throbber
Hindawi Publishing Corporation
`Journal of Ophthalmology
`Volume 2012, Article ID 483034, 13 pages
`doi:10.1155/2012/483034
`
`Review Article
`Development of Anti-VEGF Therapies for Intraocular Use:
`A Guide for Clinicians
`
`Pearse A. Keane1 and Srinivas R. Sadda2
`
`1 NIHR Biomedical Research Centre for Ophthalmology, Moorfields Eye Hospital NHS Foundation Trust and UCL Institute of
`Ophthalmology, London EC1V 2PD, UK
`2 Doheny Eye Institute and Department of Ophthalmology, Keck School of Medicine, University of Southern California,
`Los Angeles, CA 90033, USA
`
`Correspondence should be addressed to Srinivas R. Sadda, ssadda@doheny.org
`
`Received 8 September 2011; Accepted 1 November 2011
`
`Academic Editor: Toshiaki Kubota
`
`Copyright © 2012 P. A. Keane and S. R. Sadda. This is an open access article distributed under the Creative Commons Attribution
`License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
`cited.
`
`Angiogenesis is the process by which new blood vessels form from existing vessel networks. In the past three decades, significant
`progress has been made in our understanding of angiogenesis; progress driven in large part by the increasing realization that
`blood vessel growth can promote or facilitate disease. By the early 1990s, it had become clear that the recently discovered “vascular
`endothelial growth factor” (VEGF) was a powerful mediator of angiogenesis. As a result, several groups targeted this molecule
`as a potential mediator of retinal ischemia-induced neovascularization in disorders such as diabetic retinopathy and retinal vein
`occlusion. Around this time, it also became clear that increased intraocular VEGF production was not limited to ischemic retinal
`diseases but was also a feature of choroidal vascular diseases such as neovascular age-related macular degeneration (AMD). Thus,
`a new therapeutic era emerged, utilizing VEGF blockade for the management of chorioretinal diseases characterized by vascular
`hyperpermeability and/or neovascularization. In this review, we provide a guide for clinicians on the development of anti-VEGF
`therapies for intraocular use.
`
`1. Introduction
`
`In 1948, Isaac Michaelson proposed that a diffusible factor
`(named afterward “factor X”) could be responsible, not
`only for the development of the normal retinal vasculature
`but also for pathological neovascularization in prolifera-
`tive diabetic retinopathy and other ocular disorders [1].
`By the early 1990s, it had become clear that the recently
`discovered “vascular endothelial growth factor” (VEGF)
`possessed many of the requisite characteristics of a “factor
`X” [2]. As a result, several groups targeted this molecule
`as a potential mediator of retinal ischemia-induced neovas-
`cularization in disorders such as diabetic retinopathy and
`retinal vein occlusion (RVO) [3, 4]. Around this time, it also
`became clear that increased intraocular VEGF production
`was not limited to ischemic retinal diseases but was also a
`feature of choroidal vascular diseases such as neovascular
`age-related macular degeneration (AMD) [5, 6]. Thus, a
`new therapeutic era emerged, utilizing VEGF blockade for
`
`the management of chorioretinal diseases characterized by
`vascular hyperpermeability and/or neovascularization.
`In this review, we begin by providing an overview of
`angiogenesis, the manner in which VEGF was discovered to
`be central to this process, and then a summary of VEGF
`biology. In this manner, we aim to provide the clinician with
`an understanding of the clinical scenarios in which VEGF
`blockade is likely to be successful and of patient benefit. We
`continue by describing the development of four key anti-
`VEGF therapies (pegaptanib, bevacizumab, ranibizumab,
`and aflibercept) and the results of their application in a
`selection of pioneering clinical trials. By describing the main
`features of their development in a manner accessible to clin-
`icians, we aim to highlight those molecular characteristics,
`of each agent, with implications for clinical outcomes and
`patient safety. We conclude the review by describing likely
`future directions in the application of anti-VEGF therapy in
`chorioretinal disease.
`
`Novartis Exhibit 2025.001
`Regeneron v. Novartis, IPR2021-00816
`
`

`

`2
`
`2. Angiogenesis
`
`2.1. Overview. Angiogenesis is the process by which new
`blood vessels form from existing vessel networks (by com-
`parison, vasculogenesis is a form of de novo blood vessel
`formation that is typically seen in the embryo) [7–9].
`Angiogenesis begins with vasodilatation and increases in vas-
`cular permeability, followed by activation and proliferation
`of vascular endothelial cells; these changes are accompa-
`nied by degradation of the surrounding extracellular matrix
`(ECM), facilitating endothelial cell migration. The migrat-
`ing endothelial cells assemble, form cords, and ultimately
`acquire lumens; further differentiation to accommodate local
`requirements then occurs and a network of similarly dif-
`ferentiated periendothelial cells and matrix develops. After
`further remodeling a complex vascular network is ultimately
`formed.
`
`2.2. Role of Angiogenesis in Disease. In the past three decades,
`significant progress has been made in our understanding of
`angiogenesis: progress driven in large part by the increasing
`realization that blood vessel growth can promote or facilitate
`disease [10]. This major conceptual advance first occurred
`in the 1930s and 1940s, when it was hypothesized that
`induction of blood vessel growth through release of vasopro-
`liferative factors would confer a growth advantage on tumor
`cells [11]. Subsequently, in the 1970s, Folkman hypothesized
`that blockade of angiogenesis could be a strategy to treat
`cancer and other disorders [12]. However, adoption of such a
`strategy first required the identification and characterization
`of the mediators of angiogenesis—a major technological
`challenge at that point.
`
`2.3. Putative Regulators of Angiogenesis. In the subsequent
`years, advances in molecular biology led to the identifica-
`tion of many putative regulators of angiogenesis, with well-
`known examples including basic fibroblast growth factor
`transforming growth factor (TGF)-β, and the
`(bFGF),
`angiopoietins [7]. In the 1980s, bFGF was thought to be the
`major angiogenic factor in the pituitary and other organs.
`However, this model was called into question when, in
`1986, it became clear that bFGF lacks a peptide sequence
`necessary for secretion and is thus confined intracellularly
`(angiogenesis is a process that requires diffusion in an
`extracellular environment) [13].
`
`2.4. Discovery of VEGF. In the mid-1980s, Ferrara and Hen-
`zel cultured a population of nonhormone secreting follicular
`cells—with unusual characteristics—from bovine pituitary
`glands (follicular cells have cytoplasmic projections that
`establish intimate connections with perivascular spaces and
`were thought to have a role in regulating growth and main-
`tenance of pituitary vasculature) [14]. Ferrara discovered
`that culture medium conditioned by these cells strongly
`promoted endothelial cell growth. He hypothesized that this
`mitogenic activity may be the result of a secreted protein;
`the subsequent isolation and sequencing of this protein led
`to discovery of the most important mediator of angiogenesis
`currently known—VEGF [15].
`
`Journal of Ophthalmology
`
`2.5. Vascular Permeability Factor. Independently, in the early
`1980s, Senger et al. had reported the identification of a per-
`meability-enhancing protein (in the supernatant of a guinea
`pig tumor cell line), which they named “vascular permeabil-
`ity factor” (VPF) [16]. In 1989, at the same time Ferrara
`and coworkers were reported their discovery of VEGF. Keck
`et al. reported the isolation and sequencing of VPF [17].
`Surprisingly, their findings indicated that VEGF and VPF
`were, in fact, the same molecule.
`
`2.6. Clinical Role for VEGF Blockade. Although multiple
`growth factors other than VEGF have been implicated in
`the angiogenic process (e.g., bFGF), VEGF appears critical
`for a number of reasons: its production is driven by hyp-
`oxia; it is highly selective for endothelial cells, it possesses
`diffusion characteristics that allow it to reach its target,
`and it affects multiple aspects of the angiogenic process
`[18, 19]. VEGF also causes vascular dilatation and promotes
`vasopermeability, both of which facilitate a rich environment
`for the growth of new vessels. Thus, despite the complexity
`of the angiogenic process, and the potential redundancy of
`the growth factors involved, VEGF blockade was quickly
`recognized as a promising approach for the restriction of
`blood vessel formation in a variety of pathologic scenarios
`[8].
`
`3. VEGF Biology
`
`3.1. Gene Family. VEGF-A, first discovered in 1989 (see
`above), is the prototype member of a gene family (i.e.,
`a group of genes with shared sequences and with similar
`biochemical functions) that also includes placental growth
`factor (PLGF), VEGF-B, VEGF-C, VEGF-D, and VEGF-E
`(prior to the discovery of other family members, VEGF-A
`was known simply as VEGF; the terms are used interchange-
`ably in this review) [10, 18]. Of note, VEGF-C and VEGF-
`D are involved in the regulation of lymphatic angiogenesis
`[20], demonstrating the unique role of this gene family in
`controlling multiple structural components of the vascular
`system.
`
`3.2. Regulation of VEGF Gene Expression. Oxygen tension
`has a key role in regulating the production of VEGF. VEGF
`mRNA expression is induced by exposure to low oxygen
`tension under a variety of pathophysiological circumstances,
`and it is now well established that a transcription factor,
`hypoxia-inducible factor-1 (HIF-1), is a key mediator of this
`response [21, 22]. Recent studies have also shown that Von
`Hippel Lindau (VHL) protein, a product of the VHL tumor
`suppressor gene, provides negative regulation of VEGF and
`other hypoxia-inducible genes (inactivation of this gene leads
`to development of capillary hemangioblastomas in the retina
`and cerebellum, and in many cases, renal cell carcinomas)
`[23].
`Several major growth factors, such as epidermal growth
`factor, also upregulate VEGF mRNA expression, suggesting
`that paracrine or autocrine release of such factors works
`in concert with local hypoxia to increase production of
`VEGF [24, 25]. In addition, inflammatory cytokines, such as
`
`Novartis Exhibit 2025.002
`Regeneron v. Novartis, IPR2021-00816
`
`

`

`Journal of Ophthalmology
`
`3
`
`interleukin-1α and interleukin-6, induce expression of VEGF
`in several cell types (an observation in agreement with the
`hypothesis that VEGF plays a role in the angiogenesis and
`hyperpermeability seen in some inflammatory disorders)
`[25].
`
`3.3. VEGF Isoforms. The human VEGFA gene is organized as
`eight exons separated by seven introns (i.e., eight expressed
`regions that are joined together in the final mature RNA)
`[26]. Alternative splicing of the VEGFA gene results in
`the generation of four major isoforms (VEGF121, VEGF165,
`VEGF189, and VEGF206), having, respectively 121, 165, 189,
`and 206 amino acids. VEGF165 is the predominant isoform
`[27].
`Native VEGF is a heparin-binding glycoprotein (heparin
`is commonly used during protein purification due to its
`structural similarity to RNA and DNA), with a protein
`molecular weight of 45 kDa, the properties of which corre-
`spond closely to those of VEGF165 [27]. Loss of the heparin-
`binding domain of VEGF results in a significant loss in
`its mitogenic activity [28]. VEGF121, while freely diffusible
`in the ECM, is acidic and does not bind heparin [27].
`Conversely, VEGF189 and VEGF206, while being highly basic
`and capable of binding heparin with high affinity, are almost
`completely sequestered in the ECM. Thus, VEGF165, with
`intermediary properties, possesses the optimal characteris-
`tics of bioavailability and biological potency [27].
`
`3.4. VEGF Receptors. VEGF binds to two, related, receptor
`tyrosine kinases: VEGF Receptor 1 (VEGFR1) and VEGF
`Receptor 2 (VEGFR2) [27]. Both VEGFR1 and VEGFR2
`have seven immunoglobulin-like domains in the ECM, a
`single transmembrane region, and a tyrosine kinase sequence
`interrupted by a kinase-insert domain. In 1992, VEGFR1 was
`the first VEGF receptor discovered and was found to bind
`VEGF with high affinity [29]. However, despite its lower
`binding affinity for VEGF relative to VEGFR1, there is now
`agreement that VEGFR2 is the major mediator of the mito-
`genic, angiogenic, and permeability-enhancing effects of
`VEGF (the precise function of VEGFR1 is still under debate
`but may provide a “decoy effect” on VEGF signaling) [27]. In
`addition, VEGF interacts with a family of nonsignaling core-
`ceptors, the neuropilins—neuropilin-1 (NRP-1) appears to
`present VEGF165 to VEGFR2 in a configuration that increases
`the effectiveness of VEGFR2-mediated signal transduction
`[18, 30].
`
`[33]. Coverage by pericytes is thought to be one of the key
`events, resulting in loss of VEGF dependence [34].
`VEGF has also been shown to act as a chemotactic agent
`for bone marrow-derived monocytes [35], a pro-inflam-
`matory cytokine through upregulation of intercellular adhe-
`sion molecule-1 (ICAM-1) with consequent leukocyte adhe-
`sion [36], and a promoter of blood vessel extravasation
`through the upregulation of matrix metalloproteinases and
`decreased release of metalloproteinase inhibitors [37].
`The effects of VEGF on the promotion of vascular leak-
`age, both in inflammation and in other pathologic circum-
`stances, are also well established (prior to its isolation and
`sequencing, VEGF was initially characterized as “vascular
`permeability factor” by Senger et al. (see above)) [16]. Con-
`sistent with this role, VEGF has been shown to promote
`dissolution of tight junctions between endothelial cells and to
`induce endothelial fenestration in a number of vascular beds
`[38]. VEGF also induces vasodilatation in a dose-dependent
`fashion as a result of release of endothelial cell-derived nitric
`oxide—systemic blockade of VEGF may thus result in a
`clinically significant adverse hypertensive effect [39].
`Taken together, blockade of the biologic effects of VEGF
`results in rapid vessel remodeling with regression of pericyte-
`poor capillaries, reductions in vascular lumen diameter, and
`reductions in vascular permeability [33, 34]. More recently,
`evidence has suggested that VEGF could have additional
`neuroprotective effects [40].
`
`3.6. Role of VEGF in Ocular Disease. In 1994, Aiello et al.
`found a striking correlation between intraocular VEGF con-
`centrations and active proliferative retinopathy in patients
`with diabetes and ischemic central retinal vein occlusion
`(CRVO) [3]. Around the same time, Adamis et al. reported
`increased concentrations of VEGF in the vitreous of patients
`with diabetic retinopathy [4]. In 1996, it also became clear
`that increased intraocular levels of VEGF were not limited to
`ischemic retinal disorders: in a pair of influential studies, the
`localization of VEGF to choroidal neovascular membranes
`in patients with neovascular AMD was reported [5, 6].
`Proof-of-concept studies then demonstrated that blockade
`of VEGF, in animal models, led to marked decreases in
`retinal and iris neovascularization [41, 42]. Furthermore,
`exogenous administration of VEGF was demonstrated to
`produce retinal ischemia and vascular hyperpermeability in
`primates [43].
`
`4. Pegaptanib
`
`3.5. Activities of VEGF. Vascular endothelial cells are the
`primary targets for VEGF biologic activity, with their mito-
`genic effects well documented, both in vitro and in vivo [27].
`In particular VEGF induces a potent angiogenic effect in a
`variety of animal models in vivo [15, 31].
`VEGF also acts as a survival factor for endothelial cells
`in a variety of circumstances. Inhibition of VEGF results in
`extensive apoptotic changes in the vasculature of neonatal,
`but not adult mice [32]; furthermore, a marked VEGF de-
`pendence has been demonstrated in the endothelial cells of
`newly formed but not of established vessels within tumors
`
`Pegaptanib sodium is an RNA aptamer that binds to the
`heparin-binding domain of VEGF and, thus, prevents the
`predominant VEGF165 isoform from binding to VEGF recep-
`tors [44]. Pegaptanib was licensed to EyeTech Pharmaceuti-
`cals (now OSI Pharmaceuticals) for late stage development
`and marketing in the United States as “Macugen” (outside
`the USA, pegaptanib is marketed by Pfizer Inc.).
`
`4.1. Chemistry. Aptamers (from the Latin aptus, to fit, and
`the Greek meros, part or region) are oligonucleotides that
`bind to specific target molecules and that are usually created
`
`Novartis Exhibit 2025.003
`Regeneron v. Novartis, IPR2021-00816
`
`

`

`4
`
`Journal of Ophthalmology
`
`by selection from a large random sequence pool [45]. In
`this manner, aptamers are commonly used for basic research
`and clinical purposes as macromolecular drugs. Aptamers
`constitute one of four classes of oligonucleotide reagents,
`the others being antisense oligonucleotides, ribozymes, and
`small interfering RNAs (siRNAs) [44]. However, in contrast
`with these other entities, aptamers can act on extracellular
`targets and, therefore, are not required to cross cell mem-
`branes to exert their therapeutic effects.
`The selection of aptamers has become relatively straight-
`forward with the advent of “systematic evolution of ligands
`by exponential enrichment” (SELEX) [46]; in this process,
`aptamers are engineered to bind to various target molecules
`through repeated rounds of in vitro selection. Aptamers
`offer molecular recognition properties that rival that of
`antibodies, but with a number of advantages: (1) they can be
`engineered completely in vitro, (2) they are readily produced
`by chemical synthesis, (3) they possess desirable storage
`properties, and (4) they elicit little or no immunogenicity
`[44, 45]. Pegaptanib has the distinction of being the first
`aptamer therapeutic approved for use in humans [44].
`Having chosen VEGF165 as the target for selection of
`a prospective anti-VEGF aptamer, three separate iterations
`of the SELEX methodology were carried out by scientists
`at NeXstar Pharmaceuticals [44]. By 1998, three, stable,
`high-affinity anti-VEGF165 aptamers had been characterized,
`one of which was selected for development as pegaptanib
`(initially designated NX1838, and then, EYE001) (all three
`aptamers demonstrated little or no binding to VEGF121)
`[47].
`
`4.2. Preclinical Studies. The fact that pegaptanib offers se-
`lective inhibition of a single isoform offers the theoretical
`advantage that “normal” vessels may be maintained by
`VEGF121 and other isoforms, while pathologic neovascular-
`ization may be suppressed [18, 44, 48]. Indeed, prior to
`clinical trials in humans, basic research demonstrated that
`administration of EYE001 (pegaptanib) could lead to both
`reduced vascular permeability and inhibition of both corneal
`and retinal neovascularization [49]. It has subsequently been
`shown, however, that various proteases activated during
`angiogenesis may cleave VEGF165 (and longer isoforms) to
`generate nonheparin binding fragments—such fragments
`may be sufficient to drive angiogenesis while evading pegap-
`tanib blockade [50, 51].
`
`4.3. Pharmacokinetics and Metabolism. Nonmodified aptam-
`ers are rapidly cleared from the body, with a half-life of mi-
`nutes to hours, as a result of nuclease degradation and renal
`clearance (a result of the inherently low molecular weight
`of aptamers). Therefore, modification of aptamers, such as
`(cid:2)
`-fluorine-substituted pyrimidines, and polyethylene glycol
`2
`(PEG) linkage, can be used to increase their stability and
`terminal half-life (both approaches are used in the case of
`pegaptanib) [44]. Using these approaches pegaptanib has
`been found to be stable in human plasma, at ambient tem-
`peratures, for more than 18 hours [52].
`Pegaptanib pharmacokinetics have been evaluated fol-
`lowing intravitreal injection in monkeys and rabbits [49,
`
`52, 53]. In both animal models, pegaptanib was detected in
`the vitreous at biologically active levels for at least 28 days
`following a single 0.5 mg intravitreal injection. In rabbits,
`after a single dose of pegaptanib, the initial vitreous humor
`levels were approximately 350 μg/mL and decreased by an
`apparent first-order elimination process to approximately
`1.7 μg/mL by day 28. By comparison, the plasma concentra-
`tions of pegaptanib were significantly lower, ranging from
`0.092 μg/mL to 0.005 μg/mL (day 1 to day 21). Plasma levels
`also declined by an apparent first-order elimination. In a
`human pharmacokinetic study, pegaptanib was not found
`to accumulate in the plasma after multiple doses (i.e.,
`systemic exposures were similar at different time-points);
`furthermore, no antipegaptanib antibodies (IgG or IgM)
`were detected [54].
`
`4.4. Selected Clinical Studies: Neovascular AMD. In 2004, fol-
`lowing publication of results from two, concurrent, phase
`III clinical trials (the VEGF Inhibition Study in Ocular Neo-
`vascularization, or VISION, trials), pegaptanib was licensed
`for use in the USA by the Food and Drug Administration
`(FDA) [55]. The VISION trials—two large-scale, multicen-
`ter, randomized, controlled, clinical trials—demonstrated
`that intravitreal administration of 0.3 mg of pegaptanib at
`six weekly intervals, for a period of 48 weeks (a total of nine
`treatments), was effective in reducing moderate vision loss
`in patients with neovascular AMD (higher doses were not
`shown to provide clinical benefit). In these studies, 70% of
`pegaptanib-treated patients avoided further moderate visual
`loss (defined in most AMD studies as a loss of fewer than 15
`letters of visual acuity) compared with 55% of sham-treated
`patients. However, treated eyes still lost, on average, 1.5
`lines of visual acuity over the course of a year of treatment.
`There was no evidence of either systemic toxicity or an
`increased risk of potential VEGF inhibition-related adverse
`events (a safety profile confirmed following three years of
`treatment/follow-up) [56].
`
`4.5. Selected Clinical Studies: Diabetic Macular Edema. In
`2011, the results of a phase II/III-randomized controlled
`trial, of intravitreal pegaptanib for the treatment of diabetic
`macular edema (DME), were published [57]. In this study,
`subjects with DME received injections of 0.3 mg of intravit-
`real pegaptanib, or sham injections, every six weeks for a year,
`and then according to prespecified criteria in a second year.
`In all, 36.8% of patients receiving pegaptanib, versus 19.7%
`of those in the sham group, experienced an improvement
`in visual acuity greater than 10 letters when compared to
`baseline. After two years, pegaptanib-treated patients gained,
`on average, 6.1 letters of visual acuity (versus 1.3 letters
`for controls). Pegaptanib-treated patients also received fewer
`focal/grid laser treatments (subjects were eligible for this
`beginning at week 18).
`
`5. Bevacizumab
`
`Bevacizumab (Avastin, Genentech, South San Francisco, CA)
`is a full-length monoclonal antibody, first derived from a
`murine source and prepared for intravenous administration,
`
`Novartis Exhibit 2025.004
`Regeneron v. Novartis, IPR2021-00816
`
`

`

`Journal of Ophthalmology
`
`5
`
`which binds to and inhibits all isoforms of VEGF [18, 58].
`Bevacizumab was originally developed and approved for the
`treatment of metastatic colorectal cancer but may also be
`of benefit in the treatment of nonsmall cell lung cancer,
`metastatic breast cancer, and glioblastoma multiforme [59].
`Use of bevacizumab in these contexts has been associated
`with increased incidences of hypertension, bleeding, and
`thromboembolic events [59]. However the doses employed
`for intraocular use are many times lower than those used
`systemically, and the efficacy and safety of bevacizumab,
`for the treatment of neovascular AMD, has recently been
`demonstrated in phase III clinical trials [60, 61].
`
`5.1. Chemistry. Bevacizumab was originally developed from
`a mouse antihuman VEGF antibody (A.4.6.1), generated
`from mice immunized with the VEGF165 isoform [58].
`A.4.6.1 recognizes all isoforms of VEGF and, in 1992, was
`shown to inhibit growth of human tumor cell
`lines in
`vivo [62]. Subsequently, in 1996, intraocular administration
`of A.4.6.1 was found to inhibit
`iris neovascularization
`occurring secondary to retinal ischaemia in a primate model
`[42]. In 1997, bevacizumab was developed by humaniza-
`tion of A.4.6.1 [63]. In this process, six complementarity-
`determining regions (CDRs) (i.e., regions that determine
`antibody-binding) were transferred from A.4.6.1 to a human
`antibody framework previously used for humanizations.
`However, this transfer reduced VEGF binding over 1000
`fold—to reduce this effect, eight framework residues were
`changed from human to murine.
`Bevacizumab is produced in Chinese hamster ovary cells
`using expression plasmids (plasmids are DNA molecules sep-
`arate from chromosomal DNA that can be used to manu-
`facture large quantities of proteins) [58]. Bevacizumab is a
`149 kDa full-length antibody, composed of two light chains
`and two heavy chains, and with a 93% human amino acid
`sequence.
`
`5.2. Preclinical Studies. The effects of bevacizumab have been
`examined in a number of in vitro and in vivo studies [58];
`as bevacizumab was not developed with the intention of
`intraocular administration, many of these studies were per-
`formed only after its widespread adoption in this manner for
`clinical practice. In both murine and porcine models, beva-
`cizumab has been demonstrated to reduce VEGF-induced
`permeability and proliferation of choroidal endothelial cells
`and to inhibit VEGF-induced migration of human umbilical
`vein endothelial cells [64–66]. In addition, bevacizumab has
`been demonstrated as nontoxic, or not to alter the viability
`of, neurosensory retinal cells, retinal ganglion cells, and
`human retinal pigment epithelium (RPE) cells [58]. Concern
`has also been raised about the Fc component present in
`full-length antibodies such as bevacizumab—Fc domains
`are known to initiate complement activation and immune
`cell destruction [18]. Recent studies have demonstrated that
`choroidal neovascular membranes from patients with neo-
`vascular AMD treated with bevacizumab are characterized by
`significantly higher inflammatory activity [67]. Preliminary
`results have also demonstrated that bevacizumab Fc domains
`are capable of binding effectively to human RPE and human
`
`umbilical vascular endothelial cell (HUVEC) membranes via
`Fc receptors, activating the complement cascade and leading
`to cell death [58].
`
`5.3. Pharmacokinetics and Metabolism. Bevacizumab was
`developed for intravenous administration in diseases such as
`colorectal cancer [59]. As a result, compounding into smaller
`doses is required for intraocular administration. Studies have
`demonstrated differences in bevacizumab concentration and
`the presence of particulate contaminants following this pro-
`cess, emphasizing the need for implementation of optimal
`protocols when compounding pharmacies prepare this drug
`for intravitreal use [58, 68].
`The pharmacokinetics of bevacizumab, following intrav-
`itreal administration, have not been well characterized.
`Knowledge of the vitreous half-life is an important consid-
`eration when optimizing retreatment frequencies, whereas
`serum concentrations are an important factor with respect
`to systemic adverse effects (e.g., stroke). In rabbits receiving
`1.25 mg of bevacizumab, the vitreous half-life was 4.32 days
`(versus 2.88 days for ranibizumab), and the maximum serum
`concentrations were reached after eight days [69, 70]. Small
`amounts of bevacizumab were also detected in the vitreous of
`the fellow, uninjected eye. In a more recent study performed
`in humans, an aqueous half-life of 9.82 days was found after
`intravitreal injection of 1.5 mg of bevacizumab [71].
`The retinal penetration of bevacizumab has also been
`studied in animal models (experience with retinal pene-
`tration of other, full-length antibodies suggested that their
`large size would act as a limiting factor). In rabbits, Shahar
`et al. demonstrated, using confocal immunohistochemistry,
`that full thickness retinal penetration occurred 24 hours
`after intravitreal injection; this study also demonstrated the
`essential absence of bevacizumab from the retina by four
`weeks post injection [72].
`
`5.4. Selected Clinical Studies: Neovascular AMD. In 2010, the
`results of the ABC (Avastin (Bevacizumab) for treatment
`of Choroidal Neovascularization) trial provided the first
`evidence from a phase III-randomized controlled study
`for the efficacy of intravitreal bevacizumab in neovascular
`AMD [60, 73]. In this single year trial, 32% of patients
`treated with bevacizumab gained 15 or more letters from
`baseline visual acuity (initial three month loading phase, and
`then retreatment as required). In addition, 91% of patients
`receiving bevacizumab lost fewer than 15 letters of visual
`acuity from baseline, and mean visual acuity increased by 7.0
`letters over the study period.
`In 2011, the results of the CATT (Comparison of Age-
`Related Macular Degeneration Treatments Trials) study pro-
`vided further evidence, from larger phase III trial, for the
`efficacy of bevacizumab in neovascular AMD [61]. In this
`trial, 31.3% of patients treated with bevacizumab on a fixed,
`monthly regimen gained 15 or more letters from baseline
`visual acuity (28.0% for patients treated with bevacizum-
`ab as required). In addition, 94.0% of patients receiving
`bevacizumab on a fixed, monthly regimen lost fewer than
`15 letters of visual acuity from baseline (91.5% in the
`bevacizumab as required group). Finally, mean visual acuity
`
`Novartis Exhibit 2025.005
`Regeneron v. Novartis, IPR2021-00816
`
`

`

`6
`
`Journal of Ophthalmology
`
`increased by 8.0 letters over the study period in those re-
`ceiving bevacizumab monthly (5.9 letters in bevacizumab as
`required group). Of note, statistical comparisons between
`bevacizumab given as needed, and given on a fixed, monthly
`regimen, were inconclusive.
`
`6. Ranibizumab
`
`Ranibizumab (formerly known as rhuFAb V2) is an antibody
`fragment that binds and inhibits all
`isoforms of VEGF
`[18]. Specific development of ranibizumab for intraocular
`use was driven, in part, by preliminary studies suggesting
`that full-length monoclonal antibodies would not distribute
`across all retinal
`layers [74]. Furthermore, the relatively
`long systemic half-life of full-length antibodies (versus an-
`tibody fragments) raised concerns about systemic toxicity in
`patients requiring long-term anti-VEGF blockade [18].
`
`6.1. Chemistry. Bevacizumab is constructed from A.4.6.1
`using one of 12 possible Fab (fragment antigen binding)
`variants: “Fab 12” [58]. Ranibizumab is constructed using
`a different Fab variant from A.4.6.1: “Fab MB1.6”, in an
`effort to obtain higher binding affinities for VEGF [75].
`Ranibizumab is produced as a 48 kDa antibody fragment,
`in E. coli, using expression plasmids [58]. It is a chimeric
`molecule, consisting of an antigen-binding murine compo-
`nent, and a nonbinding human component that serves to
`make it less antigenic (in Greek mythology, the chimera was a
`monster with a lion’s head, a goat’s body, and a serpent’s tail).
`On a molar basis, ranibizumab is between five- and 20-times
`more potent than bevacizumab at binding of VEGF [75].
`
`6.2. Preclinical Studies. Preclinical studies have demonstrat-
`ed the safety, tolerability, and efficacy of ranibizumab in an-
`imal models. In particular, intravitreal administration of
`ranibizumab reduced vascular leakage in a monkey model
`of choroidal neovascularization (CNV), while pretreatment
`with ranibizumab prevented laser-induced development of
`CNV in this model [76].
`
`6.3. Pharmacokinetics and Metabolism. The pharmacokinet-
`ics of ranibizumab, after intravitreal administration, have
`been studied both in animal models and in human trials [69,
`77, 78]. Ranibizumab is thought to exit the vitreous cavity
`posterior via retinal penetration and choroidal vascular
`drainage or anteriorly via the aqueous drainage route. In
`animal studies, ranibizumab is cleared from the vitreous
`with a half-life of approximately three days [69]. Therefore,
`ranibizumab is thought to maintain biologically active retinal
`concentrations for approximately one month. After reaching
`a maximum at approximately one day, the serum concen-
`tration of ranibizumab declines in parallel with this. In
`human studies, following monthly intravitreal ranibizumab
`administration, maximum serum concentrations were dose
`dependent but low (0.3 ng/mL to 2.36 ng/mL—levels more
`than 1000 fold lower than in the vitreous and thought to be
`below the concentrations necessary for reduction in biolog-
`ical activity of VEGF by 50%) (http://www.gene.com/). In a
`recent study by Bakri et al., no ranibizumab was detected in
`
`the serum, or the fellow uninjected eye, of rabbits injected
`with 0.5 mg of intravitreal ranibizumab; by comparison,
`small amounts of bevacizumab were detected, both in

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket