles: Ocular Tolerance and In Vivo Drug Release in Rabbit Eyes. Curr. Eye Res. 2009, 34, 996–1003. [Google Scholar] [CrossRef] [PubMed]
  140. Huang, H.Y.; Wang, M.C.; Chen, Z.Y.; Chiu, W.Y.; Chen, K.H.; Lin, I.C.; Yang, W.V.; Wu, C.C.; Tseng, C.L. Gelatin-epigallocatechin gallate nanoparticles with hyaluronic acid decoration as eye drops can treat rabbit dry-eye syndrome effectively via inflammatory relief. Int. J. Nanomed. 2018, 13, 7251–7273. [Google Scholar] [CrossRef] [Green Version]
  141. Luo, L.-J.; Lai, J.-Y. Epigallocatechin Gallate-Loaded Gelatin-g-Poly(N-Isopropylacrylamide) as a New Ophthalmic Pharmaceutical Formulation for Topical Use in the Treatment of Dry Eye Syndrome. Sci. Rep. 2017, 7, 9380. [Google Scholar] [CrossRef] [Green Version]
  142. López-Machado, A.; Díaz, N.; Cano, A.; Espina, M.; Badía, J.; Baldomà, L.; Calpena, A.C.; Biancardi, M.; Souto, E.B.; García, M.L.; et al. Development of topical eye-drops of lactoferrin-loaded biodegradable nanoparticles for the treatment of anterior segment inflammatory processes. Int. J. Pharm. 2021, 609, 121188. [Google Scholar] [CrossRef] [PubMed]
  143. Varela-Fernández, R.; García-Otero, X.; Díaz-Tomé, V.; Regueiro, U.; López-López, M.; González-Barcia, M.; Isabel Lema, M.; Otero-Espinar, F.J. Mucoadhesive PLGA Nanospheres and Nanocapsules for Lactoferrin Controlled Ocular Delivery. Pharmaceutics 2022, 14, 799. [Google Scholar] [CrossRef] [PubMed]
  144. López-Machado, A.; Díaz-Garrido, N.; Cano, A.; Espina, M.; Badia, J.; Baldomà, L.; Calpena, A.C.; Souto, E.B.; García, M.L.; Sánchez-López, E. Development of Lactoferrin-Loaded Liposomes for the Management of Dry Eye Disease and Ocular Inflammation. Pharmaceutics 2021, 13, 1698. [Google Scholar] [CrossRef] [PubMed]
  145. Varela-Fernández, R.; García-Otero, X.; Díaz-Tomé, V.; Regueiro, U.; López-López, M.; González-Barcia, M.; Isabel Lema, M.; Javier Otero-Espinar, F. Lactoferrin-loaded nanostructured lipid carriers (NLCs) as a new formulation for optimized ocular drug delivery. Eur. J. Pharm. Biopharm. 2022, 172, 144–156. [Google Scholar] [CrossRef]
  146. He, W.; Guo, X.; Feng, M.; Mao, N. In vitro and in vivo studies on ocular vitamin A palmitate cationic liposomal in situ gels. Int. J. Pharm. 2013, 458, 305–314. [Google Scholar] [CrossRef] [PubMed]
  147. Castro, B.F.M.; Fulgêncio, G.d.O.; Domingos, L.C.; Cotta, O.A.L.; Silva-Cunha, A.; Fialho, S.L. Positively charged polymeric nanoparticles improve ocular penetration of tacrolimus after topical administration. J. Drug Deliv. Sci. Technol. 2020, 60, 101912. [Google Scholar] [CrossRef]
  148. Mohammad; Garg, V.; Nirmal, J.; Warsi, M.H.; Pandita, D.; Kesharwani, P.; Jain, G.K. Topical Tacrolimus Progylcosomes Nano-Vesicles As a Potential Therapy for Experimental Dry Eye Syndrome. J. Pharm. Sci. 2022, 111, 479–484. [Google Scholar] [CrossRef]
  149. Zhang, C.; Zheng, Y.; Li, M.; Zhang, Z.; Chang, L.; Ai, M.; Wang, J.; Zhao, S.; Li, C.; Zhou, Z. Carboxymethyl Cellulose-Coated Tacrolimus Nonspherical Microcrystals for Improved Therapeutic Efficacy of Dry Eye. Macromol. Biosci. 2020, 20, 2000079. [Google Scholar] [CrossRef]
  150. Chen, X.; Wu, J.; Lin, X.; Wu, X.; Yu, X.; Wang, B.; Xu, W. Tacrolimus Loaded Cationic Liposomes for Dry Eye Treatment. Front. Pharmacol. 2022, 13, 838168. [Google Scholar] [CrossRef]
  151. Liu, D.; Wu, Q.; Chen, W.; Lin, H.; Zhu, Y.; Liu, Y.; Liang, H.; Zhu, F. A novel FK506 loaded nanomicelles consisting of amino-terminated poly(ethylene glycol)-block-poly(D,L)-lactic acid and hydroxypropyl methylcellulose for ocular drug delivery. Int. J. Pharm. 2019, 562, 1–10. [Google Scholar] [CrossRef]
  152. Rebibo, L.; Tam, C.; Sun, Y.; Shoshani, E.; Badihi, A.; Nassar, T.; Benita, S. Topical tacrolimus nanocapsules eye drops for therapeutic effect enhancement in both anterior and posterior ocular inflammation models. J. Control. Release 2021, 333, 283–297. [Google Scholar] [CrossRef]
  153. Soiberman, U.; Kambhampati, S.P.; Wu, T.; Mishra, M.K.; Oh, Y.; Sharma, R.; Wang, J.; Al Towerki, A.E.; Yiu, S.; Stark, W.J.; et al. Subconjunctival injectable dendrimer-dexamethasone gel for the treatment of corneal inflammation. Biomaterials 2017, 125, 38–53. [Google Scholar] [CrossRef] [Green Version]
  154. Bian, F.; Shin, C.S.; Wang, C.; Pflugfelder, S.C.; Acharya, G.; De Paiva, C.S. Dexamethasone Drug Eluting Nanowafers Control Inflammation in Alkali-Burned Corneas Associated with Dry Eye. Investig. Ophthalmol. Vis. Sci. 2016, 57, 3222–3230. [Google Scholar] [CrossRef] [Green Version]
  155. Tan, G.; Yu, S.; Li, J.; Pan, W. Development and characterization of nanostructured lipid carriers based chitosan thermosensitive hydrogel for delivery of dexamethasone. Int. J. Biol. Macromol. 2017, 103, 941–947. [Google Scholar] [CrossRef]
  156. Taheri, S.L.; Rezazadeh, M.; Hassanzadeh, F.; Akbari, V.; Dehghani, A.; Talebi, A.; Mostafavi, S.A. Preparation, physicochemical, and retinal anti-angiogenic evaluation of poloxamer hydrogel containing dexamethasone/avastin-loaded chitosan-N-acetyl-L-cysteine nanoparticles. Int. J. Biol. Macromol. 2022, 220, 1605–1618. [Google Scholar] [CrossRef]
  157. Alami-Milani, M.; Zakeri-Milani, P.; Valizadeh, H.; Sattari, S.; Salatin, S.; Jelvehgari, M. Evaluation of anti-inflammatory impact of dexamethasone-loaded PCL-PEG-PCL micelles on endotoxin-induced uveitis in rabbits. Pharm. Dev. Technol. 2019, 24, 680–688. [Google Scholar] [CrossRef]
  158. Kassem, M.A.; Abdel Rahman, A.A.; Ghorab, M.M.; Ahmed, M.B.; Khalil, R.M. Nanosuspension as an ophthalmic delivery system for certain glucocorticoid drugs. Int. J. Pharm. 2007, 340, 126–133. [Google Scholar] [CrossRef] [PubMed]
  159. Fialho, S.L.; Da Silva-Cunha, A. New vehicle based on a microemulsion for topical ocular administration of dexamethasone. Clin. Exp. Ophthalmol. 2004, 32, 626–632. [Google Scholar] [CrossRef] [PubMed]
  160. Swaminathan, S.; Vavia, P.R.; Trotta, F.; Cavalli, R. Nanosponges Encapsulating Dexamethasone for Ocular Delivery: Formulation Design, Physicochemical Characterization, Safety and Corneal Permeability Assessment. J. Biomed. Nanotechnol. 2013, 9, 998–1007. [Google Scholar] [CrossRef] [PubMed]
  161. Wang, T.-Z.; Guan, B.; Liu, X.-X.; Ke, L.-N.; Wang, J.-J.; Nan, K.-H. A topical fluorometholone nanoformulation fabricated under aqueous condition for the treatment of dry eye. Colloids Surf. B Biointerfaces 2022, 212, 112351. [Google Scholar] [CrossRef] [PubMed]
  162. Sabzevari, A.; Adibkia, K.; Hashemi, H.; De Geest, B.G.; Mohsenzadeh, N.; Atyabi, F.; Ghahremani, M.H.; Khoshayand, M.-R.; Dinarvand, R. Improved Anti-Inflammatory Effects in Rabbit Eye Model Using Biodegradable Poly Beta-Amino Ester Nanoparticles of Triamcinolone Acetonide. Investig. Ophthalmol. Vis. Sci. 2013, 54, 5520–5526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Zimmer, A.K.; Maincent, P.; Thouvenot, P.; Kreuter, J. Hydrocortisone delivery to healthy and inflamed eyes using a micellar polysorbate 80 solution or albumin nanoparticles. Int. J. Pharm. 1994, 110, 211–222. [Google Scholar] [CrossRef]
  164. Ibrahim, H.K.; El-Leithy, I.S.; Makky, A.A. Mucoadhesive Nanoparticles as Carrier Systems for Prolonged Ocular Delivery of Gatifloxacin/Prednisolone Bitherapy. Mol. Pharm. 2010, 7, 576–585. [Google Scholar] [CrossRef] [PubMed]
  165. Katzer, T.; Chaves, P.; Bernardi, A.; Pohlmann, A.; Guterres, S.S.; Ruver Beck, R.C. Prednisolone-loaded nanocapsules as ocular drug delivery system: Development, in vitro drug release and eye toxicity. J. Microencapsul. 2014, 31, 519–528. [Google Scholar] [CrossRef] [PubMed]
  166. Schopf, L.; Enlow, E.; Popov, A.; Bourassa, J.; Chen, H. Ocular Pharmacokinetics of a Novel Loteprednol Etabonate 0.4% Ophthalmic Formulation. Ophthalmol. Ther. 2014, 3, 63–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Asasutjarit, R.; Theerachayanan, T.; Kewsuwan, P.; Veeranodha, S.; Fuongfuchat, A.; Ritthidej, G.C. Development and Evaluation of Diclofenac Sodium Loaded-N-Trimethyl Chitosan Nanoparticles for Ophthalmic Use. AAPS PharmSciTech 2015, 16, 1013–1024. [Google Scholar] [CrossRef] [Green Version]
  168. Agnihotri, S.M.; Vavia, P.R. Diclofenac-loaded biopolymeric nanosuspensions for ophthalmic application. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 90–95. [Google Scholar] [CrossRef]
  169. Abrego, G.; Alvarado, H.L.; Egea, M.A.; Gonzalez-Mira, E.; Calpena, A.C.; Garcia, M.L. Design of Nanosuspensions and Freeze-Dried PLGA Nanoparticles as a Novel Approach for Ophthalmic Delivery of Pranoprofen. J. Pharm. Sci. 2014, 103, 3153–3164. [Google Scholar] [CrossRef]
  170. Luo, Y.; Yang, L.; Feng, P.; Qiu, H.; Wu, X.; Lu, S.; Zhou, M.; Xu, L.; Zhu, Y. Pranoprofen Nanoparticles with Poly(L-Lactide)-b-Poly(Ethylene Glycol)-b-Poly(L-Lactide) as the Matrix Toward Improving Ocular Anti-inflammation. Front. Bioeng. Biotechnol. 2020, 8, 581621. [Google Scholar] [CrossRef]
  171. Sánchez-Santos, I.; García-Sánchez, G.A.; Gonzalez-Salinas, R.; Linares-Alba, M.A.; Rodríguez-Reyes, A.A.; García-Santisteban, R.; Tirado-González, V.; Hernández-Piñamora, E.; García-Arzate, D.; Morales-Cantón, V.; et al. Intravitreal bromfenac liposomal suspension (100 μg/0.1 mL). A safety study in rabbit eyes. Exp. Eye Res. 2020, 194, 108020. [Google Scholar] [CrossRef]
  172. Otake, H.; Goto, R.; Ogata, F.; Isaka, T.; Kawasaki, N.; Kobayakawa, S.; Matsunaga, T.; Nagai, N. Fixed-Combination Eye Drops Based on Fluorometholone Nanoparticles and Bromfenac/Levofloxacin Solution Improve Drug Corneal Penetration. Int. J. Nanomed. 2021, 16, 5343–5356. [Google Scholar] [CrossRef]
  173. Warsi, M.H. Development and optimization of vitamin E TPGS based PLGA nanoparticles for improved and safe ocular delivery of ketorolac. J. Drug Deliv. Sci. Technol. 2021, 61, 102121. [Google Scholar] [CrossRef]
  174. Tauber, J.; Karpecki, P.; Latkany, R.; Luchs, J.; Martel, J.; Sall, K.; Raychaudhuri, A.; Smith, V.; Semba, C.P. Lifitegrast Ophthalmic Solution 5.0% versus Placebo for Treatment of Dry Eye Disease: Results of the Randomized Phase III OPUS-2 Study. Ophthalmology 2015, 122, 2423–2431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Hovanesian, J.; Epitropoulos, A.; Donnenfeld, E.D.; Holladay, J.T. The Effect of Lifitegrast on Refractive Accuracy and Symptoms in Dry Eye Patients Undergoing Cataract Surgery. Clin. Ophthalmol. (Auckl. N.Z.) 2020, 14, 2709–2716. [Google Scholar] [CrossRef] [PubMed]
  176. Cheng, Y.-H.; Ko, Y.-C.; Chang, Y.-F.; Huang, S.-H.; Liu, C.J.-l. Thermosensitive chitosan-gelatin-based hydrogel containing curcumin-loaded nanoparticles and latanoprost as a dual-drug delivery system for glaucoma treatment. Exp. Eye Res. 2019, 179, 179–187. [Google Scholar] [CrossRef]
  177. Dang, H.; Dong, C.; Zhang, L. Sustained latanoprost release from PEGylated solid lipid nanoparticle-laden soft contact lens to treat glaucoma. Pharm. Dev. Technol. 2022, 27, 127–133. [Google Scholar] [CrossRef]
  178. Xu, J.; Ge, Y.; Bu, R.; Zhang, A.; Feng, S.; Wang, J.; Gou, J.; Yin, T.; He, H.; Zhang, Y.; et al. Co-delivery of latanoprost and timolol from micelles-laden contact lenses for the treatment of glaucoma. J. Control. Release 2019, 305, 18–28. [Google Scholar] [CrossRef] [PubMed]
  179. Bessone, C.D.V.; Akhlaghi, S.P.; Tártara, L.I.; Quinteros, D.A.; Loh, W.; Allemandi, D.A. Latanoprost-loaded phytantriol cubosomes for the treatment of glaucoma. Eur. J. Pharm. Sci. 2021, 160, 105748. [Google Scholar] [CrossRef] [PubMed]
  180. Schnichels, S.; Hurst, J.; de Vries, J.W.; Ullah, S.; Gruszka, A.; Kwak, M.; Löscher, M.; Dammeier, S.; Bartz-Schmidt, K.-U.; Spitzer, M.S.; et al. Self-assembled DNA nanoparticles loaded with travoprost for glaucoma-treatment. Nanomed. Nanotechnol. Biol. Med. 2020, 29, 102260. [Google Scholar] [CrossRef]
  181. Masse, F.; Ouellette, M.; Boisselier, E. Ultrastable gold nanoparticles as a new drug vector for glaucoma therapy. Investig. Ophthalmol. Vis. Sci. 2018, 59, 3512. [Google Scholar]
  182. Shukr, M.H.; Ismail, S.; El-Hossary, G.G.; El-Shazly, A.H. Design and evaluation of mucoadhesive in situ liposomal gel for sustained ocular delivery of travoprost using two steps factorial design. J. Drug Deliv. Sci. Technol. 2021, 61, 102333. [Google Scholar] [CrossRef]
  183. Shukr, M.H.; Ismail, S.; El-Hossary, G.G.; El-Shazly, A.H. Spanlastics nanovesicular ocular insert as a novel ocular delivery of travoprost: Optimization using Box–Behnken design and in vivo evaluation. J. Liposome Res. 2022, 32, 354–364. [Google Scholar] [CrossRef]
  184. Goldstein, M.H.; Goldberg, D.; Walters, T.R.; Vantipalli, S.; Braun, E.; Metzinger, J.L. Evaluating Safety, Tolerability and Efficacy of an Intracameral Hydrogel-Based Travoprost Implant in Subjects with Glaucoma-Phase 1 Trial. Investig. Ophthalmol. Vis. Sci. 2020, 61, 4266. [Google Scholar]
  185. Wadetwar, R.N.; Agrawal, A.R.; Kanojiya, P.S. In situ gel containing Bimatoprost solid lipid nanoparticles for ocular delivery: In-vitro and ex-vivo evaluation. J. Drug Deliv. Sci. Technol. 2020, 56, 101575. [Google Scholar] [CrossRef]
  186. Li, Q.; Ma, C.; Ma, Y.; Ma, Y.; Mao, Y.; Meng, Z. Sustained bimatoprost release using gold nanoparticles laden contact lenses. J. Biomater. Sci. Polym. Ed. 2021, 32, 1618–1634. [Google Scholar] [CrossRef]
  187. Meany, E.L.; Andaya, R.; Tang, S.; Kasse, C.M.; Fuji, R.N.; Grosskopf, A.K.; d’Aquino, A.L.; Bartoe, J.T.; Ybarra, R.; Shelton, A.; et al. Injectable Polymer-Nanoparticle Hydrogel for the Sustained Intravitreal Delivery of Bimatoprost. Adv. Ther. 2023, 6, 2200207. [Google Scholar] [CrossRef]
  188. Xu, W.; Jiao, W.; Li, S.; Tao, X.; Mu, G. Bimatoprost loaded microemulsion laden contact lens to treat glaucoma. J. Drug Deliv. Sci. Technol. 2019, 54, 101330. [Google Scholar] [CrossRef]
  189. Maulvi, F.A.; Soni, P.D.; Patel, P.J.; Desai, A.R.; Desai, D.T.; Shukla, M.R.; Shah, S.A.; Shah, D.O.; Willcox, M.D.P. Controlled bimatoprost release from graphene oxide laden contact lenses: In vitro and in vivo studies. Colloids Surf. B Biointerfaces 2021, 208, 112096. [Google Scholar] [CrossRef]
  190. Seal, J.R.; Robinson, M.R.; Burke, J.; Bejanian, M.; Coote, M.; Attar, M. Intracameral Sustained-Release Bimatoprost Implant Delivers Bimatoprost to Target Tissues with Reduced Drug Exposure to Off-Target Tissues. J. Ocul. Pharmacol. Ther. 2019, 35, 50–57. [Google Scholar] [CrossRef] [Green Version]
  191. Yadav, M.; Guzman-Aranguez, A.; Perez de Lara, M.J.; Singh, M.; Singh, J.; Kaur, I.P. Bimatoprost loaded nanovesicular long-acting sub-conjunctival in-situ gelling implant: In vitro and in vivo evaluation. Mater. Sci. Eng. C 2019, 103, 109730. [Google Scholar] [CrossRef] [PubMed]
  192. Franca, J.R.; Foureaux, G.; Fuscaldi, L.L.; Ribeiro, T.G.; Rodrigues, L.B.; Bravo, R.; Castilho, R.O.; Yoshida, M.I.; Cardoso, V.N.; Fernandes, S.O.J.P.O. Bimatoprost-loaded ocular inserts as sustained release drug delivery systems for glaucoma treatment: In vitro and in vivo evaluation. PLoS ONE 2014, 9, e95461. [Google Scholar] [CrossRef] [PubMed]
  193. Nagai, N.; Yamada, S.; Kawasaki, J.; Koyanagi, E.; Saijo, S.; Kaji, H.; Nishizawa, M.; Nakazawa, T.; Abe, T. Pharmacokinetic and Safety Evaluation of a Transscleral Sustained Unoprostone Release Device in Monkey Eyes. Investig. Ophthalmol. Vis. Sci. 2018, 59, 644–652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Khallaf, A.M.; El-Moslemany, R.M.; Ahmed, M.F.; Morsi, M.H.; Khalafallah, N.M. Exploring a Novel Fasudil-Phospholipid Complex Formulated as Liposomal Thermosensitive in situ Gel for Glaucoma. Int. J. Nanomed. 2022, 17, 163–181. [Google Scholar] [CrossRef] [PubMed]
  195. Mietzner, R.; Kade, C.; Froemel, F.; Pauly, D.; Stamer, W.D.; Ohlmann, A.; Wegener, J.; Fuchshofer, R.; Breunig, M. Fasudil Loaded PLGA Microspheres as Potential Intravitreal Depot Formulation for Glaucoma Therapy. Pharmaceutics 2020, 12, 706. [Google Scholar] [CrossRef] [PubMed]
  196. Kusuhara, S.; Nakamura, M. Ripasudil Hydrochloride Hydrate in the Treatment of Glaucoma: Safety, Efficacy, and Patient Selection. Clin. Ophthalmol. 2020, 14, 1229–1236. [Google Scholar] [CrossRef]
  197. Mehran, N.A.; Sinha, S.; Razeghinejad, R. New glaucoma medications: Latanoprostene bunod, netarsudil, and fixed combination netarsudil-latanoprost. Eye 2020, 34, 72–88. [Google Scholar] [CrossRef]
  198. Jung, H.J.; Abou-Jaoude, M.; Carbia, B.E.; Plummer, C.; Chauhan, A. Glaucoma therapy by extended release of timolol from nanoparticle loaded silicone-hydrogel contact lenses. J. Control. Release 2013, 165, 82–89. [Google Scholar] [CrossRef]
  199. Huang, J.; Peng, T.; Li, Y.; Zhan, Z.; Zeng, Y.; Huang, Y.; Pan, X.; Wu, C.-Y.; Wu, C. Ocular Cubosome Drug Delivery System for Timolol Maleate: Preparation, Characterization, Cytotoxicity, Ex Vivo, and In Vivo Evaluation. AAPS PharmSciTech 2017, 18, 2919–2926. [Google Scholar] [CrossRef]
  200. Cuggino, J.C.; Tártara, L.I.; Gugliotta, L.M.; Palma, S.D.; Alvarez Igarzabal, C.I. Mucoadhesive and responsive nanogels as carriers for sustainable delivery of timolol for glaucoma therapy. Mater. Sci. Eng. C 2021, 118, 111383. [Google Scholar] [CrossRef]
  201. Hathout, R.M.; Gad, H.A.; Abdel-Hafez, S.M.; Nasser, N.; Khalil, N.; Ateyya, T.; Amr, A.; Yasser, N.; Nasr, S.; Metwally, A.A. Gelatinized core liposomes: A new Trojan horse for the development of a novel timolol maleate glaucoma medication. Int. J. Pharm. 2019, 556, 192–199. [Google Scholar] [CrossRef]
  202. Wei, N.; Dang, H.; Huang, C.; Sheng, Y. Timolol loaded microemulsion laden silicone contact lens to manage glaucoma: In vitro and in vivo studies. J. Dispers. Sci. Technol. 2021, 42, 742–750. [Google Scholar] [CrossRef]
  203. Kumar, N.; Aggarwal, R.; Chauhan, M.K. Extended levobunolol release from Eudragit nanoparticle-laden contact lenses for glaucoma therapy. Future J. Pharm. Sci. 2020, 6, 109. [Google Scholar] [CrossRef]
  204. Marchal-Heussler, L.; Sirbat, D.; Hoffman, M.; Maincent, P. Poly(ε-Caprolactone) Nanocapsules in Carteolol Ophthalmic Delivery. Pharm. Res. 1993, 10, 386–390. [Google Scholar] [CrossRef] [PubMed]
  205. Nagai, N.; Yamaoka, S.; Fukuoka, Y.; Ishii, M.; Otake, H.; Kanai, K.; Okamoto, N.; Shimomura, Y. Enhancement in Corneal Permeability of Dissolved Carteolol by Its Combination with Magnesium Hydroxide Nanoparticles. Int. J. Mol. Sci. 2018, 19, 282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Zafar, A.; Alruwaili, N.K.; Imam, S.S.; Alsaidan, O.A.; Alharbi, K.S.; Yasir, M.; Elmowafy, M.; Ansari, M.J.; Salahuddin, M.; Alshehri, S. Formulation of carteolol chitosomes for ocular delivery: Formulation optimization, ex-vivo permeation, and ocular toxicity examination. Cutan. Ocul. Toxicol. 2021, 40, 338–349. [Google Scholar] [CrossRef]
  207. Losa, C.; Marchal-Heussler, L.; Orallo, F.; Jato, J.L.V.; Alonso, M.J. Design of New Formulations for Topical Ocular Administration: Polymeric Nanocapsules Containing Metipranolol. Pharm. Res. 1993, 10, 80–87. [Google Scholar] [CrossRef]
  208. Huang, Y.; Tao, Q.; Hou, D.; Hu, S.; Tian, S.; Chen, Y.; Gui, R.; Yang, L.; Wang, Y. A novel ion-exchange carrier based upon liposome-encapsulated montmorillonite for ophthalmic delivery of betaxolol hydrochloride. Int. J. Nanomed. 2017, 12, 1731–1745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Zhao, Y.; Li, J.; Han, X.; Tao, Q.; Liu, S.; Jiang, G.; Zhu, G.; Yang, F.; Lv, Z.; Chen, Y.; et al. Dual controlled release effect of montmorillonite loaded polymer nanoparticles for ophthalmic drug delivery. Appl. Clay Sci. 2019, 180, 105167. [Google Scholar] [CrossRef]
  210. Allam, A.; Elsabahy, M.; El Badry, M.; Eleraky, N.E. Betaxolol-loaded niosomes integrated within pH-sensitive in situ forming gel for management of glaucoma. Int. J. Pharm. 2021, 598, 120380. [Google Scholar] [CrossRef]
  211. Sakr, M.G.; El-Zahaby, S.A.; Al-Mahallawi, A.M.; Ghorab, D.M. Fabrication of betaxolol hydrochloride-loaded highly permeable ocular bilosomes (HPOBs) to combat glaucoma: In vitro, ex vivo & in vivo characterizations. J. Drug Deliv. Sci. Technol. 2023, 82, 104363. [Google Scholar] [CrossRef]
  212. Sun, J.; Lei, Y.; Dai, Z.; Liu, X.; Huang, T.; Wu, J.; Xu, Z.P.; Sun, X. Sustained Release of Brimonidine from a New Composite Drug Delivery System for Treatment of Glaucoma. ACS Appl. Mater. Interfaces 2017, 9, 7990–7999. [Google Scholar] [CrossRef]
  213. Shivakumar, H.N.; Desai, B.G.; Subhash, P.G.; Ashok, P.; Hulakoti, B. Design of ocular inserts of brimonidine tartrate by response surface methodology. J. Drug Deliv. Sci. Technol. 2007, 17, 421–430. [Google Scholar] [CrossRef]
  214. Emad Eldeeb, A.; Salah, S.; Ghorab, M. Proniosomal gel-derived niosomes: An approach to sustain and improve the ocular delivery of brimonidine tartrate; formulation, in-vitro characterization, and in-vivo pharmacodynamic study. Drug Deliv. 2019, 26, 509–521. [Google Scholar] [CrossRef] [Green Version]
  215. Chiang, B.; Kim, Y.C.; Doty, A.C.; Grossniklaus, H.E.; Schwendeman, S.P.; Prausnitz, M.R. Sustained reduction of intraocular pressure by supraciliary delivery of brimonidine-loaded poly(lactic acid) microspheres for the treatment of glaucoma. J. Control. Release 2016, 228, 48–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Bigdeli, A.; Makhmalzadeh, B.S.; Feghhi, M.; SoleimaniBiatiani, E. Cationic liposomes as promising vehicles for timolol/brimonidine combination ocular delivery in glaucoma: Formulation development and in vitro/in vivo evaluation. Drug Deliv. Transl. Res. 2022, 13, 1035–1047. [Google Scholar] [CrossRef]
  217. Zhao, Y.; Huang, C.; Zhang, Z.; Hong, J.; Xu, J.; Sun, X.; Sun, J. Sustained release of brimonidine from BRI@SR@TPU implant for treatment of glaucoma. Drug Deliv. 2022, 29, 613–623. [Google Scholar] [CrossRef] [PubMed]
  218. Shahab, M.S.; Rizwanullah, M.; Alshehri, S.; Imam, S.S. Optimization to development of chitosan decorated polycaprolactone nanoparticles for improved ocular delivery of dorzolamide: In vitro, ex vivo and toxicity assessments. Int. J. Biol. Macromol. 2020, 163, 2392–2404. [Google Scholar] [CrossRef] [PubMed]
  219. Kassem, A.A.; Salama, A.; Mohsen, A.M. Formulation and optimization of cationic nanoemulsions for enhanced ocular delivery of dorzolamide hydrochloride using Box-Behnken design: In vitro and in vivo assessments. J. Drug Deliv. Sci. Technol. 2022, 68, 103047. [Google Scholar] [CrossRef]
  220. Fu, J.; Sun, F.; Liu, W.; Liu, Y.; Gedam, M.; Hu, Q.; Fridley, C.; Quigley, H.A.; Hanes, J.; Pitha, I. Subconjunctival Delivery of Dorzolamide-Loaded Poly(ether-anhydride) Microparticles Produces Sustained Lowering of Intraocular Pressure in Rabbits. Mol. Pharm. 2016, 13, 2987–2995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Fouda, N.H.; Abdelrehim, R.T.; Hegazy, D.A.; Habib, B.A. Sustained ocular delivery of Dorzolamide-HCl via proniosomal gel formulation: In-vitro characterization, statistical optimization, and in-vivo pharmacodynamic evaluation in rabbits. Drug Deliv. 2018, 25, 1340–1349. [Google Scholar] [CrossRef] [Green Version]
  222. Özdemir, S.; Çakırlı, E.; Sürücü, B.; Aygüler, C.İ.; Üner, B.; Çelebi, A.R.C. Preparation and characterization studies of dorzolamide loaded ophthalmic implants for the treatment of glaucoma. Turk. J. Pharm. Sci. 2022. [Google Scholar] [CrossRef]
  223. Franca, J.R.; Foureaux, G.; Fuscaldi, L.L.; Ribeiro, T.G.; Castilho, R.O.; Yoshida, M.I.; Cardoso, V.N.; Fernandes, S.O.A.; Cronemberger, S.; Nogueira, J.C.; et al. Chitosan/hydroxyethyl cellulose inserts for sustained-release of dorzolamide for glaucoma treatment: In vitro and in vivo evaluation. Int. J. Pharm. 2019, 570, 118662. [Google Scholar] [CrossRef] [PubMed]
  224. Wu, W.; Li, J.; Wu, L.; Wang, B.; Wang, Z.; Xu, Q.; Xin, H. Ophthalmic Delivery of Brinzolamide by Liquid Crystalline Nanoparticles: In Vitro and In Vivo Evaluation. AAPS PharmSciTech 2013, 14, 1063–1071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Tuomela, A.; Liu, P.; Puranen, J.; Rönkkö, S.; Laaksonen, T.; Kalesnykas, G.; Oksala, O.; Ilkka, J.; Laru, J.; Järvinen, K.; et al. Brinzolamide nanocrystal formulations for ophthalmic delivery: Reduction of elevated intraocular pressure in vivo. Int. J. Pharm. 2014, 467, 34–41. [Google Scholar] [CrossRef] [PubMed]
  226. Li, H.; Liu, Y.; Zhang, Y.; Fang, D.; Xu, B.; Zhang, L.; Chen, T.; Ren, K.; Nie, Y.; Yao, S.; et al. Liposomes as a Novel Ocular Delivery System for Brinzolamide: In Vitro and In Vivo Studies. AAPS PharmSciTech 2016, 17, 710–717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Dubey, V.; Mohan, P.; Dangi, J.S.; Kesavan, K. Brinzolamide loaded chitosan-pectin mucoadhesive nanocapsules for management of glaucoma: Formulation, characterization and pharmacodynamic study. Int. J. Biol. Macromol. 2020, 152, 1224–1232. [Google Scholar] [CrossRef] [PubMed]
  228. Bhalerao, H.; Koteshwara, K.B.; Chandran, S. Design, optimisation and evaluation of in situ gelling nanoemulsion formulations of brinzolamide. Drug Deliv. Transl. Res. 2020, 10, 529–547. [Google Scholar] [CrossRef]
  229. Cegielska, O.; Sierakowski, M.; Sajkiewicz, P.; Lorenz, K.; Kogermann, K. Mucoadhesive brinzolamide-loaded nanofibers for alternative glaucoma treatment. Eur. J. Pharm. Biopharm. 2022, 180, 48–62. [Google Scholar] [CrossRef]
  230. Smith, S.M.; Salmon, J.H.; Abbaraju, S.; Amin, R.; Gilger, B.C. Tolerability, pharmacokinetics, and pharmacodynamics of a brinzolamide episcleral sustained release implant in normotensive New Zealand white rabbits. J. Drug Deliv. Sci. Technol. 2021, 61, 102123. [Google Scholar] [CrossRef]
  231. Teba, H.E.; Khalil, I.A.; El Sorogy, H.M. Novel cubosome based system for ocular delivery of acetazolamide. Drug Deliv. 2021, 28, 2177–2186. [Google Scholar] [CrossRef]
  232. Abdel-Rashid, R.S.; Helal, D.A.; Omar, M.M.; El Sisi, A.M. Nanogel loaded with surfactant based nanovesicles for enhanced ocular delivery of acetazolamide. Int. J. Nanomed. 2019, 14, 2973–2983. [Google Scholar] [CrossRef] [Green Version]
  233. Mazyed, E.A.; Abdelaziz, A.E. Fabrication of Transgelosomes for Enhancing the Ocular Delivery of Acetazolamide: Statistical Optimization, In Vitro Characterization, and In Vivo Study. Pharmaceutics 2020, 12, 465. [Google Scholar] [CrossRef] [PubMed]
  234. Morais, M.; Coimbra, P.; Pina, M.E. Comparative Analysis of Morphological and Release Profiles in Ocular Implants of Acetazolamide Prepared by Electrospinning. Pharmaceutics 2021, 13, 260. [Google Scholar] [CrossRef] [PubMed]
  235. El-Menshawe, S.F. A novel approach to topical acetazolamide/PEG 400 ocular niosomes. J. Drug Deliv. Sci. Technol. 2012, 22, 295–299. [Google Scholar] [CrossRef]
  236. Mohsen, A.M.; Salama, A.; Kassem, A.A. Development of acetazolamide loaded bilosomes for improved ocular delivery: Preparation, characterization and in vivo evaluation. J. Drug Deliv. Sci. Technol. 2020, 59, 101910. [Google Scholar] [CrossRef]
  237. Obiedallah, M.M.; Abdel-Mageed, A.M.; Elfaham, T.H. Ocular administration of acetazolamide microsponges in situ gel formulations. Saudi Pharm. J. 2018, 26, 909–920. [Google Scholar] [CrossRef] [PubMed]
  238. Mishra, V.; Jain, N.K. Acetazolamide encapsulated dendritic nano-architectures for effective glaucoma management in rabbits. Int. J. Pharm. 2014, 461, 380–390. [Google Scholar] [CrossRef]
  239. Lin, H.-R.; Yu, S.-P.; Kuo, C.-J.; Kao, H.-J.; Lo, Y.-L.; Lin, Y.-J. Pilocarpine-loaded chitosan-PAA nanosuspension for ophthalmic delivery. J. Biomater. Sci. Polym. Ed. 2007, 18, 205–221. [Google Scholar] [CrossRef]
  240. Suketu, D.; Desai, J.B. Pluronic® F127-Based Ocular Delivery System Containing Biodegradable Polyisobutylcyanoacrylate Nanocapsules of Pilocarpine. Drug Deliv. 2000, 7, 201–207. [Google Scholar] [CrossRef]
  241. Vandamme, T.F.; Brobeck, L. Poly(amidoamine) dendrimers as ophthalmic vehicles for ocular delivery of pilocarpine nitrate and tropicamide. J. Control. Release 2005, 102, 23–38. [Google Scholar] [CrossRef]
  242. Patel, C.C.; Mandava, N.; Oliver, S.C.N.; Braverman, R.; Quiroz-Mercado, H.; Olson, J.L. Treatment of Intractable Posterior Uveitis in Pediatric Patients with the Fluocinolone Acetonide Intravitreal Implant (Retisert). J. Retin. Vitr. Dis. 2012, 32, 537–542. [Google Scholar] [CrossRef]
  243. Salama, A.H.; Mahmoud, A.A.; Kamel, R. A Novel Method for Preparing Surface-Modified Fluocinolone Acetonide Loaded PLGA Nanoparticles for Ocular Use: In Vitro and In Vivo Evaluations. AAPS PharmSciTech 2016, 17, 1159–1172. [Google Scholar] [CrossRef] [PubMed]
  244. Shelley, H.; Annaji, M.; Grant, M.; Fasina, O.; Babu, R.J. Sustained Release Biodegradable Microneedles of Difluprednate for Delivery to Posterior Eye. J. Ocul. Pharmacol. Ther. 2022, 38, 449–458. [Google Scholar] [CrossRef] [PubMed]
  245. Gonzalez-Pizarro, R.; Parrotta, G.; Vera, R.; Sánchez-López, E.; Galindo, R.; Kjeldsen, F.; Badia, J.; Baldoma, L.; Espina, M.; García, M.L. Ocular penetration of fluorometholone-loaded PEG-PLGA nanoparticles functionalized with cell-penetrating peptides. Nanomedicine 2019, 14, 3089–3104. [Google Scholar] [CrossRef] [PubMed]
  246. Baba, K.; Hashida, N.; Tujikawa, M.; Quantock, A.J.; Nishida, K. The generation of fluorometholone nanocrystal eye drops, their metabolization to dihydrofluorometholone and penetration into rabbit eyes. Int. J. Pharm. 2021, 592, 120067. [Google Scholar] [CrossRef]
  247. Nirbhavane, P.; Sharma, G.; Singh, B.; Begum, G.; Jones, M.-C.; Rauz, S.; Vincent, R.; Denniston, A.K.; Hill, L.J.; Katare, O.P. Triamcinolone acetonide loaded-cationic nano-lipoidal formulation for uveitis: Evidences of improved biopharmaceutical performance and anti-inflammatory activity. Colloids Surf. B Biointerfaces 2020, 190, 110902. [Google Scholar] [CrossRef] [PubMed]
  248. Chen, Z.; Yang, M.; Wang, Q.; Bai, J.; McAlinden, C.; Skiadaresi, E.; Zhang, J.; Pan, L.; Mei, C.; Zeng, Z.; et al. Hydrogel eye drops as a non-invasive drug carrier for topical enhanced Adalimumab permeation and highly efficient uveitis treatment. Carbohydr. Polym. 2021, 253, 117216. [Google Scholar] [CrossRef] [PubMed]
  249. Zhang, R.; Qian, J.; Li, X.; Yuan, Y. Treatment of experimental autoimmune uveoretinitis with intravitreal injection of infliximab encapsulated in liposomes. Br. J. Ophthalmol. 2017, 101, 1731–1738. [Google Scholar] [CrossRef] [PubMed]
  250. Manna, S.; Donnell, A.M.; Faraj, R.Q.C.; Riemann, B.I.; Riemann, C.D.; Augsburger, J.J.; Correa, Z.M.; Banerjee, R.K. Pharmacokinetics and Toxicity Evaluation of a PLGA and Chitosan-Based Micro-Implant for Sustained Release of Methotrexate in Rabbit Vitreous. Pharmaceutics 2021, 13, 1227. [Google Scholar] [CrossRef]
  251. Paiva, M.R.B.D.; Vasconcelos-Santos, D.V.; Vieira, L.C.; Fialho, S.L.; Silva-Cunha, A. Sirolimus-Loaded Intravitreal Implant for Effective Treatment of Experimental Uveitis. AAPS PharmSciTech 2021, 22, 35. [Google Scholar] [CrossRef]
  252. Wu, W.; He, Z.; Zhang, Z.; Yu, X.; Song, Z.; Li, X. Intravitreal injection of rapamycin-loaded polymeric micelles for inhibition of ocular inflammation in rat model. Int. J. Pharm. 2016, 513, 238–246. [Google Scholar] [CrossRef]
  253. Li, H.; Zhang, Z.; Li, Y.; Su, L.; Duan, Y.; Zhang, H.; An, J.; Ni, T.; Li, X.; Zhang, X. Therapeutic Effect of Rapamycin-Loaded Small Extracellular Vesicles Derived from Mesenchymal Stem Cells on Experimental Autoimmune Uveitis. Front. Immunol. 2022, 13, 864956. [Google Scholar] [CrossRef]
  254. Silva, N.C.; Silva, S.; Sarmento, B.; Pintado, M. Chitosan nanoparticles for daptomycin delivery in ocular treatment of bacterial endophthalmitis. Drug Deliv. 2015, 22, 885–893. [Google Scholar] [CrossRef] [PubMed]
  255. Jounaki, K.; Makhmalzadeh, B.S.; Feghhi, M.; Heidarian, A. Topical ocular delivery of vancomycin loaded cationic lipid nanocarriers as a promising and non-invasive alternative approach to intravitreal injection for enhanced bacterial endophthalmitis management. Eur. J. Pharm. Sci. 2021, 167, 105991. [Google Scholar] [CrossRef] [PubMed]
  256. Cardoso, J.F.; Perasoli, F.B.; Almeida, T.C.; Marques, M.B.D.F.; Toledo, C.R.; Gil, P.O.; Tavares, H.D.S.; Da Paz, M.C.; Mussel, W.D.N.; Magalhães, J.T.; et al. Vancomycin-loaded N,N-dodecyl,methyl-polyethylenimine nanoparticles coated with hyaluronic acid to treat bacterial endophthalmitis: Development, characterization, and ocular biocompatibility. Int. J. Biol. Macromol. 2021, 169, 330–341. [Google Scholar] [CrossRef] [PubMed]
  257. Dosmar, E.; Liu, W.; Patel, G.; Rogozinski, A.; Mieler, W.F.; Kang-Mieler, J.J. Controlled Release of Vancomycin From a Thermoresponsive Hydrogel System for the Prophylactic Treatment of Postoperative Acute Endophthalmitis. Transl. Vis. Sci. Technol. 2019, 8, 53. [Google Scholar] [CrossRef] [Green Version]
  258. Abrishami, M.; Motamed Shariati, M.; Malaekeh-Nikouei, B.; Tajani, A.S.; Mahmoudi, A.; Abrishami, M.; Khameneh, B. Preparation and in vivo evaluation of nanoliposomes containing vancomycin after intravitreal injection in albino rabbits. Iran. J. Basic Med. Sci. 2020, 23, 551–555. [Google Scholar] [CrossRef]
  259. Resende, A.F.C.; Pereira, A.F.; Moreira, T.P.; Patrício, P.S.O.; Fialho, S.L.; Cunha, G.M.F.; Silva-Cunha, A.; Magalhães, J.T.; Silva, G.R. PLGA Implants containing vancomycin and dexamethasone: Development, characterization and bactericidal effects. Pharm.-Int. J. Pharm. Sci. 2016, 71, 439–446. [Google Scholar] [CrossRef]
  260. Allam, A.; El-Mokhtar, M.A.; Elsabahy, M. Vancomycin-loaded niosomes integrated within pH-sensitive in-situ forming gel for treatment of ocular infections while minimizing drug irritation. J. Pharm. Pharmacol. 2019, 71, 1209–1221. [Google Scholar] [CrossRef]
  261. Mohammadpour, M.; Jabbarvand, M.; Karimi, N. Therapeutic possibilities of ceftazidime nanoparticles in devastating pseudomonas ophthalmic infections; keratitis and endophthalmitis. Med. Hypothesis Discov. Innov. Ophthalmol. J. 2012, 1, 6–9. [Google Scholar]
  262. Bae, J.H.; Lee, S.C. Intravitreal liposomal amphotericin B for treatment of endogenous candida endophthalmitis. Jpn. J. Ophthalmol. 2015, 59, 346–352. [Google Scholar] [CrossRef]
  263. Üstündağ Okur, N.; Yozgatlı, V.; Okur, M.E.; Yoltaş, A.; Siafaka, P.I. Improving therapeutic efficacy of voriconazole against fungal keratitis: Thermo-sensitive in situ gels as ophthalmic drug carriers. J. Drug Deliv. Sci. Technol. 2019, 49, 323–333. [Google Scholar] [CrossRef]
  264. Kumar, R.; Sinha, V.R. Solid lipid nanoparticle: An efficient carrier for improved ocular permeation of voriconazole. Drug Dev. Ind. Pharm. 2016, 42, 1956–1967. [Google Scholar] [CrossRef]
  265. Bhosale, R.; Bhandwalkar, O.; Duduskar, A.; Jadhav, R.; Pawar, P. Water soluble chitosan mediated voriconazole microemulsion as sustained carrier for ophthalmic application: In vitro/ex vivo/in vivo evaluations. Open Pharm. Sci. J. 2016, 3, 215–234. [Google Scholar] [CrossRef] [Green Version]
  266. Fahmy, A.M.; Hassan, M.; El-Setouhy, D.A.; Tayel, S.A.; Al-mahallawi, A.M. Voriconazole Ternary Micellar Systems for the Treatment of Ocular Mycosis: Statistical Optimization and In Vivo Evaluation. J. Pharm. Sci. 2021, 110, 2130–2138. [Google Scholar] [CrossRef] [PubMed]
  267. de Sá, F.A.P.; Taveira, S.F.; Gelfuso, G.M.; Lima, E.M.; Gratieri, T. Liposomal voriconazole (VOR) formulation for improved ocular delivery. Colloids Surf. B Biointerfaces 2015, 133, 331–338. [Google Scholar] [CrossRef] [PubMed]
  268. Ma, F.; Nan, K.; Lee, S.; Beadle, J.R.; Hou, H.; Freeman, W.R.; Hostetler, K.Y.; Cheng, L. Micelle formulation of hexadecyloxypropyl-cidofovir (HDP-CDV) as an intravitreal long-lasting delivery system. Eur. J. Pharm. Biopharm. 2015, 89, 271–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Kuppermann, B.D.; Assil, K.K.; Vuong, C.; Besen, G.; Wiley, C.A.; De Clercq, E.; Bergeron-Lynn, G.; Connor, J.D.; Pursley, M.; Munguia, D.; et al. Liposome-Encapsulated (S)-1-(3-Hydroxy-2-Phosphonylmethoxypropyl)cytosine for Long-Acting Therapy of Viral Retinitis. J. Infect. Dis. 1996, 173, 18–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Claro, C.; Ruiz, R.; Cordero, E.; Pastor, M.T.; López-Cortés, L.F.; Jiménez-Castellanos, M.R.; Lucero, M.J. Determination and pharmacokinetic profile of liposomal foscarnet in rabbit ocular tissues after intravitreal administration. Exp. Eye Res. 2009, 88, 528–534. [Google Scholar] [CrossRef]
  271. Peter, S.; Mathews, M.M.; Saju, F.; Paul, S. Development, Optimization and In Vitro Characterization of Eudragit-Ganciclovir Nanosuspension or Treating Herpes Simplex Keratitis. J. Pharm. Innov. 2023, 1–10. [Google Scholar] [CrossRef]
  272. Naguib, M.J.; Hassan, Y.R.; Abd-Elsalam, W.H. 3D printed ocusert laden with ultra-fluidic glycerosomes of ganciclovir for the management of ocular cytomegalovirus retinitis. Int. J. Pharm. 2021, 607, 121010. [Google Scholar] [CrossRef]
  273. Choudhari, M.; Nayak, K.; Nagai, N.; Nakazawa, Y.; Khunt, D.; Misra, M. Role of mucoadhesive agent in ocular delivery of ganciclovir microemulsion: Cytotoxicity evaluation in vitro and ex vivo. Int. Ophthalmol. 2022, 43, 1153–1167. [Google Scholar] [CrossRef]
  274. Yasukawa, T.; Ogura, Y.; Kimura, H.; Sakurai, E.; Tabata, Y. Drug delivery from ocular implants. Expert Opin. Drug Deliv. 2006, 3, 261–273. [Google Scholar] [CrossRef]
  275. Choonara, Y.E.; Pillay, V.; Carmichael, T.; Danckwerts, M.P. An in vitro study of the design and development of a novel doughnut-shaped minitablet for intraocular implantation. Int. J. Pharm. 2006, 310, 15–24. [Google Scholar] [CrossRef]
  276. Yan, J.; Peng, X.; Cai, Y.; Cong, W. Development of facile drug delivery platform of ranibizumab fabricated PLGA-PEGylated magnetic nanoparticles for age-related macular degeneration therapy. J. Photochem. Photobiol. B Biol. 2018, 183, 133–136. [Google Scholar] [CrossRef] [PubMed]
  277. Elsaid, N.; Jackson, T.L.; Elsaid, Z.; Alqathama, A.; Somavarapu, S. PLGA Microparticles Entrapping Chitosan-Based Nanoparticles for the Ocular Delivery of Ranibizumab. Mol. Pharm. 2016, 13, 2923–2940. [Google Scholar] [CrossRef]
  278. Joseph, R.R.; Tan, D.W.N.; Ramon, M.R.M.; Natarajan, J.V.; Agrawal, R.; Wong, T.T.; Venkatraman, S.S. Characterization of liposomal carriers for the trans-scleral transport of Ranibizumab. Sci. Rep. 2017, 7, 16803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Qian, C.; Yan, P.; Wan, G.; Liang, S.; Dong, Y.; Wang, J. Facile synthetic Photoluminescent Graphene Quantum dots encapsulated β-cyclodextrin drug carrier system for the management of macular degeneration: Detailed analytical and biological investigations. J. Photochem. Photobiol. B Biol. 2018, 189, 244–249. [Google Scholar] [CrossRef] [PubMed]
  280. Campochiaro, P.A.; Gune, S.; Maia, M.; Ding, H.T.; Maass, K. Pharmacokinetic profile of the Port Delivery System with ranibizumab (PDS) in the phase 3 Archway trial. Investig. Ophthalmol. Vis. Sci. 2021, 62, 350. [Google Scholar]
  281. Yandrapu, S.K.; Upadhyay, A.K.; Petrash, J.M.; Kompella, U.B. Nanoparticles in Porous Microparticles Prepared by Supercritical Infusion and Pressure Quench Technology for Sustained Delivery of Bevacizumab. Mol. Pharm. 2013, 10, 4676–4686. [Google Scholar] [CrossRef] [Green Version]
  282. Jiang, P.; Chaparro, F.J.; Cuddington, C.T.; Palmer, A.F.; Ohr, M.P.; Lannutti, J.J.; Swindle-Reilly, K.E. Injectable biodegradable bi-layered capsule for sustained delivery of bevacizumab in treating wet age-related macular degeneration. J. Control. Release 2020, 320, 442–456. [Google Scholar] [CrossRef]
  283. Zhang, X.-P.; Sun, J.-G.; Yao, J.; Shan, K.; Liu, B.-H.; Yao, M.-D.; Ge, H.-M.; Jiang, Q.; Zhao, C.; Yan, B. Effect of nanoencapsulation using poly (lactide-co-glycolide) (PLGA) on anti-angiogenic activity of bevacizumab for ocular angiogenesis therapy. Biomed. Pharmacother. 2018, 107, 1056–1063. [Google Scholar] [CrossRef] [PubMed]
  284. Badiee, P.; Varshochian, R.; Rafiee-Tehrani, M.; Abedin Dorkoosh, F.; Khoshayand, M.R.; Dinarvand, R. Ocular implant containing bevacizumab-loaded chitosan nanoparticles intended for choroidal neovascularization treatment. J. Biomed. Mater. Res. 2018, 106, 2261–2271. [Google Scholar] [CrossRef] [PubMed]
  285. Jiang, P.; Jacobs, K.M.; Ohr, M.P.; Swindle-Reilly, K.E. Chitosan–Polycaprolactone Core–Shell Microparticles for Sustained Delivery of Bevacizumab. Mol. Pharm. 2020, 17, 2570–2584. [Google Scholar] [CrossRef]
  286. Karumanchi, D.K.; Skrypai, Y.; Thomas, A.; Gaillard, E.R. Rational design of liposomes for sustained release drug delivery of bevacizumab to treat ocular angiogenesis. J. Drug Deliv. Sci. Technol. 2018, 47, 275–282. [Google Scholar] [CrossRef]
  287. Kelly, S.J.; Hirani, A.; Shahidadpury, V.; Solanki, A.; Halasz, K.; Varghese Gupta, S.; Madow, B.; Sutariya, V. Aflibercept Nanoformulation Inhibits VEGF Expression in Ocular In Vitro Model: A Preliminary Report. Biomedicines 2018, 6, 92. [Google Scholar] [CrossRef] [Green Version]
  288. Liu, W.; Lee, B.-S.; Mieler, W.F.; Kang-Mieler, J.J. Biodegradable Microsphere-Hydrogel Ocular Drug Delivery System for Controlled and Extended Release of Bioactive Aflibercept In Vitro. Curr. Eye Res. 2019, 44, 264–274. [Google Scholar] [CrossRef] [PubMed]
  289. Bhatt, P.; Narvekar, P.; Lalani, R.; Chougule, M.B.; Pathak, Y.; Sutariya, V. An in vitro Assessment of Thermo-Reversible Gel Formulation Containing Sunitinib Nanoparticles for Neovascular Age-Related Macular Degeneration. AAPS PharmSciTech 2019, 20, 281. [Google Scholar] [CrossRef]
  290. Streets, J.; Bhatt, P.; Bhatia, D.; Sutariya, V. Sunitinib-Loaded MPEG-PCL Micelles for the Treatment of Age-Related Macular Degeneration. Sci. Pharm. 2020, 88, 30. [Google Scholar] [CrossRef]
  291. Narvekar, P.; Bhatt, P.; Fnu, G.; Sutariya, V.J.A.; technologies, d.d. Axitinib-loaded poly (lactic-co-glycolic acid) nanoparticles for age-related macular degeneration: Formulation development and in vitro characterization. ASSAY Drug Dev. Technol. 2019, 17, 167–177. [Google Scholar] [CrossRef]
  292. Chen, C.W.; Yeh, M.K.; Shiau, C.Y.; Chiang, C.H.; Lu, D.W. Efficient downregulation of VEGF in retinal pigment epithelial cells by integrin ligand-labeled liposome-mediated siRNA delivery. Int. J. Nanomed. 2013, 8, 2613–2627. [Google Scholar] [CrossRef] [Green Version]
  293. Ryoo, N.K.; Lee, J.; Lee, H.; Hong, H.K.; Kim, H.; Lee, J.B.; Woo, S.J.; Park, K.H.; Kim, H. Therapeutic effects of a novel siRNA-based anti-VEGF (siVEGF) nanoball for the treatment of choroidal neovascularization. Nanoscale 2017, 9, 15461–15469. [Google Scholar] [CrossRef] [PubMed]
  294. Wang, Z.; Liu, A.; Zhang, H.; Wang, M.; Tang, Q.; Huang, Y.; Wang, L. Inhibition of retinal neovascularization by VEGF siRNA delivered via bioreducible lipid-like nanoparticles. Graefe’s Arch. Clin. Exp. Ophthalmol. 2020, 258, 2407–2418. [Google Scholar] [CrossRef] [PubMed]
  295. Kim, H.; Csaky, K.G. Nanoparticle–integrin antagonist C16Y peptide treatment of choroidal neovascularization in rats. J. Control. Release 2010, 142, 286–293. [Google Scholar] [CrossRef]
  296. Jayaraman, M.S.; Bharali, D.J.; Sudha, T.; Mousa, S.A. Nano chitosan peptide as a potential therapeutic carrier for retinal delivery to treat age-related macular degeneration. Mol. Vis. 2012, 18, 2300–2308. [Google Scholar] [PubMed]
  297. Nguyen, D.D.; Luo, L.-J.; Yang, C.-J.; Lai, J.-Y. Highly Retina-Permeating and Long-Acting Resveratrol/Metformin Nanotherapeutics for Enhanced Treatment of Macular Degeneration. ACS Nano 2023, 17, 168–183. [Google Scholar] [CrossRef]
  298. Ayyagari, S.; Dar, H.; Morton, V.; Moy, K.; Patel, C.; Ramasubramanian, L.; Ravi, N.; Wood, S.; Zhao, A.; Zheng, M. Evaluation of Curcumin-Loaded Nanoliposomes for the Treatment and Prevention of Age-Related Macular Degeneration. Ph.D. Thesis, University of Maryland, College Park, MD, USA, 2017. [Google Scholar]
  299. Sun, R.; Zhang, A.; Ge, Y.; Gou, J.; Yin, T.; He, H.; Wang, Y.; Zhang, G.; Kong, J.; Shang, L.; et al. Ultra-small-size Astragaloside-IV loaded lipid nanocapsules eye drops for the effective management of dry age-related macular degeneration. Expert Opin. Drug Deliv. 2020, 17, 1305–1320. [Google Scholar] [CrossRef]
  300. Oh, E.J.; Choi, J.-S.; Kim, H.; Joo, C.-K.; Hahn, S.K. Anti-Flt1 peptide–Hyaluronate conjugate for the treatment of retinal neovascularization and diabetic retinopathy. Biomaterials 2011, 32, 3115–3123. [Google Scholar] [CrossRef]
  301. Wang, C.; Seo, S.-J.; Kim, J.-S.; Lee, S.-H.; Jeon, J.-K.; Kim, J.-W.; Kim, K.-H.; Kim, J.-K.; Park, J. Intravitreal implantable magnetic micropump for on-demand VEGFR-targeted drug delivery. J. Control. Release 2018, 283, 105–112. [Google Scholar] [CrossRef]
  302. Qiu, F.; Meng, T.; Chen, Q.; Zhou, K.; Shao, Y.; Matlock, G.; Ma, X.; Wu, W.; Du, Y.; Wang, X.; et al. Fenofibrate-Loaded Biodegradable Nanoparticles for the Treatment of Experimental Diabetic Retinopathy and Neovascular Age-Related Macular Degeneration. Mol. Pharm. 2019, 16, 1958–1970. [Google Scholar] [CrossRef]
  303. Laddha, U.D.; Kshirsagar, S.J. Formulation of PPAR-gamma agonist as surface modified PLGA nanoparticles for non-invasive treatment of diabetic retinopathy: In vitro and in vivo evidences. Heliyon 2020, 6, e04589. [Google Scholar] [CrossRef]
  304. Radwan, S.E.; El-Kamel, A.; Zaki, E.I.; Burgalassi, S.; Zucchetti, E.; El-Moslemany, R.M. Hyaluronic-Coated Albumin Nanoparticles for the Non-Invasive Delivery of Apatinib in Diabetic Retinopathy. Int. J. Nanomed. 2021, 16, 4481–4494. [Google Scholar] [CrossRef] [PubMed]
  305. Jo, D.H.; Kim, J.H.; Yu, Y.S.; Lee, T.G.; Kim, J.H. Antiangiogenic effect of silicate nanoparticle on retinal neovascularization induced by vascular endothelial growth factor. Nanomed. Nanotechnol. Biol. Med. 2012, 8, 784–791. [Google Scholar] [CrossRef] [PubMed]
  306. Paiva, M.R.B.; Andrade, G.F.; Dourado, L.F.N.; Castro, B.F.M.; Fialho, S.L.; Sousa, E.M.B.; Silva-Cunha, A. Surface functionalized mesoporous silica nanoparticles for intravitreal application of tacrolimus. J. Biomater. Appl. 2021, 35, 1019–1033. [Google Scholar] [CrossRef]
  307. Dave, V.; Sharma, R.; Gupta, C.; Sur, S. Folic acid modified gold nanoparticle for targeted delivery of Sorafenib tosylate towards the treatment of diabetic retinopathy. Colloids Surf. B Biointerfaces 2020, 194, 111151. [Google Scholar] [CrossRef] [PubMed]
  308. Amato, R.; Giannaccini, M.; Dal Monte, M.; Cammalleri, M.; Pini, A.; Raffa, V.; Lulli, M.; Casini, G. Association of the Somatostatin Analog Octreotide with Magnetic Nanoparticles for Intraocular Delivery: A Possible Approach for the Treatment of Diabetic Retinopathy. Front. Bioeng. Biotechnol. 2020, 8, 144. [Google Scholar] [CrossRef]
  309. Venkatesan, A.; Roy, A.; Kulandaivel, S.; Natesan, V.; Kim, S.-J. p-Coumaric Acid Nanoparticles Ameliorate Diabetic Nephropathy via Regulating mRNA Expression of KIM-1 and GLUT-2 in Streptozotocin-Induced Diabetic Rats. Metabolites 2022, 12, 1166. [Google Scholar] [CrossRef]
  310. Huang, D.; Chen, Y.-S.; Thakur, S.S.; Rupenthal, I.D. Ultrasound-mediated nanoparticle delivery across ex vivo bovine retina after intravitreal injection. Eur. J. Pharm. Biopharm. 2017, 119, 125–136. [Google Scholar] [CrossRef]
  311. Rassu, G.; Pavan, B.; Mandracchia, D.; Tripodo, G.; Botti, G.; Dalpiaz, A.; Gavini, E.; Giunchedi, P. Polymeric nanomicelles based on inulin D α-tocopherol succinate for the treatment of diabetic retinopathy. J. Drug Deliv. Sci. Technol. 2021, 61, 102286. [Google Scholar] [CrossRef]
  312. Bogdanov, P.; Sampedro, J.; Solà-Adell, C.; Simó-Servat, O.; Russo, C.; Varela-Sende, L.; Simó, R.; Hernández, C. Effects of Liposomal Formulation of Citicoline in Experimental Diabetes-Induced Retinal Neurodegeneration. Int. J. Mol. Sci. 2018, 19, 2458. [Google Scholar] [CrossRef] [Green Version]
  313. Mostafa, M.; Alaaeldin, E.; Aly, U.F.; Sarhan, H.A. Optimization and Characterization of Thymoquinone-Loaded Liposomes with Enhanced Topical Anti-inflammatory Activity. AAPS PharmSciTech 2018, 19, 3490–3500. [Google Scholar] [CrossRef]
  314. Zheng, C.; Luo, W.; Liu, Y.; Chen, J.; Deng, H.; Zhou, Z.; Shen, J. Killing three birds with one stone: Multi-stage metabolic regulation mediated by clinically usable berberine liposome to overcome photodynamic immunotherapy resistance. Chem. Eng. J. 2023, 454, 140164. [Google Scholar] [CrossRef]
  315. Huang, X.; Li, M.; Bruni, R.; Messa, P.; Cellesi, F. The effect of thermosensitive liposomal formulations on loading and release of high molecular weight biomolecules. Int. J. Pharm. 2017, 524, 279–289. [Google Scholar] [CrossRef]
  316. López-Cano, J.J.; González-Cela-Casamayor, M.A.; Andrés-Guerrero, V.; Herrero-Vanrell, R.; Molina-Martínez, I.T. Liposomes as vehicles for topical ophthalmic drug delivery and ocular surface protection. Expert Opin. Drug Deliv. 2021, 18, 819–847. [Google Scholar] [CrossRef] [PubMed]
  317. Christensen, G.; Barut, L.; Urimi, D.; Schipper, N.; Paquet-Durand, F. Investigating Ex Vivo Animal Models to Test the Performance of Intravitreal Liposomal Drug Delivery Systems. Pharmaceutics 2021, 13, 1013. [Google Scholar] [CrossRef]
  318. Salas-Ambrosio, P.J.; Bernad-Bernad, M.J.; Linares-Alba, M.A.; García-Santisteban, R.; Tonix-Aburto, L.A.; Ornelas-Lobato, G.J.; Gracia-Mora, I.; Rivera-Huerta, M.; Sánchez-Bartez, F.; Rico-Morales, H.; et al. Toxicity Evaluation of a Novel Rapamycin Liposomal Formulation After Subconjunctival and Intravitreal Injection. J. Ocul. Pharmacol. Ther. 2021, 37, 261–276. [Google Scholar] [CrossRef]
  319. Blazaki, S.; Pachis, K.; Tzatzarakis, M.; Tsilimbaris, M.; Antimisiaris, S.G. Novel Liposome Aggregate Platform (LAP) system for sustained retention of drugs in the posterior ocular segment following intravitreal injection. Int. J. Pharm. 2020, 576, 118987. [Google Scholar] [CrossRef] [PubMed]
  320. Terreni, E.; Chetoni, P.; Tampucci, S.; Burgalassi, S.; Al-kinani, A.A.; Alany, R.G.; Monti, D. Assembling Surfactants-Mucoadhesive Polymer Nanomicelles (ASMP-Nano) for Ocular Delivery of Cyclosporine-A. Pharmaceutics 2020, 12, 253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  321. Binkhathlan, Z.; Alomrani, A.H.; Hoxha, O.; Ali, R.; Kalam, M.A.; Alshamsan, A. Development and Characterization of PEGylated Fatty Acid-Block-Poly(ε-caprolactone) Novel Block Copolymers and Their Self-Assembled Nanostructures for Ocular Delivery of Cyclosporine A. Polymers 2022, 14, 1635. [Google Scholar]
  322. Lynch, C.; Kondiah, P.P.D.; Choonara, Y.E.; du Toit, L.C.; Ally, N.; Pillay, V. Advances in Biodegradable Nano-Sized Polymer-Based Ocular Drug Delivery. Polymers 2019, 11, 1371. [Google Scholar] [CrossRef] [Green Version]
  323. Suri, R.; Neupane, Y.R.; Mehra, N.; Nematullah, M.; Khan, F.; Alam, O.; Iqubal, A.; Jain, G.K.; Kohli, K. Sirolimus loaded chitosan functionalized poly (lactic-co-glycolic acid) (PLGA) nanoparticles for potential treatment of age-related macular degeneration. Int. J. Biol. Macromol. 2021, 191, 548–559. [Google Scholar] [CrossRef]
  324. Romeo, A.; Musumeci, T.; Carbone, C.; Bonaccorso, A.; Corvo, S.; Lupo, G.; Anfuso, C.D.; Puglisi, G.; Pignatello, R. Ferulic Acid-Loaded Polymeric Nanoparticles for Potential Ocular Delivery. Pharmaceutics 2021, 13, 687. [Google Scholar] [CrossRef]
  325. Bhatt, P.; Fnu, G.; Bhatia, D.; Shahid, A.; Sutariya, V. Nanodelivery of Resveratrol-Loaded PLGA Nanoparticles for Age-Related Macular Degeneration. AAPS PharmSciTech 2020, 21, 291. [Google Scholar] [CrossRef]
  326. Mahaling, B.; Baruah, N.; Ahamad, N.; Maisha, N.; Lavik, E.; Katti, D.S. A non-invasive nanoparticle-based sustained dual-drug delivery system as an eyedrop for endophthalmitis. Int. J. Pharm. 2021, 606, 120900. [Google Scholar] [CrossRef]
  327. Bu, H.-Z.; Gukasyan, H.J.; Goulet, L.; Lou, X.-J.; Xiang, C.; Koudriakova, T. Ocular Disposition, Pharmacokinetics, Efficacy and Safety of Nanoparticle-Formulated Ophthalmic Drugs. Curr. Drug Metab. 2007, 8, 91–107. [Google Scholar] [CrossRef]
  328. Yan, R.; Xu, L.; Wang, Q.; Wu, Z.; Zhang, H.; Gan, L. Cyclosporine A Nanosuspensions for Ophthalmic Delivery: A Comparative Study between Cationic Nanoparticles and Drug-Core Mucus Penetrating Nanoparticles. Mol. Pharm. 2021, 18, 4290–4298. [Google Scholar] [CrossRef]
  329. Shah, S.; Bhanderi, B.; Soniwala, M.; Chavda, J. Lutein-Loaded Solid Lipid Nanoparticles for Ocular Delivery: Statistical Optimization and Ex Vivo Evaluation. J. Pharm. Innov. 2022, 17, 584–598. [Google Scholar] [CrossRef]
  330. Onugwu, A.L.; Attama, A.A.; Nnamani, P.O.; Onugwu, S.O.; Onuigbo, E.B.; Khutoryanskiy, V.V. Development and optimization of solid lipid nanoparticles coated with chitosan and poly(2-ethyl-2-oxazoline) for ocular drug delivery of ciprofloxacin. J. Drug Deliv. Sci. Technol. 2022, 74, 103527. [Google Scholar] [CrossRef]
  331. Ghasemiyeh, P.; Mohammadi-Samani, S. Solid lipid nanoparticles and nanostructured lipid carriers as novel drug delivery systems: Applications, advantages and disadvantages. Res. Pharm. Sci. 2018, 13, 288–303. [Google Scholar] [CrossRef]
  332. Alhakamy, N.A.; Hosny, K.M.; Aldryhim, A.Y.; Rizg, W.Y.; Eshmawi, B.A.; Bukhary, H.A.; Murshid, S.S.A.; Khallaf, R.A. Development and optimization of ofloxacin as solid lipid nanoparticles for enhancement of its ocular activity. J. Drug Deliv. Sci. Technol. 2022, 72, 103373. [Google Scholar] [CrossRef]
  333. Eid, H.M.; Elkomy, M.H.; El Menshawe, S.F.; Salem, H.F. Development, Optimization, and In Vitro/In Vivo Characterization of Enhanced Lipid Nanoparticles for Ocular Delivery of Ofloxacin: The Influence of Pegylation and Chitosan Coating. AAPS PharmSciTech 2019, 20, 183. [Google Scholar] [CrossRef] [PubMed]
  334. Shahab, M.S.; Rizwanullah, M.; Sarim Imam, S. Formulation, optimization and evaluation of vitamin E TPGS emulsified dorzolamide solid lipid nanoparticles. J. Drug Deliv. Sci. Technol. 2022, 68, 103062. [Google Scholar] [CrossRef]
  335. Fathi, M.; Barar, J.; Aghanejad, A.; Omidi, Y. Hydrogels for ocular drug delivery and tissue engineering. BioImpacts BI 2015, 5, 159–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Annala, A.; Ilochonwu, B.C.; Wilbie, D.; Sadeghi, A.; Hennink, W.E.; Vermonden, T. Self-Healing Thermosensitive Hydrogel for Sustained Release of Dexamethasone for Ocular Therapy. ACS Polym. Au 2023, 3, 118–131. [Google Scholar] [CrossRef] [PubMed]
  337. Ilochonwu, B.C.; Mihajlovic, M.; Maas-Bakker, R.F.; Rousou, C.; Tang, M.; Chen, M.; Hennink, W.E.; Vermonden, T. Hyaluronic Acid-PEG-Based Diels–Alder In Situ Forming Hydrogels for Sustained Intraocular Delivery of Bevacizumab. Biomacromolecules 2022, 23, 2914–2929. [Google Scholar] [CrossRef]
  338. Pakzad, Y.; Fathi, M.; Omidi, Y.; Mozafari, M.; Zamanian, A. Synthesis and characterization of timolol maleate-loaded quaternized chitosan-based thermosensitive hydrogel: A transparent topical ocular delivery system for the treatment of glaucoma. Int. J. Biol. Macromol. 2020, 159, 117–128. [Google Scholar] [CrossRef]
  339. Xu, Q.; Kambhampati, S.P.; Kannan, R.M. Nanotechnology approaches for ocular drug delivery. Middle East Afr. J. Ophthalmol. 2013, 20, 26–37. [Google Scholar] [CrossRef] [Green Version]
  340. Honda, M.; Asai, T.; Oku, N.; Araki, Y.; Tanaka, M.; Ebihara, N. Liposomes and nanotechnology in drug development: Focus on ocular targets. Int. J. Nanomed. 2013, 8, 495–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Rodríguez Villanueva, J.; Navarro, M.G.; Rodríguez Villanueva, L. Dendrimers as a promising tool in ocular therapeutics: Latest advances and perspectives. Int. J. Pharm. 2016, 511, 359–366. [Google Scholar] [CrossRef]
  342. Yavuz, B.; Pehlivan, S.B.; Vural, İ.; Ünlü, N. In Vitro/In Vivo Evaluation of Dexamethasone—PAMAM Dendrimer Complexes for Retinal Drug Delivery. J. Pharm. Sci. 2015, 104, 3814–3823. [Google Scholar] [CrossRef]
  343. Wang, J.; Li, B.; Huang, D.; Norat, P.; Grannonico, M.; Cooper, R.C.; Gui, Q.; Nam Chow, W.; Liu, X.; Yang, H. Nano-in-Nano dendrimer gel particles for efficient topical delivery of antiglaucoma drugs into the eye. Chem. Eng. J. 2021, 425, 130498. [Google Scholar] [CrossRef]
  344. Deng, S.; Gigliobianco, M.R.; Censi, R.; Di Martino, P. Polymeric Nanocapsules as Nanotechnological Alternative for Drug Delivery System: Current Status, Challenges and Opportunities. Nanomaterials 2020, 10, 847. [Google Scholar] [CrossRef] [PubMed]
  345. Gao, L.; Liu, G.; Ma, J.; Wang, X.; Zhou, L.; Li, X.; Wang, F. Application of Drug Nanocrystal Technologies on Oral Drug Delivery of Poorly Soluble Drugs. Pharm. Res. 2013, 30, 307–324. [Google Scholar] [CrossRef] [PubMed]
  346. McGuckin, M.B.; Wang, J.; Ghanma, R.; Qin, N.; Palma, S.D.; Donnelly, R.F.; Paredes, A.J. Nanocrystals as a master key to deliver hydrophobic drugs via multiple administration routes. J. Control. Release 2022, 345, 334–353. [Google Scholar] [CrossRef] [PubMed]
  347. Sharma, O.P.; Patel, V.; Mehta, T. Nanocrystal for ocular drug delivery: Hope or hype. Drug Deliv. Transl. Res. 2016, 6, 399–413. [Google Scholar] [CrossRef] [PubMed]
  348. Romero, G.B.; Keck, C.M.; Müller, R.H.; Bou-Chacra, N.A. Development of cationic nanocrystals for ocular delivery. Eur. J. Pharm. Biopharm. 2016, 107, 215–222. [Google Scholar] [CrossRef]
  349. Tetyczka, C.; Brisberger, K.; Reiser, M.; Zettl, M.; Jeitler, R.; Winter, C.; Kolb, D.; Leitinger, G.; Spoerk, M.; Roblegg, E. Itraconazole Nanocrystals on Hydrogel Contact Lenses via Inkjet Printing: Implications for Ophthalmic Drug Delivery. ACS Appl. Nano Mater. 2022, 5, 9435–9446. [Google Scholar] [CrossRef]
  350. Kalam, M.A.; Iqbal, M.; Alshememry, A.; Alkholief, M.; Alshamsan, A. Fabrication and Characterization of Tedizolid Phosphate Nanocrystals for Topical Ocular Application: Improved Solubilization and In Vitro Drug Release. Pharmaceutics 2022, 14, 1328. [Google Scholar] [CrossRef]
  351. Tan, C.; Hosseini, S.F.; Jafari, S.M. Cubosomes and Hexosomes as Novel Nanocarriers for Bioactive Compounds. J. Agric. Food Chem. 2022, 70, 1423–1437. [Google Scholar] [CrossRef]
  352. Gaballa, S.A.; El Garhy, O.H.; Moharram, H.; Abdelkader, H. Preparation and Evaluation of Cubosomes/Cubosomal Gels for Ocular Delivery of Beclomethasone Dipropionate for Management of Uveitis. Pharm. Res. 2020, 37, 198. [Google Scholar] [CrossRef]
  353. Karami, Z.; Hamidi, M. Cubosomes: Remarkable drug delivery potential. Drug Discov. Today 2016, 21, 789–801. [Google Scholar] [CrossRef]
  354. Han, S.; Shen, J.-Q.; Gan, Y.; Geng, H.-M.; Zhang, X.-X.; Zhu, C.-L.; Gan, L. Novel vehicle based on cubosomes for ophthalmic delivery of flurbiprofen with low irritancy and high bioavailability. Acta Pharmacol. Sin. 2010, 31, 990–998. [Google Scholar] [CrossRef]
  355. Eldeeb, A.E.; Salah, S.; Ghorab, M. Formulation and evaluation of cubosomes drug delivery system for treatment of glaucoma: Ex-vivo permeation and in-vivo pharmacodynamic study. J. Drug Deliv. Sci. Technol. 2019, 52, 236–247. [Google Scholar] [CrossRef]
  356. Yasamineh, S.; Yasamineh, P.; Ghafouri Kalajahi, H.; Gholizadeh, O.; Yekanipour, Z.; Afkhami, H.; Eslami, M.; Hossein Kheirkhah, A.; Taghizadeh, M.; Yazdani, Y.; et al. A state-of-the-art review on the recent advances of niosomes as a targeted drug delivery system. Int. J. Pharm. 2022, 624, 121878. [Google Scholar] [CrossRef]
  357. Saettone, M.F.; Perini, G.; Carafa, M.; Santucci, E.; Alhaique, F. Nonionic surfactant vesicles as ophthalmic carriers for cyclopentolate: Preliminary evaluation. STP Pharma. Sci. 1996, 6, 94–98. [Google Scholar]
  358. Shukr, M.H. Novel in situ gelling ocular inserts for voriconazole-loaded niosomes: Design, in vitro characterisation and in vivo evaluation of the ocular irritation and drug pharmacokinetics. J. Microencapsul. 2016, 33, 71–79. [Google Scholar] [CrossRef]
  359. Aggarwal, D.; Pal, D.; Mitra, A.K.; Kaur, I.P. Study of the extent of ocular absorption of acetazolamide from a developed niosomal formulation, by microdialysis sampling of aqueous humor. Int. J. Pharm. 2007, 338, 21–26. [Google Scholar] [CrossRef]
  360. Abdelbary, G.; El-gendy, N. Niosome-Encapsulated Gentamicin for Ophthalmic Controlled Delivery. AAPS PharmSciTech 2008, 9, 740–747. [Google Scholar] [CrossRef]
  361. Gupta, P.; Yadav, K.S. Formulation and evaluation of brinzolamide encapsulated niosomal in-situ gel for sustained reduction of IOP in rabbits. J. Drug Deliv. Sci. Technol. 2022, 67, 103004. [Google Scholar] [CrossRef]
  362. Owodeha-Ashaka, K.; Ilomuanya, M.O.; Iyire, A. Evaluation of sonication on stability-indicating properties of optimized pilocarpine hydrochloride-loaded niosomes in ocular drug delivery. Prog. Biomater. 2021, 10, 207–220. [Google Scholar] [CrossRef]
  363. Zeng, W.; Li, Q.; Wan, T.; Liu, C.; Pan, W.; Wu, Z.; Zhang, G.; Pan, J.; Qin, M.; Lin, Y.; et al. Hyaluronic acid-coated niosomes facilitate tacrolimus ocular delivery: Mucoadhesion, precorneal retention, aqueous humor pharmacokinetics, and transcorneal permeability. Colloids Surf. B Biointerfaces 2016, 141, 28–35. [Google Scholar] [CrossRef] [PubMed]
  364. Puras, G.; Mashal, M.; Zárate, J.; Agirre, M.; Ojeda, E.; Grijalvo, S.; Eritja, R.; Diaz-Tahoces, A.; Martínez Navarrete, G.; Avilés-Trigueros, M.; et al. A novel cationic niosome formulation for gene delivery to the retina. J. Control. Release 2014, 174, 27–36. [Google Scholar] [CrossRef] [PubMed]
  365. Peng, C.C.; Bengani, L.C.; Jung, H.J.; Leclerc, J.; Gupta, C.; Chauhan, A. Emulsions and microemulsions for ocular drug delivery. J. Drug Deliv. Sci. Technol. 2011, 21, 111–121. [Google Scholar] [CrossRef]
  366. Tiwari, R.; Pandey, V.; Asati, S.; Soni, V.; Jain, D. Therapeutic challenges in ocular delivery of lipid based emulsion. Egypt. J. Basic Appl. Sci. 2018, 5, 121–129. [Google Scholar] [CrossRef] [Green Version]
  367. Nair, R.; Chakrapani, M.; Kaza, R. Preparation and Evaluation of Vancomycin Microemulsion for Ocular Drug Delivery. Drug Deliv. Lett. 2012, 2, 26–34. [Google Scholar]
  368. Ameta, R.K.; Soni, K.; Bhattarai, A. Recent Advances in Improving the Bioavailability of Hydrophobic/Lipophilic Drugs and Their Delivery via Self-Emulsifying Formulations. Colloids Interfaces 2023, 7, 16. [Google Scholar] [CrossRef]
  369. Li, Y.; Guan, Q.; Xu, J.; Zhang, H.; Liu, S.; Ding, Z.; Wang, Q.; Wang, Z.; Liu, M.; Zhao, Y. Comparative study of cyclosporine A liposomes and emulsions for ophthalmic drug delivery: Process optimization through response surface methodology (RSM) and biocompatibility evaluation. Colloids Surf. B Biointerfaces 2023, 225, 113267. [Google Scholar] [CrossRef]
  370. Ying, L.; Tahara, K.; Takeuchi, H. Drug delivery to the ocular posterior segment using lipid emulsion via eye drop administration: Effect of emulsion formulations and surface modification. Int. J. Pharm. 2013, 453, 329–335. [Google Scholar] [CrossRef]
  371. Liu, C.; Lan, Q.; He, W.; Nie, C.; Zhang, C.; Xu, T.; Jiang, T.; Wang, S. Octa-arginine modified lipid emulsions as a potential ocular delivery system for disulfiram: A study of the corneal permeation, transcorneal mechanism and anti-cataract effect. Colloids Surf. B Biointerfaces 2017, 160, 305–314. [Google Scholar] [CrossRef]
  372. Ahad, A.; Raish, M.; Ahmad, A.; Al-Jenoobi, F.I.; Al-Mohizea, A.M. Eprosartan mesylate loaded bilosomes as potential nano-carriers against diabetic nephropathy in streptozotocin-induced diabetic rats. Eur. J. Pharm. Sci. 2018, 111, 409–417. [Google Scholar] [CrossRef]
  373. Mondal, D.; Mandal, R.P.; De, S. Addressing the Superior Drug Delivery Performance of Bilosomes─A Microscopy and Fluorescence Study. ACS Appl. Bio. Mater. 2022, 5, 3896–3911. [Google Scholar] [CrossRef]
  374. Csaba, N.; Garcia-Fuentes, M.; Alonso, M.J. The performance of nanocarriers for transmucosal drug delivery. Expert Opin. Drug Deliv. 2006, 3, 463–478. [Google Scholar] [CrossRef]
  375. Abdelbary, A.A.; Abd-Elsalam, W.H.; Al-mahallawi, A.M. Fabrication of novel ultradeformable bilosomes for enhanced ocular delivery of terconazole: In vitro characterization, ex vivo permeation and in vivo safety assessment. Int. J. Pharm. 2016, 513, 688–696. [Google Scholar] [CrossRef]
  376. Alsaidan, O.A.; Zafar, A.; Yasir, M.; Alzarea, S.I.; Alqinyah, M.; Khalid, M. Development of Ciprofloxacin-Loaded Bilosomes In-Situ Gel for Ocular Delivery: Optimization, In-Vitro Characterization, Ex-Vivo Permeation, and Antimicrobial Study. Gels 2022, 8, 687. [Google Scholar] [CrossRef] [PubMed]
  377. Nemr, A.A.; El-Mahrouk, G.M.; Badie, H.A. Hyaluronic acid-enriched bilosomes: An approach to enhance ocular delivery of agomelatine via D-optimal design: Formulation, in vitro characterization, and in vivo pharmacodynamic evaluation in rabbits. Drug Deliv. 2022, 29, 2343–2356. [Google Scholar] [CrossRef] [PubMed]
  378. Quintanar-Guerrero, D.; Allémann, E.; Fessi, H.; Doelker, E. Preparation Techniques and Mechanisms of Formation of Biodegradable Nanoparticles from Preformed Polymers. Drug Dev. Ind. Pharm. 1998, 24, 1113–1128. [Google Scholar] [CrossRef] [PubMed]
  379. Formica, M.L.; Legeay, S.; Bejaud, J.; Montich, G.G.; Ullio Gamboa, G.V.; Benoit, J.-P.; Palma, S.D. Novel hybrid lipid nanocapsules loaded with a therapeutic monoclonal antibody–Bevacizumab–and Triamcinolone acetonide for combined therapy in neovascular ocular pathologies. Mater. Sci. Eng. C 2021, 119, 111398. [Google Scholar] [CrossRef]
  380. Eldesouky, L.M.; El-Moslemany, R.M.; Ramadan, A.A.; Morsi, M.H.; Khalafallah, N.M. Cyclosporine Lipid Nanocapsules as Thermoresponsive Gel for Dry Eye Management: Promising Corneal Mucoadhesion, Biodistribution and Preclinical Efficacy in Rabbits. Pharmaceutics 2021, 13, 360. [Google Scholar] [CrossRef]
  381. Kakkar, S.; Kaur, I.P. Spanlastics—A novel nanovesicular carrier system for ocular delivery. Int. J. Pharm. 2011, 413, 202–210. [Google Scholar] [CrossRef]
  382. Liu, Y.; Wang, Y.; Yang, J.; Zhang, H.; Gan, L. Cationized hyaluronic acid coated spanlastics for cyclosporine A ocular delivery: Prolonged ocular retention, enhanced corneal permeation and improved tear production. Int. J. Pharm. 2019, 565, 133–142. [Google Scholar] [CrossRef]
  383. Abdelbari, M.A.; El-Mancy, S.S.; Elshafeey, A.H.; Abdelbary, A.A. Implementing Spanlastics for Improving the Ocular Delivery of Clotrimazole: In vitro Characterization, Ex vivo Permeability, Microbiological Assessment and In vivo Safety Study. Int. J. Nanomed. 2021, 16, 6249–6261. [Google Scholar] [CrossRef]
  384. Ibrahim, S.S.; Abd-allah, H. Spanlastic nanovesicles for enhanced ocular delivery of vanillic acid: Design, in vitro characterization, and in vivo anti-inflammatory evaluation. Int. J. Pharm. 2022, 625, 122068. [Google Scholar] [CrossRef] [PubMed]
  385. Kumari, A.; Sharma, P.; Garg, V.; Garg, G. Ocular inserts-Advancement in therapy of eye diseases. J. Adv. Pharm. Technol. Res. 2010, 1, 291–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  386. Abdelkader, H.; Fathalla, Z.; Seyfoddin, A.; Farahani, M.; Thrimawithana, T.; Allahham, A.; Alani, A.W.G.; Al-Kinani, A.A.; Alany, R.G. Polymeric long-acting drug delivery systems (LADDS) for treatment of chronic diseases: Inserts, patches, wafers, and implants. Adv. Drug Deliv. Rev. 2021, 177, 113957. [Google Scholar] [CrossRef] [PubMed]
  387. Jadhav, C.; Yadav, K.S. Formulation and evaluation of polymer-coated bimatoprost-chitosan matrix ocular inserts for sustained lowering of IOP in rabbits. J. Drug Deliv. Sci. Technol. 2022, 77, 103885. [Google Scholar] [CrossRef]
  388. Shanmugam, S.; Valarmathi, S.; Kumar, S.S.; Shanmugasundaram, P.; Senthilkumar, M. Formulation development and evaluation of opthalmic ocusert containing aciclovir. Res. J. Pharm. Technol. 2016, 9, 1858–1862. [Google Scholar] [CrossRef]
  389. Tatke, A.; Dudhipala, N.; Janga, K.Y.; Soneta, B.; Avula, B.; Majumdar, S. Melt-Cast Films Significantly Enhance Triamcinolone Acetonide Delivery to the Deeper Ocular Tissues. Pharmaceutics 2019, 11, 158. [Google Scholar] [CrossRef] [Green Version]
  390. Rakhmetova, A. Development of a New Polymer Ocular Insert to Treat Fungal Infections Threatening the Cornea; Nazarbayev University: Astana, Kazakhstan, 2020. [Google Scholar]
  391. Bertens, C.J.F.; Martino, C.; van Osch, M.C.; Lataster, A.; Dias, A.J.A.A.; van den Biggelaar, F.J.H.M.; Tuinier, R.; Nuijts, R.M.M.A.; Gijs, M. Design of the ocular coil, a new device for non-invasive drug delivery. Eur. J. Pharm. Biopharm. 2020, 150, 120–130. [Google Scholar] [CrossRef]
  392. Taghe, S.; Mirzaeei, S.; Alany, R.G.; Nokhodchi, A. Polymeric Inserts Containing Eudragit® L100 Nanoparticle for Improved Ocular Delivery of Azithromycin. Biomedicines 2020, 8, 466. [Google Scholar] [CrossRef]
  393. Hennig, R.; Goepferich, A. Nanoparticles for the treatment of ocular neovascularizations. Eur. J. Pharm. Biopharm. 2015, 95, 294–306. [Google Scholar] [CrossRef]
  394. Lim, C.-B.; Abuzar, S.M.; Karn, P.R.; Cho, W.; Park, H.J.; Cho, C.-W.; Hwang, S.-J. Preparation, Characterization, and In Vivo Pharmacokinetic Study of the Supercritical Fluid-Processed Liposomal Amphotericin B. Pharmaceutics 2019, 11, 589. [Google Scholar] [CrossRef] [Green Version]
  395. Robinson, R.F.; Nahata, M.C. A comparative review of conventional and lipid formulations of amphotericin B. J. Clin. Pharm. Ther. 1999, 24, 249–257. [Google Scholar] [CrossRef]
  396. Barratt, G.; Bretagne, S. Optimizing efficacy of Amphotericin B through nanomodification. Int. J. Nanomed. 2007, 2, 301–313. [Google Scholar]
  397. Grandi, G.; Cavallo, R.; Zanotto, E.; Cipriani, R.; Panico, C.; Protti, R.; Scapagnini, G.; Davinelli, S.; Costagliola, C. In vitro antimicrobial activity of ozonated oil in liposome eyedrop against multidrug-resistant bacteria. Open Med. 2022, 17, 1057–1063. [Google Scholar] [CrossRef]
  398. Garrigue, J.S.; Amrane, M.; Faure, M.O.; Holopainen, J.M.; Tong, L. Relevance of Lipid-Based Products in the Management of Dry Eye Disease. J. Ocul. Pharmacol. Ther. Off. J. Assoc. Ocul. Pharmacol. Ther. 2017, 33, 647–661. [Google Scholar] [CrossRef]
  399. Nagai, N.; Otake, H. Novel drug delivery systems for the management of dry eye. Adv. Drug Deliv. Rev. 2022, 191, 114582. [Google Scholar] [CrossRef] [PubMed]
  400. Aragona, P.; Spinella, R.; Rania, L.; Postorino, E.; Roszkowska, A.; Versura, P.; Profazio, V.; Rolando, M. Assessment of the efficacy of Cationorm® in patients with moderate dry eye compared with Optive® and Emustil® eye drops. Acta Ophthalmol. 2011, 89. [Google Scholar] [CrossRef]
  401. Di Pascuale, M.A.; Goto, E.; Tseng, S.C.G. Sequential changes of lipid tear film after the instillation of a single drop of a new emulsion eye drop in dry eye patients. Ophthalmology 2004, 111, 783–791. [Google Scholar] [CrossRef] [PubMed]
  402. Halder, A.; Khopade, A.J. Pharmacokinetics and Pharmacology of Latanoprost 0.005% without Benzalkonium Chloride Vs Latanoprost 0.005% with Benzalkonium Chloride in New Zealand White Rabbits and Beagles. Curr. Eye Res. 2021, 46, 1031–1037. [Google Scholar] [CrossRef]
  403. Mazet, R.; Yaméogo, J.B.G.; Wouessidjewe, D.; Choisnard, L.; Gèze, A. Recent Advances in the Design of Topical Ophthalmic Delivery Systems in the Treatment of Ocular Surface Inflammation and Their Biopharmaceutical Evaluation. Pharmaceutics 2020, 12, 570. [Google Scholar] [CrossRef]
  404. Luchs, J. Azithromycin in DuraSite® for the treatment of blepharitis. Clin. Ophthalmol. 2010, 4, 681–688. [Google Scholar] [CrossRef] [Green Version]
  405. Tobin, K.A. Macugen treatment for wet age-related macular degeneration. Insight 2006, 31, 11–14. [Google Scholar]
  406. Couch, S.M.; Bakri, S.J. Intravitreal triamcinolone for intraocular inflammation and associated macular edema. Clin. Ophthalmol. 2009, 3, 41–47. [Google Scholar] [CrossRef] [Green Version]
  407. Han, H.; Li, S.; Xu, M.; Zhong, Y.; Fan, W.; Xu, J.; Zhou, T.; Ji, J.; Ye, J.; Yao, K. Polymer- and lipid-based nanocarriers for ocular drug delivery: Current status and future perspectives. Adv. Drug Deliv. Rev. 2023, 196, 114770. [Google Scholar] [CrossRef] [PubMed]
  408. Anderson, H.; Liu, C.K.-R.; Wakabayashi, T.; Mahmoudzadeh, R.; Salabati, M.; Spirn, M. Ocular Complications After Dexamethasone Implant Versus Intravitreal Triamcinolone in Patients with Post Vitrectomy Macular Edema. Investig. Ophthalmol. Vis. Sci. 2022, 63, 700-F0225. [Google Scholar]
  409. Hosseini, K.; Hutcheson, J.; Bowman, L. Aqueous humor concentration of Bromfenac 0.09% (Bromday) compared with Bromfenac in DuraSite 0.075% (Bromsite) in cataract patients undergoing phacoemulsification after 3 days dosing. Investig. Ophthalmol. Vis. Sci. 2013, 54, 5061. [Google Scholar]
  410. McCarey, B.E.; Cockrum, P.C. Topical Besivance, Vigamox and BSS Affects on Corneal Epithelial Wound Closure. Investig. Ophthalmol. Vis. Sci. 2010, 51, 397. [Google Scholar]
  411. Costello, M.A.; Liu, J.; Wang, Y.; Qin, B.; Xu, X.; Li, Q.; Lynd, N.A.; Zhang, F. Reverse engineering the Ozurdex dexamethasone intravitreal implant. Int. J. Pharm. 2023, 634, 122625. [Google Scholar] [CrossRef]
  412. García-Estrada, P.; García-Bon, M.A.; López-Naranjo, E.J.; Basaldúa-Pérez, D.N.; Santos, A.; Navarro-Partida, J. Polymeric Implants for the Treatment of Intraocular Eye Diseases: Trends in Biodegradable and Non-Biodegradable Materials. Pharmaceutics 2021, 13, 701. [Google Scholar] [CrossRef]
  413. Zeroni, J.; Slager, J.; Hegenrother, R.; Kloke, T.; Varner, S. Sustained Delivery of Bioactive Protein with the I-Vation Intravitreal Implant. Investig. Ophthalmol. Vis. Sci. 2007, 48, 5805. [Google Scholar]
  414. Gaballa, S.A.; Kompella, U.B.; Elgarhy, O.; Alqahtani, A.M.; Pierscionek, B.; Alany, R.G.; Abdelkader, H. Corticosteroids in ophthalmology: Drug delivery innovations, pharmacology, clinical applications, and future perspectives. Drug Deliv. Transl. Res. 2021, 11, 866–893. [Google Scholar] [CrossRef]
  415. Yasukawa, T.; Ogura, Y. Medical Devices for the Treatment of Eye Diseases. In Drug Delivery; Schäfer-Korting, M., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 469–489. [Google Scholar] [CrossRef]
  416. Kuppermann, B.D.; Blumenkranz, M.S.; Haller, J.A.; Williams, G.A.; Weinberg, D.V.; Chou, C.; Whitcup, S.M.; Group, D.D.P.I.S. Randomized Controlled Study of an Intravitreous Dexamethasone Drug Delivery System in Patients with Persistent Macular Edema. Arch. Ophthalmol. 2007, 125, 309–317. [Google Scholar] [CrossRef] [PubMed]
  417. Seah, S.K.L.; Husain, R.; Gazzard, G.; Lim, M.C.C.; Hoh, S.-T.; Oen, F.T.S.; Aung, T. Use of Surodex in Phacotrabeculectomy Surgery. Am. J. Ophthalmol. 2005, 139, 927–928. [Google Scholar] [CrossRef] [PubMed]
  418. Wróblewska, K.B.; Jadach, B.; Muszalska-Kolos, I. Progress in drug formulation design and delivery of medicinal substances used in ophthalmology. Int. J. Pharm. 2021, 607, 121012. [Google Scholar] [CrossRef]
  419. Allyn, M.M.; Luo, R.H.; Hellwarth, E.B.; Swindle-Reilly, K.E. Considerations for Polymers Used in Ocular Drug Delivery. Front. Med. 2022, 8, 2963. [Google Scholar] [CrossRef]
  420. Nyamweya, N.N. Applications of polymer blends in drug delivery. Future J. Pharm. Sci. 2021, 7, 18. [Google Scholar] [CrossRef]
  421. Cao, Y.; Samy, K.E.; Bernards, D.A.; Desai, T.A. Recent advances in intraocular sustained-release drug delivery devices. Drug Discov. Today 2019, 24, 1694–1700. [Google Scholar] [CrossRef]
  422. Fayzullin, A.; Bakulina, A.; Mikaelyan, K.; Shekhter, A.; Guller, A. Implantable Drug Delivery Systems and Foreign Body Reaction: Traversing the Current Clinical Landscape. Bioengineering 2021, 8, 205. [Google Scholar] [CrossRef]
  423. Saettone, M.F.; Salminen, L. Ocular inserts for topical delivery. Adv. Drug Deliv. Rev. 1995, 16, 95–106. [Google Scholar] [CrossRef]
  424. Banker, G.S.; Siepmann, J.; Rhodes, C. Design and evaluation of ophthalmic pharmaceutical products. In Modern pharmaceutics; CRC Press: Boca Raton, FL, USA, 2002; pp. 647–738. [Google Scholar]
  425. Brandt, J.D.; DuBiner, H.B.; Benza, R.; Sall, K.N.; Walker, G.A.; Semba, C.P.; Budenz, D.; Day, D.; Flowers, B.; Lee, S.; et al. Long-term Safety and Efficacy of a Sustained-Release Bimatoprost Ocular Ring. Ophthalmology 2017, 124, 1565–1566. [Google Scholar] [CrossRef] [Green Version]
  426. Belamkar, A.; Harris, A.; Zukerman, R.; Siesky, B.; Oddone, F.; Verticchio Vercellin, A.; Ciulla, T.A. Sustained release glaucoma therapies: Novel modalities for overcoming key treatment barriers associated with topical medications. Ann. Med. 2022, 54, 343–358. [Google Scholar] [CrossRef] [PubMed]
  427. Cagini, C.; Caricato, A.; Tosi, G.; Pascale, A.; Cesari, C.; Fiore, T. Evaluation of the efficacy and safety of the ophthalmic insert Mydriasert in patients undergoing retinal angiography. Eur. J. Ophthalmol. 2014, 24, 728–734. [Google Scholar] [CrossRef] [PubMed]
  428. Kansara, V.S.; Hancock, S.E.; Muya, L.W.; Ciulla, T.A. Suprachoroidal delivery enables targeting, localization and durability of small molecule suspensions. J. Control. Release 2022, 349, 1045–1051. [Google Scholar] [CrossRef] [PubMed]
  429. Lee, K.; Park, S.H.; Lee, J.Y.; Ryu, W.H. Commercialized Microneedles. In Microneedling in Clinical Practice; CRC Press: Boca Raton, FL, USA, 2020; pp. 91–108. [Google Scholar] [CrossRef]
  430. Novack, G.D. US Regulatory Approval of a Drug-Eluting Contact Lens. Eye Contact Lens Sci. Clin. Pract. 2023, 49, 136–138. [Google Scholar] [CrossRef] [PubMed]
  431. Miller, R.A.; Brady, J.M.; Cutright, D.E. Degradation rates of oral resorbable implants (polylactates and polyglycolates): Rate modification with changes in PLA/PGA copolymer ratios. J. Biomed. Mater. Res. 1977, 11, 711–719. [Google Scholar] [CrossRef] [PubMed]
  432. Haghjou, N.; Soheilian, M.; Abdekhodaie, M.J. Sustained release intraocular drug delivery devices for treatment of uveitis. J. Ophthalmic Vis. Res. 2011, 6, 317–329. [Google Scholar] [PubMed]
  433. Desai, A.R.; Maulvi, F.A.; Desai, D.M.; Shukla, M.R.; Ranch, K.M.; Vyas, B.A.; Shah, S.A.; Sandeman, S.; Shah, D.O. Multiple drug delivery from the drug-implants-laden silicone contact lens: Addressing the issue of burst drug release. Mater. Sci. Eng. C 2020, 112, 110885. [Google Scholar] [CrossRef]
  434. Maulvi, F.A.; Desai, D.T.; Shetty, K.H.; Shah, D.O.; Willcox, M.D.P. Advances and challenges in the nanoparticles-laden contact lenses for ocular drug delivery. Int. J. Pharm. 2021, 608, 121090. [Google Scholar] [CrossRef]
  435. Maulvi, F.A.; Soni, T.G.; Shah, D.O. A review on therapeutic contact lenses for ocular drug delivery. Drug Deliv. 2016, 23, 3017–3026. [Google Scholar] [CrossRef]
  436. Kudryavtseva, V.; Otero, M.; Zhang, J.; Bukatin, A.; Gould, D.; Sukhorukov, G.B. Drug-Eluting Sandwich Hydrogel Lenses Based on Microchamber Film Drug Encapsulation. ACS Nanosci. Au 2023. [Google Scholar] [CrossRef]
  437. Bin Sahadan, M.Y.; Tong, W.Y.; Tan, W.N.; Leong, C.R.; Bin Misri, M.N.; Chan, M.; Cheng, S.Y.; Shaharuddin, S. Phomopsidione nanoparticles coated contact lenses reduce microbial keratitis causing pathogens. Exp. Eye Res. 2019, 178, 10–14. [Google Scholar] [CrossRef]
  438. Horne, R.R.; Rich, J.T.; Bradley, M.W.; Pitt, W.G. Latanoprost uptake and release from commercial contact lenses. J. Biomater. Sci. Polym. Ed. 2020, 31, 1–19. [Google Scholar] [CrossRef] [PubMed]
  439. Torres-Luna, C.; Hu, N.; Tammareddy, T.; Domszy, R.; Yang, J.; Wang, N.S.; Yang, A. Extended delivery of non-steroidal anti-inflammatory drugs through contact lenses loaded with Vitamin E and cationic surfactants. Contact Lens Anterior Eye 2019, 42, 546–552. [Google Scholar] [CrossRef] [PubMed]
  440. Andreadis, I.I.; Karavasili, C.; Thomas, A.; Komnenou, A.; Tzimtzimis, M.; Tzetzis, D.; Andreadis, D.; Bouropoulos, N.; Fatouros, D.G. In Situ Gelling Electrospun Ocular Films Sustain the Intraocular Pressure-Lowering Effect of Timolol Maleate: In Vitro, Ex Vivo, and Pharmacodynamic Assessment. Mol. Pharm. 2022, 19, 274–286. [Google Scholar] [CrossRef] [PubMed]
  441. I Tartara, L.; D Palma, S.; Allemandi, D.; I Ahumada, M.; M Llabot, J. New Mucoadhesive Polymeric Film for Ophthalmic Administration of Acetazolamide. Recent Pat. Drug Deliv. Formul. 2014, 8, 224–232. [Google Scholar] [CrossRef]
  442. Sanap, S.N.; Bisen, A.C.; Kedar, A.; Yadav, K.S.; Krishna, A.; Akhir, A.; Chopra, S.; Mugale, M.N.; Bhatta, R.S. Chitosan/HPMC-based mucoadhesive film co-loaded with fluconazole and ofloxacin for management of polymicrobial keratitis. Int. J. Biol. Macromol. 2022, 222, 2785–2795. [Google Scholar] [CrossRef]
  443. Tandale, Y.N.; Wagh, V.D.J.I.J.P.R. Formulation and evaluation of dorzolamide hydrochloride polymeric film. Int. J. PharmTech Res. 2011, 3, 1817–1824. [Google Scholar]
  444. Gupta, P.; Yadav, K.S. Applications of microneedles in delivering drugs for various ocular diseases. Life Sci. 2019, 237, 116907. [Google Scholar] [CrossRef]
  445. Tariq, N.; Ashraf, M.W.; Tayyaba, S. A Review on Solid Microneedles for Biomedical Applications. J. Pharm. Innov. 2022, 17, 1464–1483. [Google Scholar] [CrossRef]
  446. Patel, S.R.; Edelhauser, H.F.; Prausnitz, M.R. Targeted Drug Delivery to the Eye Enabled by Microneedles. In Drug Product Development for the Back of the Eye; Kompella, U.B., Edelhauser, H.F., Eds.; Springer: Boston, MA, USA, 2011; pp. 331–360. [Google Scholar] [CrossRef]
  447. Jiang, J.; Gill, H.S.; Ghate, D.; McCarey, B.E.; Patel, S.R.; Edelhauser, H.F.; Prausnitz, M.R. Coated Microneedles for Drug Delivery to the Eye. Investig. Ophthalmol. Vis. Sci. 2007, 48, 4038–4043. [Google Scholar] [CrossRef] [Green Version]
  448. Kim, Y.C.; Grossniklaus, H.E.; Edelhauser, H.F.; Prausnitz, M.R. Intrastromal Delivery of Bevacizumab Using Microneedles to Treat Corneal Neovascularization. Investig. Ophthalmol. Vis. Sci. 2014, 55, 7376–7386. [Google Scholar] [CrossRef]
  449. Cárcamo-Martínez, Á.; Mallon, B.; Domínguez-Robles, J.; Vora, L.K.; Anjani, Q.K.; Donnelly, R.F. Hollow microneedles: A perspective in biomedical applications. Int. J. Pharm. 2021, 599, 120455. [Google Scholar] [CrossRef] [PubMed]
  450. Kim, Y.-C.; Park, J.-H.; Prausnitz, M.R. Microneedles for drug and vaccine delivery. Adv. Drug Deliv. Rev. 2012, 64, 1547–1568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  451. Gilger, B.C.; Abarca, E.M.; Salmon, J.H.; Patel, S. Treatment of Acute Posterior Uveitis in a Porcine Model by Injection of Triamcinolone Acetonide Into the Suprachoroidal Space Using Microneedles. Investig. Ophthalmol. Vis. Sci. 2013, 54, 2483–2492. [Google Scholar] [CrossRef] [Green Version]
  452. Thakur, R.R.S.; Tekko, I.A.; Al-Shammari, F.; Ali, A.A.; McCarthy, H.; Donnelly, R.F. Rapidly dissolving polymeric microneedles for minimally invasive intraocular drug delivery. Drug Deliv. Transl. Res. 2016, 6, 800–815. [Google Scholar] [CrossRef] [Green Version]
  453. Ali, M.; Namjoshi, S.; Benson, H.A.E.; Mohammed, Y.; Kumeria, T. Dissolvable polymer microneedles for drug delivery and diagnostics. J. Control. Release 2022, 347, 561–589. [Google Scholar] [CrossRef]
  454. Albadr, A.A.; Tekko, I.A.; Vora, L.K.; Ali, A.A.; Laverty, G.; Donnelly, R.F.; Thakur, R.R.S. Rapidly dissolving microneedle patch of amphotericin B for intracorneal fungal infections. Drug Deliv. Transl. Res. 2022, 12, 931–943. [Google Scholar] [CrossRef]
  455. Datta, D.; Roy, G.; Garg, P.; Venuganti, V.V.K. Ocular delivery of cyclosporine A using dissolvable microneedle contact lens. J. Drug Deliv. Sci. Technol. 2022, 70, 103211. [Google Scholar] [CrossRef]
  456. Isreb, A.; Baj, K.; Wojsz, M.; Isreb, M.; Peak, M.; Alhnan, M.A. 3D printed oral theophylline doses with innovative ‘radiator-like’ design: Impact of polyethylene oxide (PEO) molecular weight. Int. J. Pharm. 2019, 564, 98–105. [Google Scholar] [CrossRef]
  457. Pereira, B.C.; Isreb, A.; Forbes, R.T.; Dores, F.; Habashy, R.; Petit, J.-B.; Alhnan, M.A.; Oga, E.F. ‘Temporary Plasticiser’: A novel solution to fabricate 3D printed patient-centred cardiovascular ‘Polypill’ architectures. Eur. J. Pharm. Biopharm. 2019, 135, 94–103. [Google Scholar] [CrossRef] [PubMed]
  458. Tan, Y.J.N.; Yong, W.P.; Kochhar, J.S.; Khanolkar, J.; Yao, X.; Sun, Y.; Ao, C.K.; Soh, S. On-demand fully customizable drug tablets via 3D printing technology for personalized medicine. J. Control. Release 2020, 322, 42–52. [Google Scholar] [CrossRef] [PubMed]
  459. Andreadis, I.I.; Gioumouxouzis, C.I.; Eleftheriadis, G.K.; Fatouros, D.G. The Advent of a New Era in Digital Healthcare: A Role for 3D Printing Technologies in Drug Manufacturing? Pharmaceutics 2022, 14, 609. [Google Scholar] [CrossRef] [PubMed]
  460. Larochelle, R.D.; Mann, S.E.; Ifantides, C. 3D Printing in Eye Care. Ophthalmol. Ther. 2021, 10, 733–752. [Google Scholar] [CrossRef] [PubMed]
  461. Kim, B.R.; Kim, S.H.; Ko, J.; Baek, S.W.; Park, Y.K.; Kim, Y.J.; Yoon, J.S. A Pilot Clinical Study of Ocular Prosthesis Fabricated by Three-dimensional Printing and Sublimation Technique. Korean J. Ophthalmol. 2021, 35, 37–43. [Google Scholar] [CrossRef] [PubMed]
  462. Ulag, S.; Ilhan, E.; Sahin, A.; Karademir Yilmaz, B.; Kalaskar, D.M.; Ekren, N.; Kilic, O.; Nuzhet Oktar, F.; Gunduz, O. 3D printed artificial cornea for corneal stromal transplantation. Eur. Polym. J. 2020, 133, 109744. [Google Scholar] [CrossRef]
  463. Holland, G.; Pandit, A.; Sánchez-Abella, L.; Haiek, A.; Loinaz, I.; Dupin, D.; Gonzalez, M.; Larra, E.; Bidaguren, A.; Lagali, N.; et al. Artificial Cornea: Past, Current, and Future Directions. Front. Med. 2021, 8, 770780. [Google Scholar] [CrossRef]
  464. Garg, U.; Jain, N.; Kaul, S.; Rai, V.K.; Pandey, M.; Nagaich, U.; Dua, K. The emerging role of 3D-printing in ocular drug delivery: Challenges, current status, and future prospects. J. Drug Deliv. Sci. Technol. 2022, 76, 103798. [Google Scholar] [CrossRef]
  465. Shi, P.; Tan, E.Y.S.; Yeong, W.Y.; Laude, A. Hybrid three-dimensional (3D) bioprinting of retina equivalent for ocular research. Int. J. Bioprint. 2017, 3, 9. [Google Scholar] [CrossRef] [Green Version]
  466. Thompson, J.R.; Worthington, K.S.; Green, B.J.; Mullin, N.K.; Jiao, C.; Kaalberg, E.E.; Wiley, L.A.; Han, I.C.; Russell, S.R.; Sohn, E.H.; et al. Two-photon polymerized poly(caprolactone) retinal cell delivery scaffolds and their systemic and retinal biocompatibility. Acta Biomater. 2019, 94, 204–218. [Google Scholar] [CrossRef]
  467. Tagami, T.; Goto, E.; Kida, R.; Hirose, K.; Noda, T.; Ozeki, T. Lyophilized ophthalmologic patches as novel corneal drug formulations using a semi-solid extrusion 3D printer. Int. J. Pharm. 2022, 617, 121448. [Google Scholar] [CrossRef]
  468. Mohamdeen, Y.M.G.; Tabriz, A.G.; Tighsazzadeh, M.; Nandi, U.; Khalaj, R.; Andreadis, I.; Boateng, J.S.; Douroumis, D. Development of 3D printed drug-eluting contact lenses. J. Pharm. Pharmacol. 2021, 74, 1467–1476. [Google Scholar] [CrossRef]
  469. Yang, S.-A.; Mitchell, W.; Hall, N.; Elze, T.; Lorch, A.C.; Miller, J.W.; Zebardast, N. Trends and Usage Patterns of Minimally Invasive Glaucoma Surgery in the United States: IRIS® Registry Analysis 2013–2018. Ophthalmol. Glaucoma 2021, 4, 558–568. [Google Scholar] [CrossRef]
  470. Culmone, C.; Lussenburg, K.; Alkemade, J.; Smit, G.; Sakes, A.; Breedveld, P. A Fully 3D-Printed Steerable Instrument for Minimally Invasive Surgery. Materials 2021, 14, 7910. [Google Scholar] [CrossRef]
  471. Thakur, R.R.S.; McMillan, H.L.; Jones, D.S. Solvent induced phase inversion-based in situ forming controlled release drug delivery implants. J. Control. Release 2014, 176, 8–23. [Google Scholar] [CrossRef]
  472. Kempe, S.; Mäder, K. In situ forming implants—An attractive formulation principle for parenteral depot formulations. J. Control. Release 2012, 161, 668–679. [Google Scholar] [CrossRef]
  473. Sheshala, R.; Hong, G.C.; Yee, W.P.; Meka, V.S.; Thakur, R.R.S. In situ forming phase-inversion implants for sustained ocular delivery of triamcinolone acetonide. Drug Deliv. Transl. Res. 2019, 9, 534–542. [Google Scholar] [CrossRef] [Green Version]
  474. Bisht, R.; Jaiswal, J.K.; Rupenthal, I.D. Nanoparticle-loaded biodegradable light-responsive in situ forming injectable implants for effective peptide delivery to the posterior segment of the eye. Med. Hypotheses 2017, 103, 5–9. [Google Scholar] [CrossRef] [PubMed]
  475. Abdelmohsen, H.A.M.; Copeland, N.A.; Hardy, J.G. Light-responsive biomaterials for ocular drug delivery. Drug Deliv. Transl. Res. 2022, 1–24. [Google Scholar] [CrossRef]
  476. Hsu, X.-L.; Wu, L.-C.; Hsieh, J.-Y.; Huang, Y.-Y. Nanoparticle-Hydrogel Composite Drug Delivery System for Potential Ocular Applications. Polymers 2021, 13, 642. [Google Scholar] [CrossRef]
  477. Danckwerts, M.; Fassihi, A. Implantable Controlled Release Drug Delivery Systems: A Review. Drug Dev. Ind. Pharm. 1991, 17, 1465–1502. [Google Scholar] [CrossRef]
  478. Baino, F.; Kargozar, S. Regulation of the Ocular Cell/Tissue Response by Implantable Biomaterials and Drug Delivery Systems. Bioengineering 2020, 7, 65. [Google Scholar] [CrossRef] [PubMed]
  479. Smith, T.J.; Pearson, P.A.; Blandford, D.L.; Brown, J.D.; Goins, K.A.; Hollins, J.L.; Schmeisser, E.T.; Glavinos, P.; Baldwin, L.B.; Ashton, P. Intravitreal Sustained-Release Ganciclovir. Arch. Ophthalmol. 1992, 110, 255–258. [Google Scholar] [CrossRef] [PubMed]
  480. Prata, A.I.; Coimbra, P.; Pina, M.E. Preparation of dexamethasone ophthalmic implants: A comparative study of in vitro release profiles. Pharm. Dev. Technol. 2018, 23, 218–224. [Google Scholar] [CrossRef] [PubMed]
  481. Solano, A.G.R.; de Fátima Pereira, A.; de Faria, L.G.A.; Fialho, S.L.; de Oliveira Patricio, P.S.; da Silva-Cunha, A.; Fulgêncio, G.O.; da Silva, G.R.; Pianetti, G.A. Etoposide-Loaded Poly(Lactic-co-Glycolic Acid) Intravitreal Implants: In Vitro and In Vivo Evaluation. AAPS PharmSciTech 2018, 19, 1652–1661. [Google Scholar] [CrossRef] [PubMed]
  482. Natu, M.V.; Gaspar, M.N.; Fontes Ribeiro, C.A.; Cabrita, A.M.; de Sousa, H.C.; Gil, M.H. In vitro and in vivo evaluation of an intraocular implant for glaucoma treatment. Int. J. Pharm. 2011, 415, 73–82. [Google Scholar] [CrossRef] [Green Version]
  483. Stanković, M.; Frijlink, H.W.; Hinrichs, W.L.J. Polymeric formulations for drug release prepared by hot melt extrusion: Application and characterization. Drug Discov. Today 2015, 20, 812–823. [Google Scholar] [CrossRef]
  484. Breitenbach, J. Melt extrusion: From process to drug delivery technology. Eur. J. Pharm. Biopharm. 2002, 54, 107–117. [Google Scholar] [CrossRef]
  485. Bhattarai, R.S.; Das, A.; Alzhrani, R.M.; Kang, D.; Bhaduri, S.B.; Boddu, S.H.S. Comparison of electrospun and solvent cast polylactic acid (PLA)/poly(vinyl alcohol) (PVA) inserts as potential ocular drug delivery vehicles. Mater. Sci. Eng. C 2017, 77, 895–903. [Google Scholar] [CrossRef]
  486. Thakkar, S.; Misra, M. Electrospun polymeric nanofibers: New horizons in drug delivery. Eur. J. Pharm. Sci. 2017, 107, 148–167. [Google Scholar] [CrossRef]
  487. Sommer, A.C.; Blumenthal, E.Z. Implementations of 3D printing in ophthalmology. Graefe’s Arch. Clin. Exp. Ophthalmol. 2019, 257, 1815–1822. [Google Scholar] [CrossRef]
  488. Reddy Dumpa, N.; Bandari, S.A.; Repka, M. Novel Gastroretentive Floating Pulsatile Drug Delivery System Produced via Hot-Melt Extrusion and Fused Deposition Modeling 3D Printing. Pharmaceutics 2020, 12, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  489. Korelidou, A.; Domínguez-Robles, J.; Magill, E.R.; Eleftheriadou, M.; Cornelius, V.A.; Donnelly, R.F.; Margariti, A.; Larrañeta, E. 3D-printed reservoir-type implants containing poly(lactic acid)/poly(caprolactone) porous membranes for sustained drug delivery. Biomater. Adv. 2022, 139, 213024. [Google Scholar] [CrossRef] [PubMed]
  490. Muwaffak, Z.; Goyanes, A.; Clark, V.; Basit, A.W.; Hilton, S.T.; Gaisford, S. Patient-specific 3D scanned and 3D printed antimicrobial polycaprolactone wound dressings. Int. J. Pharm. 2017, 527, 161–170. [Google Scholar] [CrossRef]
  491. Annuryanti, F.; Domínguez-Robles, J.; Anjani, Q.K.; Adrianto, M.F.; Larrañeta, E.; Thakur, R.R.S. Fabrication and Characterisation of 3D-Printed Triamcinolone Acetonide-Loaded Polycaprolactone-Based Ocular Implants. Pharmaceutics 2023, 15, 243. [Google Scholar] [CrossRef] [PubMed]
  492. Alzahrani, A.; Youssef, A.A.A.; Nyavanandi, D.; Tripathi, S.; Bandari, S.; Majumdar, S.; Repka, M.A. Design and optimization of ciprofloxacin hydrochloride biodegradable 3D printed ocular inserts: Full factorial design and in-vitro and ex-vivo evaluations: Part II. Int. J. Pharm. 2023, 631, 122533. [Google Scholar] [CrossRef]
  493. Atkinson, H.V.; Davies, S. Fundamental aspects of hot isostatic pressing: An overview. Metall. Mater. Trans. A 2000, 31, 2981–3000. [Google Scholar] [CrossRef]
  494. Adelmann, B.; Hellmann, R. Mechanical Properties of LPBF-Built Titanium Lattice Structures—A Comparative Study of As-Built and Hot Isostatic Pressed Structures for Medical Implants. Metals 2022, 12, 2072. [Google Scholar]
  495. Hassanin, H.; Al-Kinani, A.A.; ElShaer, A.; Polycarpou, E.; El-Sayed, M.A.; Essa, K.J.J.O.M.C.B. Stainless steel with tailored porosity using canister-free hot isostatic pressing for improved osseointegration implants. J. Mater. Chem. B 2017, 5, 9384–9394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  496. Mullen, L.; Stamp, R.C.; Fox, P.; Jones, E.; Ngo, C.; Sutcliffe, C.J. Selective laser melting: A unit cell approach for the manufacture of porous, titanium, bone in-growth constructs, suitable for orthopedic applications. II. Randomized structures. J. Biomed. Mater. Res. 2010, 92B, 178–188. [Google Scholar] [CrossRef]
  497. Abdelkader, H.; Wertheim, D.; Pierscionek, B.; Alany, R.G. Curcumin In Situ Gelling Polymeric Insert with Enhanced Ocular Performance. Pharmaceutics 2020, 12, 1158. [Google Scholar] [CrossRef]
  498. Bode, C.; Kranz, H.; Siepmann, F.; Siepmann, J. In-situ forming PLGA implants for intraocular dexamethasone delivery. Int. J. Pharm. 2018, 548, 337–348. [Google Scholar] [CrossRef]
  499. Allababidi, S.; Shah, J.C. Kinetics and mechanism of release from glyceryl monostearate-based implants: Evaluation of release in a gel simulating in vivo implantation. J. Pharm. Sci. 1998, 87, 738–744. [Google Scholar] [CrossRef]
  500. Balasubramaniam, J.; Srinatha, A.; Pandit, J.K. Studies on Indomethacin Intraocular Implants Using Different in vitro Release Methods. Indian J. Pharm. Sci. 2008, 70, 216–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  501. Chennamaneni, S.R.; Mamalis, C.; Archer, B.; Oakey, Z.; Ambati, B.K. Development of a novel bioerodible dexamethasone implant for uveitis and postoperative cataract inflammation. J. Control. Release 2013, 167, 53–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  502. Awwad, S.; Lockwood, A.; Brocchini, S.; Khaw, P.T. The PK-Eye: A Novel In Vitro Ocular Flow Model for Use in Preclinical Drug Development. J. Pharm. Sci. 2015, 104, 3330–3342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  503. Stein, S.; Auel, T.; Kempin, W.; Bogdahn, M.; Weitschies, W.; Seidlitz, A. Influence of the test method on in vitro drug release from intravitreal model implants containing dexamethasone or fluorescein sodium in poly (d,l-lactide-co-glycolide) or polycaprolactone. Eur. J. Pharm. Biopharm. 2018, 127, 270–278. [Google Scholar] [CrossRef] [PubMed]
  504. Verhoeven, R.S.; Garcia, A.; Robeson, R.; Gilger, B.C.; Culp, D.; Struble, C.; Hamm, L.; Navratil, T.; Yerxa, B. Nonclinical Development of ENV905 (Difluprednate) Ophthalmic Implant for the Treatment of Inflammation and Pain Associated with Ocular Surgery. J. Ocul. Pharmacol. Ther. 2017, 34, 161–169. [Google Scholar] [CrossRef]
  505. Sirinek, P.E.; Lin, M.M. Intracameral sustained release bimatoprost implants (Durysta). Semin. Ophthalmol. 2022, 37, 385–390. [Google Scholar] [CrossRef]
  506. Salama, H.A.; Ghorab, M.; Mahmoud, A.A.; Abdel Hady, M. PLGA Nanoparticles as Subconjunctival Injection for Management of Glaucoma. AAPS PharmSciTech 2017, 18, 2517–2528. [Google Scholar] [CrossRef] [PubMed]
  507. FDA. Regulatory Controls. Available online: https://www.fda.gov/medical-devices/overview-device-regulation/regulatory-controls (accessed on 28 April 2023).
  508. Premarket Notification 510(k)|FDA. Available online: https://www.fda.gov/medical-devices/premarket-submissions-selecting-and-preparing-correct-submission/premarket-notification-510k (accessed on 28 April 2023).
  509. Premarket Approval (PMA)|FDA. Available online: https://www.fda.gov/medical-devices/premarket-submissions-selecting-and-preparing-correct-submission/premarket-approval-pma (accessed on 29 April 2023).
  510. Design Control Guidance for Medical Device Manufacturers|FDA. Available online: https://www.fda.gov/regulatory-information/search-fda-guidance-documents/design-control-guidance-medical-device-manufacturers (accessed on 29 April 2023).
  511. Use of International Standard ISO 10993-1, “Biological Evaluation of Medical Devices-Part 1: Evaluation and Testing within a Risk Management Process”|FDA. Available online: https://www.fda.gov/regulatory-information/search-fda-guidance-documents/use-international-standard-iso-10993-1-biological-evaluation-medical-devices-part-1-evaluation-and (accessed on 29 April 2023).
  512. Guidance for the Preparation of a Premarket Notification Application for a Surgical Mesh-Guidance for Industry and/or for FDA Reviewers/Staff and/or Compliance|Guidance Portal. Available online: https://www.hhs.gov/guidance/document/guidance-preparation-premarket-notification-application-surgical-mesh-guidance-industry (accessed on 29 April 2023).
Figure 1. Major pathophysiological causes for the development of dry eye diseases.
Figure 1. Major pathophysiological causes for the development of dry eye diseases.
Pharmaceutics 15 01746 g001
Figure 2. Schematic diagram for the aqueous humor drainage mechanisms in open-angle and angle-closure glaucoma. Intraocular pressure elevation may be a result of (a) open-angle or (b) angle-closure glaucoma. Open-angle glaucoma occurs as a result of pathological changes in the outflow tissues, which increase the resistance to aqueous humor outflow. Angle-closure glaucoma prevents aqueous humor outflow due to blocking of iridocorneal angle.
Figure 2. Schematic diagram for the aqueous humor drainage mechanisms in open-angle and angle-closure glaucoma. Intraocular pressure elevation may be a result of (a) open-angle or (b) angle-closure glaucoma. Open-angle glaucoma occurs as a result of pathological changes in the outflow tissues, which increase the resistance to aqueous humor outflow. Angle-closure glaucoma prevents aqueous humor outflow due to blocking of iridocorneal angle.
Pharmaceutics 15 01746 g002
Figure 3. Schematic diagram for the process of choroidal neovascularization along with drugs, supplements, and biologics used for treatment of AMD.
Figure 3. Schematic diagram for the process of choroidal neovascularization along with drugs, supplements, and biologics used for treatment of AMD.
Pharmaceutics 15 01746 g003
Figure 4. Schematic diagram for the progressive stages for diabetic retinopathy showing the early and late stages, along with drugs, supplements, and biologics used for treating diabetic retinopathy.
Figure 4. Schematic diagram for the progressive stages for diabetic retinopathy showing the early and late stages, along with drugs, supplements, and biologics used for treating diabetic retinopathy.
Pharmaceutics 15 01746 g004
Figure 5. Nanocarrier systems investigated for ophthalmic uses.
Figure 5. Nanocarrier systems investigated for ophthalmic uses.
Pharmaceutics 15 01746 g005
Figure 6. Different types of microneedles frequently used for ocular drug delivery. Ocularly applied microneedles can be classified into (a) solid-coated, (b) hollow, and (c) polymeric microneedles. (a) Solid coated microneedles are used to puncture the ocular tissue and allow the coated medication to disperse. (b) Hollow microneedles are tiny needles that completely contain the medication formulation. (c) Polymeric microneedles are constructed of a variety of polymers that are biocompatible and biodegradable and are easy to install into the ocular tissue, where they then dissolve upon ocular application and initiate drug release.
Figure 6. Different types of microneedles frequently used for ocular drug delivery. Ocularly applied microneedles can be classified into (a) solid-coated, (b) hollow, and (c) polymeric microneedles. (a) Solid coated microneedles are used to puncture the ocular tissue and allow the coated medication to disperse. (b) Hollow microneedles are tiny needles that completely contain the medication formulation. (c) Polymeric microneedles are constructed of a variety of polymers that are biocompatible and biodegradable and are easy to install into the ocular tissue, where they then dissolve upon ocular application and initiate drug release.
Pharmaceutics 15 01746 g006
Figure 7. Different techniques/technologies used to generate implantable drug delivery systems. Implantable drug delivery systems might be generated via the use of (a) solvent casting, (b) melt extrusion, (c) electrospinning, and (d) other techniques.
Figure 7. Different techniques/technologies used to generate implantable drug delivery systems. Implantable drug delivery systems might be generated via the use of (a) solvent casting, (b) melt extrusion, (c) electrospinning, and (d) other techniques.
Pharmaceutics 15 01746 g007
Figure 8. In vitro testing of implantable drug delivery systems. Implantable drug delivery systems might be characterized for drug dissolution and release using (a) static diffusion, (b) agar diffusion, (c) dialysis bags, and (d) other models.
Figure 8. In vitro testing of implantable drug delivery systems. Implantable drug delivery systems might be characterized for drug dissolution and release using (a) static diffusion, (b) agar diffusion, (c) dialysis bags, and (d) other models.
Pharmaceutics 15 01746 g008
Table 2. Commercial ocular drug delivery systems/devices for different chronic eye conditions.
Table 2. Commercial ocular drug delivery systems/devices for different chronic eye conditions.
Platform/DeviceCommercial BrandTherapeutic Agent (Approval Year, Country)Excipient CompositionClinical ImplicationRoute of AdministrationRefs.
LiposomesVisudyne®Verteporfin (2000, USA)Dimyristoylphosphatidylcholine and egg yolk phosphatidylglycerolChoroidal neovascularization in AMDIV[393]
Amphotec®Amphotericin B (1996, USA)Cholesteryl sulfateFungal endophthalmitisIV[394]
Abelcet®Amphotericin B (1995, USA)Dimyristoylphosphatidylcholine and dimyristoylphosphatidylglycerolFungal endophthalmitisIV[395]
AmBisome®Amphotericin B (1997, USA)Hydrogenated soy phosphatidylcholine, cholesterol, and distearoylphosphatidylglycerolFungal endophthalmitisIV[396]
Ozodrop®Sunflower ozonized oil (NA, Italy)LipozonEye, hypromellose and polyhexamethylene biguanidePost-cataract surgery inflammationTopical[397]
Lacrisek®Vitamin A palmitate, vitamin E (NA, Italy)Hydrogenated phospholipidsDry eye syndromeTopical[398]
Tears Again®Vitamin A palmitate, vitamin E (NA, USA)Soy lecithin, phenoxyethanolDry eye syndromeTopical[398]
EmulsionsRestasis®Cyclosporine A (2002, USA)Polysorbate 80, castor oilDry eye syndromeTopical[399]
Emustil®Cyclosporine A (NA, Italy)Soybean oil, egg yolk phospholipidsDry eye syndromeTopical[400]
Refresh Endura®Cyclosporine A (2020, USA)Polysorbate 80, castor oilDry eye syndromeTopical[401]
Xelpros®Latanoprost (2018, USA) Castor oil, propylene glycolGlaucomaTopical[402]
Durezol®Difluprednate (2008, USA)Castor oil, polysorbate 80Diabetic macular edemaTopical[85]
Verkazia®Cyclosporine A (2017, UK)Medium-chain triglyceride, Tyloxapol, and poloxamer 188Vernal keratoconjunctivitisTopical[403]
MicellesCequa®Cyclosporin A (2018, USA)Polyoxyl hydrogenated castor oil, polyalkoxylated alcoholDry eye syndromeTopical[54]
AzaSite®Azithromycin (2007, USA)PolycarbophilBlepharitisTopical[404]
Micro and nanoparticlesMacugen®Pegaptanib (2004, USA)PEG-40kDaAMDIVT[405]
Trivaris®Triamcinolone acetonide (2005, USA)Sodium hyaluronate UveitisIVT[406]
Inveltys®Loteprednol etabonate (2018, USA)Pluronic F127Post-operative inflammationsTopical[407]
Eysuvis®Loteprednol etabonate (2020, USA)Pluronic F127Dry eye syndromeTopical[407]
Triesence®Triamcinolone acetonide (2005, USA)Carboxymethyl cellulose Dry eye syndromeIVT[408]
Tobradex ST®Tobramycin and dexamethasone (2003, USA)Xanthan gumBacterial conjunctivitisTopical[403]
BromSite®Bromfenac (2016, USA)PolycarbophilPost-operative inflammation and pain reliverTopical[409]
Besivance®Besifloxacin (2009, USA)PolycarbophilBacterial conjunctivitisTopical[410]
ImplantsOzurdex®Dexamethasone (2009, USA)Acid-terminated PLGA (30%) + ester-terminated PLGA (10%)Macular edemaIVT[411]
Retisert®Fluocinolone acetonide (2005, USA)Ethylene-vinyl acetate coated with polyvinyl alcoholUveitisIVT[412]
Vitrasert® Ganciclovir (1996, USA)Ethylene-vinyl acetate coated with polyvinyl alcoholCytomegalovirus retinitisIVT[412]
I-vation®Triamcinolone acetonide (2007, USA)Poly (methyl methacrylate) and ethylene vinyl acetateDiabetic macular edemaIVT[413]
Iluvien®Fluocinolone acetonide (2014, USA)Polyimide tube coated with Polyvinyl alcoholDiabetic macular edemaIVT[414]
Medidur®Fluocinolone acetonide (2014, USA)Polyvinyl alcoholDiabetic macular edemaIVT[415]
Posurdex®Dexamethasone (2009, USA)PLGAMacular edemaIVT[416]
Surodex®Dexamethasone (2008, USA)PLGAPost-operative inflammationSubscleral placement[417]
Renexus®Encapsulated cell technology (NA)Polyethylene terephthalateAMDIVT[418]
Yutiq®Fluocinolone acetonide (2018, USA)Polyimide/polyvinyl alcoholDiabetic macular edemaIVT[419]
Durysta®Bimatoprost (2020, USA)Poly(D,L-lactide), PLGA, and poly (D,L-lactide) with an acid end groupGlaucomaICI[420]
Dexycu®Dexamethasone (2018, USA)Acetyl triethyl citratePost-operative inflammationPosterior chamber injection[421]
Susvimo®Ranibizumab (2020, Germany)Polysulphone, siliconeAMDIVT[422]
InsertsOcusert® *Pilocarpine (1972, USA)Polyethylene co-vinyl acetateGlaucomaCI[423]
Lacrisert®Hydroxypropyl methyl cellulose (1992, USA)Hydroxypropyl methyl celluloseModerate-to-severe dry eye syndromeCI[424]
BIM ring®Bimatoprost (NA)Support made of polypropylene and covered in a silicone matrix.GlaucomaCI[425,426]
Dextenza®Dexamethasone (2018, USA)Polyethylene glycolPost-operative inflammationCI[414]
Mydriasert®Tropicamid, phenylephrine hydrochloride, and hydroxypropyl methyl cellulose (2015, UK)Ammonium methacrylate copolymerDiagnosis (pupil dilator)Intracanalicular insertion[427]
MicroneedlesXipere™Triamcinolone acetonide (2019, USA)Carboxymethylcellulose sodium, and polysorbate 80Macular edema associated with uveitisSCS[428,429]
Drug eluting contact lensAcuvue®Ketotifen fumarate (2017, USA)Etafilcon AOcular allergic itchTopical[430]
Abbreviations: IVT: intravitreal injection; IV: intravenous; PLGA: poly (lactic-co-glycolic acid); SCS: suprachoroidal space injection; ICI: intracameral injections; CI: cul-de-sac insertion. * The pilocarpine-loaded non-biodegradable insert (Ocusert®) is obsolete due to lack of clinical outcomes.
Table 3. List of polymers that the FDA has authorized for use in the manufacture of ocular formulations, (https://www.accessdata.fda.gov/scripts/cder/iig/index.cfm; accessed on 25 September 2022).
Table 3. List of polymers that the FDA has authorized for use in the manufacture of ocular formulations, (https://www.accessdata.fda.gov/scripts/cder/iig/index.cfm; accessed on 25 September 2022).
PolymerRoute Pharmaceutical FormsCAS Number
Carbomer a Eye surfaceEmulsion
Carbomer b Eye surfaceEmulsion
Carbomer b Eye surfaceGel
Carbomer b Eye surfaceSuspension
Carbomer b Eye surfaceSuspension/drops
Carbomer b Eye surfaceSuspension
Carbomer b Eye surfaceSuspension/drops
Carbomer c Eye surfaceGel
Ethylene-vinyl acetate copolymers (EVA)Eye surfaceInsert, extended release24937788
Ethylene-vinyl acetate copolymers (EVA)Eye surfaceSolution24937788
PEG/PPG-4/30 copolymerEye surfaceSolution
PLGAIntravitrealImplant26780507
PLGAIntravitrealInjection26780507
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

Mostafa, M.; Al Fatease, A.; Alany, R.G.; Abdelkader, H. Recent Advances of Ocular Drug Delivery Systems: Prominence of Ocular Implants for Chronic Eye Diseases. Pharmaceutics 2023, 15, 1746. https://doi.org/10.3390/pharmaceutics15061746

AMA Style

Mostafa M, Al Fatease A, Alany RG, Abdelkader H. Recent Advances of Ocular Drug Delivery Systems: Prominence of Ocular Implants for Chronic Eye Diseases. Pharmaceutics. 2023; 15(6):1746. https://doi.org/10.3390/pharmaceutics15061746

Chicago/Turabian Style

Mostafa, Mahmoud, Adel Al Fatease, Raid G. Alany, and Hamdy Abdelkader. 2023. "Recent Advances of Ocular Drug Delivery Systems: Prominence of Ocular Implants for Chronic Eye Diseases" Pharmaceutics 15, no. 6: 1746. https://doi.org/10.3390/pharmaceutics15061746

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