3,650
Views
13
CrossRef citations to date
0
Altmetric
Review

Macro- and Microscale Properties of the Vitreous Humor to Inform Substitute Design and Intravitreal Biotransport

, &
Pages 429-444 | Received 30 Jun 2020, Accepted 11 Sep 2020, Published online: 12 Oct 2020

ABSTRACT

Research on the vitreous humor and development of hydrogel vitreous substitutes have gained a rapid increase in interest within the past two decades. However, the properties of the vitreous humor and vitreous substitutes have yet to be consolidated. In this paper, the mechanical properties of the vitreous humor and hydrogel vitreous substitutes were systematically reviewed. The number of publications on the vitreous humor and vitreous substitutes over the years, as well as their respective testing conditions and testing techniques were analyzed. The mechanical properties of the human vitreous were found to be most similar to the vitreous of pigs and rabbits. The storage and loss moduli of the hydrogel vitreous substitutes developed were found to be orders of magnitude higher in comparison to the native human vitreous. However, the reported modulus for human vitreous, which was most commonly tested in vitro, has been hypothesized to be different in vivo. Future studies should focus on testing the mechanical properties of the vitreous in situ or in vivo. In addition to its mechanical properties, the vitreous humor has other biotransport mechanisms and biochemical functions that establish a redox balance and maintain an oxygen gradient inside the vitreous chamber to protect intraocular tissues from oxidative damage. Biomimetic hydrogel vitreous substitutes have the potential to provide ophthalmologists with additional avenues for treating and controlling vitreoretinal diseases while preventing complications after vitrectomy. Due to the proximity and interconnectedness of the vitreous humor to other ocular tissues, particularly the lens and the retina, more interest has been placed on understanding the properties of the vitreous humor in recent years. A better understanding of the properties of the vitreous humor will aid in improving the design of biomimetic vitreous substitutes and enhancing intravitreal biotransport.

Introduction to the vitreous humor

The vitreous humor is a gel-like, soft, and transparent ocular tissue located between the lens and the retina, occupying 80% of the eye’s volume.Citation1 It is composed of 98–99% water and a framework of collagen fibers and hyaluronic acid. The hyaluronan coils intersperse and swell in the network of type II and type IX collagen, creating a hypothetical “internal tension” by Donnan swelling.Citation2,Citation3 Due to its soft and viscoelastic nature, the vitreous humor serves as a mechanical damper for the eye, holding the lens and retina in place and protecting these tissues from physical insults. The vitreous also contributes to the growth processes of the ocular tissues during development.Citation4,Citation5 Another vital role of the vitreous is to establish and maintain an oxygen gradient between the lens and retina, where a high concentration of oxygen is found near the metabolically active retinal-pigmented epithelial cells and a low concentration of oxygen is found near the oxygen-sensitive lens epithelial cells.Citation6–9 This biochemical function of the vitreous is accomplished by both a high concentration of the antioxidant ascorbic acid (also known as vitamin C), which consumes oxygen, as well as the gel-like nature of the vitreous humor, which limits transport of oxygen by convection.Citation10,Citation11 The properties of the vitreous are important not only for physical protection from mechanical insults but also for biochemical protection from oxidative damage for other tissues in the eye. Thus, a better understanding of the properties of the vitreous humor may lead to better appreciation of ocular development and disease processes, improved understanding of pharmacokinetics and biotransport of drugs in the vitreous, and enhancement of the design of vitreous substitutes. This review summarizes the material properties of the vitreous humor in comparison to hydrogel vitreous substitutes, focusing particularly on the mechanical properties including storage modulus (G’) and loss modulus (G”), which describe the solid-like/elastic and liquid-like/viscous qualities of a material, respectively. The review concludes by highlighting challenges associated with current testing techniques and suggesting ways to improve vitreous substitutes and better understand intravitreal biotransport.

Studies on the mechanical properties of the vitreous humor

The earliest studies found on the mechanical properties of the vitreous humor were published in the 1970s; however, this topic garnered little attention, generating 10 publications within a 30-year span from 1975 to 2005 ( and ). Since the publication of works by Nickerson et al,Citation3,Citation38 there has been a rise in interest in the mechanical properties of the vitreous humor over the last 15 years (2005–2020), with 27 publications. The mechanical properties of the vitreous humor have commonly been studied using human (n = 13 studies), pig (n = 17 studies), or cow (n = 14 studies) eyes (). Sheep (n = 2 studies), goat (n = 1 study), and rabbits (n = 3 studies) have also been included. The experiments were predominately performed in vitro (n = 31 studies), followed by in situ/ex vivo (n = 15 studies), and finally in vivo (n = 6 studies) (). Rheology was the technique used in most of these studies (n = 28 studies), with shear rheology predominately used (n = 12 studies) (). Other techniques used include magnetic resonance imaging (MRI), ultrasound, and light scattering (n = 10 studies).

Table 1. Summary of publications on the mechanical studies of the vitreous humor. The species, condition, and testing technique are listed for each publication

Figure 1. There has been an increase in publications on the mechanical properties of the vitreous humor over the last 15 years (2005–2020) (A). Human, pig, and cow are the most common species used (B). In vitro testing of the vitreous humor remains the most common testing condition (C). Rheometry is a common technique used to characterize the mechanical properties of the vitreous humor, with shear rheology being the most frequently used technique (D)

Figure 1. There has been an increase in publications on the mechanical properties of the vitreous humor over the last 15 years (2005–2020) (A). Human, pig, and cow are the most common species used (B). In vitro testing of the vitreous humor remains the most common testing condition (C). Rheometry is a common technique used to characterize the mechanical properties of the vitreous humor, with shear rheology being the most frequently used technique (D)

Shear rheology, commonly used to characterize the viscoelasticity of hydrogels and other non-Newtonian fluids, has become a popular technique for investigating the viscoelasticity of the vitreous humor. This technique can report various rheological properties of the vitreous humor including storage and loss modulus (G’ and G”, respectively). Nickerson et al. (2005 and 2008) were the first to use shear rheology to determine the rheological properties of the vitreous humor.Citation3,Citation38 Their works arguably provided a practical methodology for later investigators to study the mechanical properties of the vitreous, particularly with the introduction of the cleated geometry to overcome slippage effects that would otherwise underestimate the mechanical properties of the vitreous in shear testing. Before Nickerson et al. 2005, common techniques used were ultrasound,Citation39 light scattering,Citation35,Citation45 periodic oscillations,Citation43,Citation44 compression chucks,Citation46 MRI,Citation42 or microrheology.Citation20,Citation40 Following the studies by Nickerson et al.,Citation3,Citation38 there have been at least 10 studies that used shear rheology. Some other rheometric techniques used were capillary rheology,Citation33,Citation34 cavitation rheology,Citation32 microrheology,Citation22 and creep testing.Citation14,Citation21

Material properties of the vitreous can change based on the testing temperature and are also rate and strain dependent.Citation3,Citation17,Citation31,Citation34,Citation43 Unfortunately, the variety of methodologies, shear frequencies, and amplitudes used in previous studies make direct comparisons difficult. Even within a single study, the reported modulus can range over a couple orders of magnitude. Additionally, some studies did not report a range of modulus but instead reported a single value for the storage and loss moduli. We attempted to consider both the testing conditions and the modulus range reported, and therefore simplified the information by reporting the average value for the storage modulus, loss modulus, and elastic modulus. If the average modulus values were not published, the average of the highest and lowest value reported was calculated. If a paper reported both a range of the values and the mean, we used the reported mean calculated by the authors. The results were presented as the average of the means ± standard error, with n signifying the number of papers that reported modulus values.

There are significant variations in the mechanical properties of the vitreous humor from different species (human, pig, cow, rabbit, and goat) ( – note the log scale). The storage, loss, and elastic modulus of human (G’ = 78.8 ± 73.7 Pa, n = 4; G” = 26.3 ± 21.8 Pa, n = 4; and E = 1.17 ± 0.56 Pa, n = 3) were much different than those of pig (G’ = 3.57 ± 0.96 Pa, n = 10; G” = 1.04 ± 0.35 Pa, n = 10; and E = 31.2 ± 12.2 Pa, n = 4), rabbit (G’ = 2.76 ± 1.87 Pa, n = 3; G” = 0.73 ± 0.46 Pa, n = 3; E not reported), goat (G’ = 1000 Pa, n = 1; G” = 400 Pa, n = 1; E not reported), sheep (G’ = 47.1 ± 42.9 Pa, n = 2; G” = 46.4 ± 44.1 Pa, n = 2; E not reported), and cow (G’ = 72.7 ± 54.2 Pa, n = 12; G” = 4.44 ± 1.79 Pa, n = 9; and E = 1.95 Pa, n = 1) vitreous. Piccirelli et al. (2012), however, reported human storage and loss moduli both at 300 ± 100 Pa,Citation30 which were significantly larger than values reported in the other studies (ranging from 1 to 10 Pa). Thus, these data points were removed from this analysis after using Grubb’s outlier test (p < .001). The removal of these data points resulted in the human moduli (G’ = 5.09 ± 1.86 Pa, n = 3 and G” = 1.81 ± 0.98 Pa, n = 3) having a similar storage and loss modulus to pig and rabbit samples ( and – note the linear scale). It should be noted that the human storage and loss moduli reported by Piccirelli et al (2012) were not directly measured but calculated using MRI measurements of in vivo human vitreous.Citation30 The assumptions used in their analytical model (homogeneous and spherical vitreous), limited image resolution, and patient compliance might explain the high standard deviation (300 ± 100 Pa) and drastic difference from the moduli directly measured using shear rheology on in vitro dissected vitreous samples reported by other studies.Citation15–17 However, to the best of the authors’ knowledge, Piccirelli et al (2012) is the first study that reported the in vivo storage and loss moduli of the human vitreous humor. It is likely that the measured moduli from in vitro and ex vivo experiments are slightly lower than in vivo due to tissue degradation and destruction of the tissue architecture during testing.

Table 2. Summary of the reported storage, loss, and elastic moduli of the vitreous humor by species

Figure 2. There are significant differences in the modulus between species (A), with human, pig, and rabbit samples having the most similar storage and loss modulus to each other (B). Note the log scale in (A) and the linear scale in (B). The wide variation in the modulus between species results in large variations in the reported modulus using shear rheology and other rheological techniques (C). Combining data for human, pig, and rabbit (excluding cow, sheep, and goat) resulted in reported moduli with a much smaller range of variation (D). Again, note the differences in scale between (C and D). The storage and loss moduli of human, pig, and rabbit vitreous are similar to each other, unlike those of cows, goats, and sheep (E and the zoomed-in plot F). *Human data analyzed includes the results reported by Piccirelli et al. 2012

Figure 2. There are significant differences in the modulus between species (A), with human, pig, and rabbit samples having the most similar storage and loss modulus to each other (B). Note the log scale in (A) and the linear scale in (B). The wide variation in the modulus between species results in large variations in the reported modulus using shear rheology and other rheological techniques (C). Combining data for human, pig, and rabbit (excluding cow, sheep, and goat) resulted in reported moduli with a much smaller range of variation (D). Again, note the differences in scale between (C and D). The storage and loss moduli of human, pig, and rabbit vitreous are similar to each other, unlike those of cows, goats, and sheep (E and the zoomed-in plot F). *Human data analyzed includes the results reported by Piccirelli et al. 2012

The large differences in species resulted in a wide range of moduli reported using shear rheology (G’ = 51.5 ± 41.4 Pa, n = 24 and G” = 23.7 ± 17.5 Pa, n = 23) and other rheometric techniques (G’ = 112.9 ± 92.7 Pa, n = 7 and G” = 1.07 ± 0.50 Pa, n = 5) (). Excluding the data from sheep, goat, and cow studies, the reported moduli for human, rabbit, and pig studies using shear and other rheological techniques were much more uniform compared to data from all species (G’ = 4.16 ± 0.84 Pa, n = 13 and G” = 1.23 ± 0.34 Pa, n = 13 compared to G’ = 1.75 ± 1.23 Pa, n = 3 and G” = 0.68 ± 0.48 Pa, n = 3, respectively) ( and – note the difference in scale compared to ). The studies that used shear rheology generally used similar testing conditions (strain 1–100%, frequency 0.1–100 Hz, temperature 20–37°C), which might explain the similarity in storage and loss moduli reported in these studiesCitation3,Citation17,Citation20,Citation31 (). The studies that used other rheology techniques are more varied (). The scatter plot of different species () and the zoomed-in plot () highlight the similarities in both storage and loss modulus between vitreous samples from humans, pigs, and rabbits. It should be noted that studies utilizing other techniques (ultrasound, light scattering, etc.) have reported other meaningful mechanical properties, but have not directly measured or reported storage and loss modulus, making comparison challenging. The data were obtained in vivo as opposed to data obtained in vitro using rheology (which could also be rate dependent as aforementioned) and, therefore, could not be directly compared to each other.

Table 3. Summary of the reported storage and loss moduli of the vitreous humor based on testing technique (excluding cow, sheep, and goat)

Aging of the vitreous humor

With increasing age, the homogeneous vitreous humor phase separates, becoming a heterogeneous mixture of a stiffened phase composed of aggregated collagen fibrils and a loosened phase composed of dissociated hyaluronans.Citation3,Citation17 This occurs due to a variety of factors such as oxidative damage, digestion by enzymes, and collagen mutations.Citation48,Citation49 The degradation of collagen type IX has been particularly shown to be one of the causes of vitreous degradation.Citation2 Collagen type IX normally coats the outer surface of collagen type II in the vitreous, preventing the collagen fibrils from adhering to each other and preventing collapse of the collagen network. Collagen type IX has a half-life of 11 years, with an exponential decrease in concentration in human vitreous starting at a young age (~ 5 years old).Citation50 With increasing age, more collagen type IX is lost on the surface of the collagen fibrils of the vitreous, leading to an increased probability that the fibrils aggregate. This aggregation of the collagen fibrils causes the collapse of the homogeneous vitreous, resulting in the creation of the solid and liquid vitreous phases due to the increase in the degree of crosslinking of the collagen network and the expulsion of the hyaluronic acid from inside the collagen-hyaluronic acid network, respectively.Citation3,Citation17

The phase separation of the vitreous compromises its ability to function as a protective structure for the surrounding ocular tissues and causes complications including rhegmatogenous retinal detachment, macular holes, vitreous hemorrhage, and vitreous floaters.Citation51 Vitreous floaters, which can be seen as floating shadows in one’s field of vision, are large fibrous aggregations floating in pockets of liquefied vitreous that cast shadows on the retina, interfering with vision. When the liquid pockets develop at the back of the eye, areas of high stress develop where the vitreous connects to the retina, potentially causing retinal tears or macular holes. Thereafter, the liquid component of the vitreous might leak underneath the retina layer through the tear in the retina, lifting the retina away from the choroid and causing retinal detachment. Blood from the retinal vessels may bleed into the vitreous, creating vitreous hemorrhage which might necessitate surgical removal of the tissue if the vitreous hemorrhage does not clear on its own. The degree of phase separation of the vitreous humor has been correlated to signs of cataract formationCitation52 due to the diminished oxygen gradient and the ability of oxygen to travel from the retina to the lens via convectional mixing. Considering the serious complications with the phase-separated vitreous humor, it is important to understand the changing mechanical properties of the aging vitreous.

While the vitreous humor is known to degrade with age, much less is known about its age-related changes in the mechanical properties of the vitreous humor, mostly due to the difficulty in obtaining human donor tissue at younger ages. Zimmerman (1980) reported the in vivo modulus of the vitreous from patients aged 18, 26, 28, 38, 47, and 50 years old and found no discernible differences due to age, potentially due to the unreliable measurements primarily based on patient compliance.Citation45 Weber et al. (1982) calculated the spring constant and damping factor of human vitreous from donors aged 44 to 73 years but did not report any significant age-related changes.Citation44 Shafaie et al. (2018) was the first paper that used shear rheology to measure the modulus of the human vitreous of multiple patients aged 48, 61, 70, 71, and 92, but also did not report any age-related changes in the rheological properties of the vitreous.Citation16 Tram and Swindle-Reilly (2018) was the first study that reported the age-related rheological changes of the human vitreous.Citation17 By testing both the solid and liquid phase of the human vitreous humor with shear rheology, it was determined that the vitreous gel became stiffer while the vitreous liquid became less elastic with increasing age.

There have been two other studies to date that have reported age-related changes in the rheological properties of the vitreous. Colter et al. (2015) investigated age-related changes in the mechanical properties of the vitreous humor using sheep eyes.Citation23 By testing whole ovine vitreous with the hyaloid membrane, they found that the dynamic moduli of the adult vitreous was about 2 times lower than infant vitreous, albeit not statistically significant. More recently, Shultz et al. (2019) analyzed 190 whole human vitreous body samples extracted from donors aged 33 to 92 via shear rheology. The authors found the storage and loss moduli both significantly decreased with age. While these observations appeared contradictory to the findings of Tram and Swindle-Reilly (2018), it should be noted that the entire vitreous was tested in these studies. The hallmark of vitreous liquefaction is phase separation of the vitreous humor, resulting in pockets of liquid inside the vitreous. The older vitreous samples in these studies likely had those pockets of liquid which, when tested as a whole with the vitreous gel, might have lowered the dynamic moduli compared to those from the younger vitreous samples, which would have less liquid phase in the vitreous. In other words, the human vitreous becomes macroscopically softer, likely due to the increase of liquid volume with age. When the solid and liquid phases of the vitreous humor were tested separately, the heterogeneity of the aging vitreous was captured.Citation17

The strive for improved vitreous substitutes and intraocular therapeutic delivery

As the vitreous humor is incapable of regeneration or transplantation, it must be removed and replaced with a substitute during vitrectomy. In this surgical procedure, the vitreous humor is homogenized using a rotating blade and removed via suction. The empty vitreous chamber is then backfilled with a vitreous substitute. Commonly used vitreous substitutes include silicone oil, saline solutions, perfluorocarbon liquids, and gases. These materials serve as temporary tamponades that provide tension on the ocular tissues, keeping the retina in place for healing. It should be noted that these materials are fluids and therefore provide less protection against retinal detachments than the natural viscoelastic vitreous.Citation53 The gold standard for a long-term vitreous substitute is silicone oil, despite its many known complications and drawbacks.Citation54 For example, since silicone oil floats on top of aqueous humor in the eye, patients receiving silicone oil as a vitreous substitute typically have to lay face down for days to weeks to position the silicone oil layer in proper contact with the damaged retina. This reduces patients’ quality of life and compliance, resulting in suboptimal retinal reattachment rates (60%-70%).Citation55 Additionally, silicone oil’s hydrophobic nature renders it more permeable to oxygen than the natural vitreous humor. The diffusivity of oxygen in silicone oil is 8 × 10−9 m2/s,Citation56 which is double its diffusivity in water (4 x 10−9 m2/s).Citation10 This higher diffusivity allows for increased oxygen mixing in the vitreous chamber and the flattening the oxygen gradient,Citation10,Citation11 exposing the sensitive lens to oxidative damage,Citation6,Citation7 and potentially causing a high incidence of cataract formation after vitrectomy.Citation8 Furthermore, the hydrophobicity of silicone oil can cause it to emulsify with the aqueous humor or get trapped in the trabecular meshwork, causing cytotoxicity and potentially inducing glaucoma.Citation55,Citation57 There is a demonstrated clinical need for a new generation of vitreous substitutes that behave like the natural vitreous humor and reduce the side effects associated with the current vitrectomy procedure and use of silicone oil.

Due to its similarities to the natural vitreous, hydrogels have garnered much attention as potential vitreous substitutes. Since hydrogels are hydrophilic, they eliminate the emulsification problem seen with silicone oil vitreous substitutes. The rise in popularity of hydrogel vitreous substitutes is similar to the rise in published research of the mechanical properties of the vitreous humor (). A big discrepancy in the number of publications on hydrogel vitreous substitutes and vitreous humor in the years 1995–2000 (9 publications and 0 publications, respectively) should be noted, considering the approval of silicone oil for use as a vitreous substitute by the Food and Drug Administration (FDA) in 1994.Citation58 SyntheticCitation33,Citation34,Citation59–91 and semisyntheticCitation92–108 hydrogels have been more common materials for experimental vitreous substitutes in recent years, while natural hydrogelsCitation37,Citation109–123 were much more prevalent in the early years (). The foldable capsular vitreous body (FCVB) is a different polymeric design for a vitreous substitute that is currently undergoing clinical trials in China.Citation124–143 The FCVB comprises a polymeric bag that is surgically inserted into the vitreous body and subsequently inflated with the use of filling materials (saline or silicone oil). Overall, a synthetic hydrogel (n = 36) is the most commonly reported type of experimental vitreous substitute (), followed by FCVB (n = 20), semisynthetic hydrogel (n = 17), and natural hydrogel (n = 16) vitreous substitutes. Direct injection (n = 41) is the most common mechanism for instilling the vitreous substitute into the eye (), followed by FCVB (n = 20), in situ crosslinking (n = 16), thermogelling (n = 6), and thermogelling/in situ crosslinking (n = 5). The different types of hydrogel vitreous substitutes and mechanisms of injection are listed in Supplemental Table 1. More details about these vitreous substitutes can be found in several review articles.Citation53,Citation144–154

Figure 3. Number of publications on the mechanical properties of the vitreous humor (from ) compared to the number of publications on hydrogel vitreous substitutes (A). There has been an increase in the publications on hydrogel vitreous substitutes concurrent with publications on the mechanical properties of the vitreous humor. Natural hydrogels attracted the most attention in the early publications, while synthetic and semisynthetic hydrogels have become more prevalent in vitreous substitute research in recent years (B). Foldable capsular vitreous body (FCVB) is another prominent candidate for vitreous substitution. Synthetic hydrogels remain the most common type of material for experimental vitreous substitutes (C). Direct injection of the hydrogel is the most frequently reported mechanism for delivering the vitreous substitute into the eye (D)

Figure 3. Number of publications on the mechanical properties of the vitreous humor (from Figure 1a) compared to the number of publications on hydrogel vitreous substitutes (A). There has been an increase in the publications on hydrogel vitreous substitutes concurrent with publications on the mechanical properties of the vitreous humor. Natural hydrogels attracted the most attention in the early publications, while synthetic and semisynthetic hydrogels have become more prevalent in vitreous substitute research in recent years (B). Foldable capsular vitreous body (FCVB) is another prominent candidate for vitreous substitution. Synthetic hydrogels remain the most common type of material for experimental vitreous substitutes (C). Direct injection of the hydrogel is the most frequently reported mechanism for delivering the vitreous substitute into the eye (D)

Most current experimental hydrogel vitreous substitutes have moduli that are orders of magnitude higher than those of the human vitreous ( and ). However, stiff hydrogels can be difficult to inject through a small gauge needle into the vitreous chamber. The storage and loss moduli of synthetic hydrogelsCitation34,Citation59–62,Citation69,Citation73,Citation74,Citation76,Citation81–83 (G’ = 1203 ± 755 Pa, n = 18 and G” = 373 ± 330 Pa, n = 12) and semisynthetic hydrogelsCitation92–94,Citation97,Citation98,Citation100–102,Citation104 (G’ = 1330 ± 1090 Pa, n = 9 and G” = 291 ± 223 Pa, n = 9) are approximately two orders of magnitude higher, while natural hydrogels’Citation37,Citation110 (G’ = 99.4 ± 53.4 Pa, n = 3 and G” = 51.5 ± 48.5 Pa, n = 2) are about one order of magnitude higher compared to the human vitreousCitation15–17 (G’ = 5.09 ± 1.86 Pa, n = 3 and G” = 1.81 ± 0.98 Pa, n = 3). Hydrogels with different mechanisms of injection into the eye also have a wide variation in storage and loss moduli ( and ). Hydrogel vitreous substitutes with direct injection mechanismCitation59,Citation62,Citation73,Citation76,Citation81–83,Citation94,Citation110 (G’ = 49.9 ± 31.3 Pa, n = 9 and G” = 26.3 ± 21.8 Pa, n = 9) have the closest modulus to the human vitreous, followed by thermogelling/in situ crosslinking hydrogelsCitation37,Citation96,Citation100,Citation101 (G’ = 136 ± 33.1 Pa, n = 4 and G” = 28.9 ± 23.8 Pa, n = 4). The moduli of hydrogels with in situ crosslinkingCitation34,Citation69,Citation93,Citation97,Citation98,Citation104 (G’ = 1626 ± 1031 Pa, n = 13 and G” = 338 ± 332 Pa, n = 6) or thermogellingCitation60,Citation61,Citation74,Citation102 (G’ = 2946 ± 2360 Pa, n = 4 and G” = 1206 ± 937 Pa, n = 4) mechanisms are, on average, more than two orders of magnitude higher than those of the natural human vitreous. Scatter plots of different hydrogel types () and injection mechanisms () highlight the widespread differences in both storage and loss modulus for current experimental hydrogel vitreous substitutes.

Table 4. Summary of the reported storage and loss moduli of the hydrogel vitreous substitutes based on hydrogel type

Table 5. Summary of the reported storage and loss moduli of the hydrogel vitreous substitutes based on delivery mechanism

Figure 4. Hydrogel vitreous substitutes have higher modulus compared to the human vitreous (A), regardless of what type of hydrogel was used. Natural hydrogels are, on average, softer than synthetic and semisynthetic hydrogels, but are still much stiffer than the human vitreous. Different mechanisms for instilling the hydrogel vitreous substitute into the eye also have a wide range of moduli that are also larger than the moduli of the human vitreous. Hydrogels that employ direct injection and thermogelling/in situ crosslinking mechanisms have the most similar modulus compared to the human vitreous, as opposed to hydrogels employing in situ crosslinking or thermogelling mechanisms (B). The storage and loss modulus of different hydrogel types (C) and injection mechanisms (D) differed by orders of magnitude (note that the axes are in logarithmic scale)

Figure 4. Hydrogel vitreous substitutes have higher modulus compared to the human vitreous (A), regardless of what type of hydrogel was used. Natural hydrogels are, on average, softer than synthetic and semisynthetic hydrogels, but are still much stiffer than the human vitreous. Different mechanisms for instilling the hydrogel vitreous substitute into the eye also have a wide range of moduli that are also larger than the moduli of the human vitreous. Hydrogels that employ direct injection and thermogelling/in situ crosslinking mechanisms have the most similar modulus compared to the human vitreous, as opposed to hydrogels employing in situ crosslinking or thermogelling mechanisms (B). The storage and loss modulus of different hydrogel types (C) and injection mechanisms (D) differed by orders of magnitude (note that the axes are in logarithmic scale)

Foldable Capsular Vitreous Body (FCVB) is another prominent candidate for vitreous substitutes. The FCVB, designed to treat severe retinal detachment, consists of a vitreous-shaped capsule with a tube-valve system made of silicone rubber that can be triple folded and implanted into the vitreous cavity of the eye. The mechanical properties of the FCVB depend on the filling materials (saline, silicone oil, or polyvinyl alcohol (PVA)). Since 2008, there have been at least 20 publications on the FCVB, ranging from original design and in vitro mechanical testing to animal testing and human clinical trials.Citation124–143 While radically different from the natural vitreous humor as well as other hydrogel vitreous substitute designs, the FCVB has been shown to be a safe and effective vitreous substitute in humans over a 1- and 3-year observation period. With clinical trials currently underway in China, the FCVB is arguably closest to routine clinical use in humans compared to other experimental hydrogel vitreous substitutes. However, due to its radical design and a more invasive implantation approach (a transscleral port is made for the tube-valve system postoperatively), it remains unknown whether the FCVB will be readily accepted by clinicians and patients.

Loading therapeutics into vitreous substitutes also has the potential to improve their utility. Notably, due to the presence of nano-scale apertures on its capsule surface, the FCVB has been used to sustain release of various ophthalmic therapeutics such as levofloxacin, 5-fluorouracil, siRNA-PKCα, and dexamethasone sodium phosphate.Citation130–132,Citation136,Citation137,Citation141 More recently, a PVA/chitosan hydrogel loaded with fluorouracil-containing poly(lactic-co-glycolic acid) microspheres was proposed as a new vitreous substitute for proliferative vitreoretinopathy.Citation93 Besides providing treatments for vitreoretinal diseases, it may be worthwhile to consider vitreous substitutes with the ability to prevent cataract formation post-vitrectomy. Hydrogel vitreous substitutes loaded with antioxidants such as ascorbic acid have recently been reported to reduce reactive oxygen species activity for lens cells, potentially preventing or reducing the incidence of cataract formation.Citation59 Glutathione, another antioxidant found at high concentrations in the ocular lens,Citation155–157 has recently been used in conjunction with ascorbate in a hydrogel vitreous substitute and was shown to curtail the potential cytotoxicity of ascorbate, prolong its antioxidant activity, and improve its stability in an aqueous environment.Citation158 These developments provide a novel avenue for ophthalmologists to better treat and control vitreoretinal and other ocular diseases while reducing the frequency of therapeutic administration for patients.

The need for in vivo measurements of the vitreous humor

The vitreous connects to specific locations inside the eye in vivo, creating a complex mechanical response that depends not only on the material properties of the vitreous humor but also on the properties of the corneoscleral shell and other intraocular tissues, particularly the lens and retina. A common problem with shear rheology and other rheological techniques is that they are destructive, particularly for irreversibly shear thinning materials like the vitreous.Citation3 The vitreous samples were usually dissected from the eye, disrupting and damaging the internal connection of the vitreous to the eye. Nickerson et al. and various other authors have pointed out that the properties of the vitreous rapidly change following dissection, going from a homogeneous vitreous to a heterogeneous gel surrounded by a pool of liquid.Citation3,Citation17,Citation20,Citation38 This expulsion of liquid from the vitreous humor occurs rapidly, within 10 minutes of dissection. It is likely that the measured properties of the dissected in vitro vitreous samples are different from the properties of intact in vivo vitreous in the eye. Therefore, future studies should focus on the in vivo or in situ mechanical properties of the intact vitreous humor.

Recent attempts to measure the undissected vitreous utilize volume-controlled cavity expansion, cavitation rheology, and creep testing with a submerged rotating rheometric head or microprobes.Citation13,Citation14,Citation22,Citation26,Citation32,Citation159 Volume-controlled cavity expansion is a needle-based technique that employs a large deformation viscoelastic model to capture a material response measured after several cycles of expansion-relaxation at controlled stretch rates in a cavity expansion setting.Citation159 Similarly, cavitation rheology involves inserting a syringe needle into a vitreous sample and inducing an elastic instability via slow pressurization that correlates with the local mechanical properties of the sample.Citation32 In situ rheological creep tests, in which a rotating rheometric head is submerged into the vitreous chamber, were recently used to measure the changes in the mechanical behavior of the vitreous after degradation using collagenase.Citation13,Citation14 Minimally invasive rheological tests of the vitreous humor have also been attempted using microprobes controlled magnetically or optically.Citation22,Citation26 While still relatively invasive (an opening has to be made in the sclera for the insertion of cavitating needle, rheometric head, silica or magnetic beads, etc.), these techniques ensure that the vitreous remains mostly intact, thereby preserving the internal structure of the vitreous humor inside the eye.

Ideally, the mechanical properties of the intact vitreous humor should be tested in vivo. Non-invasive techniques used to study the material properties of the vitreous include MRI,Citation12,Citation19,Citation30,Citation36,Citation42 light scattering,Citation35,Citation45 microbubble-based acoustic radiation forceCitation28, ultrasound,Citation29,Citation39,Citation47 and Brillouin spectroscopy.Citation160 With the advantageous ability of non-destructive in vivo vitreous assessment, the disadvantage of these methods is that they do not provide data (storage modulus and loss modulus) that are comparable to the current rheological data on the vitreous humor or reported experimental hydrogel vitreous substitutes. For example, using Brillouin microscopy, the storage and loss moduli for the porcine vitreous humor were calculated to be 2.3 GPa and 0.7 GPa, which are six orders of magnitude higher than what have been reported in other papers using shear rheology.Citation161

Alternatively, MRI was used to predict the storage and loss moduli of in vitro vitreous-mimicking gels and ex vivo porcine vitreous humor using transverse relaxation time (T2). Correlation between MRI-predicted values and the actual values measured using shear rheology still has room for improvement.Citation12 It should be noted that the T2 relaxation time measurements follow the local motion of water. Since the vitreous is composed of 99% water, T2 relaxation time has great potential to be utilized in future investigations to study the rates of diffusion in the vitreous. This can potentially have a broad impact on our understanding of intravitreal biotransport phenomena including drug delivery, oxygen gradient formation, and antioxidant sequestration. Studying the high-frequency harmonics or vibrational modes of the vitreous humor may also provide some information noninvasively, particularly on the upper bound of the elasticity of the vitreous without the influence of relaxation, deformation, or flow on longer timescales of milliseconds to minutes.Citation162,Citation163 While these techniques are promising and should continue to be investigated in future studies, more work is needed to improve our understanding of the in vivo properties of the vitreous humor.

In vivo mechanical properties of the vitreous humor are useful for improved biomimetic vitreous substitutes. Currently, there is minimal agreement between the known rheological properties of the vitreous humor and the hydrogel vitreous substitutes being developed. Ideally, the material properties of the vitreous substitute should be measured using the same methods as the vitreous humor would be measured.Citation17,Citation34,Citation59 Only a handful of hydrogel vitreous substitutes (n = 13) have similar storage and loss moduli to those reported for the human vitreous humor (less than 20 Pa).Citation34,Citation59,Citation62–64,Citation69,Citation73,Citation76,Citation81,Citation84,Citation110,Citation128,Citation129 In contrast, a majority of the previously proposed vitreous substitutes have modulus values that are orders of magnitude higher than those of the reported human vitreous humor. While the in vivo results for some of these hydrogel vitreous substitutes were promising for short- and intermediate-term applications (less than 1 or 2 years), it is unknown whether the high modulus of these stiff gels would have long-term adverse effects on ocular tissues. However, the modulus of the human vitreous has been hypothesized to be underestimated by current in vitro techniques,Citation3,Citation38 so it may warrant investigation to develop vitreous substitutes with moduli slightly higher than the currently reported moduli for the vitreous humor.

Microscale factors to consider

This review primarily focuses on the macroscale properties of the vitreous humor and hydrogel vitreous substitutes. The microscale properties of the vitreous humor and substitutes are also important factors that need to be considered for the restoration of the natural oxygen gradient and establishment of redox homeostasis in the vitreous humor after vitrectomy. The viscosity, solute concentration, interaction between solute and diffusion medium, mesh size and tortuosity of the diffusion network all contribute to a solute’s diffusion through a medium.Citation164,Citation165 Age-related changes in the vitreous structure adds complexity to the diffusion problem by influencing the hyaluronic acid and collagen network through vitreous liquefaction and fiber aggregation.Citation166 The solute’s surface charge has also been shown to affect the diffusion of a solute in the vitreous humor.Citation167 Shui et al (2009) showed that the ascorbate concentration and oxygen consumption decrease with increasing vitreous liquefaction,Citation11 and Filas et al (2013) showed that the ascorbate content and the structure of the vitreous gel are critical determinants of lens oxygen exposure.Citation10 The phase separation of the vitreous into collagen-rich and hyaluronic-acid-rich vitreous phasesCitation17 further complicates the intravitreal biotransport and pharmacokinetics of therapeutics. The interactions between the vitreous phases (intact or liquefied) and the solutes (oxygen, antioxidants, or other therapeutics) remains largely unknown and warrant further investigation in the future.

A better understanding of the micro- and macrostructures of the vitreous humor can be used to inform designs of hydrogels with appropriate properties, both macroscopically and microscopically, that can better serve as a biomimetic vitreous substitute. Shafaie et al (2018) showed that, even with similar rheological properties between the human, porcine, and bovine vitreous, the steady-state flux of fluorescein (a model drug) through the human vitreous was significantly greater than through both porcine and bovine.Citation16 It should be noted, however, that the human vitreous samples tested were from older adults (aged 48–92 years old) with presumably liquefied vitreous. The ages of bovine and porcine samples were not reported, but the tissues were by-products collected from abattoirs and derived from healthy animals with presumably intact vitreous humor. Therefore, although the macroscopic rheological properties of the human, porcine, and bovine vitreous were similar, the microscale properties of the liquefied human vitreous might be different from those of the intact bovine and porcine vitreous. This might explain why the rate of diffusion and flux in human vitreous was significantly higher than in bovine or porcine vitreous as reported. Future investigations should focus more on how mechanical properties, particularly microscale properties, relate to the mass transfer, antioxidant balance, and establishment of the oxygen gradient in the vitreous humor.

Finally, there has been increased interest in the biotransport of the vitreous humor and vitreous substitutes, particularly due to increased clinical use of intravitreal injection to deliver therapeutics.Citation10,Citation16,Citation18,Citation168–183 The motion of fluid inside the vitreous (intact or liquefied) due to eye rotation or saccadic eye movement has been investigated by many groups,Citation168,Citation169,Citation171,Citation180,Citation183 with particular implications shown for drug delivery and pharmacokinetics in the vitreous.Citation170,Citation172,Citation173,Citation178,Citation184 A model of enzymatic degradation of the vitreous was created and used to study the mobility of intravitreal nanoparticles, which could better simulate the biotransport of a therapeutic injected into aged and liquefied vitreous.Citation18 An in vitro ocular flow model called the PK-Eye was proposed for use in preclinical drug development,Citation175 which can allow for testing of therapeutics in vitreous substitutes.Citation174,Citation181,Citation182 These advances will enable significant progress in determining pharmacokinetics and therapeutic feasibility at earlier stages. However, these advancements are still mostly based on the in vitro properties of the vitreous humor, which likely limits their usefulness and accuracy in in vivo applications. Understanding the in vivo properties of the vitreous will provide insights on the biotransport of intact or degraded vitreous humor and hydrogel vitreous substitutes, particularly on the intravitreal injection of therapeutics or incorporation of therapeutics inside a vitreous substitute.

Future directions

The degradation of the vitreous humor causes complications that could benefit from further investigations. The heterogenization of vitreous humor can disrupt the internal structure of the vitreous humor and unbalance the stress distribution in the eye. As the vitreous phase-separates, the retina experiences increased traction at locations with strong adhesion to the vitreous humor. Interestingly, the adhesion between the vitreous and the retina has been shown to decrease with age.Citation185 This reduction in vitreoretinal adhesion coincides with the timeline of vitreous liquefaction and has been hypothesized to facilitate posterior vitreous detachment, which might protect the retina from tearing or detaching. Induction of posterior vitreous detachment was recently reported, with particular focuses on both vitreous liquefaction and dehiscence of vitreoretinal adhesion.Citation186 A mechanical model of posterior vitreous detachment and generation of vitreoretinal tractions has also been published.Citation187 These are all interesting biomechanical problems that should be further investigated in future studies.

Additionally, the phase-separated vitreous humor alters the diffusion and transport of various nutrient and waste molecules, contributing to complications including cataract formation and proliferative vitreoretinopathy. For example, the gel quality of the vitreous humor has been shown to negatively correlate with the concentration of ascorbic acid in the vitreous,Citation11 and that a liquefied vitreous allows for more mixing of oxygen via convection,Citation10 thereby exposing the lens to more oxidative damage and increased incidence of cataract formation.Citation8,Citation52,Citation188 Advanced glycation end-products (AGEs) formation, a reactive-oxygen-species-dependent process that increases the accumulation of collagen crosslinks, has been implicated in vision loss associated with cataract formation, diabetic retinopathy, and glaucoma.Citation189 AGEs has been shown to increase in the vitreous with age, especially in patients with diabetesCitation190 and/or rhegmatogenous retinal detachment.Citation191,192 AGE accumulation has also been shown to reduce vitreous permeability,Citation192 which might explain the disruption to the antioxidant balance and the oxygen gradient in liquefied vitreous. The connection between a healthy oxygen gradient, homogeneous vitreous, rate of AGE formation, and biotransport of molecules (ascorbate, glutathione, other therapeutics, etc.) provides an opportunity to tie together multiple threads in this review and hopefully brings together interdisciplinary teams of clinicians, engineers, and scientists to answer these multifaceted problems in future investigations.

Future studies can expand upon limitations described above. In this review, the reported moduli were simplified and do not reflect the diversity in the testing conditions. A closer investigation of the testing methods, strain, rate, and temperature might help identify the source of variability between the studies. This review also mostly focused on the macroscopic properties of the vitreous humor and hydrogel vitreous substitutes. However, microscale factors are also important to the structure and function of the vitreous and will need to be further investigated. A better understanding of the macro- and microscale properties and age-related changes of the human vitreous humor will aid in the design of biomimetic vitreous substitutes, enhancement in analyzing transport of therapeutics in the vitreous, and comprehension of the pathological conditions of the vitreous humor.

Conclusions

Studies on the vitreous humor and vitreous substitutes have increased in recent years. Samples from several species have been evaluated, with the mechanical properties of pig and rabbit vitreous found to be most similar to the native human vitreous, warranting the preferred use of pig and rabbit models for future studies. The vitreous has mostly been tested in vitro, despite known limitations relating to the rapid changes of the vitreous humor once removed from the eye. With increasing age, the vitreous phase-separates and the stiffness of the whole vitreous was found to decrease, while there was localized stiffening of the solid-like components of the aged vitreous. Future work should focus on determining the in vivo properties of the human vitreous humor and the biotransport of nutrients, waste, and injected therapeutics in the intact and degraded vitreous humor and hydrogel vitreous substitutes. The recent advancements in hydrogel vitreous substitutes are promising; however, more focus should be placed on ensuring the biomimicry of the substitutes, especially relating to the rheological properties. Finally, drug-eluting hydrogel vitreous substitutes are a promising new generation of vitreous substitutes that have the potential to shift the design paradigm of current vitreous substitutes to a more holistic strategy that considers not only the mechanical roles but also the biochemical functions of the natural vitreous humor.

Patent Applications

Reilly KE, Reilly MA, Tram NK. Antioxidant-Releasing Vitreous Substitutes and Uses Thereof. US Patent Application PCT/US2020/017525, Feb. 10, 2020.

Supplemental material

Supplemental Material

Download PDF (148.2 KB)

Acknowledgments

We would like to acknowledge the past and present members of the Swindle-Reilly Lab for Biomimetic Polymeric Biomaterials for help and encouragement, and Drs. Matthew Reilly, Cynthia Roberts, and Jun Liu for advice and research guidance.

Data availability statement

The data that support the findings of this study are available from the corresponding author, KESR, upon request.

Supplementary material

Supplemental data for this article can be accessed on the publisher’s website.

Additional information

Funding

We would like to acknowledge The Ohio State University College of Engineering for funding.

References

  • Swindle-Reilly KE, Reilly MA, Ravi N. Chapter 5: current concepts in the design of hydrogels as vitreous substitutes. In: Chirila T, Harkin D editors. Biomaterials and regenerative medicine in ophthalmology. Second. Oxford (UK): Woodhead Publishing; 2016. p. 101–30.
  • Le Goff MM. Adult vitreous structure and postnatal changes. Eye (Lond). 2008;22:1214–22.
  • Nickerson CS, Park J, Kornfield JA, Karageozian H. Rheological properties of the vitreous and the role of hyaluronic acid. J Biomech. 2008;41:1840–46.
  • Coulombre AJ. The role of intraocular pressure in the development of the chick eye. I. Control of Eye Size. J Exper Zool. 1956;133:211–25.
  • Coulombre AJ, Coulombre JL. The role of intraocular pressure in the development of the chick eye. IV. Corneal Curvature. AMA Arch Ophthalmol. 1958;59:502–06.
  • Siegfried CJ, Shui YB. Intraocular oxygen and antioxidant status: new insights on the effect of vitrectomy and glaucoma pathogenesis. Am J Ophthalmol. 2019;203:12–25.
  • Siegfried CJ, Shui YB, Tian B, Nork TM, Heatley GA, Kaufman PL. Effects of vitrectomy and lensectomy on older rhesus macaques: oxygen distribution, antioxidant status, and aqueous humor dynamics. Invest Ophthalmol Vis Sci. 2017;58:4003–14.
  • Beebe DC, Shui Y, Siegfried CJ, Holekamp NM, Bai F. Preserve the (intraocular) environment: the importance of maintaining normal oxygen gradients in the eye. Jpn J Ophthalmol. 2014;58:225–31.
  • Siegfried CJ, Shui YB, Holekamp NM, Bai F, Beebe DC. Oxygen distribution in the human eye: relevance to the etiology of open-angle glaucoma after vitrectomy. Invest Ophthalmol Vis Sci. 2010;51:5731–38.
  • Filas BA, Shui Y-B, Beebe DC. Computational model for oxygen transport and consumption in human vitreous. Invest Ophthalmol Vis Sci. 2013;54(10):6549–59. doi:10.1167/iovs.13-12609.
  • Shui YB, Holekamp NM, Kramer BC, Crowley JR, Wilkins MA, Chu F, Malone PE, Mangers SJ, Hou JH, Siegfried CJ, et al. The gel state of the vitreous and ascorbate-dependent oxygen consumption: relationship to the etiology of nuclear cataracts. Arch Ophthalmol. 2009;127(4):475–82.
  • Thakur SS, Pan X, Kumarasinghe GL, Yin N, Pontré BP, Vaghefi E, Rupenthal ID. Relationship between rheological properties and transverse relaxation time (T2) of artificial and porcine vitreous humour. Exp Eye Res. 2020;194:108006.
  • Rangchian A, Hubschman JP, Kavehpour HP. Time dependent degradation of vitreous gel under enzymatic reaction: polymeric network role in fluid properties. J Biomech. 2020;109:109921.
  • Rangchian A, Francone A, Farajzadeh M, Hosseini H, Connelly K, Hubschman JP, Kavehpour P. Effects of collagenase type II on vitreous humor, an in-situ rheological study. J Biomech Eng. 2019;141(8):081007.
  • Schulz A, Wahl S, Rickmann A, Ludwig J, Stanzel BV, von Briesen H, Szurman P. Age-related loss of human vitreal viscoelasticity. Transl Vis Sci Technol. 2019;8:56.
  • Shafaie S, Hutter V, Brown MB, Cook MT, Chau DY. Diffusion through the ex vivo vitreal body–Bovine, porcine, and ovine models are poor surrogates for the human vitreous. Int J Pharm. 2018;550:207–15.
  • Tram NK, Swindle-Reilly KE. Rheological properties and age-related changes of the human vitreous humor. Front Bioeng Biotechnol. 2018. 6.
  • Huang D, Chen YS, Xu Q, Hanes J, Rupenthal ID. Effects of enzymatic degradation on dynamic mechanical properties of the vitreous and intravitreal nanoparticle mobility. Eur J Pharm Sci. 2018;118:124–33.
  • Stein S, Hadlich S, Langner S, Biesenack A, Zehm N, Kruschke S, Oelze M, Grimm M, Mahnhardt S, Weitschies W, Seidlitz A. 7.1 T MRI and T2 mapping of the human and porcine vitreous body post mortem. Eur J Pharm Biopharm. 2018;131:82–91.
  • Silva AF, Alves MA, Oliveira MSN. Rheological behaviour of vitreous humour. Rheol Acta. 2017;56:377–86.
  • Shah NS, Beebe DC, Lake SP, Filas BA. On the spatiotemporal material anisotropy of the vitreous body in tension and compression. Ann Biomed Eng. 2016;44:3084–95.
  • Pokki J, Ergeneman O, Sevim S, Enzmann V, Torun H, Nelson BJ. Measuring localized viscoelasticity of the vitreous body using intraocular microprobes. Biomed Microdevices. 2015;17:85.
  • Colter J, Williams A, Moran P, Coats B. Age-related changes in dynamic moduli of ovine vitreous. J Mech Behav Biomed Mater. 2015;41:315–24.
  • Abdelkawi SA, Abdel-Salam AM, Ghoniem DF, Ghaly SK. Vitreous humor rheology after nd: yAGlaser photo disruption. Cell Biochem Biophys. 2014;68:267–74.
  • Filas BA, Zhang Q, Okamoto RJ, Shui Y, Beebe DC. Enzymatic degradation identifies components responsible for the structural properties of the vitreous body. Invest Ophthalmol Vis Sci. 2014;55:55–63.
  • Watts F, Tan LE, Wilson CG, Girkin JM, Tassieri M, Wright AJ. Investigating the micro-rheology of the vitreous humor using an optically trapped local probe. J Opt. 2014;16:015301.
  • Zhang Q, Filas BA, Roth R, Heuser J, Ma N, Sharma S, Panitch A, Beebe DC, Shui YB. Preservation of the structure of enzymatically-degraded bovine vitreous using synthetic proteoglycan mimics. Invest Ophthalmol Vis Sci. 2014;55:8153–62.
  • Yoon S, Aglyamov S, Karpiouk A, Correspondence: ES. Spatial variations of viscoelastic properties of porcine vitreous humors. IEEE Trans Ultrason Ferroelectr Freq Control. 2013;60:2453–60.
  • Rossi T, Querzoli G, Pasqualitto G, Iossa M, Placentino L, Repetto R, Stocchino A, Ripandelli G. Ultrasound imaging velocimetry of the human vitreous. Exp Eye Res. 2012;99:98–104.
  • Piccirelli M, Bergamin O, Landau K, Boesiger P, Luechinger R. Vitreous deformation during eye movement. NMR Biomed. 2012;25:59–66.
  • Sharif-Kashani P, Hubschman JP, Sassoon D, Kavehpour HP. Rheology of the vitreous gel: effects of macromolecule organization on the viscoelastic properties. J Biomech. 2011;44:419–23.
  • Zimberlin JA, McManus JJ, Crosby AJ. Cavitation rheology of the vitreous: mechanical properties of biological tissue. Soft Matter. 2010;6:3632–35.
  • Swindle-Reilly KE, Shah M, Hamilton PD, Eskin TA, Kaushal S, Ravi N. Rabbit study of an in situ forming hydrogel vitreous substitute. Invest Ophthalmol Vis Sci. 2009;50:4840–46.
  • Swindle KE, Hamilton PD, Ravi N. In situ formation of hydrogels as vitreous substitutes: viscoelastic comparison to porcine vitreous. J Biomed Mater Res A. 2008;87:656–65.
  • Sebag J, Ansari RR, Suh KI. Pharmacologic vitreolysis with microplasmin increases vitreous diffusion coefficients. Graefes Arch Clin Exp Ophthalmol. 2007;245:576–80.
  • Patz S, Bert RJ, Frederick E, Freddo TF. T(1) and T(2) measurements of the fine structures of the in vivo and enucleated human eye. J Magn Reson Imaging. 2007;26:510–18.
  • Suri S, Banerjee R. In vitro evaluation of in situ gels as short term vitreous substitutes. J Biomed Mater Res A. 2006;79:650–64.
  • Nickerson CS, Karageozian HL, Park J, Kornfield JA. Internal tension: a novel hypothesis concerning the mechanical properties of the vitreous humor. Macromol Symp. 2005;227:183–89.
  • Walton KA, Meyer CH, Harkrider CJ, Cox TA, Toth CA. Age-related changes in vitreous mobility as measured by video B scan ultrasound. Exp Eye Res. 2002;74:173–80.
  • Lee B, Litt M, Buchsbaum G. Rheology of the vitreous body: part 2. Viscoelasticity of Bovine and Porcine Vitreous. Biorheology. 1994;31:327–38.
  • Lee B, Litt M, Buchsbaum G. Rheology of the vitreous body. Part I: Viscoelasticity of Human Vitreous. Biorheology. 1992;29:521–33.
  • Aguayo J, Glaser B, Mildvan A, Cheng HM, Gonzalez RG, Brady T. Study of vitreous liquifaction by NMR spectroscopy and imaging. Invest Ophthalmol Vis Sci. 1985;26:692–97.
  • Tokita M, Fujiya Y, Hikichi K. Dynamic viscoelasticity of bovine vitreous body. Biorheology. 1984;21:751–56.
  • Weber H, Landwehr G, Kilp H, Neubauer H. The mechanical properties of the vitreous of pig and human donor eyes. Ophthalmic Res. 1982;14:335–43.
  • Zimmerman RL. In vivo measurements of the viscoelasticity of the human vitreous humor. Biophys J. 1980;29:539–44.
  • Bettelheim FA, Wang TJY. Dynamic viscoelastic properties of bovine vitreous. Exp Eye Res. 1976;23:435–41.
  • Oksala A. Ultrasonic findings in the vitreous body at different ages and in patients with detachment of the retina. Albrecht Von Graefes Arch Klin Exp Ophthalmol. 1975;197:83–87.
  • Okada Y, Konomi H, Yada T, Kimata K, Nagase H. Degradation of type IX collagen by matrix metalloproteinase 3 (stromelysin) from human rheumatoid synovial cells. FEBS Lett. 1989;244:473–76.
  • Brown DJ, Bishop P, Hamdi H, Kenney MC. Cleavage of structural components of mammalian vitreous by endogenous matrix metalloproteinase-2. Curr Eye Res. 1996;15:439–45.
  • Sebag J. Ageing of the vitreous. Eye. 1987;1:254–62.
  • Los LI, van der Worp RJ, van Luyn MJ, Hooymans JM. Age-related liquefaction of the human vitreous body: LM and TEM evaluation of the role of proteoglycans and collagen. Invest Ophthalmol Vis Sci. 2003;44:2828–33.
  • Harocopos GJ, Shui YB, McKinnon M, Holekamp NM, Gordon MO, Beebe DC. Importance of vitreous liquefaction in age-related cataract. Invest Ophthalmol Vis Sci. 2004;45:77–85.
  • Kleinberg TT, Tzekov RT, Stein L, Ravi N, Kaushal S. Vitreous substitutes: a comprehensive review. Surv Ophthalmol. 2011 Jul-Aug;56(4):300–23.
  • Federman JL, Schubert HD. Complications associated with the use of silicone oil in 150 eyes after retina-vitreous surgery. Ophthalmology. 1988;95:870–76.
  • Giordano GG, Refojo MF. Silicone oils as vitreous substitutes. Prog Polym Sci. 1998;23:509–32.
  • Ding C, Fan Y. Measurement of diffusion coefficients of air in silicone oil and in hydraulic oil. Chinese Journal of Chemical Engineering. 2011;19:205–11.
  • Foulds WS. Is your vitreous really necessary? Eye (Lond). 1987;1:641–64.
  • Snider S. FDA approves silicone oil for retinal reattachment. Rockville, MD: US Food and Drug Administration; 1994.
  • Tram NK, Jiang P, Torres‐Flores TC, Jacobs KM, Chandler HL, Swindle‐Reilly KE. A hydrogel vitreous substitute that releases antioxidant. Macromol Biosci. 2020;20:e1900305.
  • Xue K, Liu Z, Jiang L, Kai D, Li Z, Su X, Loh XJ. A new highly transparent injectable PHA-based thermogelling vitreous substitute. Biomater Sci. 2020;8:926–36.
  • Liu Z, Liow SS, Lai SL, Alli-Shaik A, Holder GE, Parikh BH, Krishnakumar S, Li Z, Tan MJ, Gunaratne J, et al. Retinal-detachment repair and vitreous-like-body reformation via a thermogelling polymer endotamponade. Nat Biomed Eng. 2019;3(8):598–610.
  • Wang H, Wu Y, Cui C, Yang J, Liu W. Antifouling super water absorbent supramolecular polymer hydrogel as an artificial vitreous body. Adv Sci. 2018;5:1800711.
  • Hayashi K, Okamoto F, Hoshi S, Katashima T, Zujur DC, Li X, Shibayama M, Gilbert EP, Chung UI, Ohba S, et al. Fast-forming hydrogel with ultralow polymeric content as an artificial vitreous body. Nat Biomed Eng. 2017;1:0044.
  • Liang J, Struckhoff JJ, Du H, Hamilton PD, Ravi N. Synthesis and characterization of in situ forming anionic hydrogel as vitreous substitutes. J Biomed Mater Res B Appl Biomater. 2017;105:977–88.
  • Davis JT, Hamilton PD, Poly RN. (acrylamide co-acrylic acid) for use as an in situ gelling vitreous substitute. J Bioact Compat Polym. 2017;32:528–41.
  • Morandim-Giannetti A, Silva RC, Magalhães O, Junior OM, Schor P, Bersanetti PA. Conditions for obtaining polyvinyl alcohol/trisodium trimetaphosphate hydrogels as vitreous humor substitute. J Biomed Mater Res B Appl Biomater. 2016;104:1386–95.
  • Chang J, Tao Y, Wang B. An in situ-forming zwitterionic hydrogel as vitreous substitute. J Mater Chem B. 2015;3:1097–105.
  • Chang J, Tao Y, Wang B. Evaluation of a redox-initiated in situ hydrogel as vitreous substitute. Polymer. 2014;55:4627–33.
  • Tao Y, Tong X, Zhang Y. Evaluation of an in situ chemically crosslinked hydrogel as a long-term vitreous substitute material. Acta Biomater. 2013;9:5022–30.
  • Strotmann F, Wolf I, Galla HJ. The biocompatibility of a polyelectrolyte vitreous body substitute on a high resistance in vitro model of the blood-retinal barrier. J Biomater Appl. 2013 Sep;28(3):334–42.
  • Lamponi S, Leone G, Consumi M, Greco G, Magnani A. In vitro biocompatibility of new PVA-based hydrogels as vitreous body substitutes. J Biomater Sci Polym Ed. 2012;23:555–75.
  • Bohm I, Strotmann F, Koopmans C, Wolf I, Galla HJ, Ritter H. Two-component in situ forming supramolecular hydrogels as advanced biomaterials in vitreous body surgery. Macromol Biosci. 2012 Apr;12(4):432–37.
  • Pritchard CD, Crafoord S, Andreasson S, Arner KM, O’Shea TM, Langer R, Ghosh FK. Evaluation of viscoelastic poly(ethylene glycol) sols as vitreous substitutes in an experimental vitrectomy model in rabbits. Acta Biomater. 2011 Mar;7(3):936–43.
  • Annaka M, Mortensen K, Vigild ME, Matsuura T, Tsuji S, Ueda T, Tsujinaka H. Design of an injectable in situ gelation biomaterials for vitreous substitute. Biomacromolecules. 2011 Nov 14;12(11):4011–21.
  • Strotmann F, Bezdushna E, Ritter H, Galla HJ. In situ forming hydrogels: a Thermo‐Responsive polyelectrolyte as promising liquid artificial vitreous body replacement. Adv Eng Mater. 2011;13:B172–80.
  • Leone G, Consumi M, Aggravi M, Donati A, Lamponi S, Magnani A. PVA/STMP based hydrogels as potential substitutes of human vitreous. J Mater Sci Mater Med. 2010 Aug;21(8):2491–500.
  • Maruoka S, Matsuura T, Kawasaki K, Okamoto M, Yoshiaki H, Kodama M, Sugiyama M, Annaka M. Biocompatibility of polyvinylalcohol gel as a vitreous substitute. Curr Eye Res. 2006;31:599–606.
  • Foster WJ, Aliyar HA, Hamilton P, Ravi N. Internal osmotic pressure as a mechanism of retinal attachment in a vitreous substitute. J Bioact Compat Polym. 2006;21:221–35.
  • Aliyar HA, Foster WJ, Hamilton PD, Ravi N. Towards the development of an artificial human vitreous. Polym Prep. 2004;45:469–70.
  • Bruining MJ, Edelbroek-Hoogendoorn PS, Blaauwgeers HG, Mooy CM, Hendrikse FH, Koole LH. New biodegradable networks of poly(N-vinylpyrrolidinone) designed for controlled nonburst degradation in the vitreous body. J Biomed Mater Res. 1999;47:189–97.
  • Chirila TV, Hong YE. Poly (1‐vinyl‐2‐pyrrolidinone) hydrogels as vitreous substitutes: a rheological study. Polym Int. 1998;46:183–95.
  • Hong Y, Chirila TV, Vijayasekaran S, Shen W, Lou X, Dalton PD. Biodegradation in vitro and retention in the rabbit eye of crosslinked poly(1-vinyl-2-pyrrolidinone) hydrogel as a vitreous substitute. J Biomed Mater Res. 1998;39:650–59.
  • Hong Y, Chirila TV, Cuypers MJ, Constable IJ. Polymers of 1-vinyl-2-pyrrolidinone as potential vitreous substitutes: physical selection. J Biomater Appl. 1996;11:135–81.
  • Hong Y, Chirila TV, Vijayasekaran S. Crosslinked poly(1-vinyl-2-pyrrolidinone) as a vitreous substitute. J Biomed Mater Res. 1996;30:441–48.
  • Vijayasekaran S, Chirila TV, Hong Y, Tahija SG, Dalton PD, Constable IJ, McAllister IL. Poly(1-vinyl-2-pyrrolidinone) hydrogels as vitreous substitutes: histopathological evaluation in the animal eye. J Biomater Sci Polym Ed. 1996;7:685–96.
  • Chirila TV, Constable IJ, Hong Y, Vijayasekaran S, Humphrey MF, Dalton PD, Tahija SG, Maley MAL, Cuypers MJH, Sharp C, et al. Synthetic hydrogel as an artificial vitreous body. A One-year Animal Study of Its Effects on the Retina. Eur Cell Mater. 1995;5(1):83–96.
  • Yamauchi A. Synthetic vitreous body of PVA hydrogel. In: DeRossi D, Kajiwara K, Osada Y, Yamauchi A, editors. Polymer gels. Boston (MA): Springer. 1991, p. 127–134.
  • Hogen-Esch TE, Shah KR, Fitzgerald CR. Development of injectable poly(glyceryl methacrylate) hydrogels for vitreous prosthesis. J Biomed Mater Res. 1976;10:975–76.
  • Mueller-Jensen K. Polyacrylamide as an alloplastic vitreous implant. Albrecht Von Graefes Arch Klin Exp Ophthalmol. 1973;189:147–258.
  • Refojo MF, Zauberman H. Optical properties of gels designed for vitreous implantation. Invest Ophthalmol. 1973;12:465–67.
  • Daniele S, Refojo MF, Schepens CL, Freeman HM. Glyceryl methacrylate hydrogel as a vitreous implant. An Experimental Study. Arch Ophthalmol. 1968;80:120–27.
  • Laradji A, Shui YB, Karakocak BB, Evans L, Hamilton P, Ravi N. Bioinspired thermosensitive hydrogel as a vitreous substitute: synthesis, properties, and progress of animal studies. Materials. 2020;13:E1337.
  • Yu Z, Ma S, Wu M, Cui H, Wu R, Chen S, Xu C, Lu X, Feng S. Self‐assembling hydrogel loaded with 5‐FU PLGA microspheres as A novel vitreous substitute for proliferative vitreoretinopathy. J Biomed Mater Res A. Epub ahead of print. PMID: 32419359.
  • Januschowski K, Schnichels S, Hurst J, Hohenadl C, Reither C, Rickmann A, Pohl L, Bartz-Schmidt KU, Spitzer MS. Ex vivo biophysical characterization of a hydrogel-based artificial vitreous substitute. PLoS One. 2019;14:e0209217.
  • Barth H, Crafoord S, Arnér K, Ghosh F. Inflammatory responses after vitrectomy with vitreous substitutes in a rabbit model. Graefes Arch Clin Exp Ophthalmol. 2019;257:769–83.
  • Santhanam S, Shui YB, Struckhoff J, Karakocak BB, Hamilton PD, Harocopos GJ, Ravi N. Bioinspired fibrillary hydrogel with controlled swelling behavior: applicability as an artificial vitreous. ACS Appl Bio Mater. 2018;2:70–80.
  • Jiang X, Peng Y, Yang C, Liu W, Han B. The feasibility study of an in situ marine polysaccharide‐based hydrogel as the vitreous substitute. J Biomed Mater Res A. 2018;106(7):1997–2006.
  • Schnichels S, Schneider N, Hohenadl C, Hurst J, Schatz A, Januschowski K, Spitzer MS. Efficacy of two different thiol-modified crosslinked hyaluronate formulations as vitreous replacement compared to silicone oil in a model of retinal detachment. PLoS One. 2017;12:e0172895.
  • Barth H, Crafoord S, Andréasson S, Ghosh F. A cross-linked hyaluronic acid hydrogel (healaflow®) as a novel vitreous substitute. Graefes Arch Clin Exp Ophthalmol. 2016;254:697–703.
  • Morozova S, Hamilton P, Ravi N, Muthukumar M. Development of a vitreous substitute: incorporating charges and fibrous structures in synthetic hydrogel materials. Macromolecules. 2016;49:4619–26.
  • Santhanam S, Liang J, Struckhoff JJ, Hamilton PD, Ravi N. Biomimetic hydrogel with tunable mechanical properties for vitreous substitutes. Acta Biomater. 2016;43:327–37.
  • Lin YK, Chen KH, Kuan CY. The synthesis and characterization of a thermally responsive hyaluronic acid/pluronic copolymer and an evaluation of its potential as an artificial vitreous substitute. J Bioact Compat Polym. 2013;28:355–57.
  • Schramm C, Spitzer MS, Henke-Fahle S, Steinmetz G, Januschowski K, Heiduschka P, Geis-Gerstorfer J, Biedermann T, Bartz-Schmidt KU, Szurman P. The cross-linked biopolymer hyaluronic acid as an artificial vitreous substitute. Invest Ophthalmol Vis Sci. 2012 Feb 2;53(2):613–21.
  • Su W, Chen K, Chen Y, Lee Y, Tseng C, Lin F. An injectable oxidated hyaluronic acid/adipic acid dihydrazide hydrogel as a vitreous substitute. J Biomater Sci Polym Ed. 2011;22:1777–97.
  • De Jong C, Bali E, Libert J, Caspers-Velu -LADCON-L. hydrogel as a vitreous substitute: preliminary results. Bull Soc Belge Ophtalmol. 2000;278:71–75.
  • Liang C, Peyman GA, Serracarbassa P, Calixto N, Chow AA, Rao P. An evaluation of methylated collagen as a substitute for vitreous and aqueous humor. Int Ophthalmol. 1998;22:13–18.
  • Nakagawa M, Tanaka M, Miyata T. Evaluation of collagen gel and hyaluronic acid as vitreous substitutes. Ophthalmic Res. 1997;29:409–20.
  • Fernandez-Vigo J, Refojo MF, Verstraeten T. Evaluation of a viscoelastic solution of hydroxypropyl methylcellulose as a potential vitreous substitute. Retina. 1990;10:148–52.
  • Raia NR, Jia D, Ghezzi CE, Muthukumar M, Kaplan DL. Characterization of silk-hyaluronic acid composite hydrogels towards vitreous humor substitutes. Biomaterials. 2020;233:119729.
  • Uesugi K, Sakaguchi H, Hayashida Y. A self-assembling peptide gel as a vitreous substitute: a rabbit study. Invest Ophthalmol Vis Sci. 2017;58:4068–75.
  • Koster R, Stilma JS. Healon as intravitreal substitute in retinal detachment surgery in 40 patients. Doc Ophthalmol. 1986;64:13–17.
  • Pruett RC, Schepens CL, Swann DA. Hyaluronic acid vitreous substitute. A Six-year Clinical Evaluation. Arch Ophthalmol. 1979;97:2325–30.
  • Pruett RC, Schepens CL, Freeman HM. Collagen vitreous substitute. II. Preliminary Clinical Trials. Arch Ophthalmol. 1974;91:29–32.
  • Balazs EA, Freeman MI, Klöti R, Meyer-Schwickerath G, Regnault F, Sweeney DB. Hyaluronic acid and replacement of vitreous and aqueous humor. Mod Probl Ophthalmol. 1972;10:3–21.
  • Pruett RC, Calabria GA, Schepens CL. Collagen vitreous substitute. I. Experimental Study. Arch Ophthalmol. 1972;88:540–43.
  • Dunn M, Shafer D, Stenzel KH, Rubin AL, Miyata T, Hopkins L, Powers MJ. Collagen as a vitreous heterograft. Trans Am Soc Artif Intern Organs. 1971;17:421–23.
  • Dunn MW, Stenzel KH, Rubin AL, Miyata T. Collagen implants in the vitreous. Arch Ophthalmol. 1969;82:840–44.
  • Stenzel KH, Dunn MW, Rubin AL, Miyata T. Collagen gels: design for a vitreous replacement. Science. 1969;164:1282–83.
  • Dunn MW, Miyata T, Stenzel KH, Rubin AL. Studies on collagen implants in the vitreous. Surg Forum. 1968;19:492–94.
  • Gombos GM, Berman ER. Chemical and clinical observations on the fate of various vitreous substitutes. Acta Ophthalmol. 1967;45:794–806.
  • Balazs EA, Sweeney DB. The replacement of the vitreous body in the monkey by reconstituted vitreous and by hyaluronic acid. Bibl Ophthalmol. 1966;70:230–32.
  • Oosterhuis JA. Polygeline as a vitreous substitute. II. Clinical Results. Arch Ophthalmol. 1966;76:374–77.
  • Oosterhuis JA, van Haeringen NJ, Jeltes IG, Glasius E. Polygeline as a vitreous substitute. I. Observations in Rabbits. Arch Ophthalmol. 1966;76:258–65.
  • Zhang X, Tian X, Zhang B, Guo L, Li X, Jia Y. Study on the effectiveness and safety of foldable capsular vitreous body implantation. BMC Ophthalmol. 2019;19:260.
  • Zhang B, Li C, Jia Y, Li X, Guo L, Wang C, Tian X. A pilot clinical study of treating rhegmatogenous retinal detachment by silicone rubber balloon scleral buckling. Retina. 2019;40(10):1918–28.
  • Lin X, Sun X, Wang Z. Three-year efficacy and safety of a silicone oil-filled foldable-capsular-vitreous-body in three cases of severe retinal detachment. Transl Vis Sci Technol. 2016;5:2.
  • Yang W, Yuan Y, Zong Y, Huang Z, Mai S, Li Y, Qian X, Liu Y, Gao Q. Preliminary study on retinal vascular and oxygen-related changes after long-term silicone oil and foldable capsular vitreous body tamponade. Sci Rep. 2014;4:5272.
  • Feng S, Chen H, Liu Y, Huang Z, Sun X, Zhou L, Lu X, Gao Q. A novel vitreous substitute of using a foldable capsular vitreous body injected with polyvinylalcohol hydrogel. Sci Rep. 2013;3:1838.
  • Chen H, Feng S, Liu Y, Huang Z, Sun X, Zhou L, Lu X, Gao Q. Functional evaluation of a novel vitreous substitute using polyethylene glycol sols injected into a foldable capsular vitreous body. J Biomed Mater Res A. 2013;101:2538–47.
  • Wang T, Huang X, Gao Q, Feng L, Xie Z, Jiang Z, Liu Y, Li Y, Lin X, Lin J. A preliminary study to treat severe endophthalmitis via a foldable capsular vitreous body with sustained levofloxacin release in rabbits. Invest Ophthalmol Vis Sci. 2013;54:804–12.
  • Jiang Z, Wang T, Pan B, Xie Z, Wang P, Liu Y, Gao Q. Evaluation of the levofloxacin release characters from a rabbit foldable capsular vitreous body. Int J Nanomedicine. 2012;7:1–10.
  • Jiang Z, Wang P, Pan B, Xie Z, Li D, Wang T, Liu Y, Yuan Z, Gao Q. Evaluation of levofloxacin release characteristics from a human foldable capsular vitreous body in vitro. J Ocul Pharmacol Ther. 2012;28:33–40.
  • Lin X, Wang Z, Jiang Z, Long C, Liu Y, Wang P, Jin C, Yi C, Gao Q. Preliminary efficacy and safety of a silicone oil-filled foldable capsular vitreous body in the treatment of severe retinal detachment. Retina. 2012;32:729–41.
  • Wang P, Gao Q, Lin X, Zhang S, Hu J, Liu Y, Xu N, Ge J. Comprehensive analysis of inflammatory immune mediators of the intraocular fluid aspirated from the foldable capsular vitreous body filled-eyes. PloS One. 2012;7:e46384.
  • Wang P, Gao Q, Jiang Z, Lin J, Liu Y, Chen J, Zhou L, Li H, Yang Q, Wang T. Biocompatibility and retinal support of a foldable capsular vitreous body injected with saline or silicone oil implanted in rabbit eyes. Clin Exp Ophthalmol. 2012;40:e67–75.
  • Zheng H, Wang Z, Wang P, Liu Y, Jiang Z, Gao Q. Evaluation of 5-fluorouracil released from a foldable capsular vitreous body in vitro and in vivo. Graefes Arch Clin Exp Ophthalmol. 2012;250:751–59.
  • Chen X, Liu Y, Jiang Z, Zhou L, Ge J, Gao Q. Protein kinase cα downregulation via siRNA-PKCα released from foldable capsular vitreous body in cultured human retinal pigment epithelium cells. Int J Nanomedicine. 2011;6:1303–11.
  • Chen J, Gao Q, Liu Y, Ge J, Cao X, Luo Y, Huang D, Zhou G, Lin S, Lin J, et al. Clinical device-related article evaluation of morphology and functions of a foldable capsular vitreous body in the rabbit eye. J Biomed Mater Res B Appl Biomater. 2011;97(2):396–404.
  • Lin X, Ge J, Gao Q, Wang Z, Long C, He L, Liu Y, Jiang Z. Evaluation of the flexibility, efficacy, and safety of a foldable capsular vitreous body in the treatment of severe retinal detachment. Invest Ophthalmol Vis Sci. 2011;52:374–81.
  • Zhang R, Wang T, Xie C, Lin X, Jiang Z, Wang Z, Liu Y, Luo Y, Long C, He L, et al. Evaluation of supporting role of a foldable capsular vitreous body with magnetic resonance imaging in the treatment of severe retinal detachment in human eyes. Eye. 2011;25:794–802.
  • Liu Y, Ke Q, Chen J, Wang Z, Xie Z, Jiang Z, Ge J, Gao Q. Sustained mechanical release of dexamethasone sodium phosphate from a foldable capsular vitreous body. Invest Ophthalmol Vis Sci. 2010;51:1636–42.
  • Liu Y, Jiang Z, Gao Q, Ge J, Chen J, Cao X, Shen Q, Ma P. Technical standards of a foldable capsular vitreous body in terms of mechanical, optical, and biocompatible properties. Artif Organs. 2010;34:836–45.
  • Gao Q, Mou S, Ge J, To CH, Hui Y, Liu A, Wang Z, Long C, Tan J. A new strategy to replace the natural vitreous by a novel capsular artificial vitreous body with pressure-control valve. Eye. 2008;22:461–68.
  • Chan IM, Tolentino FI, Refojo MF, Fournier G, Albert DM. Vitreous substitute. Experimental Studies and Review. Retina. 1984;4:51–59.
  • Chirila TV, Tahija S, Hong Y, Vijayasekaran S, Constable IJ. Synthetic polymers as materials for artificial vitreous body: review and recent advances. J Biomater Appl. 1994;9:121–37.
  • Chirila TV, Hong YE, Dalton PD, Constable IJ, Refojo MF. The use of hydrophilic polymers as artificial vitreous. Prog Polym Sci. 1998;23:475–508.
  • Soman N, Banerjee R. Artificial vitreous replacements. Biomed Mater Eng. 2003;13:59–74.
  • Swindle K, Ravi N. Recent advances in polymeric vitreous substitutes. Expert Rev Ophthalmol. 2007;2:255–66.
  • Baino F. The use of polymers in the treatment of retinal detachment: current trends and future perspectives. Polymers. 2010;2:286–322.
  • Baino F. Towards an ideal biomaterial for vitreous replacement: historical overview and future trends. Acta Biomater. 2011 Mar;7(3):921–35.
  • Donati S, Caprani SM, Airaghi G, Vinciguerra R, Bartalena L, Testa F, Mariotti C, Porta G, Simonelli F, Azzolini C. Vitreous substitutes: the present and the future. Biomed Res Int. 2014;2014:351804.
  • Gao QY, Fu Y, Hui YN. Vitreous substitutes: challenges and directions. Int J Ophthalmol. 2015;8:437–40.
  • Alovisi C, Panico C, de Sanctis U, Eandi CM. Vitreous substitutes: old and new materials in vitreoretinal surgery. J Ophthalmol. 2017;2017:3172138.
  • Cai P, Chen X. Hydrogels for artificial vitreous: from prolonged substitution to elicited regeneration. ACS Materials Lett. 2019;1:285–89.
  • Flohé L, editor. Glutathione. Proceedings of the 16th conference of the german society of biological chemistry; March 1973; Tübingen, Germany: Biochem Soc Trans; 1974.
  • Michael R, Bron AJ. The ageing lens and cataract: a model of normal and pathological ageing. Philos Trans R Soc Lond B Biol Sci. 2011;366:1278–92.
  • Whitson JA, Sell DR, Goodman MC, Monnier VM, Fan X. Evidence of dual mechanisms of glutathione uptake in the rodent lens: a novel role for vitreous humor in lens glutathione homeostasis. Invest Ophthalmol Vis Sci. 2016;57:3914–25.
  • Tram NK, McLean RM, Swindle-Reilly KE. Glutathione improves the antioxidant activity of vitamin C in human lens and retinal epithelial cells: implications for vitreous substitutes. Curr Eye Res. 2020. Epub ahead of print. PMID: 32838548.
  • Chockalingam S, Roth C, Henzel T, Cohen T. Probing local nonlinear viscoelastic properties in soft materials. arXiv. 2020;2007:11090.
  • Scarcelli G, Seok HY. In vivo brillouin optical microscopy of the human eye. Opt Express. 2012;20:9197–202.
  • Reiß S, Burau G, Stachs O, Guthoff R, Stolz H. Spatially resolved brillouin spectroscopy to determine the rheological properties of the eye lens. Biomed Opt Express. 2011;2:2144–59.
  • Aloy MA, Adsuara JE, Cerdá-Durán P, Obergaulinger M, Esteve-Taboada JJ, Ferrer-Blasco T, Montés-Micó R. Estimation of the mechanical properties of the eye through the study of its vibrational modes. PloS One. 2017;12:e0183892.
  • High-frequency harmonic imaging of the eye.; 2005.
  • Santoro M, Marchetti P, Rossi F, Perale G, Castiglione F, Mele A, Masi M. Smart approach to evaluate drug diffusivity in injectable agar− carbomer hydrogels for drug delivery. J Phys Chem B. 2011;115:2503–10.
  • Tong J, Anderson JL. Partitioning and diffusion of proteins and linear polymers in polyacrylamide gels. Biophys J. 1996;70:1505–13.
  • Spitzer MS, Januschowski K. Aging and age-related changes of the vitreous body. Ophthalmologe. 2015;112:552–54.
  • Koo H, Moon H, Han H, Na JH, Huh MS, Park JH, Woo SJ, Park KH, Kwon IC, Kim K, et al. The movement of self-assembled amphiphilic polymeric nanoparticles in the vitreous and retina after intravitreal injection. Biomaterials. 2012;33(12):3485–93.
  • Stocchino A, Repetto R, Siggers JH. Mixing processes in the vitreous chamber induced by eye rotations. Phys Med Biol. 2009;55:453–67.
  • Repetto R, Siggers JH, Stocchino A. Mathematical model of flow in the vitreous humor induced by saccadic eye rotations: effect of geometry. Biomech Model Mechanobiol. 2010;9:65–76.
  • Balachandran RK, Barocas VH. Contribution of saccadic motion to intravitreal drug transport: theoretical analysis. Pharm Res. 2011;28:1049–64.
  • Abouali O, Modareszadeh A, Ghaffariyeh A, Tu J. Numerical simulation of the fluid dynamics in vitreous cavity due to saccadic eye movement. Med Eng Phys. 2012;34:681–92.
  • Modareszadeh A, Abouali O, Ghaffarieh A, Ahmadi G. Saccade movements effect on the intravitreal drug delivery in vitreous substitutes: a numerical study. Biomech Model Mechanobiol. 2013 Apr;12(2):281–90.
  • Bonfiglio A, Repetto R, Siggers JH, Stocchino A. Investigation of the motion of a viscous fluid in the vitreous cavity induced by eye rotations and implications for drug delivery. Phys Med Biol. 2013;58:1969–82.
  • Isakova K, Pralits JO, Repetto R, Romano MR. Mechanical models of the dynamics of vitreous substitutes. Biomed Res Int. 2014;2014:672926.
  • Awwad S, Lockwood A, Brocchini S, Khaw PT. The PK-eye: a novel in vitro ocular flow model for use in preclinical drug development. J Pharm Sci. 2015;104:3330–42.
  • Murali K, Kang D, Nazari H. Spatial variations in vitreous oxygen consumption. PLoS One. 2016;11:e0149961.
  • Wu HTD, Howse LA, Vaghefi E. Effect of age-related human lens sutures growth on its fluid dynamics. Invest Ophthalmol Vis Sci. 2017;58:6351–57.
  • Awwad S, Mohamed Ahmed AH, Sharma G, Heng JS, Khaw PT, Brocchini S, Lockwood A. Principles of pharmacology in the eye. Br J Pharmacol. 2017;174:4205–23.
  • Romano MR, Stocchino A, Ferrara M, Lagazzo A, Repetto R. Fluidics of single and double blade guillotine vitrectomy probes in balanced salt solution and artificial vitreous. Transl Vis Sci Technol. 2018;7:19.
  • Ferroni M, Cereda MG, Boschetti F. A combined approach for the analysis of ocular fluid dynamics in the presence of saccadic movements. Ann Biomed Eng. 2018;46:2091–101.
  • Henein C, Awwad S, Ibeanu N, Vlatakis S, Brocchini S, Tee Khaw P, Bouremel Y. Hydrodynamics of intravitreal injections into liquid vitreous substitutes. Pharmaceutics. 2019;11:371.
  • Stocchino A, Nepita I, Repetto R, Dodero A, Castellano M, Ferrara M, Romano MR. Fluid dynamic assessment of hypersonic and guillotine vitrectomy probes in viscoelastic vitreous substitutes. Transl Vis Sci Technol. 2020;9:9.
  • Bayat J, Emdad H, Abouali O. 3D numerical investigation of the fluid mechanics in a partially liquefied vitreous humor due to saccadic eye movement. Comput Biol Med. 2020;125:103955.
  • Awwad S, Henein C, Ibeanu N, Khaw PT, Brocchini S. Preclinical challenges for developing long acting intravitreal medicines. Eur J Pharm Biopharm. 2020;S0939–6411:30132–36.
  • Creveling CJ, Colter J, Coats B. Changes in vitreoretinal adhesion with age and region in human and sheep eyes. Front Bioeng Biotechnol. 2018;6:153.
  • Santra M, Sharma M, Katoch D, Jain S, Saikia UN, Dogra MR, Luthra-Guptasarma M. Induction of posterior vitreous detachment (PVD) by non-enzymatic reagents targeting vitreous collagen liquefaction as well as vitreoretinal adhesion. Sci Rep. 2020;10:1–14.
  • Di Michele F, Tatone A, Romano MR, Repetto R. A mechanical model of posterior vitreous detachment and generation of vitreoretinal tractions. Biomech Model Mechanobiol. 2020;1–15. Epub ahead of print. PMID: 32642790
  • Beebe DC, Holekamp NM, Shui YB. Oxidative damage and the prevention of age-related cataracts. Ophthalmic Res. 2010;44:155–65.
  • Kandarakis SA, Piperi C, Topouzis F, Papavassiliou AG. Emerging role of advanced glycation-end products (AGEs) in the pathobiology of eye diseases. Prog Retin Eye Res. 2014;42:85–102.
  • Stitt AW, Moore JE, Sharkey JA, Murphy G, Simpson DA, Bucala R, Vlassara H, Archer DB. Advanced glycation end products in vitreous: structural and functional implications for diabetic vitreopathy. Invest Ophthalmol Vis Sci. 1998;39:2517–23.
  • Fokkens BT, Mulder DJ, Schalkwijk CG, Scheijen JL, Smit AJ, Los LI. Vitreous advanced glycation endproducts and α-dicarbonyls in retinal detachment patients with type 2 diabetes mellitus and non-diabetic controls. PloS One. 2017;12:e0173379.
  • Lee OT, Good SD, Lamy R, Kudisch M, Stewart JM. Advanced glycation end-product accumulation reduces vitreous permeability. Invest Ophthalmol Vis Sci. 2015;56:2892–97.