Vascular endothelial growth factor in eye disease

https://doi.org/10.1016/j.preteyeres.2008.05.001Get rights and content

Abstract

Collectively, angiogenic ocular conditions represent the leading cause of irreversible vision loss in developed countries. In the US, for example, retinopathy of prematurity, diabetic retinopathy and age-related macular degeneration are the principal causes of blindness in the infant, working age and elderly populations, respectively. Evidence suggests that vascular endothelial growth factor (VEGF), a 40 kDa dimeric glycoprotein, promotes angiogenesis in each of these conditions, making it a highly significant therapeutic target. However, VEGF is pleiotropic, affecting a broad spectrum of endothelial, neuronal and glial behaviors, and confounding the validity of anti-VEGF strategies, particularly under chronic disease conditions. In fact, among other functions VEGF can influence cell proliferation, cell migration, proteolysis, cell survival and vessel permeability in a wide variety of biological contexts. This article will describe the roles played by VEGF in the pathogenesis of retinopathy of prematurity, diabetic retinopathy and age-related macular degeneration. The potential disadvantages of inhibiting VEGF will be discussed, as will the rationales for targeting other VEGF-related modulators of angiogenesis.

Introduction

Vascular endothelial growth factor (VEGF), a dimeric glycoprotein of approximately 40 kDa, is a potent, endothelial cell mitogen that stimulates proliferation, migration and tube formation leading to angiogenic growth of new blood vessels. It is essential for angiogenesis during development; the deletion of a single allele arrests angiogenesis and causes embryonic lethality (Ferrara et al., 1996). In mammals, the VEGF family consists of seven members: VEGF-A (typically, and hereafter, referred to as VEGF), VEGF-B, VEGF-C, VEGF-D, VEGF-E, VEGF-F and PlGF (placental growth factor) (Fig. 1A). Alternative splicing results in several VEGF variants. In humans, these include the relatively abundant VEGF121, VEGF165, VEGF189 and VEGF206, and several less-abundant forms (Fig. 1B). The solubility of these splice variants (collectively referred to as VEGFXXX) is dependent on heparin-binding affinity. VEGF206 and VEGF189 bind very tightly to heparin and, thus, remain sequestered in the extracellular matrix. VEGF165 binds heparin with less affinity, but also can be associated with the matrix, and VEGF121 lacks heparin-binding capacity, rendering it highly soluble. Investigations in mice genetically engineered to express less than the full complement of splice variants confirm that the relative solubility of VEGF splice variants strongly affects their specific bioactivities (Takahashi and Shibuya, 2005). Moreover, plasmin and various metalloproteinases can cleave VEGF165, resulting in an N-terminal 113-amino acid peptide that is non-heparin binding, but retains its bioactivity (Keyt et al., 1996; Ferrara et al., 2003). Our understanding of the relative expression of the different VEGF isoforms under normal or pathological conditions and the molecular regulators of VEGF alternative splicing is relatively limited.

Recently, discovery of the so called “VEGFXXXb isoforms” has sparked new interest in the molecular events that regulate VEGF expression (for review, see Ladomery et al., 2007). The VEGFXXXb isoforms share approximately 94–98% homology with the corresponding VEGFXXX isoforms and have the same length. However, due to alterations in the C-terminus they bind to VEGF receptors, but do not fully activate them and act as “dominant negative splice variants” (Bates et al., 2002). Studies showing that VEGF165b is downregulated in angiogenic tissues (Ladomery et al., 2007) suggest a primary role for this isoform in controlling VEGF activity in health and disease. Administration of VEGF165b has been shown to inhibit retinal angiogenesis in the mouse model of oxygen-induced retinopathy (Konopatskaya et al., 2006).

VEGFR-1/Flt-1 (fms-like tyrosine kinase) and VEGFR-2/KDR/Flk-1 (kinase insert domain-containing receptor/fetal liver kinase), along with structurally related receptors, Flt-3/Flk-2 and VEGFR-3/Flt-4, belong to the receptor tyrosine kinase family (Fig. 1A) (Hanks and Quinn, 1991; Blume-Jensen and Hunter, 2001). VEGFR-1 and -2 are primarily involved in angiogenesis (Yancopoulos et al., 2000) whereas Flt-3 and Flt-4 are involved in hematopoiesis and lymphangiogenesis (Jussila and Alitalo, 2000). The VEGFRs contain an approximately 750-amino-acid-residue extracellular domain, which is organized into seven immunoglobulin-like folds. Adjacent to the extracellular domain is a single transmembrane region, followed by a juxtamembrane domain, a split tyrosine kinase domain that is interrupted by a 70-amino-acid kinase insert and a C-terminal tail.

VEGF receptor activation requires dimerization. Guided by the binding properties of the ligands, VEGFRs form both homodimers and heterodimers (Rahimi, 2006). The signal transduction properties of the VEGFR heterodimers, compared with homodimers, remain to be fully elucidated. Dimerization of VEGFR is accompanied by activation of receptor kinase activity, leading to autophosphorylation. Site-directed mutagenesis studies have demonstrated that the Tyr1214 residue, located in the carboxy terminus of VEGFR-2, is required for the ligand-dependent autophosphorylation of the receptor and its ability to activate signaling proteins. Signal transduction is propagated when activated VEGF receptors phosphorylate SH2 domain-containing protein substrates.

In addition to VEGFRs, VEGF serves as a ligand to another family of receptors, the neuropilins. Neuropilins are 120–130-kDa non-tyrosine kinase receptors that mediate critical functions in tumor cells and the nervous and vascular systems. In endothelial cells, neuropilins serve as receptors for the class 3 semaphorins and co-receptors for VEGF family members. The role of Neuropilin-1 (NRP-1) in mediating VEGF activity is now being elucidated. VEGF signaling through NRP-1 stimulates endothelial cell migration and adhesion. The addition of an anti-NRP-1 antibody suppressed the mitogenic effects of VEGF165 on bovine retinal endothelial cells (RECs) (Oh et al., 2002). In another in vitro model, the VEGF-dependent differentiation of a subset of human bone marrow-derived cells into vascular precursors, and subsequent proliferation of these cells, required the activation of a VEGFR-2/NRP-1-dependent signaling pathway (Fons et al., 2004). Finally, VEGF promotion of the synthesis and release of prostacyclin (PGI2), an important mediator of angiogenesis, is thought to be mediated via NRP-1 binding (Neagoe et al., 2005). The angiogenic effects regulated through VEGF binding to NRP-2 are less well characterized and appear to be modulated differently from the effects controlled by NRP-1. For example, VEGF selectively upregulates NRP-1, but not NRP-2, on endothelial cells (Oh et al., 2002).

BIAcore analysis has shown NRP-1 to interact with VEGFR-1, greatly reducing its binding affinity for VEGF165 (Fuh et al., 2000). Co-culture systems of endothelial cells and breast carcinoma cells indicate that NRP-1 significantly enhances VEGF165 binding to VEGFR-2 (Soker et al., 2002). In aortic endothelial cells, NRP-2 interacted with VEGFR-1, but less is known at present about how this influences VEGF bioactivity (Gluzman-Poltorak et al., 2001). Finally, using multiple in vitro systems, NRP-2 was shown to interact with VEGFR-3, leading to lymphangiogenic activity, but no interaction was seen between NRP-2 and VEGFR-2 (Karpanen et al., 2006).

Few SH2 domain-containing proteins have been shown to interact directly with VEGFR-2. Phospholipase C-γ (PLCγ binds to phosphorylated Tyr1175 (Tyr1173 in the mouse), and mediates the activation of the mitogen-activated protein kinase (MAPK) cascade, leading to proliferation of endothelial cells (Takahashi et al., 2001) (Fig. 2). PLCγ activates protein kinase C via the production of diacylglycerol and increased concentrations of intracellular calcium. A Tyr1173Phe mutation of VEGFR-2 causes embryonic lethality due to vascular defects, mimicking the defects of VEGFR-2−/− mice (Sakurai et al., 2005). These data demonstrate an essential function of the Tyr1173 residue during vascular development.

In addition to PLCγ, the adaptor molecule, Shb, also binds to phosphorylated Tyr1175. VEGF-induced migration and PI3K activation are inhibited by small interfering RNA (siRNA)-mediated knockdown of Shb in endothelial cells (Holmqvist et al., 2004). The serine/threonine kinase, Akt, is activated downstream of PI3K and mediates endothelial cell survival (Fujio and Walsh, 1999). Akt also regulates nitric oxide (NO) production by direct phosphorylation and activation of endothelial NO synthase (eNOS). Finally, phosphorylated Tyr1175 is known to interact with Sck (Igarashi et al., 1998; Sakai et al., 2000), an adaptor molecule that binds Grb2, and participates in MAPK signaling in the epidermal growth factor pathway (Thelemann et al., 2005).

Another important phosphorylation site in VEGFR-2 is Tyr951 (Tyr949 in the mouse), a binding site for the signaling adaptor VEGF receptor-associated protein (VRAP) (Matsumoto et al., 2005). The phosphorylated Tyr951–VRAP pathway has been shown to regulate endothelial cell migration (Matsumoto et al., 2005; Zeng et al., 2001). Reduced microvessel density and tumor growth in VRAP−/− mice confirm an essential function for this residue in endothelial cells of the angiogenic phenotype (Matsumoto et al., 2005). VEGF induces the formation of a complex between VRAP and Src (Matsumoto et al., 2005), indicating that VRAP might regulate Src activation and its signaling downstream of VEGFR-2.

Mice that express a Tyr1212Phe (corresponding to the human Tyr1214) VEGFR-2 mutant are viable and fertile (Sakurai et al., 2005). However, phosphorylation of Tyr1212/1214 has been implicated in VEGF-induced actin remodeling through the sequential activation of CDC42 and p38 MAPK (Lamalice et al., 2004). Inhibition of p38 MAPK augments VEGF-induced angiogenesis in the chicken chorioallantoic membrane (CAM) assay (Issbrucker et al., 2003; Matsumoto, T., et al., 2002), without an accompanying increase in vascular permeability (Issbrucker et al., 2003). Moreover, p38 MAPK induces phosphorylation of heat-shock protein-27 (HSP27), a molecular chaperone that positively regulates VEGF-induced actin reorganization and migration (McMullen et al., 2005; Rousseau et al., 1997).

The existence of multiple ligands and receptors provides initial diversity to VEGF bioactivity. Groupings of receptor homo- and heterodimers, activated by both common and specific ligands, further augment the diversity of VEGF signaling. A final level of diversity is provided by activation of distinct signaling intermediates downstream of each VEGF receptor. The combination of these features yields an elaborate signaling network capable of regulating the extremely complex angiogenic cascade.

During hypoxia, VEGF gene expression increases via several different mechanisms (Dor et al., 2001). These mechanisms include increased transcription, mRNA stability, and protein translation using an internal ribosomal entry site (IRES), as well as increased expression of oxygen-regulated protein 150 (ORP 150), a chaperone required for intracellular transport of proteins from the endoplasmic reticulum to the Golgi apparatus prior to secretion (Chen and Shyu, 1995; Forsythe et al., 1996; Levy and Levy et al., 1996, Levy and Chung et al., 1998; Ozawa et al., 2001).

The increase in VEGF transcription is largely mediated via hypoxia inducible factor-1 (HIF-1) (Fig. 3). HIF-1 is a heterodimeric transcription factor composed of two subunits—the constitutively produced HIF-1β subunit and the inducible component, HIF-1α (Wang and Semenza, 1995). Under normoxic conditions, HIF-1α is inactivated and targeted for proteosomal degradation by hydroxylation, whereas under hypoxic conditions the specific hydroxylases are inhibited, resulting in the rescue of HIF-1α from degradation (Schofield and Ratcliffe, 2004). When this occurs, HIF-1α complexes with HIF-1β, translocates into the nucleus and binds to a specific sequence in the 5′ flanking region of the VEGF gene, the hypoxia responsive element (HRE) (Ikeda et al., 1995; Laughner et al., 2001; Shima et al., 1996; Wenger, 2002; Forsythe et al., 1996). The importance of interaction between HIF-1α and the VEGF promoter has been confirmed in studies of HIF-1α−/− mouse embryonic stem cells, in which basal expression of VEGF mRNA remains low in response to hypoxia (Carmeliet et al., 1998; Iyer et al., 1998).

Two additional isoforms of HIF, known as HIF-2α and HIF-3α, have been identified by screening for proteins that associate with HIF-1β (Ratcliffe, 2007). HIF-2α appears to be closely related to HIF-1α and can promote HRE-dependent gene transcription. While structurally and functionally similar, HIF-1α and HIF-2α appear to exert different biological functions, as demonstrated in studies using knockout mice (Hu et al., 2003). For example, while HIF-1α antagonizes c-Myc function, inhibiting renal cell carcinoma (RCC) growth, HIF-2α promotes cell cycle progression in hypoxic RCC and many other cell lines (Gordan et al., 2007). Interestingly, the most distantly related isoform, HIF-3α, appears to antagonize HRE-dependent gene expression, suggesting a possible negative influence on hypoxia-induced gene expression. Additional study is needed to determine if HIF-2α or HIF-3α is involved in the regulation of retinal VEGF expression.

Clearly, post-transcriptional events are also important in the regulation of VEGF production in the diseased retina, as underscored by the correlation of polymorphisms within the 5′-untranslated region (UTR) of the VEGF gene with the occurrence of age-related macular degeneration (AMD). The 3′UTR and the 5′UTR of the VEGF gene are important sites of regulation controlling mRNA stability and the rate of translation through the IRES (for review, see Yoo et al., 2006).

Evidence for the increase in VEGF mRNA stability in response to hypoxia comes from in vitro mRNA degradation assays that have led to the identification of adenylate/uridylate-rich elements (AREs) in the 3′ UTR of VEGF mRNA. VEGF mRNA is extremely labile in normoxic conditions, with a half-life of less than 1 h, as compared with the average half-life of 10–12 h for eukaryotic mRNA. During hypoxia, the half-life of VEGF mRNA increases by 2–3-fold (Levy et al., 1996) due to a stabilizing effect of HuR, a 36 kDa RNA-binding protein, which binds with high affinity to AREs in the 3′-UTR of VEGF mRNA, protecting it from degradation by endonucleases (Brennan and Steitz, 2001; Robinow et al., 1988).

Post-transcriptional regulation can also occur in the 5′-UTR of VEGF mRNA. This region contains multiple IRES. These are specific sites of attachment to the ribosomal machinery, which provide sites for initiation of translation alternative to the classical 5′ cap- and elF-dependent translational system (for review, see van der Velden and Thomas, 1999). Several IRES have been identified in the 5′-UTR of VEGF mRNA, and these provide alternative sites of translational control of VEGF expression (Bornes et al., 2004). Notably, evidence suggests that IRES sites in the VEGF 5′-UTR can potentially control the generation of alternatively spliced VEGF (Bornes et al., 2004; Huez et al., 2001).

Another regulatory mechanism consists of increased production of ORP150 in response to hypoxia. Studies using human macrophages transfected with adenovirus coding for ORP150 showed that overexpression of ORP150 resulted in increased VEGF secretion in hypoxia. Evidence suggests that under hypoxic conditions, ORP150 functions as a molecular chaperone to facilitate VEGF protein transport and secretion (Ozawa et al., 2001).

VEGF is not only regulated by hypoxia. VEGF function is also affected by insulin-like growth factor 1 (IGF-1) which plays an important role in retinal vascularization. Several lines of evidence, including in vitro studies, support the notion that IGF-1 is critical for vessel development (King et al., 1985; Grant et al., 1993). Preterm infants with reduced serum levels of IGF-1 have a higher incidence of development of retinopathy (Hellstrom et al., 2003). Mice null for the IGF-1 gene have retarded retinal vascular growth compared to wild-type controls (Hellstrom et al., 2001). However, the action of IGF-1 is not mediated by decreasing VEGF expression, as the amount of VEGF mRNA is similar in knockout and wild-type control mice; instead IGF-1 acts by decreasing VEGF activation of the Akt signaling pathway. Both MAPK and Akt pathways have been shown to be necessary for endothelial cell survival (Smith et al., 1999).

The influence of VEGF in retinal diseases is profound. It has been implicated in a large number of retinal diseases and conditions including, but not limited to, highly prevalent conditions like age-related macular degeneration (AMD) and diabetic retinopathy; less common disorders such as retinopathy of prematurity, sickle cell retinopathy and retinal vascular occlusion; and as a non-causal, but important, secondary influence in neovascular glaucoma (Bock et al., 2007) and inherited retinal dystrophies (Penn et al., 2000). Collectively, these conditions, all of which have critical angiogenic components, account for the vast majority of irreversible vision loss in developed countries.

At least five retinal cell types have the capacity to produce and secrete VEGF. These include the retinal pigmented epithelium (RPE) (Miller et al., 1997), astrocytes (Stone et al., 1995), Müller cells (Robbins et al., 1997), vascular endothelium (Aiello et al., 1995a) and ganglion cells (Ida et al., 2003). These cells differ widely in their responses to hypoxia; in vitro studies show that Müller cells and astrocytes generally produce the greatest amounts of VEGF under hypoxic conditions (Morrison et al., 2007; Aiello et al., 1995a; Hata et al., 1995). To date, the relative capacity of these cells to produce specific splice variants remains unclear, as do the patterns of splice variant production throughout retinal development and aging.

The distinct roles of the different VEGF splice variants in retinal vascular development is being explored, however, in mice expressing only a single variant (Stalmans et al., 2002). Vascular development was normal in the retinas of mice expressing only VEGF164 (VEGF164/164), indicating that this variant is sufficient for directing normal vascular growth and remodeling. In contrast, retinas of VEGF120/120 mice exhibited severe vascular defects, displaying retarded venous and severely flawed arterial development. VEGF188/188 mice had normal development of retinal veins but little or no arterial growth.

Evidence for the expression patterns and roles of VEGFR in retinal tissues comes from a variety of species and experimental venues. In the human retina VEGFR-1 and -2 can be expressed by neural, glial and vascular elements. In adults, expression is generally restricted to the inner nuclear layer (Müller cells and amacrine cells), the ganglion cell layer, and the retinal vasculature (Stitt et al., 1998). However, during retinal neurogenesis VEGFR-2 is also expressed by neural progenitor cells (Hashimoto et al., 2006). Notably, neural cell VEGFR-2 can be activated by VEGF in vitro (Yang and Cepko, 1996). In cultured retinal pericytes VEGFR-1, but not -2, is expressed (Takagi et al., 1996), whereas in cultured RPE cells, both receptors are expressed and are induced by oxidative stress (Sreekumar et al., 2006). In the mouse, ganglion cells express both receptors, but only VEGFR-2 is increased by intraocular inoculation with herpesvirus (Vinores et al., 2001). Studies in newborn mice using the VEGFR-specific kinase inhibitor, SU5416, indicate that Müller cell survival or proliferation during retinal development is VEGFR- and MAPK-dependent (Robinson et al., 2001). In a study of patients with diabetic retinopathy, VEGFR-1 expression dominated in normal retina, but was not increased in the diabetic retina, while VEGFR-2 levels were increased, particularly in the vascular elements (Smith et al., 1999). Finally, VEGFR-1 and -2 are found on uterine smooth muscle cells in vivo. When these cells are cultured in vitro, VEGFR-1 can be phosphorylated and is capable of inducing smooth muscle cell proliferation (Brown et al., 1997). To date, neither VEGFR-1 nor -2 has been identified in retinal smooth muscle cells.

This article will review the role of VEGF in angiogenesis related to three blinding conditions: retinopathy of prematurity, diabetic retinopathy and age-related macular degeneration. These conditions constitute the leading causes of irreversible vision loss in infants, and working age and elderly Americans, respectively. That VEGF is believed to play a causal role in all three of these disorders underscores its profound impact in eye disease. VEGF antagonists have already proven their value in tumor angiogenesis and choroidal neovascularization, and new VEGF antagonists are being tested pre-clinically and clinically for other ocular indications. Only with a more complete understanding of VEGF and its retinal and choroidal activities can we hope to develop better strategies to prevent, retard or repair the damage caused by ocular neovascularization.

Section snippets

Retinopathy of prematurity

Retinopathy of prematurity (ROP), a neovascularizing disease affecting preterm infants, is one of the most common causes of childhood blindness in the world. Recent estimates indicate that each year in the United States, 68% of the approximately 10,000 babies born with a birth weight of less than 1250 g will develop ROP. Thirty-six percent of these infants will progress to severe ROP, a condition that can lead to retinal detachment and blindness. The incidence of the disease is highly correlated

Clinical features of diabetic retinopathy

Diabetic retinopathy is the most frequent complication of diabetes and the leading cause of blindness in developed countries worldwide. Approximately 75% of all diabetic patients show clinical signs of retinopathy within 15 years after onset of diabetes and more than 10% develop visual impairment within this period (Klein et al., 1984; Klein and Klein, 1995; Sjolie et al., 1997). In the United States, diabetic retinopathy accounts for 8% of legal blindness, making diabetes the leading cause of

AMD overview

Age-related macular degeneration (AMD) is one of the leading causes of vision loss worldwide (Resnikoff et al., 2004). The neovascular form comprises only 10% of AMD, but accounts for close to 90% of legal blindness from it. Recently, chemotherapeutic agents that inhibit actions of VEGF have been found to be effective in improving vision in approximately 40% of eyes with neovascular AMD, whereas previous modalities (i.e., photocoagulation, photodynamic therapy, radiation therapy, angiostatic

Future directions

Our initial experience with pegaptanib and ranibizumab has demonstrated the validity of targeting VEGF in the treatment of ocular angiogenesis. Currently, the use of anti-VEGF therapies to treat increased vascular permeability resulting in macular edema is also being investigated, as are other strategies for VEGF inhibition. RNA interference, for example, is a powerful strategy for silencing the expression of specific genes, and its application to VEGF inhibition for CNV is ongoing (Fattal and

Acknowledgments

The authors are indebted to Susan E. Yanni for her contributions to the editing and processing of this review. Its preparation was funded in part by EY007533, EY08126, EY07135 and support from RPB (JSP); EY04618, EY011766, VA Merit Review Award (RBC); EY015130, EY017011 (MEH).

References (547)

  • S. Bornes et al.

    Control of the vascular endothelial growth factor internal ribosome entry site (IRES) activity and translation initiation by alternatively spliced coding sequences

    J. Biol. Chem.

    (2004)
  • M.S. Brown et al.

    Association between higher cumulative doses of recombinant erythropoietin and risk for retinopathy of prematurity

    J. AAPOS.

    (2006)
  • J. Cai et al.

    Oxidative damage and protection of the RPE

    Prog. Retin. Eye Res.

    (2000)
  • Y. Cao et al.

    Placenta growth factor: identification and characterization of a novel isoform generated by RNA alternative splicing

    Biochem. Biophys. Res. Commun.

    (1997)
  • C.K. Chan et al.

    Differential expression of pro- and antiangiogenic factors in mouse strain-dependent hypoxia-induced retinal neovascularization

    Lab. Invest.

    (2005)
  • C.Y. Chen et al.

    AU-rich elements: characterization and importance in mRNA degradation

    Trends Biochem. Sci.

    (1995)
  • M. Clauss et al.

    The vascular endothelial growth factor receptor Flt-1 mediates biological activities. Implications for a functional role of placenta growth factor in monocyte activation and chemotaxis

    J. Biol. Chem.

    (1996)
  • S. Claxton et al.

    Role of arteries in oxygen induced vaso-obliteration

    Exp. Eye Res.

    (2003)
  • T.D. Cohen et al.

    Interleukin 6 induces the expression of vascular endothelial growth factor

    J. Biol. Chem.

    (1996)
  • S.E. Connolly et al.

    Characterization of vascular development in the mouse retina

    Microvasc. Res.

    (1988)
  • A.M. Abu El-Asrar et al.

    Inducible nitric oxide synthase and vascular endothelial growth factor are colocalized in the retinas of human subjects with diabetes

    Eye

    (2004)
  • A.P. Adamis et al.

    Inhibition of vascular endothelial growth factor prevents retinal ischemia-associated iris neovascularization in a nonhuman primate

    Arch. Ophthalmol.

    (1996)
  • 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)
  • A simplified severity scale for age-related macular degeneration: AREDS Report No. 18

    Arch. Ophthalmol.

    (2005)
  • L.P. Aiello et al.

    Vascular endothelial growth factor in ocular fluid of patients with diabetic retinopathy and other retinal disorders

    N. Engl. J. Med.

    (1994)
  • L.P. Aiello et al.

    Hypoxic regulation of vascular endothelial growth factor in retinal cells

    Arch. Ophthalmol.

    (1995)
  • L.P. Aiello et al.

    Suppression of retinal neovascularization in vivo by inhibition of vascular endothelial growth factor (VEGF) using soluble VEGF-receptor chimeric proteins

    Proc. Natl. Acad. Sci. USA

    (1995)
  • L.P. Aiello et al.

    Diabetic retinopathy

    Diabetes Care

    (1998)
  • L.P. Aiello et al.

    Effect of ruboxistaurin on visual loss in patients with diabetic retinopathy

    Ophthalmology

    (2006)
  • K. Alexandraki et al.

    Inflammatory process in type 2 diabetes: the role of cytokines

    Ann. N.Y. Acad. Sci.

    (2006)
  • T. Alon et al.

    Vascular endothelial growth factor acts as a survival factor for newly formed retinal vessels and has implications for retinopathy of prematurity

    Nat. Med.

    (1995)
  • M. Al-Shabrawey et al.

    Preservation of the Blood Retinal Barrier (BRB) in Diabetes by Deletion of gp91phox or Inhibition of NAD(P)H Oxidase

    Invest. Ophthalmol. Vis. Sci.

    (2006)
  • Al-Shabrawey, M., Bartoli, M., et al., 2008. Mechanisms of Statin's protective actions in diabetic retinopathy: role of...
  • J. Ambati et al.

    An animal model of age-related macular degeneration in senescent Ccl-2- or Ccr-2-deficient mice

    Nat. Med.

    (2003)
  • C.G. Anderson et al.

    Retinopathy of prematurity and pulse oximetry: a national survey of recent practices

    J. Perinatol.

    (2004)
  • D.A. Antonetti et al.

    Vascular permeability in experimental diabetes is associated with reduced endothelial occludin content: vascular endothelial growth factor decreases occludin in retinal endothelial cells. Penn State Retina Research Group

    Diabetes

    (1998)
  • D.A. Antonetti et al.

    Molecular mechanisms of vascular permeability in diabetic retinopathy

    Semin. Ophthalmol.

    (1999)
  • D.A. Antonetti et al.

    Diabetic retinopathy: seeing beyond glucose-induced microvascular disease

    Diabetes

    (2006)
  • T. Asahara et al.

    Isolation of putative progenitor endothelial cells for angiogenesis

    Science

    (1997)
  • N. Ashton

    Neovascularization in ocular disease

    Trans. Ophthalmol. Soc. UK

    (1961)
  • N. Ashton

    Donders lecture, 1967. Some aspects of the comparative pathology of oxygen toxicity in the retina

    Br. J. Ophthalmol.

    (1968)
  • N. Ashton

    Retinal angiogenesis in the human embryo

    Br. Med. Bull.

    (1970)
  • N. Ashton et al.

    Studies on developing retinal vessels. V. Mechanism of vaso-obliteration; a preliminary report

    Br. J. Ophthalmol.

    (1957)
  • L.M. Askie et al.

    Restricted versus liberal oxygen exposure for preventing morbidity and mortality in preterm or low birth weight infants

    Cochrane Database Syst. Rev. CD

    (2000)
  • R.L. Avery et al.

    Intravitreal bevacizumab (Avastin) in the treatment of proliferative diabetic retinopathy

    Ophthalmology

    (2006)
  • J. Baffi et al.

    Choroidal neovascularization in the rat induced by adenovirus mediated expression of vascular endothelial growth factor

    Invest. Ophthal. Vis. Sci.

    (2000)
  • E. Bancalari et al.

    Influence of transcutaneous oxygen monitoring on the incidence of retinopathy of prematurity

    Pediatrics

    (1987)
  • A.J. Barber et al.

    Neural apoptosis in the retina during experimental and human diabetes

    Early onset and effect of insulin. J. Clin. Invest.

    (1998)
  • A.J. Barber et al.

    Altered expression of retinal occludin and glial fibrillary acidic protein in experimental diabetes. The Penn State Retina Research Group

    Invest. Ophthalmol. Vis. Sci.

    (2000)
  • A.J. Barber et al.

    The Ins2Akita mouse as a model of early retinal complications in diabetes

    Invest. Ophthalmol. Vis. Sci.

    (2005)
  • Cited by (584)

    • Microneedles for advanced ocular drug delivery

      2023, Advanced Drug Delivery Reviews
    View all citing articles on Scopus
    View full text