13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
Annu. Rev. Cell Dev. Biol. 2001. 17:255–96
c 2001 by Annual Reviews. All rights reserved
Copyright °
EARLY EYE DEVELOPMENT IN VERTEBRATES
Robert L. Chow
Program in Developmental Biology, The Research Institute, Hospital for Sick Children,
555 University Avenue, Toronto, Ontario, Canada, M5G 1X8;
e-mail: bobchow@sickkids.on.ca
Richard A. Lang
Division of Developmental Biology, Department of Ophthalmology, Children’s
Hospital Research Foundation, 3333 Burnet Avenue, Cincinnati, Ohio 45229;
e-mail: richard.lang@chmcc.org
Key Words induction, retina, lens, patterning, specification
■ Abstract This review provides a synthesis that combines data from classical
experimentation and recent advances in our understanding of early eye development.
Emphasis is placed on the events that underlie and direct neural retina formation
and lens induction. Understanding these events represents a longstanding problem in
developmental biology. Early interest can be attributed to the curiosity generated by
the relatively frequent occurrence of disorders such as cyclopia and anophthalmia, in
which dramatic changes in eye development are readily observed. However, it was the
advent of experimental embryology at the turn of the century that transformed curiosity
into active investigation. Pioneered by investigators such as Spemann and Adelmann,
these embryological manipulations have left a profound legacy. Questions about early
eye development first addressed using tissue manipulations remain topical as we try to
understand the molecular basis of this process.
CONTENTS
INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Overview of Eye Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Fate Mapping of the Eye Primordia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Defining Axes Within the Developing Eye . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
CONSERVATION IN DROSOPHILA EYE SPECIFICATION . . . . . . . . . . . . . . . . . .
EYE DEVELOPMENT IS CLOSELY ASSOCIATED
WITH NEURAL INDUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The First Events of Eye Development Occur
Within the Anterior Neural Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Anterior Neural Plate Transcription Factors
are Required for Normal Eye Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
GENERATING TWO EYES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A Single Eye Field is Bisected by Signals from
the Underlying Prechordal Mesoderm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1081-0706/01/1115-0255$14.00
256
256
258
258
259
262
262
263
267
267
255
24 Sep 2001
12:24
256
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
cyclops is Essential for Bisection of the Eye Field . . . . . . . . . . . . . . . . . . . . . . . . . .
hedgehog is Important for Ventral Neural Specification and
May Be Regulated by Cyc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
DEVELOPMENT OF THE OPTIC VESICLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Establishing Dorsal/Ventral Polarity in the Optic Vesicles . . . . . . . . . . . . . . . . . . . .
Interaction with Surface Ectoderm is Required for Neural Retina
Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lhx2 is Essential for the Transition of Optic Vesicle to Optic Cup . . . . . . . . . . . . .
FGF Signaling Can Specify Retinal Development . . . . . . . . . . . . . . . . . . . . . . . . . .
Early Patterning of the Neural Retina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
RPE Specification is Dependent on Signals from Extraocular
Mesenchyme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Dorsal/Ventral Polarity in the Neural Retina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Anterior-Posterior Polarity in the Developing Eye . . . . . . . . . . . . . . . . . . . . . . . . . .
LENS INDUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Lens Induction is a Multi-Step Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Pax6 is Essential for Lens Induction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Bmp4 and Bmp7 Play Important Roles in Lens Development . . . . . . . . . . . . . . . . .
FGF Receptor Signaling is Required for Lens Induction . . . . . . . . . . . . . . . . . . . . .
Sox1, Sox2, and Sox3 Participate in Lens Development . . . . . . . . . . . . . . . . . . . . . .
fork head Transcription Factors Foxe3 and Lens1 Lie Downstream
of Pax6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The sine oculis–Related Transcription Factor Six3 Can Transform Otic
Vesicle to Lens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Maf Family Members in Lens Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
CONCLUDING REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
268
269
270
270
271
272
272
273
273
274
275
275
276
277
280
281
282
283
284
284
285
INTRODUCTION
Overview of Eye Development
The eye is essentially a highly specialized extension of the brain. The first morphological sign of eye development in vertebrates is the bilateral evagination of
diencephalon in the early neurula (Figure 1A). In mammals, this is marked by
the appearance of the optic pit, whereas in fish and amphibians a bulging of
the optic primordia is observed. Continued evagination of the optic primordia
leads to the formation of the optic vesicles. These extend towards the overlying,
non-neural surface ectoderm that will ultimately give rise to the lens and cornea.
Mesenchyme between the optic vesicle and the surface ectoderm (apparent in
mammals and chick) is displaced as the two tissues come into close physical
contact (Figure 1B). This is a critical period in eye development during which
inductive signals between the optic vesicle and the surface ectoderm are thought
to exchange. Transmission electron microscopy studies in the rat have shown that
during this period of close contact, the optic vesicle and lens placode are tightly
associated through a network of collagenous fibrils and cytoplasmic processes
(McAvoy 1980). At this stage, the presumptive lens also shows the first morphological signs of development. This is characterized by formation of the lens
24 Sep 2001
12:24
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
257
placode, a thickening of the surface ectoderm that comes into contact with the optic
vesicle. Molecularly, lens placode formation coincides with the onset of crystallin
expression. These protein families are expressed at high levels in a lens preferential
manner and are required for generating and maintaining lens transparency (Cvekl
& Piatigorsky 1996, Graw 1996, Wistow & Piatigorsky 1988, Wride 1996).
Coordinated invagination of the lens placode and the optic vesicle results in the
formation of the lens vesicle and a double-layered optic cup and provides the first
indication of the final shape of the eye (Figure 1C). The inner layer of the optic
cup (facing the lens) forms the neural retina, while the outer layer of the optic
cup gives rise to the retinal pigment epithelium (RPE) (Figure 1D). At the ventral extremity of the optic vesicle, the process of invagination forms a groove
that runs continuously from the ventral-most region of the neural retina and
along the ventral aspect of the optic stalk to the junction with the neural tube
(Figure 2C). The point at which the laterally growing edges of the optic cup fuse
is known as the choroidal (or optic) fissure. This structure provides a channel for
blood vessels within the eye and an exit route for projecting axons (Bron et al.
1997).
Formation of a primary lens fiber cell mass from posterior cells of the lens
vesicle (i.e., those facing the optic cup) establishes polarity within the lens. This
is defined by the presence of an anterior monolayer of proliferating epithelial
cells that overlies a posteriorly positioned terminally differentiated fiber cell mass
(Figure 1D). Secondary lens fiber cells originate from the equatorial region or
transitional zone of the lens in a process that continues throughout adulthood. Fiber
cell differentiation is characterized by the onset of fiber cell-specific crystallin
expression (e.g., β and γ crystallins in the mouse), elongation, and the loss of
organelles including the nucleus, mitochondria and endoplasmic reticulum.
Development of the cornea from the surface ectoderm overlying the lens and
from migrating, neural crest–derived mesenchyme is a highly coordinated, multistep process (Hay 1986). Briefly, it first involves the secretion of a collagen-rich
extracellular matrix by the corneal epithelium (ectoderm overlying the lens). This
primary stroma attracts a wave of neural crest–derived mesenchymal cells from the
region surrounding the eye that coincides with its hydration and the migration of a
second wave of neural crest–derived mesenchymal cells. Eventually, an increased
level of thyroxine triggers the dehydration and compaction of the posterior stroma
that ultimately leads to the formation of the mature, transparent cornea.
The ciliary body and iris are derived from the distal tip of the optic cup at the
point where the inner and outer optic cup layers meet (Beebe 1986) (Figure 1C).
Although a detailed description of the development of these structures is beyond
the scope of this review, it is interesting to note that Otx1, an orthodenticlerelated transcription factor, appears to play an essential role during their development. These structures are missing or reduced in Otx1 loss-of-function
mutants, whereas the retina is otherwise normal (Acampora et al. 1996).
Within the limits of this review, we have chosen to focus on the better characterized elements of early eye development, and as a consequence, sizeable subject
areas have been excluded, including the regulation of lens polarization (Lang 1999,
24 Sep 2001
12:24
258
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
McAvoy et al. 1999), anterior segment development (Beebe & Coats 2000, Kidson
et al. 1999, Kume et al. 1998, Pressman et al. 2000), and the mechanism of retinal
differentiation (Cepko et al. 1996, Harris 1997, Kageyama et al. 1997).
Fate Mapping of the Eye Primordia
Fate mapping studies reveal that there is species-to-species variation in the location
of the retinal primordium. Studies in amphibians (Brun 1981, Eagleson & Harris
1990, Jacobson 1959, Manchot 1929, Woerdeman 1929) and chick (Couly & Le
Douarin 1985, Couly & Le Douarin 1987) have demonstrated that the retinal
primordia are positioned in two distinct regions on either side of the midline of
the anterior neural plate. Although some degree of lateral cell movement from the
anterior neural plate to the anterior neural folds has been observed in amphibians
(Brun 1981), studies in Xenopus have shown that cells positioned in the midline
of embryos at early neural plate stages do not migrate laterally (Li et al. 1997).
In contrast, fate mapping studies performed in zebrafish at the early neural shield
stage have suggested that retinal precursors are derived from a single domain in
the anteromedial forebrain region (Woo & Fraser 1995). This region is thought to
undergo lateral movements as a result of being physically displaced by invading
ventral diencephalic precursors (Woo & Fraser 1995). Zebrafish mutants such
as trilobite, silberblick and knypek that affect convergent extension movements
show defects in this process. Mutations in these genes result in varying degrees
of cyclopia, which is thought to result from the inability of ventral diencephalic
precursors to divide the presumptive retinal field (Hammerschmidt et al. 1996b,
Heisenberg et al. 1996, Heisenberg & Nusslein-Volhard 1997, Marlow et al. 1998,
Solnica-Krezel et al. 1996).
Defining Axes Within the Developing Eye
The developing eye undergoes many complex morphological changes that are
sometimes difficult to conceptualize spatially. For example, although the RPE
develops from the dorsal optic vesicle (Figures 1B, 2A–C) as morphogenesis proceeds, presumptive RPE eventually spreads ventrally to completely surround the
neural retina (Figure 2D). Because the derivatives of the eye primordium (i.e., optic
stalk, neural retina, and RPE) and the patterning events responsible for their formation are essentially determined along the dorsal/ventral axis of the eye, it is
convenient to think of the developing eye as a simple neuroepithelial sheet defined
by this axis (Figures 2A and 3). As the dorsal regions of the developing eye can
also be referred to as distal and the ventral regions as proximal, we herein refer
to this axis as the dorso-distal (DD) and the proximo-ventral (PV) axis. Although
poorly understood and omitted from our model, one should keep in mind that early
patterning events also act along the anterior (nasal) and posterior (temporal) axis
of the developing eye (see below).
In conceptualizing the developing eye field as a simple PV-DD sheet, the PVmost domain will give rise to the ventral floor of the optic stalk (Figures 2, 3). It
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
259
is arguable that the DD-most region would form the dorsal half of the optic stalk
(Figures 2, 3); however, we have chosen to define the DD-most region as that
which gives rise to RPE for two reasons. First, while the ventral half of the optic
stalk invaginates (Figure 2C), differentiates into glial cells, and forms the tissue
through which projecting retinal ganglion cell axons project (Bron et al. 1997),
the dorsal half of the optic stalk gives rise to the non-neural tissue that eventually
encases the optic nerve. Second, genes that pattern the optic stalk such as Pax2
(Nornes et al. 1990, Torres et al. 1996), Six3 (Oliver et al. 1995), Optx2 (Toy et al.
1998), and Vax1 (Hallonet et al. 1998) are restricted to the PV half of the optic
stalk, suggesting that it and not the DD-half of the optic stalk responds to signals
involved in patterning of the eye field (see below). While this conceptualization of
the PV-DD axis is highly simplified, it does provide a convenient framework for
discussing inductive processes within the developing eye (Figure 3).
CONSERVATION IN DROSOPHILA EYE SPECIFICATION
Much of our understanding of how early developmental genes direct eye formation has come from work on Drosophila eye development. Although the architecture of vertebrate and fly eyes are dramatically different, it is now evident
that similar regulatory networks and pathways direct eye development and function in both phyla. This has led to the debated hypothesis that rather than having
evolved independently, eyes from different phyla evolved monophyletically from
a common ancestral precursor (Callaerts et al. 1997, Fernald 2000, Pichaud et al.
2001). Although the mechanism behind this hypothesis is unclear, it has been proposed that the same regulatory networks and pathways directing eye formation in
flies and vertebrates may have been recruited independently throughout evolution,
thereby allowing eyes to evolve different strategies for perceiving light (Pichaud
et al. 2001).
Several transcription factors believed to play a critical role in the early stages
of Drosophila eye development have been identified in misexpression studies by
their ability to generate ectopic eyes in non-eye imaginal discs. These include eyeless (ey) (Halder et al. 1995) and twin of eyeless (toy) (Czerny et al. 1999), both
homologues of the vertebrate Pax6 (Czerny et al. 1999, Quiring et al. 1994), eyes
absent (eya), and sine oculis (so) (Bonini et al. 1997, Pignoni et al. 1997),
dachshund(dac) (Chen et al. 1997, Pignoni et al. 1997, Shen & Mardon 1997),
teashirt (ts) (Pan & Rubin 1998), and optix (opt) (Seimiya & Gehring 2000). A
pathway ordering these genes is beginning to emerge (Heberlein & Treisman 2000,
Wawersik & Maas 2000) (Figure 4). Genetic and biochemical evidence indicate
that Toy may lie at the beginning of this pathway because it can directly regulate
the eye-specific enhancer of ey (Czerny et al. 1999). ey has been proposed to be upstream of eya, so, and dac because ey expression precedes the expression of these
genes (Bonini et al. 1993, Cheyette et al. 1994, Halder et al. 1995, Mardon et al.
1994). Furthermore, eya and so are not required for ey expression, whereas ey is
13 Sep 2001
13:49
260
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Figure 4 A self-regulatory network of genes directs Drosophila eye specification.
Based on its expression pattern and its ability to generate ectopic eye formation in noneye imaginal discs, twin of eyeless (toy) is believed to act at the apex of the Drosophila
eye specification network (indicated by shaded box). Ectopic eye formation, timing
of expression and mutant expression studies place eyeless (ey) downstream of toy,
followed by eyes absent (eya), sine oculis (so), and then dachshund (dac). Additional
studies show that ey, eya, so, and dac can up-regulate one another (reverse arrows).
Teashirt and optix appear to function in this network based on their ability to generate
ectopic eyes. Drx function in the fly eye has not been established, although its vertebrate
homologue, Rx, is essential during early eye development. Notch, Egfr, Wingless,
and Hedgehog signaling pathways act upstream and are believed to contribute to the
coexpression of genes within Drosophila eye specification network.
required for so and eya expression in the eye disc during normal eye development
(Halder et al. 1998). dac has been proposed to lie downstream of eya and so based
on the observation that dac is not required for eya or so expression and that so and
eya are required for dac expression (Chen et al. 1997). While Opt requires so and
eya for ectopic eye formation, it does not require ey, suggesting that during ectopic eye formation it can function in a partially different pathway (Seimiya &
Gehring 2000).
The misexpression studies described above suggest that transcription factors
involved in Drosophila eye specification act in a hierarchical manner with toy and
ey at the beginning of the pathway. These studies, however, have also shown that
misexpression of eya, eya + so, dac, and dac + eya can upregulate ey expression as
well as each other’s expression (Chen et al. 1997, Pignoni et al. 1997, Shen & Mardon 1997), suggesting the presence of a positive feedback loop within this pathway.
This has led to the hypothesis that the transcription factors involved in Drosophila
eye specification may function within a self-regulating network that locks into an
eye development program (Heberlein & Treisman 2000). The ability of Eya to
physically interact with either Dac or So has raised the interesting possibility that
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
261
these factors function as part of a transcriptional complex in vivo (Chen et al. 1997,
Pignoni et al. 1997).
If a network of transcription factors does control eye specification in the fly,
one prediction would be that these genes show overlapping expression domains in
the developing fly eye. It has been shown that overlapping expression (although
incomplete) occurs between these transcription factors within the eye field between
larval stage 1 and 2 (Kumar & Moses 2001). This stage of fly eye development has
been shown to coincide with a critical period during which up-regulation of Notch
signaling and down-regulation of Egfr signaling in the eye half of the eye-antennal
disc is believed to specify eye fates (Kumar & Moses 2001). The obsevation
that Hedgehog and Wingless signaling can also participate in eye-antennal fate
decisions (Royet & Finkelstein 1996, 1997) has led to the hypothesis that Notch,
Egfr, Wingless, and Hedgehog signals function upstream of at least some components of the network of transcription factors controlling eye specification. Signals
from these pathways would contribute to the overlapping expression of these transcription factors, which in turn would be predicted to lock into an eye development
program (Kumar & Moses 2001) (Figure 4).
It will be interesting to see the extent to which regulatory pathways have been
shared in eye development throughout evolution. Although beyond the scope of
this review, one remarkable finding is the observation that retinal differentiation
in vertebrates and ommatidial differentiation in flies both appear to follow a wave
of Hedgehog signaling (Neumann & Nuesslein-Volhard 2000). During early eye
development, experimental evidence in vertebrates supports the existence of a similar, self-regulatory network of transcription factors controlling eye specification.
This has come partly from misexpression studies in which transcription factors
believed to play an important early role in vertebrate eye formation, such as Pax6,
Six3, Optx2, and Rx, can up-regulate one another’s expression and induce ectopic
retinal tissue (Andreazzoli et al. 1999, Bernier et al. 2000, Chow et al. 1999, Loosli
et al. 1999, Zuber et al. 1999).
Several differences exist, however, between the putative regulatory networks
directing fly and vertebrate eye specification. For example, Drosophila has undergone a duplication of the ancestral Pax6 gene (Czerny et al. 1999), giving rise to
toy and ey, whereas in chick, mice, or humans there appears to be only one Pax6
orthologue. In addition, the role of vertebrate eya homologues, EYA1-4, is unclear.
Although EYA1-3 are expressed in the developing eye, mouse Eya1 (Johnson et al.
1999, Xu et al. 1999a), human EYA1 (Abdelhak et al. 1997), and EYA4 (Wayne
et al. 2001) are not essential for eye development and instead play essential roles in
ear and kidney development (it is interesting to note, however, that missense mutations in EYA1 have been associated with cataractogenesis and anterior segment
defects; Azuma et al. 2000). Finally, although the Rx gene plays an essential role
early in vertebrate eye formation, its Drosophila homologue, Drx, is not expressed
in the eye-antennal disc (Mathers et al. 1997). These observations indicate that not
all the components that make up the regulatory network directing eye specification
are necessarily conserved in flies and vertebrates.
13 Sep 2001
13:49
262
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
EYE DEVELOPMENT IS CLOSELY ASSOCIATED
WITH NEURAL INDUCTION
The First Events of Eye Development Occur
Within the Anterior Neural Plate
While the first morphological indication of eye development occurs with the bilateral evagination of the developing forebrain, the events involved in specifying eye formation begin earlier. Although some early support for this idea came
from explant studies in amphibians by Adelmann (Adelmann 1936a,b), the elegant studies of Nieuwkoop on explanted presumptive neuroectoderm from frog
late blastulae were especially informative. Normally this type of ectodermal explant (also known as an animal cap) gives rise to non-neural epidermis if cultured
alone, but Nieuwkoop discovered that if the cells of these explants were dissociated
and then re-aggregated, they differentiated into anterior neural tissue (Nieuwkoop
1963). The important point in these experiments was that these so-called activated
explants underwent neural induction in the absence of mesoderm and endoderm
(possible inductive sources) and gave rise to anterior brain, olfactory placodes, and
eye structures (Nieuwkoop 1963). Recent work probing the molecular basis for
neural induction (and revisiting Nieuwkoop’s observation) have shown that this
process is mediated by molecules such as Chordin, Noggin, Follistatin, Cerberus,
and xnr3 (reviewed in Harland 2000, Weinstein & Hemmati-Brivanlou 1999) that
antagonize the bone morphogenetic protein (BMP) signaling pathways. When ectodermal explants from Xenopus are microinjected with RNA encoding any of
these BMP signaling pathway inhibitors, they behave exactly as Nieuwkoop’s explants did and give rise to anterior neural structures including eyes. Because the
potential inductive influence of mesoderm or endoderm is absent in these neuralized ectodermal explants, these results demonstrate that the molecular mechanisms
directing eye specification lie downstream of neural induction and are inherent to
developing anterior neural tissue (symbolized in Figure 3). It should be pointed
out, however, that the anterior neural plate is patterned by complex interactions
(Papalopulu & Kintner 1996) and, therefore, the derivation of eyes from the anterior neural plate will likely involve more complex mechanisms and interactions.
Several of the transcription factors that pattern the anterior neural plate of the
early vertebrate embryo such as Pax6, Rx, Six3, and Hesx1/Rpx are essential for
normal eye development (see below). Although the expression of these factors lies
downstream of neural induction, the precise molecular events that initiate their
expression are not well understood. Initiation of Pax6, Six3, and Otx2 (Otx2 expression precedes neural induction) expression occurs in the absence of Rx (Zhang
et al. 2000); initiation of Rx, Six3, and Otx2 expression occurs in the absence of
Pax6 (Oliver et al. 1995, Stoykova et al. 1996); cell-autonomous initiation of Six3
and Hesx1/Rpx expression occurs in the absence of Otx2 (Rhinn et al. 1998); and
initiation of Pax6 and Six3 expression occurs in the absence Hesx1/Rpx (Dattani
et al. 1998). Although these observations seemingly contradict others indicating
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
263
that this set of genes can co-activate (see above), it is possible that their transcriptional initiation in the anterior neural plate is regulated independently by currently
unidentified factors downstream of neural induction. Although factors that appear upstream of Drosophila eye specification, such as Notch, Egfr, and Hh, have
not been implicated in vertebrate eye specification, there is evidence that signaling through Wnts (vertebrate Wingless homologues) may be involved. Studies in
Xenopus have shown that misexpression of the Wnt receptor, Frizzled 3 (Xfz3),
results in ectopic expression of Pax6, Rx, and Otx2 and leads to ectopic eye formation. Furthermore, inhibition of Wnt signaling by overexpression of a truncated
Xfrz3 receptor or intracellular Xfrz3-interacting protein inhibited endogenous eye
formation (Rasmussen et al. 2001). These results suggest an early role for Wnt
signaling during vertebrate eye development, possibly in the regulation of transcription factors essential for eye formation.
Anterior Neural Plate Transcription Factors
are Required for Normal Eye Development
Otx2 is a member of the orthodenticle-related family of transcription factors, which contain a homeodomain of the bicoid class (Simeone et al. 1993). Otx2
expression precedes that of other early eye transcription factors (i.e., Pax6, Six3,
Rx) and is expressed in the entire embryonic ectoderm of the mouse prior to gastrulation (Simeone et al. 1993). Although Otx2 is eventually restricted to the anterior
end as gastrulation proceeds, double in situ hybridization analysis in Xenopus has
shown that in the anterior neural plate, Otx2 is not present in the eye field defined
by Rx expression (Andreazzoli et al. 1999). This observation is interesting in light
of the finding that Rx misexpression at early neurula stages can repress endogenous Otx2 (Andreazzoli et al. 1999). The down-regulation of Otx2 in the early eye
field suggests that it does not participate in early specification. At later stages of
eye development, however, Otx2 is expressed in the optic vesicles and eventually
becomes restricted to the RPE (Bovolenta et al. 1997, Simeone et al. 1993). In
Otx2−/− mutants, forebrain and midbrain structures are completely missing, possibly due to a defect in neural induction (Matsuo et al. 1995, Pannese et al. 1995).
Therefore, it is difficult to determine the precise role of Otx2 in the developing eye
because the defect in these mutants manifests itself well before any morphological
signs of eye formation would have occurred. Interestingly, in chimeric embryos in
which the neuroectoderm is made up of Otx2−/− cells, the expression of the forebrain/eye marker, Six3 and Hesx1/Rpx, is initiated but not maintained (Rhinn et al.
1998), suggesting that Otx2 is not required cell-autonomously for the induction of
anterior neural tissue but is required for its maintenance and regional specification
(Rhinn et al. 1998).
Otx2
Hesx1/Rpx encodes a member of the paired-like domain family of transcription factors and is expressed early in the mouse gastrula and anterior neural
plate. Hesx1 expression becomes restricted to the developing anterior pituitary
Hesx1/Rpx
13 Sep 2001
13:49
264
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
(Rathke’s pouch) by embryonic day 9.5 (Hermesz et al. 1996). In humans, mutations in HESX1 lead to septo-optic dysplasia, a condition that is characterized by
variable phenotypes including optic nerve hypoplasia, pituitary gland hypoplasia,
and midline abnormalities of the brain (Dattani et al. 1998). In mice, null mutations
of Hesx1 lead to similarly variable anterior CNS defects and pituitary dysplasia,
as well as anophthalmia or microphthalmia (Dattani et al. 1998). It has been suggested that Hesx1 is required for establishing the size of the forebrain, perhaps
by regulating proliferation in the forebrain primordium. Since Hesx1 expression
is later restricted to the pituitary, it has also been suggested that the ocular defects
observed in Hesx1 mutants arise as a consequence of the reduction in forebrain
size rather than a direct regulatory role by Hesx1 (Dattani et al. 1998).
Pax6, is a paired box and a paired-like homeobox gene that has maintained
an extremely high level of conservation throughout evolution (reviewed in Callaerts
et al. 1997, Cvekl & Piatigorsky 1996). Pax6 orthologues have been identified in
vertebrates (Hirsch & Harris 1997, Krauss et al. 1991, Li et al. 1997, Martin et al.
1992, Walther & Gruss 1991), flies (Quiring et al. 1994), and squid (Tomarev
et al. 1997). Interestingly, Pax6 homologues are also expressed in the sense organ
of nematodes (Zhang & Emmons 1995) and in the simple photosensitive ocellus
of ascidians (Glardon et al. 1997), which suggests a conserved role for Pax6 in
sensory detection.
In vertebrates, Pax6 is first expressed toward the end of gastrulation in the
anterior neural plate (Del Rio-Tsonis et al. 1995; Grindley et al. 1995; Hirsch &
Harris 1997; Li et al. 1994, 1997; Puschel et al. 1992; Walther & Gruss 1991). As
neurulation proceeds, Pax6 expression resolves to the DD optic vesicles (Figure 3),
as well as to the presumptive lens ectoderm. In the developing optic cup, Pax6 is
initially expressed throughout (Grindley et al. 1995), but in the differentiated retina
it is expressed strongly only in ganglion and amacrine (Belecky-Adams et al. 1997).
In addition to a role in eye formation, Pax6 is essential for normal development in
many regions outside the eye, including the nasal epithelium (Hogan et al. 1986),
brain (Grindley et al. 1997, Schmahl et al. 1993), pancreas (St-Onge et al. 1997),
and lacrimal gland (Makarenkova et al. 2000).
Loss-of-function mutations in Pax6 are semi-dominant and in the heterozygous
state lead to the Small eye (Sey) phenotype in mice (Hill et al. 1991, Hogan et al.
1986) and rats (Matsuo et al. 1993), which is characterized by cataractogenesis,
iris hypoplasia, and microphthalmia. Similar mutations in humans cause the ocular
syndrome aniridia (Glaser et al. 1992, Hanson et al. 1993, Jordan et al. 1992, Ton
et al. 1991). The idea that Pax6 gene dosage is important for normal eye development is illustrated by studies in which additional copies of the Pax6 locus were introduced in transgenic mice (Schedl et al. 1996). Although a single copy of the Pax6
locus could rescue the Sey (heterozygous) phenotype, mice that had more than two
normal copies of the Pax6 locus also displayed defects in eye development (Schedl
et al. 1996). This revealed that an excess of Pax6 is detrimental to normal eye development (Schedl et al. 1996) and highlights the importance of Pax6 gene dosage.
Pax6
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
265
Homozygous mouse or rat Sey mutants are anophthalmic and die at birth
(Grindley et al. 1995, Hogan et al. 1986, Matsuo et al. 1993). Optic vesicles
are present in Sey/Sey mutants but fail to constrict proximally, resulting in the
persistence of a luminal optic stalk. In addition, while the optic vesicles do come
into contact with the surface ectoderm in Sey/Sey mutants (Grindley et al. 1995),
lens placode thickening does not occur (Grindley et al. 1995, Hogan et al. 1986,
Matsuo et al. 1993). Eventually, a structure that resembles the optic cup does form,
although there is no indication of differentiation in presumptive neural retina or
RPE, and both layers appear identical (Grindley et al. 1995).
The ability of Pax6 to induce ectopic eyes supports the idea that it plays a critical
role in the early stages of eye development. Misexpression in Drosophila non-eye
imaginal discs of eyeless or mouse (Halder et al. 1995), squid (Tomarev et al.
1997), or ascidian (Glardon et al. 1997) Pax6 all lead to ectopic eye formation and
reveal the high degree of functional conservation of Pax6 throughout evolution.
Pax6 misexpression can also induce fully differentiated ectopic eyes in Xenopus
and the ectopic expression of early eye development genes such as Otx2, Rx, and
Six3 as well as endogenous Pax6 (Chow et al. 1999). The observation that ectopic
eyes form only when Pax6 mRNA is microinjected into the dorsal animal pole
blastomeres (that give rise to neural and ectodermal tissues) indicates that factors
in addition to Pax6 are required for the induction of ectopic eyes.
Since Pax6 loss-of-function mutants fail to develop eyes, the use of chimeric
studies and conditional gene targeting has been important in determining its role at
later stages of eye development (see below). Studies in chimeric mice derived from
Sey/Sey and wild-type cells have revealed a cell-autonomous requirement for Pax6
in the lens (as early as embryonic day 9.5) and DD optic vesicle (Collinson et al.
2000, Quinn et al. 1996). In the DD region of the optic vesicle and in the neural
retina, wild-type and mutant cells segregate, and there is a failure of the mutant
cell patches to differentiate. The use of a conditional Pax6 null allele activated by
a Pax6 neural retina enhancer-driven Cre expression has shown that Pax6 plays
an essential role later in eye development in maintaining the multipotent state of
retinal progenitor cells (Marquardt et al. 2001). In the absence of Pax6 neural retina
expression, the only retinal cell types to differentiate are a subset of non-glycinergic
amacrine cells. Pax6, however, does not appear to be required for the maintenance
of retinal identity in these experiments, as expression of Rx1, Six3, Six6, and Lhx2
was not lost in Pax6-deficient neural retina tissue (Marquardt et al. 2001).
Six3/Optx2 Six3 and Optx2 (alternatively named Six6 or Six9) are closely related
members of the SIX-homeodomain family (of which sine oculis is the founding member) and share over 90% homeodomain amino acid identity with Optix,
another Drosophila SIX-homeodomain gene (Toy et al. 1998). Heterozygous
mutations in humans for SIX3 have recently been identified in six cases of holoprosencephaly (Wallis et al. 1999). The presence of severe brain defects and microphthalmia in these cases support the role for SIX3 in the developing anterior
neural plate and eye.
13 Sep 2001
13:49
266
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Six3 and Optx2 have expression patterns suggestive of a role in early eye development. Although both genes are expressed in the gastrula, only Six3 is expressed
in the eye field (Bovolenta et al. 1998, Granadino et al. 1999, Jean et al. 1999,
Loosli et al. 1998, Oliver et al. 1995, Zuber et al. 1999). At later stages, Optx2 and
Six3 are expressed throughout the optic vesicle. Expression of both genes continues in the optic stalk and neural retina (Bovolenta et al. 1998, Granadino et al.
1999, Jean et al. 1999, Loosli et al. 1998, Oliver et al. 1995, Zuber et al. 1999).
Whereas Six3 is up-regulated in the lens placode, Optx2 expression appears to be
either absent (Jean et al. 1999, Zuber et al. 1999) or down-regulated (Toy et al.
1998) in this region.
Evidence from misexpression studies suggests that both Optx2 and Six3 play
important roles during retinal determination. In Medaka, misexpression of Six3
results in the conversion of regions of the midbrain and prospective cerebellum
to RPE-containing, optic cup–like structures, and in the appearance of ectopic
Rx, Pax6, and endogenous Six3 expression (Loosli et al. 1999). Furthermore, Six3
misexpression in zebrafish results in an enlargement of the rostral forebrain and
an enhanced expression of Pax2 in the optic stalks, consistent with the normal
expression of Six3 in this region, and suggesting that it may also play an important
role in determining cell fate in the optic stalk (Kobayashi et al. 1998). In chick
RPE cells, Optx2 misexpression leads to the expression of Chx10 (see below) and
visinin, two neural retinal markers normally excluded from the RPE (Toy et al.
1998). Optx2 misexpression in the eye field of Xenopus embryos results in an
expansion of the retinal territory or a dramatic increase in eye size, suggesting
that Optx2 may also regulate the proliferation of retinoblasts (Bernier et al. 2000,
Zuber et al. 1999). Together, these results suggest important roles for both Six3
and Optx2 in the determination of retinal fates.
Rx A essential role in the early events of eye development has been established
for the paired-like homeobox gene Rx based on its expression pattern, and gainand loss-of-function phenotypes (Furukawa et al. 1997, Mathers et al. 1997). Rx
is first expressed throughout the anterior neural plate, and following neurulation it
is expressed most abundantly in the optic vesicles and later throughout the neural
retina. By post-natal day 6.5 Rx is restricted to photoreceptor cells and the inner
nuclear layer, and by post-natal day 13.5, is no longer detectable.
Misexpression of Rx in Xenopus embryos results in the extension of ectopic
RPE along the optic nerve region and in the hyperproliferation of the neural retina
(Mathers et al. 1997), and in zebrafish results in the expansion of retinal tissue
into the forebrain (Chuang & Raymond 2001). These results are consistent with
observations that Rx misexpression in Xenopus can expand endogenous Pax6, Six3,
and Otx2 expression in the optic region at late neurula stages (Andreazzoli et al.
1999) and suggest that Rx may play an important role in regulating the initial
specification of retinal cells and perhaps their subsequent proliferation.
Mice homozygous for a null mutation of the Rx gene are anophthalmic (Mathers
et al. 1997). In contrast to the phenotype of Sey/Sey mice, in which optic vesicle
24 Sep 2001
12:27
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
267
formation does occur (albeit abnormally), optic vesicles fail to develop in Rx−/−
mice. Rx−/− mutants display a variable forebrain phenotype, which, in extreme
cases, leads to the complete loss of forebrain. In less extreme cases, the forebrain
is still present, but optic vesicle formation is absent and the region where the optic
vesicles would have formed is marked by the absence of Pax6, Six3, and Otx2
expression (Zhang et al. 2000). This observation is particularly interesting with
respect to the conditional medaka fish eye mutant eyeless (Winkler et al. 2000) in
which optic vesicle evagination also does not occur. In contrast to Rx−/− mutants,
however, the ventral region of the brain (where evagination would have occurred)
is patterned by distal eye markers Pax6, Rx1, and Rx2, and the proximal eye
marker Vax1 (Winkler et al. 2000). The medaka eyeless mutant demonstrates that
the morphological events behind optic vesicle formation can be uncoupled from
the events that pattern it. Rx might therefore be predicted to function upstream of
eyeless because both patterning and morphogenesis of the optic vesicles are lost
in Rx mutants.
GENERATING TWO EYES
A Single Eye Field is Bisected by Signals from
the Underlying Prechordal Mesoderm
In the early neural plate, transcription factors critical for eye development (such as
Pax6 and Rx) are generally expressed across the entire width of the anterior region.
Interestingly, these patterns of expression do not correlate with fate-mapping studies that suggest DD eye fates originate only from the lateral regions of the anterior
neural plate. As neurulation continues, however, the expression of some of these
transcription factors resolves into two lateral domains. A simple, retrospective interpretation of this observation is that at early neural plate stages, the entire anterior
neural plate has the potential to give rise to eye structures. Remarkably, in the late
1800s, sparked by interest in the frequent occurrence of cyclopia in newborns, it
had already been hypothesized that a single eye field exists in the early embryo
that resolves into two fields as development progresses (Adelmann 1936a).
Embryological manipulations provided evidence in support of the single eye
field hypothesis. Much of this came from the pioneering work of Adelmann, who,
in the late 1920s, demonstrated that transplanted regions of the anterior neural
plate midline possess an eye-forming potential (Adelmann 1936a). In addition,
Mangold made the striking observation that if the prechordal mesoderm underlying the anterior neural plate of Triton embryos was removed, cyclopic tadpoles
would develop (Mangold 1931) (Figure 5A–B). This work supported the single
eye field hypothesis and also provided some of the first evidence suggesting that
the prechordal mesoderm functions to prevent eye formation in the midline.
Several recent observations also support the idea that a single eye field exists in
the early embryo. Experiments similar to those of Mangold’s have recently been
repeated in Xenopus (Li et al. 1997) and chick (Pera & Kessel 1997), confirming
13 Sep 2001
13:49
268
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
the proposed inhibitory role of the prechordal mesoderm. Additionally, expression
of the transcription factors Pax6 and ET (a member of the T-box family) was
examined in these experiments. Pax6 and ET are both normally expressed in a
broad domain spanning the dorsal/anterior midline following the completion of
gastrulation. During neurulation, expression of these genes is down-regulated in
the midline (the future ventral forebrain), which results in the formation of two
distinct domains that demarcate the presumptive eye. Removal of the prechordal
mesoderm in embryos or in explants of the anterior neural plate results in the
persistence of midline expression of Pax6 and ET (Li et al. 1997, Pera & Kessel
1997) (Figure 5C, D). This demonstrated, at the molecular level, the importance of
prechordal mesoderm in patterning the anterior neural plate. Furthermore, when
donor prechordal plate tissue was grafted beneath the chick presumptive retinal
region, Pax6 expression was suppressed (Li et al. 1997, Pera & Kessel 1997).
Together, these experiments provide strong evidence supporting the existence of
a single eye field early in development that is bisected under the influence of
underlying prechordal mesoderm.
cyclops is Essential for Bisection of the Eye Field
Recent studies in zebrafish have shown that the cyclops (cyc) gene is expressed in
prechordal mesoderm and is necessary for the generation of two eyes. Zebrafish
with a loss-of-function mutation in cyc are characterized by a deficiency in midline
signaling that leads to cyclopia and failure of ventral forebrain formation (Ekker
et al. 1995, Hatta et al. 1991, Macdonald et al. 1995). cyc encodes a secreted,
Nodal-related member of the transforming growth factor-β (TGF-β) superfamily
(Feldman et al. 1998, Sampath et al. 1998). The observation that cyc expression
is restricted to the dorsal mesendoderm during gastrulation and to prechordal
mesoderm and notochord until early somitogenesis suggests that during normal
development, secreted Cyc may signal overlying neural tissue to adopt ventral
fates and consequently bisect the eye field (Feldman et al. 1998, Sampath et al.
1998).
Although the Cyc receptor has not yet been identified, there is strong evidence
that Cyc activity may be regulated by one-eyed pinhead (oep), a novel membraneassociated member of the EGF-CFC family of extracellular factors that includes
criptic, cripto, and FRL-1 (Zhang et al. 1998). Mutations in oep result in a failure of
prechordal mesoderm to develop and also lead to severe cyclopia (Hammerschmidt
et al. 1996a). Although misexpression of cyc mRNA cannot rescue oep mutants,
misexpression of putative downstream components of the Cyc/Nodal pathway
(such as an activated form of the putative Cyc receptor ActRIB or the transcription
factor Smad2) can rescue them (Gritsman et al. 1999) (Figure 5E, F). As Oep is
membrane associated and expressed in both ventral neuroectoderm and prechordal
mesoderm, it has been proposed to function cell-autonomously as a co-factor with
prechordal mesoderm-derived Cyc to activate an activin-like signaling cascade.
This would in turn initiate ventral forebrain specification and bisection of the eye
field (Gritsman et al. 1999) (Figure 3B).
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
269
hedgehog is Important for Ventral Neural Specification
and May Be Regulated by Cyc
It has been of interest to identify factors within the neuroectoderm that act downstream of Cyc. Members of the Hedgehog (Hh) family of signaling molecules,
which have been implicated in the specification of ventral neural tissue are excellent
candidates for such factors. hh was originally identified in Drosophila as a segment
patterning molecule and has since been implicated in many aspects of Drosophila
development, including eye morphogenetic furrow progression (Heberlein et al.
1995, 1993; Ma et al. 1993). In vertebrates, a large body of work has implicated
Hedgehog family members in specification of the ventral neural tube (Ericson et al.
1997, Patten & Placzek 2000) and in limb development (Capdevila & Johnson
2000). Like oep, sonic hedgehog (shh), a mammalian hedgehog homologue, is
expressed in both the ventral forebrain and underlying prechordal mesoderm
(Echelard et al. 1993, Ekker et al. 1995, Krauss et al. 1993, Macdonald et al.
1995, Riddle et al. 1993).
In humans, loss-of-function mutations in SHH are associated with holoprosencephaly, a condition that results in abnormal development of midline structures
including cyclopia in extreme cases (Belloni et al. 1996, Roessler et al. 1996).
Similar phenotypes are also observed in mice lacking Shh; cyclopia presumably
occurs as a result of ventral neural tube specification failure (Chiang et al. 1996).
Interestingly, shh loss-of-function in the zebrafish mutant, sonic-you, does not display cyclopia, although other defects such as abnormal retinal-tectal projections
are observed (Schauerte et al. 1998). One possible explanation for this mild phenotype is that another zebrafish hedgehog family member, tiggy-winkle hedgehog
(twhh), may compensate for the loss of shh because the two genes share overlapping expression patterns (Ekker et al. 1995, Macdonald et al. 1995).
Gain-of-function experiments demonstrate that Hh activity is likely to play a
role in ventralization of the eye fields. For example, misexpression of shh or twhh
in wild-type zebrafish embryos results in a reduction of eye pigment, as well as
the complete absence of lenses in some cases (Ekker et al. 1995, Macdonald et al.
1995). This coincides with a loss of Pax6 expression and ectopic expression of the
PV marker, Pax2, in the DD region of the optic vesicles (Figure 5G, H) (Ekker
et al. 1995). This result is consistent with the graded repression of Pax6 by Shh in
the developing neural tube (Ericson et al. 1997). A similar phenotype occurs when
a dominant-negative subunit of protein kinase A, which has been shown to mimic
Hh signaling, is misexpressed in the fish (Hammerschmidt et al. 1996a, Ungar &
Moon 1996).
Experiments performed in zebrafish support the idea that shh expression in
the ventral midline of the neural tube and brain lies downstream of cyc and that
Cyc may directly regulate shh. In cyc mutants, shh expression is lost in the ventral diencephalon but not in the underlying prechordal mesenchyme, suggesting
that Shh does not regulate its own expression in overlying neural tissue (Barth &
Wilson 1995, Krauss et al. 1993). This is further supported by misexpression studies showing that Shh is not sufficient to induce medial floor plate cells (i.e., ventral
13 Sep 2001
13:49
270
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
neural tube) (Schauerte et al. 1998) and is a poor inducer of shh expression in
neuroectoderm (Muller et al. 2000). cyc misexpression, however, can act in a cell
non-autonomous manner to rescue shh expression in cells of the ventral brain
and floorplate of cyc mutants (Sampath et al. 1998). In addition, misexpression
of constitutively active or dominant-negative downstream components of the Cyc
signaling pathway result in either ectopic up-regulation or down-regualtion, respectively, of shh in the neural ectoderm in a cell-autonomous manner (Muller
et al. 2000). The possibility that Cyc directly up-regulates shh expression has been
suggested from electroporation studies in chick neural tubes. In these studies, it
was shown that there is at least one enhancer present in an upstream region of the
shh gene that is able to direct expression to ventral neuroectoderm and is responsive to either Cyc or the downstream Cyc signaling molecules Smad2 and FAST-1
(Muller et al. 2000).
Although Shh signaling from the ventral midline may play an important role in
ventralizing the eye field, other sources of Hh may also be involved. Evidence for
this comes from an analysis of loss-of-function mutants of the fork head domain
transcription factor encoded by BF-1/Foxg1, which display an extreme dorsalization of the eyes such that the optic stalk is completely missing and is replaced by
an expanded retina (Huh et al. 1999). Although Shh expression was normal in the
ventral midline of the diencephalon, Shh expression in the ventral telencephalic
neuroepithelium, which first appears between E9.5 and E10, was absent in BF-1
mutants. This relatively late domain of Shh expression may therefore play a critical
role in the ventral patterning of the optic vesicle (Huh et al. 1999) in addition to
the earlier role of Shh derived from the ventral midline.
In summary, the anterior neural plate of the early embryo appears to possess
a single eye field that, if not resolved into two, will result in cyclopia. Important
ventralizing signals coming from prechordal mesoderm, such as Cyc, may induce
hh (sonic, tiggy-winkle) expression in overlying neural midline tissue. Hedgehog,
in turn, specifies proximo-ventral eye fates and leads to the repression of genes such
as Pax6 that direct DD eye development. Additional elements of these pathways
(Figure 3) presumably remain to be identified.
DEVELOPMENT OF THE OPTIC VESICLES
Establishing Dorsal/Ventral Polarity in the Optic Vesicles
Regardless of the source, Hh signaling as described above plays a critical role in
specifying the ventral region of the developing eye. Patterning of the ventral optic
vesicles is highlighted by the genes such as Pax2 (a paired homeobox gene)
(Dressler et al. 1990, Nornes et al. 1990), Vax1 (Hallonet et al. 1998), and tailless (an orphan nuclear receptor) (Hollemann et al. 1998). While several studies
have used Pax2 as a downstream marker of Hh signaling (Ekker et al. 1995,
Hammerschmidt et al. 1996a, Macdonald et al. 1995), the exact molecular mechanisms responsible for the initiation of Pax2 expression are not known.
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
271
Pax2 plays an important role in establishing proximo-ventral character in the
developing eye. This is supported by the observation that a frameshift mutation in
Pax2 is present in families with heritable optic nerve colobomas (Narahara et al.
1997; Sanyanusin et al. 1995a,b; Schimmenti et al. 1995, 1997). Similarly, genetargeted mice lacking Pax2 (Torres et al. 1996) and KRD (kidney renal defects)
mice harboring a naturally occurring mutation of Pax2 (Favor et al. 1996, Keller
et al. 1994, Torres et al. 1996) have similar optic nerve abnormalities, consistent
with a role for Pax2 in proximo-ventral eye development. These include coloboma
(failure to close) of the ventral fissure of the eyes, an extension of the RPE into the
optic nerve region, and a failure of optic chiasma to form.
The boundary between the presumptive optic stalk and neural retina in the
optic vesicles may form as a result of reciprocal transcriptional repression between Pax6 and Pax2 (Schwarz et al. 2000). Indirect evidence for this comes from
the observation that there is a dorsal expansion of Pax2 expression in Pax6 mutants and a ventral expansion of Pax6 expression in Pax2 mutants (Schwarz et al.
2000). Furthermore, gel shift and β-galactosidase-reporter assays have shown that
Pax2 and Pax6 can reciprocally bind and inhibit the activity of upstream regulatory regions in the Pax6 and Pax2 genes, respectively (Schwarz et al. 2000).
These results have led to the prediction that following the activation of Pax2
expression in the ventral optic region (presumably somewhere downstream of
hh), Pax2 and Pax6 expression would initially overlap (Figure 3B, C). Eventually, owing to the reciprocal transcriptional repression of these genes, they would
resolve into two distinct expression domains thereby establishing a distinct
boundary at the optic stalk–neural retina junction (Figure 3D) (Schwarz et al.
2000).
Interaction with Surface Ectoderm is Required
for Neural Retina Development
Formation of the neural retina follows a transient period of close contact between the DD region of the optic vesicle and the surface ectoderm/presumptive
lens. Early evidence that the surface ectoderm might play a role in neural retina
specification was provided by the observation that in tissue explantation experiments, neural retina would not develop if surface ectoderm was removed (Holtfreter
1939). More recent experiments have confirmed this observation (Hyer et al. 1998,
Nguyen & Arnheiter 2000) In addition, it was shown that if the amphibian optic
cup was rotated by 180◦ such that the presumptive RPE was now facing the surface ectoderm, it would transform into a secondary neural retina (Detwiler 1953,
Dragomirov 1937, Lopashov & Stroeva 1964, Mikami 1939). These observations
indicate a critical role for the surface ectoderm in directing neural retina formation. Interestingly, transplantation experiments have indicated that lens-to-retina
signaling is important for retinal maintenance in the mature fish eye, and its absence is a step in the evolution of a species of blind cave fish (Yamamoto & Jeffery
2000).
13 Sep 2001
13:49
272
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Lhx2 is Essential for the Transition of Optic
Vesicle to Optic Cup
Lhx2 encodes a LIM homeodomain-containing transcription factor essential for
eye development. Within the eyes, Lhx2 is expressed in the optic vesicles and is
subsequently localized to the neural retina where it becomes restricted to the inner
nuclear layer of the mature retina (Porter et al. 1997, Xu et al. 1993). Loss of
Lhx2 function in knock-out mice results in anophthamlia. Although optic vesicle formation occurs in these mice, optic cup and lens placode formation fails
to occur (Porter et al. 1997). The timing of the Lhx2 mutant phenotype suggests
that Lhx2 plays a critical role during the period of close contact between the
surface ectoderm and optic vesicle. Little is known, however, about the downstream and upstream targets of Lhx2. The observation that Pax6 expression is not
affected in the arrested optic vesicles of Lhx2−/− mutants and that Lhx2 expression is not affected in Sey/Sey mutants has led to the suggestion that Lhx2 and
Pax6 are independently essential for normal development of the eye (Porter et al.
1997).
FGF Signaling Can Specify Retinal Development
Considerable evidence suggests that members of the fibroblast growth factor (FGF)
family play a role in signaling neural retina formation during the period of close
contact. Much of this evidence has come from studies demonstrating that presumptive chick or frog RPE can transdifferentiate into neural retina when cultured
in the presence of FGF (Guillemot & Cepko 1992, Opas & Dziak 1994, Park &
Hollenberg 1989, Pittack et al. 1991, Reh et al. 1991, Vogel-Hopker et al. 2000).
Furthermore, addition of FGF2 neutralizing antibodies to cultured chick optic
vesicles can inhibit neural retina development (Pittack et al. 1997). In contrast,
neural retina development can be rescued in explants of the optic vesicle in which
the overlying surface ectoderm is removed by the addition of FGF-coated beads,
FGF-secreting fibroblasts, or injection of FGF retroviral expression vectors (Hyer
et al. 1998, Nguyen & Arnheiter 2000).
Although activity of exogenously applied FGF suggests a role in neural retina
induction, compound null mutants of two Fgfs (Fgf1 and Fgf2) known to be present
in the surface ectoderm (de Iongh & McAvoy 1993, Nguyen & Arnheiter 2000,
Pittack et al. 1997) have no overt eye phenotype (Miller et al. 2000). One possible explanation is that there are other Fgf family members or non-FGF ligands
(Kinoshita et al. 1995) expressed in surface ectoderm. Another more interesting possibility is that non-FGF signaling from the surface ectoderm may activate
autocrine FGF signaling within the presumptive neural retina. The upregulation
of Fgf15 expression in the neural retina region around the time of close contact (Nguyen & Arnheiter 2000) suggests that this factor is a good candidate for
fulfilling this function (Figure 3B).
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
273
Early Patterning of the Neural Retina
The earliest known transcription factor expressed in the presumptive retinal region
upon close contact with the surface ectoderm is encoded by the paired-like homeobox gene Chx10 (Liu et al. 1994). Chx10 is essential for normal eye development,
as mutations in humans and mice lead to microphthalmia, cataracts, and abnormal iris development (Burmeister et al. 1996, Ferda Percin et al. 2000). Chx10
expression first appears in the presumptive neural retina during the period of close
contact between the surface ectoderm and optic vesicle. Explant studies support
the idea that Chx10 expression in the presumptive neural retina occurs in response
to inductive signals from the presumptive lens ectoderm (Nguyen & Arnheiter
2000). The appearance of differentiated retinal cell types (with the exception of
bipolar cells) in Chx10 mutants suggests that Chx10 does not play a major role in
specifying cell fates within the neural retina. Instead, Chx10 appears to regulate
cell proliferation because mouse Chx10 mutants are characterized by a reduction
in the proliferation of retinal progenitor cells (Burmeister et al. 1996).
In mice, the basic helix-loop-helix transcription factor Mitf (microphthalmia
associated transcription factor) is initially expressed throughout the dorsal optic
vesicle, but is down-regulated in the presumptive retinal region shortly after the
onset of Chx10 expression (Bora et al. 1998). Interestingly, the dorsal RPE region of some mouse Mitf loss-of-function mutants transdifferentiates into neural
retina, suggesting that Mitf down-regulation is required for retinal specification
(Bumsted & Barnstable 2000). It has been speculated that formation of the secondary, transdifferentiated neural retina in Mitf mutants depends on signals such
as BMP7, which are expressed on the dorsal side of the optic cup (Nguyen &
Arnheiter 2000, Wawersik et al. 1999). If this were the case, then one might also
predict a similar role for BMP signaling during formation of the normal neural retina. This possibility is interesting in light of the observation that Bmp7−/−
mutants are anophthalmic. These mutants are believed to harbor a defect that manifests itself during the period of close contact between the optic vesicles and surface
ectoderm (Luo et al. 1995, Wawersik et al. 1999).
RPE Specification is Dependent on Signals
from Extraocular Mesenchyme
A large body of work has focused on the role of surface ectoderm and FGF
in neural retina formation; however, very little is known about the role of the
mesenchyme surrounding the eye. The transcription factors Foxc1/Mf1 and
Foxc2/Mfh1 are expressed in the extraocular mesenchyme, but neither gene plays
an essential role in early eye development (Bellusci et al. 1997, Kidson et al.
1999, Kume et al. 1998). Explant studies in chick suggest that extraocular mesenchyme may play an important role in RPE development (Fuhrmann et al. 2000).
In optic vesicle explants lacking extraocular mesenchyme, initiation of RPE
markers such as Mitf (which in the chick does not show an initial pan-optic
13 Sep 2001
13:49
274
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
vesicular expression pattern as it does in the mouse) does not occur. Moreover, the
expression of the neural retinal marker Chx10 (already up-regulated at the time of
explantation) expands throughout the entire explant, suggesting that extraocular
mesenchyme is required for both RPE development and for inhibition of neural
retina development. The ability of exogenously applied activin A to rescue Mitf
expression and RPE development in explants lacking mesenchyme makes it a good
candidate for a secreted mesenchymal factor that induces RPE (Fuhrmann et al.
2000). This possibility is supported by the presence of activin type IIA and IIB
receptors in the optic vesicle (Stern et al. 1995) and activin βA and βB subunits
in extraocular mesenchyme (Dohrmann et al. 1993, Feijen et al. 1994, Fuhrmann
et al. 2000).
Dorsal/Ventral Polarity in the Neural Retina
Several genes display restricted expression domains along the dorso-ventral axis
of the neural retina (Figure 3D). Pax2 (Dressler et al. 1990, Nornes et al. 1990)
and Vax2 (Barbieri et al. 1999, Schulte et al. 1999) (an Emx-related homeobox
gene) are both expressed early in the ventral half, whereas Xbr-1 (Papalopulu &
Kintner 1996) (a Xenopus homeobox gene) and Tbx5 (a T box gene) (GibsonBrown et al. 1998, Isaac et al. 1998, Logan et al. 1998) are restricted to the dorsal
half. Misexpression of Vax2 in the chick dorsal neural retina leads to the ectopic
expression of the ventral markers Pax2, EphB2, and EphB3 (EphB2 and EphB3 are
receptor tyrosine kinases), and the down-regulation of the dorsal markers Tbx5,
EphrinB2, and EphrinB3 (Ephrins B2 and B3 are Eph ligands) (Schulte et al.
1999). In contrast, Tbx5 misexpression in the ventral neural retina leads to the
down-regulation of both Vax2 and Pax2 and the ectopic induction of EphrinB2 and
EphrinB3 (Koshiba-Takeuchi et al. 2000). In both cases, Vax2 and Tbx5 misexpression leads to the abnormal projection of dorsal and ventral retinal ganglion cells,
indicating that retinal cell fate has been altered (Koshiba-Takeuchi et al. 2000,
Schulte et al. 1999). These experiments suggest that Vax2 and Tbx5 play important
roles in specifying ventral and dorsal fates and also highlight the potential role of
Ephrin signaling in the determination of dorsal/ventral polarity in the neural retina.
Experiments in zebrafish and mice have provided evidence suggesting that
retinoic acid (RA) plays a role in establishing ventral polarity within the developing
optic cup (Figure 3D). This is illustrated by the ability of RA-soaked beads to
induce a secondary choroidal fissure in any region of the zebrafish optic cup
(Hyatt et al. 1996). RA treatment also results in an enlarged optic stalk and in the
expansion of Pax [b] (a zebrafish homologue of Pax2) expression in dorsal regions
of the eye (Hyatt et al. 1996). Interestingly, experiments using retinoid-responsive
transgenic reporter mice revealed prominent dorsal neural retina expression at
early stages of development (E11.5) that relocates to ventral regions at later stages
(E14.5) (Schulte et al. 1999). In contrast to zebrafish data, this study also showed
that although RA is important for the morphological changes involved in choroidal
fissure formation, it can not ventralize the neural retina (Schulte et al. 1999). While
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
275
these differing results suggest the possibility of species or experimental differences,
they still demonstrate an important role for retinoids in defining cell fate along the
dorso-ventral axis. This is further supported by expression studies showing the
prominent expression of genes involved in retinoid synthesis in the ventral retina
(Grun et al. 2000, Li et al. 2000, Mic et al. 2000, Suzuki et al. 2000).
A role for RA in establishing ventral characteristics of the mammalian eye has
also been demonstrated in RA receptor loss-of-function mutants (Kastner et al.
1994). RA receptors are developmentally important ligand-inducible transcription
factors that are divided into two classes: the RARs (which bind to all-trans-RA and
9-cis-RA) and the RXRs (which only bind 9-cis-RA) (Chambon 1996). Rxrα −/−
mice are characterized by a reduction in size of the ventral retina and by the failure
of the optic fissure to close, which results in coloboma of the optic nerve (Kastner
et al. 1994). Furthermore, in several Rar and Rxr double-knockout mice, these
ventral phenotypes are accentuated by additional defects including the bilateral
eversion of the ventral retina and the absence of ventral iris (Kastner et al. 1994).
Anterior-Posterior Polarity in the Developing Eye
The developing optic vesicles and neural retina also display patterning along
the anterior (nasal)–posterior (temporal) axis. This is evident at the time of
optic evagination for the genes encoding the winged-helix transcription factors
BF-1/Foxg1 (see above), which is restricted to the anterior half, and BF-2/Foxd2,
which is restricted to the posterior half (Hatini et al. 1994). In BF-1-deficient mice,
ectopic BF-2 expression is present in the anterior optic vesicle (Huh et al. 1999),
suggesting that these two genes may regulate each other in the same manner as
Pax2 and Pax6 (see above).
Two homeodomain transcription factors of the HMX family, SOHo1 and GH6,
are expressed in the anterior neural retina (Deitcher et al. 1994, Stadler & Solursh
1994). Misexpression studies in chick have shown that while either SOHo1 or
GH6 can down-regulate the expression of EphA3, which is normally restricted
to the posterior neural retina, neither has major impact on the positional identity
of the temporal retina as assessed by mapping of retinal ganglion axons to the optic
tectum (Schulte & Cepko 2000). Therefore, unlike Tbx5 and Vax2, which appear
to be sufficient to specify dorsal/ventral fates, it is unclear whether SOHo1 and
GH6 function in specifying anterior/posterior fates in the neural retina.
LENS INDUCTION
In 1901, Spemann made the profound observation that ablation of the presumptive
retinal region of neural plate-stage Rana temporaria embryos resulted not only in
the absence of retinal development but also in the loss of lens formation (Spemann
1901). As early fate mapping had shown that the lens was derived from a region
outside of the one ablated, Spemann argued that development of the lens depended
13 Sep 2001
13:49
276
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
on presumptive retina. This observation not only illustrated the importance of
tissue interactions during eye development but also provided the first experimental
evidence leading to the idea of embryonic induction (Hamburger 1988).
Spemann’s conclusions on lens induction rapidly became controversial, however, with the observation by Mencl (1903) that a mutant salmon without an optic
cup could still generate lenses. Indeed, to date there remain a series of studies
in which optic cup ablation experiments have produced different outcomes, all
leading to the idea that there is genuine species-to-species variation in the timing
with which lens induction signals are delivered.
Lens Induction is a Multi-Step Process
Using modern techniques, Grainger and colleagues have reexamined the role of
early tissue interactions in lens induction and have operationally defined several
stages that correspond to (a) a period of lens-forming competence in the mid/late
gastrula ectoderm; (b) the acquisition of a lens-forming bias throughout the head
ectoderm during neurulation; (c) specification of lens cell fate toward the end of
neurulation; and (d ) differentiation, an aspect of lens development that continues
throughout life (Grainger 1992).
A brief period of lens forming competence is maximal in ectoderm of the midto-late gastrula and has been defined based on the ability of ectoderm to form
lenses when transplanted to the presumptive lens-forming region of neural plate–
stage embryos (Servetnick & Grainger 1991). Lens-forming competence may be a
cell-autonomous event since its timing appears to remain unchanged in ectoderm
that is pre-cultured in isolation (Servetnick & Grainger 1991).
The acquisition of a lens-forming bias throughout the head ectoderm is thought
to be mediated, in part, by planar signals derived from the anterior neural plate,
which are qualitatively different from those provided by the optic vesicle (Grainger
1992, Grainger et al. 1997). Evidence for this comes from experiments demonstrating that competent ectoderm displays a stronger lens-induction response when
grafted to the lens-forming region of neural plate–stage embryos than when grafted
to the lens-forming region overlying the optic vesicle (Henry & Grainger 1990).
The lens-inducing activity of the anterior neural plate may be enhanced by dorsal
mesoderm but inhibited by ventral mesoderm (Henry & Grainger 1990), consistent
with the proposal that lens induction is the result of cumulative inductive signals
from several tissue types (Jacobson 1966). Grafting experiments similar to those
performed by Lewis (Lewis 1904) showed that the optic vesicle is able to induce
ectopic lenses in head ectoderm outside of the presumptive lens region, but cannot
do so in the flank, suggesting that the entire head ectoderm of neural tube–stage
embryos has a lens-forming bias (Grainger et al. 1997).
Lens specification is defined as the ability of ectoderm to form lens tissue
when cultured in isolation and is usually characterized by the formation of small,
crystallin-expressing, lens-like structures referred to as lentoids. In Xenopus, lens
specification is established just as the optic vesicles come into contact with the
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
277
overlying lens ectoderm (Henry & Grainger 1990). In mouse and chick, this has
been reported to occur before close contact (Karkinen-Jaaskelainen 1978), although a more recent report suggests that specification occurs after the contact between the optic vesicle and the surface ectoderm is established (Furuta & Hogan
1998). The embryological manipulations described in this section indicate that
there are multiple stages to lens induction. The current challenge is to determine
which molecular events accompany these distinct stages.
Pax6 is Essential for Lens Induction
Pax6 expression in the presumptive lens region of vertebrates is first detected at
early neurula stages (Grindley et al. 1995; Li et al. 1997, 1994; Walther & Gruss
1991). In Xenopus, this expression appears shortly after that of Pax6 in the anterior neural plate and is clearly visible as a distinct band of expression lateral
and adjacent to the region of Pax6 expression in the anterior neural plate (Li
et al. 1997). Pax6 expression continues throughout the surface ectoderm overlying
the optic vesicles, but as the lens begins to differentiate, it becomes restricted
to the proliferating lens epithelial cells (Grindley et al. 1995). This is consistent
with the involvement of PAX6 in the activation of crystallins such as αB-, αA-, δ1-,
and ζ -crystallins expressed in the lens epithelium (Cvekl et al. 1995a,b, 1994;
Richardson et al. 1995) and in the repression of the fiber cell–specific βB1crystallin (Duncan et al. 1998).
Expression of Pax6 in the developing lens is directed, in part, by a 341 base pair
enhancer located in the 50 region 4.6 kb upstream of the Pax6 gene (Williams et al.
1998). This enhancer is highly conserved in the Pax6 gene of mouse, humans,
and puffer fish (Fugu) (Kammandel et al. 1999, Williams et al. 1998), raising
the possibility that it has played an important role during the evolution of lenslike structures. In reporter transgenic mice, this enhancer first begins to direct
expression in a few cells of the surface ectoderm overlying the optic vesicles at
around the time of close contact (Williams et al. 1998). Following separation of the
lens vesicle, enhancer-mediated expression continues in surface ectoderm destined
to become corneal epithelium (Kammandel et al. 1999, Williams et al. 1998). At
later stages, enhancer-dependent expression is observed in the epithelia of the
lacrimal gland and conjunctiva (Kammandel et al. 1999), and in the fully formed
lens, expression becomes restricted to the undifferentiated epithelium (Kammandel
et al. 1999, Williams et al. 1998). At all stages, the expression pattern coincides
with the normal expression pattern of Pax6 in the lens placode and its derivatives
(Grindley et al. 1995).
Several observations suggest a cell-autonomous requirement for Pax6 in the
developing lens. In animals chimeric for Sey/Sey and wild-type cells, Sey/Sey cells
are excluded from the presumptive lens at E9.5 (Collinson et al. 2000) and from
the maturing lens at E12.5 (Quinn et al. 1996). Furthermore, in explant recombination experiments using the rat Pax6 mutant rSey, presumptive lens ectoderm from
rSey/rSey animals did not form lenses when cultured with wild-type optic vesicles,
13 Sep 2001
13:49
278
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
whereas wild-type presumptive lens ectoderm did when recombined with either
wild-type or rSey/rSey optic vesicles (Fujiwara et al. 1994). This experiment
demonstrates an essential, autonomous component for Pax6 in the developing
lens and also shows that rSey mutant optic vesicles retain their lens-inducing capacities. In Xenopus ectodermal explants, Pax6 misexpression can lead to the cellautonomous formation of β-crystallin-expressing lentoids in the absence of neural
or mesodermal tissues, which are possible lens-inducing sources (Altmann et al.
1997). The majority of ectopic lenses in Pax6-misexpressing explants and embryos
are not fully formed, suggesting the requirement of additional lens differentiating
factors (Altmann et al. 1997).
Conditional mouse mutants in which Pax6 function in the prospective lens ectoderm is lost at the onset of lens placode formation have confirmed the requirement
for Pax6 in lens development (Ashery-Padan et al. 2000). These mutants were
characterized by the absence of lens placode formation, as well Six3 and Prox1 expression in the tissue that would have given rise to the lens placode. This suggests
that Six3 and Prox1 lie downstream of Pax6 placode in the lens pathway (Figure 6).
Expression of Sox2, however, was present in the surface ectoderm overlying the
optic vesicles at E10. Although Sox2 expression is lost in the surface ectoderm of
Sey/Sey unconditional mutants at E9.5, it remains unclear whether Pax6 is essential
for the low level of Sox2 expression in the preplacode (Figure 6) (Furuta & Hogan
1998, Wawersik et al. 1999).
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
→
Figure 6 A model for the genetic pathway regulating lens induction. The arrows indicate genetic interactions. Where the arrows are gray, this indicates a possible pathway
(i.e. inferred from expression pattern or misexpression studies) rather than a demonstrated pathway element (black). At the apex of the hierarchy is the pre-placodal phase
of Pax6 expression (Pax6 pre-placode). It is understood that the later phase of Pax6 expression in the lens placode (Pax6 placode) is dependent upon earlier activity of Pax6.
The genetic pathway described reflects this interaction. Pax6 placode is apparently regulated by at least two enhancers. One has been identified and characterized (ectoderm
enhancer), whereas the existence of the other (enhancer 2) is implied. Because FGF
and Bmp7 signaling cooperate to maintain the placodal phase of Pax6 expression, it
follows that their input to the pathway must be upstream of Pax6placode. Previous analysis has shown that the early phase of Pax6 expression is unaffected in Bmp7 null
mice, and thus FGF and Bmp7 signaling must converge on the pathway downstream
of Pax6 pre-placode. Based on the absence of Foxe3 expression in mice carrying a deletion of the Pax6 ectoderm enhancer (designated Pax61EE/1EE), Foxe3 is downstream
of Pax6 placode. The similar phenotypes (reduced lens lineage proliferation and lens
vesicle closure and separation failure) of the dysgenetic lens (Foxe3) and Pax61EE/1EE
mutant mice argue that Foxe3 is upstream of these cellular responses. Although we do
not understand the genetic relationship between Foxe3 and Sox2, it is clear from this
and previous analyses that Sox2 lies downstream of Pax6 placode. Because Sox2 (but not
Pax6) expression is diminished in Bmp4 null mice, Bmp4 signaling must contribute to
the pathway between Pax6 placode and Sox2.
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
279
Studies of the Sey Neu mutant have defined two distinct phases of Pax6 expression. The SeyNeu allele is a point mutation, and as a result, it is still possible to follow
transcription of Pax6 using in situ hybridization. Transcription in the head ectoderm is undisturbed in the absence of functional Pax6 product, but Pax6 expression
later in the presumptive lens region is absent (Grindley et al. 1995). This indicates
that the late phase of Pax6 expression (defined as Pax6 placode) depends on the activity of Pax6 in the head ectoderm (a phase defined previously as Pax6 pre-placode)
(see Wawersik et al. 1999). The genetic pathway for lens induction thus includes
two distinct phases of Pax6 expression (Figure 6).
Based on an understanding of the activity of the Pax6 upstream ectoderm enhancer, it is reasonable to propose that this element is responsible, at least in
part, for the placodal phase of Pax6 expression. This has been borne out by genetargeting experiments in which the enhancer has been deleted (P. Dimanlig &
R.A. Lang, submitted). Interestingly, lens development in homozygous deletion
mice (designated Pax61EE/1EE) was not prevented completely but was inhibited
13 Sep 2001
13:49
280
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
at all stages. The lens placode was thinner, the lens pit smaller than usual, and
the level of proliferation in the lens lineage was diminished. However, Pax6 immunoreactivity was still detectable in the lens placode, although at a lower level
than in wild-type. This suggests that besides the upstream ectoderm enhancer,
at least one other transcriptional control element regulates the placodal phase of
Pax6 expression. Thus we can extend the model of the genetic pathway controlling lens development to include at least two placodal phase enhancers (Figure 6).
We predict that a continuation of studies that have identified transcription control
elements for the Pax6 gene (Kammandel et al. 1999, Williams et al. 1998, Xu et al.
1999b) will identify an additional element with activity in the lens lineage.
Bmp4 and Bmp7 Play Important Roles in Lens Development
Bmp4 and Bmp7 display overlapping expression patterns within the developing
eye and appear to play important roles in the early stages of lens development. The
consequence of Bmp7 gene targeting in homozygous mice is a variably penetrant
phenotype that ranges from mild microphthalmia to anophthalmia (Dudley et al.
1995). In severely affected null animals, the lens placode fails to form, and the
placodal phase of Pax6 expression is lost (Wawersik et al. 1999). This has led
to the suggestion that Bmp7 plays an important role upstream of Pax6 during
lens placode formation and specification (Wawersik et al. 1999). Furthermore, the
timing of this proposed role for Bmp7 in lens induction and in the control of Pax6
expression in the lens placode argues that Bmp7 signaling during this period of lens
development is directed toward activating the Pax6 enhancers active in placodal
ectoderm (Figure 6).
Bmp4 null mouse embryos do not survive past E10.5 (Furuta & Hogan 1998).
Despite this, explant culture studies have shown that lens formation from the
presumptive lens ectoderm of null mutants could be rescued if cultured in the
presence of wild-type optic vesicles (Furuta & Hogan 1998). This indicates that
Bmp4 expression in the ectoderm is not essential for lens formation (Furuta &
Hogan 1998). Lens formation in Bmp4 null tissue could also be rescued if complete
eye rudiments were explanted in the presence of Bmp4. In contrast, lens formation
was absent if isolated ectoderm was cultured in the presence of Bmp4. These results
suggest that Bmp4 activity combines with at least one other signal to permit optic
vesicle-induced lens induction. It is possible that Bmp4 and the proposed second
signal stimulate presumptive lens ectoderm simultaneously or equally and that
Bmp4 acts within the optic vesicle to activate production of a signal that in turn
stimulates lens formation (Furuta & Hogan 1998).
Bmp4 and Bmp7 mutant embryos display similar lens phenotypes in that the lens
placode does not form and expression of the transcription factor Sox2 (see above)
in lens lineage ectoderm is not up-regulated in the normal way. There are some
distinct differences, however, between the two mutants. Whereas Pax6 expression
in the presumptive lens ectoderm is lost in Bmp7−/− mutants just prior to the time
lens placode formation should be occurring (Wawersik et al. 1999), the expression
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
281
of Pax6 (as well as Six3, Eya1, Eya2) in Bmp4−/− mutants is normal at comparable
stages (Furuta & Hogan 1998). These observations indicate that Bmp4 and Bmp7
may have non-redundant functions in the developing eye (Wawersik et al. 1999)
(Figure 6).
FGF Receptor Signaling is Required for Lens Induction
It is well established that FGF signaling is important for differentiation within the
lens lineage. Both loss- and gain-of-function experiments have shown that FGF
signaling is necessary and sufficient for lens fiber cell differentiation (reviewed
in McAvoy et al. 1991). Recent results reveal that FGF signaling is also required
for the early phases of lens development including lens induction (S. Faber &
R.A. Lang, in press) (Figure 6).
The involvement of FGF was addressed by expressing a dominant-negative FGF
receptor in the cells of interest by establishing transgenic mouse lines in which the
mutant FGF receptor was expressed from the Pax6 ectoderm enhancer coupled to
the P0 promoter (S. Faber and R.A. Lang, in press). Transgenic lines (designated
TFR7) expressing this construct showed defects in lens development that included
diminished proliferation at E13.5 and a delay in fiber cell differentiation. Of most
interest, however, was the observation of early lens development defects, including
a small lens pit and a failure of the lens vesicle to separate from the surface
ectoderm. These defects suggest a role for FGF receptor signaling in the inductive
phases of lens development.
The question of whether FGF signaling is involved in lens induction was
pursued by determining whether Bmp7, an established lens induction stimulus
(Wawersik et al. 1999), cooperates with FGF receptor activity. Indeed, crosses
between Bmp7 null (Dudley et al. 1995) and TFR7 mice produced compound
genotype mice with more severe lens development defects. In particular, TFR7,
Bmp7+/− mice showed a very small lens pit and failed to separate and close the
lens vesicle. When combined with the observation that compound genotype mice
show diminished expression levels of the lens lineage markers Pax6, Foxe3, and
Sox2, this indicates a genetic interaction between FGF and Bmp7 signaling. Thus
FGF receptor signaling, Bmp7, and the pre-placodal phase of Pax6 expression are
likely to converge in activating the placodal phase of Pax6 expression (Figure 6).
Analysis of Foxe3 and Sox2 expression in TFR7 Bmp7+/− mice (S. Faber &
R.A. Lang, submitted) implies that these genes lie downstream of FGF and Bmp7
signaling in the genetic control of lens induction (Figure 6).
The identity of FGF ligands involved in lens induction remains unclear and
represents a future challenge. Good candidates include Fgf8 (Lovicu & Overbeek
1998, Vogel-Hopker et al. 2000) and Fgf15 (McWhirter et al. 1997). Although
expressed only at low levels, Fgf8 mRNA is detectable in the eye at an early stage
of lens development (S. Faber, H. Makarenkova & R.A. Lang, unpublished observation). The ability of Fgf8 to induce fiber cell differentiation when misexpressed
(Lovicu & Overbeek 1998, Vogel-Hopker et al. 2000) indicates that it can be active
13 Sep 2001
13:49
282
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
within the lens lineage. Fgf15 is expressed in the optic vesicle and from this
point of view may be a good candidate for an FGF family ligand involved in lens
induction.
Sox1, Sox2, and Sox3 Participate in Lens Development
Members of the Group B1 SOX family of transcription factors are implicated
in lens development. Sox genes encode transcription factors containing a sexdetermining factor (SRY)–related HMG box. They were originally implicated in
lens differentiation based on the expression cloning of Sox2 in a screen to detect
proteins that could interact with a region of the chicken δ1-crystallin enhancer
(Kamachi et al. 1995). SOX2 has been shown to bind cooperatively with Pax6
to the δ-crystallin minimal enhancer DC5, and in transactivation studies, SOX2
and Pax6 can synergize (Kamachi et al. 2001). When transfected into chick embryos, SOX2 alone is not sufficient to induce ectopic lens tissue; however, when
co-misexpressed with Pax6, lens tissue is induced cell-autonomously in surface
ectoderm outside of the eye (Kamachi et al. 2001). SOX3 misexpression in Medaka
is sufficient on its own to induce ectopic lens formation cell-autonomously (Koster
et al. 2000).
Sox1, 2, and 3 display broad patterns of expression throughout the developing
embryo and possess similar expression patterns in the developing lens. During
lens development in mice and chick, Sox1 is first expressed during the invagination of the lens placode and later is predominantly expressed in lens fiber cells.
Sox2 and Sox3 expression appear prior to lens placode formation in ventral surface ectoderm bordering the presumptive lens region (Kamachi et al. 1998). Tissue
ablation experiments have demonstrated that Sox2 and 3 expression in the presumptive lens ectoderm depends on the presence of the optic vesicle (Kamachi et al.
1998).
A possible mechanistic explanation for the optic vesicle dependence of Sox2
expression has been provided by analysis of both Bmp4 null mice (Furuta & Hogan
1998) and those in which FGF signaling is suppressed in the presumptive lens (see
above). Where either of these signaling pathways is suppressed, Sox2 expression in
the lens lineage is diminished. Thus it is reasonable to propose that Bmp4 and one
or more FGF ligands directly or indirectly signal Sox2 up-regulation in presumptive
lens cells. When combined with the observation that Sox2 is down-regulated in
Sey/Sey embryos, these analyses suggest a downstream placement for Sox2 in a
genetic pathway regulating lens induction (Figure 6). The ability of SOX2 and
SOX3 misexpression to up-regulate Pax6 (Koster et al. 2000, Kamachi et al. 2001)
raises the interesting possibility that reciprocal gene regulation by SOX2/3 and
Pax6 might help to restrict and maintain the expression of these genes within the
lens placode (Kamachi et al. 2001) (Figure 6).
Homozygous mutant mice lacking Sox1 display an inhibition of fiber cell elongation that leads to cataract formation and microphthalmia (Nishiguchi et al. 1998).
Mutant lenses are characterized by the absence of γ -crystallins but the presence of
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
283
α- and β-crystallins, raising the possibility that SOX1 binding to the γ -crystallin
promoter is required in vivo for normal γ -crystallin gene activation. Furthermore,
the failure of fiber cell elongation in Sox1−/− mice is probably caused by the specific absence of γ -crystallin expression (Nishiguchi et al. 1998). Precedent for
this comes from the observation that a point mutation in the gene encoding γ E-crystallin is responsible for the mouse mutant, CatElo, in which lens fiber cell
differentiation is abnormal (Cartier et al. 1992).
fork head Transcription Factors Foxe3 and
Lens1 Lie Downstream of Pax6
Foxe3 and Lens1 are members of a sub-family of the fork head family of transcription factors that have restricted expression patterns limited to the lens lineage
(Blixt et al. 2000, Brownell et al. 2000, Kenyon et al. 1999) as well as the developing midbrain (Blixt et al. 2000, Brownell et al. 2000). Foxe3 is first expressed
at E8.75 in a small patch of presumptive lens ectoderm, which at this stage represents a subset of the Pax6-positive ectoderm. This is also true at E9.5 when the
lens placode has formed, and it becomes obvious at E10.5 when Pax6 is expressed
broadly in presumptive lens ectoderm and Foxe3 expression is restricted to the lens
pit. Foxe3 expression is absent in Sey/Sey animals, implying that it is positioned
downstream of Pax6 during lens induction (Brownell et al. 2000). This result is
consistent with the observation that Lens1 can be up-regulated by Pax6 in Xenopus
ectodermal explants and that Lens1, itself, can not induce Pax6 expression (Kenyon
et al. 1999).
In Xenopus misexpression studies, it was found that Lens1 can completely suppress lens differentiation but that genes indicative of presumptive lens ectoderm
such as Pax6 and Six3 are still expressed in the presumptive lens region. This
suggests that Lens1 may play a role in undifferentiated lens ectoderm/epithelia
following lens specification (Kenyon et al. 1999). A mutation in the fork head domain of the Foxe3 (predicted to be involved in DNA binding) is responsible for the
dysgenetic lens (dyl) mouse, and this leads to defects in lens vesicle closure and separation, as well as a reduction in the proliferation level of lens epithelial cells (Blixt
et al. 2000, Brownell et al. 2000). In the dyl mouse, expression of Prox1, which
encodes a transcription factor essential for lens fiber cell differentiation (Wigle
et al. 1999) and is normally most abundant in the transitional, equatorial region
of the lens, is expressed at high levels throughout the lens epithelium (Blixt et al.
2000, Brownell et al. 2000). This suggests that Foxe3 may function as a repressor
of Prox1 expression in the lens epithelium (Blixt et al. 2000, Brownell et al. 2000).
The phenotypic resemblance of the dyl mouse with those in which the Pax6 upstream ectoderm enhancer had been deleted (see above) prompted an examination
of a possible genetic relationship. This revealed that Foxe3 expression was diminished in the enhancer null mice (P. Dimanlig & R.A. Lang, in press). A more
precise genetic pathway can be suggested in which Foxe3 is located downstream
of the placodal phase of Pax6 expression and plays essential roles in regulating
13 Sep 2001
13:49
284
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
proliferation and differentiation in the lens epithelium following lens induction
(Figure 6).
The sine oculis–Related Transcription Factor
Six3 Can Transform Otic Vesicle to Lens
A role for Six3 in lens induction has been suggested from its expression in the
developing lens. Interestingly, however, there appear to be species differences in
the timing of its expression. In mice, Six3 expression in the developing lens first
appears during the formation of the lens placode and is later restricted to the lens
epithelium (Oliver et al. 1995). In Medaka fish, Six3 is expressed in the presumptive
lens ectoderm and like Optx2, is down-regulated in the lens placode prior to lens
differentiation (Loosli et al. 1998). As in Medaka, chick Six3 is expressed in the
presumptive lens ectoderm overlying the optic vesicles but persists in the lens
placode and is later localized to the lens epithelium (Bovolenta et al. 1998).
Misexpression studies in Medaka have shown that Six3 can lead to formation
of ectopic lenses (Oliver et al. 1996). In contrast to Pax6-induced ectopic lens
formation in Xenopus, Six3 misexpression is characterized by the transformation
to the otic vesicle into lens (Oliver et al. 1996). In addition, lineage tracing experiments revealed that Six3 can direct ectopic lens formation in a cell non-autonomous
manner, which led the authors to speculate that Six3 misexpression may induce a
soluble factor that changes the bias of the otic placode toward a lens fate. The timing of Six3 expression and its dependence on Pax6 expression in the lens placode
place it downstream of Pax6 placode (Figure 6).
Maf Family Members in Lens Development
Several members of the maf family of basic-leucine zipper transcription factors
have been implicated in lens induction and differentiation. L-maf is first expressed
in the lens placode and can up-regulate several crystallin genes when misexpressed
in Xenopus embryos and ectodermal explants or in chick neural retinal cultures
(Ishibashi & Yasuda 2001, Ogino & Yasuda 1998). c-maf has been shown to play
an essential role in lens differentiation: Homozygous c-maf loss-of-function mouse
mutants are microphthalmic, have a reduction in crystallin expression, and have
defects in lens fiber cell elongation (possibly as a consequence of reduced crystallin
expression) (Kawauchi et al. 1999, Kim et al. 1999, Ring et al. 2000). Xenopus
mafB is different from L-maf and c-maf in that it is expressed much earlier in
development in the presumptive lens ectoderm and is later restricted to the lens
epithelium of the mature lens (Ishibashi & Yasuda 2001). In Xenopus ectodermal
explants, XmafB can upregulate Pax6, Lens1, Six3, Sox3, and L-maf, as well as
several crystallin genes, although it is unable to induce ectopic lens formation when
misexpressed in whole embryos (Ishibashi & Yasuda 2001). These observations
suggest that XmafB may play a role in lens induction and in maintenance of the
lens epithelium; however, it is unclear how XmafB expression is regulated during
these stages of lens development (Figure 6).
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
285
CONCLUDING REMARKS
The last decade has witnessed tremendous progress in the effort to better understand the molecular mechanisms underlying vertebrate eye development, including
the identification of a large number of genes expressed in the developing eye and
the identification and characterization of genes responsible for eye disorders. We
are just beginning to understand how these molecules interact with one another
to regulate different aspects of eye development. A convergence of information
describing conserved vertebrate and invertebrate pathways is speeding progress.
The completion of genome sequencing projects should further facilitate the identification of genes that play a role in the developing eye.
Visit the Annual Reviews home page at www.AnnualReviews.org
LITERATURE CITED
Abdelhak S, Kalatzis V, Heilig R, Compain S,
Sampson D, et al. 1997. A human homologue
of the Drosophila eyes absent gene underlies branchio- and oto-renal (BOR) syndrome and identifies a novel gene family. Nat.
Genet. 15:157–64
Acampora D, Mazan S, Avantaggiato V, Barone P, Tuorto F, et al. 1996. Epilepsy and
brain abnormalities in mice lacking the Otx1
gene. Nat. Genet. 14:218–22
Adelmann HB. 1936a. The problem of cyclopia, Pt. I. Q. Rev. Biol. 11:161–82
Adelmann HB. 1936b. The problem of cyclopia, Pt. II. Q. Rev. Biol. 11:284–304
Altmann CR, Chow RL, Lang RA, HemmatiBrivanlou A. 1997. Lens induction by Pax-6
in Xenopus laevis. Dev. Biol. 185:119–23
Andreazzoli M, Gestri G, Angeloni D, Menna
E, Barsacchi G. 1999. Role of Xrx1 in Xenopus eye and anterior brain development. Development 126:2451–60
Ashery-Padan R, Marquardt T, Zhou X, Gruss
P. 2000. Pax6 activity in the lens primordium
is required for lens formation and for correct
placement of a single retina in the eye. Genes
Dev. 14:2701–11
Asuma N, Hirakiyama A, Inoue T, Asaka A,
Yamada M. 2000. Mutations of a human homologue of the Drosophila eyes absent gene
(EYA1) detected in patients with congenital
cataracts and ocular anterior segment abnormalities. Hum. Mol. Genet. 9:363–66
Barbieri AM, Lupo G, Bulfone A, Andreazzoli M, Mariani M, et al. 1999. A homeobox
gene, vax2, controls the patterning of the eye
dorsoventral axis. Proc. Natl. Acad. Sci. USA
96:10729–34
Barth KA, Wilson SW. 1995. Expression of
zebrafish nk2.2 is influenced by sonic hedgehog/vertebrate hedgehog-1 and demarcates a
zone of neuronal differentiation in the embryonic forebrain. Development 121:1755–68
Beebe DC. 1986. Development of the ciliary
body: a brief review. Trans. Ophthalmol. Soc.
UK 105:123–30
Beebe DC, Coats JM. 2000. The lens organizes
the anterior segment: specification of neural
crest cell differentiation in the avian eye. Dev.
Biol. 220:424–31
Belecky-Adams T, Tomarev S, Li HS, Ploder
L, McInnes RR, et al. 1997. Pax-6, Prox 1,
and Chx10 homeobox gene expression correlates with phenotypic fate of retinal precursor
cells. Invest. Ophthalmol. Vis. Sci. 38:1293–
303
Belloni E, Muenke M, Roessler E, Traverso G,
Siegel-Bartelt J, et al. 1996. Identification
of Sonic hedgehog as a candidate gene responsible for holoprosencephaly. Nat. Genet.
14:353–56
13 Sep 2001
13:49
286
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Bellusci S, Furuta Y, Rush MG, Henderson R,
Winnier G, Hogan BL. 1997. Involvement
of Sonic hedgehog (Shh) in mouse embryonic lung growth and morphogenesis. Development 124:53–63
Bernier G, Panitz F, Zhou X, Hollemann T,
Gruss P, Pieler T. 2000. Expanded retina territory by midbrain transformation upon overexpression of Six6 (Optx2) in Xenopus embryos. Mech. Dev. 93:59–69
Blixt A, Mahlapuu M, Aitola M, PeltoHuikko M, Enerback S, Carlsson P. 2000. A
forkhead gene, FoxE3, is essential for lens epithelial proliferation and closure of the lens
vesicle. Genes Dev. 14:245–54
Bonini NM, Bui QT, Gray-Board GL, Warrick JM. 1997. The Drosophila eyes absent
gene directs ectopic eye formation in a pathway conserved between flies and vertebrates.
Development 124:4819–26
Bonini NM, Leiserson WM, Benzer S.
1993. The eyes absent gene: genetic control of cell survival and differentiation in
the developing Drosophila eye. Cell 72:379–
95
Bora N, Conway SJ, Liang H, Smith SB.
1998. Transient overexpression of the microphthalmia gene in the eyes of Microphthalmia vitiligo mutant mice. Dev. Dynam.
213:283–92
Bovolenta P, Mallamaci A, Briata P, Corte G,
Boncinelli E. 1997. Implication of OTX2 in
pigment epithelium determination and neural
retina differentiation. J. Neurosci. 17:4243–
52
Bovolenta P, Mallamaci A, Puelles L, Boncinelli E. 1998. Expression pattern of cSix3,
a member of the Six/sine oculis family of
transcription factors. Mech. Dev. 70:201–3
Bron AJ, Tripathi RC, Tripathi BJ. 1997.
Wolff’s Anatomy of the Eye and Orbit. London: Chapman & Hall Med. 8th ed.
Brownell I, Dirksen M, Jamrich M. 2000. Forkhead Foxe3 maps to the dysgenetic lens locus
and is critical in lens development and differentiation. Genesis 27:81–93
Brun RB. 1981. The movement of the prospective eye vesicles from the neural plate into
the neural fold in Ambystoma mexicanum and
Xenopus laevis. Dev. Biol. 88:192–99
Bumsted KM, Barnstable CJ. 2000. Dorsal
retinal pigment epithelium differentiates as
neural retina in the microphthalmia (mi/mi)
mouse. Invest. Ophthalmol. Vis. Sci. 41:903–
8
Burmeister M, Novak J, Liang MY, Basu S,
Ploder L, et al. 1996. Ocular retardation
mouse caused by Chx10 homeobox null allele: impaired retinal progenitor proliferation
and bipolar cell differentiation. Nat. Genet.
12:376–84
Callaerts P, Halder G, Gehring WJ. 1997. PAX6 in development and evolution. Annu. Rev.
Neurosci. 20:483–532
Capdevila J, Johnson RL. 2000. Hedgehog signaling in vertebrate and invertebrate limb patterning. Cell. Mol. Life Sci. 57:1682–94
Cartier M, Breitman ML, Tsui LC. 1992. A
frameshift mutation in the gamma Ecrystallin gene of the Elo mouse. Nat. Genet.
2:42–45
Cepko CL, Austin CP, Yang X, Alexiades M,
Ezzeddine D. 1996. Cell fate determination
in the vertebrate retina. Proc. Natl. Acad. Sci.
USA 93:589–95
Chambon P. 1996. A decade of molecular biology of retinoic acid receptors. FASEB J.
10:940–54
Chen R, Amoui M, Zhang Z, Mardon G. 1997.
Dachshund and eyes absent proteins form a
complex and function synergistically to induce ectopic eye development in Drosophila.
Cell 91:893–903
Cheyette BN, Green PJ, Martin K, Garren
H, Hartenstein V, Zipursky SL. 1994. The
Drosophila sine oculis locus encodes a
homeodomain-containing protein required
for the development of the entire visual system. Neuron 12:977–96
Chiang C, Litingtung Y, Lee E, Young KE,
Corden JL, et al. 1996. Cyclopia and defective axial patterning in mice lacking Sonic
hedgehog gene function. Nature 383:407–
13
Chow RL, Altmann CR, Lang RA, HemmatiBrivanlou A. 1999. Pax6 induces ectopic
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
eyes in a vertebrate. Development 126:4213–
22
Chuang JC, Raymond PA. 2001. Zebrafish
genes rx1 and rx2 help define the region of
forebrain that gives rise to retina. Dev. Biol.
231:13–30
Collinson JM, Hill RE, West JD. 2000. Different roles for Pax6 in the optic vesicle and facial epithelium mediate early morphogenesis
of the murine eye. Development 127:945–56
Couly GF, Le Douarin NM. 1985. Mapping of
the early neural primordium in quail-chick
chimeras. I. Developmental relationships between placodes, facial ectoderm, and prosencephalon. Dev. Biol. 110:422–39
Couly GF, Le Douarin NM. 1987. Mapping of
the early neural primordium in quail-chick
chimeras. II. The prosencephalic neural plate
and neural folds: implications for the genesis
of cephalic human congenital abnormalities.
Dev. Biol. 120:198–214
Cvekl A, Kashanchi F, Sax CM, Brady JN,
Piatigorsky J. 1995a. Transcriptional regulation of the mouse alpha A-crystallin
gene: activation dependent on a cyclic AMPresponsive element (DE1/CRE) and a Pax6–binding site. Mol. Cell. Biol. 15:653–60
Cvekl A, Piatigorsky J. 1996. Lens development and crystallin gene expression: many
roles for Pax-6. BioEssays 18:621–30
Cvekl A, Sax CM, Bresnick EH, Piatigorsky
J. 1994. A complex array of positive and
negative elements regulates the chicken alpha A-crystallin gene: involvement of Pax-6,
USF, CREB and/or CREM, and AP-1 proteins. Mol. Cell. Biol. 14:7363–76
Cvekl A, Sax CM, Li X, McDermott JB, Piatigorsky J. 1995b. Pax-6 and lens-specific
transcription of the chicken delta 1-crystallin
gene. Proc. Natl. Acad. Sci. USA 92:4681–85
Czerny T, Halder G, Kloter U, Souabni A,
Gehring WJ, Busslinger M. 1999. twin of
eyeless, a second Pax-6 gene of Drosophila,
acts upstream of eyeless in the control of eye
development. Mol. Cell 3:297–307
Dattani MT, Martinez-Barbera JP, Thomas PQ,
Brickman JM, Gupta R, et al. 1998. Mutations in the homeobox gene HESX1/Hesx1
287
associated with septo-optic dysplasia in
human and mouse. Nat. Genet. 19:125–
33
de Iongh R, McAvoy JW. 1993. Spatio-temporal distribution of acidic and basic FGF
indicates a role for FGF in rat lens morphogenesis. Dev. Dyn. 198:190–202
Deitcher DL, Fekete DM, Cepko CL. 1994.
Asymmetric expression of a novel homeobox
gene in vertebrate sensory organs. J. Neurosci. 14:486–98
Del Rio-Tsonis K, Washabaugh CH, Tsonis
PA. 1995. Expression of pax-6 during urodele eye development and lens regeneration.
Proc. Natl. Acad. Sci. USA 92:5092–96
Detwiler SR, van Dyke RH. 1953. The induction of the neural retina from the pigment
epithelial layer of the eye. J. Exp. Zool.
128:367–83
Dohrmann CE, Hemmati-Brivanlou A, Thomsen GH, Fields A, Woolf TM, Melton DA.
1993. Expression of activin mRNA during
early development in Xenopus laevis. Dev.
Biol. 157:474–83
Dragomirov N. 1937. The influence of neighboring ectoderm on the organization of eye
rudiment. Dokl. AN SSSR 15:61–64
Dressler GR, Deutsch U, Chowdhury K,
Nornes HO, Gruss P. 1990. Pax2, a new murine paired-box-containing gene and its expression in the developing excretory system.
Development 109:787–95
Dudley AT, Lyons KM, Robertson EJ. 1995. A
requirement for bone morphogenetic protein7 during development of the mammalian kidney and eye. Genes Dev. 9:2795–807
Duncan MK, Haynes JI 2nd, Cvekl A, Piatigorsky J. 1998. Dual roles for Pax-6: a transcriptional repressor of lens fiber cell- specific beta-crystallin genes. Mol. Cell. Biol.
18:5579–86
Eagleson GW, Harris WA. 1990. Mapping of
the presumptive brain regions in the neural plate of Xenopus laevis. J. Neurobiol.
21:427–40
Echelard Y, Epstein D, St-Jacques B, Shen
L, Mohler J, et al. 1993. Sonic hedgehog,
a member of a family of putative signaling
13 Sep 2001
13:49
288
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
molecules, is implicated in the regulation of
CNS polarity. Cell 75:1417–30
Ekker SC, Ungar AR, Greenstein P, von Kessler DP, Porter JA, et al. 1995. Patterning activities of vertebrate hedgehog proteins in the
developing eye and brain. Curr. Biol. 5:944–
55
Ericson J, Rashbass P, Schedl A, BrennerMorton S, Kawakami A, et al. 1997. Pax6
controls progenitor cell identity and neuronal
fate in response to graded Shh signaling. Cell
90:169–80
Favor J, Sandulache R, Neuhauser-Klaus A,
Pretsch W, Chatterjee B, et al. 1996. The
mouse Pax2(1Neu) mutation is identical to a
human PAX2 mutation in a family with renalcoloboma syndrome and results in developmental defects of the brain, ear, eye, and kidney. Proc. Natl. Acad. Sci. USA 93:13870–75
Feijen A, Goumans MJ, van den Eijnden-van
Raaij AJ. 1994. Expression of activin subunits, activin receptors and follistatin in
postimplantation mouse embryos suggests
specific developmental functions for different activins. Development 120:3621–37
Feldman B, Gates MA, Egan ES, Dougan ST,
Rennebeck G, et al. 1998. Zebrafish organizer development and germ-layer formation require nodal- related signals. Nature
395:181–85
Ferda Percin E, Ploder LA, Yu JJ, Arici K,
Horsford DJ, et al. 2000. Human microphthalmia associated with mutations in the retinal homeobox gene CHX10. Nat. Genet.
25:397–401
Fernald RD. 2000. Evolution of eyes. Curr.
Opin. Neurobiol. 10:444–50
Fuhrmann S, Levine EM, Reh TA. 2000. Extraocular mesenchyme patterns the optic
vesicle during early eye development in the
embryonic chick. Development 127:4599–
609
Fujiwara M, Uchida T, Osumi-Yamashita N,
Eto K. 1994. Uchida rat (rSey): a new mutant rat with craniofacial abnormalities resembling those of the mouse Sey mutant. Differentiation 57:31–38
Furukawa T, Kozak CA, Cepko CL. 1997. rax,
a novel paired-type homeobox gene, shows
expression in the anterior neural fold and developing retina. Proc. Natl. Acad. Sci. USA
94:3088–93
Furuta Y, Hogan BLM. 1998. BMP4 is essential for lens induction in the mouse embryo.
Genes Dev. 12:3764–75
Gibson-Brown JJ, Silver LM, Papaioannou
VE. 1998. Expression of T-box genes Tbx2Tbx5 during chick organogenesis. Mech.
Dev. 74:165–69
Glardon S, Callaerts P, Halder G, Gehring
WJ. 1997. Conservation of Pax-6 in a lower
chordate, the ascidian Phallusia mammillata.
Development 124:817–25
Glaser T, Walton DS, Maas RL. 1992. Genomic structure, evolutionary conservation
and aniridia mutations in the human PAX6
gene. Nat. Genet. 2:232–39
Grainger RM. 1992. Embryonic lens induction:
shedding light on vertebrate tissue determination. Trends Genet. 8:349–55
Grainger RM, Mannion JE, Cook TLJ, Zygar CA. 1997. Defining intermediate stages
in cell determination: acquisition of a lensforming bias in head ectoderm during lens
determination. Dev. Genet. 20:246–57
Granadino B, Gallardo ME, Lopez-Rios J,
Sanz R, Ramos C, et al. 1999. Genomic
cloning, structure, expression pattern, and
chromosomal location of the human SIX3
gene. Genomics 55:100–5
Graw J. 1996. Genetic aspects of embryonic
eye development in vertebrates. Dev. Genet.
18:181–97
Grindley JC, Davidson DR, Hill RE. 1995. The
role of Pax-6 in eye and nasal development.
Development 121:1433–42
Grindley JC, Hargett LK, Hill RE, Ross A,
Hogan BL. 1997. Disruption of PAX6 function in mice homozygous for the Pax6Sey1Neu mutation produces abnormalities in the
early development and regionalization of the
diencephalon. Mech. Dev. 64:111–26
Gritsman K, Zhang J, Cheng S, Heckscher
E, Talbot WS, Schier AF. 1999. The EGFCFC protein one-eyed pinhead is essential
for nodal signaling. Cell 97:121–32
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
Grun F, Hirose Y, Kawauchi S, Ogura T,
Umesono K. 2000. Aldehyde dehydrogenase
6, a cytosolic retinaldehyde dehydrogenase
prominently expressed in sensory neuroepithelia during development. J. Biol. Chem.
275:41210–18
Guillemot F, Cepko CL. 1992. Retinal fate and
ganglion cell differentiation are potentiated
by acidic FGF in an in vitro assay of early
retinal development. Development 114:743–
54
Halder G, Callaerts P, Flister S, Walldorf U,
Kloter U, Gehring WJ. 1998. Eyeless initiates the expression of both sine oculis and
eyes absent during Drosophila compound
eye development. Development 125:2181–
91
Halder G, Callaerts P, Gehring WJ. 1995. Induction of ectopic eyes by targeted expression of the eyeless gene in Drosophila. Science 267:1788–92
Hallonet M, Hollemann T, Pieler T, Gruss
P. 1999. Vax1, a novel homeobox-containing
gene, directs development of the basal
forebrain and visual system. Genes Dev.
13:3106–14
Hamburger V. 1988. The Heritage of Experimental Embryology: Hans Spemann and the
Organizer. New York: Oxford Univ. Press
Hammerschmidt M, Bitgood MJ, McMahon
AP. 1996a. Protein kinase A is a common
negative regulator of Hedgehog signaling in
the vertebrate embryo. Genes Dev. 10:647–
58
Hammerschmidt M, Pelegri F, Mullins MC,
Kane DA, Brand M, et al. 1996b. Mutations
affecting morphogenesis during gastrulation
and tail formation in the zebrafish, Danio rerio. Development 123:143–51
Hanson IM, Seawright A, Hardman K, Hodgson S, Zaletayev D, et al. 1993. PAX6 mutations in aniridia. Hum. Mol. Genet. 2:915–20
Harland R. 2000. Neural induction. Curr. Opin.
Genet. Dev. 10:357–62
Harris WA. 1997. Cellular diversification in the
vertebrate retina. Curr. Opin. Genet. Dev.
7:651–58
Hatini V, Tao W, Lai E. 1994. Expression of
289
winged helix genes, BF-1 and BF-2, defines
adjacent domains within the developing forebrain and retina. J. Neurobiol. 25:1293–309
Hatta K, Kimmel CB, Ho RK, Walker C. 1991.
The cyclops mutation blocks specification of
the floor plate of the zebrafish central nervous
system. Nature 350:339–41
Hay ED. 1986. Development of the vertebrate
cornea. Int. Rev. Cytol. 63:263–322
Heberlein U, Singh CM, Luk AY, Donohoe
TJ. 1995. Growth and differentiation in the
Drosophila eye coordinated by hedgehog.
Nature 373:709–11
Heberlein U, Wolff T, Rubin GM. 1993.
The TGF beta homolog dpp and the segment polarity gene hedgehog are required for
propagation of a morphogenetic wave in the
Drosophila retina. Cell 75:913–26
Heisenberg CP, Brand M, Jiang YJ, Warga
RM, Beuchle D, et al. 1996. Genes involved
in forebrain development in the zebrafish,
Danio rerio. Development 123:191–203
Heisenberg CP, Nusslein-Volhard C. 1997. The
function of silberblick in the positioning of
the eye anlage in the zebrafish embryo. Dev.
Biol. 184:85–94
Henry JJ, Grainger RM. 1990. Early tissue interactions leading to embryonic lens formation in Xenopus laevis. Dev. Biol. 141:149–63
Hermesz E, Mackem S, Mahon KA. 1996.
Rpx: a novel anterior-restricted homeobox
gene progressively activated in the prechordal plate, anterior neural plate and
Rathke’s pouch of the mouse embryo. Development 122:41–52
Hill RE, Favor J, Hogan BL, Ton CC, Saunders GF, et al. 1991. Mouse small eye results
from mutations in a paired-like homeoboxcontaining gene. Nature 354:522–25
Hirsch N, Harris WA. 1997. Xenopus Pax-6 and
retinal development. J. Neurobiol. 32:45–61
Hogan BL, Horsburgh G, Cohen J, Hetherington CM, Fisher G, Lyon MF. 1986. Small
eyes (Sey): a homozygous lethal mutation on
chromosome 2 which affects the differentiation of both lens and nasal placodes in the
mouse. J. Embryol. Exp. Morphol. 97:95–
110
13 Sep 2001
13:49
290
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Hollemann T, Bellefroid E, Pieler T. 1998. The
Xenopus homologue of the Drosophila gene
tailless has a function in early eye development. Development 125:2425–32
Holtfreter J. 1939. Gewebeaffinitat, ein Mittel
der embryonalen Formbildung. Arch. Exp.
Zellforsch 23:169–209
Huh S, Hatini V, Marcus RC, Li SC, Lai E.
1999. Dorsal-ventral patterning defects in the
eye of BF-1-deficient mice associated with a
restricted loss of shh expression. Dev. Biol.
211:53–63
Hyatt GA, Schmitt EA, Marsh-Armstrong N,
McCaffery P, Drager UC, Dowling JE. 1996.
Retinoic acid establishes ventral retinal characteristics. Development 122:195–204
Hyer J, Mima T, Mikawa T. 1998. FGF1 patterns the optic vesicle by directing the placement of the neural retina domain. Development 125:869–77
Isaac A, Rodriguez-Esteban C, Ryan A, Altabef M, Tsukui T, et al. 1998. Tbx genes and
limb identity in chick embryo development.
Development 125:1867–75
Ishibash S, Yasuda K. 2001. Distinct roles of
maf genes during Xenopus lens development.
Mech. Dev. 101:155–66
Jacobson AG. 1966. Inductive processes in
embryonic development. Science 152:25–34
Jacobson C-O. 1959. The localization of the
presumptive cerebral regions in the neural
plate of the Axolotl larva. J. Embryol. Exp.
Morphol. 7:1–21
Jean D, Bernier G, Gruss P. 1999. Six6
(Optx2) is a novel murine Six3-related homeobox gene that demarcates the presumptive
pituitary/hypothalamic axis and the ventral
optic stalk. Mech. Dev. 84:31–40
Johnson KR, Cook SA, Erway LC, Matthews
AN, Sanford LP, et al. 1999. Inner ear and
kidney anomalies caused by IAP insertion in
an intron of the Eya1 gene in a mouse model
of BOR syndrome. Hum. Mol. Genet. 8:645–
53
Jordan T, Hanson I, Zaletayev D, Hodgson S,
Prosser J, et al. 1992. The human PAX6 gene
is mutated in two patients with aniridia. Nat.
Genet. 1:328–32
Kageyama R, Ishibashi M, Takebayashi K,
Tomita K. 1997. bHLH transcription factors
and mammalian neuronal differentiation. Int.
J. Biochem. Cell. Biol. 29:1389–99
Kamachi Y, Sockanathan S, Liu Q, Breitman
M, Lovell-Badge R, Kondoh H. 1995. Involvement of SOX proteins in lens-specific
activation of crystallin genes. EMBO J. 14:
3510–19
Kamachi Y, Uchikawa M, Collignon J, LovellBadge R, Kondoh H. 1998. Involvement of
Sox1, 2 and 3 in the early and subsequent
molecular events of lens induction. Development 125:2521–32
Kamachi Y, Uchikawa M, Tanouchi A, Sekido
R, Kondoh H. 2001. Pax6 and SOX2 form a
co-DNA-binding partner complex that regulates initiation of lens development. Genes
Dev. 15:In press
Kammandel B, Chowdhury K, Stoykova A,
Aparicio S, Brenner S, Gruss P. 1999. Distinct cis-essential modules direct the timespace pattern of the Pax6 gene activity. Dev.
Biol. 205:79–97
Karkinen-Jaaskelainen M. 1978. Permissive
and directive interactions in lens induction.
J. Embryol. Exp. Morphol. 44:167–79
Kastner P, Grondona JM, Mark M, Gansmuller
A, LeMeur M, et al. 1994. Genetic analysis
of RXR alpha developmental function: convergence of RXR and RAR signaling pathways in heart and eye morphogenesis. Cell
78:987–1003
Kawauchi S, Takahashi S, Hakajima O, Ogino
H, Morita M, et al. 1999. Regulation of lens
fiber cell differentiation by transcription factor c-Maf. J. Biol. Chem. 274:19254–60
Keller SA, Jones JM, Boyle A, Barrow LL,
Killen PD, et al. 1994. Kidney and retinal
defects (Krd), a transgene-induced mutation
with a deletion of mouse chromosome 19 that
includes the Pax2 locus. Genomics 23:309–
20
Kenyon KL, Moody SA, Jamrich M. 1999. A
novel fork head gene mediates early steps
during Xenopus lens formation. Development 126:5107–16
Kidson SH, Kume T, Deng K, Winfrey V,
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
Hogan BL. 1999. The forkhead/winged-helix
gene, Mf1, is necessary for the normal development of the cornea and formation of the
anterior chamber in the mouse eye. Dev. Biol.
211:306–22
Kim JI, Li T, Ho IC, Grusby MJ, Glimcher LH.
1999. Requirement for the c-Maf transcription factor in crystallin gene regulation and
lens development. Proc. Natl. Acad. Sci. USA
96:3781–85
Kinoshita N, Minshull J, Kirschner MW. 1995.
The identification of two novel ligands of the
FGF receptor by a yeast screening method
and their activity in Xenopus development.
Cell 83:621–30
Kobayashi M, Toyama R, Takeda H, Dawid
IB, Kawakami K. 1998. Overexpression of
the forebrain-specific homeobox gene six3
induces rostral forebrain enlargement in zebrafish. Development 125:2973–82
Koshiba-Takeuchi K, Takeuchi JK, Matsumoto K, Momose T, Uno K, et al. 2000.
Tbx5 and the retinotectum projection. Science 287:134–37
Koster RW, Kuhnlein RP, Wittbrodt PW. 2000.
Ectopic Sox3 activity elicits sensory placode
formation. Mech. Dev. 95:175–87
Krauss S, Concordet JP, Ingham PW. 1993.
A functionally conserved homolog of the
Drosophila segment polarity gene hh is expressed in tissues with polarizing activity in
zebrafish embryos. Cell 75:1431–44
Krauss S, Johansen T, Korzh V, Moens U, Ericson JU, Fjose A. 1991. Zebrafish pax[zfa]: a paired box-containing gene expressed
in the neural tube. EMBO J. 10:3609–19
Kumar JP, Moses K. 2001. EFG receptor
and Notch signaling act upstream of Eyeless/Pax6 to control eye specification. Cell
104:687–97
Kume T, Deng KY, Winfrey V, Gould DB,
Walter MA, Hogan BL. 1998. The forkhead/winged helix gene Mf1 is disrupted in
the pleiotropic mouse mutation congenital
hydrocephalus. Cell 93:985–96
Lang RA. 1999. Which factors stimulate lens
fiber cell differentiation in vivo? Invest. Ophthalmol. Vis. Sci. 40:3075–78
291
Lewis WH. 1904. Experimental studies on the
development of the eye in Amphibia. I. On
the origin of the lens, Rana palustris. Am. J.
Anat. 3:505–36
Li H, Tierney C, Wen L, Wu JY, Rao Y. 1997. A
single morphogenetic field gives rise to two
retina primordia under the influence of the
prechordal plate. Development 124:603–15
Li HS, Yang JM, Jacobson RD, Pasko D,
Sundin O. 1994. Pax-6 is first expressed in
a region of ectoderm anterior to the early
neural plate: implications for stepwise determination of the lens. Dev. Biol. 162:181–94
Li Z, Hu M, Ochocinska MJ, Joseph NM,
Easter SS, Jr. 2000. Modulation of cell proliferation in the embryonic retina of zebrafish
(Danio rerio). Dev. Dyn. 219:391–401
Liu IS, Chen JD, Ploder L, Vidgen D, van
der Kooy D, et al. 1994. Developmental expression of a novel murine homeobox gene
(Chx10): evidence for roles in determination
of the neuroretina and inner nuclear layer.
Neuron 13:377–93
Logan M, Simon HG, Tabin C. 1998. Differential regulation of T-box and homeobox transcription factors suggests roles in controlling chick limb-type identity. Development
125:2825–35
Loosli F, Koster R, Carl M, Krone A, Wittbrodt J. 1998. Six3, a medaka homologue of
the Drosophila homeobox gene sine oculis
is expressed in the anterior embryonic shield
and the developing eye. Mech. Dev. 74:159–
64
Loosli F, Winkler S, Wittbrodt J. 1999. Six3
overexpression initiates the formation of ectopic retina. Genes Dev. 13:649–54
Lopashov GV, Stroeva OG. 1964. Development
of the Eye. New York: Davey & Co.
Lovicu FJ, Overbeek PA. 1998. Overlapping
effects of different members of the FGF family on lens fiber differentiation in transgenic
mice. Development 125:3365–77
Luo G, Hofmann C, Bronckers AL, Sohocki
M, Bradley A, Karsenty G. 1995. BMP-7 is
an inducer of nephrogenesis, and is also required for eye development and skeletal patterning. Genes Dev. 9:2808–20
13 Sep 2001
13:49
292
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Ma C, Zhou Y, Beachy PA, Moses K. 1993.
The segment polarity gene hedgehog is required for progression of the morphogenetic
furrow in the developing Drosophila eye.
Cell 75:927–38
Macdonald R, Barth KA, Xu Q, Holder N,
Mikkola I, Wilson SW. 1995. Midline signalling is required for Pax gene regulation
and patterning of the eyes. Development
121:3267–78
Makarenkova HP, Ito M, Venkatesh G, Faber
SC, Sun L, et al. 2000. FGF10 is an inducer
and Pax6 a competence factor for lacrimal
gland development. Development 127:2563–
72
Manchot E. 1929. Abgranzung der Augenmaterials und anderer Teilbezirke in der Medullarplatte; die Teilbewegungen wahrend
der Auffaltung (Farbmarkierungsversuche
an Keimen von Urodelen). Arch. Entw.-Mech
116:689–708
Mangold O. 1931. Das Determinationproblem.
III. Das Wirbeltierauge in der Entwicklung
und Regeneration. Ergeb. Biol. 7:193–403
Mardon G, Solomon NM, Rubin GM. 1994.
dachshund encodes a nuclear protein required for normal eye and leg development in Drosophila. Development 120:3473–
86
Marlow F, Zwartkruis F, Malicki J, Neuhauss
SC, Abbas L, et al. 1998. Functional interactions of genes mediating convergent extension, knypek and trilobite, during the partitioning of the eye primordium in zebrafish.
Dev. Biol. 203:382–99
Marquardt T, Ashery-Padan R, Andrejewski N,
Scardigli R, Guillemot F, Gruss P. 2001. Pax6
is required for the multipotent state of retinal
progenitor cells. Cell 105:43–55
Martin P, Carriere C, Dozier C, Quatannens
B, Mirabel MA, et al. 1992. Characterization
of a paired box- and homeobox-containing
quail gene (Pax-QNR) expressed in the neuroretina. Oncogene 7:1721–28
Mathers PH, Grinberg A, Mahon KA, Jamrich
M. 1997. The Rx homeobox gene is essential for vertebrate eye development. Nature
387:603–7
Matsuo I, Kuratani S, Kimura C, Takeda N,
Aizawa S. 1995. Mouse Otx2 functions in
the formation and patterning of rostral head.
Genes Dev. 9:2646–58
Matsuo T, Osumi-Yamashita N, Noji S, Ohuchi H, Koyama E, et al. 1993. A mutation
in the Pax-6 gene in rat small eye is associated with impaired migration of midbrain
crest cells. Nat. Genet. 3:299–304
McAvoy JW. 1980. Cytoplasmic processes interconnect lens placode and optic vesicle
during eye morphogenesis. Exp. Eye Res.
31:527–34
McAvoy JW, Chamberlain CG, de Iongh RU,
Hales AM, Lovicu FJ. 1999. Lens development. Eye 13:425–37
McAvoy JW, Chamberlain CG, de Iongh RU,
Richardson NA, Lovicu FJ. 1991. The role
of fibroblast growth factor in eye lens development. Ann. NY Acad. Sci. 638:256–74
McWhirter JR, Goulding M, Weiner JA, Chun
J, Murre C. 1997. A novel fibroblast growth
factor gene expressed in the developing nervous system is a downstream target of the
chimeric homeodomain oncoprotein E2APbx1. Development 124:3221–32
Mencl E. 1903. Ein Fall von beiderseitiger Augenlinsenausbildung während der Abwesenheit von Augenblasen. Arch. Entw. Mech.
Org. 16:328–39
Mic FA, Molotkov A, Fan X, Cuenca AE,
Duester G. 2000. RALDH3, a retinaldehyde
dehydrogenase that generates retinoic acid,
is expressed in the ventral retina, otic vesicle
and olfactory pit during mouse development.
Mech. Dev. 97:227–30
Mikami Y. 1939. Reciprocal transformation of
the parts in the developing eye-vesicle, with
special reference to the inductive influence
of the lens-ectoderm on the retinal determination. Zool. Mag. 51:253–56
Miller DL, Ortega S, Bashayan O, Basch R,
Basilico C. 2000. Compensation by fibroblast growth factor 1 (FGF1) does not account for the mild phenotypic defects observed in FGF2 null mice. Mol. Cell. Biol.
20:2260–68. Erratum 2000. Mol. Cell. Biol.
20:3752
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
Muller F, Albert S, Blader P, Fischer N, Hallonet M, Strahle U. 2000. Direct action of
the nodal-related signal cyclops in induction
of sonic hedgehog in the ventral midline of
the CNS. Development 127:3889–97
Narahara K, Baker E, Ito S, Yokoyama Y, Yu
S, et al. 1997. Localisation of a 10q breakpoint within the PAX2 gene in a patient with
a de novo t(10;13) translocation and optic
nerve coloboma-renal disease. J. Med. Genet.
34:213–16
Neumann CJ, Nuesslein-Volhard C. 2000. Patterning of the zebrafish retina by a wave of
sonic hedgehog activity. Science 298:2137–
39
Nguyen M, Arnheiter H. 2000. Signaling
and transcriptional regulation in early mammalian eye development: a link between FGF
and MITF. Development 127:3581–91
Nieuwkoop PD. 1963. Pattern formation in artificially activated ectoderm (Rana pipiens and
Ambystoma punctatum). Dev. Biol. 7:255–
79
Nishiguchi S, Wood H, Kondoh H, LovellBadge R, Episkopou V. 1998. Sox1 directly
regulates the gamma-crystallin genes and is
essential for lens development in mice. Genes
Dev. 12:776–81
Nornes HO, Dressler GR, Knapik EW,
Deutsch U, Gruss P. 1990. Spatially and temporally restricted expression of Pax2 during murine neurogenesis. Development 109:
797–809
Ogino H, Yasuda K. 1998. Induction of lens
differentiation by activation of a bZIP transcription factor, L-Maf. Science 280:115–18
Oliver G, Loosli F, Koester R, Wittbrodt J,
Gruss P. 1996. Ectopic lens induction in fish
in response to the murine homeobox gene
Six3. Mech. Dev. 60:233–39
Oliver G, Mailhos A, Wehr R, Copeland NG,
Jenkins NA, Gruss P. 1995. Six3, a murine
homologue of the sine oculis gene, demarcates the most anterior border of the developing neural plate and is expressed during
eye development. Development 121:4045–
55
Opas M, Dziak E. 1994. bFGF-induced trans-
293
differentiation of RPE to neuronal progenitors is regulated by the mechanical properties
of the substratum. Dev. Biol. 161:440–54
Pan D, Rubin GM. 1998. Targeted expression
of teashirt induces ectopic eyes in Drosophila. Proc. Natl. Acad. Sci. USA 95:15508–
12
Pannese M, Polo C, Andreazzoli M, Vignali
R, Kablar B, et al. 1995. The Xenopus homologue of Otx2 is a maternal homeobox gene
that demarcates and specifies anterior body
regions. Development 121:707–20
Papalopulu N, Kintner C. 1996. A Xenopus
gene, Xbr-1, defines a novel class of homeobox genes and is expressed in the dorsal ciliary margin of the eye. Dev. Biol. 174:104–14
Park CM, Hollenberg MJ. 1989. Basic fibroblast growth factor induces retinal regeneration in vivo. Dev. Biol. 134:201–5
Patten I, Placzek M. 2000. The role of Sonic
hedgehog in neural tube patterning. Cell.
Mol. Life Sci. 57:1695–708
Pera EM, Kessel M. 1997. Patterning of the
chick forebrain anlage by the prechordal
plate. Development 124:4153–62
Pichaud F, Treisman J, Desplan C. 2001. Reinventing a common strategy for patterning the
eye. Cell 105:9–12
Pignoni F, Hu B, Zavitz KH, Xiao J, Garrity
PA, Zipursky SL. 1997. The eye-specification proteins So and Eya form a complex
and regulate multiple steps in Drosophila eye
development. Cell 91:881–91
Pittack C, Grunwald GB, Reh TA. 1997. Fibroblast growth factors are necessary for neural retina but not pigmented epithelium differentiation in chick embryos. Development
124:805–16
Pittack C, Jones M, Reh TA. 1991. Basic fibroblast growth factor induces retinal pigment epithelium to generate neural retina in
vitro. Development 113:577–88
Porter FD, Drago J, Xu Y, Cheema SS, Wassif
C, et al. 1997. Lhx2, a LIM homeobox gene,
is required for eye, forebrain, and definitive erythrocyte development. Development
124:2935–44
Pressman CL, Chen H, Johnson RL. 2000.
13 Sep 2001
13:49
294
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
LMX1B, a LIM homeodomain class transcription factor, is necessary for normal development of multiple tissues in the anterior
segment of the murine eye. Genesis 26:15–25
Puschel AW, Gruss P, Westerfield M. 1992.
Sequence and expression pattern of pax-6
are highly conserved between zebrafish and
mice. Development 114:643–51
Quinn JC, West JD, Hill RE. 1996. Multiple
functions for Pax6 in mouse eye and nasal
development. Genes Dev. 10:435–46
Quiring R, Walldorf U, Kloter U, Gehring WJ.
1994. Homology of the eyeless gene of
Drosophila to the Small eye gene in mice and
Aniridia in humans. Science 265:785–89
Rasmussen JT, Deardorff MA, Tan C, Rao MS,
Klein PS, Vetter ML. 2001. Regulation of eye
development by frizzled signaling in Xenopus. Proc. Natl. Acad. Sci. USA 98:3861–66
Reh TA, Jones M, Pittack C. 1991. Common
mechanisms of retinal regeneration in the larval frog and embryonic chick. Ciba. Found.
Symp. 160:192–204
Rhinn M, Dierich A, Shawlot W, Behringer
RR, Le Meur M, Ang SL. 1998. Sequential roles for Otx2 in visceral endoderm
and neuroectoderm for forebrain and midbrain induction and specification. Development 125:845–56
Richardson J, Cvekl A, Wistow G. 1995. Pax6 is essential for lens-specific expression of
zeta-crystallin. Proc. Natl. Acad. Sci. USA
92:4676–80
Riddle RD, Johnson RL, Laufer E, Tabin C.
1993. Sonic hedgehog mediates the polarizing activity of the ZPA. Cell 75:1401–16
Ring BZ, Cordes SP, Overbeek PS, Barsh GS.
2000. Regulation of mouse lens fiber cell
development and differentiation by the Maf
gene. Development 127:307–17
Roessler E, Belloni E, Gaudenz K, Jay P,
Berta P, et al. 1996. Mutations in the human Sonic Hedgehog gene cause holoprosencephaly. Nat. Genet. 14:357–60
Royet J, Finkelstein R. 1996. hedgehog, wingless and orthodenticle specify adult head
development in Drosophila. Development
122:1849–58
Royet J, Finkelstein R. 1997. Establishing primordia in the Drosophila eye-antennal imaginal disc: the roles of decapentaplegic, wingless, and hedgehog. Development 124:4793–
800
Sampath K, Rubinstein AL, Cheng AM, Liang
JO, Fekany K, et al. 1998. Induction of the
zebrafish ventral brain and floorplate requires
cyclops/nodal signalling. Nature 395:185–
89
Sanyanusin P, McNoe LA, Sullivan MJ,
Weaver RG, Eccles MR. 1995a. Mutation of
PAX2 in two siblings with renal-coloboma
syndrome. Hum. Mol. Genet. 4:2183–84
Sanyanusin P, Schimmenti LA, McNoe LA,
Ward TA, Pierpont ME, et al. 1995b. Mutation of the PAX2 gene in a family with optic
nerve colobomas, renal anomalies and vesicoureteral reflux. Nat. Genet. 9:358–63
Schauerte HE, van Eeden FJ, Fricke C, Odenthal J, Strahle U, Haffter P. 1998. Sonic
hedgehog is not required for the induction
of medial floor plate cells in the zebrafish.
Development 125:2983–93
Schedl A, Ross A, Lee M, Engelkamp D,
Rashbass P, et al. 1996. Influence of PAX6
gene dosage on development: overexpression
causes severe eye abnormalities. Cell 86:71–
82
Schimmenti LA, Cunliffe HE, McNoe LA,
Ward TA, French MC, et al. 1997. Further
delineation of renal-coloboma syndrome in
patients with extreme variability of phenotype and identical PAX2 mutations. Am. J.
Hum. Genet. 60:869–78
Schimmenti LA, Pierpont ME, Carpenter BL,
Kashtan CE, Johnson MR, Dobyns WB.
1995. Autosomal dominant optic nerve
colobomas, vesicoureteral reflux, and renal
anomalies. Am. J. Med. Genet. 59:204–8
Schmahl W, Knoedlseder M, Favor J, Davidson D. 1993. Defects of neuronal migration
and the pathogenesis of cortical malformations are associated with Small eye (Sey) in
the mouse, a point mutation at the Pax-6locus. Acta Neuropathol. 86:126–35
Schulte D, Cepko CL. 2000. Two homeobox
genes define the domain of EphA3 expression
13 Sep 2001
13:49
AR
ar139-09.tex
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
VERTEBRATE EYE DEVELOPMENT
in the developing chick retina. Development
127:5033–45
Schulte D, Furukawa T, Peters MA, Kozak CA,
Cepko CL. 1999. Misexpression of the Emxrelated homeobox genes cVax and mVax2
ventralizes the retina and perturbs the retinotectal map. Neuron 24:541–53
Schwarz M, Cecconi F, Bernier G, Andrejewski N, Kammandel B, et al. 2000. Spatial
specification of mammalian eye territories by
reciprocal transcriptional repression of pax2
and pax6. Development 127:4325–34
Seimiya M, Gehring WJ. 2000. The Drosophila
homeobox gene optix is capable of inducing ectopic eyes by an eyeless-independent
mechanism. Development 127:1879–86
Servetnick M, Grainger RM. 1991. Changes in
neural and lens competence in Xenopus ectoderm: evidence for an autonomous developmental timer. Development 112:177–88
Shen W, Mardon G. 1997. Ectopic eye development in Drosophila induced by directed dachshund expression. Development
124:45–52
Simeone A, Acampora D, Mallamaci A, Stornaiuolo A, D’Apice MR, et al. 1993. A vertebrate gene related to orthodenticle contains a
homeodomain of the bicoid class and demarcates anterior neuroectoderm in the gastrulating mouse embryo. EMBO J. 12:2735–47
Solnica-Krezel L, Stemple DL, MountcastleShah E, Rangini Z, Neuhauss SC, et al.
1996. Mutations affecting cell fates and cellular rearrangements during gastrulation in
zebrafish. Development 123:67–80
Spemann H. 1901. Uber Correlationen in der
Entwickelung des Auges. Ver. Anat. Ges.
15:61–79
St-Onge L, Sosa-Pineda B, Chowdhury K,
Mansouri A, Gruss P. 1997. Pax6 is required
for differentiation of glucagon-producing
alpha-cells in mouse pancreas. Nature 387:
406–9
Stadler HS, Solursh M. 1994. Characterization of the homeobox-containing gene GH6
identifies novel regions of homeobox gene
expression in the developing chick embryo.
Dev. Biol. 161:251–62
295
Stern CD, Yu RT, Kakizuka A, Kintner CR,
Matthews LS, et al. 1995. Activin and its
receptors during gastrulation and the later
phases of mesoderm development in the
chick embryo. Dev. Biol. 172:192–205
Stoykova A, Fritsch R, Walther C, Gruss P.
1996. Forebrain patterning defects in Small
eye mutant mice. Development 122:3453–65
Suzuki R, Shintani T, Sakuta H, Kato A,
Ohkawara T, et al. 2000. Identification of
RALDH-3, a novel retinaldehyde dehydrogenase, expressed in the ventral region of the
retina. Mech. Dev. 98:37–50
Tomarev SI, Callaerts P, Kos L, Zinovieva
R, Halder G, et al. 1997. Squid Pax-6 and
eye development. Proc. Natl. Acad. Sci. USA
94:2421–26
Ton CC, Hirvonen H, Miwa H, Weil MM,
Monaghan P, et al. 1991. Positional cloning
and characterization of a paired box- and
homeobox- containing gene from the aniridia
region. Cell 67:1059–74
Torres M, Gomez-Pardo E, Gruss P. 1996. Pax2
contributes to inner ear patterning and optic nerve trajectory. Development 122:3381–
91
Toy J, Yang JM, Leppert GS, Sundin OH.
1998. The optx2 homeobox gene is expressed
in early precursors of the eye and activates
retina-specific genes. Proc. Natl. Acad. Sci.
USA 95:10643–48
Ungar AR, Moon RT. 1996. Inhibition of protein kinase A phenocopies ectopic expression
of hedgehog in the CNS of wild-type and cyclops mutant embryos. Dev. Biol. 178:186–
91
Vogel-Hopker A, Momose T, Rohrer H, Yasuda K, Ishihara L, Rapaport DH. 2000.
Multiple functions of fibroblast growth
factor-8 (FGF-8) in chick eye development.
Mech. Dev. 94:25–36
Wallis DW, Roessler E, Hehr U, Nanni L, Wiltshire T, et al. 1999. Mutations in the homeodomain of the human SIX3 gene cause holoprosencephaly. Nat. Genet. 22:196–98
Walther C, Gruss P. 1991. Pax-6, a murine
paired box gene, is expressed in the developing CNS. Development 113:1435–49
13 Sep 2001
13:49
296
AR
CHOW
ar139-09.tex
¥
ar139-09.sgm
ARv2(2001/05/10)
P1: GJB
LANG
Wawersik S, Maas RL. 2000. Vertebrate eye development as modeled in Drosophila. Hum.
Mol. Genet. 9:917–25
Wawersik S, Purcell P, Rauchman M, Dudley
AT, Robertson EJ, Maas R. 1999. BMP7 acts
in murine lens placode development. Dev.
Biol. 207:176–88
Wayne S, Robertson NG, DeClau F, Chen N,
Verhoeven K, et al. 2001. Mutation in the
transcriptional activator EYA4 causes lateonset deafness at the DFNA10 locus. Hum.
Mol. Genet. 10:195–200
Weinstein DC, Hemmati-Brivanlou A. 1999.
Neural induction. Annu. Rev. Cell Dev. Biol.
15:411–33
Wigle JT, Chowdhury K, Gruss P, Oliver G.
1999. Prox1 function is crucial for mouse
lens-fibre elongation. Nat. Genet. 21:318–
22
Williams SC, Altmann CR, Chow RL, Hemmati-Brivanlou A, Lang RA. 1998. A highly
conserved lens transcriptional control element from the Pax-6 gene. Mech. Dev.
73:225–29
Winkler S, Loosli F, Henrich T, Wakamatsu Y, Wittbrodt J. 2000. The conditional
medaka mutation eyeless uncouples patterning and morphogenesis of the eye. Development 127:1911–9
Wistow GJ, Piatigorsky J. 1988. Lens crystallins: the evolution and expression of proteins for a highly specialized tissue. Annu.
Rev. Biochem. 57:479–504
Woerdeman MW. 1929. Experimentelle Untersuchungen uber Lage and Bau der Augenbildenden Bezirke in der Medullarplatte
beim Axolotl. Arch. Entw. Mech. Vol. 116
Woo K, Fraser SE. 1995. Order and coherence
in the fate map of the zebrafish nervous system. Development 121:2595–609
Wride MA. 1996. Cellular and molecular features of lens differentiation: a review of recent advances. Differentiation 61:77–93
Xu PX, Adams J, Peters H, Brown MC, Heaney
S, Maas R. 1999a. Eya1-deficient mice lack
ears and kidneys and show abnormal apoptosis of organ primordia. Nat. Genet. 23:113–
17
Xu PX, Zhang X, Heaney S, Yoon A, Michelson AM, Maas RL. 1999b. Regulation of
Pax6 expression is conserved between mice
and flies. Development 126:383–95
Xu Y, Baldassare M, Fisher P, Rathbun G, Oltz
EM, et al. 1993. LH-2: a LIM/homeodomain
gene expressed in developing lymphocytes
and neural cells. Proc. Natl. Acad. Sci. USA
90:227–31
Yamamoto Y, Jeffery WR. 2000. Central role
for the lens in cave fish eye degeneration.
Science 289:631–33
Zhang J, Talbot WS, Schier AF. 1998. Positional cloning identifies zebrafish one-eyed
pinhead as a permissive EGF-related ligand
required during gastrulation. Cell 92:241–51
Zhang L, Mathers PH, Jamrich M. 2000. Function of rx, but not pax6, is essential for the
formation of retinal progenitor cells in mice.
Genesis 28:135–42
Zhang Y, Emmons SW. 1995. Specification
of sense-organ identity by a Caenorhabditis
elegans Pax-6 homologue. Nature 377:55–
59
Zuber ME, Perron M, Philpott A, Bang A, Harris WA. 1999. Giant eyes in Xenopus laevis
by overexpression of XOptx2. Cell 98:341–
52
25 Sep 2001
15:45
AR
AR139-09-COLOR.tex
AR139-09-COLOR.SGM
ARv2(2001/05/10)
P1: GDL
Figure 1 Schematic overview of vertebrate eye development. In panels A–D, presumptive
or differentiated eye tissues are color-coded in the following manner: blue, lens/cornea;
green, neural retina; yellow, retinal pigmented epithelium (RPE); purple, optic stalk (for
explanation why only the ventral region of optic stalk is colored, see the text section on
defining axes in the developing eye); red, ventral forebrain/prechordal mesenchyme; grey,
mesenchyme. (A) Formation of the optic vesicle is initiated by an evagination (indicated
by arrow) of the presumptive forebrain region resulting in the formation of the optic pit
(OP). The optic vesicle region is divided into dorso-distal region (green), which contains the
presumpitve neural retina (PNR) and RPE (not shown), and the proximo-ventral region, which
gives rise to the presumptive ventral optic stalk (POS); PLE, presumptive lens ectoderm; M,
mesenchyme; VF, ventral forebrain; PCM, prechordal mesoderm. (B) Continued growth of
the optic vesicle culminates with a period of close contact between the lens placode (LP) and
the presumptive neural retina (NR) during which important inductive signal likely exchange:
RPE, presumptive retinal pigmented epithelium; VOS, ventral optic stalk; DOS, dorsal optic
stalk. (C) Invagination of the optic vesicle results in formation of the lens vesicle (LV) and
neural retina (NR) and establishes the overall structure of the eye. The point at which the
neural retina and RPE meet gives rise to components of the ciliary body and iris (C/I). (D)
Mature eye: C, cornea; LE, lens epithelium; LF, lens fiber cells; I, iris; CB, ciliary body; GCL,
ganglion cell layer; INL, inner nuclear layer; ONL, outer nuclear layer; ON, optic nerve.
17:36
AR
AR139-09-COLOR.tex
Figure 2 Defining the proximo-ventral/dorso-distal axis in the developing eye. This figure represents a ventral view
of four different stages of eye morphogenesis. (A) The early eye field can be conceptualized as a flat sheet in which the
proximo-ventral (PV) region adjacent to the midline will give rise to the ventral optic stalk, and the dorso-distal (DD) domain
will give rise to the RPE (see text). (B) Optic evagination (arrow) is the first morphological event that alters the shape of
presumptive eye tissues. The PV-DD axis within the eye field is beginning to curve at this stage. (C) Optic invagination occurs
along a continuous line extending from the presumptive neural retina (green and top arrow) and along the ventral optic stalk
(bottom arrow). The PV-DD axis is reflexed as the presumptive RPE folds back over the presumptive neural retina. (D) Closure
of the ventral choroidal fissure (asterisks) occurs as the laterally extending edges of the neural retina and RPE (top set of
arrows) sweep in a ventral movement and eventually fuse. The laterally extending edges of the dorsal optic stalk (bottom set
of arrows) move ventrally and fuse at the midline to become the outer wall of the optic stalk, while the former ventral optic
stalk pinches off at the point of closure and becomes the inner wall. Although this representation of the developing eye field
could accommodate anterior (nasal) and posterior (temporal) influences present on either side of the PV-DD axis, it has been
omitted because little is known about these influences at early stages of eye development.
26 Sep 2001
AR139-09-COLOR.SGM
ARv2(2001/05/10)
P1: GDL
10:51
AR
AR139-09-COLOR.tex
Figure 3 Patterning and inductive events in the developing eye. The developing eye field is represented as a simple sheet defined by
proximo-ventral/dorso-distal boundaries. Anterior (nasal) and posterior (temporal) boundaries have been excluded from this simplified
representation. (A) As a downstream response to neural induction and possible Wnt signaling (green triangle), many transcription factors
important for eye development are expressed in the anterior neural plate in a single eye field. (B) Cyc signaling from underlying prechordal
mesoderm (red triangle) is believed to induce ventral midline specification and up-regulation of hh expression. Hedgehog signals from the
midline and possibly other sources (purple triangle) specify the ventral optic vesicle. Pax2 and Pax6 reciprocal repression at the PV-DD
boundary is believed to occur gradually (B and C) until a sharp boundary (D) forms between neural retina and optic stalk; m, mouse. (C)
Signals from the surface ectoderm (possibly BMPs or FGFs; green, lower triangle) are important for neural retina specification, whereas
signals from extraocular mesenchyme (possibly BMPs or activins; yellow, upper triangle) are believed to function in RPE specification;
ch, chick. (D) Distinct PV-DD patterning occurs within the neural retina possibly under the influence of retinoic acid (RA; green triangle).
Note that while Pax2 is normally present in the ventral neural retina, it is down-regulated in later development as depicted in (D).
26 Sep 2001
AR139-09-COLOR.SGM
ARv2(2001/05/10)
P1: GDL
26 Sep 2001
10:51
AR
AR139-09-COLOR.tex
AR139-09-COLOR.SGM
ARv2(2001/05/10)
P1: GDL
Figure 5 Generating two eyes. (A) Removal of the prechordal mesoderm (Urd.) from
Triton embryos at the neural plate stage resulted in the cyclopic embryo shown in (B)
(Mangold 1931). Recent experiments using Xenopus anterior neural plate explants
demonstrated that in contrast to the control in (C) (arrowhead ), removal of the prechordal mesoderm in explants leads to the persistence of Pax6 expression in the midline
(D) (arrowhead ) (Li et al. 1997). The one-eyed pinhead phenotype in zebrafish that
leads to cyclopia (E ) (arrowhead ) can be rescued by activating the cyc/nodal signaling
pathway in oep mutants misexpressing Smad2 (F ) (arrowheads); arrow in (F ) indicates restoration of the ventral forebrain (Gritsman et al. 1999). (G) Optic vesicle (ov)
restricted Pax6 expression in the zebrafish embryo. (H ) Sonic hedgehog misexpression
in zebrafish leads to the loss of Pax6 expression in the optic vesicle (ov) and (J ) the
expansion of Pax2 expression. Pax2 expression is (I ) restricted to the optic stalks (os)
in uninjected controls (Ekker et al. 1995). The asterisks in (I ) and (J ) indicate Pax2
expression at the midbrain-hindbrain boundary.