You are on page 1of 14

Anatomy and Physiology Section 1

Retinal and Choroidal Vasculature: Chapter

Retinal Oxygenation
Maria B. Grant, Gerard A. Lutty 18 
INTRODUCTION vasculature is supplied with blood directly by the central retinal
artery in humans. The retinal vasculature has a traditional end-
Oxygen is necessary for the existence of mammals because it is arterial hierarchy: arteries branch to arterioles, which supply a
required to generate adenosine triphosphate (ATP) oxidatively. capillary network that is drained by venules, and then veins
Although the partial pressure of oxygen (pO2) is 149 mmHg (21% remove the blood from the retina (Fig. 18.1). The retinal vascu-
of atmospheric oxygen) at sea level, arterial oxygen content is as lature forms the inner blood–retinal barrier (BRB), restricting
low as 75–100 mmHg (10–14%) and tissue pO2 is much lower. The passage of molecules that do not have receptors or transporters
oxygen level in inner segments of photoreceptors (mitochondria- on the luminal surface of the endothelial cells. Capillaries have
rich) after dark adaptation is between 0 and 5 mmHg (0.7%) but a lumen diameter of 3.5–6 µm, permitting passage of red blood
up to 20 mmHg in the light. Inner retinal oxygen is normally cells only after deformation of their disc shape. The retinal capil-
20 mmHg, so normoxia depends on the area of retina and dark/ laries and venules have perivascular pericytes and the retina has
light state.1,2 Hypoxia is an oxygen level below normoxia while the highest endothelial cell-to-pericyte ratio in the body, 1 : 1.8
hyperoxia is achieved by inhaling high levels of oxygen as in the The choroidal vasculature forms well before the retinal vessels
isolette of the neonatal intensive care unit. (6–9 WG), although its maturation is only completed after
The retina is one of the most metabolically active tissues in the 20 WG. It develops by hemovasculogenesis, formation of blood
body. It has two unique zones of oxygenation.1 The inner retina cells and blood vessel cells from a common progenitor, the
is supplied with oxygen by the retinal vasculature. The retinal hemangioblast.9 The choroidal vasculature provides oxygen and
vasculature is autoregulated because it is responsive to changes nutrients to the photoreceptors. The capillary system, the chorio-
in systemic oxygen levels, keeping the inner retina at a relatively capillaris, lies directly under the Bruch’s membrane, while inter-
constant level. If the retinal vasculature is compromised, as in mediate and large blood vessels of the system lie posterior to the
ischemic retinopathies, the retina becomes hypoxic in that area. capillaries. The short and long ciliary arteries supply blood to
The outer retina is supplied solely by the choroidal vasculature. the choroidal vasculature while 4–6 vortex veins remove blood
Unlike retina, choroidal vessels are not autoregulated, so sys- from this vast system. Unlike the retina, the hierarchy in choroid
temic levels of oxygen control the level of oxygen in choroid. is lobular, similar to kidney glomeruli (Fig. 18.1). The lobules
Supply of oxygen to choroid is diminished by stenosis of the change in shape, vascular density, and size depending upon area
ophthalmic artery, which is the branch off the internal carotid and the location of feeding arterioles and draining venules also
that is most likely to be stenosed because it is a right-angle varies by geographic location of the lobule.10 The capillaries are
branch. broad and flat, having luminal diameters ranging from 10 to
38 µm in diameter. Another major difference from retina is that
Comparison of retinal and the capillaries are fenestrated, allowing the passage of small
choroidal vasculatures molecules and solutes through these 60–70-nm pores. The cho-
riocapillaris is sided in that the majority of the fenestrations are
Although less than 300 µm apart in distance, the retinal and
on the retinal side as well as all three types of vascular endothe-
choroidal vasculatures are vastly different in many attributes in
lial growth factor (VEGF) receptors.11 Pericytes, however, are
addition to autoregulation. The initial retinal vasculature in
mostly on the scleral side of these capillaries. Control of vascular
human starts forming around 14 weeks’ gestation (WG) by vas-
tone in choroid may be accomplished by mast cells, which lie
culogenesis, development by differentiation and assembly of
abluminal to arteries and arterioles, or choroidal ganglion cells.12
vascular precursors, angioblasts.3–5 The deep capillary network
forms after 20 WG by angiogenesis, development by migration,
and proliferation of endothelial cells from existing blood vessels.
HISTORY OF RETINAL ISCHEMIA
The driving force for vascular development is physiological Ischemia is the restriction in oxygen supply without considering
hypoxia; metabolic requirements of developing neurons are only actual levels of oxygen. Michaelson13,14 and Wise15 hypothesized
met by stimulating the development of a retinal vasculature.6,7 that areas of vascular loss in retina must be hypoxic because the
It is mostly a bilayered system, a superficial network, and a deep high metabolic rate requires a continuous supply of oxygen.
capillary plexus; however, there are multiple layers of capillaries They observed that neovascularization always formed adjacent
in the peripapillary region, the radial peripapillary capillaries. to these nonperfused areas and, therefore, an angiogenic factor
There is only one layer of blood vessels in the periphery at must be produced by the hypoxic retina. They hypothesized
the ora serrata where the retina thins to 100 µm. The retinal that this factor X must be hypoxia-inducible and diffusible.
Fig. 18.1 Comparison of retinal (A-C) and
choroidal (D-F) vasculatures. (A) The retinal
vasculature, stained here with adenosine
434 diphosphatase (ADPase) enzyme
histochemistry, has an end-arterial hierarchy:
arteries (arrow) have a capillary-free zone,
and arterioles and then capillaries, which
Section 1

are drained by venules, and then veins


(arrowhead). (B) A capillary in the deep
capillary network has a pericyte and
endothelial cell. (C) Transmission electron
A D microscopy (TEM) of a retinal capillary shows
a pericyte (p) within the shared basement
membrane (arrow) with an endothelial cell
(e). (Courtesy of DB Archer, TA Gardiner,
Anatomy and Physiology
Basic Science and Translation to Therapy

and AW Stitt.) (D) The choriocapillaris,


p stained by alkaline phosphatase (APase)
enzyme histochemistry, has a lobular
e hierarchy. (E) When sectioned, the
endothelial cell APase activity in
c
c choriocapillaris (c) shows the broad flat
B E lumens of the choriocapillaris which are
positioned under retinal pigment epithelial
cells (top) and Bruch’s membrane between
them. (F) With TEM the thin processes of the
choriocapillaris endothelial cells (arrow) are
p apparent and, at higher magnification (inset),
the fenestrations in these endothelial cells
are visible.

e
C F

Subsequently, oxygen was measured directly in retinas of several children are placed in 40% oxygen, making their tissue further
species and it demonstrated that nonperfused areas were indeed hyperoxic, which yields vaso-obliteration (endothelial cells die
hypoxic.1,2 It was not until 1989 that factor X was discovered, and pericytes and progenitors survive).19 The only direct mea-
purified, and characterized as vascular endothelial growth factor surements of oxygen in a model of retinopathy of prematurity
(VEGF).16 This factor was first shown to be responsible for the (ROP) were performed by Ernest and Goldstick.20 They found in
increased vascular permeability seen in some retinopathies.17 kitten after 80–90% O2 that preretinal pO2 over avascular retina
was close to zero but was normal over vascularized retina. Vaso-
NORMOXIA obliteration from hyperoxia does not occur in the choroid of
The studies of Wangsa-Wirawan and Linsenmeier1 and Yu and humans and dogs19 but does occur when rats are exposed to
Cringle2 using oxygen electrodes directly assessed oxygen levels hyperoxia.21 Loss of vasculature in vaso-obliteration makes the
from choroid to vitreous in various species. The oxygen tension retina hypoxic when the child is returned to room air. Exposure
is around 70 mmHg in choroid and plummets to zero at the of the adult vasculature to hyperoxia causes constriction but not
inner segments in the dark (Fig. 18.2). Inner retina is around vaso-obliteration. During hyperoxia breathing (100% oxygen),
10–20 mmHg. There are regional variations in oxygen concentra- the inner retinal pO2 remains unchanged due to autoregulation
tion within the retina. Yu et al.18 showed that oxygen consump- while the choroidal pO2 rises to 250 mmHg in cat22 and 220 mmHg
tion in outer retina is highest in the parafoveal region while inner in the minipig, due to a lack of metabolic control of the choroidal
retinal oxygen in the fovea (approximately 5 mmHg) reflected vasculature.23
the lack of a retinal vasculature and the predominantly choroidal
source of oxygen. HYPOXIA
Complex homeostatic mechanisms are designed to maintain
HYPEROXIA O2 concentration in each cell within a narrow range. While
Life in utero is hypoxic, so when a child is born prematurely, the O2 consumption increases with the metabolic activity of the
normoxic environment is actually hyperoxic. Prematurely born organism, exposure to O2 must be limited due to the potentially
Fig. 18.2 Oxygen profile measured with
microelectrodes in cat retina during light
and dark adaptations. The schematic
at the top shows where, anatomically, 435
80 the measurements were taken from
choriocapillaris (left, retinal depth = 100)
to the internal limiting membrane (right,

Chapter 18
retinal depth = 0). (Reproduced with
Light permission from Wangsa-Wirawan ND,
60 Linsenmeier RA. Retinal oxygen. Arch
Dark
Ophthalmol 2003;121:547–55.)
Oxygen tension, mmHg

40

Retinal and Choroidal Vasculature:


20

110 100 90 80 70 60 50 40 30 20 10 0
Retinal depth, %

damaging effects of reactive oxygen species (ROS). Hypoxia, which facilitates heterodimerization, and a C-terminus, which
the state of low oxygen concentration, promotes the formation recruits transcriptional coregulatory proteins.
of blood vessels and is important for the formation of a vascular The activity of HIF depends on the intracellular levels of its
system in embryos.24 Disease occurs when the retina and choroid inducible alpha subunit. In the presence of oxygen, HIF-1α is
are deprived of adequate oxygen supply; this can also be hydroxylated on two critical proline residues (Pro402 and Pro564)
described as a mismatch of oxygen supply versus demand at in the so-called oxygen-dependent degradation domain. Three
the cellular level within ocular tissues. prolyl hydroxylases have been identified in mammalian cells
The blood O2-carrying capacity is maintained by the O2- and use O2 as a substrate to generate 4-hydroxyproline at resi-
regulated production of erythropoietin (EPO), which stimulates dues 402 and/or 564 of HIF-1α. The hydroxylation reaction
the proliferation and survival of red blood cell progenitors. also requires 2-oxoglutarate (α-ketoglutarate) as a substrate and
Semenza and coworkers25,26 performed seminal studies to iden- generates succinate as a side product. These prolyl hydroxylases
tify hypoxia-inducible factor-1 (HIF-1). HIF-1 orchestrates a have a high Km for O2 that is slightly above atmospheric con-
pleiotropic adaptive response to hypoxia by inducing the expres- centration; thus O2 is rate-limiting for enzymatic activity under
sion of more than 100 genes encoding glycolytic enzymes and physiological conditions and any change in cellular O2 concen-
glucose transporters (thereby facilitating the glycolytic switch in tration is directly transduced into changes in the rate of HIF-1α
energy metabolism typically observed under hypoxic condi- hydroxylation.29
tions), matrix metalloproteinases, and angiogenic, mitogenic, Factor-inhibiting HIF-1 (FIH-1), which was identified in a
and survival factors, including EPO.27,28 Other molecules upreg- yeast two-hybrid screen as a protein that interacts with, and
ulated by HIF-1 that have profound effects on vasculature inhibits the activity of the HIF-1α transactivation domain,30
include 5’ nucleotidase, an enzyme that is the major source of functions as asparaginyl hydroxylase.31 As in the case of the
the potent vasodilator adenosine in the body, and VEGF. HIFs prolyl hydroxylases, FIH-1 appears to use O2 and 2-oxoglutarate,32
are vital to development and, in mammals, deletion of the HIF-1 although it has a Km for O2 that is three times lower than
genes results in perinatal death. HIF-1 is expressed in all cell the prolyl hydroxylases.33 Hydroxylation provides a mechanism
types and functions as a master regulator of oxygen homeostasis for regulating protein–protein interactions, similar to the effect
by playing critical roles in embryonic development and post­ of phosphorylation and other posttranslational modifications.
natal physiology. However, this hydroxylation occurs in an O2-dependent manner,
thus establishing a direct link between cellular oxygenation and
Hypoxia-inducible factor HIF-1 activity. Following HIF-1α hydroxylation, the protein
HIF is a highly conserved transcriptional complex which is a becomes targeted for ubiquitination by an E3 ligase complex
heterodimer composed of an alpha and a beta subunit. HIF-1 (including the von Hippel–Lindau (VHL) tumor suppressor
belongs to the PER-ARNT-SIM (PAS) subfamily of the basic protein) and subsequent proteasomal degradation.
helix–loop–helix (bHLH) family of transcription factors. The Under hypoxic conditions, the HIF prolyl-hydroxylases are
alpha and beta subunit both contain an N-terminus bHLH inhibited, because these HIF prolyl-hydroxylases utilize oxygen
domain for DNA binding, a central region with PAS domain, as a cosubstrate. Hypoxia results in an increase in succinate,
Fig. 18.3 Hypoxia-inducible factor 1 (Hif-1α)
in hypoxia and hyperoxia.
Hyperoxia Ub
436 OH OH
Prolyl E3 Ligase
hydroxylase Degradation
complex
Hif-1α Hif-1α Hif-1α
Section 1

Hypoxia

Angiogenic growth factors


Hif-1β
Angiogenic survival factors
(ARNT)
Anatomy and Physiology
Basic Science and Translation to Therapy

Cytoskeletal proteins
Proapoptotic proteins
Glucose transporters
Glycolytic enzymes
Nucleus Hif-1β
(ARNT)
Transcription

Hif-1α
Hif-1β
(ARNT)

due to inhibition of the electron transport chain in the mito- system is required for embryonic survival by embryonic day 9
chondria, which serves to inhibit further HIF prolyl-hydroxylase (E9) in the mouse. In wild-type mouse embryos, HIF-1α expres-
activity. When stabilized by hypoxic conditions, HIF increases sion increases dramatically between E8.5 and E9.5, whereas
the expression of critical genes that promote survival in low- embryos that lack HIF-1α expression die between E9.5 and E10.5
oxygen conditions, including glycolytic enzymes, which allow and show cardiac malformations, vascular regression, and
ATP synthesis in an oxygen-independent manner. HIF activates massive cell death.40 Complete HIF-2α deficiency is also associ-
the transcription of genes encoding secreted signaling mole- ated with embryonic lethality41 and because the embryos survive
cules, including angiogenic growth factors and survival factors, longer than HIF-1a–/– mice, effects on multiple organ systems can
cell surface receptors, extracellular matrix proteins and modify- be demonstrated.42
ing enzymes, transcription factors, cytoskeletal proteins, pro- Complete HIF-1α deficiency results in developmental defects;
apoptotic proteins, and glucose transporters and glycolytic however, partial HIF-1α deficiency is sufficient to result in
enzymes (Fig. 18.3).29 impaired responses to physiological stimuli. A particularly dra-
HIF-induced VEGF, stromal-derived factor-1 (SDF-1), and matic example is the loss of O2 sensing in the carotid body of
EPO promote neovascularization. HIF-1 acts by binding to HIF- HIF-1a+/– mice.43 Although the carotid bodies are anatomically
responsive elements in promoters that contain the sequence and histologically normal and depolarize normally in response
NCGTG, which is present in the promoters for VEGF, SDF-1, to cyanide application, they show essentially no response to
EPO, and many other genes. In addition to hypoxia, other factors hypoxia. Thus partial HIF-1α deficiency in the carotid body
such as nuclear factor κB (NF-κB) modulate HIF-1α expression results in a complete loss of the ability to sense and/or respond
in the presence of normal oxygen pressure. Thus, conditions to changes in the arterial PO2 by stimulation of the central
such as tissue inflammation can lead to local HIF-1α expres- nervous system cardiorespiratory centers. The HIF-1 target
sion.34 HIF-1 DNA-binding activity and target gene expression genes that are critical for O2 sensing and/or efferent responses
are induced in cells exposed not only to hypoxia but also to the by the carotid body have not been identified.
iron chelator desferrioxamine or to cobalt chloride.35 Mice with HIF-1α conditionally knocked out using PAX6-Cre
A structurally and functionally related protein to HIF-1α, des- have delayed development of the outer retinal plexus but not
ignated HIF-2α, is the product of the EPAS1 gene. HIF-2α can the superficial or deep plexus.44 However, when HIF-1α was
also heterodimerize with HIF-1ß.36 HIF-1α:HIF-1ß and HIF- knocked down only in Müller cells using a Cre-LOX system, and
2α:HIF-1ß heterodimers have overlapping yet distinct target the animals were made diabetic with streptozotocin, vascular
gene specificities.37 HIF-2α, unlike HIF-1α, is not expressed in all permeability in retina was reduced and leukostasis and overpro-
cell types and HIF-2α can be inactivated by cytoplasmic seques- duction of VEGF and intercellular adhesion molecule (ICAM)-1
tration. This “compartmentalization” of oxygen-sensitive signal- were attenuated in adult mice.45
ing components also influences the hypoxic response.38,39 Another dramatic phenotype is the complete inability of HIF-
1α–/– myeloid cells (granulocytes and macrophages) to respond
HIF deficiency and its resultant pathology to inflammatory stimuli.46 Myeloid cells are dependent on gly-
O2 delivery to cells of the developing embryo becomes limited colysis for ATP generation, perhaps reflecting the hypoxic
by diffusion such that establishment of a functioning circulatory microenvironment that is often associated with inflammation
and infection. HIF-1α deficiency results in ATP deficiency, which While VEGF is critical to maintaining normal ocular func-
impairs critical myeloid cell functions such as aggregation, tion, overproduction of VEGF is deleterious. Elevated levels
motility, invasion, and bacterial killing. HIF-1 also plays of VEGF have been strongly implicated in the pathogenesis 437
critical roles in B-lymphocyte development47 and T-lymphocyte of ocular neovascular diseases such as neovascular age-related
activation.48 macular degeneration (NV in AMD)66 and proliferative diabetic

Chapter 18
retinopathy67 as well as diabetic macular edema.68 Elevated
HIF-activated genes relevant to VEGF levels are observed in central and branch retinal vein
physiological and pathological occlusion (CRVO and BRVO),69 neovascular glaucoma,70 and
ROP.71 Blocking VEGF action is now an established strategy
ocular angiogenesis for the treatment of NV in AMD, with two agents (the RNA
The paragraphs above provide a brief summary of the critical aptamer pegaptanib sodium72 and the humanized murine
role of HIF-1α in oxygen sensing, development, and physiology. monoclonal antibody antigen-binding fragment ranibizumab73)

Retinal and Choroidal Vasculature:


HIF-1α plays an equally important role in disease pathophysiol- having received regulatory approval for the intravitreal treat-
ogy, including retinal diseases. As a result, there is considerable ment of NV in AMD.
interest in HIF-1 α as a therapeutic target.49 In cardiovascular While current standard of care for NV in AMD uses intravitreal
diseases, increased HIF-1α activity induced as a result of HIF-1α delivery of an anti-VEGF agent on a repeated basis, this routine
gene therapy,50 small-molecule inhibitors of prolyl hydroxylase clinical practice does not address all the nuances of VEGF biology.
activity,51 or inhibitors of HIF-1α–VHL interaction52 may provide The biology of the VEGF proteins is extremely complex. The
a means to stimulate neovascularization in ischemic tissue. In VEGF family members are part of a superfamily of cysteine
contrast, small-molecule inhibitors of HIF-1 activity may be knot proteins and include VEGF-B, -C, -D, and placental growth
useful as antiangiogenic agents. However, because HIF-1α func- factor (PlGF). Alternative splicing events for VEGFA give rise to
tions as a global regulator of oxygen homeostasis, it may not be at least 14 subtypes of VEGF, namely, VEGF111, VEGF121, VEGF121b,
a useful therapeutic target if the treatment results in unintended VEGF145, VEGF145b, VEGF148, VEGF162, VEGF165, VEGF165b, VEGF183,
and undesirable side-effects. VEGF183b, VEGF189, VEGF189b,74 and VEGF206.75,76 Following the
An alternative therapeutic approach that may be particularly discovery of the antiangiogenic isoform of VEGF, VEGF165b, and
relevant to the treatment of ocular pathology is to focus on its associated family of isoforms, a further layer of complexity
modulation of HIF-1α target genes. However, the protein prod- was added to understanding the regulation of VEGF.
ucts of these target genes must also be delivered in a precise and All VEGF isoforms are essential regulators of angiogenesis
perfectly timed manner. EPO is an oxygen-regulated hormone and vascular permeability. VEGFs elicit their intracellular activi-
stimulating erythrocyte production and is critical for retinal ties via the activation of two receptor tyrosine kinases (RTKs):
angiogenesis. Increasing EPO expression in phase 1 of the VEGFR-1 and -2. VEGFR-1, a high-affinity fms-like tyrosine
murine ROP model (postnatal days 7–12) is protective and kinase-1,77 and VEGFR-2, a kinase insert domain-containing
results in less neovascularization during phase 2 (postnatal days receptor,78 are transmembrane glycoproteins consisting of a
12–17).53 In contrast, EPO mRNA expression levels in retina are seven-tandem immunoglobulin-like domains, which serves as
highly elevated during the hypoxia-induced proliferation phase the extracellular ligand-binding region, a single-transmembrane
of retinopathy (phase 2) and inhibition of retinal EPO mRNA domain, and a cytoplasmic domain consisting of two tyrosine
expression with RNA interference results in suppressed retinal kinase catalytic domains. Moreover, it has also been reported
neovascularization.53 that a family of cell surface glycoproteins, particularly
The best-known gene activated by HIF-1α is VEGF, first iden- neuropilin-1, act as isoform-specific coreceptors for VEGF-A.79
tified as a potent promoter of vascular permeability17 and endo- The VEGF family and their respective receptors are outlined in
thelial cell proliferation.15 VEGF has become known as a master Fig. 18.4.
regulator of angiogenesis.54 Tight control of physiologic VEGF Ligand binding to the extracellular domain of VEGFR-2 results
levels is required for proper embryological development.55 in a maximal increase of kinase activity following the induction
Although it was initially thought that the postembryonic role of of receptor dimerization and subsequent phosphorylation of
VEGF was restricted to a few processes, it is now quite clear that tyrosine residues on the intracellular domain of the receptor.80
VEGF acts as a pluripotent growth factor essential for a wide This event is crucial for the recruitment of additional signaling
variety of physiological processes,56 including maintenance of molecules that contain Src homology 2 or phosphotyrosine
the adult microvasculature,57 neuronal survival,58 and trophic binding domains, which mediate further downstream signaling
maintenance of ocular tissues. cascades.81 The association of RTKs with coreceptors, such as
NP-1, in the case of VEGFR-2:VEGF165 signaling/interaction,
VEGF in health and in ocular disease can enhance the functional signal transduction and facilitate
VEGF is produced by many cell types in the retina, including diverse cellular responses.80 VEGFR1/R2 signaling activates
retinal pigment epithelium (RPE),59 vascular endothelial cells,60 RAS, raf1, MEK1, and ERK1/ERK2 and stimulates PI3K/AKT/
pericytes,60 retinal neurons,61 Müller cells,61 and astrocytes,62 sug- PKCz/MAPK pathways to mediate proliferation, migration, and
gesting that VEGF has important functions in ocular homeosta- cell survival (Fig. 18.4).
sis. RPE-secreted VEGF plays an important role in maintaining VEGFR-2 is the major mediator of angiogenic signaling in
the choriocapillaris.11,63,64 VEGF secretion by retinal cells and the endothelial cells and is required for de novo vessel formation,
RPE is stimulated in response to hypoxia.60 VEGF administration vasculogenesis, and for angiogenesis, the formation of vessels
protects retinal neurons from apoptosis.65 Moreover, chronic from pre-existing vasculature.82 The pathways leading to VEGFR
VEGF inhibition can lead to a significant loss of retinal ganglion internalization and the role of receptor degradation in VEGF
cells in normal adult animals.65 signaling remain controversial and differ for VEGFR-1 and -2.
VEGFRs generate signal output at the plasma membrane and on including the nucleus,88 where receptors encounter distinct sig-
their way to degradation through endocytic vesicles,83 whereas, naling molecules.89
438 in unstimulated cells, VEGFR-2 is predominantly located in recy- In addition to EPO and VEGF, SDF-1 is hypoxia-regulated and
cling endosomes identified by Rab4 and/or Rab5.84 VEGFR inter- ischemic tissues express high levels of SDF-1 to recruit repara-
nalization is clathrin-mediated and transport is further directed tive cells to the injured region. SDF-1 activation of either CXCR-4
by the endosomal sorting complex required for transport pro- or CXCR-7 results in stimulation of the Ras/Raf/Mek/ERK
Section 1

teins.85 VEGFR signaling is also regulated by ubiquitination, not pathway and the PI3K/Akt pathway to promote endothelial cell
only of the receptor itself, but also of receptor-associated signal- proliferation and neovascularization. Thus ligand–receptor
ing molecules.86 Specific VEGFR trafficking regulates biological interaction of VEGF to VEGF-R1 and VEGF-R2 and SDF-1 to
output, as shown for arterial morphogenesis, for example.87 CXCR4/CXCR-7 and their subsequent internalization sets in
The molecular basis for ligand specificity of VEGFR signaling motion the cascade of cellular effects of VEGF and SDF-1
is poorly understood. It is well accepted however that VEGF (Fig. 18.4). The VEGF and SDF-1 signaling pathways appear
Anatomy and Physiology
Basic Science and Translation to Therapy

receptors can associate with distinct coreceptors such as neuro- to be intimately connected with HIF-1 activation (Fig. 18.3), as
pilins, integrins, semaphorins, or heparan sulfate glycosamino- the promoters of each of these factors contain a HIF response
glycans, and engage distinct signaling molecules giving rise to element. Because hypoxic tissue releases SDF-1 and VEGF,
specific signal output. Ligand-specific signaling may also result varying O2 concentrations would have an effect on expression of
from receptor trafficking to specific cellular compartments, the receptors for these factors.

VEGF-A
VEGF-B VEGF-E
Insulin
PIGF
IGF-1

IGF-2 SDF-1

VEGF-R1 VEGF-R2 IGF-1R CXCR4/CXCR7

Signaling by
intracellular RAS PI3K/Akt PLCg
translocation

RAF1 PKCz PKC

MEK1 MAPK

ERK1

ERK2

Fig. 18.4 Vascular endothelial growth factor (VEGF) family and its receptors. PlGF, placental growth factor; IGF, insulin-like growth factor;
SDF, stromal-derived factor.
Bone marrow-derived progenitor cells endothelial-like cells109; however, the blood vessels formed by
the endothelial cells of CD14+ origin eventually generate patho-
(BMPC) and vascular repair
logical blood vessels with increased permeability and contribute 439
In conditions like diabetic retinopathy and ROP, areas of retinal to the pathology of diabetic retinopathy.
vasodegeneration occur and lead to retinal ischemia which in We and others have been particularly interested in a novel
turn induces the expression of hypoxia-regulated angiogenic

Chapter 18
factor expressed in increased concentrations in hypoxic tissue,
factors. Typically, BMPCs robustly respond to these factors, insulin-like growth factor-binding protein 3 (IGFBP-3). While
including VEGF and SDF-1.90 the standard IGF-dependent actions of the family of IGF-binding
Importantly, all hypoxia-regulated angiogenic factors are proteins have been well described, recently several IGF-
modulators of bone marrow-derived stem cells. Specifically, independent actions have been discovered for IGFBPs, including
hematopoietic stem cells and other BMPCs reside in the bone IGFBP-3. These IGF-1 independent actions have been charac­
marrow and are mobilized into the peripheral circulation by terized as regulating cell fate and apoptosis.110–113 The role of

Retinal and Choroidal Vasculature:


increased levels of EPO, SDF-1, and VEGF released by the isch- IGFBP-3 in the control of cell growth remains an area under
emic tissue. Hematopoietic stem cell/BMPC mobilization occurs intense study, since IGFBP-3 may enhance or suppress cell
in response to vascular injury throughout the body and, when growth, depending on specific conditions.112,114 In the retina,
functioning properly, leads to revascularization of injured multiple forms of IGFBP (2–5) are secreted by retinal endothelial
areas.91 In healthy individuals, bone marrow-derived CD34+ cells.113 IGFBP-3 has been shown to enhance cell proliferation
endothelial progenitor cells as well as other BMPC successfully in retinal endothelial cells and to decrease the formation of
orchestrate the reparative process. These BMPC demonstrate a neovascular tufts in a murine model of oxygen-induced reti-
limited ability to differentiate directly into endothelial cells and nopathy.115,116 These studies suggest that IGFBP-3 may be
form components of new blood vessels by vasculogenesis, but vascular-protective in the retina. Recently, we also have dem-
show marked ability to provide paracrine support for the resi- onstrated that IGFBP-3 is neuroprotective in the retina and
dent vasculature. This paracrine support facilitates resident reduces injury-induced retinal inflammation.117
endothelial cell recovery.

Disease-associated BMPC dysfunction Key factors that modulate VEGF function


In diabetes, for example, BMPC are dramatically altered and in the retina
cannot facilitate the repair process. Diabetic individuals have The retina is one of the last organs to become vascularized in the
fewer circulating CD34+ cells and an increased number of fetus. The vasculogenic part of the process of human retinal
inflammatory BMPC such as CD14+ cells than nondiabetics.92–94 vascularization begins in about gestational week 14 and appears
This diabetes-related bone marrow dysfunction is closely linked to depend on a physiological hypoxia7,118 brought about by an
to the impaired healing response experienced by many diabetic increase in metabolic demand within the developing retina.3–5
patients and to the vasodegenerative aspect of diabetic macro- This physiological hypoxia induces the local release of VEGF
and microvascular complications.94–96 Diabetes-induced BMPC which, together with IGF-1, regulates angiogenesis and therefore
defects occur in part due to uncoupling of nitric oxide synthases, normal vascularization of the retina.118,119 Retinal vascularization
enhanced NADPH oxidase activity, and increased generation of is complete by 36–40 WG.
ROS such as superoxide and peroxynitrite (ONOO-)97 within ROP is a two-phase disease in which the phases are mirror
BMPC. While stem and progenitor cells are deemed more resis- images; the controlling growth factors are deficient in phase 1
tant to oxidative stress,98 the highly oxidative diabetic milieu has and in excess in phase 2. Phase 2 involves uncontrolled prolifera-
a clearly detrimental effect on the function of these cells.99 Pro- tive growth of retinal blood vessels in response to hypoxia. The
longed oxidant exposure reduces reparative function100 by therapeutic intention would be to prevent the first phase of ces-
impairing antioxidant defense enzymes. Previously, we and sation of vessel and neural retinal development and then the
others have shown that diabetic CD34+ cells exhibit decreased second destructive phase would be prevented. This has been
migration and adhesion activities in vitro, and consequently successfully performed using exogenous EPO, VEGF, and
reduce recruitment to areas of injury.96 In addition to oxidative IGFBP-3 during phase 1 to prevent phase 2. These two phases,
stress, other key mechanisms implicated in diabetes-induced seen in premature infants, can be duplicated in animal models
BMPC dysfunction include a reduction of cathepsin L activity101 of ROP. Hyperoxia causes vaso-obliteration and cessation of
and an upregulation of thrombospondin-1.100,102 While these normal retinal blood vessel development, which mimics phase
functional defects are profound, strategies that successfully 1 of ROP. IGF-1 levels rise in the third trimester of pregnancy
reverse BMPC defects in diabetics have included: (1) enhance- but not in the preterm infant. The low IGF-1 in the preterm infant
ment of angiogenic stimulus by increasing BMPC mobilization is due to the infant no longer being exposed to the support of
using granulocyte colony-stimulating factor and targeting the maternal environment, including maternal sources of IGF-1.
SDF-1103; (2) use of nitric oxide donors to correct migration and IGF-1 levels rise slowly after preterm birth, as these babies are
promote cell deformability104; (3) enhancing cell interactions unable to produce adequate IGF-1 compared to term infants.120
with substrate proteins to increase attachment to basement In these premature infants, IGF-1 is further reduced by poor
membranes105; (4) reducing high levels of endogenous trans- nutrition,121 acidosis, hypothyroxinemia, and sepsis.122 IGF-1
forming growth factor-β to normal levels106; and (5) treatment appears important for retinal and brain growth.119 Low levels of
with rosiglitazone107 or atorvastatin.108 IGF-1 appear to play an important role in the early cessation of
In addition to CD34+ cells, other populations of bone retinal growth that precipitates ROP.
marrow cells may attempt vascular repair, such as CD14+ cells, IGF-1 mediates its effects through activation of the IGF-1 recep-
discussed above, which can, under select circumstances, form tor (IGF-1R) (Fig. 18.3). IGF-1R is a receptor tyrosine kinase and
is well established as a key regulator of cell growth and survival independent predictors.128 These observations are supported by
with activation of Ras-ERK pathway, the PI3K/Akt pathway and several other prospective studies, in which tissue plasminogen
440 PKC (Fig. 18.4). Insulin and IGF-2 can also signal using the IGF-R. activator, another marker of reduced fibrinolysis,129 and von
There is also a growing body of data to support a role for the Willebrand factor, a marker of endothelial injury, were predic-
structurally and functionally related insulin receptor (IR) in cell tive.130 In one clinical study of type 2 diabetics, adhesion mol-
survival, even though its major function has been to modulate ecules were higher in subjects with retinopathy than those
Section 1

metabolism. Bidirectional cross-talk between IGF-1R and IR is without,131 and in another population-based cohort, composite
observed, where specific inhibition of either receptor confers a scores of both inflammatory and endothelial function markers
compensatory increase in activity for the reciprocal receptor. were strongly associated with the presence of diabetic retinopa-
Although fluctuating oxygen has long been associated with thy.132 Similarly, E-selectin values were found to be increased
the development of ROP, oxygen-regulated factors like VEGF in a group with type 1 diabetes and retinopathy.133 Adiponectin
appear to be directly modulated by IGF-1. IGF-1 is a key growth was increased in the advanced stages of retinopathy.134 These
Anatomy and Physiology
Basic Science and Translation to Therapy

factor in early retinal development. IGF-1 controls maximum results should be interpreted cautiously, however, as serum
VEGF activation of the Akt endothelial cell survival pathway markers are not necessarily indicators of tissue events. However,
(Fig. 18.3). Thus loss of IGF-1 leads to loss of VEGF signaling P-selectin, ICAM-1, and polymorphonuclear leukocyte numbers
and retinal vaso-obliteration. This vaso-obliteration leads to are all elevated in human diabetic retina.135 Experimental work
phase 2 of ROP which is the proliferative phase. At this point, in diabetic rat retinas later demonstrated that inflammatory
suppression of IGF-1 and VEGF can reduce neovascularization. cytokine-mediated leukostasis occurs early in diabetic retina
Thus IGF-1 is critical to normal retinal vascular development and neutralizing ICAM-1 and CD-18 prevents it.136–138 From these
and a lack of IGF-1 in the early neonatal period is associated with studies, it appears that inflammation-mediated EC injury and
lack of vascular growth and with subsequent proliferative ROP. vaso-occlusion may cause nonperfusion and subsequrent
In IGF-1-null mice, the retinal blood vessels grow more slowly hypoxia in diabetic retina.139,140
than in those of normal mice, a pattern very similar to that seen The hypothesis that inflammation is critical to the develop-
in premature babies with ROP. ment of diabetic retinopathy arose from initial reports that dia-
betic patients taking salicylates to treat rheumatoid arthritis had
ADULT RETINAL HYPOXIA AND ETIOLOGY a lower-than-expected incidence of diabetic retinopathy.141 The
subsequent decades demonstrated an increase in inflammatory
Diabetic retinopathy markers and growth factors in the diabetic vitreous and retina.
Much like ROP, diabetic retinopathy has a vaso-obliteration Recently microarray analyses substantiated a marked inflamma-
phase that leads to a proliferative phase. The vaso-obliteration tory response in the retinas of diabetic rodents.142 Confirmation
is not due to hyperoxia but rather to vaso-occlusion and vaso­ of the importance of inflammatory factors and growth factors is
degeneration of the microvasculature, setting up the ischemic supported by additional rodent studies that show that blocking
environment that leads to the vasoproliferative end-stage condi- these factors prevents the development of lesions characteristic
tion. Clinically the early pathology has been classified as non- of the retinopathy in animals. Specific inflammatory molecules
proliferative (microaneurysms, exudates, leakage, capillary that have been shown to contribute to structural or functional
nonperfusion) resulting in hypoxia and the end-stage pathology alterations that are characteristic of the retinopathy include
as proliferative (preretinal neovascularization). The purely vaso- NF-κβ143; inducible nitric oxide synthase143; cytochrome c
degenerative, nonproliferative form of the disease is by far the oxidase143; ICAM140; 5-lipoxygenase144; interleukin-1β145; tumor
most common and represents a disease of the neurovascular necrosis factor (TNF)-α146; and VEGF.67,147 Inflammation can
unit, resulting in dysfunction and eventual death of several of intensify the generation of AGEs that are produced in response
the key cells that maintain the BRB: pericytes, vascular endothe- to hyperglycemia and increased oxidative stress.
lial cells, Müller glia and neurons. Kohner and Henkind ele-
gantly demonstrated that, in diabetic individuals, areas of Retinal vein occlusion (RVO)
nonperfusion on fluorescein angiography are associated with Occlusion of large retinal blood vessels is a common occurrence,
acellular capillaries in trypsin digests.123 Direct measurement of which results in retinal hypoxia. RVO is the second most
oxygen in diabetic cat retinas demonstrated that even small common sight-threatening retinal vascular disorder after dia-
aneurysms can result in a decrease in retinal interstitial oxygen.124 betic retinopathy.148 RVO represents an obstruction of the retinal
There are many mechanisms implicated in the pathogenesis venous system that involves either the central retinal vein or
of diabetic retinopathy but one that has gained considerable a branch retinal vein. RVO is typically due to external compres-
attention in the last decade is inflammation. The environment sion or disease of the vein wall, such as is seen in vasculitis.149
of hyperglycemia, abnormal lipids, increased oxidative stress, Central retinal artery occlusion (CRAO) results in sudden, cata-
elevated serum and tissue advanced glycation endproducts strophic visual loss and branch retinal arteriolar occlusion
(AGE)/receptor for AGE, increased serum/tissue cytokines, (BRAO) causes sudden segmental visual loss and may recur to
elevated blood pressure, and endoplasmic reticulum stress are involve other branch retinal arterioles. CRAO studies have
the likely initiators of inflammation.125 Pathways of inflamma- shown that the ischemic retinal whitish opacity and swelling
tion converge with pathways of endothelial dysfunction and of CRAO are essentially located in the perifoveolar region of
coagulation to accelerate the pathogenesis of this disease.126 the macula. Oxygen supply and nutrition from the choroidal
Initially nonspecific indicators of inflammation such as white- vascular bed to the thinner peripheral retina help in its much
cell count and fibrinogen were found to be predictive of incident longer survival and the maintenance of peripheral visual fields.
diabetes.127 Subsequently plasminogen activator inhibitor-1 The diagnosis is clinical and based on the observation of the
(PAI-1), C-reactive protein, and fibrinogen were shown to be ocular fundus: venous dilatation and tortuosity, flame-shaped
retinal hemorrhages, retinal edema and cotton-wool exudates It is also important to consider that fibrinolytic agents can dis-
affecting all the retinal sectors (in CRVO) or the sector of the solve only platelet fibrin emboli.157 Retinal emboli are made of
retina drained by the affected vein in BRVO. Open-angle glau- 74% cholesterol, 10.5% calcific material, and only 15.5% of plate- 441
coma is the most frequent local alteration predisposing to RVO let fibrin. Fibrinolytic agents cannot dissolve cholesterol or calci-
as it compromises venous outflow by increasing intraocular fied material. Therefore, there is no scientific rationale for the use

Chapter 18
pressure. Raised intraocular pressure causes external compres- of fibrinolytic agents in at least 85% of CRAO cases. For the
sion of the central retinal vein as it passes through the lamina remaining 15%, if the diagnosis is made within 15 days from
cribrosa, resulting in turbulent blood flow distal to the compres- onset of clinical manifestations, low-molecular-weight heparins
sion leading to thrombus formation. at anticoagulant doses are typically used 10–15 days followed by
The natural history of RVO is highly variable; in some cases half dose for a total of 90 days.158 No data are available on the
the retinal findings progressively disappear and there is a possible role of antithrombotic/antiplatelet strategies in the
good visual outcome, while in other cases severe complications long-term prevention of recurrent RVO159; however, they are

Retinal and Choroidal Vasculature:


like ocular neovascularization (proliferative retinopathy), vitre- routinely prescribed to most patients with diabetes, hyperten-
ous hemorrhage, neovascular glaucoma, and macular edema sion, or arterial disease. Of Virchow’s three classical factors that
develop. Using retinal oximetry in BRVO, venular saturation play a role in thrombogenesis – stasis, vessel wall damage, and
was found to be highly variable between patients (12–93%)150; hypercoagulability – the first two have long been reported in
this was attributed to variable severity of the disease, recana- patients with RVO, whereas the third has not been sufficiently
lization, degree of occlusion, collateral vasculature, or tissue investigated until recently.
atrophy. The majority of patients with CRVO have signs of
macular edema at presentation whereas only 5–15% of eyes Sickle-cell disease (SCD)
with BRVO develop macular edema over the first year. Venous The pathological processes involved in SCD can affect virtually
collateral channels represent tortuous vessels that develop every vascular bed including the retina and in its advanced
locally, mainly around the optic disc, and are usually associ- stages has the potential to cause blindness.160 Classification of
ated with a long-standing vascular obstruction. Vitreous hem- ophthalmic manifestations of SCD in the retina is based on the
orrhage develops in 10% of eyes with CRVO within 9 months presence or absence of vascular proliferation, which is the most
of presentation and in about 40% of eyes with BRVO.148 If important precursor of blinding complications and precedes
there is restoration of circulation in the central retinal artery, development of a vitreous hemorrhage or retinal detachment.
the retinal capillaries in the central, thickest part of the macular Ischemia occurs due to obstruction of capillaries with sickle red
region do not refill because of compression by the surround- blood cells and subsequent thrombus formation. The subsequent
ing swollen superficial retinal tissue, resulting in the “no-reflow scenario for disease pathogenesis is similar to that for ROP,
phenomenon,”151and consequently in permanent ganglion cell diabetic retinopathy, and RVO, described above, except that
death in the nonperfused retina; the area of central retinal there is no retinal leakage of retinal blood vessels only from
capillary nonfilling may vary from eye to eye depending upon preretinal neovascularization, even though there is loss of
the severity of retinal swelling in the macular region. This peripheral vasculature and elevation of VEGF.161 Neovascular-
results in the variable size of the permanent central scotoma. ization that occurs can lead to fibrosis that can then result in
In considering the outcome of CRAO, it is important to con- retinal detachment.
sider retinal tolerance time to acute retinal ischemia. The chance
of recovery of vision only exists as long as the retina has Ocular ischemic syndrome (OIS)
reversible ischemic damage. The retina suffers no detectable OIS occurs at a mean age of 65 years and is rare before the
damage with CRAO of up to 97 minutes, but after that, the age of 50. Men are affected twice as often as women,162 reflect-
longer the CRAO, the more extensive the irreversible ischemic ing their higher incidence of atherosclerotic disease; however,
retinal damage. no racial predilection exists. Bilateral involvement may occur
It is generally believed that the CRAO is always either embolic in up to 22% of cases.163,164 Sturrock and Mueller estimated 7.5
or thrombotic in origin. Embolism is far more common than cases per million persons every year,165 but this is likely an
thrombosis,152 as was pointed out almost a century ago by underestimation as OIS can be easily misdiagnosed. Kearns166
Coats.153,154 An inflammatory etiology is also postulated in RVO reported that, of patients with occlusion of the internal carotid
as RVO is associated with immunological diseases in young artery undergoing surgical anastomosis between the superficial
patients. To support this contention, patients can experience temporal artery and the middle cerebral artery, 18% presented
prompt resolution of symptoms with the use of periocular ste- with OIS. Up to 29% of patients with a symptomatic carotid
roids.155 Giant cell arteritis is an important and well-known cause artery occlusion manifest retinal vascular changes that are
of CRAO, and is an ophthalmic emergency because of the high usually asymptomatic; however, 1.5% of them progress per year
risk of bilateral visual loss, which is preventable. to symptomatic OIS.167 OIS develops especially in patients with
Elevated levels of PAI-1, lipoprotein(a), and hyperhomocyste- poor collateral circulation between the internal and external
inemia, and low circulating levels of folic acid, vitamin B12 carotid arterial systems. Insufficient collateral vascular flow in
and vitamin B6 have been implicated in the pathogenesis of OIS patients explains the frequent association with cerebral
this disease.156 While the pathogenesis is complex and largely infarctions and the poor neurologic outcomes.168 Degree of inter-
unknown, medical treatment includes identification and correc- nal carotid artery stenosis, presence of collateral vessels, and
tion of vascular risk factors. The use of fluorescein fundus angi- compensation by collaterals are important in assessing OIS
ography before (showing occlusion of the central retinal artery) disease severity. Also if the OIS is bilateral or there are associ-
and immediately after thrombolysis may show improvement in ated systemic vascular diseases, this tends to worsen the
and/or restoration of retinal circulation and retinal function. prognosis.
GA: Nonatrophic GA: Atrophic Wet AMD

442
Section 1

A D G
Anatomy and Physiology
Basic Science and Translation to Therapy

B E H

c c c
C F I

Fig. 18.5 The choroid and retinal pigment epithelium (RPE) in geographic atrophy (A-F) and wet (neovascular) age-related macular degeneration
(AMD) (G-I). (A) Alkaline phosphatase (APase)-stained choroidal blood vessels in a nonatrophic region of geographic atrophy (GA) choroid.
(B) When sectioned, this area has normal-appearing RPE cells (arrowhead) and viable choriocapillaris lumens filled with serum APase.
(C) Higher magnification shows the intimate relationship between RPE, Bruch’s membrane, and choriocapillaris (c). (D) The atrophic region of
this GA choroid has no RPE and an attenuated choriocapillaris with narrow lumens. (E) Cross-sections of this area demonstrate the APase
reaction product in a few remaining choriocapillaris lumens and the endothelial cells of artery (bottom). (F) At higher magnification it is apparent
that the APase– capillaries (c) are only collagenous tubes between intercapillary septa. (G) In a flat choroid preparation of a wet AMD subject,
a large fan-shaped choroidal neovascular formation is present and the RPE monolayer to the left appears normal. (H) A section of this
subject immediately in advance of the choroidal neovascularization demonstrates viable hypertrophic RPE (arrowhead) over an attenuated
choriocapillaris. (I) At higher magnification, the remnants of atrophic choriocapillaris lumens (c) are present between intercapillary septa and
a single APase+ lumen remains, yet RPE are still present.

Retinal detachment center of our understanding of ocular angiogenesis. Retinal


neovascularization is defined as a state where new pathologic
Retinal detachment causes the sensory retina to be distant from
vessels originate from the existing retinal veins and extend
choriocapillaris, thus reducing its oxygen supply and resulting
along the inner surface of the retina.
in photoreceptor degeneration. Linsenmeier and Padnick-
Silver169 demonstrated in cat that retinal detachment resulted in Vascular permeability
a significant decrease in outer retinal oxygen, having a serious
Growth factors such as VEGF have been implicated in both
metabolic effect on photoreceptors. In subsequent work, Linsen-
retinal neovascularization and vascular hyperpermeability.173
meier demonstrated that hyperoxia may have clinical benefit
Antibodies to VEGF improve visual function in patients with
after retinal detachment because it normalized photoreceptor
diabetic macular edema.174 In experimental diabetes and in
oxygen consumption and prevented photoreceptor dysfunc-
VEGF-induced permeability, alterations of the tight junction (TJ)
tion.170 Furthermore, hyperoxia prevents proliferation and reac-
complex of microvascular endothelial cells alters the BRB.175
tivity of retinal Müller cells in the detached feline retina, limiting
A role of classical PKC isoforms (cPKCs) but especially PKCβ
retinal injury.171
in regulating VEGF-induced vascular permeability is well
accepted.176 VEGF activation of PKCβ leads to phosphorylation
Consequences of retinal ischemia and reorganization of the TJ complex, increasing vessel wall
In 1971, Judah Folkman reported in the New England Journal permeability.177 VEGF increases the phosphorylation of the
of Medicine that all cancer tumors are angiogenesis-dependent.172 TJ protein, occludin, at multiple sites,178 including Ser490.179
If a tumor could be stopped from growing its own blood Phosphorylation at Ser490 allows subsequent ubiquitination
supply, he surmised, it would wither and die. Though his and endocytosis of occludin and fosters breaks in the TJ.180
hypothesis was initially disregarded by most experts in the Despite this well-supported mechanism, the PKCβ inhibitor
field, Folkman persisted with his research. After more than a ruboxistaurin failed to achieve Food and Drug Administration
decade, his theory became widely accepted and is now at the approval for diabetic retinopathy. cPKC inhibition also failed to
prevent TNF-α-induced permeability,181 a proinflammatory REFERENCES
cytokine also implicated in diabetic retinopathy. Thus targeting 1. Wangsa-Wirawan ND, Linsenmeier RA, Retinal oxygen. Fundamental and
clinical aspects. Arch Ophthalmol 2003;121:547–57.
permeability in the retina continues to be a difficult clinical 2. Yu DY, Cringle SJ. Oxygen distribution and consumption within the retina in 443
problem. vascularised and avascular retinas and in animal models of retinal disease.
Prog Retin Eye Res 2001;20:175–208.
3. Chan-Ling T, McLeod DS, Hughes S, et al. Astrocyte–endothelial cell relation-
ADULT CHOROIDAL ISCHEMIA

Chapter 18
ships during human retinal vascular development. Invest Ophthalmol Vis Sci
2004;45:2020–32.
The choroidal vasculature is an expansive vascular plexus that 4. Hasegawa T, McLeod DS, Prow T, et al. Vascular precursors in developing
human retina. Invest Ophthalmol Vis Sci 2008;49:2178–92.
handles 90% of the blood from the ophthalmic artery. It was 5. McLeod DS, Hasegawa T, Prow T, et al. The initial fetal human retinal vas-
often assumed that the choriocapillaris was so vast that genera- culature develops by vasculogenesis. Dev Dyn 2006;235:3336–47.
tion of choroidal ischemia rarely occurred. However, there are 6. Stone J, Itin A, Alon T, et al. Development of retinal vasculature is mediated
by hypoxia-induced vascular endothelial growth factor (VEGF) expression by
several diseases in which the loss of choroidal vasculature is neuroglia. J Neurosci 1995;15:4738–47.

Retinal and Choroidal Vasculature:


extensive, resulting in apparent choroidal ischemia. 7. Chan-Ling T, Gock B, Stone J. The effect of oxygen on vasoformative cell
division: Evidence that ‘physiological hypoxia’ is the stimulus for normal
Diabetic choroidopathy was first described by Hidiyat and retinal vasculogenesis. Invest Ophthalmol Vis Sci 1995;36:1201–14.
Fine.182 The use of alkaline phosphatase enzyme histochemistry 8. Cogan DG, Kuwabara T. The mural cell in perspective. Arch Ophthalmol
1967;78:133–9.
on human choroid permitted quantification of vascular loss in 9. Hasegawa T, McLeod DS, Bhutto IA, et al. The embryonic human choriocapil-
choroid. Viable vessels were positive for alkaline phosphatase laris develops by hemo-vasculogenesis. Dev Dyn 2007;236:2089–100.
10. McLeod DS, Lutty GA. High resolution histologic analysis of the human
while acellular, dysfunctional capillaries lacked it, and choroidal choroidal vasculature. Invest Ophthalmol Vis Sci 1994;35:3799–811.
neovascularization had the greatest activity.10 This technique 11. Blaauwgeers HG, Holtkamp GM, Rutten H, et al. Polarized vascular
demonstrated that there was four times greater loss in chorio- endothelial growth factor secretion by human retinal pigment epithelium and
localization of vascular endothelial growth factor receptors on the inner cho-
capillaris in diabetic subjects than in aged control subjects.183 riocapillaris. Evidence for a trophic paracrine relation. Am J Pathol 1999;
Loss of alkaline phosphatase activity was associated with the 155:421–8.
12. Lutjen-Drecoll E. Choroidal innervation in primate eyes. Exp Eye Res 2006;82:
presence of polymorphonuclear leukocytes.184 Areas of chorio- 357–61.
capillaris loss were also associated with choroidal neovascular- 13. Michaelson IC. The mode of development of the vascular system of the retina,
with some observations on its significance for certain retinal diseases.
ization and Bruch’s membrane deposits. A consequence of Trans Ophthalmol Soc UK 1948;68:137–80.
choriocapillaris dropout is loss of outer retinal oxygenation. This 14. Michaelson IC. Retinal circulation in man and animals. Springfield: CC
may be the reason for loss of blue cones and other cones in Thomas; 1954.
15. Wise GN. Retinal neovascularization. Trans Am Ophthalmol Soc 1956;96:
diabetic retina when there is no retinopathy present.185,186 729–826.
Decreased choroidal blood flow has been measured in 16. Leung DW, Cachianes G, Kuang WJ, et al. Vascular endothelial growth factor
is a secreted angiogenic mitogen. Science 1989;246:1306–9.
AMD.187–190 Using alkaline phosphatase activity and quantifica- 17. Senger DR, Galli SJ, Dvorak AM, et al. Tumor cells secrete a vascular perme-
tion of RPE in AMD, the loss of choriocapillaris has been dem- ability factor that promotes accumulation of ascites fluid. Science 1983;219:
983–5.
onstrated in exudative or wet AMD as well as geographic 18. Yu DY, Cringle SJ, Su EN. Intraretinal oxygen distribution in the monkey
atrophy (Fig. 18.5). RPE loss occurs first in geographic atrophy retina and the response to systemic hyperoxia. Invest Ophthalmol Vis Sci
2005;46:4728–33.
and then choriocapillaris degeneration occurs. Although some 19. McLeod DS, Brownstein R, Lutty GA. Vaso-obliteration in the canine model
capillaries survive in the area of RPE atrophy, a 50% loss of of oxygen-induced retinopathy. Invest Ophthalmol Vis Sci 1996;37:300–11.
vasculature occurs and the surviving capillaries are highly con- 20. Ernest JT, Goldstick TK. Retinal oxygen tension and oxygen reactivity in reti-
nopathy of prematurity in kittens. Invest Ophthalmol Vis Sci 1984;25:
stricted.191,192 This undoubtedly contributes to photoreceptor loss 1129–34.
in the area of degeneration. Capillary loss occurs in exudative 21. Hardy P, Peri KG, Lahaie I, et al. Increased nitric oxide synthesis and action
preclude choroidal vasoconstriction to hyperoxia in newborn pigs. Circ Res
AMD as well but in the presence of a complete RPE monolayer 1996;79:504–11.
(Fig. 18.5). This suggests that the RPE in this area are hypoxic 22. Linsenmeier RA, Yancy CM. Effects of hyperoxia on the oxygen distribution
in the intact cat retina. Invest Ophthalmol Vis Sci 1989;30:612–8.
and VEGF production may be increased, causing choroidal neo- 23. Pournaras CJ, Riva CE, Tsacopoulos M, et al. Diffusion of O2 in the retina of
vascularization. Large-vessel stenosis is also often observed in anesthetized miniature pigs in normoxia and hyperoxia. Exp Eye Res 1989;
AMD choroid, suggesting that the choriocapillaris blood supply 49:347–60.
24. Benizri E, Ginouves A, Berra E. The magic of the hypoxia-signaling cascade.
may be limited as well. Unfortunately, we have no way of mea- Cell Mol Life Sci 2008;65:1133–49.
suring oxygen directly in the choroid so we can only assume 25. Semenza GL, Wang GL. A nuclear factor induced by hypoxia via de novo
protein synthesis binds to the human erythropoietin gene enhancer at a site
ischemia occurs when there is loss of viable choriocapillaris and required for transcriptional activation. Mol Cell Biol 1992;12:5447–54.
stenosis of large and intermediate choroidal blood vessel. 26. Wang GL, Jiang BH, Rue EA, et al. Hypoxia-inducible factor 1 is a basic-helix-
loop-helix-PAS heterodimer regulated by cellular O2 tension. Proc Natl Acad
Sci U S A 1995;92:5510–4.
CONCLUSIONS 27. Gariboldi MB, Ravizza R, Monti E. The IGFR1 inhibitor NVP-AEW541 dis-
rupts a pro-survival and pro-angiogenic IGF-STAT3-HIF1 pathway in human
Both too little and too much oxygen can be damaging to the glioblastoma cells. Biochem Pharmacol 2010;80:455–62.
28. Jiang BH, Zheng JZ, Leung SW, et al. Transactivation and inhibitory domains
retina. Retina cannot function properly nor survive without the of hypoxia-inducible factor 1alpha. Modulation of transcriptional activity by
support of two vasculatures: retinal and choroidal. If either vas- oxygen tension. J Biol Chem 1997;272:19253–60.
29. Semenza GL. Hydroxylation of HIF-1: oxygen sensing at the molecular level.
culature is dysfunctional, retinal hypoxia occurs and HIF-1α Physiol (Bethesda) 2004;19:176–82.
becomes stable and induces expression of key factors to assist 30. Mahon PC, Hirota K, Semenza GL. FIH-1: a novel protein that interacts with
in retinal cell survival. Reduced retinal oxygenation causes HIF-1alpha and VHL to mediate repression of HIF-1 transcriptional activity.
Genes Dev 2001;15:2675–86.
impaired neuroretinal activity in young healthy persons.193 One 31. Lando D, Peet DJ, Whelan DA, et al. Asparagine hydroxylation of the HIF
key factor upregulated by HIF-1α is VEGF, which stimulates transactivation domain a hypoxic switch. Science 2002;295:858–61.
32. Lee C, Kim SJ, Jeong DG, et al. Structure of human FIH-1 reveals a unique
new angiogenesis and increased vascular permeability. These active site pocket and interaction sites for HIF-1 and von Hippel–Lindau.
pathological events can be controlled by destroying ischemic J Biol Chem 2003;278:7558–63.
33. Kukkola L, Koivunen P, Pakkanen O, et al. Collagen prolyl 4-hydroxylase
retina (photocoagulation), inhibiting VEGF, or reducing levels of tetramers and dimers show identical decreases in Km values for peptide sub-
HIF-1α. strates with increasing chain length: mutation of one of the two catalytic sites
in the tetramer inactivates the enzyme by more than half. J Biol Chem 64. Saint-Geniez M, Kurihara T, Sekiyama E, et al. An essential role for RPE-
2004;279:18656–61. derived soluble VEGF in the maintenance of the choriocapillaris. Proc Natl
34. van Uden P, Kenneth NS, Rocha S. Regulation of hypoxia-inducible factor- Acad Sci U S A 2009;106:18751–6.
444 1alpha by NF-kappaB. Biochem J 2008;412:477–84. 65. Nishijima K, Ng YS, Zhong L, et al. Vascular endothelial growth factor-A is
35. Wang GL, Semenza GL. Desferrioxamine induces erythropoietin gene expres- a survival factor for retinal neurons and a critical neuroprotectant during the
sion and hypoxia-inducible factor 1 DNA-binding activity: implications for adaptive response to ischemic injury. Am J Pathol 2007;171:53–67.
models of hypoxia signal transduction. Blood 1993;82:3610–5. 66. Ng EW, Adamis AP. Targeting angiogenesis, the underlying disorder in neo-
36. Tian H, McKnight SL, Russell DW. Endothelial PAS domain protein 1 vascular age-related macular degeneration. Can J Ophthalmol 2005;40:
Section 1

(EPAS1), a transcription factor selectively expressed in endothelial cells. 352–68.


Genes Dev 1997;11:72–82. 67. Kunz Mathews M, Merges C, McLeod DS, et al. Vascular endothelial growth
37. Sowter HM, Raval RR, Moore JW, et al. Predominant role of hypoxia-inducible factor (VEGF) and vascular permeability changes in human diabetic retinopa-
transcription factor (Hif)-1alpha versus Hif-2alpha in regulation of the tran- thy. Invest Ophthalmol Vis Sci 1997;38:2729–41.
scriptional response to hypoxia. Cancer Res 2003;63:6130–4. 68. Starita C, Patel M, Katz B, et al. Vascular endothelial growth factor and the
38. Park SK, Dadak AM, Haase VH, et al. Hypoxia-induced gene expression potential therapeutic use of pegaptanib (Macugen) in diabetic retinopathy.
occurs solely through the action of hypoxia-inducible factor 1alpha (HIF- Dev Ophthalmol 2007;39:122–48.
1alpha): role of cytoplasmic trapping of HIF-2alpha. Mol Cell Biol 2003;23: 69. Aiello LP, Cahill MT, Cavallerano JD. Growth factors and protein kinase C
Anatomy and Physiology

4959–71. inhibitors as novel therapies for the medical management diabetic retinopa-
Basic Science and Translation to Therapy

39. Makino Y, Cao R, Svensson K, et al. Inhibitory PAS domain protein is a nega- thy. Eye (Lond) 2004;18:117–25.
tive regulator of hypoxia-inducible gene expression. Nature 2001;414:550–4. 70. Tripathi RC, Li J, Tripathi BJ, et al. Increased level of vascular endothelial
40. Compernolle V, Brusselmans K, Franco D, et al. Cardia bifida, defective heart growth factor in aqueous humor of patients with neovascular glaucoma.
development and abnormal neural crest migration in embryos lacking Ophthalmology 1998;105:232–7.
hypoxia-inducible factor-1alpha. Cardiovasc Res 2003;60:569–79. 71. Lashkari K, Hirose T, Yazdany J, et al. Vascular endothelial growth factor
41. Peng J, Zhang L, Drysdale L, et al. The transcription factor EPAS-1/hypoxia- and hepatocyte growth factor levels are differentially elevated in patients
inducible factor 2alpha plays an important role in vascular remodeling. with advanced retinopathy of prematurity. Am J Pathol 2000;156:
Proc Natl Acad Sci U S A 2000;97:8386–91. 1337–44.
42. Scortegagna M, Morris MA, Oktay Y, et al. The HIF family member EPAS1/ 72. Chakravarthy U. Age related macular degeneration. Br Med J 2006;333:
HIF-2alpha is required for normal hematopoiesis in mice. Blood 2003;102: 869–70.
1634–40. 73. Rosenfeld PJ, Brown DM, Heier JS, et al. Ranibizumab for neovascular age-
43. Kline DD, Peng YJ, Manalo DJ, et al. Defective carotid body function and related macular degeneration. N Engl J Med 2006;355:1419–31.
impaired ventilatory responses to chronic hypoxia in mice partially deficient 74. Miller-Kasprzak E, Jagodzinski PP. 5-Aza-2’-deoxycytidine increases the
for hypoxia-inducible factor 1 alpha. Proc Natl Acad Sci U S A 2002;99: expression of anti-angiogenic vascular endothelial growth factor 189b variant
821–6. in human lung microvascular endothelial cells. Biomed Pharmacother 2008;
44. Caprara C, Thiersch M, Lange C, et al. HIF1A is essential for the development 62:158–63.
of the intermediate plexus of the retinal vasculature. Invest Ophthalmol Vis 75. Anthony FW, Wheeler T, Elcock CL, et al. Short report: identification of
Sci 2011; 52:2109–17. a specific pattern of vascular endothelial growth factor mRNA expression
45. Lin M, Chen Y, Jin J, et al. Ischaemia-induced retinal neovascularisation and in human placenta and cultured placental fibroblasts. Placenta 1994;15:
diabetic retinopathy in mice with conditional knockout of hypoxia-inducible 557–61.
factor-1 in retinal Muller cells. Diabetologia 2011;54:1554–66. 76. Woolard J, Bevan HS, Harper SJ, et al. Molecular diversity of VEGF-A as a
46. Cramer T, Yamanishi Y, Clausen BE, et al. HIF-1alpha is essential for myeloid regulator of its biological activity. Microcirculation 2009;16:572–92.
cell-mediated inflammation. Cell 2003;112:645–57. 77. de Vries C, Escobedo JA, Ueno H, et al. The fms-like tyrosine kinase, a receptor
47. Kojima H, Gu H, Nomura S, et al. Abnormal B lymphocyte development and for vascular endothelial growth factor. Science 1992;255:989–91.
autoimmunity in hypoxia-inducible factor 1alpha -deficient chimeric mice. 78. Quinn TP, Peters KG, De Vries C, et al. Fetal liver kinase 1 is a receptor for
Proc Natl Acad Sci U S A 2002;99:2170–4. vascular endothelial growth factor and is selectively expressed in vascular
48. Makino Y, Nakamura H, Ikeda E, et al. Hypoxia-inducible factor regulates endothelium. Proc Natl Acad Sci U S A 1993;90:7533–7.
survival of antigen receptor-driven T cells. J Immunol 2003;171:6534–40. 79. Soker S, Takashima S, Miao HQ, et al. Neuropilin-1 is expressed by endothe-
49. Giaccia A, Siim BG, Johnson RS. HIF-1 as a target for drug development. lial and tumor cells as an isoform-specific receptor for vascular endothelial
Nat Rev Drug Discov 2003;2:803–11. growth factor. Cell 1998;92:735–45.
50. Kelly BD, Hackett SF, Hirota K, et al. Cell type-specific regulation of angio- 80. Schlessinger J. New roles for Src kinases in control of cell survival and angio-
genic growth factor gene expression and induction of angiogenesis in non­ genesis. Cell 2000;100:293–6.
ischemic tissue by a constitutively active form of hypoxia-inducible factor 1. 81. Hubbard SR. Structural analysis of receptor tyrosine kinases. Prog Biophys
Circ Res 2003;93:1074–81. Mol Biol 1999;71:343–58.
51. Hirsila M, Koivunen P, Gunzler V, et al. Characterization of the human prolyl 82. Shalaby F, Rossant J, Yamaguchi TP, et al. Failure of blood-island formation
4-hydroxylases that modify the hypoxia-inducible factor. J Biol Chem 2003; and vasculogenesis in Flk-1-deficient mice. Nature 1995;376:62–6.
278:30772–80. 83. Scott A, Mellor H. VEGF receptor trafficking in angiogenesis. Biochem Soc
52. Willam C, Masson N, Tian YM, et al. Peptide blockade of HIFalpha degrada- Trans 2009;37:1184–8.
tion modulates cellular metabolism and angiogenesis. Proc Natl Acad Sci 84. Jopling HM, Odell AF, Hooper NM, et al. Rab GTPase regulation of VEGFR2
U S A 2002;99:10423–8. trafficking and signaling in endothelial cells. Arterioscler Thromb Vasc Biol
53. Chen J, Smith LE. A double-edged sword: erythropoietin eyed in retinopathy 2009;29:1119–24.
of prematurity. J AAPOS 2008;12:221–2. 85. Salikhova A, Wang L, Lanahan AA, et al. Vascular endothelial growth factor
54. Ferrara N. Vascular endothelial growth factor as a target for anticancer and semaphorin induce neuropilin-1 endocytosis via separate pathways.
therapy. Oncologist 2004;9(Suppl 1):2–10. Circ Res 2008;103:e71–79.
55. Carmeliet P, Ferreira V, Breier G, et al. Abnormal blood vessel development 86. Ewan LC, Jopling HM, Jia H, et al. Intrinsic tyrosine kinase activity is required
and lethality in embryos lacking a single VEGF allele. Nature 1996;380: for vascular endothelial growth factor receptor 2 ubiquitination, sorting and
435–9. degradation in endothelial cells. Traffic 2006;7:1270–82.
56. Alon T, Hemo I, Itin A, Pe’er J, et al. Vascular endothelial growth factor acts 87. Lanahan AA, Hermans K, Claes F, et al. VEGF receptor 2 endocytic trafficking
as a survival factor for newly formed retinal vessels and has implications for regulates arterial morphogenesis. Dev Cell 2010;18:713–24.
retinopathy of prematurity. Nat Med 1995;1:1024–8. 88. Cai J, Jiang WG, Ahmed A, et al. Vascular endothelial growth factor-induced
57. Baffert F, Le T, Thurston G, et al. Angiopoietin-1 decreases plasma leakage by endothelial cell proliferation is regulated by interaction between VEGFR-2,
reducing number and size of endothelial gaps in venules. Am J Physiol Heart SH-PTP1 and eNOS. Microvasc Res 2006;71:20–31.
Circ Physiol 2006;290:H107–118. 89. Chen J, Braet F, Brodsky S, et al. VEGF-induced mobilization of caveolae and
58. Saint-Geniez M, Maharaj AS, Walshe TE, et al. Endogenous VEGF is required increase in permeability of endothelial cells. Am J Physiol Cell Physiol
for visual function: evidence for a survival role on muller cells and photore- 2002;282:C1053–1063.
ceptors. PLoS ONE 2008;3:e3554. 90. Grant MB, Boulton ME, Ljubimov AV. Erythropoietin: when liability becomes
59. Adamis AP, Shima DT, Yeo KT, et al. Synthesis and secretion of vascular asset in neurovascular repair. J Clin Invest 2008;118:467–70.
permeability factor/vascular endothelial growth factor by human retinal 91. Takahashi T, Ueno H, Shibuya M. VEGF activates protein kinase C-dependent,
pigment epithelial cells. Biochem Biophys Res Commun 1993;193:631–8. but Ras-independent Raf-MEK-MAP kinase pathway for DNA synthesis in
60. Aiello LP, Northrup JM, Keyt BA, et al. Hypoxic regulation of vascular endo- primary endothelial cells. Oncogene 1999;18:2221–30.
thelial growth factor in retinal cells. Arch Ophthalmol 1995;113:1538–44. 92. Fadini GP, Miorin M, Facco M, et al. Circulating endothelial progenitor cells
61. Famiglietti EV, Stopa EG, McGookin ED, et al. Immunocytochemical localiza- are reduced in peripheral vascular complications of type 2 diabetes mellitus.
tion of vascular endothelial growth factor in neurons and glial cells of human J Am Coll Cardiol 2005;45:1449–57.
retina. Brain Res 2003;969:195–204. 93. Tepper OM, Galiano RD, Capla JM, et al. Human endothelial progenitor cells
62. Sandercoe TM, Geller SF, Hendrickson AE, et al. VEGF expression by gan- from type II diabetics exhibit impaired proliferation, adhesion, and incorpora-
glion cells in central retina before formation of the foveal depression in tion into vascular structures. Circulation 2002;106:2781–6.
monkey retina: evidence of developmental hypoxia. J Comp Neurol 2003;462: 94. Loomans CJ, de Koning EJ, Staal FJ, et al. Endothelial progenitor cell dysfunc-
42–54. tion: a novel concept in the pathogenesis of vascular complications of type 1
63. Marneros AG, Fan J, Yokoyama Y, et al. Vascular endothelial growth factor diabetes. Diabetes 2004;53:195–9.
expression in the retinal pigment epithelium is essential for choriocapillaris 95. Schatteman GC, Hanlon HD, Jiao C, et al. Blood-derived angioblasts acceler-
development and visual function. Am J Pathol 2005;167:1451–9. ate blood-flow restoration in diabetic mice. J Clin Invest 2000;106:571–8.
96. Caballero S, Sengupta N, Afzal A, et al. Ischemic vascular damage can be 125. Tang J, Kern TS. Inflammation in diabetic retinopathy. Prog Retin Eye Res
repaired by healthy, but not diabetic, endothelial progenitor cells. Diabetes 2011; 30:343–58.
2007;56:960–7. 126. Kern TS. Contributions of inflammatory processes to the development of the
97. Jarajapu YP, Caballero S, Verma A, et al. Blockade of NADPH oxidase restores early stages of diabetic retinopathy. Exp Diabetes Res 2007;2007:95103. 445
vasoreparative function in diabetic CD34+ cells. Invest Ophthalmol Vis Sci 127. Schmidt MI, Duncan BB, Sharrett AR, et al. Markers of inflammation and
2011;52:5093–104. prediction of diabetes mellitus in adults (Atherosclerosis Risk in Communities
98. Case J, Ingram DA, Haneline LS. Oxidative stress impairs endothelial progeni- study): a cohort study. Lancet 1999;353:1649–52.

Chapter 18
tor cell function. Antioxid Redox Signal 2008;10:1895–907. 128. Festa A, D’Agostino Jr R, Tracy RP, et al. Elevated levels of acute-phase pro-
99. Togliatto G, Trombetta A, Dentelli P, et al. Unacylated ghrelin rescues endo- teins and plasminogen activator inhibitor-1 predict the development of type
thelial progenitor cell function in individuals with type 2 diabetes. Diabetes 2 diabetes: the insulin resistance atherosclerosis study. Diabetes 2002;51:
2010;59:1016–25. 1131–7.
100. Ingram DA, Lien IZ, Mead LE, et al. In vitro hyperglycemia or a diabetic 129. Eliasson MC, Jansson JH, Lindahl B, et al. High levels of tissue plasminogen
intrauterine environment reduces neonatal endothelial colony-forming cell activator (tPA) antigen precede the development of type 2 diabetes in a
numbers and function. Diabetes 2008;57:724–31. longitudinal population study. The Northern Sweden MONICA study.
101. Urbich C, Dernbach E, Rossig L, et al. High glucose reduces cathepsin L activ- Cardiovasc Diabetol 2003;2:19.
ity and impairs invasion of circulating progenitor cells. J Mol Cell Cardiol 130. Meigs JB, Dupuis J, Liu C, et al. PAI-1 gene 4G/5G polymorphism and risk
2008;45:429–36. of type 2 diabetes in a population-based sample. Obesity (Silver Spring)

Retinal and Choroidal Vasculature:


102. Li SS, Liu Z, Uzunel M, et al. Endogenous thrombospondin-1 is a cell-surface 2006;14:753–8.
ligand for regulation of integrin-dependent T-lymphocyte adhesion. Blood 131. Orasanu G, Plutzky J. The pathologic continuum of diabetic vascular disease.
2006;108:3112–20. J Am Coll Cardiol 2009;53:S35–42.
103. Butler JM, Guthrie SM, Koc M, et al. SDF-1 is both necessary and sufficient to 132. van Hecke MV, Dekker JM, Nijpels G, et al. Inflammation and endothelial
promote proliferative retinopathy. J Clin Invest 2005;115:86–93. dysfunction are associated with retinopathy: the Hoorn study. Diabetologia
104. Segal MS, Shah R, Afzal A, et al. Nitric oxide cytoskeletal-induced alterations 2005;48:1300–6.
reverse the endothelial progenitor cell migratory defect associated with 133. Spijkerman AM, Gall MA, Tarnow L, et al. Endothelial dysfunction and low-
diabetes. Diabetes 2006;55:102–9. grade inflammation and the progression of retinopathy in type 2 diabetes.
105. Bhatwadekar AD, Glenn JV, Li G, et al. Advanced glycation of fibronectin Diabet Med 2007;24:969–76.
impairs vascular repair by endothelial progenitor cells: implications for vaso- 134. Hadjadj S, Aubert R, Fumeron F, et al. Increased plasma adiponectin concen-
degeneration in diabetic retinopathy. Invest Ophthalmol Vis Sci 2008;49: trations are associated with microangiopathy in type 1 diabetic subjects.
1232–41. Diabetologia 2005;48:1088–92.
106. Bhatwadekar AD, Guerin EP, Jarajapu YP, et al. Transient inhibition of trans- 135. McLeod DS, Lefer DJ, Merges C, et al. Enhanced expression of intracellular
forming growth factor-beta1 in human diabetic CD34+ cells enhances vascular adhesion molecule-1 and P-selectin in the diabetic human retina and choroid.
reparative functions. Diabetes 2010; 59:2010–9. Am J Pathol 1995;147:642–53.
107. Sorrentino SA, Bahlmann FH, Besler C, et al. Oxidant stress impairs in vivo 136. Joussen AM, Murata T, Tsujikawa A, et al. Leukocyte-mediated endothelial
reendothelialization capacity of endothelial progenitor cells from patients cell injury and death in the diabetic retina. Am J Pathol 2001;158:147–52.
with type 2 diabetes mellitus: restoration by the peroxisome proliferator- 137. Joussen AM, Poulaki V, Qin W, et al. Retinal vascular endothelial growth
activated receptor-gamma agonist rosiglitazone. Circulation 2007;116: factor induces intracellular adhesion molecule-1 and endothelial nitric oxude
163–73. synthase expression and initiates early diabetic retinal leukocyte adhesion in
108. Mohler 3rd ER, Shi Y, Moore J, et al. Diabetes reduces bone marrow and vivo. Am J Pathol 2002;160:501–9.
circulating porcine endothelial progenitor cells, an effect ameliorated by ator- 138. Joussen AM, Poulaki V, Mitsiades N, et al. Nonsteroidal anti-inflammatory
vastatin and independent of cholesterol. Cytometry A 2009;75:75–82. drugs prevent early diabetic retinopathy via TNF-α suppression. FASEB J
109. Anghelina M, Krishnan P, Moldovan L, et al. Monocytes/macrophages coop- 2002;16:438–40.
erate with progenitor cells during neovascularization and tissue repair: con- 139. Adamis AP. Is diabetic retinopathy an inflammatory disease? Br J Ophthalmol
version of cell columns into fibrovascular bundles. Am J Pathol 2006;168: 2002;86:363–5.
529–41. 140. Joussen AM, Poulaki V, Le ML, et al. A central role for inflammation in the
110. Chang KH, Chan-Ling T, McFarland EL, et al. IGF binding protein-3 regulates pathogenesis of diabetic retinopathy. Faseb J 2004;18:1450–2.
hematopoietic stem cell and endothelial precursor cell function during 141. Powell ED, Field RA. Diabetic retinopathy and rheumatoid arthritis. Lancet
vascular development. Proc Natl Acad Sci U S A 2007;104:10595–600. 1964;2:17–8.
111. Granata R, Trovato L, Garbarino G, et al. Dual effects of IGFBP-3 on endothe- 142. Brucklacher RM, Patel KM, VanGuilder HD, et al. Whole genome assessment
lial cell apoptosis and survival: involvement of the sphingolipid signaling of the retinal response to diabetes reveals a progressive neurovascular inflam-
pathways. FASEB J 2004;18:1456–8. matory response. BMC Med Genomics 2008;1:26.
112. Franklin SL, Ferry Jr RJ, Cohen P. Rapid insulin-like growth factor (IGF)- 143. Kern TS, Miller CM, Du Y, et al. Topical administration of nepafenac inhibits
independent effects of IGF binding protein-3 on endothelial cell survival. diabetes-induced retinal microvascular disease and underlying abnormalities
J Clin Endocrinol Metab 2003;88:900–7. of retinal metabolism and physiology. Diabetes 2007;56:373–9.
113. Liu LQ, Sposato M, Liu HY, et al. Functional cloning of IGFBP-3 from human 144. Gubitosi-Klug RA, Talahalli R, Du Y, et al. 5-Lipoxygenase, but not 12/15-
microvascular endothelial cells reveals its novel role in promoting prolifera- lipoxygenase, contributes to degeneration of retinal capillaries in a mouse
tion of primitive CD34+CD38– hematopoietic cells in vitro. Oncol Res model of diabetic retinopathy. Diabetes 2008;57:1387–93.
2003;13:359–71. 145. Vincent JA, Mohr S. Inhibition of caspase-1/interleukin-1beta signaling
114. Vasylyeva TL, Chen X, Ferry Jr RJ. Insulin-like growth factor binding protein-3 prevents degeneration of retinal capillaries in diabetes and galactosemia.
mediates cytokine-induced mesangial cell apoptosis. Growth Horm IGF Res Diabetes 2007;56:224–30.
2005;15:207–14. 146. Behl Y, Krothapalli P, Desta T, et al. Diabetes-enhanced tumor necrosis factor-
115. Lofqvist C, Chen J, Connor KM, et al. IGFBP3 suppresses retinopathy through alpha production promotes apoptosis and the loss of retinal microvascular
suppression of oxygen-induced vessel loss and promotion of vascular cells in type 1 and type 2 models of diabetic retinopathy. Am J Pathol
regrowth. Proc Natl Acad Sci U S A 2007;104:10589–94. 2008;172:1411–8.
116. Giannini S, Cresci B, Pala L, et al. IGFBPs modulate IGF-I- and high glucose- 147. Lutty GA, McLeod DS, Merges C, et al. Localization of VEGF in human retina
controlled growth of human retinal endothelial cells. J Endocrinol 2001;171: and choroid. Arch Ophthalmol 1996;114:971–7.
273–84. 148. Rogers SL, McIntosh RL, Lim L, et al. Natural history of branch retinal vein
117. Kielczewski JL, Hu P, Shaw LC, et al. Novel protective properties of IGFBP-3 occlusion: an evidence-based systematic review. Ophthalmology 2010;
result in enhanced pericyte ensheathment, reduced microglial activation, 117:1094–101 e1095.
increased microglial apoptosis, and neuronal protection after ischemic retinal 149. Laouri M, Chen E, Looman M, et al. The burden of disease of retinal vein
injury. Am J Pathol 2011; 178:1517–28. occlusion: review of the literature. Eye (Lond) 2011; 25(8):981–8.
118. Chan-Ling TL, Halasz P, Stone J. Development of retinal vasculature in the 150. Hardarson HS, Stefansson E. Oxygen saturation in branch vein occlusion.
cat: processes and mechanisms. Curr Eye Res 1990;9:459–78. Acta Ophthalmol 2011;Epub ahead of printing.
119. Hellstrom A, Engstrom E, Hard AL, et al. Postnatal serum insulin-like growth 151. Hayreh SS, Weingeist TA. Experimental occlusion of the central artery of the
factor I deficiency is associated with retinopathy of prematurity and other retina. IV: Retinal tolerance time to acute ischaemia. Br J Ophthalmol
complications of premature birth. Pediatrics 2003;112:1016–20. 1980;64:818–25.
120. Giudice LC, de Zegher F, Gargosky SE, et al. Insulin-like growth factors and 152. Hayreh SS, Podhajsky P. Ocular neovascularization with retinal vascular
their binding proteins in the term and preterm human fetus and neonate with occlusion. II. Occurrence in central and branch retinal artery occlusion.
normal and extremes of intrauterine growth. J Clin Endocrinol Metab Arch Ophthalmol 1982;100:1585–96.
1995;80:1548–55. 153. Coats G. Obstruction of the central artery of the retina. R. Lond Ophthalmic
121. Smith WJ, Underwood LE, Keyes L, et al. Use of insulin-like growth factor I Hosp Rep 1905;16:262–306.
(IGF-I) and IGF-binding protein measurements to monitor feeding of prema- 154. Coats G. Pathology of obstruction of the centeral artery of the retina. R. Lond
ture infants. J Clin Endocrinol Metab 1997;82:3982–8. Ophthalmic Hosp Rep 1913;19:45–70.
122. Smith LE. IGF-1 and retinopathy of prematurity in the preterm infant. 155. Fong AC, Schatz H, McDonald HR, et al. Central retinal vein occlusion in
Biol Neonate 2005;88:237–44. young adults (papillophlebitis). Retina 1992;12:3–11.
123. Kohner EM, Henkind P. Correlation of fluorescein angiogram and retinal 156. Sofi F, Marcucci R, Fedi S, et al. High lipoprotein (a) levels are associated with
digest in diabetic retinopathy. Am J Ophthalmol 1970;69:403–14. an increased risk of retinal vein occlusion. Atherosclerosis 2010; 210:278–81.
124. Linsenmeier RA, Braun RD, McRipley MA, et al. Retinal hypoxia in long term 157. Arruga J, Sanders MD. Ophthalmologic findings in 70 patients with evidence
diabetic cats. Invest Ophthalmol Vis Sci 1998;39:1647–57. of retinal embolism. Ophthalmology 1982;89:1336–47.
158. Ageno W, Cattaneo R, Manfredi E, et al. Parnaparin versus aspirin in the 177. Barber AJ, Antonetti DA, Gardner TW. Altered expression of retinal
treatment of retinal vein occlusion. A randomized, double blind, controlled occludin and glial fibrillary acidic protein in experimental diabetes. The
study. Thromb Res 2010; 125:137–41. Penn State Retina Research Group. Invest Ophthalmol Vis Sci 2000;41:
446 159. Marcucci R, Sofi F, Grifoni E, et al. Retinal vein occlusions: a review for the 3561–8.
internist. Intern Emerg Med 2011; 6:307–14. 178. Harhaj NS, Felinski EA, Wolpert EB, et al. VEGF activation of protein kinase
160. Nagpal KC, Goldberg MF, Rabb MF. Ocular manifestations of sickle hemo- C stimulates occludin phosphorylation and contributes to endothelial perme-
globinopathies. Surv Ophthalmol 1977;21:391–411. ability. Invest Ophthalmol Vis Sci 2006;47:5106–15.
161. Cao J, Kunz Mathews M, McLeod DS, et al. Angiogenic factors in human 179. Sundstrom JM, Sundstrom CJ, Sundstrom SA, et al. Phosphorylation site
Section 1

proliferative sickle cell retinopathy. Br J Ophthalmol 1999;83:838–46. mapping of endogenous proteins: a combined MS and bioinformatics
162. Brown GC, Magargal LE. The ocular ischemic syndrome. Clinical, fluorescein approach. J Proteome Res 2009;8:798–807.
angiographic and carotid angiographic features. Int Ophthalmol 1988;11: 180. Murakami T, Felinski EA, Antonetti DA. Occludin phosphorylation
239–51. and ubiquitination regulate tight junction trafficking and vascular
163. Boto de los Bueis A, Fernandez-Prieto A, Ruiz-Martin MM, et al. [Bilateral endothelial growth factor-induced permeability. J Biol Chem 2009;284:
carotid occlusion in young woman. Clinical and hemodynamic ocular results.] 21036–46.
Arch Soc Esp Oftalmol 2003;78:227–30. 181. Aveleira CA, Lin CM, Abcouwer SF, et al. TNF-alpha signals through
164. Baatz H, Lange S, Buchner H, et al. [Pseudoangiitis in bilateral ocular isch- PKCzeta/NF-kappaB to alter the tight junction complex and increase retinal
Anatomy and Physiology

emia.] Ophthalmologe 2007;104: 243–245. endothelial cell permeability. Diabetes 2010; 59:2872–82.
Basic Science and Translation to Therapy

165. Sturrock GD, Mueller HR. Chronic ocular ischaemia. Br J Ophthalmol 182. Hidayat A, Fine B. Diabetic choroidopathy: light and electron microscopic
1984;68:716–23. observations of seven cases. Ophthalmology 1985;67:512–22.
166. Kearns TP. Differential diagnosis of central retinal vein obstruction. 183. Cao J, McLeod S, Merges CA, et al. Choriocapillaris degeneration and related
Ophthalmology 1983;90:475–80. pathologic changes in human diabetic eyes. Arch Ophthalmol 1998;116:
167. Klijn CJ, Kappelle LJ, van Schooneveld MJ, et al. Venous stasis retinopathy in 589–97.
symptomatic carotid artery occlusion: prevalence, cause, and outcome. Stroke 184. Lutty GA, Cao J, McLeod DS. Relationship of polymorphonuclear leukocytes
2002;33:695–701. (PMNs) to capillary dropout in the human diabetic choroid. Am J Pathol
168. Costa VP, Kuzniec S, Molnar LJ, et al. Clinical findings and hemo­ 1997;151:707–14.
dynamic changes associated with severe occlusive carotid artery disease. 185. Lovasik J, Kergoat H. Electroretinographic results and ocular vascular perfu-
Ophthalmology 1997;104:1994–2002. sion in type I diabetes. Invest Ophthalmol Vis Sci 1993;34:1731–43.
169. Linsenmeier RA, Padnick-Silver L. Mtabolic dependence of photoreceptors on 186. Holopigian K, Greenstein VC, Seiple W, et al. Evidence for photoreceptor
the choroid in the normal and detached retina. Invest Ophthalmol Vis Sci changes in patients with diabetic retinopathy. Invest Ophthalmol Vis Sci
2000;41:3117–23. 1997;38:2355–65.
170. Wang S, Linsenmeier RA. Hyperoxia improves oxygen consumption in the 187. Grunwald J, Hariprasad S, DuPont J, et al. Foveolar choroidal blood flow in
detached feline retina. Invest Ophthalmol Vis Sci 2007;48:1335–41. age-related macular degeneration. Invest Ophthalmol Vis Sci 1998;39:
171. Lewis G, Mervin K, Valter K, et al. Limiting the proliferation and reactivity 385–90.
of retinal Muller cells during experimental retinal detachment: the value of 188. Grunwald JE, Metelitsina TI, Dupont JC, et al. Reduced foveolar choroidal
oxygen supplementation. Am J Ophthalmol 1999;128:165–72. blood flow in eyes with increasing AMD severity. Invest Ophthalmol Vis Sci
172. Folkman, J. Tumor angiogenesis: therapeutic implications. N. Engl J Med 285: 2005;46:1033–8.
1182–6, 1971. 189. Metelitsina TI, Grunwald JE, DuPont JC, et al. Foveolar choroidal circulation
173. Penn JS, Madan A, Caldwell RB, et al. Vascular endothelial growth factor in and choroidal neovascularization in age-related macular degeneration. Invest
eye disease. Prog Retin Eye Res 2008;27:331–71. Ophthalmol Vis Sci 2008;49:358–63.
174. Salam A, Mathew R, Sivaprasad S. Treatment of proliferative diabetic reti- 190. Pournaras CJ, Rungger-Brandle E, Riva CE, et al. Regulation of retinal blood
nopathy with anti-VEGF agents. Acta Ophthalmol 2011; 89:405–11. flow in health and disease. Prog Retin Eye Res 2008;27:284–330.
175. Antonetti DA, Barber AJ, Khin S, et al. Vascular permeability in experimental 191. McLeod DS, Taomoto M, Otsuji T, et al. Quantifying changes in RPE
diabetes is associated with reduced endothelial occludin content: vascular and choriocapillaris in eyes with age-related macular degeneration. Invest
endothelial growth factor decreases occludin in retinal endothelial cells. Penn Ophthalmol Vis Sci 2002;43:1986–93.
State Retina Research Group. Diabetes 1998;47:1953–9. 192. McLeod DS, Grebe R, Bhutto I, et al. Relationship between RPE and chorio-
176. Aiello L, Bursell S, Clermont A, et al. Vascular endothelial growth factor- capillaris in age-related macular degeneration. Invest Ophthalmol Vis Sci
induced retinal permeability is mediated by protein kinase C in vivo and 2009;50:4982–91.
suppressed by an orally effective beta-isoform-selective inhibitor. Diabetes 193. Feigl B. Age-related maculopathy-linking aetiology and pathophysiological
1997;46:1473–80. changes to the ischaemia hypothesis. Prog Retin Eye Res 2009;28:63–86.

You might also like