Standard PDF - Wiley Online Library

advertisement
THE ANATOMICAL RECORD 297:137–160 (2014)
Retinal Stem Cells and Regeneration of
Vision System
1
HENRY K. YIP1,2,3*
Department of Anatomy, Li Ka Shing Faculty of Medicine, The University of Hong Kong,
Pokfulam, Hong Kong Special Adminstrative Region, People’s Republic of China
2
Research Center of Heart, Brain, Hormone and Healthy Aging, Li Ka Shing Faculty of
Medicine, The University of Hong Kong, Pokfulam, Hong Kong Special Adminstrative
Region, People’s Republic of China
3
State Key Laboratory of Brain and Cognitive Sciences, The University of Hong Kong,
Pokfulam, Hong Kong Special Adminstrative Region, People’s Republic of China
ABSTRACT
The vertebrate retina is a well-characterized model for studying neurogenesis. Retinal neurons and glia are generated in a conserved order from
a pool of mutlipotent progenitor cells. During retinal development, retinal
stem/progenitor cells (RPC) change their competency over time under the
influence of intrinsic (such as transcriptional factors) and extrinsic factors
(such as growth factors). In this review, we summarize the roles of these
factors, together with the understanding of the signaling pathways that
regulate eye development. The information about the interactions between
intrinsic and extrinsic factors for retinal cell fate specification is useful to
regenerate specific retinal neurons from RPCs. Recent studies have identified RPCs in the retina, which may have important implications in health
and disease. Despite the recent advances in stem cell biology, our understanding of many aspects of RPCs in the eye remains limited. PRCs are
present in the developing eye of all vertebrates and remain active in lower
vertebrates throughout life. In mammals, however, PRCs are quiescent
and exhibit very little activity and thus have low capacity for retinal regeneration. A number of different cellular sources of RPCs have been identified in the vertebrate retina. These include PRCs at the retinal margin,
pigmented cells in the ciliary body, iris, and retinal pigment epithelium,
and M€
uller cells within the retina. Because PRCs can be isolated and
expanded from immature and mature eyes, it is possible now to study these
cells in culture and after transplantation in the degenerated retinal tissue.
We also examine current knowledge of intrinsic RPCs, and human embryonic stems and induced pluripotent stem cells as potential sources for cell
transplant therapy to regenerate the diseased retina. Anat Rec, 297:137–
C 2013 Wiley Periodicals, Inc.
160, 2014. V
Key words: tissue engineering; stem cell; regeneration
Grant sponsor: The University of Hong Kong Seed Funding
Program for Basic Research; Grant numbers: 200611159203 and
201011159065; Grant sponsor: The University of Hong Kong
Small Project Funding; Grant number: 200807170103; Grant
sponsor: General Research Fund, Regional Grant Council of
Hong Kong; Grant number: 10208603.
*Correspondence to: Henry K. Yip, Department of Anatomy,
Li Ka Shing Faculty of Medicine, The University of Hong Kong,
C 2013 WILEY PERIODICALS, INC.
V
21 Sassoon Road, Pokfulam, Hong Kong SAR, China. Fax:
1852-28170857. E-mail: hkfyip@hku.hk
Received 13 September 2013; Accepted 13 September 2013.
DOI 10.1002/ar.22800
Published online 2 December 2013 in Wiley Online Library
(wileyonlinelibrary.com).
138
YIP
During development, the nervous system arises from
the neuroepithelium of the neural plate located along
the dorsal midline of the embryo and then folds into the
neural tube before undergoing various patterning events
and specification. Most cells in the emerging nervous
system during early embryonic development are multipotent and they can give rise to both neurons and glia.
The population and the neurogenic potential of neural
stem/progenitor cells (NPCs) decrease progressively with
age in higher vertebrates, for example, avian and mammal, so that NPCs become restricted to two neurogenic
areas in the adult mammalian brain: the subgranular
zone (SGZ) of the dentate gyrus in the hippocampus,
and subventricular zone (SVZ) lining the lateral ventricles (Miller and Gauthier-Fisher, 2009; Weinandy
et al., 2011) as well as in non-neurogenic regions such as
cerebral cortex (Gould et al., 1999; Magavi et al., 2000),
spinal cord (Xiao et al., 2010; Sabelstrom et al., 2013)
and in various structures of the eye (Ffrench-Constant
and Raff, 1986; Ahmad et al., 2000; Tropepe et al., 2000;
Haruta et al., 2001). NSCs in the SGZ generate granule
neurons and NSCs in the SVZ give rise to neurons and
glia in the developing telencephalon and to ensure a lifelong contribution of neural progenitors that migrate long
distance along the rostral migratory stream to reach
their final destination in the olfactory bulb, a major area
of adult neurogenesis (Costa et al., 2010). The SVZ in
the adult brain has the highest neurogenic rate, from
which NSCs are first isolated, and characterized (Reynolds and Weiss, 1992). These cells resemble a radial glia
during neurogenesis, after that they acquire an astroglial stem cell or ependymal identity (Mori et al., 2005).
This glial identity of NSCs is important since it has
been suggested that adult glial cells such as NG2 glia or
even astrocytes, may acquire stem cell properties of self
renew and multipotency following brain injury to participate in local tissue repair (Robel et al., 2011; Bonaguidi
et al., 2012). NPCs in the non-neurogenic regions remain
dormant and quiescent, but can be activated by exogenous factors, for example, after injury (Palmer et al.,
1995, 1999; Weiss et al., 1996; Magavi et al., 2000; Kernie et al., 2001; Ming and Song, 2005).
The regenerative potential of the vertebrate retinal
has been observed in a variety of species during either
their development or for some even during their adult
life. In mammals, the neuroretina and the retinal pigmented epithelium (RPE) show no evidence of regeneration in the adult as observed in fish and amphibian.
Although the eye continues to grow for some time after
birth in mammals and birds, this is largely due to the
stretching of the retina associated with the growth of
the sclera rather than addition of new cells to the retina
(Perron and Harris, 2000). Fish, amphibian, and birds
are among those with the ability to regenerate during
adulthood (Perron et al., 1998; Reh and Levine, 1998;
Fischer and Reh, 2000; Perron and Harris, 2000; Amato
et al., 2004; Klassen et al., 2004a,b; Moshiri et al. 2004;
Fischer and Omar, 2005). It has been well documented
that in the cold-blooded vertebrates, like fish and
amphibian, the retina continues to grow throughout
their lifetime by addition of new neurons at the peripheral rim of the retina at the ciliary margin (Reh and
Levine, 1998; Otteson and Hitchcock, 2003). In this
review, we provide an overview of retinal progenitor cells
(RPCs) identified in different regions of the vertebrate
eye, in particular, in the epithelia of the retina, ciliary
body, and iris, and the retinal radial glial cells. Before
we could use them for the possible treatment of retinal
diseases, it is important to understand their basic characteristic features. In addition, we discuss the intrinsic
properties of the RPCs and the key extrinsic signaling
molecules that regulate RPC behavior in proliferation
and differentiation. An integrating knowledge of the
molecular and genetic processes underlying the development of the retina is essential for understanding not
only normal developmental mechanisms, but also in
future therapeutic strategies aiming at restoring vision
loss. Recent advances in the field of stem cell research
have raised the feasibility of using stem cell-based therapies as a potential avenue for treatment of retinal diseases. Finally, this review will also focus on this
research in identifying suitable sources of pluripotent
stem/progenitor cells for cell replacement or transplantation therapies, including both human embryonic stem
cells (hESCs) and human induced pluripotent stem cells
(hiPSCs).
RPC IN THE DEVELOPING BRAIN
Intrinsic Properties of the RPCs in Retinal Cell
Specification
The retina of the vertebrate arises from an evagination of the diencepalon that constitutes the anterior neural tube during the early neurulation stage of the
embryo development. The continued evagination of the
optic primordia results in the formation of the optic
vesicles. The overlying surface ectoderm and the optic
vesicles undergo a coordinated invagination resulting in
the formation of the lens vesicle and the bilayered optic
cup. The inner layer of the optic cup, closest to the lens
vesicle, eventually forms the multilayered neural retina,
while the outer layer remains as a single epithelial layer
and gives arises to the retinal pigment epithelium
(RPE). The periphery of the optic cup, where the inner
layer and outer layer meet, become the ciliary epithelium (CE) and the iris (Beebe, 1986). The pigmented
part of the CE and the iris arises from the outer layer of
the optic cup, and it is continuous with the pigment epithelium of the retina. The nonpigment part of the CE
represents the extension of the neural layer of the retina, in which muscle and connective tissue developed.
Before the formation of eye field, expression of the
orthodenticle homolog transcription factor Otx2 is found
at the anterior end of the neural tube and later in the
optic vesicle and optic cup. Otx2 is essential to “initiate”
the anterior neural plate cells to form the embryonic eye
fields (Chuang and Raymond, 2002). The single eye field
across the midline region is separated into two distinct
lateral eye primodia, by a number of soluble factors such
as sonic hedgehog, Shh, (MacDonald et al., 1995; Li
et al., 1997) and bone morphogenetic proteins (BMPs) 4
and 5 (Golden et al., 1999) released by the underlying
prechordal mesoderm. The cells of the developing retinal
epithelium are analogous to the neural progenitors from
other neurogenic regions of the CNS and are mutlipotent
retinal stem/progenitors cells (RPCs). These cells express
a unique set of transcription factors called eye field transcription factors (EFTFS), which set them molecularly
apart from the surrounding cells at the anterior edge of
RSC AND REGENERATION OF VISION SYSTEM
the neural plate. The two earliest EFTFS expressed in
the eye field are both paired-like homeodomain genes,
Rax/Rx and Pax-6 (Hill and Hanson, 1992; Furukawa
et al., 1997a; Mathers et al., 1997) (Fig. 1). Homozygous
mutation of both Rax/Rx and Pax-6 result in anophthamia (no eye formation) observed across species in Xenopus, mice and rats (Matsuo et al., 1993; Grindley et al.,
1995; Mathers et al., 1997; Andreazzoli et al., 2003). The
Rx-deficient mice also have a reduction in the expression
of other EFTFS include Pax-6 and Six3, suggesting that
Rx may have an induction role on these genes. However,
Pax-6 knockout mice have normal Rx expression, indicating that Pax-6 is downstream of Rx (Zhang et al.,
2000). Overexpression of Rx and Pax-6 in Xenopus
results in the formation of ectopic retinal tissue (Mathers et al., 1997; Zuber and Harris, 2006). Pax-6 overexpression also induces ectopic expression of Rx,
suggesting that Pax-6 also has an induction role on Rx.
Thus, these loss- and gain-of-function studies support
the idea that each of these transcription factors regulates the activity of each other or a number of other
EFTFS genes such as ET, Six3, Lhx2, T11, and Optx2
(Six6) (Zuber et al., 2003). It would appear that all the
EFTFS work together at the beginning of the developmental network of transcription factors to establishing
the initial commitment of anterior neural plate cells to a
retinal fate (Fig. 1).
Cell lineage analysis of the progeny of these actively
dividing RPCs shows that all the different types of retinal cells, including all six types of retinal neurons and
together with the radial M€
uller glial cells, arise from a
common pool of multipotent RPCs (Turner and Cepko,
1987; Holt et al., 1988; Wetts and Fraser, 1988; Turner
et al., 1990; Fekete et al., 1994; Cepko et al., 1996).
Birthdating experiments have demonstrated that retinal
cell types are generated by the RPCs in a relatively fixed
chronological order that is evolutionary conserved,
although the time scale of retinogenesis could vary
greatly between species (review see Altshuler et al.,
1991). These studies have determined that RGCs, cone
photoreceptors, amacrine cells, and horizontal cells are
generated during the first wave of neurogenesis in the
retina, followed, in a second wave, by rod photoreceptors, bipolar cells, and M€
uller glial cells, with considerable overlap in the appearance among these different
cell types (Holt et al., 1988; Wetts and Fraser, 1988;
Prada et al., 1991; Hu and Easter, 1999; Galli-Resta,
2002; Marquardt and Gruss, 2002; Das et al., 2003) (Fig.
1). Once various cell types are generated, they migrate
into the developing retina forming three cellular layers:
(i) the outer nuclear layer (ONL), which contains photoreceptors (rod and cone); (ii) the inner nuclear layer
(INL), which contains interneurons (horizontal, bipolar,
and amacrine cells) and M€
uller glial cells; and the inner
most layer (iii) the ganglion cell layer (GCL), which contains projecting neurons retinal ganglion cells (RGCs).
Retinogenesis in vertebrates begins in central retina and
spread toward the periphery, such that cells in the
periphery retina are the last to born and differentiate,
while those in central retina are older (Perron and Harris, 2000; Amato et al., 2004).
The vertebrate retina consists of seven major classes
of cells, and several of which can be further divided into
multiple distinct subtypes (Harris, 1997; Masland,
2001). It has been shown that retinal cell fate specifica-
139
tion and differentiation are controlled by temporally
varying intrinsic properties of RPCs and extrinsic signals from the environment (Cepko 1999). The basic
helix-loop-helix (bHLH) transcription factors are likely
candidates as intrinsic factors in regulating retinal cell
fate (Cepko, 1999) (Fig. 1). In Drosophila, proneural
genes of atonal (ato) and achaete-scute complex (AS-C)
are required for the selection of sense organ precursors
(Jan and Jan, 1993). Atonal is also the proneural gene
for photoreceptor neurons in the developing Drosophila
retina (Jarman et al., 1995; White and Jarman, 2000).
Many neurogenic bHLH genes such as the AS-C homolog Mash1 and ato homologs Math3, Ngn2, Math5, and
NeuroD are expressed by the RPCs in the developing
vertebrate retina and bias RPCs toward retinal cell fates
(Cepko et al., 1996; Brown et al., 1998). Loss- and gainof-function studies indicate that retinal cell fate determination is regulated by multiple bHLH transcription factors. Hes1 and Hes5 are chief regulators of proliferation
process during early retinal development to ensure a
continuous supply of RPCs for the successive production
of differentiated retinal cells during retinogenesis. In
Hes1 mutant murine embryos, cell proliferation in the
retina is severely impaired. In addition, precocious neurogenesis and disruption of laminar structures are
observed (Ishibashi et al., 1994; Takatsuka et al., 2004;
Lee et al., 2005). Double Hes1/Hes5 mutant animals
have more eye formation abnormalities, including
absence of optic vesicles (Hatakeyama et al., 2001), suggesting that Hes1 and Hes5 work cooperatively to maintain RPCs in an undifferentiated state. Furthermore,
Hes1 and Hes5 are downstream targets of Notch signaling (Ohtsuka et al., 1999). Notch signaling plays a critical role in maintaining stem cells in an undifferentiated
state (Gaiano et al., 2000) without affecting the competency of the RPCs over the course of development (Jadhav et al., 2006). In Xenopus, targeted expression of
Xath5, a Xenopus ortholog of Math5, promotes ganglion
cell differentiation (Kanekar et al., 1997). Math5, a
murine ortholog of the atonal, is involved in ganglion
cell specification (Brown et al., 2001; Wang et al., 2001)
through the activation of downstream Brn3b, a POU
domain transcription factor (Liu et al., 2001). In Math5
null mice, there is a significant decrease of ganglion cells
accompanied by an increase in amacrine cells and cone
photoreceptors (Brown et al., 2001), indicating that
Math5, in addition to direct cell fate specifications, also
involved in controlling the number of retinal cells and
may regulate successive stages of retinal cell differentiation (Marquardt and Gruss, 2002). In contrast, Math3
and NeuroD double-mutant retina displays selective loss
of amacrine cells, while ganglion cells are significantly
increased in number. In addition, Math5 expression is
upregulated in the absence of Math3 and NeuroD, suggesting that Math5 and Math3/NeuroD are acting antagonistically in regulating ganglion and amacrine cell fate
specification (Inoue et al., 2002). Mash1 and Math3
(Tomita et al., 2000), and the homeobox gene Chx10
(Hatakeyama et al., 2001) are required for bipolar cell
specification. In either Mash1- or Math3-mutant retinas,
no apparent defect in bipolar cell fate specification is
observed (Tomita et al., 1996). However, in Mash1 and
Math3 double-mutant, bipolar cells disappear and the
RPCs adopted the M€
uller glial cell fate (Tomita et al.,
2000). Taken together, these findings indicate that
140
YIP
retinal cell fate depends on the expression of multiple
bHLH genes and the downregulation of these genes
could be one of the mechanisms to initiate glial cell
fate determination. In addition, these studies also
revealed the intricate cross-regulation among bHLH
genes. Furthermore, a combination of the bHLH and
the homeobox genes are required for the specification
of bipolar cells. It is suggested that bHLH genes determine the neuronal cell fate, while the homebox gene
confer the positional identity to the RPCs in the INL
(Hatakeyama et al., 2001). The homeobox gene Crx and
Otx2 direct RPCs toward a photoreceptor cell fate.
Otx2 activates the replication of Crx gene and deletion
of Otx2 results in the conversion of photoreceptor cells
to amacrine-like cells (Chen et al., 1997; Furukawa
et al., 1997b; Nishida et al., 2003). Similarly, Crx-null
mutant retinas exhibit defects in the generation of photoreceptors (Furukawa et al., 1999). Recent studies
have shown that Nrl, a basic leucine zipper transcription factor, is required for the regulation of cell fate
determination between rod and cone photoreceptors
through the activation of an orphan nuclear receptor
Nr2e3 (Cheng et al., 2006). Deletion of Nrl resulted in
the loss of rod photoreceptors and Nr2e3 expression in
the mutant retinas (Mears et al., 2001). Nr2e3 was
shown to activate rod-specific genes, but works in concert with CrX, represses cone-specific genes (Cheng
et al., 2006; Chen et al., 2005; Peng et al., 2005). Horizontal cell specification is regulated by the core group
of Foxn4, Ptf1a and Prox1. In the absence of any of
these three genes, horizontal cell genesis is impaired
(Dyer et al., 2003; Li et al., 2004; Fujitani et al., 2006).
Hes1 and Hes5, which is a negative regulators of neurogenic bHLH genes such as Mash1 and Math3 (Akazawa et al., 1992; Ohtsuka et al., 1999), induce M€
uller
glial cell fate (Furukawa et al., 2000; Hojo et al., 2000).
In Hes5-null mice, there is a significant decrease in the
number of M€
uller glial cells due to insufficient expansion of RPCs that could differentiate into M€
uller glial
cells (Hojo et al., 2000; Deneen et al., 2006). Hes1 or
Hes5 misexpression induces M€
uller glial cell generation
in the postnatal retina. Notch (Dorsky et al., 1995;
Ohnuma et al., 1999; Furukawa et al., 2000) and the
homeobox gene rax (Furukawa et al., 2000) are also
reported to promote the formation of M€
uller glial. Rax
has been shown to promote proliferation of RPCs (Furukawa et al., 1997a,b; Mathers et al., 1997). Thus,
Hes1 and Hes5 as well as Notch and Rax have dual
functions in maintaining RPCs and promoting gliogenesis during retinal development. Interestingly, M€
uller
glial cells share many morphological characteristics of
neural/glial progenitors in many other regions of the
CNS. They have a simple bipolar morphology, with
extending processes contacting both the ventricular and
vitreal surfaces of the neuroepithelium. Furthermore,
these cells are competent to generate neurons and glia
(Fischer and Reh, 2001a,b; Dyer et al., 2003). It is conceivable that RPCs retain Hes1 and Hes5 expression in
the late stage of retinogenesis adopt the M€
uller glial
cell fate. While our understanding of the molecular
mechanisms of retinal cell fate specification has
advance dramatically in recent year, we probably just
have the crudest outline of the whole process describing
how the orchestrated expression of these intrinsic factors act to generate a functional retina. Nevertheless,
such knowledge is fundamental to an understanding
the role of RPCs in regenerating the retina.
Extrinsic Signaling Molecules in the Regulation
of RPC Proliferation and Differentiation
One of the remaining challenges is to elucidate the
mechanisms by which extrinsic environmental cues
impact on the intrinsic cell determinants in RPC fate
determination. Two important common events emerge
from the studies of RPC specification during retinal
development: (1) seven major types of retinal cells are
generated is a sequential and yet overlapping order that
is conserved among many species (Altshuler et al.,
1991); (2) vertebrate RPCs are multipotent at different
developmental stages and the progeny derived from individual RPC can assume a variety of cell fate (Turner
and Cepko, 1987; Holt et al., 1988; Wetts and Fraser,
1988; Turner et al., 1990; Fekete et al., 1994). The multipotency of the RPCs throughout retinogenesis indicates
that the local environment plays critical roles in cell fate
determinations. There are abundant evidence to support
that RPCs are not homogenous and do not stay static
during development (Lillien, 1998). Heterotopic transplantation studies have demonstrated that early and
late RPCs have different differentiation capacities when
placed in similar environment (Watanabe and Raff,
1990; Morrow et al., 1998; Belliveau and Cepko, 1999;
Belliveau et al., 2000) and they have different gene
expression profiles (Jasoni and Reh, 1996; Yang and
Cepko, 1996; Alexiades and Cepko, 1997; Perron et al.,
1998; Matter-Sadzinski et al., 2001). Thus, the intrinsic
determinants that define the properties of RPCs, including surface receptors, intracellular signaling pathways
and nuclear transcription factors, undergo adaptive
changes with the progression of developmental events in
the retina. It was proposed the RPCs advance through a
series of “competent state” during development, and
each “competent state” support specification of one or
more cell fates. The “competent state” is defined as the
intrinsic cell properties that determine the responsiveness of the RPCs to extrinsic cues and the potential of
the RPCs. Hence, cell fate specification is determined by
both the intrinsic properties of RPCs and the extrinsic
cues from the environment from the developing retina
(Cepko et al., 1996; Livesey and Cepko, 2001).
The development of the eye is known to involve many
different interactions and signaling events between the
intrinsic factors and extrinsic environment. Extrinsic
molecules help to specify and ensure a balanced production of the types of retinal cells, establishing the laminar
structures, defining dorsoventral and nasotemporal differences and establishing topographical connections
between the RGCs and the visual centers of the brain
(Fig. 1).
Hedgehog signaling pathways. Multiple signaling pathways have been demonstrated to regulate the
development of vertebrate eye. The family of vertebrate
hedgehog (Hh) proteins is important secreted signaling
molecules that have multiple roles in a variety of developmental processes including pattern formation, tissue
specification, neurogenesis, cell survival, and cell proliferation (Huh et al., 1999; Ingham and McMahon, 2001;
Zhang and Yang, 2001; Perron et al., 2003). Hh
RSC AND REGENERATION OF VISION SYSTEM
molecules are expressed in dynamic pattern by the anterior ventral midline tissues, RPE, and specific types of
postmitotic retinal neurons during vertebrate eye morphogenesis and retinogenesis. Hedgehog mediates signaling through two transmembrane receptors Patched
(Ptc) and Smoothened (Smo), and activated Smo is translocated to the nucleus and affects the transcription of
target genes (Murone et al., 1999). The cellular
responses to Hh signals are determined by the local Hh
concentration gradient and the intrinsic properties of
the cells within the Hh gradient. Studies in fish, Xenopus, chick, and murine have demonstrated that expression of Shh, a member of Hh family, in the ventral
midline of the CNS plays a critical role in vertebrate eye
pattern formation. Persistent Shh signals are required
for the temporal transition from the optic vesicle to optic
cup and after eye cup formation (Huh et al., 1999; Zhang
and Yang, 2001; Perron et al., 2003). Shh expression in
the anterior ventral midline of the neural plate, are
involved in the formation of separate eye fields. Mutations in the Shh gene in the murine results in a single
centrally positioned primitive eye vesicle (Chiang et al.,
1996). Similarly, disruption of Shh gene leads to the formation of cyclopic eye in human (Muenke and Beachy,
2000). Distinct Shh signal thresholds emanating along
the ventral midline control the expression of various
transcription factors and contribute to the pattern formation of the eye primordium along the proximodistal
axis. Shh signals activates the paired domain genes
Pax2 (Nornes et al., 1990) and Vax (Bertuzzi et al., 1999;
Hallonet et al., 1999; Schulte et al., 1999; Mui et al.,
2002; Take-uchi et al., 2003) in portions of the optic primordium proximal to the midline and specify the formation of optic stalk and ventral retinal fates, whereas
Pax6 (Walther and Gruss, 1991) and Rx/Rax (Furukawa
et al., 1997a,b; Mathers et al., 1997) expressed in the
primodium further distal to the midline that forms the
optic cup. In addition, Shh also regulates expression of
bHLH zipper gene MITF (Mochii et al., 1998) and the
homeodomain gene Otx2 (Martinez-Morales et al., 2001)
in the RPE. Furthermore, Shh signals are involved in
the dorsoventral compartmentation of the vertebrate eye
primodium (Chiang et al., 1996; Huh et al., 1999; Zhang
and Yang, 2001). In zebrafish, Shh secreted by the differentiated RGCs is necessary for the propagation of the
neurogenic wave from the center of the retina towards
the
undifferentiated
periphery
(Neumann
and
Nuesslein-Volhard, 2000). Beside initiating the entry of
RPCs from a proliferative state to a differentiated state
in the retina primordium, Shh produced by the differentiated RGCs have also been demonstrated to suppress
further production of RGCs from the early RPC pool in
the chick retina (Zhang and Yang, 2001). Furthermore,
Shh molecule suppresses both the number and the
length of neurites form retinal explants in the chick
(Trousse et al., 2001b). Interestingly, conditional deletion
of Shh in murine RGCs shows a loss of astrocyte precursors in the optic disc and defective RGC axon guidance
(Trousse et al., 2001b), as well as conversion of the RPCs
in the optic stalk to RPE cells (Dakubo et al., 2003).
Conditional ablation of Shh gene in RGCs also results in
lamination defects as indicated by the disorganized photoreceptor cell layer and the malformation of M€
uller glia
(Wang et al., 2001). In addition, members of the Hh family are also produced by RPE; for example, Indian hedge-
141
hog (Ihh) is expressed in both embryonic and mature rat
RPE (Levine et al., 1997), and Banded hedgehog and
Cephalic hedgehog, homologs to the mammalian Ihh and
desert hedgehog (Dhh), respectively, are found in the
embryonic Xenopus RPE (Perron et al., 2003). In rat and
murine retinal cultures, Shh stimulates RPC proliferation and enhances the generation of later born retinal
cell types (Jansen and Wallace, 1997; Levinie et al.,
1997). In zebrafish, RPE-derived Hh may be implicated
in promoting differentiation of photoreceptor cells (Stenkamp et al., 2000).
Transforming growth factor beta (TGF-b) signaling pathways. Transforming growth factor beta
(TGF-b) constitutes a large super family of pleiotropic
growth factors, participates in the development of the
nervous system, including neural induction, dorsoventral
patterning of the neural tube, cell fate determination
and differentiation, and neuronal survival. TGF-b superfamily members transduce signals through the Type I
(BMPR-1) and Type II (BMPR-II) transmembrane serine/threonine kinase receptors. The signals are transduced from the receptors by Smad transcription factors
to the nucleus to regulate gene transcription (Messague
and Kelly, 1986; Moustakas et al., 2001). Emerging evidence has been accumulating recently suggesting a role
of TGF-b family of members in retingogensis and eye
development. BMPs, members of the TGF-b family, are
essential for the development of nervous system. Several
BMP members are expressed during murine eye development (Dudley and Robertson, 1997; Du et al., 2010).
Gene deletion studies have shown that BMPs are critical
for early morphogenesis of the eye (Dudley et al., 1995;
Luo et al., 1995; Furuta and Hogan, 1998; Wawersik
et al., 1999). In the eye, specific BMPs contribute to multiple aspects of early retinal and lens development.
BMP4 specifies domain-specific gene expression and cell
identity in the dorsal retina and induces the surface
ectoderm overlying the optic cup to form lens tissue
(Furuta and Hogan, 1998; Belecky-Adams and Adler,
2001; Sasagawa et al., 2002; Beebe et al., 2004). Absence
of BMP7 in the surface ectoderm overlying the optic
vesicles in the BMP7 null mice frequently displays an
eyeless phenotype and retarded lens development (Dudley et al., 1995; Jena et al., 1997; Wawersik et al., 1999;
Lang, 2004), probably due to the disruption of interaction between the pre-lens ectoderm and the optic vesicle
during eye morphogenesis (Hyer et al., 2003). These
studies showed that target deletion of murine BMP4 and
BMP7 resulted in failure of lens placode formation, suggesting these molecules are essential in lens induction.
There is also evidence that BMPs and TGF-bs contribute
to the differentiation of lens fiber cells and lens fiber
elongation (Belecky-Adams et al., 2002; Faber et al.,
2002), the formation of ciliary body (Zhao et al.,
2002b,c), the differentiation of ventral optic cup structures (Adler and Belecky-Adams, 2002), programmed
cell death and axon guidance (Dunker et al., 2001;
Sakuta et al., 2001; Trousse et al., 2001a; Liu et al.,
2003; Franke et al., 2006), the formation of corneal epithelium and stroma (Sanford et al., 1997; Saika et al.,
2001) and the participation in the development of pigmented epithelium (Fuhrmann et al., 2000; Idelson
et al., 2009; Hongisto et al., 2012). In addition to the
142
YIP
signaling in ocular tissue specification and retinal pattern formation, TGF-b protein Activin promotes differentiation of RPCs into photoreceptor (Jaffe et al., 1994;
Davis et al., 2000) and BMP7 stimulates chick photoreceptor outer segment formation (Sehgal et al., 2006).
Recently, BMPs has been demonstrated to stimulate
RPCs and neuroblastoma cells to differentiate into neuronal linage whereas astrocyte differentiation was inhibited. In addition, it was shown that the effect BMPs is
mediated through the activation of inhibitor of DNAbinding (Id) target genes by the BMP/Smad signaling
pathway (Du et al., 2010; Du and Yip, 2010, 2011).
The Wnt signaling pathway. The canonical Wnt
signaling pathway is highly conserved among various species and is known to regulate multiple eye development
events, such as the formation of eye field, cell proliferation, differentiation, polarity, and movement during different stages of ocular development and growth. Wnt
proteins belong to a large family of secreted glycoproteins,
bind to the Frizzled (Fz) family of transmembrane receptors and co-receptor of the lipoprotein receptor-related
protein (LRP) and activate the canonical b-catenindependent Wnt (Wnt/b-catenin) pathways or the noncanonical b-catenin-independent Wnt signaling pathways.
The noncanonical Wnt signaling pathways include the
Wnt/planar cell polarity (PCP) and the Wnt/Calcium
pathway. In the Wnt/b-catenin pathway, binding of Wnt
to Fzs leads to the activation of intracellular protein
Disheveled (Dsh), which results in the inhibition of GSK3b. This blocks phosphorylation and degradation of bcatenin resulting in its translocation into the nucleus and
the formation of b-catenin and the TCF/LEF complex that
initiates target gene transcription (Huelsken and Behrens, 2002; Wharton, 2003). In the Wnt/PCP pathway,
Dsh activates Rho/Rac small GTPase and JNK in the subsequent regulation of cytoskeletal organization and gene
expression (Tree et al., 2002; Zallen, 2007; Simons and
Mlodzik, 2008). Activation of Wnt/Calcium pathway leads
to a transient increase in the concentration of intracellular molecules, such as inositol 1,4,5-triphosphate (IP3)
and 1,2 diacylglycerol (DAG). IP3 interacts with the calcium channels on the membrane of endoplasmic reticulum (ER) causing it to release calcium ions. The released
calcium together with the cytosolic expressed calmodulin
activates calcium calmodulin-dependent protein kinase II
(CaMKII). The DAG along with the ER-released calcium
activates Protein kinase C (PKC). Both CaMKII and PKC
then activate nuclear factors NF-AT and other transcription factors like NFjB and CREB (Sheldahl et al., 1999;
K€
uhl et al., 2000; Hogan et al., 2003; De, 2011).
The dynamic expression pattern of various components of the Wnt/Fz signaling at different stages of eye
development suggested that Wnt/Fz signaling is involved
in coordinating numerous critical processes during ocular tissue development. A graded Wnt signaling in the
anterior neural plate allows specification of forebrain
subdomains (Yamaguchi, 2001; Nordstrom et al., 2002;
Satoh et al., 2004). Inhibitory molecules, such as EFTFs
(Six3), secreted frizzled related protein 1 (SFRP1) and
transcription factor-3 (TCF-3), expressed in the anterior
neural plate allow the development of the forebrain,
including the eye field, by suppressing the Wnt/b-catenin pathway (Kim et al., 2000; Houart et al., 2002;
Lagutin et al., 2003; Esteve et al., 2004). Several studies
provided evidence to support the role of noncanonical
Wnt signaling in the formation and the maintenance of
eye field, mediated by the morphogenetic movements of
RPCs into the eye field (Cavodeassi et al., 2005; Lee
et al., 2006) and the activation of eye field-specific genes
Pax6 and Rx (Rasmussen et al., 2001; Maurus et al.,
2005). In the chick and murine, TCF/LEF binding sites,
which shows activation of Wnt/b-catenin signaling, is
detected in the dorsal optic vesicle (Maretto et al., 2003;
Smith et al., 2005; Cho and Cepko, 2006). Deletion of
Wnt/b-catenin co-receptor LRP6 leads to the downregulaton of TCF/LEF expression and the loss of the dorsal
marker Tbx5, indicating that Wnt/b-catenin pathway
might be involved in the dorsoventral patterning in the
optic vesicles (Maretto et al., 2003). In frog, Wnt/b-catenin pathway is implicated in the regulation of RPC proliferation and neurogenesis in the developing retina
(Ladher et al., 2000; Galy et al., 2002; Van Raay et al.,
2005). In contrast, loss of function of b-catenin in the
murine retina does not affect proliferation or differentiation of RPCs, suggesting that Wnt/b-catenin pathway is
not required in mammalian retingogenesis (Ouchi et al.,
2005; Fu et al., 2006). However, Wnt/b-catenin signaling
pathway may play a role in retinal regeneration, since it
enhances stem cell-like properties of M€
uller glial cells in
adult mammalian eye (Das et al., 2006; Osakada et al.,
2007). In murine retinal explants, ectopic expression of
b-catenin inhibits neurite outgrowth, indicating that
Wnt/b-catenin pathway might also be involved in the
regulation of axonal guidance of projecting neurons,
RGCs, in the retina (Ouchi et al., 2005; Rodriguez et al.,
2005). Furthermore, Wnt signaling could also participate
in the medial-lateral retinotectal topographic mapping
in the murine (Schmitt et al., 2006). Interestingly, in
contrast, noncanonical Wnt signaling promotes RGC
axon outgrowth in chick and frog mediated by the interaction of Wnt antagonists SFRP1 and Fzd2 and the activation of pertussis toxin-sensitive Ga protein (Rodriguez
et al., 2005). It has been suggested that Wnt/b-catenin
in addition to FGF and TGF-b could participate in the
formation of ciliary body and iris (Zhao et al., 2002b,c;
Dias da Silva, 2007; Liu et al., 2007). In addition, Wnt/
b-catenin signaling may also function to keep PRCs in
the adult ciliary margin in an undifferentiated state so
as to maintain a continuous stem cell pool (Inoue et al.,
2006). Reports have demonstrated that suppression of
Wnt/b-catenin signaling is required in the lens ectoderm
to initiate lens formation, but is necessary during the
later stage of the lens morphogenesis for the proper
development of the lens epithelium and differentiation of
lens fibers (Smith et al., 2005; Kreslova et al., 2007).
Inhibition of Wnt/b-catenin pathway is also found to be
essential for the differentiation of cornea in the murine
(Mukhopadhyay et al., 2006). Finally, studies have demonstrated that Wnt/b-catenin signaling pathway controls
both the regression of hyaloids vessels and the development retinal vasculature, although by a different mechanism (Xu et al., 2004; Lobov et al., 2005; Masckauchan
et al., 2007; Rao et al., 2007).
Fibroblast growth factor (FGF) signaling
pathways. The FGF is a large family of neurtrophic
signaling molecules that are characterized by their
RSC AND REGENERATION OF VISION SYSTEM
ability to bind heparin and heparin sulfate proteoglycans
(HSPG) as cofactors for the activation of FGF tyrosine
kinase receptors (FGFR1–4) (Venkataraman et al., 1999;
Schlessinger, 2000). The interactions of the growth factors, proteoglycan, and FGFR mediate FGF dimerization
and activate multiple signal pathways, including those
involving Ras, mitogen-activated protein kinase
(MAPKS), extracellular signal-regulated kinases (ERKs),
phosolipase-gamma (PLC-Ç), Jun N-terminal kinase
(JNK), and protein kinase-C (PKC). FGFR activation
induces tyrosine phosphorylation of FGFR substrate-2
(FRS2) in a complex with growth factor receptor bound
protein-2 (GRB2) and Src Homology 2 Phosphatase-2
(SHP2), which promotes RAS activation and in turn activates the Raf1/MEK/MAPK pathway leading to change
in gene transcription (Turner and Grose, 2010). Acidic
FGF (aFGF or FGF-1) and basic FGF (bFGF or FGF-2)
are the two prototypical members of the FGF family
named because of their different isoelectric points.
FGF signaling has long been recognized for its neural
induction role in the developing embryo (Storey et al.,
1998; Stern, 2006) as well as for its role as the key mitogen for self-renewal of NSCs in vitro and in vivo (Gritti
et al., 1996; Reuss and von Bohlen und Halback, 2003;
Sirko et al., 2010). Indeed, FGF-2 in combination with
EGF is universally used to expand NSCs in the neurosphere assay. At least 10 of the 23 FGF ligands have
been described to be expressed in the brain. FGFR1 is
expressed as early as E8.5–9.5 in murine telencephalon
and persists in the ventricular zone and dentate gyrus
later on (Tropepe et al., 1999; Beer et al., 2000). Expression of FGFR2 and 3 have also been reported and seem
to be highly expressed by glial cells, mostly in the subependymal zone and SGZ but also around brain lesions
following trauma (Reuss and von Bohlen und Halback,
2003; Chadashvili and Peterson, 2006). The expression
of FGFRs and their ligands appears to be very dynamic
and their effects may depend on specific stages during
development and adult life (Ford-Perriss et al., 2001;
Temple, 2001; Fortin et al., 2005). Interestingly, bFGF
has been found to be closely associated with HSPGs in
NSC proliferation in vivo and may also in the regulation
of NSCs self-renewal in vitro (Kerever et al., 2007; Sirko
et al., 2010). Moreover, increased FGF signaling in the
radial glia along the ventricular zone in zebrafish does
not correlate with proliferative activity but rather correlates with the radial glia nature of ventricular cells
(Topp et al., 2008).
In the eye development, several FGFs are implicated
in playing a role in separating the neuroepithelium of
the optic vesicle into the neuroretinal and the RPE
domains. bFGF is highly expressed in the surface ectoderm overlying the optic vesicle in the chick (Pittack
et al., 1997). Presence of FGFs in optic vesicle culture
causes the pigmented epithelial cells to undergo neuronal differentiation resulting in the formation of a
double-layered neural retina. Conversely, when the
optic vesicles are cultured in FGF neutralizing antibodies neural differentiation in the retina is blocked but
the RPE is not affected (Pittack et al., 1997). Surgical
removal of the surface ectoderm results in a mixture of
mingled neural and pigmented cells in the optic vesicle.
Addition of FGF after ectoderm removal partially
restore the segregated neural and RPE domains, with
the neural domain being formed near to the FGF source
143
(Hyer et al., 2003). The results of these studies suggest
that FGFs are required for neural retina specification
and FGFs released from the overlying surface ectoderm
provide the positional cues that induce the neuroepithelium of optic vesicle to develop into an inner neural retina and an outer RPE. The specification of neural
retina and pigmented epithelium of the optic vesicle is
in part mediated by the inhibition of FGFs on specific
genes required for RPE determination. The expression
of specific bHLH-zipper transcription factor MITF is
initially found ubiquitously in the undifferentiated
murine optic vesicle and later become restricted to the
pigmented epithelium (Nguyen and Arnheiter, 2000).
Furthermore, in MITF mutant mice, parts of the future
pigmented epithelium are converted to a laminated
neural retina. Implantation of FGF-coated beads in the
presumptive RPE region leads to a downregulation of
MITF and the formation of neural retina. Conversely,
after the removal of the surface ectoderm, expression of
MITF is retained, neuroretinal-specific CHX10 expression is lost and the epithelium is converted to a pigmented monolayer. These effects can be reversed by the
application of FGF. Similar results has been found in
chick retina, overexpression of MITF causes hyperpigmentation and inhibits FGF-induced dedifferentiation
and transdifferentiation of RPE into neural retina
(Mochii et al., 1998). In addition to the FGF derived
from the surface ectoderm, FGF9 expression is also
found in the distal optic vesicle, which gives rise to the
neural retina. Ectopic or misexpression of FGF9 expression induces the conversion of RPE into neural retina
(Zhao and Overbeck, 1999; Zhao, 2001). In the FGF9
knockout mice, RPE extends into the outer neural
retina, suggesting a role of FGF9 in defining the boundary between the RPE and retina. In addition, in these
FGF9 deficient mice, the lens fiber cells are underdeveloped; indicating that FGF9 may involve in the
differentiation of lens fiber cells. Moreover, the transdifferentiation of the RPE into neural retina requires the
activation of RAS and MAPK/ERK signaling pathway
(Zhao et al., 2001; Galy et al., 2002).
FGFs are not only involved in the RPE and retina
specification, but also are implicated in retinal cell fate
determination. Blocking of FGF signaling with a FGFR
inhibitor retards the wave of RGC differentiation in
chick retinal explants, whereas exogenous FGF1 but not
FGF8 induces ectopic development of RGCs in the
peripheral retina, suggesting that FGFs play a role in
RGC differentiation and progression of wave of RGC differentiation (McCabe et al., 1999). Interestingly, in Xenopus, inhibition of FGF signaling with a dominant
negative FGFR causes a loss of photoreceptor and amacrine cells, with an increase of M€
uller glia (McFarlane
et al., 1998). However, overexpression of FGF2 in Xenopus RPCs increases the number of RGCs and decreases
the number of M€
uller glia. Although the proportion of
photoreceptors is unchanged, there is a two-fold increase
in rod photoreceptor cells compared with cone photoreceptors (Patel and McFarlane, 2000). Similar results
were obtained when FGF2 was added to embryonic rat
retinal explants cultures, accelerated the RGC differentiation of the uncommitted RPC, whereas anti-FGF antibodies delay their appearance (Zhao and Barnstable,
1996). In the same study, it was also reported that even
though differentiation of photoreceptors were not
144
YIP
affected, the rosette formation of rod photoreceptor is
suppressed. In addition to the role of FGF2 in the timing
of RGC differentiation, other studies have also shown
that FGF2 can regulate the proliferation of RPCs, and
shift the bias of retinal cell differentiation (Hicks and
Courtois, 1992; Lillien and Cepko, 1992). Furthermore,
FGFs are expressed in conjunction with Hhs in various
regions of the developing CNS, including optic vesicles
and retina (Crossley et al., 2001). In fish, Hh requires
FGF signaling to activate the Hh target gene Spalt in
the proximal eye vesicle at the eye and mid-hindbrain
boundary (Carl and Wittbrodt, 1999). FGFs and Hhs
have similar roles in dorsoventral patterning of the eye,
and in retinal cell fate specification and neurogenic
wave progression. Thus, it is possible that FGFs and
Hhs may act together or complement each other in these
important events during eye development.
The Nature and Origin of Intrinsic Retinal
Stem/Progenitor Cells
Retinal stem/progenitor cells in the ciliary
marginal zone (CMZ). In fish and amphibians, retinal growth is coordinated with the overall increase in
body size and retinogenesis occurs continuously throughout life. The addition of new neurons is generated from
two sources (Fig. 1). First, highly proliferative multipotent RPCs are found in the CMZ along the peripheral of
the retina (Straznicky and Gaze, 1971; Johns, 1977) and
the newly differentiated retinal cells are functionally
integrated into the existing circuits between the retina
and the iris epithelium (Otteson and Hitchcock, 2003).
Second, new rod photoreceptors are added to the central
retina in order to maintain visual acuity in the expanding retina. These rod photoreceptors are generated from
the rod progenitor cells in the ONL (Otteson and Hitchcock, 2003). It was originally thought that rod progenitors can generate multiple retinal cell types in response
to injury (Raymond et al., 1988). However, recent studies
suggest that new neurons in the regenerating retina are
derived from a population of slow-dividing stem cells in
the INL (Wu et al., 2001; Yurco and Cameron, 2005;
Fausett and Goldman, 2006). Furthermore, these proliferative cells of the INL express markers of PRCs such
as Pax6, Vsx1, Notch-3, and N-cadherin (Levine et al.,
1994; Hitchcock et al., 1996; Sullivan et al., 1997; Wu
et al., 2001), and have been suggested to be the true
PRCs that in intact retina give rise to the progenitor
cells for retinal regeneration, including rod progenitor
cells in the ONL. Interestingly, M€
uller glia whose cell
bodies are also located in the INL, proliferate after retinal injury and have not been ruled out as a source of
RPCs in fish (Braisted et al., 1994; Wu et al., 2001;
Yurco and Cameron, 2005; Fausett and Goldman, 2006).
However, these putative RPCs in the INL have not been
isolated and propagated in vitro making the examination
of their stem cell potential difficult. Nevertheless, it is
possible that the INL-derived stem cells, rod progenitor
cells and M€
uller glia can all serve as sources of regenerating retinal cells and the final cell fate decisions
depends on the type and extent of damage.
The ability of the retina of fish to regenerate within
several days to weeks after subjected to various types of
injury has been well documented (see reviews in Fadool,
2003; Otteson and Hitchcock, 2003). However, it should
be noticed that the functionality of the newly regenerated neurons and reintegration of these neurons into
existing circuitry has not been fully examined. In fact, it
has been demonstrated that after injury, regenerated
fish retinal tissue fails to reform a proper cone photoreceptor mosaic pattern (Vihtelic and Hyde, 2000; Stenkamp et al., 2001; Raymond et al., 2006). In amphibians,
CMZ stem cells are also implicated for cellular regeneration after retinal injury (Keefe, 1973; Reh, 1987; Reh
and Nagy, 1987; Perron et al., 1998). Moreover, it has
demonstrated that loss of specific type of retinal neurons
after neurotoxic injury promotes the production of that
particular neuronal cell types (Reh and Tully, 1986; Reh,
1987). In contrast to fish, the size and the regenerative
potential of the CMZ in the amphibian retina reduce
drastically after metamorphosis (Moshiri et al., 2004). In
birds, retinogenesis is completed and all neurons are
generated by hatching (Prada et al., 1991). Postnatal
growth of the avian retina is due to the tissue stretching
of the growing sclera and not as a result of addition of
new neurons (Amato et al., 2004). A CMZ-like germinal
zone at the peripheral margin of the retina was nevertheless identified in the chick (Fischer and Reh, 2000)
and quail (Kubota et al., 2002). Although the cells at the
retinal margin of post-hatched chicks have the capacity
to divide, express PRC genes, such as Pax6 and Chx10,
and generate neurons, the neurogenic potential of these
cells are fairly restricted compared with the cells in the
CMZ of the fish and amphibian. In contrast to the fish
and amphibian CMZ that can generate all retinal cell
types, CMZ in the chick regenerate only amacrine and
bipolar cells, but not photoreceptor and ganglion cells
and only in small quantity (Fischer and Reh, 2000).
However, this restriction can be overcome by exogenous
growth factor stimulation, suggesting that extrinsic and
not intrinsic factors limit the neurogenic potential of the
CMZ-derived RPCs in the chick (Fischer and Reh, 2000).
Furthermore, in contrast to fish and amphibian, avian
CMZ-derived RPCs do not response to retinal damages
and contribute to retinal regeneration (Fischer and Reh,
2000).
The CMZ is either greatly reduced or absence in mammalian eye. Several studies that searched for a comparable region in the mice, rats, and macaques indicated
that there is no CMZ in mammalian retina (Ahmad
et al., 2000; Kubota et al., 2002; Moshiri et al., 2004).
However, the ciliary body in the adult murine eye contains a population of quiescent cell that can be expanded
in vitro (Ahmad et al., 2000; Tropepe et al. 2000). The
ciliary body is derived from neuroepithelium and located
behind the iris at the distal margin of the neuroretina.
The CE is composed of two layers. The inner layer is
transparent and unpigmented, and is continuous from
the neural tissue of the retina. The outer layer is highly
pigmented, continuous with the RPE. This doublelayered epithelium of the ciliary body is often regarded
to be continuous with the retina and a rudiment of the
embryological retina. Cell labeling studies with chronic
injections of BrdU revealed that a rare population (0.2%)
of proliferative cells is found in the pigment layer of the
ciliary body in adult rats (Ahmad et al., 2000). A population of nestin-expressing cells in the adult mice ciliary
body responds to RGC injury by increase in proliferation
and upregulation of homeodoamin protein Chx10 and
recoverin, which is normally expressed in photoreceptors
RSC AND REGENERATION OF VISION SYSTEM
and bipolar cells (Nickerson et al., 2007). Dissociated
cells of the ciliary body from adult mice proliferate in
vitro, forming pigmented neurospheres that can be
expanded to secondary neurospheres in subsequent passages (Tropepe et al., 2000). Furthermore, exogenous
FGF and pigment epithelium-derived factor (PEPF) can
enhance the formation and proliferation of pigmented
neurospheres (Tropepe et al., 2000; De Marzo et al.,
2010). These pigmented neurosphere cells (PNCs)
express RPC marker Chx10 and EFTFs such as Pax6,
Six3, and Rx (Ahmad et al., 2000; Tropepe, et al., 2000;
Lord-Grignon et al., 2006), and can be induced to differentiate along neuronal and glial lineages (Ahmad et al.,
2000). Wnt3a, a canonical Wnt ligand, stimulates the
proliferation and multipotency of PNCs (Inoue et al.,
2006). Moreover, PNCs can express specific markers and
differentiate into different retinal cell types, including
rod photoreceptors, bipolar neurons, RGCs, and M€
uller
glia (Ahmad et al., 2000; Tropepe et al., 2000). Although
porcine CE-derived cells express b–III-tubulin and NeuN in vitro, they fail to express specific retinal markers
(MacNeil et al., 2007). Recently, it was suggested that
the efficient differentiation of CE-derived cells to acquire
RPE-like phenotypes depends on the conditions of differentiation protocols and cellular environment in vivo;
suggesting that pre-differentiated or re-programmed porcine CE-derived cells may have better potential for retinal repair (Guduric-Fuchs et al., 2011). Several groups
have reported that pigmented cells isolated from the
adult human CE can transdifferentiate to retinal
progenitor-like cells (Ahmad et al., 2000; Fischer and
Reh, 2003; MacNeil et al., 2007). Cells with RPE features have also been differentiated from postmorteum
human eyes and can be induced to differentiate into photoreceptors after transplanted into adult mice eyes
(Coles et al., 2004). Thus, these studies indicate that
RPCs in the pigmented ciliary body display stem cell
properties and have the capacity to generate different
retinal neurons in vitro. However, it is important to
point out the fact that these PRCs in the CE are rare,
fully differentiated, and pigmented epithelial cells which
are different from the naive, undifferentiated and unpigmented stem cells found in the CMZ of lower vertebrates
or in the neurogenic regions of adult mammalian brain.
Furthermore, the capacity of these CE-derived adult
RPCs to proliferate and self-renew gradually decrease
with sequent passages and expansion (Coles et al., 2004;
Xu et al., 2007) and eventually lose the ability to differentiate into photoreceptors (Gualdoni et al., 2010).
Indeed, doubts have been raised recently over the identity and retinogenic potential of the RPCs in the CE
(Cicero et al., 2009; Gualdoni et al., 2010).
Retinal stem/progenitor cells in the RPE. RPE
arises from neuroectoderm which it shares with neural
retina in early development and plays multifunctional
roles in support of the vertebrate eye (Bok, 1993; Boulton and Dayhaw-Barker, 2001). The embryonic RPE is
capable of transdifferentiation into a neural retina in
many vertebrate species; including mammal (Zhao et al.,
1995; Layer PG, 1998; Cayouette et al., 2001; Jensen
et al., 2001; Del Rio-Tsonis and Tsonis, 2003), perhaps
because they are developed from a common origin. During the transdifferentiation process, the pigment epithe-
145
lial cells lose their pigmentation, proliferate, and begin
to express markers of RPCs (Okada, 1980). The dedifferentiated epithelial cells then proceed to generate new
retinal neurons and glia in a similar manner that resembles normal retinogenesis (Reh et al, 1987; Sakaguchi
et al., 2003).
Interestingly, it has been demonstrated that RPE in
amphibians, but not in fish, has the ability to transdifferentiatite into neurons and glia (Ikegami et al., 2002;
Susaki and Chiba, 2007). In addition, the ability of adult
RPE to transdifferentiate into retinal neurons and to
regenerate the entire retina is retained only in urodela
(newts and salamanders) (Fig. 1) and not in anura
(frogs; Raymond and Hitchcock, 1997; Reh and Fischer,
2001; Hitchcock et al., 2004; Klassen et al., 2004a).
Despite evidence of neuronal transdifferentiation of RPE
cells in adult rodent and human has been demonstrated
in cultures (Vinores et al., 1995; Chen et al., 2003; Amemiya et al., 2004; Engelhardt et al., 2005), the potential
of these cells as RPCs has not been clearly documented
in vivo. Adult rat mammalian peripheral RPE although
retains its ability to proliferate, albeit at a much slower
rate, appears to lose the ability to transdifferentiate into
diverse retinal cell types (Al-Hussaini et al., 2008).
Retinal stem/progenitor cells in the iris
pigment epithelium (IPE). IPE develops from the
inner layer of the optic cup shares a common embryonic
origin as the neural retina. The iris tissue has long been
known for its remarkable ability to regenerate lens in
newt, chick, and human under culture conditions (Eguchi 1971, 1986; Tsonis et al., 2001; Kosaka et al., 2004).
Epithelial cells of the iris from amphibians (Eguchi,
1986), birds (Sun et al., 2006), and rodents (Haruta
et al., 2001; Asami et al., 2007) exhibit stem cell-like
properties in vitro. Nestin-positive RPCs are located in
the pigmented inner layer of the iris epithelium adjacent
to the eye chamber (Yamaguchi et al., 2000; Asami
et al., 2007). When the iris tissue of adult rats was cultured in retinal cell differentiation medium with bFGF,
some of the iris-derived cells express differentiated neuronal marker, neurofilament 200, but not rhodopsin, a
specific marker for rod photoreceptors. However, ectopic
expression of Crx, a hemeobox gene critical for photoreceptor differentiation and is specifically expressed in the
photoreceptors in the mature retina, in the adult rat
iris-derived cells resulted in the expression of rhodopsin
and adoption of a photoreceptor-specific phenotype (Haruta et al., 2001; Akagi et al., 2005). A combination of
Crx and NeuroD was needed for the generation of irisderived photoreceptor cells in primate. Moreover, these
photoreceptor-like cells display electrophysiological characteristics of rod photoreceptors and are capable to integrate and survive in the cocultured embryonic retinal
explants (Akagi et al., 2005). More recently, it was demonstrated that a combination of CRx, Rx, and NeuroD
converts IPE cells into light responsive photoreceptorlike cells in human (Seko et al., 2012). Furthermore,
pure isolated IPE cells from postnatal chick and adult
rats cultured in the presence of bFGF form neurospheres, which express RPC markers Pax6, Chx10, and
vimentin (Sun et al., 2006; Asami et al., 2007). When
IPE-derived neurosphere cells were cultured on laminincoated dishes with bFGF, a subset initiated the
146
YIP
expression of TuJ1, GFAP, and O4. Some neurosphere
cells expressed retinal-specific neuronal markers, such
as rhodopsin (rod photoreceptor), iodopsin (cone photoreceptor), PKC (bipolar cell), and HPC-1 (amarcrine), suggesting that IPE-derived cells have the potential to
generate retinal-specific neurons in vitro (Sun et al.,
2006). IPE cells co-cultured with embryonic RPCs participate in the formation of spheroids and express rod photoreceptor marker rhodopsin only in the ONL and
M€
uller cell marker vimentin only in the INL (Sun et al.,
2006). The results indicate that IPE-derived cells could
respond to lamina-specific cues for differentiation and
integration similar to RPCs (Rothermel et al., 1997).
When IPE cells were grafted into the space between
RPE and the photoreceptor layer, IPE cells were incorporated in the subretinal space and differentiated into rod
photoreceptors. In the rat IPE, the inner and outer
layers of the IPE differentially expressed nestin in a
manner corresponding to their shared origins with the
neural retina and the pigmented epithelial layers,
respectively. The nestin-expressing cells are located only
in the inner IPE layer. These cells proliferate and differentiated into retina-specific cells in response to bFGF,
whereas cells that do not express nestin do not proliferate and have restricted neuronal potency and only
express pan-neural markers (Asami et al., 2007). The
results from these experiments suggest that heterogeneous populations of RPCs displaying different developmental potential exist postnatally in the IPE, and some
of them are able to differentiate into multiple retinal cell
types without gene transfer. However, just like the RPCs
derived from the ciliary pigment epithelium, recent
study has demonstrated that although human iris cells
proliferate and express low levels of RPC or neuronal
markers, many of them retained properties of differentiated epithelial cells and lack central properties of neural
stem cells to differentiate into retinal neurons (Cicero
et al., 2009; Moe et al., 2009; Gualdoni et al., 2010; Bhatia et al., 2011; Froen et al., 2011). Thus, functional
studies are essential to determine the potency of irisderived cells to differentiate into fully operative photoreceptors before it can be considered as a potential source
of cell-based therapy for retinal degenerative diseases.
Retinal stem/progenitor-like M€
uller cells in
the retina. M€
uller cells are the major glial cells in
the retina comprising about 5% of all retinal cells. During retinal histogenesis, M€
uller cells are generated last
(Marquardt and Gruss, 2002; Cayouette et al., 2006).
M€
uller cells are astrocyte-like radial glial cells with cell
bodies located in the INL and processes that span across
the retina from the vitreal surface to RPE. M€
uller glia
ensheath all retinal neurons and thus cells play crucial
roles in supporting the neurons and their functions.
From early stages of development, they are essential in
maintaining the homeostasis of the retinal tissue, providing structural support, participating neuronal signaling processes, contributing to the formation of bloodretina-barrier and regulating blood flow (Bringmann
et al., 2006, 2009; Reichenback et al., 2007). Similar to
astrocytes, M€
uller cells express GFAP and glutamate
sythetase. M€
uller cells become activated following injury
and form a protective barrier between the healthy and
damaged tissues. Reactive M€
uller glia response to injury
by increase in proliferation, and upregulation of GFAP
and neurotrophic factors expression. Reactive glia also
participate in wound healing, stabilizing damaged tissue, attracting inflammatory cells, and promoting neuronal survival. Moreover, reactive glia can also induce
apopotitic cell death (Bringmann et al., 2006).
M€
uller cells also display neurogenic properties in the
vertebrate retina. In the fish retina, M€
uller glia dedifferentiate and generate rod progenitor cells that give
rise to rods throughout life (Bernardos et al., 2007).
These stem cells can regenerate all retinal cell types,
including cone photoreceptors, in response to damage
(Easter and Hitchcock, 2000; Fausett and Goldman,
2006; Hitchcock and Raymond, 2004; Ramachandran
et al., 2010). Following injury, M€
uller cells reenter the
cell cycle, dedifferentiate, and become RPCs that generate neurons (Raymond and Hitchcock, 2000; Yurco and
Cameron, 2005; Fausett and Goldmann, 2006; Fimbel
et al., 2007) (Fig. 1). In addition, in response to ablation
of cone photoreceptor, it was shown that the subsequent
regeneration is biased towards in replacement of the
cognate cone photoreceptor cell type (Fraser et al.,
2013). Unlike the M€
uller cells in birds and mammals
uller glia in normal fish are GFAP1.
which are GFAP2, M€
Since GFAP is expressed by the NSCs in the neurogenic
regions in mammals, this could be an indication of the
“neurogenic” nature of the fish M€
uller glia compared to
birds and mammals. In the avian retina, when extensive
amacrine cell death was induced by intraocular injections of neurotoxin NMDA, M€
uller cells become activated, proliferated, and dedifferentiated (Fischer and
Reh, 2001a,b). Although the majority of the cells remain
undifferentiated or persisted as M€
uller glia, and continue to express RPC markers Chx10 and Pax6 (Fischer
and Reh, 2003), a small number of the M€
uller cells
transdifferentiated into retinal neurons, particularly
into amacrine and bipolar cells (Fischer and Reh,
2001a,b). Furthermore, there is a cell-type-specific
replacement of neurons that have been selectively
injured. Generation of Brn31 RGCs, in the presence of
bFGF and insulin, was observed after treatment of colchicine or kainate both selectively causes RGC cell
death. In addition, bFGF and insulin can also stimulate
M€
uller cells to transdifferentiate (Fischer and Reh,
2002). M€
uller glial cells of the mature mammalian retina
express genes that are specific for M€
uller glia and genes
that are also expressed in RPCs (Livesey et al., 2004;
Roesch et al., 2008). In culture, M€
uller cells generated
neurospheres and transdifferentiated into all three types
of glial cells in the CNS as well as retinal neurons (Das
et al., 2006; Monnin et al., 2007; Nickerson et al., 2008;
Takeda et al., 2008; Wan et al., 2008). M€
uller cells
remain dormant in normal adult rat retina, however, in
response to injury, they proliferate, become activated,
(Dyer and Cepko, 2000; Karl et al., 2008), change their
gene expression, for example, upregulation of GFAP,
Sox2, cyclinD, and nestin, and convert to retinal stem
cells or other retinal cell types (Bernardos et al., 2007).
Moreover, extrinsic and intrinsic cues promoted and controlled the differentiation of these cells to specific retinal
neurons (Kim et al., 1998; Lewis and Fisher, 2003; Ooto
et al., 2004; Kohno et al., 2006; Osakada et al., 2007;
Wohl et al., 2009; Xue et al., 2010). Addition growth factor treatment promoted the transdifferentiation of
M€
uller glia into retinal neurons (Karl et al., 2008),
RSC AND REGENERATION OF VISION SYSTEM
147
Fig. 1. Retinal stem cells (RSCs) at the periphery of the eye (CMZ
progenitor cells) and from the neuroepithelium spanning the width of
the developing retina (central progenitor cells) give rise to proliferating
retinal progenitor cells (RPCs). The cycling (intrinsic property) retinal
stem/progenitor cells are able to response to local environment
(extrinsic cues) to promote their transition from proliferation to differentiation. Generation of retinal cell types follows a temporal sequential
€ller cells are generated last.
order: RGCs are generated first, and Mu
Retinal cell fate commitment and specification are regulated by combinations of transcription factors (as shown). In retinal regeneration
€ller cells dedifferentiating and then give rise to differafter damage, Mu
ent retinal cell types mirrors the RSCs in the CMZ and central retina in
patterns of gene expression and cellular organization. RPE in response
to injury cues is capable of transdifferentiation into neuronal and glial
cell-phenotypes, inducing neural retinal regeneration. CMZ: ciliary marginal zone; NFL, nerve fiber layer; GCL: ganglion cell layer; IPL: inner
plexiform layer; INL, inner nuclear layer; OPL: outer plexiform layer;
ONL: outer nuclear layer; OS: outer segment of photoreceptors; RPE:
retinal pigment epithelium.
whereas growth factor treatment alone rarely activates
M€
uller cells (Close et al., 2006; Karl et al., 2008). In
human retina, a population of M€
uller glia with NSC
characteristics has been identified and they can be
induced to grow and differentiate into retinal neurons in
vitro (Limb et al., 2002; Lawrence et al., 2007; Bull
et al., 2008; Lamba et al., 2009a; Bhatia et al., 2010,
2011; Giannelli et al., 2011; Singhal et al., 2012), suggesting that these cells may be a promising source of
cells for cell-based therapies to treat retinal degenerative
diseases (Fig. 1).
tive pathway. One possible strategy for treatment of
these blinding diseases is to replace cells that are lost
via transplantation. One of challenges in this approach
is to identify and characterize sources of cells for transplantation. Several cell populations may be regarded as
potential sources for retinal transplantation, they
include NPCs derived from central nervous system
(CNS), adult stem cells such as mesenchymal stem cells,
embryonic stem cells (ESCs) and induced pluripotent
stem cells (iPSCs) (Fig. 2).
Retinal Stem Cell in Regenerative Medicine
Regeneration of retinal cells from retinal progenitor cells. Retinal progenitor cells (RPCs) derived
Treatments to repair the human retina following
degenerative diseases remain a challenge. Unlike species
of lower vertebrates, the human retina lacks a regenera-
from fetal or neonatal retinas comprise a population of
immature cells that is responsible for generation of all
retinal cells during development (Reh, 2006). RPCs have
148
YIP
been successfully isolated from several mammalian species, including rodents (Chacko, et al., 2000; Yang et al.,
2002; Akimoto et al., 2006), pigs (Klassen et al., 2007),
and human (Yang et al., 2002). By manipulating time
and environment in vitro, immature RPCs can be
expanded extensively in culture and express photoreceptor markers (Merhi-Soussi et al., 2006). Transplantation
of RPCs from the developing retina into dystrophic
mature retina promotes survival of host tissue, along
with integration into the neural retina and recovery of
light-mediated behavior (Klassen et al., 2004b; MacLaren et al., 2006; Aftab et al., 2009). However, these early
studies of photoreceptor transplantation have only met
with minimal success due to the limited ability of the
cells to invade and integrate into the recipient retina.
The environment of the degenerating retina is hostile to
the transplanted cells and has adverse effects on the
ability of the cells to migrate from the transplantation
site into the host retina (Kinouchi et al., 2003; Ma et al.,
2011). Not until recently, several publications have
shown high levels of integration of transplanted photoreceptor precursors in advanced degenerated retina. Using
transplantation of newborn rod precursors grafting in a
murine model of severe human retinitis pigmentosa,
Singh et al. (in press) provides evidence for replacement
of polarized ONL with light-sensitive outer segments,
reconnecting with host retinal circuit leading to visual
function recovery. In addition, it has been shown that
the outcome of rod-photoreceptor precursor transplantation depends on different types and stages of degeneration (Pearson et al., 2012). Thus, effective rodphotoreceptor transplantation can be achieved by tailored manipulations of the host environment and appropriate therapeutic time windows (Barber et al., 2013).
However, since the protocol involves harvesting RPCs
from fetal eyes poses an ethical issue for clinical applications, it cannot be translated to human patients.
Regeneration of retinal cells from adult stem
cells. Adult stem cells, which have been identified in a
variety of tissues, such as in the epidermis, cornea,
intestinal, peripheral blood, and bone marrow, are multipotent with a limited capacity of self-renewal and differentiation to certain cell types (Fig. 2). Thus, adult stem
cells can be obtained from the patients and used as autologus grafts without rejection. Recently, it has been demonstrated that hematopoietic stem cells of the bone
marrow can be differentiated into various lineage cells
including neural cells and astrocytes in vitro (SanchezRamos et al., 2000; Woodbury et al., 2000) and in vivo
(Eglitis and Mezey, 1997; Kopen et al., 1999; Brazelton
et al., 2000; Mezey et al., 2000). Interestingly, hematopoietic stem cells have also been reported in animal
models to have neuroprotective effects on retinal diseases (Harris et al., 2009; Marchetti et al., 2010). Adult
bone marrow-derived nonhematopoietic lineage stem
cells incorporate into the degenerating blood vessels following intravitreal injections in neonatal mice and rescue cone photoreceptor in murine models of RP (Otani
et al., 2004). Some of these bone marrow mesenchymal
stem cells migrated into the retina, differentiated into
microglia and promote vascular repair in the ischemicinduced retinopthy (Ritter et al., 2006). In addition,
RPE-induced bone marrow stem cells mobilized into
peripheral blood can home to focal damaged RPE tissue
in the subretinal space and express RPE-specific cell
markers (Li et al., 2007). Recent study demonstrates
that a rare population of very small embryonic/epiblastlike stem cells (VSEL) in the murine retina can also differentiate into RPE-like cells (Zuba-Surma et al., 2008;
Liu et al., 2009). Adult bone marrow stem cells transplanted into the adult degenerative (Kicic et al., 2003) or
mechanical injury rat eye (Tomita et al., 2002; Inoue
et al., 2007) slowed down retinal cell degeneration and
integrated into the retina and differentiated into photoreceptor cells. However, it has been demonstrated that
neuroprotective properties of the bone marrow stem cells
may be attributed to the secretion of neurotrophic factors (Crigler et al., 2006) and/or anti-inflammatory modulators (Pluchino et al., 2005) by these cells, instead of
direct functional retinal cell replacement (LevkovitchVerbin, 2010). Furthermore, although autologous transplantation of adult stem cells has the advantage of
reducing the risk of rejection and eliminating ethical
issues, the scarcity of these cells and the restriction in
their functional retinal differentiation have limited the
application of adult stem cells in stem cell therapies for
retinal diseases.
Regeneration of retinal cells from ESCs and
iPSCs. The current state of cell replacement-therapy
for the treatment of retinal diseases focus on the development of protocols on the direct differentiation of
hESCs or hiPSCs to RPCs and a photoreceptor cell phenotypes. ESCs are derived from the inner cell mass of
the embryonic blastocyst, with self-renewal capabilities
and the ability to differentiate into cell types derived
from all three embryonic germ layers (Thomson et al.,
1998; Reubinoff et al., 2000; Cowan et al., 2004). In vitro
differentiation of mouse and human ESCs into different
functional retinal cell types (Fig. 2), in particular RPE
cells and/or photoreceptors, has been demonstrated by
numerous stepwise protocols (Zhao et al., 2002a,b; Ikeda
et al., 2005; Lamba et al., 2006; Carr et al., 2009a; Idelson et al., 2009; Osakada et al., 2008, 2009a,b). Studies
have shown that transplantation of mouse, primate, and
human ESC-derived retinal cells, including RPE cells, in
rodent retinal degeneration models protected host photoreceptors, integrated into the recipient retina, differentiated into functional photoreceptors and restored visual
function (Haruta et al., 2004; Meyer et al., 2005; Lund
et al., 2006; Meyer et al., 2004, 2006; Vugler et al., 2008;
Idelson et al., 2009; Lamba et al., 2009b; Park et al.,
2011; Vaajasaari et al., 2011). Thus, transplantation of
photoreceptors with or without RPE cells derived from
the hESCs offers huge potential for cell replacement
therapy in treating retinal degenerative diseases (Jin
et al., 2009). Clinical trials in the United States using
human ESC-derived RPE to treat Stargardt’s disease
and AMD were approved by the FDA (Schwartz et al.,
2012). Furthermore, mouse ESCs can be induced to generate an eye-like structures made up of lens cells, retinal
cells and RPE cells (Hirano et al., 2003) and subsequently it has been shown that cells from these eye-like
structures can be differentiated into RGCs when transplanted into the vitreous body of an injured adult mouse
retina (Aoki et al., 2008). Most recently, in a pioneering
study by Eiraku et al., mouse ESCs aggregates can
RSC AND REGENERATION OF VISION SYSTEM
149
Fig. 2. Sources of replacement cells to provide treatments for people who suffer from retinal diseases:
(A) ESCs from the inner cell mass of the blastocyst; (B) adult stem cells from the bone marrow, brain, and
the eye; and (C) the iPSCs and induced retinal neurons (iRNs) derived from human fibroblasts. NSC: neu€ller glial cells. Dash line with
ral stem cells; RPCs: retinal progenitor cells; C: cone photoreceptor; M: Mu
question mark: suggested differentiation.
organize into stratified optic-cup in a three-dimensional
(3D) culture system (Eiraku et al., 2011; Eiraku and
Sasai, 2012a,b). Furthermore, the entire process of this
spontaneous optic-cup morphogenesis follows the typical
spatial and temporal histogenic sequence occurring during retinal development in vivo (Eiraku et al., 2011). In
the past few years, several groups have attempted to
reconstitute 3D retinal tissue in vitro using human
ESCs (Dutt and Cao, 2009; Nistor et al., 2010). The
ESCs aggregate and form sheets of differentiated retinal
cell types, but fail to organize into a typical laminated
3D retinal structure. In 2011, Meyer et al. made an
important breakthrough by demonstrating that a 3D
optic vesicle-like structure can be obtained using human
ESCs and iPSCs (Meyer et al., 2011) by the induction of
eye-field markers such as Rx and Pax6 (Meyer et al.,
2009). However, even with a high degree of neuroretinal
differentiation, RPE was rarely developed in the optic
vesicle-like structures without Activin A supplementation that mimic certain aspects of embryonic RPE devel-
opment. Although ESC-derived RPE has been shown to
have some success with transplantation and thus support the feasibility of using the human ESCs for cellbased retinal regenerative therapy, the potential of using
ESCs in cell replacement therapy for treatment of retinal diseases is limited by important ethical issues and
the risk of immune rejection. In addition, na€Ä±ve ESCs
have been associated with tetratoma formation after
transplantation (Arnhold et al., 2004; Cowan et al.,
2004) and the efficiency of generation of functional RPE
cells is too low and the timing is too slow for the narrow
window of effective therapy.
An alternate source of cells for stem cell transplantation in retinal regeneration is iPSCs (Fig. 2). iPSCs are
ESC-like pluripotent cells that are reprogrammed in
vitro from terminally differentiate somatic cell without
using embryonic tissue (Takahashi and Yamanaka, 2006;
Takahashi et al., 2007; Wernig et al., 2007; Yu et al.,
2007; Nakagawa et al., 2008) and have the potential to
differentiate into all cell types of the adult organism.
150
YIP
Thus, iPSCs, unlike ESCs, have been identified as an
unlimited source of replacement tissue for use in human
retinal cellular therapies without ethical implications. In
addition, transplantation of autologus grafts from
patient’s own iPSC-derived cells could avoid the need of
long-term immunosuppression of graft rejection. Recent
studies have shown that human iPSCs could be genetically reprogrammed using either three or four transcription factors (Oct4/Sox2/Klf4 or Oct4/Sox2/Klf4/c-Myc,
respectively) and induced to differentiate into retinal
cells by small molecules, including RGCs and photoreceptors (Osakada et al., 2009a,b; Hara et al., 2012).
iPSCs have the same potential as the ESCs to mimics
retinal development in situ (Meyer et al., 2009). In addition, similar to human ESCs, human iPSCs can spontaneously differentiate into RPE cells (Buchholz et al.,
2009; Carr et al., 2009b; Meyer et al., 2009) and the differentiation process can be greatly facilitated by addition
of compounds such as Dkk1 and Lefty-A that are
involved in the developmental signaling pathways of
RPE (Hirami et al., 2009; Osakada et al., 2009b). It is
reported that transplantation of iPSC-derived RPE
exerts protective effects and restores visual functions in
retinal dystrophic rats (Buchholz et al., 2009; Carr
et al., 2009b; Tucker et al., 2011). Thus, retinal cells
such as iPSC-derived RPE generated from patients with
retinal degenerative diseases are of particular interest
because these cells reveal a disease-specific functional
defect that can be corrected either by pharmacological
treatment or by following gene target repair (Meyer
et al., 2011). Recently studies have shown that iPSCs
from swine eye which shares a close similarity to the
human eye can differentiate into photoreceptors in vitro,
and these cells can be transplanted and integrated in
damaged swine retina, thus provides an appropriate system for the evaluation of potential therapeutic strategies
for eye degenerative diseases, including RP and AMD
(Hendrickson and Hicks, 2002; Guduric-Fuchs et al.,
2009; Zhou et al., 2011). However, major safety concerns
of using clinical application of iPSCs include the potential risk of malignant formation results from the oncogenic properties of the transcription factors used in the
reprogramming protocols and the random genomic integration of these factors after retroviral transduction.
Extensive studies have been undertaken to establish
new reprogramming protocols that use nonintegrating
gene delivery methods (Yu et al., 2009) or that replace
the application of exogenous reprogramming factors by
treatment with proteins (Kim et al., 2009) or small molecules (Li et al., 2009).
CONCLUDING REMARKS
Eye formation requires the coordination of complex
interactions from multiple cellular sources to create the
cell behaviors that progressively shape the developing
eye. Research into understanding the mechanisms that
regulate stem cells in progenitor cell fate determination
and their subsequent differentiation during eye development is still far from complete. Nevertheless, significant
progress has been made towards the identification of various extrinsic cues and intrinsic determinants involved in
RPC specification. The mechanisms of development and
differentiation of eye are remarkably similar in all vertebrates. During retinogenesis, proliferating RPCs and
newly generated cells are confined at the peripheral margin of the retina. In fish and amphibians, this region is
maintained after embryonic development and this specialized region referred to as the CMZ. The retina of many
fish and amphibians continue to grow throughout their
life. The increase in retinal size is due to in part to the
addition of new neurons, at the CMZ. In birds, neurogenesis at the CMZ decreases dramatically than that observed
in fish and amphibians. Furthermore, in rodents the retinal margin does not exhibit mitotic activities after the
first week of postnatal life. It is interesting to note that
there might be a direct correlation of the evolutionary
importance of the ability of retina to regenerate with the
presence of RPCs and their potential to generate retinal
neurons. Regeneration of retina is frequently observed
fish and amphibians. In fish, the major source of new neurons is the stem cells of the marginal zone and the rod
progenitors in the INL and ONL in association with
M€
uller glia cells. Neurogenesis in adult fish visual system
has provided valuable insights into the regulatory mechanisms and potential of adult neural stem cells, and the
basic molecular and cellular processes underlying neurogenesis and cell specification. Adult mammalian retina
has long been known to be devoid of stem cells and has
lost the ability to regenerate after damage.
Nevertheless, several groups have reported that pigmented cells isolated from the adult human CE can
transdifferentiate to retinal progenitor-like cells and
M€
uller glia cells can display characteristics of NPCs,
thus identified both cell populations as potential candidate for stem-cell based therapies to regenerate visual
function. It seems logical that it is preferable to mobilize
endogenous RPCs to drive the repair process in the retina. However, the challenge of using endogenous RPCs
for self repair will be to identify appropriate cellular
sources and molecules, including pharmacological
agents, that can expand the endogenous cell pool and
reactivate the regenerative processes similar to those
described for the lower vertebrates in the mammalian
retina. Recent advances in stem cell research have
raised the possibility to use hESCs and HiPSCs to repair
or regenerate damaged mammalian retina. Cell transplantation is the most direct approach towards replacing
damaged retinal cells and restoration of lost visual function. To achieve a breakthrough in cell replacement
therapies in retinal degenerative diseases would require
isolation and molecular characterization of human RPCs
for specific neuronal replacement in the actively degenerating adult retina and that these new cells survive
without immune suppression as well as displaying evidence of integration into host circuitry. Information on
the phenotypic potential and immunogenicity of the
donor cells would most likely benefit from future clinical
application of these cells. Findings in the understanding
the role of coordinated transcription factor expression in
fate specification should have predicative significance in
phenotypic potential and improve our ability to control
stem cell differentiation. However, the mammalian retina is a highly complex organized structure whose putative neurogenic potential we just begin to understand.
Thus, even if methods for successful cell replacement of
new neurons are established, there is no guarantee that
regenerating axons from these newly replaced cells
would grow towards the correct targets and be functional integrated into existing neural circuits.
RSC AND REGENERATION OF VISION SYSTEM
Furthermore, the inhibitory environment in the injured
adult mammalian CNS caused by myelin-associated glycoproteins and extracellular matrix molecules such as
chondroitin sulphate proteoglycans would strongly
inhibit axonal, and therefore, cellular regeneration. In
addition, inhibitory molecules secreted by activated
microgla and astrocytes in response to injury, further
exacerbate this hostile environment. Therefore, to have a
successful migration, integration, and synaptic formation
of grafted or endogenous RPCs, the inhibitory environment in the degenerating tissue has to be overcome by
pharmacological approaches. Regenerative medicine for
retinal degenerative diseases must take a combinatory
approach involving exogenous reprogramming or gene
transfer of RPCs, together with modulation of the
changes of environmental cues required for the regenerative processes within the appropriate time frame. While
much needed to be demonstrated, particularly recovery of
visual function, this challenging and multifaceted
approach to RPC transplantation provide an exciting new
strategy for the treatment of retinal disease and offers
the hope that effective treatments may be within reach in
the near future.
LITERATURE CITED
Adler R, Belecky-adams TL. 2002. The role of bone morphogenetic
proteins in the differentiation of the ventral optic cup. Development 129:3161–3171.
Aftab U, Jiang C, Tucker B, Kim JY, Klassen H, Miljan E, Sinden
J, Young M. 2009. Growth kinetics and transplantation of human
retinal progenitor cells. Exp Eye Res 89:301–310.
Ahmad I, Tang L, Pham H. 2000. Identification of neural progenitors in the adult mammalian eye. Biochem Biophy Res Commun
270:517–521.
Akazawa C, Sasai Y, Nakanishi S, Kageyama R. 1992. Molecular
characterization of a rat negative regulator with a basis helixloop-helix structures predominantly expressed in the developing
nervous system. J Biol Chem 267:21879–21885.
Akagi T, Akita J, Haruta M, Suzuki T, Honda Y, Inoue T, Yoshiura
S, Kageyama R, Yatsu T, Yamada M, Takahashi M. 2005. Invest
Ophthalmol Vis Res 46:3411–3419.
Akimoto M, Cheng H, Zhu D, Brzezinski JA, Khanna R, Filippova
E, Oh ECT, J Y, Linares J-L, Brooks M, Zareparsi S, Mears AJ,
Hero A, Glaser T, Swaroop A. 2006. Targeting of GFP to
newborn rods by Nrl promoter and temporal expression profiling
of flow-sorted photoreceptors. Proc Natl Acad Sci USA 103:3890–
3895.
Al-Hussaini H, Kam JH, Vugler A, Semo M, Jeffery G. 2008.
Mature retinal pigment epithelium cells are retained in the cell
cycle and proliferate in vivo. Mol Vis 14:1784–1791.
Alexiades MR, Cepko CL. 1997. Subsets of retinal progenitors display temporally regulated and distinct biases in the fates of their
progeny. Development 124:1119–1131.
Altshuler E, Turner DL, Cepko CL. 1991. Specification of cell type
in the vertebrate retina. In Lam M-K, Shatz CJ, editors. Development of the visual system. Cambridge: MIT Press. p 37–58.
Amato MA, Arnault E, Perron M. 2004. Retinal stem cells in vertebrates: parallels and divergence. Int J Dev Biol 48:993–1001.
Amemiya K, Haruta M, Takahashi M, Kosaka M, Eguchi G. 2004.
Adult human retinal pigment epithelial cells capable of differentiating into neurons. Biochem Biophys Res Commun 316:1–5.
Andreazzoli M, Gestri G, Cremisi F, Casarosa S, Dawid IB,
Barsacchi G. 2003. Xrx1 controls proliferation and neurogenesis
in Xenopus anterior neural plate. Development 130:5143–5154.
Aoki H, Hara A, Niwa M, Motohashi T, Suzuki T, Kunisada T.
2008. Transplantation of cells from eye-like structures differentiated from embryonic stem cells in vitro and in vivo regeneration
151
of retinal ganglion-like cells. Graefes Arch Clin Exp Ophthalmol
246:255–265.
Arnhold S, Klein H, Semkova I, Addicks K, Schraemeyer U. 2004.
Neurally selected embryonic stem cells induce tumor formation
after long-term survival following engraftment into the subretinal
space. Invest Ophthalmol Vis Sci 45:4251–4255.
Asami M, Sun G, Yamaguchi M, Kosaka M. 2007. Multipotent cells
from mammalian iris pigment epithelium. Dev Biol 304:433–446.
Barber AC, Hippert C, Duran Y, West EL, Bainbridge JWB, WarreCornish K, Luhmann UFO, Lakowski J, Sowden JC, Ali RR,
Pearson RA. 2013. Repair of the degeneration retina by photoreceptor transplantation. Proc Natl Acad Sci 110:354–359.
Beebe D, Garcia C, Wang X, Rajagopal R, Feldmeier M, Kim J-Y,
Chytil A, Moses H, Ashery-Padan R, Rauchman M. 2004. Contributions by members of the TGFbeta family to lens development.
Int J Dev Biol 48:845–856.
Beebe DC. 1986. Development of the ciliary body: a brief review.
Trans Ophthalmol Soc UK 105 (Pt 2):123–130.
Beer HD, Vindevoghel L, Gait MJ, Revest JM, Duan DR, Mason I,
Dickson C, Werner S. 2000. Fibroblast growth factor (FGF) receptor 1-IIIb is a naturally occurring functional receptor for FGFs
that is preferentially expressed in the skin and the brain. J Biol
Chem 275:16091–16097.
Belecky-Adams T, Adler R. 2001 Developmental expression patterns
of bone morphogenetic proteins, receptors, and binding proteins
in the chick retina. J Comp Neurol 430:562–572.
Belecky-Adams TL, Alder R, Beebe DC. 2002. Bone morphogenetic
protein signaling and the initiation of lens fiber differentiation.
Development 129:3795–3802.
Belliveau MJ, Cepko CL. 1999. Extrinsic and intrinsic factors control the genesis of amacrine and cone cells in the rat retina.
Development 126:555–566.
Belliveau MJ, Young TL, Cepko. 2000. Late retinal progenitor cells
show intrinsic limitations in the production of cell types and the
kinetics of opsin synthesis. J Neurosci 20:2247–2254.
Bernardos RL, Barthel, LK, Meyers JR, Raymond PA. 2007. Latestage neuronal progenitors in the retina are radial Muller glia
that function as retinal stem cells. J Neurosci 27:7028–7040.
Bertuzzi S, Hindges R, Mui SH, O’Leary DD, Lemke G. 1999. The
homeodomain protein vax 1 is required for axon guidance and
major tract formation in the developing forebrain. Genes Dev 13:
3092–3105.
Bhatia B, Jayaram H, Singhal S, Jones MF, Limb GA. 2010. Differences between the neurogenic and proliferative abilities of Muller
glia with stem cell characteristics and the ciliary epithelium from
the adult human eye. Exp Eye Res 93:852–861.
Bhatia B, Singhal S, Tadman DN, Khaw PT, Limb GA. 2011. SOX2 is
required for adult human Muller stem cell survival and maintenance of progenity in vitro. Invest Ophthalmol Vis Sci 52:136–145.
Bok D. 1993. The retinal pigment epithelium: a versatile partner of
vision. J Cell Sci Sup 17:189–195.
Bonaguidi MA, Song J, Ming G-l, Song H. 2012. A unifying hypothesis on mammalian neural stem cell properties in the adult hippocampus. Curr Opin Neurobiol 22:754–761.
Boulton M, Dayhaw-Barker P. 2001. The role of the retinal pigment
epithelium: topographical variation and aging changes. Eye 15:
384–389.
Braisted JE, Essman TF, Raymond PA. 1994. Selective regeneration
of photoreceptors in goldfish retina. Dev 120:2409–2419.
Brazelton TR, Rossi FMV, Keshet GI, Blau HM. 2000. From marrow
to brain: expression of neuronal phenotypes in adult mice. Science
290:1775–1779.
Bringmann A, Pannicke T, Grosche J, Francke M, Wiedemann P,
Skatchkov SN, Osborne NN, Reichenbach A. 2006. Muller cells in
the healthy and diseased retina. Prog Retin Eye Res 25:397–424.
Bringmann A, Iandiev I, Pannicke T, Wurm A, Hollborn M,
Wiedemann P, Osborne NN, Reichenbach A. 2009. Cellular signaling and factors involved in Muller cell gliosis: neuroprotective
and detrimental effects. Prog Retin Eye Res 28:423–451.
Brown HL, Kanekar S, Vetter ML, Tucker PK, Gemza DL, Glaser
T. 1998. Math5 encodes a murine basic helix-loop-helix
152
YIP
transcription factor expressed during early stages of retinal neurogenesis. Dev 125:4821–4833.
Brown NL, Patel S, Brzezinski J, Glaser T. 2001. Math 5 is required
for retinal ganglion cell and optic nerve formation. Development
128:2497–2508.
Buchholz DE, Hikita ST, Rowland TJ, Friedrich AM, Hinman CR,
Johnson LV, Clegg DO. 2009. Derivation of functional retinal pigmented epithelium from induced pluripotent stem cells. Stem
Cells 27:2427–2434.
Bull ND, Limb GA, Martin KR. 2008. Human Muller stem cell
(MIO-M1) transplantation in a rat model of glaucoma: survival,
differentiation, and integration. Invest Ophtalmol Vis Sci 49:
3449–3456.
Carl M, Wittbrodt J. 1999. Graded interference with FGF signaling
reveals its dorsoventral asymmetry at the mid-hindbrain boundary. Development 126:5659–5667.
Carr AJ, Vugler A, Lawrence J, Chen LL, Ahmado A, Chen FK,
Semo M, Gias C, da Cruz L, Moore HD, Walsh J, Coffey PJ.
2009a. Molecular characterization and functional analysis of
phagocytosis by human embryonic stem cell-derived RPE cells
using a novel human retinal assay. Mol Vis 15:283–295.
Carr AJ, Vugler AA, Hikita ST, Lawrence JM, Gias C, Chen LL,
Buchholz DE, Ahmado A, Semo M, Smart MJ, Hasan S, da Cruz
L, Johnson LV, Clegg DO, Coffey PJ. 2009b. Protective effects of
human iPS-derived retinal pigment epithelium cell transplantation in the retinal dystrophic rats. PLoS One 4:e8152.
Cavodeassi F, Carreira-Barbosa F, Young RM, Concha ML, Allende
ML, Houart C, Tada M, Wilson SW. 2005. Early stages of zebrafish eye formation require the coordinated activity of Wnt11, Fz5,
and Wnt/beta-catenin pathway. Neuron 47:43–56.
Cayouette M, Poggi L, Harris WA. 2006. Lineage in the vertebrate
retina. Trends Neurosci 29:563–570.
Cayouette M, Whitmore AV, Jeffery G, Raff M. 2001. Asymmetric
segregation of Numb in retinal development and the influence of
the pigmented epithelium. J Neurosci 21:5643–5651.
Cepko CL. 1999. The roles of intrinsic and extrinsic cues and bHLH
genes in the determination of retinal cell fates. Curr Opin Neurobiol 9:37–46.
Cepko CL, Austin CP, Yang X, Alexiades M, Ezzeddine D. 1996.
Cell fate determination in the vertebrate retina. Proc Natl Acad
Sci USA 93:589–595.
Chacko DM, Rogers JA, Turner JE, Ahmad I. 2000. Survival and
differentiation of cultured retinal progenitors transplanted in the
subretinal space of the rat. Biochem Biophys Res Commun 268:
842–846.
Chadashvili T, Peterson DA. 2006. Cytoarchitecture of fibroblast
growth factor receptor 2 (FGFR-2) immunoreactivity in astrocytes
of neurogenic and non-Neurogenic regions of the young adult and
aged Rat Brain. Brain 15:1–15.
Chen J, Rattner A, Nathans J. 2005. The rod photoreceptor-specific
nuclear receptor Nr2e3 represses transcription of multiple conespecific genes. J Neurosci 25:118–129.
Chen S, Samuel W, Fariss RN, Duncan T, Kutty RK, Wiggert B.
2003. Differentiation of human retinal pigment epithelial cells
into neuronal phenotypes by N-(4-hydroxyphenyl) retinamide. J
Neurochem 84:972–981.
Chen S, Wang QL, Nie Z, Sun H, Lennon G, Copeland NG, Gilbert
DJ, Jenkins NA, Zack DJ. 1997. Crx, a novel Otx-like pairedhomeodomain protein, binds to and transactivates photoreceptor
cell-specific genes. Neuron 19:1017–1030.
Cheng H, Aleman TS, Cideciyan AV, Khanna R, Jacobson SG,
Swaroop A. 2006. In vivo function of the orphan nuclear receptor
NR2E3 in establishing photoreceptor identity during mammalian
retinal development. Hum Mol Genet 15:2588–2602.
Chiang C, Litingtung Y, Lee E, Young KE, Corden JL, Westphal H,
Beachy PA. 1996. Cyclopia and defective axial patterning in mice
lacking Sonic hedgehog gene function. Nature 383:407–413.
Cho SH, Cepko CL. 2006. Wnt2b/beta-catenin-mediated canonical
Wnt signaling determines the peripheral fates of the chick eye.
Development 133:3167–3177.
Chuang JC, Raymond PA. 2002. Embryonic origin of the eyes in teleost fish. Bioessays 24:519–529.
Cicero SA, Johnson D, Reyntjens S, Frase S, Connell S, Chow LM,
Baker SJ, Sorrentino BP, Dyer MA. 2009. Cells previously identified as retinal stem cells are pigmented ciliary epithelial cells.
Proc Natl Acad Sci USA 106:6685–6690.
Close JL, Liu J, Gumuscu B, Reh TA. 2006. Epidermal growth factor receptor expression regulates proliferation in the postnatal rat
retina. Glia 54:94–104.
Coles BL K, Ang
enieux B, Inoue T, Del Rio-Tsonis K, Spence JR,
McInnes RR, Arsenijevic Y, van der Kooy D. 2004. Facile isolation
and the characterization of human retinal stem cells. PNAS 101:
15772–15777.
Costa MR, G€
otz M, Berninger B. 2010. What determines neurogenic
competence in glia? Brain Res Rev 63:47–59.
Cowan CA, Klimanskaya I, McMahon J, Atienza BS, Witmyer J,
Zucher JP, Wang S, Morton C, McMahon AP, Powers D, Melton
DA. 2004. Derivation of embryonic stem-cell lines from human
blastocysts. New Eng J Med 350:1353–1356.
Crigler L, Robey RC, Asawachaicharn A, Gaupp D, Phinney DG.
2006. Human mesenchymal stem cell subpopulations express a
variety of neuroregulatory molecules and promote neuronal cell
survival and neurogenesis. Exp Neurol 198:54–64.
Crossley, PH, Martinez S, Ohkubo Y, Rubenstein JLR. 2001. Coordinate expression of fgf8, otx2, bmp4 and shh in the rostral prosencephalon during development of the telencephalic and optic
vesicles. Neuroscience 108:183–206.
Dakubo GD, Wang YP, Mazerolle C, Campsall K, McMahon AP,
Wallace VA. 2003. Retinal ganglion cell-derived sonic hedgehog
signaling is required for optic disc and stalk neuroepithelial cell
development. Development 130:2967–2980.
Das AV, Mallya KB, Zhao X, Ahmad F, Bhattacharya S, Thoreson
WB, Hegde GV, Ahmad I. 2006. Neural stem cell properties of
Muller glia in the mammalian retina: regulation by Notch and
Wnt signaling. Dev Biol 299:283–302.
Das T, Payer B, Cayouette M, Harris WA. 2003. In vivo time-lapse
imaging of cell divisions during neurogenesis in the developing
zebrafish retina. Neuron 37:597–609.
Davis AA, Matzuk MM, Reh TA. 2000. Activin A promotes progenitors differentiation into photoreceptors in rodent retina. Mol Cell
Neurosci 15:11–21.
De A. 2011. Wnt /Ca21 signaling pathway: a brief overview. Acta
Biochim Biophy Sin 43:745–756.
De Marzo A, Aruta C, Marigo V. 2010. PEDF promotes retinal
neurosphere formation and expansion in vitro in: Retinal
Degenerative Diseases (Anderson RE, ed.). Adv Exp Med Biol
664:621–630.
Del Rio-Tsonis K, Tsonis PA. 2003. Eye regeneration at the molecular age. Dev Dyn 226:211–224.
Deneen B, Ho R, Lukaszewicz A, Hochstim CJ, Gronostajski RM,
Anderson DJ. 2006. The transcription factor NF1A controls the onset
of gliogenesis in the developing spinal cord. Neuron 52:953–968.
Dias da Silva MR, Tiffin N, Mima T, Mikawa T, Hyer J. 2007. FGFmediated induction of ciliary body tissue in the chick eye. Dev
Biol 304:272–285.
Dorsky RI, Rapaport DH, Harris WA. 1995. Xotch inhibits cell differentiation in the Xenopus retina. Neuron 14:487–496.
Du Y, Xiao Q, Yip HK. 2010. Regulation of retinal ganglion progenitor cell differentiation by bone morphogenetic protein 4 is mediated by the Smad/Id cascade. Invest Opthalmol Vis Sci 51:3764–
3773.
Du Y, Yip HK. 2010. Effects of bone morphogeneitic protein 2 on Id
expression and neuroblastoma cell differentiation. Differentiation
79:84–92.
Du Y, Yip HK. 2011. The expression and roles of inhibitor of DNA
binding helix-loop-helix proteins in the developing and adult
mouse retina. Neuroscience 175:367–379.
Dudley AT, Lyons KM, Robertson EJ. 1995. A requirement for bone
morphogenetic protein-7 during development of the mammalian
kidney and eye. Genes Dev 9:2795–2807.
Dudley AT, Robertson EJ. 1997. Overlapping expression domains of
bone morphogenetics protein family members potentially account
for limited tissue defects in BMP7 deficient embryos. Dev Dyn
208:349–362.
RSC AND REGENERATION OF VISION SYSTEM
Dunker N, Schuster N, Krieglstein K. 2001. TGF-beta modulates
programmed cell death in the retina of the developing chick
embryo. Development 128:1933–1942.
Dutt K, Cao Y. 2009. Engineering retina from human retinal progenitors (cell lines). Tissue Eng Part A 15:1401–1413.
Dyer MA, Cepko CL. 2000. Control of Muller glial cell proliferation
and activation following retinal injury. Nat Neurosci 3:873–880.
Dyer MA, Livesey FJ, Cepko CL, Oliver G. 2003. Prox1 function
controls progenitor cell proliferation and horizontal cell genesis in
the mammalian retina. Nat Genet 34:53–58.
Easter SS, Hitchcock PF. 2000. Stem cells and regeneration in the
retina. What fish have taught us about neurogenesis. Neuroscientist 6:454–464.
Eglitis MA, Mezey E. 1997. Hematopoietic cells differentiate into
both microglia and microglia in the brains of adult mice. Proc
Natl Acad Sci USA 94:4080–4085.
Eguchi G. 1971. Celllular analysis on localization of lens forming
potency in the newt iris epithelium. Dev Growth Differ 13:337–
349.
Eguchi G. 1986. Instability in cell commitment of vertebrate pigmented epithelial cells and their transdifferentiation into lens
cells. Curr Top Dev Biol 20:21–37.
Eiraku M, Takata N, Ishibashi H, Kawada M, Sakajura E,
Okuda S, Sekiguchi K, Adachi T, Sasai Y. 2011. Self-organizing
optic-cup morphogenesis in three-dimensional culture. Nature
472:51–56.
Eiraku M, Sasai Y. 2012a. Mouse embryonic stem cell culture for
generation of three-dimensional retinal and cortical tissues. Nat
Protoc 7:69–79.
Eiraku M, Sasai Y. 2012b. Self-formation of layered neural structures in three-dimensional culture of ES cells. Curr Opini Neurobiol 22:768–777.
Engelhardt M, Bogdahn U, Aigner L. 2005. Adult retinal pigment
epithelium cells express neural progenitor properties and the neuronal precursors protein doublecortin. Brain Res 1040:98–111.
Esteve P, Lopez-Rios J, Bovolenta P. 2004. SFRP1 is required for
the proper establishment of the eye field in the medaka fish.
Mech Dev 121:687–701.
Faber SC, Robinson ML, Makarenkova HP, Lang RA. 2002. Bmp
signaling is required for development of primary lens fiber cells.
Development 129:3727–3737.
Fadool JM. 2003. Rod genesis in the teleost retina as a model of
neural stem cells. Exp Neurol 184:14–27.
Fausett BV, Goldman D. A role for alpha1 tubulin-expressing Muller glia in regeneration of the injured zebrafish retina. 2006. J
Neurosci 26:6303–6313.
Fekete D, Perez-Miguelsanz J, Ryder E, Cepko C. 1994. Clonal
analysis in the chicken retina reveals tangential dispersion of clonally related cells. Dev Biol 166:666–682.
Ffrench-Constant C, Raff MC. 1986. Proliferating bipotential glial
progenitor cells in the adult rat optic nerve. Nature 319:499–502.
Fimbel SM, Montgomery JE, Burket CT, Hyde DR. 2007. Regeneration of inner retinal neurons after intravitreal injection of quabain in zebrafish. J Neurosci 27:1712–1724.
Fischer AJ, Omar G. 2005. Transitin, a nestin-related intermediate
filament, is expressed by neural progenitors and can be induced
in Muller glia in the chicken retina. J Comp Neurol 484:1–14.
Fischer AJ, Reh TA. 2000. Identification of a proliferating marginal zone
of retinal progenitors in postnatal chickens. Dev Biol 220:197–210.
Fischer AJ, Reh TA. 2001a. Muller glia are a potential source of
neural regeneration in the postnatal chicken retina. Nat Neurosci
4:247–252.
Fischer AJ, Reh TA. 2001b. Trandifferentiation of pigmented epithelial cells: a source of retinal stem cells? Dev Neurosci 23:268–276.
Fischer AJ, Reh TA. 2002. Exogenous growth factors stimulate the
regeneration of ganglion cells in the chicken retina. Dev Biol 251:
367–379.
Fishcer AJ, Reh TA. 2003. Growth factors induce neurogenesis in
the ciliary body. Dev Biol 259:225–240.
Ford-Perriss M, Abud H, Murphy M. 2001. Brief review: fibroblast
growth factors in the developing central nervous system. Clin
Exp Pharma Physiol 28:493–503.
153
Fortin D, Rom E, Sun H, Yayon A, Bansal R. 2005. Distinct fibroblast growth factor (FGF)/FGF receptor signaling pairs initiate
diverse cellular responses in the oligodendrocyte lineage. J Neurosci 25:7470–7479.
Franke A, Gubbe C, Beier M, Duenker N. 2006. Transforming
growth factor-beta and bone morphogenetic proteins: cooperative
players in chick and murine programmed retinal cell death. J
Comp Neurol 495:263–278.
Fraser B, DuVal MG, Wang H, Allison WT. 2013. Regeneration of
cone photoreceptors when cell ablation is primarily restricted to a
particular cone. PLoS ONE 8:e55410. DOI: 10.1371/journal.
pone.0055410.
Froen RC, Johnsen EO, Petrovski G, Berenyi E, Facsko A, Berta, A.
Nicolaissen B, Moe MC. 2011. Pigment epithelial cells isolated
from human peripheral iridectomies have limited properties of
retinal stem cells. Acta Ophthalmol 89:e635–e644.
Fu X, Sun H, Klein WH, Mu X. 2006. Beta-catenin is essential for
lamination but not neurogenesis in mouse retinal development.
Dev Biol 299:424–437.
Furukawa T, Morrow EM, Li T, Davis FC, Cepko CL. 1999. Retinopathy and attenuated circadian entrainment in Crx-deficient mice.
Nat Genet 23:466–470.
Fuhrmann S, Levine EM, Reh TA. 2000. Extraoculoar mesenchyme
patterns the optic vesicle during early eye development in the
embryonic chick. Development 127:4599–4609.
Fujitani Y, Fujitani S, Luo H, Qiu F, Burlison J, Long Q,
Kawaguchi Y, Edlund H, MacDonald RJ, Furukawa T, Fujikado
T, Magnuson MA, Xiang M, Wright CV. 2006. Ptf1a determines
horizontal and amacrine cell fates during mouse retinal development. Development 133:4439–4450.
Furukawa T, Kozak CA, Cepko CL. 1997a. Rax, a novel paired-type
homeobox gene, shows expression in the anterior neural fold and
developing retina. Proc Natl Acad Sci USA 94:3088–3093.
Furukawa T, Morrow EM, Cepko CL. 1997b. Crx, a novel otx-like
homeobox gene, shows photoreceptor-specific expression and regulates photoreceptor differentiation. Cell 91:531–541.
Furukawa T, Mukherjee S, Bao, ZZ, Morrow EM, Cepko CL. 2000.
Rax, Hes1, and notch1 promote the formation of M€
uller glia by
postnatal retinal progenitor cells. Neuron 26:383–394.
Furuta Y, Hogan BL. 1998. BMP4 is essential for lens induction in
the mouse embryo. Genes Dev 12:3764–3775.
Gaiano N, Nye JS, Fishell G. 2000. Radial glial identity is
promoted by Notch1 signaling in the murine forebrain. Neuron
26:395–404.
Galli-Resta L. 2002. Putting neurons in the right place: local interactions in the genesis of retinal architecture. Trends Neurosci 25:
638–643.
Galy A, Neron B, Planque N, Saule S, Eychene A. 2002. Activated
MAPK/ERK kinase (MEK-1) induces transdifferentiation of pigmented epithelium into neural retina.l Dev Biol 248:251–264.
Giannelli SG, Demontis GC, Pertile G, Rama P, Broccoli V. 2011.
Adult human M€
uller glia cells are a highly efficient source of rod
photoreceptors. Stem Cells 29:344–356.
Golden JA, Bracilovic A, McFadden KA, Beesley JS, RR JL,
Grinspan JB. 1999. Ectopic bone morphogenetic proteins 5 and 4
in the chicken forebrain lead to cyclopia and holoprosencephaly.
Proc Natl Acad Sci USA 96:2439–2444.
Gould E, Reeves AJ, Graziano MS, Gross CG. 1999. Neurogenesis
in the neocortex of adult primates. Science 286:548–552.
Grindley JC, Davidson DR, Hill RE. 1995. The role of Pax-6 in eye
and nasal development. Development 121:1433–1442.
Gualdoni, S, Baron M, Lakowski J, Decembrini S, Smith AJ,
Pearson RA, Ali RR, Sowden JC. 2010. Adult ciliary epithelial
cells, previously identified as retinal stem cells with potential for
retinal repair, fail to differentiate into new rod photoreceptors.
Stem Cells 28: 1048–1059.
Guduric-Fuchs J, Chen W, Price H, Archer DB, Cogliate T. 2011
RPE and neuronal differentiation of allotransplantated porcine
ciliary epithelium-derived cells. Mol Vis 17:2580–2595.
Hallonet M, Hollemann T, Pieler T, Gruss P. 1999. Vax 1, a novel
homeobox-containing gene, directs development of the basal forebrain and visual system. Genes Dev 13:3106–3114.
154
YIP
Hara A, Aoli H, Takamatsu M, Hatano Y, Tomita H, Kuno T, Niwa
M, Kunisada T. 2012. Human embryonic stem cells transplanted
into mouse retina induce neural differentiation. Stem Cells Cancer Stem Cells 2:291–298.
Harris JR, Fisher R, Jorgensen M, Kaushal S, Scott EW. 2009.
CD133 progenitor cells from bone marrow contribute to retinal
pigment epithelium repair. Stem Cells 27:457–466.
Harris WA. 1997. Cellular diversification in the vertebrate retina.
Curr Opin Genet Dev 7:651–658.
Haruta M, Kosaka M, Kanegae Y, Saito I, Inoue T, Kageyama R,
Nishda A, Honda Y, Takahashi M. 2001 Induction of
photoreceptor-specific phenotypes in adult mammalian iris tissue.
Nat Neurosci 4:1163–1164.
Haruta M, Sasai Y, Kawasaki H. 2004. In vitro and in vivo characterization of pigement epithelial cells differentiated from primate
embryonic stem cells. Invest Ophthalmol Vis Sci 45:1020–1025.
Hatakeyama J, Tomita K, Inoue T, Kageyama R. 2001. Roles of
homeobox and bHLH genes in specification of a retinal cell type.
Development 128:1313–1322.
Hendrickson A, Hicks D. 2002. Distribution and density of mediumand short-wavelength selective cones in the domestic pig retina.
Exp Eye Res 74:435–444.
Hicks D, Courtois Y. 1992. Fibroblast growth factor stimulates photoreceptor differentiation in vitro. J Neurosci 12:2022–2033.
Hill RE, Hanson IM. 1992. Molecular genetics of the Pax gene family. Curr Opin Cell Biol 4:967–972.
Hirami Y, Osakada F, Takahashi K, Okita K, Yamanaka S, Ikeda H,
Yoshimura N, Takahashi M. 2009. Generation of retinal cells
from mouse and human induced pluripotent stem cells. Neurosci
Lett 458:126–131.
Hirano M, Yamamato A, Yoshimura N, Tokunaga T, Motohashi T,
Ishizaki K, Yoshida H, Okazaki K, Yamazaki H, Hayashi S-I,
Kunisada T. 2003. Generation of structures formed by lens and
retinal cells differentiating from embryonic stem cells. Dev Dyn
228:664–671.
Hitchcock PF, Macdonald RE, VanDeRyt JT, Wilson SW. 1996. Antibodies against Pax6 immunostain amacrine and ganglion cells
and neuronal progenitors, but not rod precursors, in the normal
and regenerating retina of the goldfish. J Neurobiol 29:399–413.
Hitchcock P, Ochocinska M, Sieh A, Otteson D. 2004. Persistent
and injury-induced neurogenesis in the vertebrate retina. Prog
Retin Eye Res 23:183–194.
Hitchcock P, Raymond PA. 2004. The zebrafish retina as a model
for developmental and regeneration biology. Zebrafish 1:257–271.
Hogan PG, Chen L, Nardone J, Rao A. 2003. Transcriptional regulation by calcium, calcineurin, and NFATv. Genes Dev 17:2205–
2232.
Hojo M, Ohtsuka T, Hashimoto N, Gradwohl G, Guillemot F,
Kageyama R. 2000. Glial cell fate specification modulated by the
bHLH gene Hes5 in mouse retina. Development 127:2515–2522.
Holt CE, Bertsch TW, Ellis HM, Harris WA. 1988. Cellular determination in the Xenopus retina is independent of lineage and birth
date. Neuron 1:15–26.
Hongisto H, Mikhailova A, Hiidenmaa H, IImarinen T, Skottman
H. 2012. Low levels of activin A secreted by fibroblast feeder cells
accelerates early stage differentiation of retinal pigment epithelial
cells from human pluripotent stem cells. Stem Cell Dis 2:176–
186.
Houart C, Caneparo L, Heisenberg C, Barth K, Take-Uchi M,
Wilson S. 2002. Establishment of the telencephalon during gastrulation by local antagonist of Wnt signaling. Neuron 35:255–
265.
Hu M, Easter SS. 1999. Retinal neurogenesis: the formation of the
initial central patch of postmitotic cells. Dev Biol 207:309–321.
Huelsken J, Behrens J. 2002. The Wnt signaling pathway. J Cell
Sci 115:3977–3978.
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.
Hyer J, Kuhlman J, Afif E, Mikawa T. 2003. Optic cup morphogenesis requires per-lens ectoderm but not lens differentiation. Dev
Biol 259:351–363.
Idelson M, Alper R, Obolensky A, Ben-Shusham E, Hemo I,
Yachimovich-Cohen N, Khaner H, Smith Y, Wiser, O, Gropp M,
Cohen MA, Even-Ram S, Berman-Zaken Y, Matzrafi L, Rechavi
G, Banin E, Reubinoff B. 2009. Directed differentiation of human
embryonic stem cells into functional retinal pigment epithelium
cells. Cell Stem Cells 5:396–408.
Ikegami Y, Mitsuda S, Araki M. 2002. Neural cell differentiation
from retinal pigment epithelial cells of the newt: an organ culture
model for the urodele retinal regeneration. J Neurobiol 50:209–
220.
Ingham PW, McMahon AP. 2001. Hedgehog signaling in animal
development: paradigms and principles. Genes Dev 15:3059–3087.
Inoue T, Hojo M, Bessho Y, Tano Y, Lee JE, Kageyama R. 2002.
Math3 and NeuroD regulate amacrine cell fate specification in
the retina. Development 129:831–842.
Inoue Y, Iriyama A, Ueno S, Takahashi H, Kondo M, Tamaki Y,
Araie M, Yanagi Y. 2007. Subretinal transplantation of bone
marrow mesenchymal stem cells delays retinal degeneration in
the RCS rat model of retinal degeneration. Exp Eye Res 85:
234–241.
Inoue T, Kagawa T, Fukushima M, Shimizu T, Yoshinaga Y, Takada
S, Tanihara H, Taga T. 2006. Activation of canonical Wnt pathway promotes proliferation of retinal stem cells derived from
adult mouse ciliary margin. Stem Cells 24:95–104.
Ishibashi M, Moriyoshi K, Sasai Y, Shiota K, Nakanishi, S,
Kageyama R. 1994. Persistent expression of helix-loop-helix factor
HES-1 prevents mammalian neural differentiation in the central
nervous system. EMBO J 13:1799–1805.
Jadhav AP, Cho S-H, Cepko CL. 2006. Notch activity permits retinal cells to progress through multiple progenitor states and
acquire a stem cell property. Proc Natl Acad Sci USA 103:18998–
19003.
Jaffe GJ, Harrison CE, Lui GM, Roberts WL, Goldsmith PC,
Mesiano S, Jaffe RB. 1994. Activin expression by cultured human
retinal pigment epithelial cells. Invest Opthalmol Vis Sci 35:
2924–2931.
Jan YN, Jan LY. 1993. HLH proteins, fly neurogenesis, and vertebrate myogenesis. Cell 75:827–830.
Jarman AP, Sun Y, Jan LY, Jan YN. 1995. Role of the proneural
gene, atonal, in formation of Drosophila chordotonal organs and
photoreceptors. Development 121:2019–2030.
Jasoni CL, Reh TA. 1996. Temporal and spatial pattern of MASH-1
expression in the developing rat retina demonstrates progenitor
cell heterogeneity. J Comp Neurol 369:319–327.
Jena N, Martin-Seisdedos C, McCue P, Croce CM. 1997. BMP7 null
mutation in mice: developmental defects in skeleton, kidney, and
eye. Exp Cell Res 230:28–37.
Jensen AM, Wallace VA. 1997. Expression of Sonic hedgehog and its
putative role as a precursor cell mitogen in the developing mouse
retina. Dev 124:363–371.
Jensen AM, Walker C, Westerfield M. 2001. Mosaic eyes: a
zebrafish gene required in pigemented epithelium for apical localization of retinal cell division and lamination. Development 128:
95–105.
Jin ZB, Okamoto S, Mandai M, Takahashi M. 2009. Induced pluripotent stem cells for retinal degenerative diseases: a new perspective on the challenges. J Genet 88:417–424.
Johns PR. 1977. Growth of the adult goldfish eye. III. Sources of
the new retinal cells. J Comp Neurol 176:343–357.
Kanekar S, Perron M, Dorsky R, Harris WA, Jan LY, Jan YN,
Vetter ML. 1997. Xath5 participates in a network of bHLH genes
in the developing Xenopus retina. Neuron 19, 981–994.
Karl MO, Hayes S, Nelson BR, Tan K, Buckingham B, Reh TA.
2008. Stimulation of neural regeneration in the mouse retina.
Proc Natl Acad Sci USA 105:19508–19513.
Keefe JR. 1973. An analysis of urodelian retinal regeneration. J
Exp Zool 184:185–206.
Kerever A, Schnack J, Vellinga D, Ichikawa N, Moon C, ArikawaHirasawa E, Efird JT, Mercier F. 2007. Novel extracellular matrix
structures in the neural stem cell niche capture the neurogenic
factor fibroblast growth factor 2 from the extracellular milieu.
Stem cells 25:2146–2157.
RSC AND REGENERATION OF VISION SYSTEM
Kernie SG, Erwin TM, Parada LFL. 2001. Brain remodeling due to
neuronal and astrocytic proliferation after controlled cortical
injury in mice. J Neurosci Res 66:317–326.
Kicic A, Shen WY, Wilson AS, Constable IJ, Robertson T, Rakoczy
PE. 2003. Differentiation of marrow stromal cells into photoreceptors in the rat eye. J Neurosci 23:7742–7749.
Kim CH, Oda T, Itoh M, Jiang D, Artinger KB, Chandrasekharappa
SC, Driever W, Chitnis AB.2000. Repressor activity of Headless/
Tcf3 is essential for vertebrate head formation. Nature 407:913–
916.
Kim D, Kim CH, Moon JI, Chung YG, Chang MY, Han BS, Ko S,
Yang E, Cha KY, Lanza R, Kim KS. 2009. Generation of human
induced pluripotent stem cells by direct delivery of reprogramming proteins. Cell Stem Cell 4:472–476.
Kim IB, Kim KY, Joo JK, Lee MY, Oh SJ, Chung JW, Chun MH.
1998. Reaction of Muller cells after increased intraocular pressure
in the rat retina. Exp Brain Res 121:419–424.
Kinouchi R, Takeda M, Yang L, Wilhelmsson U, Lundkvist A,
Pekny M, Chen DF. 2003. Robust neural integration from retinal
transplants in mice deficient in GFAP and vimentin. Nat Neurosci 6:863–868.
Klassen H, Kiilgaard JF, Zahir T, Ziaeian B, Kirov I, Scherfig E,
Warfvinge K, Young MJ. 2007. Progenitor cells from the porcine
neural retina express photoreceptor markers after transplantation to the subretinal space of allorecipients. Stem Cells 25:1222–
1230.
Klassen H, Sakaguchi DS, Young MJ. 2004a. Stem cells and retinal
repair. Prog Retin Eye Res 23:149–181.
Klassen HJ, Ng TF, KurimotoY, Kirov I, Shatos M, Coffey P,
Young MJ. 2004b. Multipotent retinal progenitors express development markers, differentiate into retinal neurons, and preserve
might-mediated behavior. Invest Ophthalmol Vis Sci 45:4167–
4173.
Kohno H, Sakai T, Kitahara K. 2006. Induction of nestin, Ki-67,
and cyclin D1 expression in Muller cells after laser injury in adult
rat retina. Graefes Arch Clin Exp ophthalmol 244:90–95.
Kopen GC, Prockop DJ, Phinney DG. 1999. Marrow stromal cells
migrate throughout forebrain and cerebellum, and they differentiate into astrocytes after injection into neonatal mouse brain. Proc
Natl Acad Sci USA 96:10711–10716.
Kosaka, M, Sun GW, Haruta M, Takahashi M. 2004. Multipotentiality of iris pigmented epitherlial cells in vertebrate eye. In: Turksen K, editor. Adult Stem Cells. Human Press: Totowa, NJ. p
253–268.
Kreslova J, Machon O, Ruzickova J, Lachova J, Wawrousek EF,
Kemler R, Krauss S, Piatigorsky J, Kozmik Z. 2007. Abnormal
lens morphogenesis and ectopic lens formantion in the absence of
the beta-catenin function. Genesis 45:157–168.
Kubota R, Hokoc JN, Moshiri A, McGuire C, Reh TA. 2002. A comparative study of neurogenesis in the retinal ciliary margin zone
of homeothermic vertebrate. Brain Res Dev Brain Res 134:31–41.
K€
uhl M, Sheldahl LC, Malbon CC, Moon RT. 2000. Ca21/calmodulin-dependent protein kinase II is stimulated by Wnt and Frizzled
homologs and promotes ventral cell fates in Xenopus. J Biol
Chem 275:12701–12711.
Ladher RK, Church VL, Allen S, Robson L, Abdelfattah A, Brown
NA, Hattersley G, Rosen V, Luyten FP, Dale L, Francis-West PH.
2000. Cloning and expression of the Wnt antagonist Sfrp-2 and
Frzb during chick development. Dev Biol 218:183–198.
Lagutin OV, Zhu CC, Kobayashi D, Topczewski J, Shimamura K,
Puelles L, Russell HR, McKinnon PJ, Solnica-Krezel L, Oliver G.
2003. Six3 repression of Wnt signaling in the anterior neuroectoderm is essential for vertebrate forebrain development. Genes
Dev 17:368–379.
Lamba DA, Karl MO, Reh TA. 2009a. Strategies for retinal repair:
cell replacement and regeneration. Prog Brain Res 175:23–31.
Lamba DA, Gust J, Reh TA. 2009b. Transplantation of human
embryonic stem cell-derived photoreceptors restore some visual
function in Crx-deficient mice. Cell Stem Cell 4:73–79.
Lamba DA, Karl MO, Ware DB, Reh TA. 2006. Efficient generation
of retinal progenitor cells from human embryonic stem cells. Proc
Natl Acad Sci USA 103:12769–12774.
155
Lang RA. 2004. Pathways regulating lens induction in the mouse.
Int J Dev Biol 48:783–791.
Layer PG, Rothermel A, Willbold E. 1998. Inductive effects of the
retina pigmented epithelium RPE on histogenesis of the avian
retina as revealed by retinoapheroid technology. Sem Cell Dev
Biol 9:257–262.
Lee HS, Bong YS, Moore KB, Soria K, Moody SA, Daar IO. 2006.
Dishevelled mediates ephrinB1 signaling in the eye field through
the planar cell polarity pathway. Nat Cell Biol 8:55–63.
Lee HY, Wroblewski E, Philips GT, Stair CN, Conley K, Reedy M,
Mastick GS, Brown HL. 2005. Multiple requirements for Hes1
during early eye formation. Dev Biol 284:464–478.
Levine EM, Hitchcock PF, Glasgow E, Schechter N. 1994. Restricted
expression of a new paired-class homeobox gene in normal and
regenerating adult goldfish retina. J Comp Neurol 348:596–606.
Levine EM, Roelink H, Turner J, Reh TA. 1997. Sonic hedgehog
promotes photoreceptor differentiation in mammalian retinal cells
in vitro. J Neurosci 17:6277–6288.
Levkovitch-Verbin H, Sadan O, Vander S, Rosner M, Barhum Y,
Melamed E, Offen D, Melamed S. 2010. Intravitreal injections of
neurotrophic factors secreting mesenchymal stem cells are neuroprotective in rat eyes following optic nerve transaction. Invest
Ophthalmol Vis Sci 51:6394–6400.
Lewis GP, Fisher SK. 2003. Up-regulation of glial fibrillary acidic
protein in response to retinal injury: its potential role in glial
remodeling and a comparison to vimentin expression. Int Rev
Cytol 230:263–290.
Li H, Tierney C, Wen L, Wu JY, Rao Y. 1997. A single morphogenetic field gives rise to two retina primordial under the influence
of the prechordal plate. Development 124:603–615.
Li S, Mo Z, Yang X, Price SM, Shen MM, Xiang M. 2004. Foxn4
controls the genesis of amacrine and horizontal cells by retinal
progenitors. Neuron 43:795–807.
Li Y, Atmaca-Sonmez P, Schanie CL, Ildstad ST, Kaplan HJ,
Enzmann V. 2007. Endogenous bone marrow derived cells
express retinal pigment epithelium cell markers and migrate to
focal areas of RPE damage. Invest Ophthalmol Vis Sci 48:4321–
4327.
Li W, Zhou HY, Abujarour R, Zhu S, Joo JY, Lin T, Hao E, Sch€
oler
HR, Hayek A, Ding S. 2009. Generation of human-induced pluripotent stem cells in the absence of exogenous Sox2. Stem Cells
27:2992–3000.
Lillien L. 1998. Neural progenitors and stem cells: mechanisms of
progenitor heterogeneity. Curr Opin Neurobiol 8:37–44.
Lillien L, Cepko C. 1992. Control of proliferation in the retina: temporal changes in responsiveness of FGF and TGFa. Development
115:253–266.
Limb GA, Lawrence JM, Reh TA, Khaw PT. 2005. Localization of
Muller glia with neural stem cell characteristics in the adult
human retina. Invest Ophthalmol Vis Sci 43:864–869.
Liu H, Xu S, Wang Y, Mazerolle C, Thurig S, Coles BL, Ren JC,
Taketo MM, van der Kooy D, Wallace VA. 2007. Ciliary margin
transdifferentiation from neural retina is controlled by canonical
Wnt signaling. Dev Biol 308:54–67.
Liu J, Wilson S, Reh T. 2003. Bmp receptor 1b is required for axon
guidance and cell survival in the developing retina. Dev Biol 256:
34–48.
Liu W, Mo Z, Xiang M. 2001. The Ath5 proneural genes function
upstream of Brn 3 POU domain transcription factor genes to promote retinal ganglion cell development. Proc Natl Acad Sci USA
98:1649–1654.
Liu Y, Gao L, Zuba-Surma EK, Peng X, Kucia M, Ratajczak MZ,
Wang W, Enzman V, Kaplan HJ, Dean DC. 2009. Identification of
small Sca-11, Lin-, CD45- multipotential cells in the neonatal
murine retina. Exp Hematol 37:1096–1107.
Livesey FJ, Cepko CL. 2001. Vertebrate neural cell-fate determination: lessons from the retina. Nat Rev Neurosci 2:109–118.
Livesey FJ, Young TL, Cepko CLL. 2004. An analysis of the gene
expression program of mammalian neural progenitor cells. Proc
Natl Acad Sci USA 101:1374–1379.
Lobov IB, Rao S, Carroll TJ, Vallance JE, Ito M, Ondr JK, Kurup S,
Glass DA, Patel MS, Shu W, Morrisey EE, McMahon AP,
156
YIP
Karsenty G, Lang RA. 2005. WNT7b mediates macrophageinduced programmed cell death in patterning of the vasculature.
Nature 437:417–421.
Lord-Grignon J, Abdouh M, Bernier G. 2006. Identification of genes
expressed in retinal progenitor/stem cell colonies isolated from
the ocular ciliary body of adult mice. Gene Expr Patterns 6:992–
999.
Lund RD, Wang S, Klimanskaya I, Holmes T, Ramos-Kelesey R, Lu
B, Girman S, Bischoff N, Sauve Y, Lanza R. 2006. Human embryonic stem cell-derived cells rescue visual function in dystrophic
RCS rats. Cloning Stem Cells 8:189–199.
Luo G, Hofmann C, Broncker AL, Sohocki M, Bradley A, Karsenty
G. 1995. BMP-7 is an inducer of nephrogenesis, and is also
required for eye development ad skeletal patterning. Genes Dev
9:2808–2820.
Ma J, Kabiel M, Tucker BA, Ge J, Young MJ. 2011. Combining
chondroitinase ABC and growth factors promotes the integration
of murine retinal progenitor cells transplanted into Rho (-/-) mice.
Mol Vis 17:1759–1770.
MacDonald R, Barth KA, Xu Q, Holder N, Mikkola I, Wilson SW.
1995. Midline signaling is required for Pax gene regulation and
patterning of the eyes. Development 121:3267–3278.
MacLaren RE, Pearson RA, MacNeil A, Douglas RH, Salt TE,
Akimoto M, Swaroop A, Sowden JC, Ali RR. 2006. Retinal repair
by transplantation of photoreceptor precursors. Nature 444:203–
207.
MacNeil A, Pearson RA, MacLaren RE, Smith AJ, Sowden JC, Ali
RR. 2007. Comparative analysis of progenitor cells isolated from
the iris, pars plana, and ciliary body of the adult porcine eye.
Stem Cells 25:2430–2438.
Magavi SS, Leavitt BR, Macklis JD. 2000. Induction of neurogenesis in the neocortex of adult mice. Nature 405:951–955.
Marchetti V, Krohne TU, Friedlander DF, Friedlander M. 2010.
Stemming vision loss with stem cells. J Clin Invest 120:3012–
3021.
Maretto, S, Cordenonsi M, Dupont S, Braghetta P, Broccoli V,
Hassan AB, Volpin D, Bressan GM, Piccolo S. 2003. Mapping
Wnt/beta-catenin signaling during mouse development and in
colorectal tumors. Proc Natl Acad Sci USA 100:3299–3304.
Marquardt T, Gruss PL. 2002. Generating neuronal diversity in the
retina: one for nearly all. Trends Neurosci 25:32–38.
Martinez-Morales JR, Signore M, Acampora D, Simeone A,
Bovolenta P. 2001. Otx genes are required for tissue specification
in the developing eye. Development 128:2019–2030.
Masckauchan TN, Kitajewski J. 2007. Wnt pathways in angiogenesis. Adv Dev Biol 17:224–238.
Masland RH. 2001. The fundamental plan of the retina. Nat Neurosci 4:877–886.
Massague J, Kelly B. 1986. Internalization of transforming growth
factor-beta and its receptor in BALB/c 3T3 fibroblasts. J Cell
Physiol 128:216–222.
Mathers PH, Grinberg A, Mahon KA, Jamrich M. 1997. The Rx
homeobox gene is essential for vertebrate eye development.
Nature 387:603–607.
Matsuo T, Osumi-Yamashita N, Noji S, Ohuchi H, Koyama E,
Myokai F, 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.
Matter-Sadzinski L, Matter JM, Ong MT, Hernadez J, Ballivet M.
2001. Specification of neurotransmitter receptor identity in developing retina: the chick ATH5 promoter integrates the positive
and negative effects of several bHLH proteins. Development 128:
217–231.
Maurus D, Heligon C, Burger-Schwarzler A, Brandi AW, Kuhl M.
2005. Noncanonical Wnt-4 signaling and EAF2 are required for
eye development in Xenopus Laevis. EMBO J 24:1181–1191.
McCabe KL, Gunther EC, Reh TA. 1999. The development of the
pattern of retinal ganglion cells in the check retina: mechanisms
that control differentiation. Development 126:5713–5724.
McFarlane S, Zuber ME, Holt CE. 1998. A role of the fibroblast
growth factor receptor in cell fate decisions in the developing vertebrate retina. Development 125:3967–3975.
Mears AJ, Kondo M, Swain PK, Takada Y, Bush RA, Saunders TL,
Sieving PA, Swaroop A. 2001. Nrl is required for rod photoreceptor development. Nat Genet 29:447–452.
Merhi-Soussi F, Angenieux B, Canola K, Kostic C, Tekaya M,
Hornfield D, Arsenijevic Y. 2006. High yield of cells committed to
the photoreceptor fate from expanded mouse retinal stem cells.
Stem Cells 24:2060–2070.
Meyer JS, Howden SE, Wallace KA, Verhoeven AD, Wright LS,
Capowski EE, Pinilla I, Martin JM, Tian S, Stewart R, Pattnaik
B, Thomson JA, Gamm DM. 2011. Optic vesicle-like structures
derived from human pluripotent stem cells facilitate a customized
approach to retinal disease treatment. Stem Cells 29:1206–1218.
Meyer JS, Katz ML, Maruniak JA, Kirk MD. 2004. Neural differentiation of mouse embryonic stem cells in vitro and after transplantation into eyes of mutant mice with rapid retinal
degeneration. Brain Res 1014:131–144.
Meyer JS, Katz ML, Maruniak JA, Kirk MD. 2005. Embryonic stem
cell-derived neural progenitors incorporate into degenerating retina and enhance survival of host photoreceptors. Stem cells 24:
274–283.
Meyer JS, Katz ML, Maruniak JA, Kirk MD. 2006. Embryonic stem
cell-derived neural progenitors incorporate into degenerating retina and enhance survival of host photoreceptors. Stem Cells 34:
274–283.
Meyer JS, Shearer RL, Capowski EE, Wright LS, Wallace KA,
McMillan EL, Zhang SC, Gamm DM. 2009. Modeling early retinal
development with human embryonic and induced pluripotent
stem cells. Proc Natl Acad Sci USA 106:16698–16703.
Mezey E, Chandross KJ, Harta G, Maki RA, McKercher SR. 2000.
Turning blood into brain: cells bearing neuronal antigens generated in vivo from bone marrow. Science 290:1779–1782.
Miller FD, Gauthier-Fisher A. 2009. Home at last: neural stem cell
niches defined. Cell Stem Cell 4:507–510.
Ming GL, Song H. 2005. Adult neurogenesis in the mammalian central nervous system. Annu Rev Neurosci 28:223–250.
Mochii M, Mazaki Y, Mizuno N, Hayashi H, Eguchi G. 1998. Role of
Mitf in differentiation and transdifferentiation of chicken pigmented epithelial cell. Dev Biol 193:47–62.
Moe MC, Kolberg RS, Sandberg C, Vik-Mo E, Olstorn H, Varghese
M, Langmoen IA, Nicolaissen B. 2009. A comparison of epithelial
and neural properties in progenitor cells derived from the adult
human ciliary body and brain. Exp Eye Res 88: 30–38.
Monnin J, Morand-Villeneuve N, Michel G, Hicks D, VersauxBotteri C. 2007. Production of neurospheres from mammalian
Muller cells in culture. Neurosci Lett 421:22–26.
Mori T, Guffo A, Gotz M. 2005. The novel roles of glial cells revisited: the contribution of radial glia and astrocytes to neurogenesis. Curr Top Dev Biol 69:67–99.
Morrow EM, Belliveau MJ, Cepko CL. 1998. Two phases of rod photoreceptor differentiation during rat retinal development. J Neurosci 18:3738–3748.
Moshiri A, Close J, Reh TA. 2004. Retinal stem cells and regeneration. Int J Dev Biol 48:10003–1014.
Moustakas A, Souchelnytskyi S, Heldin CH. 2001. Smad regulation
in TGF-beta signal transduction. J Cell Sci 114:4359–4369.
Mui SH, Hindges R, O’Leary DD, Lemke G, Bertuzzi S. 2002. The
homeodomain protein Vax2 patterns the dorsoventral and nasotmeporal axes of the eye. Development 129:797–804.
Mukhopadhyay M, Gorivodsky M, Shtrom S, Grinberg A, Niehrs C,
Morasso MI, Westphal H. 2006. Dkk2 plays an essential role in
the corneal fate of the ocular surface epithelium. Development
133:2149–2154.
Murone M, Rosenthal A, de Sauvage FJ. 1999. Sonic hedgehog signaling by the patched-smoothened receptor complex. Curr Biol 9:
76–84.
Nakagawa M, Koyanagi M, Tanabe K, Takahashi K, Ichisaka T, Aoi
T, Okita K, Mochiduki Y, Takizawa N, Yamanaka S. 2008. Generation of induced pluripotent stem cells without Myc from mouse
and human fibroblasts. Nat Biotechnol 26:101–106.
Neumann CJ, Nuesslein-Volhard C. 2000. Patterning of the zebrafish retina by a wave of sonic hedgehog activity. Science 289:
2137–2139.
RSC AND REGENERATION OF VISION SYSTEM
Nguyen M, Arnheiter H. 2000. Signaling and transcriptional regulation in early mammalian eye development: a link between FGF
and MITF. Dev 127:3581–3591.
Nickerson PE, Emsley JG, Myers T, Clarke DB. 2007. Proliferation
and expression of progenitor and mature retinal phenotypes in
the adult mammalian ciliary body after retinal ganglion cell
injury. Invest Ophthalmic Vis Res 48:5266–5275.
Nickerson PE, Da Silva N, Myers T, Stevens K, Clarke DB. 2008.
Neural progenitor potential in cultured Muller glia: effects of passaging and exogenous growth factor exposure. Brain Res 1230:1–
12.
Nishida A, Furukawa A, Koike C, Tano Y, Aizawa S, Matsuo I,
Furukawa T. 2003. Otx2 homeobox gene controls retinal photoreceptor cell fate and pineal gland development. Nat Neurosci 6:
1255–1263.
Nistor G, Seiler MJ, Yan F, Ferguson D, Keirstead HS. 2010. Threedimensional early retinal progenitor 3D tissue contructs derived
from human embryonic stem cells. J Neurosci Methods 190:63–
70.
Nordstrom U, Jessel TM, Edlund T. 2002. Progressive induction of
caudal neural character by graded Wnt signaling. Nat Neurosci 5:
525–532.
Nornes HO, Dressler GR, Knapid EW, Deutsch U, Gruss P. 1990.
Spatially and temporally restricted expression of Pax2 during
murine neurogenesis. Development 109:797–809.
Ohnuma S, Philpott A, Wang K, Holt CE, Harris WA. 1999. p27Xic1,
a Cdk inhibitor, promotes the determination of glial cells in Xenopus retina. Cell 99:499–510.
Ohtsuka T, Ishibashi M, Gradwohl G, Nakanishi S, Guillemot F,
Kageyama R. 1999. Hes1 and Hes5 as Notch effectors in mammalian neuronal differentiation. EMBO J 18:2196–2207.
Okada TS. 1980. Cellular metaplasia or transdifferentiation as a
model for retinal cell differentiation. Curr Top Dev Biol 16:349–
380.
Ooto S, Akagi T, Kageyama R, Akita J, Mandai M, Honda Y,
Takahashi M. 2004. Potential for neural regeneration after neurotoxic injury in the adult mammalian retina. Proc Natl Acad Sci
USA 101:13654–13659.
Osakada F, Ikeda H, Mandai M, Wataya T, Watanabe K, Yoshimura
N, Akaike A, Sasai Y, Takahashi M. 2008. Toward the generation
of rod and cone photoreceptors from mouse, monkey and human
embryonic stem cells. Nat Biotech 26:215–224.
Osakada F, Ikeda H, Sasai Y, Takahashi M. 2009a. Stepwise differentiation of pluripotent stem cells into retinal cells. Nat Protoc 4:
811–824.
Osakada F, Jin ZB, Hirami Y, Ikeda H, Danjyo T, Watanabe K,
Sasai Y, Takahashi M. 2009b. In vitro differentiation of retinal
cells from human pluripotent stem cells by small-molecule induction. J Cell Sci 122:3169–3179.
Osakada F, Ooto S, Akagi T, Mandai M, Akaike A, Takahashi M.
2007. Wnt signaling promotes regeneration in the retina of adult
mammalian. J Neurosci 27:4210–4219.
Otani A, Dorrell MI, Kinder K, Moreno SK, Nusinowitz S, Banin E,
Heckenlively J, Friedlander M. 2004. Rescue of retinal
degeneration by intravitreally injected adult bone marrow-derived
lineage-negative hematopoietic stem cells. J Clin Invest 114:
765–774.
Otteson DC, Hitchcock PF. 2003. Stem cells in the teleost retina:
persistent neurogenesis and injury-induced regeneration. Vis Res
43:927–936.
Ouchi Y, Tabata Y, Arai K, Watanabe S. 2005. Negative regulation
of retinal-neurite extension by beta-catenin signaling pathway. J
Cell Sci 118:4473–4483.
Palmer TD, Markakis EA, Willhoite AR, Safar F, Gage FH. 1999.
Fibroblast growth factor-2 activates a latent neurogenic program
in neural stem cells from diverse regions of the adult CNS. J Neurosci 19:8487–8497.
Palmer TD, Ray J, Gage FH. 1995. FGF-2-responsive neuronal progenitors reside in proliferative and quiescent regions of the adult
rodent brain. Mol Cell Neurosci 6:474–486.
Park UC, Cho MS, Park JH, Kim SJ, Ku SY, Choi YM, Moon SY,
Yu HG. 2011. Subretinal transplantation of putative retinal pige-
157
ment epithelial cells derived from human embryonic stem cells in
rat retinal degeneration model. Clin Exp Reprod Med 38:216–221.
Patel A and McFarlane S. 2000. Overexpression of FGF-2 alters cell
fate specification in the developing retina of Xenopus laevis. Dev
Biol 222. 170–080.
Pearson R A, Barber AC, Rizzi M, Hippert C, Xue T, West EL,
Duran Y, Smith AJ, Chuang JZ, Azam SA, Luhmann UFO,
Benucci A, Sung CH, Bainbridge JW, Carandini M, Yau K-W,
Sowden JC, Ali RR. (2012) Restoration of vision after transplantation of photoreceptors. Nature 485:99–103.
Peng G-H, Ahmad O, Ahmad F, Liu J, Chen S. 2005. The
photoreceptor-specific nuclear receptor Nr2e3 interacts with CrX
and exerts opposing effects on the transcription of rod versus
cone genes. Hum Mol Genet 14:747–764.
Perron M, Boy S, Amato MA, Viczian A, Koebernick K, Pieler T,
Harris WA. 2003. A novel function of Hedgehog signaling in retinal pigment epithelium differentiation. Development 130:1565–
1577.
Perron M, Harris WA. 2000. Retinal stem cells in vertebrates. Bioessays 22:685–688.
Perron M, Kanekar S, Vetter ML, Harris WA. 1998. The genetic
sequence of retinal development in the ciliary margin of the Xenopus eye. Dev Biol 199:185–200.
Pittack C, Grunwald GB, Reh TA. 1997. Fibroblast growth factors
are necessary for neural retina but not pigmented epithelium differentiation in chick embryo. Development 124:805–816.
Pluchino S, Zanotti L, Rossi B, Brambilla E, Ottoboni L, Salani G,
Martinello M, Cattalini A, Bergami A, Furlan R, Comi G,
Constantin G, Martino G, 2005. Neurosphere-derived multipotent
precursors promote neuroprotection by an immunomodulatory
mechanism. Nature 436:266–271.
Prada C, Puga J, Perez-Mendez L, Lopez R, Ramirez G. 1991. Spatial and temporal patterns of neurogenesis in the chick retina.
Eur J Neurosci 3:559–569.
Ramachandran R, Reifler A, Parent JM, Goldman D. 2010. Conditional gene expression and lineage tracing of tuba1a expressing
cells during zebrafish development and retina regeneration. J
Comp Neurol 518:4196–4212.
Rao S, Lobov IB, Vallance JE, Tsujikawa K, Shiojima I, Akunuru S,
Walsh K, Benjamin LE, Lang RA. 2007. Obligatory participation
of macrophages in an angiopoietin 2-mediated cell death switch.
Development 134:4449–4458.
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 98:3861–3866.
Raymond PA, Barthel LK, Bernardos RL, Perkowski JJ. 2006.
Molecular characterization of retinal stem cells and their niches
in adult zebrafish. BMC Dev Biol 6:36.
Raymond PA, Hitchcock PF, 1997. Retinal regeneration: common
principles but a diversity of mechanisms. Adv Neurol 72:171–184.
Raymond PA, Hitchcock PF. 2000. How the neural retina regenerates. Results Probl Cell Diff 31:197–218.
Raymond PA, Reifler MJ, Rivlin PK. 1988. Regeneration of goldfish
retina: rod precursors are a likely source of regenerated cells. J
Neurobiol 19:431–463.
Reh TA, Tully T. 1986. Regulation of tyrosine hydroxylase-containing amacrine cell number in larval frog retina. Dev Biol 114:463–
469.
Reh TA. 1987. Cell-specific regulation of neuronal production in the
larval frog retina. J Neurosci 7:3371–3324.
Reh TA, Nagy T, Gretton H. 1987. Retinal pigmented epithelial cells
induced to transdifferentiate to neurons by laminin. Nature 330:
68–71.
Reh TA. 2006. Neurobiology: right timing for retina repair. Nature
444:156–157.
Reh TA, Fischer AJL. 2001. Stem cells in the vertebrate retina.
Brain Behav Evol 58:296–305.
Reh TA, Levine EML. 1998. Multipotential stem cells and progenitors in the vertebrate retina. J Neurobiol 36:206–220.
Reh TA, Nagy T. 1987. A possible role for the vascular membrane in
retinal regeneration in Rana catesbianna tadpoles. Dev Biol 122:
471–482.
158
YIP
Reichenback A, Wurm A, Pannicke T, Iandiev I, Wiedemann P,
Bringmann A. 2007. Muller cells as players in retinal
degeneration and edema. Graefes Arch Clin Exp Ophthalmol 245:
627–636.
Reubinoff BE, Pera MF, Fong C-Y, Trounson A, Bongso A. 2000.
Embryonic stem cell lines from human blastocysts: somatic differentiation in vitro. Nature Biotechnol 18:399–404.
Reuss B, von Bohlenund Halbach O. 2003. Fibroblast growth factors
and their receptors in the central nervous system. Cell Tissue Res
313:139–157.
Reynolds BA, Weiss S. 1992. Generation of neurons and astrocytes
from isolated cells of the adult mammalian central nervous system. Science 255:1707–1710.
Ritter MR, Banin E, Moreno SK, Aguilar E, Dorrell MI,
Friedlander M. 2006. Myeloid progenitors differentiate into microglia and promote vascular repair in a model of ischemic retinopathy. Clin Invest 116:3266–3276.
Robel S, Berninger B, G€
otz M. 2011. The stem cell potential of glia:
lessons from reactive gliosis. Nat Rev Neurosci 12:88–104.
Rodriguez J, Esteve P, Weinl C, Ruiz JM, Fermin Y, Trousee F,
Dwivedy A, Holt C, Bovolenta P. 2005. SFRP1 regulates the
growth of retinal ganglion cell axons through the Fz2 receptor.
Nat Neurosci 8:1301–1309.
Roesch K, Jadhav AP, Trimarchi JM, Stadler MB, Roska B, Sun
BB, Cepko CL. 2008. The transcriptome of retinal Muller glial
cells. J Comp Neurol 509:225–238.
Gritti A, Parati EA, Cova L, Frolichsthal P, Galli R, Wanke E,
Faravelli L, Morassutti DJ, Roisen F, Nickel DD, Vescovi AL.
1996. Multipotential stem cells from the adult mouse brain proliferate and self-renew in response to basic fibroblast growth factor.
J Neurosci 16:1091–1100.
Rothermel A, Willbold E, Degrip WJ, Layer PG. 1997. Pigmented
epithelium induces complete retinal reconstitution from dispersed
embryonic chick retinae in reaggregation culture. Proc Biol Sci
264:1293–1302.
Sabelstrom H, Stenudd M, Frisen J. 2013. Neural stem cells in the
adult spinal cord. Exp Neurol; DOI: 10.1016/j.expneurol.
2013.01.026.
Saika S, Liiu CY, Azhar M, Sanford LP, Doetschamn T, Gendron
RL, Kao CW, and Kao WW. 2001. Tgfbeta2 in corneal morphogenesis during mouse embryonic development. Dev Biol 240:419–432.
Sakaguchi DS, van Hoffelen SJ, Young MJ. 2003. Differentiation
and morphological integration of neural progenitor cells transplanted into the developing mammalian eye. Ann NY Acad Sci
995:127–139.
Sakuta H, Suzuki R, Takahashi H, Kato, A, Shintani T, Iemura S,
Yamamoto TS, Ueno N, Noda M. 2001. Ventroptin: a BMP-4 anatgonist expressed in a double-gradient pattern in the retina. Science 293:111–115.
Sanchez-Ramos J, Song S, Cardozo-Pelaez F, Hazzi C, Stedeford T,
Willing A, Freeman TB, Saporta S, Janssen W, Patel N, Cooper
DR, Sanberg PR. 2000. Adult bone marrow stromal cells differentiate into neural cells in vitro. Exp Neurol 164:247–256.
Sanford L, Ormsby I, Gittenberger-de Groot A, Sariola H, Friedman
R, Boivin G, Cardell E, Doetschman T. 1997. TGFbeta2 knockout
mice have multiple developmental defects that are nonoverlapping with other TGFbeta knockout phenotypes. Development 124:2659–2670.
Sasagawa S, Takabatake T, Takabatake Y, Muramatsu T,
Takeshima K. 2002. Axes establishment during eye morphogenesis in Xenopus by coordinate and antagonistic actions of BMP4,
Shh, and RA. Genesis 33:86–96.
Satoh K, Kasai M, Ishidao T, Tago K, Ohwada S, Hasegawa Y,
Senda T, Takada S, Nada S, Nakamura T, Akiyama T. 2004. Anteriorization of neural fate by inhibitor of beta-catenin and T cell
factor (ICAT), a negative regulator of Wnt signaling. Proc Natl
Acad Sci USA 101:8017–8021.
Schlessinger J. 2000. Cell signaling by receptor tyrosine kinases.
Cell 103:211–225.
Schmitt AM, Shi J, Wolf AM, Lu CC, King LA, Zou Y. 2006. WntRyk signaling mediates medial-lateral retinotectal topographic
mapping. Nature 439:31–37.
Schulte D, Furukawa T, Peters MA, Kozak CA, Cepko CL. 1999.
Misexpression of the Emx-related homeobox genes cVax and
mVax2 ventralized the retina and perturbs the retinotectal map.
Neuron 24:541–553.
Schwartz SD, Hubschman, JP, Heilwell G, Franco-Cardenas V, Pan
CK, Ostrick RM, Mickunas E, Gay R, Klimanskaya I, Lanza R.
2012. Embryonic stem cell trials for macular degeneration: a preliminary report. The Lancet 379:713–720.
Sehgal R, Andres DJ, Adler R, Belecky-Adams TL. 2006. Invest
Opthalmol Vis Sci 47:3625–3634.
Seko Y, Azuma N, Kaneda M, Nakatani K, Miyagawa Y, Noshiro Y,
Kurokawa R, Okano H, Umezawa A. 2012. Derivation of human
differential photoreceptor-like cells from the iris by defined combination of CRX, RX and NEUROD. PLoS One 7:e35611.
Sheldahl LC, Park M, Malbon CC, Moon RT. 1999. Protein kinase C
is differentially stimulated by Wnt and Frizzled homologs in a Gprotein-dependent manner. Curr Biol 9:695–698.
Simons M, Mlodzik M. 2008. Planar cell polarity signaling: from fly
development to human diseases. Annu Rev Genet 42:517–540.
Singh MS, Issa PC, Butler R, Martin C, Lipinski DM, Sekaran S,
Barnard AR, MacLaren RE. 2013. Reversal of end-stage retinal
degeneration and restoration of visual function by photoreceptor
transplantation. Proc Natl Acad Sci 110:1101–1106.
Singhal S, Bhatia B, Jayaram H, Becker S, Jone MF, Cottrill PB,
Khaw PT, Salt TE, Limb GA. 2012. Human Muller glia with
stem cell characteristics differentiate into retinal ganglion
cell (RGC) precursors in vitro and partially restore RGC function
in vivo following transplantation. Stem Cells Transl Med 1:188–
199.
Sirko S, von Holst A, Weber A, Wizenmann A, Theocharidis U, G€
otz
M, Faissner A. 2010. Chondroitin sulfates are required for fibroblast growth factor-2-dependent proliferation and maintenance in
neural stem cells and for epidermal growth factor-dependent
migration of their progeny. Stem cells 28:775–787.
Smith AN, Miller LA, Song N, Taketo MM, Lang RA. 2005. The
duality of beta-catenin function: a requirement in lens morphogenesis and signaling suppression of lens fate in periocular ectoderm. Dev Biol 285:477–489.
Stenkamp DL, Frey RA, Prabhudesai SN, Raymond PA. 2000. Function for hedgehog genes in zebrafish retinal development. Dev
Biol 220:238–252.
Stern CD. 2006. Neural induction: 10 years on since the "default
model". Curr Opin Cell Biol 18:692–697.
Sullivan SA, Barthel LK, Largent BL, Raymond PA. 1997. A goldfish notch-3 homologue is expressed in neurogenic regions of
embryonic, adult, and regenerating brain and retina. Dev Genet
20:208–223.
Sun G, Asami M, Ohta H, Kosaka J, Kosaka M. 2006. Retinal stem/
progenitor properties of iris pigment epithelial cells. Dev Biol 289:
243–252.
Starznicky K, Gaze RM. 1971. The growth of the retina in Xenopus
Laevis: an autoradiographic study. J Embryol Exp Morphol 26:
67–79.
Storey KG, Goriely A, Sargent CM, Brown JM, Burns HD, Abud
HM, Heath JK. 1998. Early posterior neural tissue is induced by
FGF in the chick embryo. Development 125:473–484.
Susaki K, Chiba C. 2007. MEK mediates in vitro neural transdifferentiation of the adult newt retinal pigment epithelium cells: is
FGF2 an induction factor? Pigment Cell Res 20:364–379.
Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda
K, Yamanaka S. 2007. Induction of pluripotent stem cells from
adult human fibroblasts by defined factors. Cell 131:861–872.
Takahashi K, Yamanaka S. 2006. Induction of pluripotent stem cells
from mouse embryonic and adult fibroblast cultures by defined
factors. Cell 126:663–676.
Takatsuka K, Hatakeyama J, Bessho Y, Kageyama R. 2004. Roles of
the bHLH gene Hes1 in retinal morphogenesis. Brain Res 1004:
148–155.
Takeda M, Takamiya A, Jiao JW, Cho KS, Trevino SG, Matsuda T,
Chen DF. 2008. Alpha-aminoadipate induces progenitor cell properties of Muller glia in adult mice. Invest Ophthalmol Vis Sci 49:
1142–1150.
RSC AND REGENERATION OF VISION SYSTEM
Take-uchi M, Clarke JD, Wilson SW. 2003. Hedgehog signaling
maintains the optic stalk-retinal interface through the regulation
of Vax gene activity. Development 130:955–968.
Temple S. 2001. The development of neural stem cells. Nature 414:
112–117.
Thomson JA, Itskovitz-Eldor J, Shapiro SS, Waknitz MA, Swiergiel
JJ, Marshall VS, Jones JM. 1998. Embryonic stem cell lines
derived from human blastocysts. Science 282:1145–1147.
Tomita K, Moriyoshi K, Nakanishi S, Guillemot F, Kageyama R.
2000. Mammalian achaete-scute and atonal homologs regulate
neuronal versus glial fate determination in the central nervous
system. EMBO J 19:5460–5472.
Tomita K, Nakanishi S, Guillemot F, Kageyama R. 1996. Mash1
promotes neuronal differentiation in the retina. Genes Cells 1:
765–774.
Tomita M, Adachi Y, Yamada H, Takahashi K, Kiuchi K, Oyaizu H,
Ikebukuro K, Kaneda H, Matsumura M, Ikehara SL. 2002. Bone
marrow-derived stem cells can differentiate into retinal cells in
injured rat retina. Stem Cells 20:279–283.
Topp S, Stigloher C, Komisarczuk AZ, Adolf B, Becker TS, BallyCuif L. 2008. Fgf signaling in the zebrafish adult brain: association of Fgf activity with ventricular zones but not cell proliferation. J Comp Neurol 510:422–439.
Tree DR, Ma D, Axelrod JD. 2002. A three-tiered mechanism for
regulation of planar cell polarity. Semin Cell Dev Biol 13:217–
224.
Tropepe V, Coles BL, Chiasson BJ, Horsford DJ, Elia AJ, McInnes
RR, van der Kooy D. 2000. Retinal stem cells in the adult mammalian eye. Science 287:2032–2036.
Tropepe V, Sibilia M, Ciruna BG, Rossant J, Wagner EF, van der
Kooy D. 1999. Distinct neural stem cells proliferate in response to
EGF and FGF in the developing mouse telencephalon. Dev Biol
208:166–188.
Trousse F, Esteve P, Bovolenta P. 2001a. Bmp4 mediates
apoptotic cell death in the developing chick eye. J Neurosci 21:
1292–1301.
Trousse F, Marti E, Gruss P, Torres M, Bovolenta P. 2001b. Control
of retinal ganglion cell axon growth: a new role for Sonic hedgehog. Development 128:3927–3936.
Tsonis PA, Jang W, Del Rio-Tsonis K, Eguchi G. 2001. A unique
aged human retinal pigmented epithelial cell line useful for
studying lens differentiation in vitro. Int J Dev Biol 45:753–758.
Tucker BA, Park IH, Qi SD, Klassen HJ, Jiang C, Yao J, Redenti S,
Daley GQ, Young MJ. 2011. Transplantation of adult mouse iPS
cell-derived photoreceptor precursors restores retinal structure
and function in degenerative mice. PLoS One 6:e18992.
Turner DL, Cepko CL. 1987. A common progenitor for neurons and
glia persists in rat retina late in development. Nature 328:131–
136.
Turner DL, Snyder EY, Cepko CL. 1990. Lineage-independent
determination of cell type in the embryonic mouse retina. Neuron
4:833–845.
Turner N, Grose R. 2010. Fibroblast growth factor signaling: from
development to cancer. Nat Rev Cancer 10:116–129.
Vaajasaari H, Ilmarinen T, Juuti-Uusitalo K, Rajala K, Onnela N,
Narkilahti S, Suuronen R, Hyttinen J, uusitalo H, Skottman H.
2011. Toward the defined and xeno-free differentiation of functional human pluripotent stem cell-derived retinal pigment epithelial cells. Mol Vis 17:558–575.
Van Raay TJ, Moore KB, Iordanova I, Steele M, Jamrich M, Harris
WA, Vetter ML. 2005. Frizzled5 signaling governs the neural
potential of progenitors in the developing Xenopus retina. Neuron
46:23–36.
Venkataraman G, Raman R, Sasisekharan V, Sakisekharan R.
1999. Molecular characteristics of fibroblast growth factorfibroblast growth factor receptor-heparin-like glycosaminoglycan
complex. Proc Natl Acad Sci USA 96:3658–3663.
Vihtelic TS, Hyde DR. 2000. Light-induced rod and cone cell death
and regeneration in the adult albino zebrafish (Danio rerio) retina. J Neurobiol 44:289–307.
Vinores SA, Derevjanik NL, Mahlow J, Hackett SF, Haller JA,
deJuan E, Frankfurter A, Campochiaro PA. 1995. Clasee III beta-
159
tubulin in human retinal pigment epithelial cells in culture and
in epiretinal membranes. Exp Eye Res 60:385–400.
Vugler AA, Carr AJ, Lawrence J, Chen LL, Burrell K, Wright A,
Lundh P, Semo M, Ahmado A, Gias C, da Cruz L, Moore H,
Andrews P, Walsh J, Coffey P. 2008. Elucidating the phenomenon
of HESC-derived RPE: anatomy of cell genesis, expansion and retinal transplantation. Exp Neurol 214:347–361.
Wan J, Zheng H, Chen ZL, Xiao HL, Shen ZJ, Zhou GM. 2008. Preferential regeneration of photoreceptor from Muller glia after retinal degeneration in adult rat. Vis Res 48:223–234.
Wang SW, Kim BS, Ding K, Wang H, Sun D, Johnson RL, Klein
WH, Gan L. 2001. Requirement of math5 in the development of
retinal ganglion cells. Genes Dev 15:24–29.
Watanabe T, Raff MC. 1990. Rod photoreceptor development in
vitro: intrinsic properties of proliferating neuroepithelial cells
changes as development proceeds in the rat retina. Neuron 4:
461–467.
Wawersik S, Prucell P, Rauchman M, Dudley AT, Robertson EJ,
Mass R. 1999. BMP7 acts in murine lens placode development.
Dev Biol 207:176–188.
Weinandy F, Ninkovic J, G€
otz M. 2011. Restrictions in time and
space-new insights into generation of specific neuronal subtypes
in the adult mammalian brain. Eur J Neurosci 33:1045–1054.
Weiss S, Dunne C, Hewson J, Wohl C, Wheatley M, Peterson AC,
Reynolds BA. 1996. Multipotent CNS stem cells are present in
the adult mammalian spinal cord and ventricular neuroaxis. J
Neurosci 16:7599–7609.
Wernig M, Meissner A, Foreman R, Brambrink T, Ku M,
Hochedlinger K, Berstein BE, Jaenisch R. 2007. In vitro reprogramming of fibroblasts into a pluripotent ES-cell-like state.
Nature 448:318–324.
Wetts R, Fraser SE. 1988. Multipotent precursors can give rise to
all major types of frog retina. Science 239:1142–1145.
Wharton KA. 2003. Runnin’ with the dvl: Proteins that associate
with Dsh/Dvl and their significance to wnt signaling transduction. Dev Biol 253:1–17.
White, NM, Jarman AP. 2000. Drosophila atonal controls
photoreceptor R8-specific properties and modulates both receptor
tyrosine kinase and Hedgehog signaling. Development 127:1681–
1689.
Wohl SG, Schmeer CW, Kretz A, Witte OW, Isenmann S. 2009.
Optic nerve lesion increases cell proliferation and nestin expression in the adult mouse eye in vivo. Exp Neurol 219:175–186.
Woodbury D, Schwarz EJ, Prockop DJ, Black IB. 2000. Adult rat
and human bone marrow stromal cells differentiate into neurons.l
J Neurosci Res 61:364–370.
Wu DM, Schneiderman T, Burgett J, Gokhale P, Barthel L,
Raymond PA. 2001. Cones regenerate from retinal stem cells
sequestered in the inner nuclear layer of adult goldfish retina.
Invest Ophthalmol Vis Sci 42:2115–2124.
Xiao Q, Du Y, Wu W, Yip HK. 2010. Bone morphogenetic proteins
mediate cellular response and, together with Noggin, regulate
astrocyte differentiation after spinal cord injury. Exp Neurol 221:
353–366.
Xu H, StaIglesia DD, Kielczewski JL, Valenta DF, Pease ME, Zack
DJ, Quigley HA. 2007. Characteristics of progenitor cels derived
from adult ciliary body in mouse, rat, and human eyes. Invest
Ophthalmol Vis Sci 48:1674–1682.
Xu Q, Wang Y, Dabdoub A, Smallwood PM, Williams J, Woods C,
Kelly MW, Jiang L, Tasman W, Zhang K, Nathans J. 2004.
Vascular development in the retina and inner ear: control by Norrin and Frizzled-4, a high-affinity ligand-receptor pair. Cell 116:
883–895.
Xue LP, Lu J, Cao Q, Hu S, Ding P, Ling EA. 2006. Muller glial
cells express nestin coupled with glial fibrillary acidic protein in
experimentally induced glaucoma in the rat retina. Neuroscience
139:723–732.
Yamaguchi M, Saito H, Suzuki M, Mori K. 2000. Visualization of
neurogenesis in the central nervous system using nestin
promoter-GFP transgenic mice. Neuroreport 11:1991–1996.
Yamaguchi TP. 2001. Heads or tails: Wnts and anterior-posterior
patterning. Curr Biol 11:713–724.
160
YIP
Yang P, Seiler MJ, Aramant RB, Whittmore SR. 2002. Differential
lineage restriction of rat retinal progenitor cells in vitro and in
vivo. J Neurosci Res 69:466–476.
Yang XY, Cepko CL. 1996. Flk-1, a receptor for vascular endothelial
growth factor (VEGF), is expressed by retinal progenitor cells. J
Neurosci 16:6089–6099.
Yu J, Hu K, Smuga-Otto K, Tian S, Stewart R, Slukvin II, Thomson
JA. 2009. Human induced pluripotent stem cells free of vector
and transgene sequences. Science 324:797–801.
Yu J, Vodyanik MA, Smuga-Otto K, Antosiewica-Bourget J, Frane
JL, Tian S, Nie J, Jonsdottir GA, Ruotti V, Stewart R, Slukvin II,
Thomson JA. 2007. Induced pluripotent stem cell lines derived
from human somatic cells. Science 318:1917–1920.
Yurco P, Cameron DA. 2005. Responses of Muller glia to retinal
injury in adult zebrafish. Vision Res 45:991–1002.
Zallen JA. 2007. Planar polarity and tissue morphogenesis. Cell
129:1051–1063.
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 28:135–142.
Zhang XM, Yang X-J. 2001. Temporal and spatial effects of Sonic
hedgehog signaling in chick eye morphogenesis. Dev Biol 233:
271–290.
Zhao S, Chen Q, Hung FC. 2002a. BMP signaling is required for
development of the ciliary body. Development 129:4435–4442.
Zhao S, Hung FC, Colvin JS, White A, Dai W, Lovicu FJ, Ornitz
DM, Overbeek PA. 2001. Patterning the optic neuroepithelium by
FGF signaling and Ras activatioin. Development 128:5051–5060.
Zhao S, Overbeck PA. 1999. Tyrosine-related protein 2 promoter
targets transgene expression to ocular and neural crest-derived
tissues. Dev Bio 216:154–163.
Zhao S, Thornquist SC, Barnstable CJ. 1995. In vitro transdifferentiation of embryonic rat retinal pigment epithelium to neural retina. Brain Res 677:300–310.
Zhao S, Barnstable CJ. 1996. Differential effects of bFGF on development of the rat retina. Brain Res 723:169–176.
Zhao X, Das AV, Thoreson WB, James J, Wattnem TE, RodriguezSierra J, Ahmad I. 2002b. Adult corneal limbal epithelium: a
model for studying neural potential of non-neural stem cells/progenitors. Dev Biol 250:317–331.
Zhao X, Liu J, Ahmad I. 2002c. Differentiation of embryonic stem cells
into retinal neurons. Biochem Biophy Res Commun 297:177–184.
Zhou L, Wang W, Liu Y, Fernandez de Castro J, Ezashi T, Telugu
BP, Roberts RM, Kaplan HJ, Dean DC. 2011. Differentiation of
induced pluripotent stem cells of swine into rod photoreceptors
and their integration into the retina. Stem Cells 29:972–980.
Zuba-Surma EK, Kucia M, Dawn B, Guo Y, Ratajczak MZ, Bolli R.
2008. Bone marrow-derived pluripotent very small embryonic –
like stem cells (VSELs) are mobilized after acute myocardial
infarction. J Mol Cell Cardiol 44. 865–873.
Zuber ME, Gestri G, Viczian AS, Barsacchi G, Harris WA. 2003.
Specification of the vertebrate eye by a network of eye field transcription factors. Development 130:5155–5167.
Zuber ME, Harris WA. 2006. Formation of the eye field. In: Retinal
development, Sernagor E, Eglen S, Harris W, Wong R, editors.
Cambridge, UK: Cambridge University Press. p. 8–29.
Download