You are on page 1of 31

Progress in Retinal and Eye Research xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Progress in Retinal and Eye Research


journal homepage: www.elsevier.com/locate/preteyeres

Recapitulating developmental mechanisms for retinal regeneration


Iqbal Ahmada,∗, Pooja Teotiaa, Helen Ericksona, Xiaohuan Xiab
a
Department of Ophthalmology and Visual Science, University of Nebraska Medical Center, Omaha, NE, 68198, USA
b
Center for Translational Neurodegeneration and Regenerative Therapy, Shanghai Tenth People's Hospital Affiliated to Tongji University School of Medicine, Shanghai,
200072, China

ARTICLE INFO ABSTRACT

Keywords: Degeneration of specific retinal neurons in diseases like glaucoma, age-related macular degeneration, and re-
Retina tinitis pigmentosa is the leading cause of irreversible blindness. Currently, there is no therapy to modify the
Development disease-associated degenerative changes. With the advancement in our knowledge about the mechanisms that
Regeneration regulate the development of the vertebrate retina, the approach to treat blinding diseases through regenerative
Pluripotent stem cells
medicine appears a near possibility. Recapitulation of developmental mechanisms is critical for reproducibly
Disease modeling
Organoid
generating cells in either 2D or 3D culture of pluripotent stem cells for retinal repair and disease modeling. It is
the key for unlocking the neurogenic potential of Müller glia in the adult retina for therapeutic regeneration.
Here, we examine the current status and potential of the regenerative medicine approach for the retina in the
backdrop of developmental mechanisms.

1. Introduction However, transformative research over the last twenty years has led to
discoveries that are promising for regenerative medicine for retinal
The retina, an integral part of the central nervous system (CNS), degeneration. These include (1) the discovery that mammalian MG
consists of seven different cell types, histologically organized in an possess neurogenic potential, raising the prospect of treating retinal
evolutionarily conserved laminar structure, which are responsible for degeneration from within (Ahmad et al., 2011; Goldman, 2014), (2) the
generating and transmitting the visual signal. For example, rod and directed differentiation of pluripotent stem cells, whether embryonic
cone photoreceptors (rods and cones, respectively) capture light re- stem (ES) cells or induced pluripotent stem (iPSC) cells (Takahashi and
flected from an object and generate an electrical signal. This is subse- Yamanaka, 2006) in 2D culture into retinal cells (Borooah et al., 2013;
quently relayed to retinal ganglion cells (RGCs) for transmission to Gasparini et al., 2019; Parameswaran et al., 2010), and (3) self orga-
higher centers in the brain for visual perception after having been nization of pluripotent stem cells into 3D retinal organoids (Eiraku
modulated by the intervening neurons, the horizontal cells (HCs), bi- et al., 2011), providing platforms for disease modeling and cells for
polar cells (BCs), and amacrine cells (ACs). Müller glia (MG), the single retinal repair. When combined with drug screening, disease modeling
glia generated by the multipotential retinal progenitor cells (RPCs), using iPSC lines from patients from different genetic backgrounds with
regulate the homeostasis of this highly metabolically active and energy retinal degeneration of familial and sporadic origins has the potential
demanding tissue (Reichenbach and Bringmann, 2013). The non-neu- for (1) clinical trial in a dish, the depth and breadth of which is not fully
ronal retinal pigment epithelium (RPE), located outside the retina, yet achievable in regular clinical trials (Haston and Finkbeiner, 2016), and
in intimate contact with photoreceptors, plays a critical role in the (2) prospectively selecting patients for personalized treatment (Berkers
structural and functional viability of these cells (Strauss, 2005). et al., 2019). These findings suggest that strategies could be formulated
The loss of the visual signal when photoreceptors degenerate in age- for practical and personalized regenerative medicine with the purpose
related macular degeneration (AMD) (Curcio et al., 1996) or retinitis of recovering and preventing vision loss due to degenerative changes in
pigmentosa (RP) (Hartong et al., 2006) or the lost ability to transmit it diverse populations of patients (Fig. 1).
to the brain when RGCs degenerate in glaucoma (Almasieh et al., 2012) This review will describe the recent progress made toward ap-
invariably leads to blindness. Unfortunately, there is no effective proaches that hold promise for therapeutic regeneration in the retina.
treatment to reverse the loss of vision when photoreceptors or RGCs die. Since these approaches are underpinned by developmental


Corresponding author. Department of Ophthalmology and Visual Science, University of Nebraska Medical Center, Durham Research Center 1, Room 4034,
985840 Nebraska Medical Center, Omaha, NE, 68198-5840, USA.
E-mail address: iahmad@unmc.edu (I. Ahmad).

https://doi.org/10.1016/j.preteyeres.2019.100824
Received 16 September 2019; Received in revised form 6 December 2019; Accepted 11 December 2019
1350-9462/ © 2019 Published by Elsevier Ltd.

Please cite this article as: Iqbal Ahmad, et al., Progress in Retinal and Eye Research, https://doi.org/10.1016/j.preteyeres.2019.100824
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 1. Schematic overview of In-vivo and Ex-vivo approaches to regenerative medicine through recapitulating developmental mechanisms. Recruitment of
mechanisms (involving TFs, epigenetic regulators, and signaling pathways and their interactions) underlying retinal development is critical for unlocking the
neurogenic potential of resident MG (in vivo approach to retinal regeneration) and reproducible generation of retinal cells in 2D or 3D culture from pluripotent stem
cells for retinal repair, disease modeling, and modeling optic nerve regeneration (ex-vivo approach to retinal regeneration). Information emerging from the ex-vivo
approaches will help formulate approaches for MG-mediated therapeutic regeneration.

mechanisms, it includes a discussion on the (1) neural induction, (2) eye field (EF), which is specified in the medial anterior neural plate
development of the eye field and optic vesicle, (3) differentiation of soon after gastrulation by overlapping expression of at least five EF
RPCs into RGC and photoreceptors, (4) directed differentiation of transcription factors (EFTFs); RX, LHX2, PAX6, SIX3, AND SIX6 (Chow
pluripotent stem cells into retinal cell types for retinal repair and dis- and Lang, 2001; Wilson and Houart, 2004; Zaghloul et al., 2005; Zuber
ease modeling via 2D monolayer culture, (5) self organization of plur- et al., 2003). A variety of approaches in different species have con-
ipotent stem cells into 3D retinal organoids for retinal repair and dis- firmed their necessary roles in the development of the optic vesicle and
ease modeling, and (6) regeneration through endogenous stem-like its derivatives. For example, ectopic expression of PAX6, a homologue
cells, MG (Fig. 1). of Drosophila eyeless gene, induces ectopic eyes in Drosophila (Halder
et al., 1995) and mouse (Chow et al., 1999). Loss of PAX6 function leads
2. Development of the optic vesicle and optic cup to complete absence of the eyes in Drosophila (Quiring et al., 1994),
small eyes in mouse (Hill et al., 1991), and abnormal eye tissues in
The first morphological sign of rudimentary eyes is bilateral in- human (Glaser et al., 1992). Overexpression of SIX3, a homologue of
dentations called the optic sulci/pits in the anterior neural plate. As the Drosophila sine oculis gene, promotes the formation of ectopic optic
neural plate folds, the sulci evaginate to become bilateral optic vesicles vesicles (Lagutin et al., 2001) and retina (Loosli et al., 1999), while its
which become bilayer optic cups due to complex morphogenetic in- inactivation results in the absence of eyes and other forebrain structures
duction. The inner layer differentiates into the neural retina while the (Carl et al., 2002; Lagutin et al., 2003). Ectopic expression of RX, en-
outer layer becomes the RPE; the two layers appear separated at the coding a paired-like homeobox TF, leads to enlarged eyes (Mathers
periphery by the prospective ciliary epithelium (Adler and Canto-Soler, et al., 1997). When RX is absent (Loosli et al., 2003; Mathers et al.,
2007; Sinn and Wittbrodt, 2013; Zagozewski et al., 2014). 1997) or mutated (Voronina et al., 2004) formation of eyes is com-
promised, leading to anophthalmia. SIX6, like SIX3, is also a homologue
of the Drosophila sine oculus gene, and when overexpressed causes the
2.1. Specification of the eye field
formation of giant eyes (Zuber et al., 1999). Its absence leads to mi-
crophthalmia, due to reduced proliferation of retinal progenitors
The optic sulci are derived from a single morphogenetic field, the

2
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 2. Schematic illustration of eye field induction in the anterior neural plate. The morphogenetic field for eye development is specified in the medial anterior
neural plate by overlapping expression of EFTFs. Neural tube is demarcated by the blue enclosure, OTX2 positive anterior neural plate is identified by grey circle, and
the dark ellipse represents the presumptive area of EFTF expression. Lower panel shows complex genetic regulatory network activated at the corresponding stages
(Adapted from Zuber et al., 2003.).

(Bernier et al., 2000; Li et al., 2002). LHX2 encodes a TF belonging to by Wnt ligands and FGFs, facilitates the regionalization process. For
LIM homeodomain family. In its absence, formation of the optic pri- example, Wnt signaling, presumably driven from the extra-ocular me-
mordia is arrested at the optic vesicle stage without the formation of the senchyme, accentuates the expression of MITF and OTX2 in the dorsal
optic cup, leading to anophthalmia (Porter et al., 1997). Subsequent optic vesicle, the presumptive RPE (Fujimura, 2016; Westenskow et al.,
studies have shown that it plays a role in suppressing extra-optic neural 2009). FGF signaling, likely driven from the apposed surface ectoderm
properties, thus important for the specificity of the optic primordium (Horsford et al., 2005), either directly or indirectly through VSX2, ex-
(Roy et al., 2013) and regionalization of the optic vesicle into the tinguishes the expression of MITF in the distal optic vesicle, giving the
presumptive retina and RPE (Tetreault et al., 2009; Yun et al., 2009). cells therein the propensity to differentiate along the retinal lineage
(Horsford et al., 2005; Rowan et al., 2004). MITF, thus confined to the
2.2. Induction and maintenance of EFTFs dorsal optic vesicle, activates the downstream RPE phenotype-specific
genes in cooperation with OTX2, commiting the cells along the RPE
While the roles of EFTFs in the development of the optic vesicles lineage (Martinez-Morales et al., 2003). Subsequently, under the in-
and its derivatives are well established, how their expression is initiated fluence of adhesion molecules (Martinez-Morales et al., 2009) and re-
and maintained to delineate the EF is not well known. Using a cocktail tinoic acid (Mic et al., 2004), the optic vesicle invaginates to form the
of EFTFs, Harris and colleagues took an ingenious approach of ectopi- bilayer optic cup. For example, optic vesicles in mice lacking retinal
cally expressing single EFTFs or in different combinations to determine dehydrogenases 1a1 (ALDHA1), an enzyme that converts retinoic acid
their relative involvement in the formation of the Xenopus EF (Zuber (RA) from vitamin A, do not invaginate to form the optic cup (Mic et al.,
et al., 2003). This and other studies led to a progressive EF induction 2004). The role of the surface ectoderm, apposed to the optic vesicle, in
model (Fig. 2) where, following the induction of the neural tube under optic cup formation remains rather uncertain (Fuhrmann, 2010; Hyer
the influence of reciprocal FGF and BMP/Wnt signaling (Munoz- et al., 2003).
Sanjuan and Brivanlou, 2002), the anterior neural plate is patterned by
OTX2 for normal specification of EF (Chuang and Raymond, 2002; 2.4. Retinal differentiation
Zuber et al., 2003). OTX2 in the anterior neural plate, most likely in
combination with SOX2 (Danno et al., 2008; Heavner and Pevny, Birth dating experiments have demonstrated that the seven different
2012), helps initiate the expression of EFTFs, which is then sustained by intrinsic retinal cell types are generated in an evolutionarily conserved
cross-regulation (Zuber et al., 2003). The EF is further delineated by the temporal sequence, spanning two distinct stages of histogenesis; as a
subsequent elimination of OTX2 expression, under the negative feed- general rule, RGCs, HCs, cones, and ACs are born during early histo-
back of RX (Zuber et al., 2003) and SIX3-mediated reduced Wnt sig- genesis, while rods, BCs, and MG are born during late histogenesis
naling (Lagutin et al., 2003). The ventral forebrain SHH (Chiang et al., (Rapaport et al., 2004; Young, 1985) (Fig. 4). Lineage tracing experi-
1996), regulated likely by SIX3(Geng et al., 2008; Jeong et al., 2008) ments carried out in different species have demonstrated that these cells
splits the single EF into bilateral eye morphogenetic regions from which types, including MG, are generated from single multipotential RPCs
emerge the optic vesicles. (Holt et al., 1988; Turner and Cepko, 1987; Turner et al., 1990; Wetts
and Fraser, 1988), their developmental potential progressively re-
2.3. Regionalization of the optic vesicle and optic cup induction stricted by stage-specific competence (Cayouette et al., 2006; Cepko
et al., 1996; Livesey and Cepko, 2001). However, the stage-specific
The evaginating optic vesicles are regionalized into prospective competency may be labile because late RPCs generate early born neu-
RPE, retina, and optic stalk by differential expression of MITF, VSX2, rons when exposed to an early histogenic environment (James et al.,
and PAX2, respectively, most likely under the influence of LHX2 (Yun 2003), exposing their plasticity vis-à-vis the niche that could be tapped
et al., 2009) (Fig. 3). Intercellular signaling, predominantly mediated into for generating a variety of retinal cell types from pluripotential/

3
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 3. Regionalization of optic vesicle


and optic cup during eye development.
The evaginating optic vesicle is regionalized
into prospective RPE, retina, and optic stalk
by differential expression of MITF, VSX2,
and PAX2, respectively, under the influence
of LHX2. Signaling mediated by Wnts (from
the extra-ocular mesenchymal cells) and
FGFs (from the apposed surface ectoderm)
facilitates the regionalization process; the
former by activating MITF and OTX2 in
prospective RPE and the latter by extin-
guishing the expression of MITF from the
prospective retina directly or indirectly
through VSX2. Later, under the influence of
adhesion molecules and RA the optic vesicle
invaginates to form the bilayer optic cup
(Adapted from Fuhrmann, 2010 and Yun
et al., 2009.).

multipotential progenitors for regenerative purposes. Riesenberg et al., 2009b) and cones (Jadhav et al., 2006; Riesenberg
The possibility of recruiting developmental processes for ther- et al., 2009b; Yaron et al., 2006). However, the mechanistic link be-
apeutic repair and regeneration is further helped by the knowledge of tween Notch signaling and commitment along a specific lineage is not
intercellular signaling pathways that mediate the influence of the niche defined (Riesenberg et al., 2009b). It is rather safe to assume that at a
for RPC maintenance and differentiation (Yang, 2004). Predominant, certain point of time after a decrease in Notch signaling, the RPCs ac-
among many, is the Notch pathway (Ahmad et al., 1997; Austin et al., quire expression of lineage-specific TFs, i.e., OTX2 for photoreceptors
1995; Dorsky et al., 1995, 1997; James et al., 2004). Through its in- and ATOH7 for RGCs. Studies carried out over the last fifteen years
tercellular effectors, belonging to the HES and HEY families of tran- have revealed that a combination of TFs belonging to the basic helix
scriptional repressors, the Notch pathway keeps RPCs uncommitted, loop helix (bHLHL) and homeodomain (HD) families act in concert to
and thus undifferentiated (Tomita et al., 1996). SHH, elaborated by determine RPC fates (Hatakeyama and Kageyama, 2004; Marquardt
differentiating RGCs, facilitates RPC proliferation through Gli-2-medi- and Gruss, 2002) (Fig. 4). A theme has emerged out of these studies that
ated HES1 expression (Wall et al., 2009). Additionally, Notch signaling bHLH TFs confer a specific fate on differentiating RPCs, which acquire
can work in concert with Wnt signaling to facilitate RPCs proliferation laminar specificity under the influence of HD TFs. Evidence suggests
(Das et al., 2008). Presumably, these mechanisms maintain RPC po- that the cell type-specific TFs can be recruited under the influence of
pulation throughout histogenesis to sustain temporal differentiation of extrinsic (James et al., 2004; Parameswaran et al., 2015; Teotia et al.,
different cell types. Thus, inhibition of Notch signaling in early RPCs 2017a) and/or intrinsic (Elliott et al., 2008) factors, allowing directed
promotes differentiation of RGCs (Ahmad et al., 1997; Austin et al., differentiation of stem cells/progenitors along specific retinal sub-
1995; Dorsky et al., 1997; James et al., 2004; Nelson et al., 2006; lineages. The mechanisms underlying RGC and photoreceptor

Fig. 4. Schematic illustration of retinal histogenesis, and transcriptional codes involved in retinal cell fate specification. Retinal cells are generated in an
evolutionarily conserved sequence of early and late histogenesis. The temporal scale of histogenesis is derived from birth-dating study by LaVail and colleagues,
carried out in the developing rat retina (Rapaport et al., 2004). The tip of an arrowhead corresponds approximately to the time when 50% of that cell type is
generated in the central retina. The approximate time corresponding to cell-type generation in human is given in fetal weeks (FWKs) according to Chen et al. (2017);
Hendrickson et al. (2008); Hoshino et al. (2017). The time refers to the emergence of cells, particularly late born cells, in the fovea. Multiple TFs are involved in cell-
fate specification and differentiation of multi-potential RPCs into early (RGCs, HCs, cone photoreceptors, and ACs)- and late (rod photoreceptors, BCs, and MG)-born
cells. Evidence has emerged in favor of cell type-specific transcriptional codes, consisting of proper combinations of bHLH (red) and HD (blue) transcription factors
for cell fate specification and lamina-specific differentiation, respectively. RGCs, retinal ganglion cells; HCs, horizontal cells; ACs, amacrine cells; BC, bipolar cells;
MG, Müller glia; ONL, outer nuclear layer; INL, Inner nuclear layer; GCL, ganglion cell layer.

4
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

differentiation are covered in detail because their recapitulation pre- 1992), is absent or low cone-specific genes are activated under the
dicts the efficiency of their generation in 2D or 3D models for re- cooperative interactions between CRX and an orphan nuclear receptor
generative purposes. RORβ (Srinivas et al., 2006), promoting the differentiation of photo-
receptor precursors along the cone sub lineage. Thus, in this model the
2.4.1. Retinal ganglion cells (RGCs) cone pathway is considered the default pathway in the absence of NRL
The differentiation of RPCs along the RGC lineage initiates when a expression. Under a less understood mechanism, which may involve the
subset acquires the expression of ATOH7, a proneural bHLH tran- influence of the altered niche due to the presence of cones, NRL ex-
scription factor, which confers competency for RGC specification pression is upregulated at a later stage of histogenesis, presumably by
(Brzezinski et al., 2012; Yang et al., 2003). ATOH7 is positively regu- RORβ, this time cooperating with other TFs known to be involved in
lated by PAX6 (Riesenberg et al., 2009a) and inhibited by Notch sig- photoreceptor such as ASCL1 (Ahmad, 1995) and NEUROD1 (Acharya
naling (Miesfeld et al., 2018) and VSX2 (Vitorino et al., 2009). Though et al., 1997; Morrow et al., 1999). NRL in cooperation with CRX induces
ATOH7 confers RGC competency on the post-mitotic precursors, it does rod-specific genes (Pittler et al., 2004; Yoshida et al., 2004), and in
not ensure their terminal differentiation into RGCs because ATOH7 cooperation with orphan nuclear receptor NR2E3 (Oh et al., 2008)
expression has been observed in the lineages of all other retinal cell suppresses S opsin and other cone genes, thus shunting the generic
types (Brown et al., 2001; Kay et al., 2001; Le et al., 2006; Ma et al., photoreceptor precursors away from the default cone pathway. There-
2004; Wang et al., 2001; Yang et al., 2003). Therefore, ATOH7 positive fore, in the absence of NRL, it is thought that the generic photoreceptor
cells serve as early generic precursors capable of generating different and rod precursors acquire the default pathway and differentiate into S
retinal cell types unless they express RGC-specific TFs that promote cones (Mears et al., 2001).
terminal differentiation and survival and suppress regulators for non- Precursors which are differentiating along the cone lineage under
RGC cell types (Le et al., 2006; Yang et al., 2003). Toward this end, the influence of TRB2, a thyroid hormone regulating TF, express M
ATOH7, presumably at a particular threshold of expression levels and/ opsin to become M cones (Ng et al., 2001). The role of extrinsic factors
or in concert with other TFs, activates the expression of three specific on the transcriptional dominant model is not well known, except that of
classes of HD-containing TFs that are necessary for RGC terminal dif- thyroid hormone in the specification of M cones (Lu et al., 2009).
ferentiation and survival. These include POU-homeodomain factor However, several in vitro studies of retinal progenitors (Ahmad et al.,
POU4F2 (BRN3B), LIM Homeodomain factor ISL1, and DLX-home- 1999; Altshuler and Cepko, 1992; Levine et al., 2000; Watanabe and
odomain factors DLX1/DLX2. In the absence of each of the transcription Raff, 1992) and neural progenitors derived from pluripotent stem cells
factors, RGCs are generated but do not survive beyond certain em- (Zhao et al., 2002) demonstrated the influence of niche on the devel-
bryonic stage; RGC loss calculated to be 80%, 67%, and 33% in opment of photoreceptors, specifically the rods. Among known rod fa-
POU4F2−/− (Erkman et al., 1996; Gan et al., 1996), ISL1 conditional cilitative factors are taurine (Altshuler et al., 1993; Young and Cepko,
knockout (CKO) (Pan et al., 2008), and DLX1/DLX2−/− (de Melo et al., 2004), activin A (Davis et al., 2000), SHH (Levine et al., 1997), FGF
2005) mice. Besides influencing differentiation and survival, these TFs (McFarlane et al., 1998), and RA (Kelley et al., 1999; Khanna et al.,
play an important role in axon growth and pathfinding (de Melo et al., 2006). In contrast, LIF (Graham et al., 2005) and CNTF (Ezzeddine
2005; Erkman et al., 2000; Pan et al., 2008). Their cooperative inter- et al., 1997) inhibit photoreceptor differentiation. This information was
actions in regulating the terminal differentiation of RGCs and whether critical in increasing the efficiency of photoreceptor generation in 2D
these happen in distinct subpopulations of ATOH7 positive precursors is and 3D cultures of pluripotent stem cells (see below).
poorly understood. A variety of approaches have demonstrated that
RGC development is influenced by cell-extrinsic factors, such as FGFs 3. 2D monolayer culture for generating target cells for repair and
(Martinez-Morales et al., 2005; McCabe et al., 1999), SHH (Neumann disease modeling
and Nuesslein-Volhard, 2000; Zhang and Yang, 2001), and GDF11 (Kim
et al., 2005). The use of these factors for stage-specific recruitment of The practical and ethical barriers associated with obtaining human
signaling pathways was instrumental in directed differentiation of RPCs-present in active form only in the fetal retina-for potential clinical
pluripotent stem cells along the RGC lineage (Parameswaran et al., applications necessitated examining whether self-renewing pluripotent
2015; Teotia et al., 2017a). stem cells can be a viable source of RPCs from which retinal neurons
can be generated (Ahmad, 2001). The directed generation of target cells
2.4.2. Photoreceptors by recapitulating developmental mechanisms in a monolayer culture
In both mice and humans, initiation of cone and rod generation facilitated preclinical examinations of ex-vivo stem cell therapy for
takes place during early histogenesis but their maturation continues blinding diseases (Gasparini et al., 2019). Additionally, and more im-
postnatally (Cornish et al., 2004; Hendrickson et al., 2008; Young, portantly, it provided a facile platform for modeling diseases to un-
1985) (Fig. 4). The transcriptional dominant model of photoreceptor derstand the mechanisms underlying degenerative changes in specific
cell fate determination (Swaroop et al., 2010) is informative and in- retinal cell types, which can provide information to formulate new
structive for obtaining cones and rods for regenerative medicine pur- approaches toward therapeutic regeneration (Liu et al., 2018).
poses. This model includes six TFs (RORβ, OTX2, NRL, CRX, NR2E3 and
TRβ2), whose contextual change in the threshold of expression de- 3.1. Generation of RPCs
termines whether the photoreceptor precursors will differentiate along
the cone or rod lineage. The process of photoreceptor differentiation The early evidence regarding the retinal potential of ES cells was the
begins when a subset of RPCs acquires the expression of HD tran- observation that cells in the mouse ES cell-derived embryoid bodies
scription factor OTX2, as it has been shown that CKO of OTX2 leads to (EBs), when exposed to FGF2 and ITSFn neural induction medium
the loss of both cones and rods (Nishida et al., 2003). OTX2 confers the (Okabe et al., 1996), began to express PAX6 and RX (Zhao et al., 2002).
generic photoreceptor potential on precursors by activating CRX, an Some of these cells co-expressed RPC marker VSX2, and when cultured
OTX-like homeobox gene, which may be important for overall differ- in conditions simulating late histogenesis, they differentiated along the
entiation of photoreceptors, but not for determining the specific fates of rod lineage, as ascertained by the expression of rod-specific markers
photoreceptor precursors (Furukawa et al., 1999; Nishida, 2005). (Zhao et al., 2002). However, in the absence of a strategy for directed
According to the transcriptional dominant model (Swaroop et al., differentiation of EBs along the retinal lineage, the number of EFTF
2010), at some point before committing to one sub lineage or the other, expressing cells were few and therefore, retinal differentiation in-
the generic photoreceptors express varying levels of the six TFs. When efficient. Subsequently, two approaches for generating RPCs from
the expression of NRL, a basic motif-leucine zipper TF (Swaroop et al., pluripotent stem cells were developed, one based on the developmental

5
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 5. Directed differentiation of pluripotent stem cell-derived retinal progenitor cells into RGC and photoreceptors. This entails neural/retinal induction,
which can be achieved either by passive (Meyer et al., 2009; Pankratz et al., 2007) or active (Lamba et al., 2006; Ying et al., 2003) recruitment of the default neural
potential of pluripotent stem cells. In both cases neural rosettes (NR) are generated with cells expressing EFTFs. These retinal progenitors are known to spontaneously
differentiate along the RGC or photoreceptor lineages. However, their efficiency of differentiation along a particular lineage can be enhanced by specific small
molecules or through the recruitment of stage-specific developmental mechanisms (Teotia et al., 2017b).

principle in which one allows the passive manifestation of the default (e.g., FGF4) and ES cell markers (e.g., OCT4 and SOX2), and activation
neural potential of the embryonic ectoderm (Munoz-Sanjuan and of SOX1, PAX6, and OTX2 expression that coincided with the NR for-
Brivanlou, 2002; Stern, 2005), and the other uses its active recruitment mation. The authors called this neural induction period the “primitive
by influencing the underlying pathways (Fig. 5). These are covered in anterior neuroepithelium” stage, which was followed by the “definitive
some detail because of their implications in the specificity and fidelity neuroepithelium stage,” consisting of cells expressing genes corre-
of the generation of retinal cells in either 2D or 3D culture. sponding to anterior neural progenitors, i.e., OTX2 and BF1, including
EFTFs. This model of neural induction did not examine the involvement
3.1.1. Passive manifestation of default neural potential of the endogenous pathways of the default mechanism, while Meyer
This approach to generate RPCs (Meyer et al., 2009; Zhao et al., et al. who used the exact protocol to generate retinal progenitors did
2002) had precedence in studies demonstrating that ES cells, when (Meyer et al., 2009). They observed that when hESCs were differ-
grown in monolayer (Ying et al., 2003) as single cells (Smukler et al., entiated into NR, 95% of cells therein were PAX6 and RX positive by the
2006; Tropepe et al., 2001) or as aggregates (Pankratz et al., 2007; 10th day into the adherent culture conditions (Meyer et al., 2009). That
Reubinoff et al., 2001; Watanabe et al., 2005) acquire neural properties the acquisition of EFTFs expression involved endogenous inhibition of
predominantly of anterior neural plate cells under conditions that are BMP and Wnt pathways was demonstrated by the upregulation of the
supportive for cell survival. The neural differentiation occurred in- expression of their inhibitors, noggin and DKK1, respectively. The loss
dependent of media supplements such as N2 (Thermo Fisher Scientific) of PAX6 and RX expression when the FGF pathway was pharmacolo-
and B27 (Thermo Fisher Scientific) that are known to support differ- gically inhibited demonstrated its involvement, similar to observations
entiation and survival of neurons (Pankratz et al., 2007; Ying et al., that the activation of SOX1 in mouse ES cells was dependent on FGF
2003) or adherence inductive substratum (Pankratz et al., 2007; signaling (Ying et al., 2003). When NRs were mechanically triturated
Tropepe et al., 2001; Watanabe et al., 2005) further demonstrating the and grown in suspension in retinal differentiation medium containing
acquisition of neural properties by default. That the intrinsic pathways FGF2, they generated neurospheres (Meyer et al., 2009). Cells in the
(BMP and FGF pathways) underlying the default neural induction was neurospheres gradually excluded the expression of MITF and acquired
engaged endogenously was revealed by an increase in the expression of the expression of VSX2 in a process reminiscent of the differentiation of
BMP signal antagonists noggin and follistatin, and a decrease in neural the inner layer of the optic cup into retinal primordium (see above),
induction when FGF signaling was pharmacologically inhibited (Ying whereas if NRs were left adherent they acquired MITF1 expression and
et al., 2003). differentiated into RPE (Meyer et al., 2009).
Similar default neural induction was reported in human ES cells. For A comparative analysis of results obtained by Pankratz et al. (2007),
example, when human ES cells were cultured for a prolonged time, foci and Meyer et al. (2009), using protocols which were almost identical,
of differentiated cells were observed expressing PAX6 and NCAM where the latter examined human ES cell differentiation along the
(Reubinoff et al., 2001). Mechanical trituration of these foci and their retinal lineage, and the former along the pan neural lineage, demon-
suspension culture in the presence of B27, bFGF and EGF generated strate the nature of cells in the NRs and their influence on the efficiency
spheres, where more than 90% of cells expressed Nestin, NCAM, and of the directed differentiation along the retinal lineage. These cells had
glial progenitor marker, A2B5 (Reubinoff et al., 2001). Pankratz et al. effectively silenced genes corresponding to endoderm (AFP) and me-
developed an in vitro model of default neural induction by culturing soderm (BRACHYURY) and activated expression of pan neuroecto-
aggregates of human ES cells, first in suspension culture (4 days in dermal genes (SOX1, PAX6, SOX3, ZIC1, NCAD, CHURCHILL) and
ESCM + 2 days in N2 supplemented medium) and then on the adherent anterior neural plate genes (OTX2, BF1) including the rest of EFTFs.
surface of laminin for 10 days where neural rosettes (NR) appeared at Unlike neural induction in vivo, PAX6 expression in these cells preceded
day 4 after plating (Pankratz et al., 2007). In this model, neural in- that of SOX1 (Pankratz et al., 2007). The expression of these genes in
duction was regulated; it was preceded by silencing of the ICM markers the NR suggests that the NR cells represent generic anterior neural plate

6
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

cells in which they co-express multiple transcription factors including resulting NRs were mechanically triturated and cultured on poly-D-ly-
EFTFs. Their anterior neural plate property may still be malleable since sine and laminin substratum in the presence of noggin and FGF2 for
exposure of these cells to RA induced posterior neural tube character- another 25 days to promote differentiation along the retinal lineage.
istics, as demonstrated by the activation of HOXB4 gene (Pankratz The following observations suggested the recruitment of developmental
et al., 2007). This malleability suggests that these cells may have retinal mechanisms toward sequential programming of EB cells along the
potential by virtue of the expression of EFTFs, however for them to give neural and then retinal lineage upon extrinsic instructions: (1) a pro-
rise to functional retinal cells the immature/hybrid properties of these gressive attenuation of germ line markers, OCT4 (ectoderm), SOX17
progenitors/precursors need to be extinguished. This may require spe- (endoderm), and BRACHYURY (mesoderm) and accentuation of ante-
cific culture conditions to engage signaling pathways that are specific to rior neural plate marker OTX2, and (2) a progressive decrease of OTX2
retinal development. Therefore, the spontaneous differentiation of and increase in the expression of EFTFs during retinal induction phase.
neural/retinal progenitors into retinal cells and RPE in generic culture That these changes were directed and not random, a reflection on the
conditions necessitates robust characterization of the differentiated fidelity of the method, was demonstrated by the global gene expression
cells not only for the presence of cell type specific markers, but also the of the ES cells at the end of the retinal induction, which was comparable
absence of immature markers, necessary for the stability and non tu- (r2 = 0.90) to that of RPCs isolated from early stages of retinal histo-
morigenicity of the transplanted cells for therapeutic purposes. genesis. Despite the similarity between iPSC-derived RPCs and em-
bryonic/fetal retinal cells, the fact that these cells also expressed some
3.1.2. Active manifestation of default neural potential pluripotency genes and those belonging to diencephalon (Lamba et al.,
This approach involved differentiating mouse and human ES cells 2010; Parameswaran et al., 2015) suggested that the transcriptomes of
along the neural/retinal lineage by active engagement of the signaling these cells were not yet hard-wired along retinal lineage. Therefore, the
pathways that facilitate the default neural induction (Lamba et al., evidence suggests that refinement of RPC generation holds the key to
2006, 2010; Osakada et al., 2009b; Parameswaran et al., 2010, 2015). increasing efficiency of the generation of target cells such as photo-
The approach was preceded by demonstrations by several labs that receptors and RGCs for regenerative medicine.
inhibition of pathways that are antagonistic to neural induction in vivo,
particularly BMP and Wnt signaling, facilitates neural differentiation in 3.2. Generation of RGCs
mouse and human ES cells (Aubert et al., 2002; Lindsley et al., 2006;
Pera et al., 2004; Tropepe et al., 2001; Ying et al., 2003). In addition, Different approaches have been taken to obtain RGCs from plur-
recruitment of FGF signaling was observed to promote neural differ- ipotent stem cell-derived RPCs. Zack's group used CRISPR-Cas9 genome
entiation of ES cells, consistent with its role in neural induction editing technology to create RGC reporter human ES cell lines in which
(Tropepe et al., 2001; Ying et al., 2003; Zhang et al., 2001) (Stavridis flurorescent reporter mCherry (Sluch et al., 2015) or tdTomato (Sluch
et al., 2007)). The protocol developed by Lamba et al. was reproducible et al., 2017) was expressed from endogenous POU4F2 promoters to
and widely adopted; EBs were generated from human ES cells in the provide a real time readout of RGC differentiation. Their initial ma-
presence of knockout serum, B27 supplement, DKK and noggin for 3 trigel-based protocol was based on passive neural induction and spon-
days to neuralize EB cells (Lamba et al., 2006). IGF1 was included to taneous differentiation of RPCs into RGCs, hence the efficiency of RGC
promote the induction of the rudiments of the eye, based on the ob- generation was low (3–5% of mCherry+ POU4F2+ cells) and the cul-
servation that injection of IGF1 in Xenopus embryos leads to the for- ture was mixed with cells expressing other retinal cell-type specific
mation of ectopic heads and eyes (Pera et al., 2001). The neuralized EBs transcripts (Sluch et al., 2015). However, temporal activation of RGC
were cultured on matrigel in “retinal determination” conditions, cre- regulatory genes was observed and electrophysiological examination of
ated by the presence of N2 and B27 supplements plus higher con- cells displayed trains of action potential in response to depolarization
centrations of DKK, noggin, and IGF1, in addition to FGF2 for 1–3 currents, an electrophysiological signature of RGCs (Sluch et al., 2015).
weeks. The EBs gave rise to NRs, which expressed EFTFs and retinal The addition of forskolin, an activator of adenylate cyclase, during the
progenitor marker VSX2. More importantly, this method demonstrated initial stage of the culture generated significantly more RGCs, preceded
that IGF signaling is evolutionarily conserved for the induction of the by increased activation of EFTF genes, which suggested that the im-
anterior neural plate cells based on the increased expression of EFTFs proved efficiency was likely due to better facilitation of neural induc-
when cultured in the presence of DKK + noggin + IGF, as compared to tion in the presence of forskolin. Zack and his colleague amended the
DKK + noggin or in the absence of inducers. Interestingly, DKK and shortcomings in their next protocol (Sluch et al., 2017) by resorting to
noggin alone did not increase the expression of EFTFs versus non-in- active neural induction by inhibiting BMP signaling and activating
ducer controls, suggesting that the culture conditions may not allow Nodal signaling, in the presence of forskolin and nicotinamide, re-
optimal inhibition of BMP4 and Wnt signaling. Therefore, examining spectively, during the initial stage of culture. This alteration increased
the role of culture conditions, particularly the matrigel, in engaging the efficiency of RGC generation to 20%. Inhibition of Notch signaling
BMP4 and Wnt signaling pathways may further help increase the effi- by DAPT later during the culture increased the efficiency further up to
ciency of the generation of RPCs. Nevertheless, a strong correlation 50%. The fidelity of RGC differentiation, measured against the presence
(r2 = 0.90) existed between the global gene expression of human ES of other retinal cell types, and the electrophysiological signature of
cell-derived RPCs and 90 post conception days (PCD) human fetal ret- RGCs were not examined. However, these protocols demonstrated the
inal cells, attesting to the fidelity of the approach to induce retinal importance of proper neural induction for efficient generation of retinal
properties (Lamba et al., 2010). This method was improvised by seg- cell types in general, and RGCs in particular.
regating retinal induction from neural induction, followed by differ- The method developed by our lab (Teotia et al., 2017a) was based
entiation of retinal progenitors along photoreceptor and RGC lineages on recapitulating the developmental mechanism underlying RGC dif-
(Parameswaran et al., 2015). EBs were generated from the mouse iPS ferentiation, which can be divided into three phases: initiation, differ-
cells in the presence of high concentration (10 ng/ml) of DKK and entiation, and maturation. Signaling pathways underlying each of these
noggin for 3 days to prime cells for neural induction. Neural induction stages were recruited using small molecules and/or recombinant li-
was achieved by culturing EBs in a supportive medium (ITSFn, B27, N2 gands. For example, it has been observed that transient FGF signaling
supplements) containing noggin. DKK was removed to promote the by FGF8 and FGF3 facilitates initial RGC differentiation in the central
proliferation of neural progenitors under the influence of endogenous retina (Martinez-Morales et al., 2005; McCabe et al., 1999). SHH is
Wnt signaling (Das et al., 2008). FGF2 was not included at this stage as thought to be similarly involved. However, beyond the initiation of
its activation before gastrulation is known to prevent anterior neural RGC differentiation, SHH may promote proliferation, and thus maintain
plate cells from expressing a retinal fate (Moore et al., 2004). The RPCs for subsequent differentiation (Martinez-Morales et al., 2005;

7
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Zhang and Yang, 2001). A coordinated decrease in Notch signaling is demonstrated that recapitulating the photoreceptor development me-
essential for RPCs to commit along the RGC lineage (James et al., 2004; chanisms could not only reduce the time of differentiation from months
Nelson et al., 2006). Based on these observations, the initiation of RGC to weeks, but also facilitate preferential differentiation of RPCs along
differentiation in vitro included transiently activating FGF and Shh cone or rod sub-lineages. For example, Ali and his colleagues adapted
signaling and inhibiting Notch activities for the first 48 h in culture, as the passive neural induction protocol and demonstrated that exposure
outlined in Fig. 5. Subsequently, during the differentiation phase, FGF8 of human ES cell-derived RPCs to a cocktail of RA, taurine, aFGF and
was removed from the culture and cyclopamine added to inhibit SHH bFGF generated 36% of NRL+ cells by day 30 in culture (Boucherie
elaborated by the nascent RGCs that could adversely affect differ- et al., 2013). This culture condition effectively silenced the expression
entiation. Notch signaling, like SHH signaling, was kept inhibited to of cone-specific genes in the favor of those regulating (NRL) and
prevent any drift of committed precursors back into the proliferative characterizing (RHODOPSIN) rods. Bernier and colleagues adapted the
mode. To keep the committed precursors on the RGC differentiation active neural induction approach and demonstrated that COCO
track, TGFβ pathway was left inhibited in the presence of follistatin, (DAN5), a member of Cerberus/Dan family of secreted inhibitors of
given the observation that activation of the TGFβ pathway by GDF11, BMP, TGFβ and Wnt pathways, in combination with IGF-1, generated
secreted by differentiating RGCs, inhibits RGC differentiation (Kim 60–80% of cells with the S-cone phenotype from human ES cell-derived
et al., 2005). The survival of nascent RGCs depends upon neurotrophins RPCs within 4–5 weeks, presumably through the default S-cone
to prevent the activation of programmed cell death (PCD) (Guerin et al., pathway (Swaroop et al., 2010; Zhou et al., 2015). The generated cones
2006). Thus, to facilitate RGC maturation without excessive cell death, were functional, as they could degrade cGMP upon light stimulation,
BDNF, NT4, and CNTF, all known to prevent PCD in RGCs (Cohen et al., incorporate into host retina upon transplantation, and elaborate outer
1994; Ji et al., 2004; Meyer-Franke et al., 1995; Spalding et al., 2004), segments. Bernier's group further demonstrated that the S-cone phe-
were included with general promoters of cell survival, forskolin and notype was malleable and that in the presence of thyroid hormone, the
ROCK inhibitor (Lingor et al., 2008; Meyer-Franke et al., 1995). This genesis of M-cones could be facilitated at the expense of S-cones, thus
led to recapitulation of hierarchical expression of RGC regulators when providing additional evidence of generating specific subtypes of cells by
the method was tested on enriched rat RPCs and mouse ES cells, with recapitulating the developmental mechanism (Zhou et al., 2015)
~60% efficiency (POU4F2+βTUBULIN+ cells). The efficiency of RGC (Fig. 5).
generation decreased by half to ~30% when the method was adapted to
human iPSC cells. It was observed that in human cells, addition of cy-
clopamine during the differentiation phase inhibited differentiation, 3.4. Retinal repair
presumably due to difference in the duration of RGC genesis in humans
and mice. The inhibition was overcome by removing cyclopamine and 3.4.1. RGC degeneration
extending the exposure to SHH over the differentiation phase, which Retinal repair for RGC degeneration entails transplantation of RGCs
increased the RGC differentiation up to 40% (Teotia et al., 2019). The derived from pluripotent stem cells, their lamina-specific integration,
hiPSC-derived RGCs fired action potentials upon injection of depolar- survival, formation of synapses with BCs, and most importantly, their
izing currents and most importantly, from a regeneration perspective, ability to elaborate guidable axons that can navigate out of the retina
expressed a battery of guidance molecules essential for intra- and extra- toward central targets (Table 1). Based on these criteria, transplanta-
retinal navigation of their axons to reach to their central targets. That tion of pluripotent stem cell-derived RGCs in the retina has met with
these RGCs have the capacity to discriminate between specific and non- limited success. For example, when rodent iPSC-derived RGCs, labeled
specific targets was demonstrated by their ability to elaborate axons with tracking dye, carboxyfluorescein deacetate (CFDA), were trans-
toward superior colliculus (SC) and retract from inferior colliculus (IC) planted in rat model of RGC degeneration due to ocular hypertension
(Teotia et al., 2017a). (Morrison et al., 1997), they integrated in the host's RGC layer but
displayed rudimentary neurites, distributed apically toward the inner
3.3. Generation of photoreceptors plexiform layer (Parameswaran et al., 2015). Whether or not they
formed synapses with BCs remained unexamined. It has been demon-
It has been observed that prolonged culture of human iPSC-derived strated recently that mTOR signaling plays a significant role in RGC
RPCs in supportive medium (N2 and B27 supplement) allowed them to neuritogenesis, including the maintenance and regeneration of den-
differentiate along the photoreceptor lineage (Lamba et al., 2006). The drites (Morquette et al., 2015; Teotia et al., 2019). When human iPSC-
efficiency of differentiation, however, was low for rods (~4% rho- derived RGCs, in which mTOR signaling was experimentally activated,
dopsin+ cells) and even lower for cones (0.01% S-opsin+ cells). Several were transplanted in the neonatal retina, they incorporated into the
labs adopted the directed differentiation approach to address this bar- host's RGC layer and elaborated significantly longer but tortuous
rier for practical evaluation of cell replacement therapy and disease neurites, compared to control RGCs (Fig. 6). In the case of either rodent
modeling for photoreceptor degeneration. For example, Takahashi and (Parameswaran et al., 2015) or human (Fig. 6) iPSC-derived RGC
her colleague systematically evaluated the effects of various activators transplantation, neurites extending toward the optic disc were not ob-
and inhibitors of pathways involved in photoreceptor differentiation in served.
mouse ES cell-derived RPCs (Osakada et al., 2008). When they exposed The lack of successful outcome of RGC transplantation could be due
human ES cell-derived RPCs to RA and taurine, factors that promote to several reasons that may include the developmental stage of the
terminal differentiation of photoreceptors (Altshuler et al., 1993; Kelley transplanted RGCs and/or the inhospitable environment of the adult
et al., 1999; Young and Cepko, 2004), they observed an increase in the retina, which may not favor axon and dendrite development. That the
efficiency of the generation of rods (~8% rhodopsin+ cells), s-cones endogenous retinal environment may not be inhibitory was shown by
(~9% s-opsin+ cells) and m-cones (~9% L/M-opsin+ cells) over a Goldberg's group; it was observed that one month post-transplantation
period of 150–200 days in culture (Osakada et al., 2008). Lako and in adult rat retina, the GFP-expressing neonatal rat RGCs integrated into
colleagues modified this protocol by adding activin A, SHH, and T3 and the host retina and acquired morphology of the host's RGCs, extending
observed that 18% of the cells expressed rhodopsin, 52% expressed S- neurites toward the optic nerve head and elaborating dendrites into the
opsin, and 60% expressed M-opsin at 45 days in culture (Mellough inner plexiform layer (Venugopalan et al., 2016). GFP+ axons were
et al., 2012). However, the number of these de novo generated photo- seen in the optic nerve, at the optic chiasm, and in the central targets
receptors, however, decreased precipitously by day 60 in culture such as the superior colliculus (SC) and lateral geniculate nucleus (LGN)
(Mellough et al., 2012). (Venugopalan et al., 2016).
Further modification of the directed differentiation approach

8
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Parameswaran et al. (2015)

Wu et al. (2018)

Li et al. (2017)
Reference

RGCs incorporated into the host RGC layer and expressed RGC-specific markers.

- Enhanced RGC survival was observed in RGC-iPSC cotransplantations to adult


- An increase in RGC survival was observed in RGC-iPSC direct cocultures and
- Transplanted in the rat model of ocular hypertension, the de novo generated

- In addition, RGCs with iPSC cotransplantation extended significantly longer

- Technique to produce the engineered human RGC-scaffold biomaterial.


- Assessed the transplantation method for biomaterial in vivo.
neurites than RGC-only transplants.
RGC-iPSC indirect cocultures

retinas ex vivo and in vivo.


Outcome

- Dissociated RGCs (1 × 105) seeded onto scaffold coated with


2.5% Matrigel injected into the intraocular environment.
- Intravitreal transplantation of iPS cell-derived RGCs.

- Purified mouse GFP+ RGCs with or without iPSCs

Fig. 6. Survival and integration of hiPSC-derived RGCs (hRGC) with acti-


Transplanted population/Injection route

vated mTOR signaling in neonatal rat retina. mTOR signaling was activated
in human iPSC-derived RGCs by silencing the expression of tuberous sclerosis
complex 2 (TSC2), an inhibitor of the mTOR complex 1, by lentivirus mediated
transplanted intravitreally.

expression of TSC2 shRNA (A). hRGCs with activated mTOR signaling and
controls, both expressing GFP, were transplanted intravitreally in postnatal day
1 (PN1) rat pups. Examination of the retinal sections, two weeks post trans-
plantation, revealed the survival and integration of grafted cells in the host's
RGC layer, where hRGCs with activated mTOR pathway were observed to
elaborate long, albeit tortuous, apical neurites toward host's inner nuclear layer,
Retinal transplantation of pluripotent stem cell-derived RGCs.

compared to controls. Vertical arrow indicates active mTOR signaling. Scale


bar = 50 μm.

3.4.2. Photoreceptor degeneration


The requirement for retinal repair through cell therapy for photo-
10–12wk Sprague Dawley
Morrison's rat model of

receptor loss is similar to that of RGCs, but relatively less daunting


Rabbit and Monkey
ocular hypertension

because the synaptic connections between the graft and the host need to
occur only within the retina. The two most important challenges be-
sides lamina specific integration and survival for functional recovery
are: development of the outer segment and synapse formation by the
Host

grafted cells for visual signal generation and transmission, respectively.


rats

In the last fifteen years, several approaches to retinal transplanta-


3D Culture System

tion with the aim to replace degenerated photoreceptors have been


Mouse/Human PSC

2D culture system

taken and these are covered in detail in two excellent reviews


(Gagliardi et al., 2019; Gasparini et al., 2019). Here, we will briefly
line used

cover a few approaches that have given new direction to this field
Table 1

miPSC

hiPSC

hiPSC

(Table 2). The most prominent among those was Ali and his colleague's,
which demonstrated that post-mitotic rod precursors from postnatal

9
Table 2
I. Ahmad, et al.

Retinal transplantation of pluripotent stem cell-derived photoreceptors.


Cell type Host Transplanted population/injection method Outcome Reference

2D culture system
hESCs adult Crx−/− mice − 50,000–80,000 GFP-labeled 3wk-old retinal cells - Average of 3,000 Nrl+ human cells in host retinas Lamba et al. (2009)
- Intravitreal and subretinal injections via trans-scleral route - Light-response restored in eyes receiving donor cells
hiPSCs WT mice treated with cyclosporine - ~50,000 4–8wk GFP-labeled retinal cells - ~ 50 cells/eye migrated to ONL and expressed Otx2, recoverin and Lamba et al. (2010)
- Subretinal injection via trans-scleral route rhodopsin
hESCs 4–6 week old WT mice − 50,000 unsorted retinal progenitors - Cells in the subretinal grafts expressed and survived without Hambright et al. (2012)
- Epiretinal and subretinal injections via trans-vitreal route immunosuppression
- The epiretinal grafts survived, but did not express markers of mature
retinal cells.
- Retinal integration into the retinal ganglion cell (RGC) layer and the inner
nuclear layer (INL) was efficient from the epiretinal grafts, but not
subretinal grafts
hiPSCs P4 immune-compromised Rag1−/− x − 150-day old photoreceptor precursor cells differentiated from - Transplanted cells matured and expressed recoverin Tucker et al. (2013a)
Crb1−/− mice autosomal reccessive RP (USH2A mutant) patient-specific hiPSCs - Cells developed axonal and inner/outer segment-like projections
- Subretinal injection
hESCs WT PN1 mice ~10,000 2–4wk retinal cells - A fraction of cells (~1.4%) that underwent differentiation for 2 weeks, but Zhou et al. (2015)
- intravitreal injection no longer, could migrate into layers of host retina and expressed s-opsin
hiPSCs adult WT and rd1 mice treated with - ~200,000 photoreceptor precursors − 5.7% of iPSC-PhRPs survived in subretinal space of rd1 mice 3 weeks post- Barnea-Cramer et al.
cyclosporine - Subretinal injection via trans-scleral route transplantation (2016)
hESCs adult WT and IL2ry−/− mice; adult - ~500,000–750,000 retinal cells - Average of 20,900 donor cells integrated within ONL of IL2ry−/− host Zhu et al. (2017)
Crxtvrm65 and IL2ry−/−/Crxtvrm65 mice - Subretinal injection via trans-corneal route and expressed pan-photoreceptor markers and mature markers such as
PNA, Ret-P1 and GCO, and light sensitivity was restored

10
- Donor cells integrated poorly in WT and non-immune-compromised Crx
mutant mice
- ~4,300 cells integrated into IL1ry/Crx mice and expressed OTX2 and
rhodopsin

3D culture system

hiPSCs SCID mice - Clinical-grade ipsc-derived organoids dissociated and injected - No sign of tumor formation in any of the 48 mice, nor evidence of Wiley et al. (2016)
intramuscularly into hind limbs infection or systemic immune reaction
− 10 mice received subretinal injections of 1 × 106 cells
hESCs, hiPSCs adult Nrl−/− mice; Aipl1−/− mice - FAC-sorted L/M cones from weeks 14–17 of differentiation − 50% of transplanted eyes contained hPSC-derived cones that remained in Gonzalez-Cordero et al.
- Injected into superior retina subretinal space near ONL in Nrl−/− mice (2017)
- Cells formed distinct layer adjacent to host INL in Aipl1−/− mice.
hiPSCs adult IL2ry−/− mice - ~500,000–750,000 clinical grade retinal cells − 0.15% of transplanted cells integrated into host retina, exhibited Zhu et al. (2018)
- Subretinal injection via trans-corneal route photoreceptor orientation and morphology, and expressed Otx2,
recoverin, and rhodopsin
hiPSCs P23H rats treated with cyclosporine − 300,000 unsorted or MAC-sorted CD73 + cells from D120 - unsorted cells remained in subretinal space of host as heterogeneous Gagliardi et al. (2018)
organoids population
- Subretinal injection via transvitreal route - CD73 + sorted cells survived and coexpressed RECOVERIN and S121
hESC WT and rd1 mice - unsorted dissociated cells and sorted double negative CD29/ - CD29-/SSEA1-sorted cells that expressed human RECOVERIN were Lakowski et al. (2018)
SSEA-1 cells -subretinal injection detected in the SRS either as cell clumps or aligned with the host ONL
hESCs Pde6brd1 mice treated with cyclosporine - FAC-sorted CRX-GFP+ photoreceptor precursors - CRX+ photoreceptor precursors integrated into ONL and made synaptic Collin et al. (2019)
- Subretinal injection via transvitreal route connections with host bipolar cells
hiPSCs Cpfl1/Rho−/− and rd1 mice - Jaws+ (optogenetically enhanced) photoreceptors from retinal - Jaws-expressing donor cells observed near host INL and expressed Garita-Hernandez et al.
organoids RECOVERIN (2019)
- subretinal injection - Human nuclear antigen (HNA) staining confirmed minimal (5%) of PMID: 31586094
GFP+ cells were due to material transfer
- Light response restored
Progress in Retinal and Eye Research xxx (xxxx) xxxx
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

day 4–7 mouse retina, as opposed to embryonic cells, integrate and Primary open angle glaucoma (POAG), characterized by an un-
survive within the host retina when transplanted in the subretinal space obstructed (=open) iridocorneal angle and cupping of the optic nerve
(MacLaren et al., 2006). They acquired the morphology of mature rods head with corresponding vision loss, is the most common of all glau-
with well-defined outer segments and were connected synaptically, and coma subtypes, (Kwon et al., 2009). Although elevated intraocular
therefore functionally, with the host retina. However, almost a decade pressure (IOP) is regarded as the most important risk factor for devel-
later, studies revealed the phenomenon of cytoplasmic material ex- oping POAG, a significant number of patients with POAG have normal
change between the donor and host cells, the former labeling the latter, IOP, and yet their optic nerves degenerate (Iwase et al., 2004; Kwon
providing the impression that the donor cells have differentiated, ma- et al., 2009). This type of POAG, where the onset of the disease is not
tured, and integrated (Pearson et al., 2016; Santos-Ferreira et al., associated with increased IOP, is called normal tension glaucoma
2016a; Singh et al., 2016). These observations have prompted re-as- (NTG). One of the familial forms of NTG is associated with the E50K
sessment of previous findings and caution in the interpretation of cur- missense mutation in the optineurin (OPTN) gene (Rezaie et al., 2002),
rent and future transplantation outcomes, with approaches such as which is responsible for encoding a scaffold protein involved in mem-
fluorescence in situ hybridization (FISH) for X/Y chromosome to rule brane trafficking and exocytosis (Bond et al., 2011; Sahlender et al.,
out the cytoplasmic material exchange, if donor and hosts were of 2005).
different sex (Gasparini et al., 2019; Nickerson et al., 2018). The study Iwata and his group using iPSC-derived generic neurons from POAG
of the transplantation outcomes of human photoreceptors became patients carrying the E50K OPTN mutation, demonstrated that the
practical with the development of methods to obtain these cells from mutation renders the protein insoluble, causing intercellular traffic
human pluripotent stem cells. Reh's and colleagues were the first to disturbance and eventually cell death (Minegishi et al., 2013). Another
transplant human ES-cell derived retinal cells into neonatal and adult familial form of NTG is associated with the duplication of the Tank
mice retina (Lamba et al., 2009). The GFP-labeled cells integrated and binding kinase-1 (TBK1) gene (Fingert et al., 2011). Tucker and col-
survived in the neonatal retina and expressed rod/cone-specific mar- leagues demonstrated that RGC-like cells derived from TBK1 patient-
kers. When transplanted subretinally in adult adult CRX−/− mice, a specific iPSCs expressed high levels of an autophagy marker LC3-II
model of Leber's Congenital Amaurosis (LCA) in which photoreceptors protein, compared to controls, suggesting that enhanced autophagy
fail to develop outer segments, some of the cells integrated in the host may underlie RGC degeneration (Tucker et al., 2014). It is important to
retina, accompanied by partial restoration of light response as judged note that the normal maintenance of the intracellular traffic involves a
by electroretinogram (ERG), although the transplanted cells did not cross talk between OPTN and TBK-1; enhanced interaction between
elaborate outer segments (Lamba et al., 2009). OPTN and TBK-1 due to E50K mutation, may contribute to the in-
With the advent of retinal organoid technology, which promises a solubility of OPTN. Therefore, in addition to rendering TBK-1-regulated
higher efficiency of generation of rods and cones than 2D technology, it cellular autophagy abnormal, this may lead to a cellular burden that
became important to develop approaches to sort and enrich these cells ultimately culminates in degeneration (Minegishi et al., 2013). The
among different cell types and from proliferating cells that can lead to modeling thus substantiated a notion that mutations in disparate genes
teratoma formation (Tucker et al., 2011a). One of the advancements may affect a common biological function, and therefore degeneration
made in this direction was the characterization of plasma membrane could be rescued by targeting that function. For example, Iwata's group
protein cluster of differentiation 73 (CD73) for the selection of trans- observed that pharmacologically inhibiting TBK-1 abrogated the ab-
plantation-competent photoreceptor cells from mouse retina (Eberle normal insolubility of OPTN due to E50K mutation (Minegishi et al.,
et al., 2011; Lakowski et al., 2011), mouse ES cells (Lakowski et al., 2013). However, Iwata and colleagues modeled the disease in generic
2015) and human iPSC-derived retinal organoids (Gagliardi et al., neurons and Tucker's group in RGC-like cells at a single time point of
2018) (see below) generation. Therefore, their studies could not examine the effects of
mutations on the development, morphology, or function of RGCs that
3.5. Disease modeling in 2D monolayer culture would have shed light on the development of the disease.
Our lab approached modeling POAG on the premise that the
The directed differentiation of pluripotent stem cells into diverse common pathology of RGC degeneration in complex disorders like
cell types by recapitulating developmental mechanisms in 2D cultures, glaucoma is due to the developmental susceptibility of these cells; i.e.,
combined with facile generation of iPSCs from patients’ somatic cells that they are intrinsically vulnerable (Teotia et al., 2017b). Using the
(Shi et al., 2017) has led to the development of a reproducible human chemically defined method that recapitulates stage-specific mechan-
disease model. This helped overcome the limitation of animal models isms underlying RGC development, we generated RGCs from POAG
that did not recapitulate human disease faithfully because of the fun- patient-specific iPSCs. These cells were re-programmed from the per-
damental inter species differences. Human disease modeling has shed ipheral blood of POAG patients carrying the missense variant
light on dysregulated genes and pathways underlying the degenerative (rs33912345; C > A; His141Asn) in the exon of a developmentally
pathology that can be targeted for therapeutic regeneration and repair. regulated gene, SIX6 (Carnes et al., 2014; Iglesias et al., 2014;
For example, dopaminergic (DA) neurons generated from Parkinson Skowronska-Krawczyk et al., 2015). SIX6 is a member of the home-
Disease (PD) patient-specific iPSCs recapitulated pathological features odomain TF family and is required for proper eye development
associated with degeneration such as mitochondrial dysfunction, oxi- (Anderson et al., 2012) (see above). A number of studies have identified
dative stress, ER stress, and alpha synuclein accumulations (Sison et al., the association of SIX6 risk allele variant (rs33912345; C > A; Hi-
2018). More importantly, it was demonstrated that the genetic cor- s141Asn) with POAG (Carnes et al., 2014; Iglesias et al., 2014; Rong
rection of the mutation in LRRK2 gene, the most common genetic cause et al., 2019). Moreover, the SIX6 locus with risk allele has been asso-
of PD, not only rescued the pathological phenotype of patient-specific ciated with thinning of the retinal nerve fiber layer (RNFL) (Carnes
DA neurons but also revealed that the ERK pathway was dysregulated, et al., 2014; Cheng et al., 2014; Yoshikawa et al., 2017), a major pa-
and its pharmacological inhibition ameliorated the PD-associated de- thological change in glaucoma. We demonstrated that the efficiency of
generative changes (Reinhardt et al., 2013). RGC differentiation was significantly reduced and accompanied by
suppressed expression of RGC regulatory genes in SIX6 risk allele RGCs
3.5.1. Modeling glaucomatous RGC degeneration compared to those derived from healthy, age- and sex-matched controls
Disease modeling is particularly suited for examining the molecular (Teotia et al., 2017b). The SIX6 risk allele RGCs exhibited numerous
and cellular basis of glaucoma, because it is not a single homogeneous abnormalities such as short and simpler neurites, elevated expression of
disease but a group of multifactorial diseases with a unifying pathology glaucoma-associated genes and dysregulation of genes related to de-
of RGC degeneration (Almasieh et al., 2012; Janssen et al., 2013). velopmentally relevant biological process and pathways, and immature

11
I. Ahmad, et al.

Table 3
Disease modeling using iPSC-derived RGCs for optic neuropathies.
Disease Gene Mutation Primary cells Reprogramming method Phenotypes and Rescue Reference

2D culture system

Normal Tension OPTN E50K PBMCs (Blood) Sendai Virus (OKSM) - E50K mutant is insoluble and was associated with the hydrophobic precipitate in Minegishi et al. (2013)
Glaucoma lysates in E50K iPSCs and E50K iPSC-derived neural cells.
- The abnormal insolubility of the E50K mutant was rescued by treatment with a
TBK1-specific inhibitor.
Normal Tension TBK1 780 kb duplication 12q14 Fibroblasts Sendai Virus (OKSM) An extra copy of the TBK1 gene leads to activation of a critical autophagy protein Tucker et al. (2014)
Glaucoma (LC3-II) in patient-specific retinal ganglion cells.
Primary open angle SIX6 rs33912345; PBMCs (Blood) Sendai Virus (OKSM) - The efficiency of RGC generation by SIX6 risk allele iPSCs was significantly lower Teotia et al. (2017b)
glaucoma C > A; His141Asn than iPSCs-derived from healthy, age- and sex-matched controls.
- The SIX6 risk allele RGCs displayed short and simple neurites, reduced expression
of guidance molecules, and immature electrophysiological signature.
- Increased expression of glaucoma-associated genes, CDKN2A and CDKN2B.
- RNA seq analysis enriched developmentally relevant biological processes and
signaling pathways involved in differentiation, metabolism, and survival.

3D culture system

12
Normal Tension OPTN E50K Fibroblasts mRNA (OKSM) - Undifferentiated OPTN iPSCs exhibited a disorganized Golgi morphology that led Ohlemacher et al. (2016)
Glaucoma to a significant increase in size.
- Caspase-3 activation was significantly increased in OPTN iPSC-derived RGCs
compared to control iPSC-derived RGCs.
- Treatment of OPTN iPSC-derived RGCs with either BDNF or PEDF partially
rescued these cells from apoptosis.
Dominant optic atrophy OPA1 intron24 c.2496 + 1 G > T Fibroblasts Retrovirus (OKSM) - Mutant OPA1-derived iPSCs exhibited a significant increase in apoptosis. Chen et al. (2016)
- Efficiency of OPA1 mutant-derived iPSCs into RGCs was increased in the presence
of Noggin, or estrogen.
LHON MT-ND4 m.4160T > C and Fibroblasts Episomal (OKSM/Lin28/ - Demonstrated generation of isogenic iPSC controls by replacing LHON mtDNA Wong et al. (2017)
m.14484T > C shRNAp53) using cybrid technology.
- RGC differentiation demonstrated increased cell death in LHON-RGCs and could
be rescued in cybrid corrected RGCs.
LHON MT-ND4 m.11778G > A PBMCs (Blood) Sendai Virus (OKSM) - Defective neuritogenesis and neurite maintenance in LHON patient-derived Wu et al. (2018)
RGCs.
- Increased apoptosis and mitochondrial biogenesis in RGCs differentiated from
LHON patient-derived hiPSCs.
- LHON-affected mitochondria have decreased spare respiratory capacity.
congenital glaucoma CYP1B1 c1403_1429dup Fibroblasts Sendai Virus (OKSM) -n/a Bolinches-Amoros et al.
(2018)
Progress in Retinal and Eye Research xxx (xxxx) xxxx
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

electrophysiological signature (Teotia et al., 2017b). Together, these axons was compromised. The microfluidic model provided the first
results suggested that the SIX6 risk allele influences RGC differentiation proof of principle that the recapitulation of developmentally relevant
and ultimately leads to abnormalities at cellular, molecular, and func- pathways can be a viable clinical approach for axon regeneration in
tional levels, which, if carried into adulthood, may make these cells glaucoma.
vulnerable to degenerative changes (Table 3).
The examination of glaucomatous RGC abnormalities in controlled
conditions offers an opportunity to model glaucomatous optic nerve 3.5.2. Modeling photoreceptor dystrophy
degeneration/regeneration. Though significant progress has been made Patient-specific hiPSCs have proven especially useful in modeling
in identifying genes and pathways involved in optic nerve regeneration genetically heterogeneous photorecepotor dystrophies, such as retinitis
(Oh et al., 2018; Park et al., 2008; Smith et al., 2009; Wang et al., pigmentosa (RP), an inherited disease caused by mutations in at least
2018b; Zhang et al., 2019) the clinical relevance of these studies re- 100 different genes (Daiger et al., 2013). RP can be autosomal domi-
mains poorly established because they have been carried out in rodents. nant, recessive dominant, x-linked, or sporadic (Ayuso and Millan,
Modeling optic nerve regeneration using hiPSC-derived RGCs may 2010; Ferrari et al., 2011). In 2011, Takahashi and her colleagues re-
identify molecular targets for therapeutic optic nerve regeneration. ported the first disease modeling of RP using iPSC technology (Jin et al.,
Recently, we established a microfluidic model in which hRGC axons are 2011). In that study iPSCs were generated from the fibroblasts of RP
cultured directionally away from the soma in microgrooves toward patients with mutations in RP1/RP9/PRPH2/RHO gene, involved in
target cells to examine the role of mTOR signaling in optic nerve re- photosensitivity and rod outer segment development, followed by their
generation (Teotia et al., 2019). The mTOR pathway, an ubiquitous directed differentiation into rod photoreceptors in 2D culture systems
nutrient sensing signaling mechanism that promotes growth and sur- (Jin et al., 2011). While cells from both control and RP patient-specific
vival of cells by regulating protein synthesis (Costa-Mattioli and cell lines similarly differentiated to mature rods by day 120, con-
Monteggia, 2013), represents an evolutionarily conserved regulatory tinuation of the culture led to their decreased survival in mutant cell
mechanism underlying RGC development (Teotia et al., 2019). Besides lines compared to controls at day 150. The decrease in survival was
regulating RGC differentiation, it facilitates the expression and function associated with oxidative and ER stresses, suggesting a pharmacological
of guidance receptors, necessary for axon pathfinding, and presumably, approach of using antioxidant alpha-tocopherol to promote survival.
regeneration (Teotia et al., 2019). We found that this pathway, which This approach resulted in varying response from different patient-spe-
has been demonstrated to facilitate optic nerve regeneration in rodents cific cell lines, with pathological phenotype rescued in only RP9-pa-
(Park et al., 2008), was observed to be dysregulated in SIX6 risk allele tient-specific rods (Jin et al., 2011). Modeling RP due to mutation in
POAG patient-specific RGCs (Teotia et al., 2017b). To test the premise RHO genes (E181k), Yoshida et al. observed reduced survival of patient-
that the effect of mTOR signaling on optic nerve regeneration is evo- specific rods compared to controls, which was attributed to ER stress
lutionarily conserved, axons of hRGCs were chemically axotomised in (Yoshida et al., 2014). Genetic correction of the E181K mutation re-
the microfluidic device. A robust regeneration of axons was observed in sulted in a normal cellular phenotype, while inducing the mutation in
hRGC in which mTOR signaling was activated by shRNA-mediated si- control lines resulted in the disease phenotype in iPSC-derived rods.
lencing of tuberous sclerosis complex 2 (TSC2), compared to controls The group went on to utilize this disease model for screening possible
(Teotia et al., 2019) (Fig. 7). TSC2 inhibits mTOR complex 1, therefore therapeutic agents that interfere with the ER stress pathway; treatment
silencing it activates the mTOR pathway (Costa-Mattioli and of an E181K mutant line with either mTOR inhibitors (rapamycin and
Monteggia, 2013). Conversely, when mTOR1 complex was pharmaco- PP242), an AMPK activator (AICAR), an ASK1 inhibitor (NQDI-1), or a
logically inhibited by rapamycin regeneration of axotomised hRGC protein synthesis suppressor (salubrinal) increased the yield of photo-
receptors and reduced the ER stress associated with mutant rhodopsin

Fig. 7. Modeling human optic nerve regenera-


tion: A schematic representation of a microfluidic
model of chemical axotomy and regeneration of
optic nerve, where RGCs (hRGCs) are generated from
human iPSCs in the soma chamber of a microfluidic
device. Axons, traversing through the microgrooves
are axotomised by saponin, followed by their re-
generation back into the axon chamber. B-D: Control
(hRGCGFP), and experimental (hRGCTSC2shRNA:GFP)
hRGCs in which mTOR signaling was activated by
shRNA-mediated silencing of TSC2, an inhibitor of
mTOR complex1, were subjected to axotomy and
allowed to regenerate. Following regeneration axons
were labeled retrogradely with CTB-CY3 to account
for regeneration of GFP+ axons. A robust regenera-
tion (an increase in the number of GFP+CY3+ axons
in the axon chamber post-axotomy) was observed in
hRGCTSC2shRNA:GFP group (activated mTOR
pathway), compared to controls. Vertical arrow in-
dicates active mTOR signaling. Scale bar = 50 μm.

13
Table 4
Disease modeling for photoreceptor dystrophies.
Disease Gene Mutation Primary cells Reprogramming method Phenotypes and Rescue Reference
I. Ahmad, et al.

2D Culture System

Retinitis pigmentosa RP1, RP9, RP1: 721Lfs722X Fibroblasts Retroviral transduction - Decreased number of rod cells. Jin et al. (2011)
PRPH2, RHO RP9: H137L PRPH2: W316G RHO: G188R - Patient-derived rod cells exhibit high expression of markers for
oxidative or ER stress.
- Treatment with α-tocopherol was beneficial to RP9-rod
photoreceptor survival, and caused different effects on
rhodposin-positive cells derived from different patients.
Retinitis pigmentosa MAK Alu insertion in exon 9 Fibroblasts Lentiviral transduction - Loss of retina-specific MAK isoform Tucker et al.
(2011b)
Retinitis pigmentosa RHO G188R Fibroblasts Sendai virus - Diffused distribution of RHO protein in cytoplasm and Jin et al. (2012)
expressions of endoplasmic reticulum (ER) stress markers in
patient-derived rod cells.
- Significant decrease in rod cell numbers after successive culture
demonstrated rod degeneration.
Retinitis pigmentosa RHO E181K Fibroblasts Retroviral transduction - Reduced survival rate in the photoreceptors and increased Yoshida et al.
expression of ER stress and apoptotic markers. (2014)
- Using helper-dependent adenoviral vector (HDAdV) gene
transfer, mutation was corrected in the patient's iPSCs and also
introduced into control iPSCs.
- Rapamycin, PP242, AICAR, NQDI-1, and salubrinal promoted
the survival of the patient's iPSC-derived photoreceptor cells
Leber congenital amaurosis CEP290 Autosomal recessive CEP290 Fibroblasts Lentiviral transduction - Fewer and shorter cilia were formed in patient derived cells Burnight et al.
than in cells from unaffected individuals. (2014)
- Lentiviral delivery of CEP290 rescued the ciliogenesis defect.

14
3D Culture System

Retinitis Pigmentosa USH2A Arg4192His Keratinocytes Sendai virus - Increased expression of protein misfolding and ER stress Tucker et al.
markers (2013b)
Microphthalmia VSX2 R200Q Activated T cells Retroviral transduction - VSX2 hiPSC-OVs demonstrated growth retardation and Phillips et al.
preferential differentiation toward an RPE fate. (2014)
- VSX2 hiPSC-OVs demonstrated delayed photoreceptor
maturation.
- Exogenous expression of WT VSX2 reduced RPE production and
enhanced photoreceptor development in (R200Q) VSX2 hiPSC-
OVs
Enhanced S-cone syndrome NR2E3 Homozygous for IVS2-2 A > C fibroblasts Sendai Virus - Patient derived retinal organoids recapitulated enhanced S- Wiley et al.
cone phenotype (2016)
- Robust expression of SW opsin, and lack of rod photoreceptors
Leber congentital amaurosis CEP290 Homozygous c.2991 + 1655A > G Fibroblasts Integration-free - CEP290 not detectable at connecting cilia in LCA optic cups Parfitt et al.
(LCA) episomal vectors - Reduced cilia incidence and length (2016)
- Increased abberant splicing in photoreceptors compared to
other cell types
- Antisense Morpholino treatment reduced aberrant splicing and
restored ciliogenesis in LCA RPE and photoreceptors
Retinitis Pigmentosa RP2 c.519C > T (p.R120X) fibroblasts Integration-free - Reduced expression of Kif7 at cilia tips of photoreceptor outer Schwarz et al.
episomal vectors segments (2017)
- Restoration of Kif7 through translational read-through inducing
drugs (TRIDs) such as PTC124
Retinitis Pigmentosa RPGR g.ORF15 + 689-692del4 Fibroblasts Lentiviral transduction - Increased actin polymerisation and phosphorylation of Megaw et al.
cytoskeletal regulatory proteins due to amelioration of actin- (2017)
severing protein Gelsolin
- Activated gelsolin rescued RPGR-deficient cells
(continued on next page)
Progress in Retinal and Eye Research xxx (xxxx) xxxx
Table 4 (continued)

Disease Gene Mutation Primary cells Reprogramming method Phenotypes and Rescue Reference
I. Ahmad, et al.

2D Culture System

Retinitis pigmentosa TRNT1 Truncated version of the protein Fibroblasts Sendai virus - Patient-specific iPSCs and iPSC-derived retinal organoids Sharma et al.
exhibited a deficit in autophagy. (2017)
- Aberrant accumulation of LC3-II and elevated levels of
oxidative stress.
Leber congenital amaurosis and CEP290 CEP290-LCA Fibroblasts Lentiviral transduction - Optic cups derived from induced pluripotent stem cells (iPSCs) Shimada et al.
Joubert-syndrome and CEP290-JSRD of CEP290-LCA patients displayed less developed photoreceptor (2017)
related disorders cilia.
- Lack of CEP290 in JSRD fibroblasts resulted in abnormal cilia
and decreased ciliogenesis.
- Reduced localization of GTPase protein required for ciliogenesis
(ADCY3 and ARL13B).
- Hedgehog signaling was augmented in CEP290-JSRD.
Retinitis pigmentosa RPGR Patient 1: exon 14, c.1685_1686delAT patients 2 Urinary cells Lentiviral transduction - Patient-specific photoreceptors displayed defects in Deng et al. (2018)
and 3: ORF15, c.2234_2235delGA and morphology, localization, transcriptional profiling, and
c.2403_2404delAG electrophysiological activity.
- Shorted cilium was found in patient iPSCs, RPE cells, and three-
dimensional retinal organoids.
- CRISPR-Cas9-mediated correction of RPGR mutation rescued
photoreceptor structure and electrophysiological property,
reversed the observed ciliopathy, and restored gene expression
Enhanced S-cone syndrome NR2E3 Patient 1: c.119-2A > C Fibroblasts Sendai virus - Patient-derived retinal cells expressed a mutant NR2E3 Bohrer et al.
Patient 2: p.Arg73Ser p.Arg311Gln transcript that contained a portion of intron 1, which caused a (2019)
frameshift and the creation of a premature stop codon.
- The expression of the wild-type NR2E3 transcript was restored

15
by monoallelic CRISPR correction.
Retinal Dystrophy CRB1 n/a Fibroblasts Lentiviral transduction - CRB1 patient iPSC retinal organoids showed disruptions at the Quinn et al.
OLM as found in Crb1 mutant mice, and developed retinal (2019)
degeneration
- AAV5 serotype, combined with CMV-promoter mediated
expression was efficient for transducing CRB gene expression in
photoreceptors and Müller glial cells in adult and iPSC-derived
retina.
Progress in Retinal and Eye Research xxx (xxxx) xxxx
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 8. Schematic illustration of two different 3D culture methods for retinal organoid generation. A: Suspension culture (Nakano et al., 2012): In this
approach, both aggregation of pluripotent stem cells and retinal induction are achieved in suspension culture in the presence ROCK inhibitor (Y27632) and Wnt
inhibitor (IWR1e), followed by organoid formation also in suspension, augmented in the presence of FBS, SHH inhibitor (SAG), and a Wnt activator (CHIR99021). B:
Suspension-adherent-suspension culture (Zhong et al., 2014): In this approach, retinal induction is achieved in adherent culture after aggregation of pluripotent stem
cells in suspension in the presence of blebbistatin. This is followed by organoid generation in suspension, augmented in the presence of FBS, taurine and retinoic acid
for photoreceptor differentiation. Abbreviation: KSR, knockout serum replacement; FBS, fetal bovine serum; SAG, smoothened agonist.

(Yoshida et al., 2014) (Table 4). niche cells. For similar reasons, neither specific cell types nor their post-
mitotic precursors, generated in 2D systems for retinal repair may be
able to functionally integrate and replace degenerating neurons. The
4. 3D organoid culture for generating target cells for repair and emerging field of organoid technology holds promise to address these
disease modeling significant barriers to the practical aspects of regenerative medicine. Its
origin is traced back to the beginning of the 20th century, when sponge
The differentiation of RPCs and the survival of their progeny takes cells were observed to self-organize into whole sponges (Wilson, 1907),
place in the intimate microenvironment created by retinal cells in dif- and its integration into regenerative medicine was established when
ferent stages of development and non-neural cells that include astro- breast epithelial cells were found to generate 3D structures resembling
cytes, endothelial cells, and microglia. For example, years ago Reh and secretory alveoli secreting milk proteins (Li et al., 1987). Organoid
Tully observed that neurotoxin-mediated ablation of tyrosine hydro- technology involves the self-organization of pluripotent stem cells, or
xylase (TH) positive amacrine cells in the developing frog retina led to progenitors derived from specific tissues, into 3D structures that possess
the generation of a greater than normal number of TH positive ama- the rudiments of structures and functions displayed by the tissue in vivo
crine cells, while the number of other amacrine cell types remained when cultured in conditions that favor 3D growth. Here, the conditions
unchanged, thus providing one of the earliest evidence of the influence defined by Lancaster and Knoblich for the generation of an organoid is
of niche on RPC differentiation (Reh and Tully, 1986). Subsequently, informative: First, the organoid must contain multiple cell types of a
with the advent of transgenic animal technology, Calof and colleagues specific tissue; second, these cells are spatially organized as in vivo; and
provided the mechanism underlying the influence of the niche on the third, it displays some function specific to that tissue (Lancaster and
generation of retinal cells (Kim et al., 2005). They observed that Knoblich, 2014). The retinal organoid fulfills all three of these condi-
GDF11, a member of the transforming growth factor-β superfamily, tions.
presumably secreted by differentiated RGCs, regulates the competence
of early RPCs to generate RGCs. Accordingly, their number increased
significantly when GDF11 was genetically knocked out in mice (Kim 4.1. Retinal organoid
et al., 2005). Such a complex environment cannot be fully replicated in
the 2D culture system, and therefore a specific cell type may not The advent of organoid technology for the brain in general and
achieve the full complement of molecular, morphological, and func- retina in particular, is owed to the pioneering work of Sasai and his
tional differentiation needed for understanding the degenerative pa- colleagues, who successfully demonstrated the self-organization of
thology and drug screening. The efficacy of a drug to rescue a neuron mouse ES cells into polarized cortical tissues containing layer-specific
undergoing degeneration in a 2D model system may not pan out in vivo, neurons whose temporal generation mimicked corticogenesis in vivo
where the effects of the drug are likely to be modulated by the other (Eiraku et al., 2008; Sasai et al., 2012). Three years later, Sasai's group

16
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 9. Schematic illustration of temporal generation of neurons in retinal organoids. The generation of retinal neurons in organoids follows an evolutionarily
conserved temporal sequence observed in vivo. The schematic is adapted from Zhong et al. (2014) and Capowski et al. (2019). The approximate time corresponding to
cell-type generation in human is given in fetal weeks (FWKs) according to studies by (Chen et al., 2017; Hendrickson et al., 2008; Hoshino et al., 2017). Broken arrow
denotes disorganization/degeneration of RGCs in stage 3 of the culture (Capowski et al., 2019). RGCs, retinal ganglion cells; HCs, horizontal cells; ACs, amacrine
cells; BC, bipolar cells.

demonstrated that by changing the culture conditions, particularly by generation of retinal cells in the human organoids followed the evolu-
adding matrigel during the aggregation of ES cells to facilitate the tionarily conserved temporal sequence observed in vivo; the generation
formation of a rigid continuous epithelium, while simultaneously re- of photoreceptors was preceded by RGCs. Variations of this method to
ducing the concentration of the knock out serum from 10% to 1.5%, generate retinal organoids from human pluripotent stem cells have
organoids can be generated with histogenic potential of the retina emerged since. For example, it was demonstrated that the efficiency of
(Eiraku et al., 2011). By day 5 in suspension culture, the cells in the retinal organoid generation can be increased if pluripotent stem cells
developing organoid segregated into RX+ SOX1- and RX− SOX1+ were exposed to increasing concentrations of BMP4 during the early
epithelia, with retinal and non-retinal neural potential, respectively. By stage of aggregate formation, as it facilitated their differentiation into
day 7, the RX+ SOX1- epithelium evaginated, and by day 9 it in- RPCs (RX+ and VSX2+ cells) (Kuwahara et al., 2015). In a significant
vaginated, a process resembling the formation of the optic cup in vivo. variation, Cantor-Soler's group introduced an intermediate adherent
The distal part of the optic cup, expressing RX, VSX2, and PAX6 re- culture stage, similar to that prevalent in 2D culture for neural/retinal
sembled the rudiment of the developing retina, and the proximal part induction (see above), between suspension culture stages for generating
expressing MITF and PAX6 appeared to be the prospective RPE. When aggregates of pluripotent stem cells and RPCs and organoids, respec-
optic cups were excised and cultured in suspension for another 10 days, tively (Zhong et al., 2014) (Fig. 8). This approach led to the generation
the distal epithelium became stratified, containing all retinal cell types of optic cups with 50–70% efficiency without exogenous modulation of
organized with laminar specificity. Importantly, cell birth followed the Wnt and SHH signaling. Briefly, human iPSCs were aggregated in the
general pattern of histogenesis in vivo: cells expressing RGC markers presence of blebbistatin, a non-muscle myosine IIA inhibitor known to
appeared first, while those expressing rod markers appeared later. prevent dissociation associated apoptosis in pluripotent cells (Chen
However, the epithelium was not stable beyond 35 days in culture. The et al., 2010; Ohgushi et al., 2010; Walker et al., 2010), instead of ROCK
adaptation of this method to generate retinal organoids from human inhibitor. The aggregates, after 7 days in suspension culture with N2
pluripotent stem cells required Sasai's group to modify their culture supplement and heparin, were seeded on matrigel-coated dishes in
conditions, due to the species-specific difference in adhesive properties medium containing B27 supplement for 2–3 weeks. During this time,
of cells, tissue size, and gestation time (~260 days in humans versus the aggregates morphed into complex neural rosette-like structures, the
~20 days in mice) (Nakano et al., 2012) (Fig. 8). Besides increasing the centers of which contained cells expressing EFTFs, surrounded by
number of cells seeded/well from 3000 to 9000 and adding ROCK in- SOX1+ cells, reminiscent of the anterior neural plate with demarcated
hibitor to prevent apoptosis, the modification included adding Wnt eye field in vivo (Zuber et al., 2003). The central RPC domains, ac-
inhibitor to counter the caudalizing effects of increased concentration quiring horseshoe shapes over time, were manually excised and cul-
of KSR (20%) and reduced concentration (1%) of matrigel for gen- tured in suspension in the presence of B27 supplement for the genera-
erating EB-like aggregates. Furthermore, induction of retinal properties tion of 3D retinal organoids. Within the stratified layers of the optic
in the epithelium required activation of SHH signaling and the presence cups retinal cell types were generated in an evolutionarily conserved
of 10% FBS. Unlike mouse retinal organoids, induction of the RPE layer central to peripheral gradient and temporal sequence (Rapaport et al.,
required stage-specific exposure to a Wnt agonist, based on the devel- 2004; Young, 1985) (Fig. 9). More importantly, from the disease-
opmental requirement of Wnt signaling (Westenskow et al., 2009). As modeling viewpoint this method is one of the few (Dorgau et al., 2019;
expected, the development process of the optic epithelium, its stratifi- Lowe et al., 2016; Parfitt et al., 2016) that led to the reproducible
cation, and generation of retinal cell types were significantly delayed generation of photoreceptors with rudimentary outer segments and
compared to that in the mouse retinal organoids. This barrier was photosensitivity (Zhong et al., 2014). Recently, it has been reported
somewhat mitigated by engaging a specific developmental signal (e.g., that RGCs generated in retinal organoids achieve a semblance of sub-
Notch signaling) that acts as a gatekeeper to differentiation; small type diversity, characteristic of RGCs in vivo (Langer et al., 2018).
molecule inhibition of Notch signaling accelerated differentiation. The The approach to retinal organoids through the optic vesicle/optic

17
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

cup stage has been met with variable success. This drawback has led to Within the self-sorted domain, spatially restricted lineage commitment,
innovation in methods that circumvent the optic vesicle/optic cup which is likely influenced by the changing niche as new cells are gen-
stage, leading to increased efficiency in the generation of retinal or- erated, may create a histological structure similar to that in vivo. For
ganoids. Karl and colleagues (Volkner et al., 2016) generated embryoid example, the emergence of the subdomains of RPE (RX− MITF+ cells)
body-like aggregates from mouse ES cells and cultured them for 10 days and retina (RX+ VSX2+ cells) within the ocular domain (RX+, PAX6+
as described by Sasai's group (Nakano et al., 2012). The mother orga- cells) may initiate due to the activation of Wnt signaling in progenitors,
noids were manually trisected and cultured in suspension in retinal which are biased toward RPE lineage (RPE precursors). A positive
maturation medium consisting of N2 supplement and 10% fetal calf feedback of Wnt signaling as MITF collaborates with LEF-1 to activate
serum (FCS), which led to remarkably efficient generation (~180% if Wnt target genes may further accentuate the RPE phenotype (Yasumoto
one considered each mother organoid trisected to be 100%) of stratified et al., 2002). In contrast, the retinal domain may emerge by suppressing
retinal organoids by day 21 in culture. Although no optic cups formed, Wnt signaling in retinal progenitors through SIX3, which keeps the
the epithelium was distinguished into prospective retina (RX+, VSX2+, expression of Wnt ligands suppressed (Liu et al., 2010), and DKK, a
MITF− cells), surrounded by the prospective RPE (MITF+, RX−, Vsx2- diffusible Wnt inhibitor (Fuhrmann, 2008).
cells). Temporal analysis of retinal cell type-specific genes from day 7 The cellular diversity within the retinal subdomain may emerge
(mother organoid stage) to day 21 (retinal maturation stage) demon- through several rounds of spatially restricted lineage commitment,
strated recapitulation of temporal aspects of retinal histogenesis; the along with other mechanisms that may include the symmetrical and
onset of RGC-specific gene expression preceded that of rods. Further- asymmetrical cell division of retinal progenitors/precursors. For ex-
more, they demonstrated that stage specific inhibition of the Notch ample, it has been observed that asymmetrical and symmetrical in-
pathway could enrich the retinal organoids with either cones or rods heritance of Numb protein, which is dependent upon the plane of
(Volkner et al., 2016), recapitulating the in vivo effect of Notch sig- cleavage of cell division, predicts whether the progenies will acquire
naling on photoreceptor differentiation (Jadhav et al., 2006). similar (photoreceptors) and dissimilar (photoreceptors and MG) phe-
notypes, thus sorting out fates in the same, or different layers
(Cayouette and Raff, 2003). This mechanism may also be influenced by
4.2. Self-organization of retinal organoid
the changing niche, as symmetrical and asymmetrical inheritance of
Numb is influenced by the environment (Dooley et al., 2003). The la-
The mechanism behind self-organization of organoids remains
minar sorting of cells in the retinal subdomain is likely to be aided by
poorly understood (Sasai, 2013). However, a scheme proposed by
cell type-specific adhesion molecules and/or their differential expres-
Lancaster and Knoblich can be adopted as a framework to build upon a
sion.
putative mechanism (Lancaster and Knoblich, 2014). At a simpler level,
the self-organization of stem cells into an organoid may involve self-
sorting and spatially restricted lineage commitment. Self-sorting im- 4.3. Retinal repair
plies that the differential adhesion properties of different cell types
within the stem cell aggregates (=embryoid body like structures), Retinal organoids are a rich source of target cells for retinal repair.
mediated by cell surface adhesion proteins, sort cells with similar ad- Single cells, or sheets of cells obtained from organoids when trans-
hesion properties into specific domains (Fig. 10). This notion is sup- planted into rodent eyes survive, incorporate, mature as photo-
ported by the important role adhesion proteins play in the development receptors, and in some instances functionally integrate within the host
and maintenance of the structural organization of tissues in vivo. For retina and lead to visual recovery (Assawachananont et al., 2014;
example, when N-cadherin is neutralized by antibodies, the embryonic Gagliardi et al., 2018; Gonzalez-Cordero et al., 2013; Kruczek et al.,
retina dissociates and loses its 3D structure (Matsunaga et al., 1988). 2017; Santos-Ferreira et al., 2016b). However, the cellular

Fig. 10. Schematic illustration of self-or-


ganization of retinal progenitors into
organoids. In a highly simplified scheme,
RPCs (green) segregate in the embryoid
bodies from other cells (yellow) based on
the differential expression of cell adhesive
molecules (dark green bars in green cells
and brown bars in yellow cells). RPCs by
default generate RGCs first as it happens in
vivo and they segregate in a rudimentary
lamina aided by the expression of cell-ad-
hesive molecules and this process is re-
iterated for the next cell types in changing
micro-niche with the emergence of new
cells, presumably aided by asymmetrical
division of RPCs, facilitated by Notch-Numb
mechanism. (Figure adapted from Lancaster
and Knoblich, 2014.)

18
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

heterogeneity of organoids consisting of proliferating and differentiated a requirement of personalized treatment through clinical trials in a dish
cells other than the target cells raises the risk of teratoma formation and for complex diseases (Berkers et al., 2019; Haston and Finkbeiner,
disruption of the normal structure of the host retina, respectively. 2016). This was recently demonstrated in the context of cystic fibrosis
Therefore, the success of organoid-based repair and regeneration re- (CF), a disease in which the lungs and digestive system cannot function
quires strategies to enrich target cells. properly due to mutations in the cystic fibrosis trans-membrane con-
Currently, two different approaches have been taken to enrich ductance regulator (CFTR) chloride channel gene (Riordan et al., 1989).
photoreceptor precursors for transplantation, both taking advantage of Van der Ent and colleagues, using a functional CFTR intestinal organoid
developmentally relevant mechanisms or factors. Ali and his colleagues assay, in which wild type organoids swell in response to Forskolin
generated organoids from mouse ES cells transduced with the AAV-GFP unlike CF organoids (Dekkers et al., 2013), observed a high correlation
reporter driven by the human red green cone opsin promoter (Kruczek of the effects of drugs on CF organoids and CF patients, demonstrating
et al., 2017). To increase the proportion of cones (FACS sorted GFP+ that organoid-based drug testing may be a reliable predictor of clinical
cells) for transplantation, generation of these cells in organoid was fa- responses in complex diseases (Berkers et al., 2019).
cilitated by a short exposure to retinoic acid (RA), based on previous
observations that RA is associated with cone opsin expression during 4.4.1. Glaucomatous RGC degeneration
retinal development (Alfano et al., 2011). These enriched cones (~15% Although a reliable in vitro glaucoma disease model using organoids
of total cells) were transplanted in the subretinal space of Aipl−/− mice, is awaited, Meyer's group developed a retinal neurosphere model of an
a model of end-stage photoreceptor degeneration, where they survived inherited form of glaucoma using hiPSC from a patient with a mutation
and displayed some semblance of maturity (Kruczek et al., 2017). in the optineurin gene (OPTN E50K) (Ohlemacher et al., 2016). The
However, this enrichment approach based on viral transduction of re- study demonstrated that patient-derived RGCs (POU4F2+ and ISLET1+
porter constructs is not clinically practical. The practical alternative is cells) in neurospheres exhibited a dramatic increase in apoptosis, which
to enrich cells based on endogenous cell surface markers. Watanabe and could be rescued by the addition of neuroprotective factors
colleagues, who made seminal contributions toward characterizing (Ohlemacher et al., 2016). However, the study did not shed light on the
such markers in developing retina (Koso et al., 2007, 2009) demon- physiological mechanisms and pathological processes in order to fully
strated that CD73 (NT5E), a cell surface nucleotidase genetically recreate the glaucoma disease event through organoid technology
downstream of CRX, is a cell surface marker of cone-rod common (Table 3).
precursors and mature rods in rodent and primate retina (Koso et al.,
2009). Ader and colleagues adopted this approach to enrich photo- 4.4.2. Photoreceptor dystrophies
receptors from mouse retinal organoids and transplant into the sub The advantage of 3D organoids in uncovering the pathology af-
retinal space of wild type mice and those with moderate (Prom1−/−) fecting structurally and functionally complex neurons such as photo-
and severe (Cplfl/Rho−/−) photoreceptor degeneration(Santos-Ferreira receptors was illustrated by modeling (1) an X-linked RP due to mu-
et al., 2016b). Transplanted cells survived in the subretinal space of all tations in the retinitis pigmentosa (RP) GTPase regulator (RPGR) gene,
mouse models. However, their morphological maturity and integration which encodes an important component of the centrosome cilium in-
within the host retina was observed only in wild type and mouse model terface (Deng et al., 2018), and (2) LCA, which is caused by mutations
of moderate, but not severe degeneration, suggesting that the ther- in cilia related gene CEP290, resulting in aberrant splicing of the pro-
apeutic outcome of retinal repair may depend upon the severity of tein (Parfitt et al., 2016). RPGR is localized to the connecting cilium of
degenerative changes. The CD73-based approach to enrich photo- the photoreceptors and plays an important role in the cilium biogenesis
receptor precursors has been successfully adopted in human retinal and maintenance (Ghosh et al., 2010; Murga-Zamalloa et al., 2010),
organoid and xenotransplantation in rat model (P23H rats) of photo- and protein transport from the inner segment to the outer segment of
receptor degeneration (Gagliardi et al., 2018). The enrichment pro- the photoreceptor (Wang and Deretic, 2014). Interference with the
tocol, based on a panel of cell surface markers including CD73, devel- ciliary function due to the mutation is likely the underlying mechanism
oped by Sowden and colleagues, represents an advancement toward of photoreceptor degeneration. A practical human model of this X-
examining the functionality of organoid-derived photoreceptors linked RP therefore requires the generation of photoreceptors with
through xenotransplantation (Lakowski et al., 2018) (Table 2). Re- inner and outer segments, which does not occur reproducibly in 2D
cently, in a proof of principle approach, optogenetically engineered monolayer culture (Jin et al., 2011; Osakada et al., 2009a). To address
cones (cones in organoids infected with adeno-associated virus to ex- this barrier, Jin and his colleagues generated retinal organoids from
press hyperpolarizing microbial opsin under a cone-specific promoter) iPSCs reprogrammed from urine cells from patients with frame shift
were transplanted subretinally in two mouse models of retinal degen- mutations in RPGR and healthy controls (Deng et al., 2018). Organoids
eration (Cpf1/Rho−/− and rd/rd) (Garita-Hernandez et al., 2019). from controls consisted of structured photoreceptors with outer seg-
Light-driven responses were detected at both photoreceptor and RGC ments with connecting cilia, properly localized opsin, and synaptic
levels, demonstrating the concept of retinal repair through optogen- connection with BCs. In contrast, they observed abnormal photo-
etically transformed photoreceptors that may lack properly formed receptors with short outer segments and mis-localized opsin in patient-
outer segments (Garita-Hernandez et al., 2019). specific organoids. Correction of the RPGR mutation in one of the pa-
tient-specific iPSC lines by CRISP/Cas9 gene editing reversed the ci-
4.4. Disease modeling in organoid liopathy observed in the organoid. Similarly, modeling LCA in orga-
noids revealed CEP290 mutation associated ciliopathy, which was
In the previous sections, we discussed the recent approaches in- corrected when organoids were treated with antisense morpholino to
vestigating 2D iPSC-based disease models, mostly comprising single-cell block aberrant splicing of CEP290 (Table 4) (Parfitt et al., 2016).
subtypes of human origin. However, the inability to mimic highly or-
ganized retinal tissue, cell-cell, cell-matrix and cell-environment inter- 4.5. Drawback of organoids
actions represents an impediment to the development of physiologically
reliable disease models in 2D culture. The capacity of organoid tech- The randomness with which neural progenitors differentiate along
nology to recapitulate in vivo organ development, physiology, and different lineages in the process of self-organization makes the speci-
functionality of original tissues has spurred their use in disease mod- ficity of neural organoids unpredictable to some extent. For example,
eling for genotype-phenotype analysis and drug screening. Combining single cell RNAseq analysis of brain organoids is revealing in that they
disease modeling in organoids with drug screening has predictive po- may generate a range of cells of the central nervous system including
tential to identify clinical responders and non-responders to treatment, the retina (Lancaster and Huch, 2019; Quadrato et al., 2017). This

19
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

relative lack of specificity of differentiation may underlie the current in that direction has been recently made by enhancing synapse for-
deficiency in retinal organoid technology, which is the heterogeneity in mation by exposing developing organoids to ECM-derived peptides
the differentiation outcomes in terms of eye field specification, devel- (Dorgau et al., 2019) and high resolution visualization of synapses in
opment of the optic cup, and generation of retinal cell types (Llonch retinal organoids by passive clearing technique (PACT) and optical
et al., 2018; Mellough et al., 2019; Wang et al., 2018a); (Capowski sectioning (Cora et al., 2019).
et al., 2019). There are at least three variables, separately or in com- In another approach to test the reproducibility of retinal organoid
bination, which may influence the efficiency of retinal organoid gen- generation, Gamm's group subjected 16 different pluripotent stem cells
eration. These are (1) sources of pluripotent stem cells, (2) culture lines (12 hiPSC lines generated from blood or fibroblasts and 4 hESC
methods, and (3) methods for determining and scoring retinal differ- lines) through a modified method of Zhong et al. (2014) and standar-
entiation. Dyer and colleagues, using a standardized scoring system dized staging of optic cup generation and retinal differentiation through
called the STEM-RET, which includes the quantitative measurement of light microscopy, optical coherence tomography, and metabolic
eye field specification, optic cup generation, and retinal differentiation, screening (Capowski et al., 2019). All 16 cell lines successfully gener-
demonstrated that not all pluripotent stem cell lines are equivalent in ated retinal organoids under three stages of development: generation of
retinal differentiation in 3D culture (Hiler et al., 2015; Wang et al., RGCs (stage 1); generation of photoreceptors and interneurons (stage
2018a). For example, iPSCs reprogrammed from rods generated retinal 2), and presence of metabolically active photoreceptors with ribbon
organoids at a higher efficiency than those from fibroblasts (Hiler et al., synapses (stage 3). However, this systematic study confirmed the pro-
2015). Extending this finding, they recently reported that different gressive disorganization and degeneration of inner retinal cells, parti-
retinal cell types are reprogrammed to pluripotency with different ease, cularly RGCs in retinal organoids, which may explain the variability in
presumably due to the retention of cell-type specific epigenetic light responsiveness observed in different labs. This study also revealed
memory, which may subsequently influence the efficiency and fidelity a lack of developmental synchrony between organoids in the same
of the generation of retinal organoids (Wang et al., 2018a). For ex- culture, another limitation to be aware of. These observations are re-
ample, retinal cell types that were more difficult to reprogram to flective of the fact that organoid technology is in the early stage of
pluripotency, such as rods and BCs compared to cones/MG/inter- development. Studies carried out on different pluripotent stem cell lines
neurons, generated retinal organoids with better efficiency, associated under different methodologies through unbiased evaluation of differ-
with the retention of DNA methylation and histone modification pat- entiation outcomes are essential for the development of standardized
terns of the cells of origin (Wang et al., 2018a). Furthermore, Dyer and protocols, particularly important for the reproducibility of the disease
colleagues observed that there were iPSC lines obtained from retinal models.
cells that failed to generate retinal organoids and this deficiency was The protocols for retinal organoids should be flexible enough to
associated with increased expression of MEIS1, a three amino acid loop accommodate the co-culture of retinal organoids with glial and en-
extension (TALE) homeodomain-containing TF (Longobardi et al., dothelial cells to reflect the in vivo cellular organization and complexity,
2014), which was lower in cell lines that generated retinal organoids necessary for accurate and sustainable disease modeling. For example,
(Wang et al., 2018a). This finding is counterintuitive to the emerging establishment of rudimentary vasculature may prevent hypoxia related
role of MEIS1 in coordinating a network of genes during the develop- cell death in 3D organoids, and is therefore essential for long term
ment of the optic cup (Marcos et al., 2015) and whether or not levels of culture to obtain matured retinal cells. Besides, emerging evidence
MESI1 expression would be predictive of retinal organoid generation suggests close interactions between vasculature and homeostasis of
needs further study (Wang et al., 2018a). That different methods have retinal neurons, and the involvement of the former in the progression of
bearing effects on retinal organoid generation from different human degenerative changes in the retina (Sun and Smith, 2018). Astrocytes,
pluripotent stem cell lines was recently demonstrated by Lako's group immigrant cells from the optic nerve, promote survival and function of
(Mellough et al., 2019). They observed a consistent generation of or- neurons by facilitating vascular development, metabolic support, and
ganoids with laminated retinal structure when the early stage of the synapse formation (Vecino et al., 2016; Zuchero and Barres, 2015).
method included mechanical dissociation of stem cell colonies to gen- Consistent with these observations, Myers’ group recently reported that
erate aggregates of pluripotent stem cells in the presence of ROCK in- in a co-culture paradigm, astrocytes enhance functional and morpho-
hibitor, as opposed to enzymatic dissociation method (Mellough et al., logical maturation of RGCs in retinal organoids (VanderWall et al.,
2019). When they paired the enzymatic dissociation method with 2019). Microglia, which are immigrant from the yolk sac, participate in
shaking culture none of the cell lines generated organoids. Even well- retinal development by removing apoptotic cells, supporting neuro-
established protocols from Sasai's group (Nakano et al., 2012) failed to genesis through growth factors and cytokines, and influencing wiring
generate organoids from a cell line (Neo1 hiPSC line) unless the initial through synaptic pruning and the maintenance of synaptic structure
seeding density was increased to 12,000 cells from the recommended and function (Casano and Peri, 2015; Rathnasamy et al., 2019;
9,000 cells, further illustrating the relative influence of different Silverman and Wong, 2018). Activated microglia have been associated
methods and cell lines on the efficiency of retinal organoid generation with pathological conditions that include glaucoma, RP and AMD,
(Mellough et al., 2019). These influences may also underlie the varia- where they may exacerbate degenerative changes by producing pro-
bility in the generation of specific cell types (Chichagova et al., 2019; inflammatory cytokines (Rathnasamy et al., 2019; Silverman and
Zerti et al., 2019; Brooks et al., 2019) and their functional respon- Wong, 2018). Therefore, co-culture of 3D organoids or cells directly
siveness (Zhong et al., 2014; Chichagova et al., 2019; Hallam et al., differentiated in 2D culture with endothelial and glial cells is not only
2018) in the retinal organoids. Associated with this current limitation is necessary to simulate normal neurogenesis and cellular complexity in
the observation that organoids in general reflect a specific develop- vitro but also to examine whether the supportive roles that they display
mental stage of fetal tissues and therefore cells therein are functionally during development can be recapitulated toward cell survival and re-
immature. For example, examination of retinal organoids across mul- generation. A co-culture paradigm of retinal organoid and RPE is in-
tiple iPSC cell lines by a high density multielectrode system revealed tuitively essential for disease modeling given the role RPE plays in the
responses of RGCs to cGMP (phototransduction capability) and light maintenance of structural and functional integrity of photoreceptors
(synaptic transmission capability), however these were immature, si- and their pathology (Nasonkin et al., 2013; Strauss, 2005). The micro
milar to responses obtained from the neonatal mouse retina (Dorgau niche provided by RPE, consisting of cell-cell contact, extracellular
et al., 2019; Hallam et al., 2018). Thus, functional maturation of pho- matrix, diffusible factors/nutrients, and extracellular vesicles may fa-
toreceptors and correct synapse formation with downstream neurons cilitate photoreceptor differentiation and maturity. This premise is
remains a challenge for the utilization of retinal organoids for re- supported by a recent report of accelerated differentiation (Akhtar
generative medicine and functional drug screening. However, progress et al., 2019) and efficient outer segment generation (Achberger et al.,

20
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

2019) when retinal organoids were co-cultured with RPE. Alternatively, and differentiated into neurons and glia (Das et al., 2006). This study
exposure of developing organoids to nutrients elaborated by RPE such provided direct evidence of the neurogenic potential of MG in higher
as polyunsaturated docosahexaenoic acid (DHA), which is a component vertebrates, compared with previous ones (Fischer and Reh, 2001; Ooto
of the outer segments and required for the maintenance of disc mor- et al., 2004) where the BrdU-based lineage-tracing approach raised the
phology (Shindou et al., 2017), may promote maturation of photo- possibility that Brdu+ neurons may represent dead neurons that might
receptors (Brooks et al., 2019). These co-culture paradigms can be have incorporated BrdU during DNA repair (Kugler et al., 2015). In an
executed in organoid-on-a-chip (Park et al., 2019) and other micro- alternate approach to obtaining evidence of neurogenic potential, we
fluidic platforms (Teotia et al., 2019) for co-culture or through simple transplanted activated MG, prospectively enriched from neurotoxin-
fusion of two organoids with cell types of different brain regions. The damaged retina, directly into neonatal rat eyes. Two weeks after
latter approach, exemplified by the fusion of the iPSC-derived thalamic transplantation, CFDA-labeled MG displayed lamina-specific differ-
and cortical organoids to study the reciprocal projection of neurons in entiation in the host retina; those in the ONL and GCL expressed rod-
different CNS domains (Xiang et al., 2019), can be used to determine and RGC-specific markers, respectively, demonstrating the differentia-
the target specificity of RGCs in retinal organoids. tion of a subset of activated MG into lamina-specific retinal neurons in
vivo (Das et al., 2006). Subsequently, different groups, using various
5. Therapeutic regeneration through MG approaches to activate MG that included activation of the SHH pathway
(Wan et al., 2007), neurotoxic injury with the coincidental activation of
5.1. MG-mediated regeneration in lower vertebrates ERK pathways (Karl et al., 2008), ONL injury mediated by N-methyl-N-
nitrosourea (Wan et al., 2008), subtoxic doses of the L-glutamate ana-
MG are the primary support cells in the vertebrate retina, regulating logue DL-alpha-aminoadipic acid (Takeda et al., 2008), and in vivo
homeostasis in one of the most metabolically active tissues. Their activation of Notch and Wnt pathways in an animal model of RP (Del
neurogenic and regenerative properties, was firmly established in lower Debbio et al., 2010) confirmed the neurogenic potential of MG in
vertebrates, particularly in teleosts. In fish, like in amphibians the re- mammals. Later, Reh and colleagues, based on the important role
tina grows continuously, allowing it to keep up with the normal growth played by the pro-neural gene, ASCL1 in MG-mediated regeneration in
of the eyes in adult. Though this homeostatic growth of the retina is zebrafish (Fausett et al., 2008; Ramachandran et al., 2010a, 2011),
different from regeneration that takes place in fish in response to injury demonstrated the influence of ectopically expressed ASCL1 in con-
(Braisted et al., 1994; Raymond et al., 1988; Hitchcock and Raymond, verting MG along the retinal neural lineages in vivo (Jorstad et al., 2017;
1992), both tap into same intrinsic cellular source, the neurogenic Ueki et al., 2015) and in vitro (Pollak et al., 2013). Recently, Chen and
clusters consisting of slow cycling cells in the inner nuclear layer, dis- colleagues directly converted β-catenin-activated MG in uninjured
covered in embryonic and larval gold fish (Johns, 1982). The location mouse retina into rods by over expressing regulators of rod photo-
of these slow-cycling cells in the inner nuclear layer (INL) where MG receptors, OTX2, CRX and NRL (Yao et al., 2018). Not only did this
reside (Julian et al., 1998), and the observation that MG re-enter the study demonstrate a significantly better efficiency of neuronal conver-
cell cycle and migrate into spaces vacated by dying photoreceptors in sion of activated MG, but also showed that MG were associated with the
laser-damaged gold fish retina (Braisted et al., 1994) suggested that MG restoration of vision in a mouse model of congenital blindness. How-
are the RPCs in the neurogenic clusters that sustain regeneration. ever, the lack of an iron clad lineage-tracing approach and the use of a
Raymond and colleagues demonstrated that the progenitor properties rhodopsin-promoter-driven tag, also active in the host's rods, may have
of zebrafish GFP-tagged MG are not confined to their reaction to injury, led to the over-estimation of the efficiency of neuronal conversion (Yao
but are also vital to homeostatic growth of the retina (Bernardos et al., et al., 2018) (Table 5).
2007). These progenitors cycle rapidly and move along the radial
process of the daughter MG to reach the ONL, where they eventually 5.3. Mechanism underlying neurogenic potential of MG
withdraw from the cell cycle and differentiate into rod photoreceptors.
Later, using the optic nerve crush and mechanical injury models in While growing evidence confirms that MG possess evolutionarily
transgenic zebrafish, using transient and permanent fate mapping, conserved neurogenic potential, studies from various labs have ob-
Goldman and colleagues demonstrated that injury activated MG display served that the injury or disease-activated mammalian MG proliferate,
stem cell properties and generate a range of retinal neurons that in- but convert to neurons rather infrequently (Ahmad et al., 2011;
cluded photoreceptors, ACs, BCs and RGCs, and MG in their respective Goldman, 2014; Lenkowski and Raymond, 2014). Whether or not the
lamina (Ramachandran et al., 2010b). converted neurons are functional, make synaptic connections, and
survive for the long-term remains poorly understood (Xia and Ahmad,
5.2. MG-mediated regeneration in mammals 2016b). The challenge that remains is how to unlock the neurogenic
potentials of the mammalian MG in vivo that can sustain retinal repair.
The regenerative potential of MG are evolutionarily conserved in This may require (1) identification of molecular axes involved in MG
mammals but unlike lower vertebrates is dormant. Takahashi and col- development and MG-mediated regeneration and (2) understanding the
leagues used an in vivo approach to examine the regenerative potential niche within which it takes place (Fig. 11).
of MG in response to NMDA-mediated injury to the inner retina in the
adult rat (Ooto et al., 2004). They observed that MG in the INL in- 5.3.1. Intrinsic regulation
corporated BrdU 2 days after neurotoxin injection demonstrating their Given that injuries lead to proliferation of MG across species, but
activation upon injury. Two weeks later, a subset of BrdU+ cells was not to their efficient differentiation into neurons in higher vertebrates,
observed expressing markers corresponding to rods, BCs, HCs, and ACs, it could be assumed that the cross-talk between transcriptional net-
suggesting that the injury activated MG are capable of generating a works subserving the activation and neuronal differentiation is either
range of retinal neurons (Ooto et al., 2004). Second, we tested the not connected to the process, or the network components are in place,
neurogenic potential of MG in vitro on the premise that MG, like the but are not epigenetically primed for optimal activity in the mammalian
radial glia, their CNS counterparts with whom they share morpholo- MG (Xia and Ahmad, 2016b). Together, approaches like genome-wide
gical and biochemical properties (Alvarez-Buylla et al., 2001; Kriegstein screening through RNAseq, single cell RNAseq, and ChIP seq of pro-
and Alvarez-Buylla, 2009; Noctor et al., 2001), possess neural stem cell spectively enriched MG in vitro or in select animal models in quiescent
properties (Das et al., 2006). Using the neurosphere assay, we demon- and activated states may provide insight into molecular axes that can be
strated that a small subset of retrospectively enriched rat MG in re- tested against the background of those operational in zebrafish
sponse to activated Notch and Wnt signaling express stem cell markers (Goldman, 2014; Gorsuch and Hyde, 2014; Lenkowski and Raymond,

21
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Table 5
Neurogenic potential of Mammalian MG.
Animals Experimental Design Outcome Reference

With injury

Wnt/Catenin reporter mice - Intravitreal injection of NMDA, followed by Wnt3 - Wnt3a increased proliferation of dedifferentiated Osakada et al.
- supplementation with RA or valproic acid MG > 20-fold (2007)
- RA/VA promoted MG dedifferentiation to rods/
cones
- Inhibition of Wnt with DKK1 prevented neuronal
regeneration
Adult WT mice - IP injection of MNU followed by daily intravitreal - Shh treatment increased proliferation of MG- Wan et al. (2007)
injection of Shh for 7 days derived progenitors, which adopted rod
photoreceptor fate
GAD67-GFP and Grm6-GFP transgenic mice - Intraocular injection of NMDA followed by single - MG dedifferentiated to progenitors Karl et al. (2008)
injection of FGF1, EGF, or FGF1+insulin 2 days - BrdU-labeled MG subset differentiated into
post-injury amacrine cells
Adult WT rats; PN7 Z/EG mice - IP injection of MNU - MG underwent gliosis and proliferation Wan et al. (2008)
- intravitreal transplantation of MG-derived - subset of MG differentiated into rod PRs and
rhodopsin+ cells into damaged mouse retinas formed possible synapses
- transplanted cells migrated and grafted in host
retina and produced rhodopsin
Adult Axin2(LacZ/+) Wnt reporter mice - Laser burn injuries on retinas of transgenic mice - subset of MG proliferated post-injury and Liu et al. (2013)
with increased sensitivity to Wnt signaling adopted RPC expression patterns
- retinal explant culture treated with Wnt3a - Wnt-responsive Muller cells showed long-term
survival
- some MG expressed rhodopsin
Adult and young transgenic mice - IP injections of tamoxifen for 5 days to induce Ascl1 - Only after injury, Ascl1 overexpression Ueki et al. (2015)
overexpressing Ascl1 in presence of expression followed by either intravitreal injection stimulated dedifferentiation and neurogenesis in
tamoxifen of NMDA or continuous exposure to 10,000 lux MG, giving rise to bipolar, amacrine, and rod
light for 8 h photoreceptor cells
Adult WT rats - Intravitreal injection of NMDA, followed by - MG proliferated post injury Ooto et al. (2004)
injections of bFGF, EGF, RA, Activin A, and Bovine - Expression of PKC, NSE, rhodopsin and recoverin
insulin in BrdU-labeled cells suggested dedifferentiation
of MG.
- RA promoted increased bipolar cell growth
Adult transgenic mice overexpressing Ascl1 - IP injections of tamoxifen for 5 days to induce Ascl1 - tamoxifen and NMDA treatment resulted in Jorstad et al.
in presence of tamoxifen expression followed by either intravitreal injection dedifferentiation of MG and neurogenesis (2017)
of NMDA (2 days later) and TSA (4 days later) - MG-derived neurons expressed bipolar and
amacrine cell markers, formed synapses with
host retinal neurons, and responded to light

Without injury

Neonatal rat - Intravitreal injection of prospectively enriched MG - Transplanted MG differentiated into cells Das et al. (2006)
from adult rat expressing rod and RGC-specific markers, located
in hosts' ONL and RGC layers, respectively
Adult WT mice - Subretinal injections of subtoxic doses of glutamate - MG dedifferentiated and a subset differentiated Takeda et al.
or its analog α-AA into rod photoreceptors (2008)
S334ter rats (rat model of PR degeneration) - intravitreal injections of Wn2b and Jag1 followed - MG activated in response to stimulation of Notch Del Debbio et al.
by intravitreal injections of Shh and DAPT and Wnt signaling (2010)
- Subset of MG differentiated into rod
photoreceptors
- improved rats' response to light
Adult WT and Rosa26-tdTomato reporter - Shh10-GFAP-mediated gene transfer of WT β- - β-catenin gene transfer stimulated MG Yao et al. (2016)
mice; catenin, Lin28a or Lin28b proliferation
Lin28aflox/flox,and Lin28bflox/flox mice - Lin28 regulated MG proliferation
- A subset of cell-cycle-reactivated MG cells
expressed markers for amacrine cells
Gnat1rd17Gnat2cpfl3 double mutant mice, a - Shh10-mdieated gene transfer of β-catenin - β-catenin activated MG proliferation Yao et al. (2018)
model of congenital blindness (intravitreal injection), followed by Otx2, Crx, and - dedifferentiated MGs reprogrammed into rod
Nrl (intravitreal injection) photoreceptors Otx2, Crx, and Nrl administration
- new rod PRs integrated into host retina and
rescued light response

2014). For example, the molecular axis defined by LIN28, a hetero- in this direction. For example, evidence suggests that LIN28 influences
chronic gene, and ASCL1, which facilitates MG-mediated regeneration different developmental function by contextual recruitment of HMGA2,
in zebrafish, is a valid target in mammals (Fausett et al., 2008; Pollak a gene encoding a DNA architecture protein (Nishino et al., 2008;
et al., 2013; Ramachandran et al., 2011). Parameswaran et al., 2014) and ASCL1 (Cimadamore et al., 2013) by
Recent studies in higher vertebrates have demonstrated LIN28's directly regulating the heterochronic miRNA let-7 (Cimadamore et al.,
regulatory involvement in the maintenance of neural progenitors and 2013; Nishino et al., 2008; Xia and Ahmad, 2016a). let-7 targets
neuroglial decision (Rehfeld et al., 2015; Xia et al., 2018), two im- HMGA2 (Nishino et al., 2008; Xia and Ahmad, 2016a) and ASCL1
portant functions likely to be recruited if MG were to generate neurons (Cimadamore et al., 2013) transcripts. Therefore, the LIN28-let-7-
in response to injury. Therefore, it is important to understand LIN28- HMGA2/ASCL1 axis may represent an evolutionary conserved axis that
based mechanisms underlying these functions. Progress has been made may be the key to unlocking neurogenic potential of MG. Early

22
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Fig. 11. A schematic representation of candidate


intrinsic and extrinsic factors regulating neuro-
genic potential of MG. A subset of MG with dor-
mant stem cell properties responds to injury by
proliferating, presumably under the influence of
Notch signaling. As the activated MG migrate out of
the INL, the changing niche, reflected in altered in-
tercellular signaling, influence them to engage both
the excitatory (Green) and inhibitory (Red) cell-in-
trinsic axes regulating the neurogenic potential. It is
possible that the imbalance between the two axes
and their inadequate niche-based recruitment pre-
vents mammalian MG from regenerating retinal
neurons. The niche could be composed of retinal
neurons, RPE, endothelial cells, and immigrant as-
trocytes and microglia. The niche-based commu-
nication for regeneration may involve diverse sig-
naling pathways, that may include those mediated by
Notch/growth factor/cytokine/neurotransmitter,
acting in concert. ONL: outer nuclear layer, OPL:
outer plexiform layer, INL: inner nuclear layer, IPL:
inner plexiform layer, GCL: ganglion cell layer.
(Figure adapted from Xia and Ahmad, 2016a,b.)

evidence supports this premise. For example, overexpression of LIN28A improvement of ASCL1-mediated reprogramming of MG by ectopic
in the enriched MG prompted these cells to acquire neuronal mor- expression of these miRNAs (Wohl et al., 2019; Wohl and Reh, 2016).
phology and express immunoreactivities and transcripts corresponding The above findings, together with a recent one (Xia et al., 2018),
to neuronal genes (Xia and Ahmad, 2016b; Xia et al., 2018). More reveal that LIN28A- and REST-related molecular axes may act in con-
importantly, neuronal differentiation was accompanied by an increase cert in regulating the neurogenic potential of the mammalian MG.
in levels of miR-124, a proneural miRNA (Stappert et al., 2015), However, the efficacy of the axes in promoting regeneration is likely to
HMGA2, and ASCL1, and a decrease in REST, a global inhibitor of depend upon chromatin accessibility, influenced by epigenetic reg-
neuronal differentiation (Qureshi et al., 2010). Furthermore, it has been ulators of histones. For example, Reh and colleagues, who had observed
demonstrated that ectopic expression of ASCL1 confers neuronal that forced expression of ASCL1 in MG is ineffective in inducing neu-
property on mouse MG in vitro and in vivo (Jorstad et al., 2017; Pollak rogenic change beyond postnatal day 16 in the mouse retina (Ueki
et al., 2013), suggesting that its endogenous activation by the recruit- et al., 2015), argued that the loss of neurogenic potential in more
ment of LIN28-let-7-HMGA2/ASCL1 axis could lead to the activation of mature MG could be due the loss of chromatin accessibility (Jorstad
the neurogenic program in MG. et al., 2017). They tested this premise by pharmacologically inhibiting
Targeting the LIN28-let-7-HMGA2/ASCL1 axis alone may not be histone deacetylase, an enzyme that takes part in chromatin con-
sufficient for the functional reprogramming of MG along the neuronal densation. Conditional activation of ASCL1 in MG in neurotoxin-da-
lineage, given that its influence may not be able to counter the in- maged adult mouse retina along with the exposure to TSA, a potent
hibitory resistance of the REST axis completely. REST, by suppressing histone deacetylase inhibitor, led to differentiation of activated MG into
the expression of proneural miRNAs, miR-124 and miR-9-9*, whose inner retinal neurons, capable of forming synapses and responding to
targets are proglial genes encoding SOX9 and the HES family of reg- light (Jorstad et al., 2017). More importantly from a mechanistic
ulators, may keep MG non-neuronal, and therefore non-regenerative viewpoint, was the observation that the neuronal differentiation was
(Stappert et al., 2015). Besides, miR-124 (Visvanathan et al., 2007) and accompanied by a progressive change in DNA accessibility at neural
miR-9-9* (Packer et al., 2008), together with other miRNAs such as miR- genes, demonstrating the critical role of cooperation between molecular
29 (Xia et al., 2019), negatively regulate the expression of REST and axes and epigenetic regulators toward unmasking the neurogenic po-
CoREST, key components of the REST complex. The existence of the tential of MG.
miR-29-REST-miR-124-9-9*-SOX9/HES1 axis in MG suggested that ei-
ther inhibiting REST or ectopically expressing miR-29/miR-124/miR-9-
9* or both, may direct MG along the neuronal lineage. This notion was 5.3.2. Extrinsic regulation
supported by the following observations. For example, over expression MG respond to injuries by proliferating and migrating out of the
of miR-124-9-9* in MG facilitated development of neuronal morphology inner nuclear layer (Del Debbio et al., 2010). However, despite pro-
with accompanied expression of neuronal genes and suppression of liferation and migration as in zebrafish retina, as mentioned, mamma-
glial-specific genes (Xia and Ahmad, 2016b). Expression of both ASCL1 lian MG do not initiate regeneration effectively. This raises the possi-
and LIN28A was upregulated and that of REST was down-regulated in bility that, in addition to cell-intrinsic constraints, the environment in
the perturbed groups, compared to controls. That miR-124 and miR-9-9* the mammalian retina might not support neurogenic conversion of MG.
could facilitate proneural function of ASCL1 was demonstrated by the This premise is supported by observations that parenchymal astrocytes
in the CNS do not display neurogenic properties unless removed from

23
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

the niche (Kriegstein and Alvarez-Buylla, 2009) and MG, enriched from cellular and molecular levels unless models are developed where target
the retina, readily generate early (e.g., RGCs) and late (e.g. rods) born cells could be aged either by extending culture time or through ectopic
neurons when exposed to culture conditions simulating early and late expression of PROGERIN to promote premature aging (Miller et al.,
retinal histogenesis, respectively (Das et al., 2006). A niche-based ap- 2013). Besides the immaturity of cells and synaptic organization, the
proach to unlock neurogenic potential of MG, therefore appears logical. current disease model suffers from the lack of cellular heterogeneity of
It will involve examining the relationship of MG with neighboring the retina that includes RPE, astrocytes, microglia, and endothelial
retinal cells, microglia, immigrant astrocytes, and endothelial cells in cells, all of which are required for the proper development and main-
the context of signaling pathways and their capacity to engage the tenance of the retina. In addition, they may play important non-cell
molecular axes involved in reprograming MG along neuronal lineage. autonomous roles in the emergence of diseases, exemplified by astro-
Past and recent studies carried out in different species have identi- cyte-mediated toxicity of motor neurons in familial and sporadic
fied several signaling pathways that may play important roles in med- amyotrophic lateral sclerosis (ALS) disease models (Meyer et al., 2014).
iating the influence of the niche in reprogramming MG. These include Therefore, development of co-culture paradigms that facilitate cell-cell
pathways mediated by Notch (Del Debbio et al., 2016; Goldman, 2014), interactions, physical or diffusional, using organoid-on-a-chip and other
Wnt (Das et al., 2006; Del Debbio et al., 2010), FGF (Das et al., 2006; microfluidic platforms would help simulate the in vivo cellular com-
Fischer et al., 2002; Goldman, 2014), insulin (Fischer et al., 2002) and plexities for a more realistic disease modeling. Like retinal organoid
insulin-like growth factor-1 (Goldman, 2014), SHH (Wan et al., 2007), technology, the discipline of MG-based regeneration is a recent one. As
and cytokines such as TNFα (Goldman, 2014; Gorsuch and Hyde, we understand more about the molecular axes underlying the activation
2014), leptin, and IL-11 (Zhao et al., 2014). However, the identity of and neurogenic potential of MG, aided by the mechanistic information
cells delivering the signals and whether these signals act in concert about retinal development emerging from 2D and 3D co-culture, the
remains largely unknown, but studies in zebrafish have begun to shed prospect of a small molecule-based recruitment of these endogenous
light on these issues. For example, it has been observed that TNFα re- progenitors for therapeutic regeneration appears practical.
leased by dying photoreceptors in mechanically damaged retina may
constitute one of the early signals for activating MG for regeneration Declaration of competing interest
(Gorsuch and Hyde, 2014). Further, it was reported that zebrafish MG
respond to retinal injury by secreting leptin and IL-11, which help re- The authors declare no competing or financial interests.
program cells in an autocrine fashion (Zhao et al., 2014). In both cases,
these cytokines facilitated injury-dependent induction of ASCL1, a key Acknowledgments
step in reprogramming of MG (Goldman, 2014). Whether signaling
mediated by these factors is involved in the activation of mammalian We are thankful to Nam Nguyen for artwork and Matt Van Hook and
MG remains to be demonstrated. On the other hand, though the identity John Sorrentino for critiques and editorial help. Biorender.com was
of cells delivering Notch signaling in MG remains speculative, evidence used for the artwork. This work was supported by National Institutes of
suggest that its role in the activation of MG is evolutionarily conserved Health (2R01EY022051-05, R01EY029778-01), Pearson Foundation,
(Ahmad et al., 2011; Conner et al., 2014; Goldman, 2014) and therefore Lincy Foundation, and Nebraska Department of Health and Human
it is a valid niche-based target for MG-dependent regeneration. One of Services (LB606).
the important components of the niche, the microglia, one of the first
responders to injury, may mediate regeneration through inflammatory References
signals (Fischer et al., 2014), presumably by cytokine-mediated mod-
ification of the epigenetic status and expression of LIN28 (Reyes- Acharya, H.R., Dooley, C.M., Thoreson, W.B., Ahmad, I., 1997. cDNA cloning and ex-
Aguirre et al., 2013). Therefore, niche-related information is crucial for pression analysis of NeuroD mRNA in human retina. Biochem. Biophys. Res.
Commun. 233, 459–463.
understanding the temporal and spatial modulation of signaling re- Achberger, K., Probst, C., Haderspeck, J., Bolz, S., Rogal, J., Chuchuy, J., Nikolova, M.,
quired for activating MG and shifting them from the state of activation Cora, V., Antkowiak, L., Haq, W., Shen, N., Schenke-Layland, K., Ueffing, M., Liebau,
to neuronal differentiation. S., Loskill, P., 2019. Merging organoid and organ-on-a-chip technology to generate
complex multi-layer tissue models in a human retina-on-a-chip platform. Elife 8.
Adler, R., Canto-Soler, M.V., 2007. Molecular mechanisms of optic vesicle development:
6. Future directions complexities, ambiguities and controversies. Dev. Biol. 305, 1–13.
Ahmad, I., 1995. Mash-1 is expressed during ROD photoreceptor differentiation and binds
an E-box, E(opsin)-1 in the rat opsin gene. Brain Res Dev Brain Res 90, 184–189.
In the past two decades, advancement made in our understanding of Ahmad, I., 2001. Stem cells: new opportunities to treat eye diseases. Investig.
mechanisms underlying retinal development (Heavner and Pevny, Ophthalmol. Vis. Sci. 42, 2743–2748.
2012) in conjunction with the evidence that the adult retina harbors Ahmad, I., Del Debbio, C.B., Das, A.V., Parameswaran, S., 2011. Muller glia: a promising
target for therapeutic regeneration. Investig. Ophthalmol. Vis. Sci. 52, 5758–5764.
latent stem cells (Ahmad et al., 2011; Goldman, 2014) and the advent of
Ahmad, I., Dooley, C.M., Polk, D.L., 1997. Delta-1 is a regulator of neurogenesis in the
iPSC technology (Shi et al., 2017) has made regenerative medicine a vertebrate retina. Dev. Biol. 185, 92–103.
realistic approach for treating retinal degeneration. We can re- Ahmad, I., Dooley, C.M., Thoreson, W.B., Rogers, J.A., Afiat, S., 1999. In vitro analysis of
producibly recruit developmental mechanisms to directly differentiate a mammalian retinal progenitor that gives rise to neurons and glia. Brain Res. 831,
1–10.
hiPSCs into specific retinal cell types in 2D culture or take advantage of Akhtar, T., Xie, H., Khan, M.I., Zhao, H., Bao, J., Zhang, M., Xue, T., 2019. Accelerated
the self-organization of iPSCs into retinal organoids in 3D culture to photoreceptor differentiation of hiPSC-derived retinal organoids by contact co-cul-
examine preclinical outcome of ex-vivo cell therapy in chimeric tissues ture with retinal pigment epithelium. Stem Cell Res. 39, 101491.
Alfano, G., Conte, I., Caramico, T., Avellino, R., Arno, B., Pizzo, M.T., Tanimoto, N., Beck,
and models of complex diseases such as glaucoma and retinitis pig- S.C., Huber, G., Dolle, P., Seeliger, M.W., Banfi, S., 2011. Vax2 regulates retinoic acid
mentosa. However, there are several barriers that need to be addressed distribution and cone opsin expression in the vertebrate eye. Development 138,
before regenerative medicine becomes clinically practical, the re- 261–271.
Almasieh, M., Wilson, A.M., Morquette, B., Cueva Vargas, J.L., Di Polo, A., 2012. The
quirement of standardization of protocols, including scoring criteria for molecular basis of retinal ganglion cell death in glaucoma. Prog. Retin. Eye Res. 31,
outcomes for iPSC lines with different epigenetic signatures and gen- 152–181.
erating safe and pure populations of cells for transplantation notwith- Altshuler, D., Cepko, C., 1992. A temporally regulated, diffusible activity is required for
rod photoreceptor development in vitro. Development 114, 947–957.
standing. For example, in most cases, cells in either 2D culture or 3D Altshuler, D., Lo Turco, J.J., Rush, J., Cepko, C., 1993. Taurine promotes the differ-
organoids are relatively immature, approximating fetal or neonatal age entiation of a vertebrate retinal cell type in vitro. Development 119, 1317–1328.
while the degenerative changes in diseases usually have late onset Alvarez-Buylla, A., Garcia-Verdugo, J.M., Tramontin, A.D., 2001. A unified hypothesis on
the lineage of neural stem cells. Nat. Rev. Neurosci. 2, 287–293.
(Rowe and Daley, 2019). Therefore, the current disease models may be
Anderson, A.M., Weasner, B.M., Weasner, B.P., Kumar, J.P., 2012. Dual transcriptional
inadequate in providing a comprehensive picture of the pathology at

24
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

activities of SIX proteins define their roles in normal and ectopic eye development. 94–102.
Development 139, 991–1000. Chen, J., Riazifar, H., Guan, M.X., Huang, T., 2016. Modeling autosomal dominant optic
Assawachananont, J., Mandai, M., Okamoto, S., Yamada, C., Eiraku, M., Yonemura, S., atrophy using induced pluripotent stem cells and identifying potential therapeutic
Sasai, Y., Takahashi, M., 2014. Transplantation of embryonic and induced pluripotent targets. Stem Cell Res. Ther. 7, 2.
stem cell-derived 3D retinal sheets into retinal degenerative mice. Stem Cell Reports Cheng, C.Y., Allingham, R.R., Aung, T., Tham, Y.C., Hauser, M.A., Vithana, E.N., Khor,
2, 662–674. C.C., Wong, T.Y., 2014. Association of common SIX6 polymorphisms with peripa-
Aubert, J., Dunstan, H., Chambers, I., Smith, A., 2002. Functional gene screening in pillary retinal nerve fiber layer thickness: the Singapore Chinese Eye Study. Investig.
embryonic stem cells implicates Wnt antagonism in neural differentiation. Nat. Ophthalmol. Vis. Sci. 56, 478–483.
Biotechnol. 20, 1240–1245. Chiang, C., Litingtung, Y., Lee, E., Young, K.E., Corden, J.L., Westphal, H., Beachy, P.A.,
Austin, C.P., Feldman, D.E., Ida Jr., J.A., Cepko, C.L., 1995. Vertebrate retinal ganglion 1996. Cyclopia and defective axial patterning in mice lacking Sonic hedgehog gene
cells are selected from competent progenitors by the action of Notch. Development function. Nature 383, 407–413.
121, 3637–3650. Chichagova, V., Dorgau, B., Felemban, M., Georgiou, M., Armstrong, L., Lako, M., 2019.
Ayuso, C., Millan, J.M., 2010. Retinitis pigmentosa and allied conditions today: a para- Differentiation of retinal organoids from human pluripotent stem cells. Curr Protoc
digm of translational research. Genome Med. 2, 34. Stem Cell Biol 50, e95.
Barnea-Cramer, A.O., Wang, W., Lu, S.J., Singh, M.S., Luo, C., Huo, H., McClements, M.E., Chow, R.L., Altmann, C.R., Lang, R.A., Hemmati-Brivanlou, A., 1999. Pax6 induces ec-
Barnard, A.R., MacLaren, R.E., Lanza, R., 2016. Function of human pluripotent stem topic eyes in a vertebrate. Development 126, 4213–4222.
cell-derived photoreceptor progenitors in blind mice. Sci. Rep. 6, 29784. Chow, R.L., Lang, R.A., 2001. Early eye development in vertebrates. Annu. Rev. Cell Dev.
Berkers, G., van Mourik, P., Vonk, A.M., Kruisselbrink, E., Dekkers, J.F., de Winter-de Biol. 17, 255–296.
Groot, K.M., Arets, H.G.M., Marck-van der Wilt, R.E.P., Dijkema, J.S., Vanderschuren, Chuang, J.C., Raymond, P.A., 2002. Embryonic origin of the eyes in teleost fish. Bioessays
M.M., Houwen, R.H.J., Heijerman, H.G.M., van de Graaf, E.A., Elias, S.G., Majoor, 24, 519–529.
C.J., Koppelman, G.H., Roukema, J., Bakker, M., Janssens, H.M., van der Meer, R., Cimadamore, F., Amador-Arjona, A., Chen, C., Huang, C.T., Terskikh, A.V., 2013. SOX2-
Vries, R.G.J., Clevers, H.C., de Jonge, H.R., Beekman, J.M., van der Ent, C.K., 2019. LIN28/let-7 pathway regulates proliferation and neurogenesis in neural precursors.
Rectal organoids enable personalized treatment of cystic fibrosis. Cell Rep. 26, Proc. Natl. Acad. Sci. U. S. A. 110, E3017–E3026.
1701–1708 e1703. Cohen, A., Bray, G.M., Aguayo, A.J., 1994. Neurotrophin-4/5 (NT-4/5) increases adult rat
Bernardos, R.L., Barthel, L.K., Meyers, J.R., Raymond, P.A., 2007. Late-stage neuronal retinal ganglion cell survival and neurite outgrowth in vitro. J. Neurobiol. 25,
progenitors in the retina are radial Muller glia that function as retinal stem cells. J. 953–959.
Neurosci. 27, 7028–7040. Collin, J., Zerti, D., Queen, R., Santos-Ferreira, T., Bauer, R., Coxhead, J., Hussain, R.,
Bernier, G., Panitz, F., Zhou, X., Hollemann, T., Gruss, P., Pieler, T., 2000. Expanded Steel, D., Mellough, C., Ader, M., Sernagor, E., Armstrong, L., Lako, M., 2019. CRX
retina territory by midbrain transformation upon overexpression of Six6 (Optx2) in expression in pluripotent stem cell-derived photoreceptors marks a transplantable
Xenopus embryos. Mech. Dev. 93, 59–69. subpopulation of early cones. Stem Cells 37, 609–622.
Bohrer, L.R., Wiley, L.A., Burnight, E.R., Cooke, J.A., Giacalone, J.C., Anfinson, K.R., Conner, C., Ackerman, K.M., Lahne, M., Hobgood, J.S., Hyde, D.R., 2014. Repressing
Andorf, J.L., Mullins, R.F., Stone, E.M., Tucker, B.A., 2019. Correction of NR2E3 notch signaling and expressing TNFalpha are sufficient to mimic retinal regeneration
associated enhanced S-cone syndrome patient-specific iPSCs using CRISPR-cas9. by inducing Muller glial proliferation to generate committed progenitor cells. J.
Genes (Basel) 10. Neurosci. 34, 14403–14419.
Bolinches-Amoros, A., Lukovic, D., Castro, A.A., Leon, M., Kamenarova, K., Kaneva, R., Cora, V., Haderspeck, J., Antkowiak, L., Mattheus, U., Neckel, P.H., Mack, A.F., Bolz, S.,
Jendelova, P., Blanco-Kelly, F., Ayuso, C., Corton, M., Erceg, S., 2018. Generation of a Ueffing, M., Pashkovskaia, N., Achberger, K., Liebau, S., 2019. A cleared view on
human iPSC line from a patient with congenital glaucoma caused by mutation in retinal organoids. Cells 8.
CYP1B1 gene. Stem Cell Res. 28, 96–99. Cornish, E.E., Hendrickson, A.E., Provis, J.M., 2004. Distribution of short-wavelength-
Bond, L.M., Peden, A.A., Kendrick-Jones, J., Sellers, J.R., Buss, F., 2011. Myosin VI and its sensitive cones in human fetal and postnatal retina: early development of spatial
binding partner optineurin are involved in secretory vesicle fusion at the plasma order and density profiles. Vis. Res. 44, 2019–2026.
membrane. Mol. Biol. Cell 22, 54–65. Costa-Mattioli, M., Monteggia, L.M., 2013. mTOR complexes in neurodevelopmental and
Borooah, S., Phillips, M.J., Bilican, B., Wright, A.F., Wilmut, I., Chandran, S., Gamm, D., neuropsychiatric disorders. Nat. Neurosci. 16, 1537–1543.
Dhillon, B., 2013. Using human induced pluripotent stem cells to treat retinal disease. Curcio, C.A., Medeiros, N.E., Millican, C.L., 1996. Photoreceptor loss in age-related ma-
Prog. Retin. Eye Res. 37, 163–181. cular degeneration. Investig. Ophthalmol. Vis. Sci. 37, 1236–1249.
Boucherie, C., Mukherjee, S., Henckaerts, E., Thrasher, A.J., Sowden, J.C., Ali, R.R., 2013. Daiger, S.P., Sullivan, L.S., Bowne, S.J., 2013. Genes and mutations causing retinitis
Brief report: self-organizing neuroepithelium from human pluripotent stem cells fa- pigmentosa. Clin. Genet. 84, 132–141.
cilitates derivation of photoreceptors. Stem Cells 31, 408–414. Danno, H., Michiue, T., Hitachi, K., Yukita, A., Ishiura, S., Asashima, M., 2008. Molecular
Braisted, J.E., Essman, T.F., Raymond, P.A., 1994. Selective regeneration of photo- links among the causative genes for ocular malformation: otx2 and Sox2 coregulate
receptors in goldfish retina. Development 120, 2409–2419. Rax expression. Proc. Natl. Acad. Sci. U. S. A. 105, 5408–5413.
Brooks, M.J., Chen, H.Y., Kelley, R.A., Mondal, A.K., Nagashima, K., De Val, N., Li, T., Das, A.V., Bhattacharya, S., Zhao, X., Hegde, G., Mallya, K., Eudy, J.D., Ahmad, I., 2008.
Chaitankar, V., Swaroop, A., 2019. Improved retinal organoid differentiation by The canonical Wnt pathway regulates retinal stem cells/progenitors in concert with
modulating signaling pathways revealed by comparative transcriptome analyses with Notch signaling. Dev. Neurosci. 30, 389–409.
development in vivo. Stem Cell Reports 13, 891–905. Das, A.V., Mallya, K.B., Zhao, X., Ahmad, F., Bhattacharya, S., Thoreson, W.B., Hegde,
Brown, N.L., Patel, S., Brzezinski, J., Glaser, T., 2001. Math5 is required for retinal G.V., Ahmad, I., 2006. Neural stem cell properties of Muller glia in the mammalian
ganglion cell and optic nerve formation. Development 128, 2497–2508. retina: regulation by Notch and Wnt signaling. Dev. Biol. 299, 283–302.
Brzezinski, J.A.t., Prasov, L., Glaser, T., 2012. Math5 defines the ganglion cell competence Davis, A.A., Matzuk, M.M., Reh, T.A., 2000. Activin A promotes progenitor differentiation
state in a subpopulation of retinal progenitor cells exiting the cell cycle. Dev. Biol. into photoreceptors in rodent retina. Mol. Cell. Neurosci. 15, 11–21.
365, 395–413. de Melo, J., Du, G., Fonseca, M., Gillespie, L.A., Turk, W.J., Rubenstein, J.L., Eisenstat,
Burnight, E.R., Wiley, L.A., Drack, A.V., Braun, T.A., Anfinson, K.R., Kaalberg, E.E., D.D., 2005. Dlx1 and Dlx2 function is necessary for terminal differentiation and
Halder, J.A., Affatigato, L.M., Mullins, R.F., Stone, E.M., Tucker, B.A., 2014. CEP290 survival of late-born retinal ganglion cells in the developing mouse retina.
gene transfer rescues Leber congenital amaurosis cellular phenotype. Gene Ther. 21, Development 132, 311–322.
662–672. Dekkers, J.F., Wiegerinck, C.L., de Jonge, H.R., Bronsveld, I., Janssens, H.M., de Winter-
Capowski, E.E., Samimi, K., Mayerl, S.J., Phillips, M.J., Pinilla, I., Howden, S.E., Saha, J., de Groot, K.M., Brandsma, A.M., de Jong, N.W., Bijvelds, M.J., Scholte, B.J.,
Jansen, A.D., Edwards, K.L., Jager, L.D., Barlow, K., Valiauga, R., Erlichman, Z., Nieuwenhuis, E.E., van den Brink, S., Clevers, H., van der Ent, C.K., Middendorp, S.,
Hagstrom, A., Sinha, D., Sluch, V.M., Chamling, X., Zack, D.J., Skala, M.C., Gamm, Beekman, J.M., 2013. A functional CFTR assay using primary cystic fibrosis intestinal
D.M., 2019. Reproducibility and staging of 3D human retinal organoids across mul- organoids. Nat. Med. 19, 939–945.
tiple pluripotent stem cell lines. Development 146. Del Debbio, C.B., Balasubramanian, S., Parameswaran, S., Chaudhuri, A., Qiu, F., Ahmad,
Carl, M., Loosli, F., Wittbrodt, J., 2002. Six3 inactivation reveals its essential role for the I., 2010. Notch and Wnt signaling mediated rod photoreceptor regeneration by
formation and patterning of the vertebrate eye. Development 129, 4057–4063. Muller cells in adult mammalian retina. PLoS One 5, e12425.
Carnes, M.U., Liu, Y.P., Allingham, R.R., Whigham, B.T., Havens, S., Garrett, M.E., Qiao, Del Debbio, C.B., Mir, Q., Parameswaran, S., Mathews, S., Xia, X., Zheng, L., Neville, A.J.,
C., Investigators, N.C., Katsanis, N., Wiggs, J.L., Pasquale, L.R., Ashley-Koch, A., Oh, Ahmad, I., 2016. Notch signaling activates stem cell properties of muller glia through
E.C., Hauser, M.A., 2014. Discovery and functional annotation of SIX6 variants in transcriptional regulation and skp2-mediated degradation of p27Kip1. PLoS One 11,
primary open-angle glaucoma. PLoS Genet. 10, e1004372. e0152025.
Casano, A.M., Peri, F., 2015. Microglia: multitasking specialists of the brain. Dev. Cell 32, Deng, W.L., Gao, M.L., Lei, X.L., Lv, J.N., Zhao, H., He, K.W., Xia, X.X., Li, L.Y., Chen, Y.C.,
469–477. Li, Y.P., Pan, D., Xue, T., Jin, Z.B., 2018. Gene correction reverses ciliopathy and
Cayouette, M., Poggi, L., Harris, W.A., 2006. Lineage in the vertebrate retina. Trends photoreceptor loss in iPSC-derived retinal organoids from retinitis pigmentosa pa-
Neurosci. 29, 563–570. tients. Stem Cell Reports 10, 1267–1281.
Cayouette, M., Raff, M., 2003. The orientation of cell division influences cell-fate choice Dooley, C.M., James, J., Jane McGlade, C., Ahmad, I., 2003. Involvement of numb in
in the developing mammalian retina. Development 130, 2329–2339. vertebrate retinal development: evidence for multiple roles of numb in neural dif-
Cepko, C.L., Austin, C.P., Yang, X., Alexiades, M., Ezzeddine, D., 1996. Cell fate de- ferentiation and maturation. J. Neurobiol. 54, 313–325.
termination in the vertebrate retina. Proc. Natl. Acad. Sci. U. S. A. 93, 589–595. Dorgau, B., Felemban, M., Hilgen, G., Kiening, M., Zerti, D., Hunt, N.C., Doherty, M.,
Chen, G., Hou, Z., Gulbranson, D.R., Thomson, J.A., 2010. Actin-myosin contractility is Whitfield, P., Hallam, D., White, K., Ding, Y., Krasnogor, N., Al-Aama, J., Asfour,
responsible for the reduced viability of dissociated human embryonic stem cells. Cell H.Z., Sernagor, E., Lako, M., 2019. Decellularised extracellular matrix-derived pep-
Stem Cell 7, 240–248. tides from neural retina and retinal pigment epithelium enhance the expression of
Chen, J., Ma, L., Wang, S., Wang, X., Sun, Y., Gao, L., Li, J., Zhou, G., 2017. Analysis of synaptic markers and light responsiveness of human pluripotent stem cell derived
expression of transcription factors in early human retina. Int. J. Dev. Neurosci. 60, retinal organoids. Biomaterials 199, 63–75.

25
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Dorsky, R.I., Chang, W.S., Rapaport, D.H., Harris, W.A., 1997. Regulation of neuronal Gonzalez-Cordero, A., West, E.L., Pearson, R.A., Duran, Y., Carvalho, L.S., Chu, C.J.,
diversity in the Xenopus retina by Delta signalling. Nature 385, 67–70. Naeem, A., Blackford, S.J.I., Georgiadis, A., Lakowski, J., Hubank, M., Smith, A.J.,
Dorsky, R.I., Rapaport, D.H., Harris, W.A., 1995. Xotch inhibits cell differentiation in the Bainbridge, J.W.B., Sowden, J.C., Ali, R.R., 2013. Photoreceptor precursors derived
Xenopus retina. Neuron 14, 487–496. from three-dimensional embryonic stem cell cultures integrate and mature within
Eberle, D., Schubert, S., Postel, K., Corbeil, D., Ader, M., 2011. Increased integration of adult degenerate retina. Nat. Biotechnol. 31, 741–747.
transplanted CD73-positive photoreceptor precursors into adult mouse retina. Gorsuch, R.A., Hyde, D.R., 2014. Regulation of Muller glial dependent neuronal re-
Investig. Ophthalmol. Vis. Sci. 52, 6462–6471. generation in the damaged adult zebrafish retina. Exp. Eye Res. 123, 131–140.
Eiraku, M., Takata, N., Ishibashi, H., Kawada, M., Sakakura, E., Okuda, S., Sekiguchi, K., Graham, D.R., Overbeek, P.A., Ash, J.D., 2005. Leukemia inhibitory factor blocks ex-
Adachi, T., Sasai, Y., 2011. Self-organizing optic-cup morphogenesis in three-di- pression of Crx and Nrl transcription factors to inhibit photoreceptor differentiation.
mensional culture. Nature 472, 51–56. Investig. Ophthalmol. Vis. Sci. 46, 2601–2610.
Eiraku, M., Watanabe, K., Matsuo-Takasaki, M., Kawada, M., Yonemura, S., Matsumura, Guerin, M.B., McKernan, D.P., O'Brien, C.J., Cotter, T.G., 2006. Retinal ganglion cells:
M., Wataya, T., Nishiyama, A., Muguruma, K., Sasai, Y., 2008. Self-organized for- dying to survive. Int. J. Dev. Biol. 50, 665–674.
mation of polarized cortical tissues from ESCs and its active manipulation by extrinsic Halder, G., Callaerts, P., Gehring, W.J., 1995. Induction of ectopic eyes by targeted ex-
signals. Cell Stem Cell 3, 519–532. pression of the eyeless gene in Drosophila. Science 267, 1788–1792.
Elliott, J., Jolicoeur, C., Ramamurthy, V., Cayouette, M., 2008. Ikaros confers early Hallam, D., Hilgen, G., Dorgau, B., Zhu, L., Yu, M., Bojic, S., Hewitt, P., Schmitt, M.,
temporal competence to mouse retinal progenitor cells. Neuron 60, 26–39. Uteng, M., Kustermann, S., Steel, D., Nicholds, M., Thomas, R., Treumann, A., Porter,
Erkman, L., McEvilly, R.J., Luo, L., Ryan, A.K., Hooshmand, F., O'Connell, S.M., Keithley, A., Sernagor, E., Armstrong, L., Lako, M., 2018. Human-induced pluripotent stem
E.M., Rapaport, D.H., Ryan, A.F., Rosenfeld, M.G., 1996. Role of transcription factors cells generate light responsive retinal organoids with variable and nutrient-depen-
Brn-3.1 and Brn-3.2 in auditory and visual system development. Nature 381, dent efficiency. Stem Cells 36, 1535–1551.
603–606. Hambright, D., Park, K.Y., Brooks, M., McKay, R., Swaroop, A., Nasonkin, I.O., 2012.
Erkman, L., Yates, P.A., McLaughlin, T., McEvilly, R.J., Whisenhunt, T., O'Connell, S.M., Long-term survival and differentiation of retinal neurons derived from human em-
Krones, A.I., Kirby, M.A., Rapaport, D.H., Bermingham, J.R., O'Leary, D.D., bryonic stem cell lines in un-immunosuppressed mouse retina. Mol. Vis. 18, 920–936.
Rosenfeld, M.G., 2000. A POU domain transcription factor-dependent program reg- Hartong, D.T., Berson, E.L., Dryja, T.P., 2006. Retinitis pigmentosa. Lancet 368,
ulates axon pathfinding in the vertebrate visual system. Neuron 28, 779–792. 1795–1809.
Ezzeddine, Z.D., Yang, X., DeChiara, T., Yancopoulos, G., Cepko, C.L., 1997. Postmitotic Haston, K.M., Finkbeiner, S., 2016. Clinical trials in a dish: the potential of pluripotent
cells fated to become rod photoreceptors can be respecified by CNTF treatment of the stem cells to develop therapies for neurodegenerative diseases. Annu. Rev.
retina. Development 124, 1055–1067. Pharmacol. Toxicol. 56, 489–510.
Fausett, B.V., Gumerson, J.D., Goldman, D., 2008. The proneural basic helix-loop-helix Hatakeyama, J., Kageyama, R., 2004. Retinal cell fate determination and bHLH factors.
gene ascl1a is required for retina regeneration. J. Neurosci. 28, 1109–1117. Semin. Cell Dev. Biol. 15, 83–89.
Ferrari, S., Di Iorio, E., Barbaro, V., Ponzin, D., Sorrentino, F.S., Parmeggiani, F., 2011. Heavner, W., Pevny, L., 2012. Eye development and retinogenesis. Cold Spring Harb
Retinitis pigmentosa: genes and disease mechanisms. Curr. Genom. 12, 238–249. Perspect Biol 4.
Fingert, J.H., Robin, A.L., Stone, J.L., Roos, B.R., Davis, L.K., Scheetz, T.E., Bennett, S.R., Hendrickson, A., Bumsted-O'Brien, K., Natoli, R., Ramamurthy, V., Possin, D., Provis, J.,
Wassink, T.H., Kwon, Y.H., Alward, W.L., Mullins, R.F., Sheffield, V.C., Stone, E.M., 2008. Rod photoreceptor differentiation in fetal and infant human retina. Exp. Eye
2011. Copy number variations on chromosome 12q14 in patients with normal tension Res. 87, 415–426.
glaucoma. Hum. Mol. Genet. 20, 2482–2494. Hiler, D., Chen, X., Hazen, J., Kupriyanov, S., Carroll, P.A., Qu, C., Xu, B., Johnson, D.,
Fischer, A.J., McGuire, C.R., Dierks, B.D., Reh, T.A., 2002. Insulin and fibroblast growth Griffiths, L., Frase, S., Rodriguez, A.R., Martin, G., Zhang, J., Jeon, J., Fan, Y.,
factor 2 activate a neurogenic program in Muller glia of the chicken retina. J. Finkelstein, D., Eisenman, R.N., Baldwin, K., Dyer, M.A., 2015. Quantification of
Neurosci. 22, 9387–9398. retinogenesis in 3D cultures reveals epigenetic memory and higher efficiency in iPSCs
Fischer, A.J., Reh, T.A., 2001. Muller glia are a potential source of neural regeneration in derived from rod photoreceptors. Cell Stem Cell 17, 101–115.
the postnatal chicken retina. Nat. Neurosci. 4, 247–252. Hill, R.E., Favor, J., Hogan, B.L., Ton, C.C., Saunders, G.F., Hanson, I.M., Prosser, J.,
Fischer, A.J., Zelinka, C., Gallina, D., Scott, M.A., Todd, L., 2014. Reactive microglia and Jordan, T., Hastie, N.D., van Heyningen, V., 1991. Mouse small eye results from
macrophage facilitate the formation of Muller glia-derived retinal progenitors. Glia mutations in a paired-like homeobox-containing gene. Nature 354, 522–525.
62, 1608–1628. Hitchcock, P.F., Raymond, P.A., et al., 1992. Retinal Regeneration. Trends Neurosci. 15,
Fuhrmann, S., 2008. Wnt signaling in eye organogenesis. Organogenesis 4, 60–67. 103–108.
Fuhrmann, S., 2010. Eye morphogenesis and patterning of the optic vesicle. Curr. Top. Holt, C.E., Bertsch, T.W., Ellis, H.M., Harris, W.A., 1988. Cellular determination in the
Dev. Biol. 93, 61–84. Xenopus retina is independent of lineage and birth date. Neuron 1, 15–26.
Fujimura, N., 2016. WNT/beta-Catenin signaling in vertebrate eye development. Front Horsford, D.J., Nguyen, M.T., Sellar, G.C., Kothary, R., Arnheiter, H., McInnes, R.R.,
Cell Dev Biol 4, 138. 2005. Chx10 repression of Mitf is required for the maintenance of mammalian neu-
Furukawa, T., Morrow, E.M., Li, T., Davis, F.C., Cepko, C.L., 1999. Retinopathy and at- roretinal identity. Development 132, 177–187.
tenuated circadian entrainment in Crx-deficient mice. Nat. Genet. 23, 466–470. Hoshino, A., Ratnapriya, R., Brooks, M.J., Chaitankar, V., Wilken, M.S., Zhang, C.,
Gagliardi, G., Ben M'Barek, K., Chaffiol, A., Slembrouck-Brec, A., Conart, J.B., Nanteau, Starostik, M.R., Gieser, L., La Torre, A., Nishio, M., Bates, O., Walton, A.,
C., Rabesandratana, O., Sahel, J.A., Duebel, J., Orieux, G., Reichman, S., Goureau, O., Bermingham-McDonogh, O., Glass, I.A., Wong, R.O.L., Swaroop, A., Reh, T.A., 2017.
2018. Characterization and transplantation of CD73-positive photoreceptors isolated Molecular anatomy of the developing human retina. Dev. Cell 43, 763–779 e764.
from human iPSC-derived retinal organoids. Stem Cell Reports 11, 665–680. Hyer, J., Kuhlman, J., Afif, E., Mikawa, T., 2003. Optic cup morphogenesis requires pre-
Gagliardi, G., Ben M’Barek, K., Goureau, O., 2019. Photoreceptor cell replacement in lens ectoderm but not lens differentiation. Dev. Biol. 259, 351–363.
macular degeneration and retinitis pigmentosa: a pluripotent stem cell-based ap- Iglesias, A.I., Springelkamp, H., van der Linde, H., Severijnen, L.A., Amin, N., Oostra, B.,
proach. Prog. Retin. Eye Res. 71, 1–25. Kockx, C.E., van den Hout, M.C., van Ijcken, W.F., Hofman, A., Uitterlinden, A.G.,
Gan, L., Xiang, M., Zhou, L., Wagner, D.S., Klein, W.H., Nathans, J., 1996. POU domain Verdijk, R.M., Klaver, C.C., Willemsen, R., van Duijn, C.M., 2014. Exome sequencing
factor Brn-3b is required for the development of a large set of retinal ganglion cells. and functional analyses suggest that SIX6 is a gene involved in an altered prolifera-
Proc. Natl. Acad. Sci. U. S. A. 93, 3920–3925. tion-differentiation balance early in life and optic nerve degeneration at old age.
Garita-Hernandez, M., Lampic, M., Chaffiol, A., Guibbal, L., Routet, F., Santos-Ferreira, Hum. Mol. Genet. 23, 1320–1332.
T., Gasparini, S., Borsch, O., Gagliardi, G., Reichman, S., Picaud, S., Sahel, J.A., Iwase, A., Suzuki, Y., Araie, M., Yamamoto, T., Abe, H., Shirato, S., Kuwayama, Y.,
Goureau, O., Ader, M., Dalkara, D., Duebel, J., 2019. Restoration of visual function by Mishima, H.K., Shimizu, H., Tomita, G., Inoue, Y., Kitazawa, Y., Tajimi Study Group,
transplantation of optogenetically engineered photoreceptors. Nat. Commun. 10, J.G.S., 2004. The prevalence of primary open-angle glaucoma in Japanese: the Tajimi
4524. Study. Ophthalmology 111, 1641–1648.
Gasparini, S.J., Llonch, S., Borsch, O., Ader, M., 2019. Transplantation of photoreceptors Jadhav, A.P., Mason, H.A., Cepko, C.L., 2006. Notch 1 inhibits photoreceptor production
into the degenerative retina: current state and future perspectives. Prog. Retin. Eye in the developing mammalian retina. Development 133, 913–923.
Res. 69, 1–37. James, J., Das, A.V., Bhattacharya, S., Chacko, D.M., Zhao, X., Ahmad, I., 2003. In vitro
Geng, X., Speirs, C., Lagutin, O., Inbal, A., Liu, W., Solnica-Krezel, L., Jeong, Y., Epstein, generation of early-born neurons from late retinal progenitors. J. Neurosci. 23,
D.J., Oliver, G., 2008. Haploinsufficiency of Six3 fails to activate Sonic hedgehog 8193–8203.
expression in the ventral forebrain and causes holoprosencephaly. Dev. Cell 15, James, J., Das, A.V., Rahnenfuhrer, J., Ahmad, I., 2004. Cellular and molecular char-
236–247. acterization of early and late retinal stem cells/progenitors: differential regulation of
Ghosh, A.K., Murga-Zamalloa, C.A., Chan, L., Hitchcock, P.F., Swaroop, A., Khanna, H., proliferation and context dependent role of Notch signaling. J. Neurobiol. 61,
2010. Human retinopathy-associated ciliary protein retinitis pigmentosa GTPase 359–376.
regulator mediates cilia-dependent vertebrate development. Hum. Mol. Genet. 19, Janssen, S.F., Gorgels, T.G., Ramdas, W.D., Klaver, C.C., van Duijn, C.M., Jansonius, N.M.,
90–98. Bergen, A.A., 2013. The vast complexity of primary open angle glaucoma: disease
Glaser, T., Walton, D.S., Maas, R.L., 1992. Genomic structure, evolutionary conservation genes, risks, molecular mechanisms and pathobiology. Prog. Retin. Eye Res. 37,
and aniridia mutations in the human PAX6 gene. Nat. Genet. 2, 232–239. 31–67.
Goldman, D., 2014. Muller glial cell reprogramming and retina regeneration. Nat. Rev. Jeong, Y., Leskow, F.C., El-Jaick, K., Roessler, E., Muenke, M., Yocum, A., Dubourg, C., Li,
Neurosci. 15, 431–442. X., Geng, X., Oliver, G., Epstein, D.J., 2008. Regulation of a remote Shh forebrain
Gonzalez-Cordero, A., Kruczek, K., Naeem, A., Fernando, M., Kloc, M., Ribeiro, J., Goh, enhancer by the Six3 homeoprotein. Nat. Genet. 40, 1348–1353.
D., Duran, Y., Blackford, S.J.I., Abelleira-Hervas, L., Sampson, R.D., Shum, I.O., Ji, J.Z., Elyaman, W., Yip, H.K., Lee, V.W., Yick, L.W., Hugon, J., So, K.F., 2004. CNTF
Branch, M.J., Gardner, P.J., Sowden, J.C., Bainbridge, J.W.B., Smith, A.J., West, E.L., promotes survival of retinal ganglion cells after induction of ocular hypertension in
Pearson, R.A., Ali, R.R., 2017. Recapitulation of human retinal development from rats: the possible involvement of STAT3 pathway. Eur. J. Neurosci. 19, 265–272.
human pluripotent stem cells generates transplantable populations of cone photo- Jin, Z.B., Okamoto, S., Osakada, F., Homma, K., Assawachananont, J., Hirami, Y., Iwata,
receptors. Stem Cell Reports 9, 820–837. T., Takahashi, M., 2011. Modeling retinal degeneration using patient-specific induced

26
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

pluripotent stem cells. PLoS One 6, e17084. Lenkowski, J.R., Raymond, P.A., 2014. Muller glia: stem cells for generation and re-
Jin, Z.B., Okamoto, S., Xiang, P., Takahashi, M., 2012. Integration-free induced plur- generation of retinal neurons in teleost fish. Prog. Retin. Eye Res. 40, 94–123.
ipotent stem cells derived from retinitis pigmentosa patient for disease modeling. Levine, E.M., Fuhrmann, S., Reh, T.A., 2000. Soluble factors and the development of rod
Stem Cells Transl Med 1, 503–509. photoreceptors. Cell. Mol. Life Sci. 57, 224–234.
Johns, P.R., 1982. Formation of photoreceptors in larval and adult goldfish. J. Neurosci. Levine, E.M., Roelink, H., Turner, J., Reh, T.A., 1997. Sonic hedgehog promotes rod
2, 178–198. photoreceptor differentiation in mammalian retinal cells in vitro. J. Neurosci. 17,
Jorstad, N.L., Wilken, M.S., Grimes, W.N., Wohl, S.G., VandenBosch, L.S., Yoshimatsu, T., 6277–6288.
Wong, R.O., Rieke, F., Reh, T.A., 2017. Stimulation of functional neuronal re- Li, K., Zhong, X., Yang, S., Luo, Z., Li, K., Liu, Y., Cai, S., Gu, H., Lu, S., Zhang, H., Wei, Y.,
generation from Muller glia in adult mice. Nature 548, 103–107. Zhuang, J., Zhuo, Y., Fan, Z., Ge, J., 2017. HiPSC-derived retinal ganglion cells grow
Julian, D., Ennis, K., Korenbrot, J.I., 1998. Birth and fate of proliferative cells in the inner dendritic arbors and functional axons on a tissue-engineered scaffold. Acta Biomater.
nuclear layer of the mature fish retina. J. Comp. Neurol. 394, 271–282. 54, 117–127.
Karl, M.O., Hayes, S., Nelson, B.R., Tan, K., Buckingham, B., Reh, T.A., 2008. Stimulation Li, M.L., Aggeler, J., Farson, D.A., Hatier, C., Hassell, J., Bissell, M.J., 1987. Influence of a
of neural regeneration in the mouse retina. Proc. Natl. Acad. Sci. U. S. A. 105, reconstituted basement membrane and its components on casein gene expression and
19508–19513. secretion in mouse mammary epithelial cells. Proc. Natl. Acad. Sci. U. S. A. 84,
Kay, J.N., Finger-Baier, K.C., Roeser, T., Staub, W., Baier, H., 2001. Retinal ganglion cell 136–140.
genesis requires lakritz, a Zebrafish atonal Homolog. Neuron 30, 725–736. Li, X., Perissi, V., Liu, F., Rose, D.W., Rosenfeld, M.G., 2002. Tissue-specific regulation of
Kelley, M.W., Williams, R.C., Turner, J.K., Creech-Kraft, J.M., Reh, T.A., 1999. Retinoic retinal and pituitary precursor cell proliferation. Science 297, 1180–1183.
acid promotes rod photoreceptor differentiation in rat retina in vivo. Neuroreport 10, Lindsley, R.C., Gill, J.G., Kyba, M., Murphy, T.L., Murphy, K.M., 2006. Canonical Wnt
2389–2394. signaling is required for development of embryonic stem cell-derived mesoderm.
Khanna, H., Akimoto, M., Siffroi-Fernandez, S., Friedman, J.S., Hicks, D., Swaroop, A., Development 133, 3787–3796.
2006. Retinoic acid regulates the expression of photoreceptor transcription factor Lingor, P., Tonges, L., Pieper, N., Bermel, C., Barski, E., Planchamp, V., Bahr, M., 2008.
NRL. J. Biol. Chem. 281, 27327–27334. ROCK inhibition and CNTF interact on intrinsic signalling pathways and differentially
Kim, J., Wu, H.H., Lander, A.D., Lyons, K.M., Matzuk, M.M., Calof, A.L., 2005. GDF11 regulate survival and regeneration in retinal ganglion cells. Brain 131, 250–263.
controls the timing of progenitor cell competence in developing retina. Science 308, Liu, B., Hunter, D.J., Rooker, S., Chan, A., Paulus, Y.M., Leucht, P., Nusse, Y., Nomoto, H.,
1927–1930. Helms, J.A., 2013. Wnt signaling promotes Muller cell proliferation and survival after
Koso, H., Minami, C., Tabata, Y., Inoue, M., Sasaki, E., Satoh, S., Watanabe, S., 2009. injury. Investig. Ophthalmol. Vis. Sci. 54, 444–453.
CD73, a novel cell surface antigen that characterizes retinal photoreceptor precursor Liu, C., Oikonomopoulos, A., Sayed, N., Wu, J.C., 2018. Modeling human diseases with
cells. Investig. Ophthalmol. Vis. Sci. 50, 5411–5418. induced pluripotent stem cells: from 2D to 3D and beyond. Development 145.
Koso, H., Satoh, S., Watanabe, S., 2007. c-kit marks late retinal progenitor cells and Liu, W., Lagutin, O., Swindell, E., Jamrich, M., Oliver, G., 2010. Neuroretina specification
regulates their differentiation in developing mouse retina. Dev. Biol. 301, 141–154. in mouse embryos requires Six3-mediated suppression of Wnt8b in the anterior
Kriegstein, A., Alvarez-Buylla, A., 2009. The glial nature of embryonic and adult neural neural plate. J. Clin. Investig. 120, 3568–3577.
stem cells. Annu. Rev. Neurosci. 32, 149–184. Livesey, F.J., Cepko, C.L., 2001. Vertebrate neural cell-fate determination: lessons from
Kruczek, K., Gonzalez-Cordero, A., Goh, D., Naeem, A., Jonikas, M., Blackford, S.J.I., the retina. Nat. Rev. Neurosci. 2, 109–118.
Kloc, M., Duran, Y., Georgiadis, A., Sampson, R.D., Maswood, R.N., Smith, A.J., Llonch, S., Carido, M., Ader, M., 2018. Organoid technology for retinal repair. Dev. Biol.
Decembrini, S., Arsenijevic, Y., Sowden, J.C., Pearson, R.A., West, E.L., Ali, R.R., 433, 132–143.
2017. Differentiation and transplantation of embryonic stem cell-derived cone pho- Longobardi, E., Penkov, D., Mateos, D., De Florian, G., Torres, M., Blasi, F., 2014.
toreceptors into a mouse model of end-stage retinal degeneration. Stem Cell Reports Biochemistry of the tale transcription factors PREP, MEIS, and PBX in vertebrates.
8, 1659–1674. Dev. Dynam. 243, 59–75.
Kugler, M., Schlecht, A., Fuchshofer, R., Kleiter, I., Aigner, L., Tamm, E.R., Braunger, Loosli, F., Staub, W., Finger-Baier, K.C., Ober, E.A., Verkade, H., Wittbrodt, J., Baier, H.,
B.M., 2015. Heterozygous modulation of TGF-beta signaling does not influence 2003. Loss of eyes in zebrafish caused by mutation of chokh/rx3. EMBO Rep. 4,
Muller glia cell reactivity or proliferation following NMDA-induced damage. 894–899.
Histochem. Cell Biol. 144, 443–455. Loosli, F., Winkler, S., Wittbrodt, J., 1999. Six3 overexpression initiates the formation of
Kuwahara, A., Ozone, C., Nakano, T., Saito, K., Eiraku, M., Sasai, Y., 2015. Generation of ectopic retina. Genes Dev. 13, 649–654.
a ciliary margin-like stem cell niche from self-organizing human retinal tissue. Nat. Lowe, A., Harris, R., Bhansali, P., Cvekl, A., Liu, W., 2016. Intercellular adhesion-de-
Commun. 6, 6286. pendent cell survival and ROCK-regulated actomyosin-driven forces mediate self-
Kwon, Y.H., Fingert, J.H., Kuehn, M.H., Alward, W.L., 2009. Primary open-angle glau- formation of a retinal organoid. Stem Cell Reports 6, 743–756.
coma. N. Engl. J. Med. 360, 1113–1124. Lu, A., Ng, L., Ma, M., Kefas, B., Davies, T.F., Hernandez, A., Chan, C.C., Forrest, D., 2009.
Lagutin, O., Zhu, C.C., Furuta, Y., Rowitch, D.H., McMahon, A.P., Oliver, G., 2001. Six3 Retarded developmental expression and patterning of retinal cone opsins in hy-
promotes the formation of ectopic optic vesicle-like structures in mouse embryos. pothyroid mice. Endocrinology 150, 1536–1544.
Dev. Dynam. 221, 342–349. Ma, W., Yan, R.T., Xie, W., Wang, S.Z., 2004. A role of ath5 in inducing neuroD and the
Lagutin, O.V., Zhu, C.C., Kobayashi, D., Topczewski, J., Shimamura, K., Puelles, L., photoreceptor pathway. J. Neurosci. 24, 7150–7158.
Russell, H.R., McKinnon, P.J., Solnica-Krezel, L., Oliver, G., 2003. Six3 repression of MacLaren, R.E., Pearson, R.A., MacNeil, A., Douglas, R.H., Salt, T.E., Akimoto, M.,
Wnt signaling in the anterior neuroectoderm is essential for vertebrate forebrain Swaroop, A., Sowden, J.C., Ali, R.R., 2006. Retinal repair by transplantation of
development. Genes Dev. 17, 368–379. photoreceptor precursors. Nature 444, 203–207.
Lakowski, J., Gonzalez-Cordero, A., West, E.L., Han, Y.T., Welby, E., Naeem, A., Marcos, S., Gonzalez-Lazaro, M., Beccari, L., Carramolino, L., Martin-Bermejo, M.J.,
Blackford, S.J., Bainbridge, J.W., Pearson, R.A., Ali, R.R., Sowden, J.C., 2015. Amarie, O., Mateos-San Martin, D., Torroja, C., Bogdanovic, O., Doohan, R., Puk, O.,
Transplantation of photoreceptor precursors isolated via a cell surface biomarker Hrabe de Angelis, M., Graw, J., Gomez-Skarmeta, J.L., Casares, F., Torres, M.,
panel from embryonic stem cell-derived self-forming retina. Stem Cells 33, Bovolenta, P., 2015. Meis1 coordinates a network of genes implicated in eye devel-
2469–2482. opment and microphthalmia. Development 142, 3009–3020.
Lakowski, J., Han, Y.T., Pearson, R.A., Gonzalez-Cordero, A., West, E.L., Gualdoni, S., Marquardt, T., Gruss, P., 2002. Generating neuronal diversity in the retina: one for nearly
Barber, A.C., Hubank, M., Ali, R.R., Sowden, J.C., 2011. Effective transplantation of all. Trends Neurosci. 25, 32–38.
photoreceptor precursor cells selected via cell surface antigen expression. Stem Cells Martinez-Morales, J.R., Del Bene, F., Nica, G., Hammerschmidt, M., Bovolenta, P.,
29, 1391–1404. Wittbrodt, J., 2005. Differentiation of the vertebrate retina is coordinated by an FGF
Lakowski, J., Welby, E., Budinger, D., Di Marco, F., Di Foggia, V., Bainbridge, J.W.B., signaling center. Dev. Cell 8, 565–574.
Wallace, K., Gamm, D.M., Ali, R.R., Sowden, J.C., 2018. Isolation of human photo- Martinez-Morales, J.R., Dolez, V., Rodrigo, I., Zaccarini, R., Leconte, L., Bovolenta, P.,
receptor precursors via a cell surface marker panel from stem cell-derived retinal Saule, S., 2003. OTX2 activates the molecular network underlying retina pigment
organoids and fetal retinae. Stem Cells 36, 709–722. epithelium differentiation. J. Biol. Chem. 278, 21721–21731.
Lamba, D.A., Gust, J., Reh, T.A., 2009. Transplantation of human embryonic stem cell- Martinez-Morales, J.R., Rembold, M., Greger, K., Simpson, J.C., Brown, K.E., Quiring, R.,
derived photoreceptors restores some visual function in Crx-deficient mice. Cell Stem Pepperkok, R., Martin-Bermudo, M.D., Himmelbauer, H., Wittbrodt, J., 2009. ojo-
Cell 4, 73–79. plano-mediated basal constriction is essential for optic cup morphogenesis.
Lamba, D.A., Karl, M.O., Ware, C.B., Reh, T.A., 2006. Efficient generation of retinal Development 136, 2165–2175.
progenitor cells from human embryonic stem cells. Proc. Natl. Acad. Sci. U. S. A. 103, Mathers, P.H., Grinberg, A., Mahon, K.A., Jamrich, M., 1997. The Rx homeobox gene is
12769–12774. essential for vertebrate eye development. Nature 387, 603–607.
Lamba, D.A., McUsic, A., Hirata, R.K., Wang, P.R., Russell, D., Reh, T.A., 2010. Matsunaga, M., Hatta, K., Takeichi, M., 1988. Role of N-cadherin cell adhesion molecules
Generation, purification and transplantation of photoreceptors derived from human in the histogenesis of neural retina. Neuron 1, 289–295.
induced pluripotent stem cells. PLoS One 5, e8763. McCabe, K.L., Gunther, E.C., Reh, T.A., 1999. The development of the pattern of retinal
Lancaster, M.A., Huch, M., 2019. Disease modelling in human organoids. Dis Model ganglion cells in the chick retina: mechanisms that control differentiation.
Mech 12. Development 126, 5713–5724.
Lancaster, M.A., Knoblich, J.A., 2014. Organogenesis in a dish: modeling development McFarlane, S., Zuber, M.E., Holt, C.E., 1998. A role for the fibroblast growth factor re-
and disease using organoid technologies. Science 345, 1247125. ceptor in cell fate decisions in the developing vertebrate retina. Development 125,
Langer, K.B., Ohlemacher, S.K., Phillips, M.J., Fligor, C.M., Jiang, P., Gamm, D.M., Meyer, 3967–3975.
J.S., 2018. Retinal ganglion cell diversity and subtype specification from human Mears, A.J., Kondo, M., Swain, P.K., Takada, Y., Bush, R.A., Saunders, T.L., Sieving, P.A.,
pluripotent stem cells. Stem Cell Reports 10, 1282–1293. Swaroop, A., 2001. Nrl is required for rod photoreceptor development. Nat. Genet.
Le, T.T., Wroblewski, E., Patel, S., Riesenberg, A.N., Brown, N.L., 2006. Math5 is required 29, 447–452.
for both early retinal neuron differentiation and cell cycle progression. Dev. Biol. 295, Megaw, R., Abu-Arafeh, H., Jungnickel, M., Mellough, C., Gurniak, C., Witke, W., Zhang,
764–778. W., Khanna, H., Mill, P., Dhillon, B., Wright, A.F., Lako, M., Ffrench-Constant, C.,

27
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

2017. Gelsolin dysfunction causes photoreceptor loss in induced pluripotent cell and regulator UHRF1 inactivates REST and growth suppressor gene expression via DNA
animal retinitis pigmentosa models. Nat. Commun. 8, 271. methylation to promote axon regeneration. Proc. Natl. Acad. Sci. U. S. A. 115,
Mellough, C.B., Collin, J., Queen, R., Hilgen, G., Dorgau, B., Zerti, D., Felemban, M., E12417–E12426.
White, K., Sernagor, E., Lako, M., 2019. Systematic comparison of retinal organoid Ohgushi, M., Matsumura, M., Eiraku, M., Murakami, K., Aramaki, T., Nishiyama, A.,
differentiation from human pluripotent stem cells reveals stage specific, cell line, and Muguruma, K., Nakano, T., Suga, H., Ueno, M., Ishizaki, T., Suemori, H., Narumiya,
methodological differences. Stem Cells Transl Med 8, 694–706. S., Niwa, H., Sasai, Y., 2010. Molecular pathway and cell state responsible for dis-
Mellough, C.B., Sernagor, E., Moreno-Gimeno, I., Steel, D.H., Lako, M., 2012. Efficient sociation-induced apoptosis in human pluripotent stem cells. Cell Stem Cell 7,
stage-specific differentiation of human pluripotent stem cells toward retinal photo- 225–239.
receptor cells. Stem Cells 30, 673–686. Ohlemacher, S.K., Sridhar, A., Xiao, Y., Hochstetler, A.E., Sarfarazi, M., Cummins, T.R.,
Meyer, J.S., Shearer, R.L., Capowski, E.E., Wright, L.S., Wallace, K.A., McMillan, E.L., Meyer, J.S., 2016. Stepwise differentiation of retinal ganglion cells from human
Zhang, S.C., Gamm, D.M., 2009. Modeling early retinal development with human pluripotent stem cells enables analysis of glaucomatous neurodegeneration. Stem
embryonic and induced pluripotent stem cells. Proc. Natl. Acad. Sci. U. S. A. 106, Cells 34, 1553–1562.
16698–16703. Okabe, S., Forsberg-Nilsson, K., Spiro, A.C., Segal, M., McKay, R.D., 1996. Development of
Meyer, K., Ferraiuolo, L., Miranda, C.J., Likhite, S., McElroy, S., Renusch, S., Ditsworth, neuronal precursor cells and functional postmitotic neurons from embryonic stem
D., Lagier-Tourenne, C., Smith, R.A., Ravits, J., Burghes, A.H., Shaw, P.J., Cleveland, cells in vitro. Mech. Dev. 59, 89–102.
D.W., Kolb, S.J., Kaspar, B.K., 2014. Direct conversion of patient fibroblasts de- Ooto, S., Akagi, T., Kageyama, R., Akita, J., Mandai, M., Honda, Y., Takahashi, M., 2004.
monstrates non-cell autonomous toxicity of astrocytes to motor neurons in familial Potential for neural regeneration after neurotoxic injury in the adult mammalian
and sporadic ALS. Proc. Natl. Acad. Sci. U. S. A. 111, 829–832. retina. Proc. Natl. Acad. Sci. U. S. A. 101, 13654–13659.
Meyer-Franke, A., Kaplan, M.R., Pfrieger, F.W., Barres, B.A., 1995. Characterization of the Osakada, F., Ikeda, H., Mandai, M., Wataya, T., Watanabe, K., Yoshimura, N., Akaike, A.,
signaling interactions that promote the survival and growth of developing retinal Sasai, Y., Takahashi, M., 2008. Toward the generation of rod and cone photoreceptors
ganglion cells in culture. Neuron 15, 805–819. from mouse, monkey and human embryonic stem cells. Nat. Biotechnol. 26, 215–224.
Mic, F.A., Molotkov, A., Molotkova, N., Duester, G., 2004. Raldh2 expression in optic Osakada, F., Ikeda, H., Sasai, Y., Takahashi, M., 2009a. Stepwise differentiation of plur-
vesicle generates a retinoic acid signal needed for invagination of retina during optic ipotent stem cells into retinal cells. Nat. Protoc. 4, 811–824.
cup formation. Dev. Dynam. 231, 270–277. Osakada, F., Jin, Z.B., Hirami, Y., Ikeda, H., Danjyo, T., Watanabe, K., Sasai, Y.,
Miesfeld, J.B., Moon, M.S., Riesenberg, A.N., Contreras, A.N., Kovall, R.A., Brown, N.L., Takahashi, M., 2009b. In vitro differentiation of retinal cells from human pluripotent
2018. Rbpj direct regulation of Atoh7 transcription in the embryonic mouse retina. stem cells by small-molecule induction. J. Cell Sci. 122, 3169–3179.
Sci. Rep. 8, 10195. Osakada, F., Ooto, S., Akagi, T., Mandai, M., Akaike, A., Takahashi, M., 2007. Wnt sig-
Miller, J.D., Ganat, Y.M., Kishinevsky, S., Bowman, R.L., Liu, B., Tu, E.Y., Mandal, P.K., naling promotes regeneration in the retina of adult mammals. J. Neurosci. 27,
Vera, E., Shim, J.W., Kriks, S., Taldone, T., Fusaki, N., Tomishima, M.J., Krainc, D., 4210–4219.
Milner, T.A., Rossi, D.J., Studer, L., 2013. Human iPSC-based modeling of late-onset Packer, A.N., Xing, Y., Harper, S.Q., Jones, L., Davidson, B.L., 2008. The bifunctional
disease via progerin-induced aging. Cell Stem Cell 13, 691–705. microRNA miR-9/miR-9* regulates REST and CoREST and is downregulated in
Minegishi, Y., Iejima, D., Kobayashi, H., Chi, Z.L., Kawase, K., Yamamoto, T., Seki, T., Huntington's disease. J. Neurosci. 28, 14341–14346.
Yuasa, S., Fukuda, K., Iwata, T., 2013. Enhanced optineurin E50K-TBK1 interaction Pan, L., Deng, M., Xie, X., Gan, L., 2008. ISL1 and BRN3B co-regulate the differentiation
evokes protein insolubility and initiates familial primary open-angle glaucoma. Hum. of murine retinal ganglion cells. Development 135, 1981–1990.
Mol. Genet. 22, 3559–3567. Pankratz, M.T., Li, X.J., Lavaute, T.M., Lyons, E.A., Chen, X., Zhang, S.C., 2007. Directed
Moore, K.B., Mood, K., Daar, I.O., Moody, S.A., 2004. Morphogenetic movements un- neural differentiation of human embryonic stem cells via an obligated primitive
derlying eye field formation require interactions between the FGF and ephrinB1 anterior stage. Stem Cells 25, 1511–1520.
signaling pathways. Dev. Cell 6, 55–67. Parameswaran, S., Balasubramanian, S., Babai, N., Qiu, F., Eudy, J.D., Thoreson, W.B.,
Morquette, B., Morquette, P., Agostinone, J., Feinstein, E., McKinney, R.A., Kolta, A., Di Ahmad, I., 2010. Induced pluripotent stem cells generate both retinal ganglion cells
Polo, A., 2015. REDD2-mediated inhibition of mTOR promotes dendrite retraction and photoreceptors: therapeutic implications in degenerative changes in glaucoma
induced by axonal injury. Cell Death Differ. 22, 612–625. and age-related macular degeneration. Stem Cells 28, 695–703.
Morrison, J.C., Moore, C.G., Deppmeier, L.M., Gold, B.G., Meshul, C.K., Johnson, E.C., Parameswaran, S., Dravid, S.M., Teotia, P., Krishnamoorthy, R.R., Qiu, F., Toris, C.,
1997. A rat model of chronic pressure-induced optic nerve damage. Exp. Eye Res. 64, Morrison, J., Ahmad, I., 2015. Continuous non-cell autonomous reprogramming to
85–96. generate retinal ganglion cells for glaucomatous neuropathy. Stem Cells 33,
Morrow, E.M., Furukawa, T., Lee, J.E., Cepko, C.L., 1999. NeuroD regulates multiple 1743–1758.
functions in the developing neural retina in rodent. Development 126, 23–36. Parameswaran, S., Xia, X., Hegde, G., Ahmad, I., 2014. Hmga2 regulates self-renewal of
Munoz-Sanjuan, I., Brivanlou, A.H., 2002. Neural induction, the default model and em- retinal progenitors. Development 141, 4087–4097.
bryonic stem cells. Nat. Rev. Neurosci. 3, 271–280. Parfitt, D.A., Lane, A., Ramsden, C.M., Carr, A.J., Munro, P.M., Jovanovic, K., Schwarz,
Murga-Zamalloa, C.A., Atkins, S.J., Peranen, J., Swaroop, A., Khanna, H., 2010. N., Kanuga, N., Muthiah, M.N., Hull, S., Gallo, J.M., da Cruz, L., Moore, A.T.,
Interaction of retinitis pigmentosa GTPase regulator (RPGR) with RAB8A GTPase: Hardcastle, A.J., Coffey, P.J., Cheetham, M.E., 2016. Identification and correction of
implications for cilia dysfunction and photoreceptor degeneration. Hum. Mol. Genet. mechanisms underlying inherited blindness in human iPSC-derived optic cups. Cell
19, 3591–3598. Stem Cell 18, 769–781.
Nakano, T., Ando, S., Takata, N., Kawada, M., Muguruma, K., Sekiguchi, K., Saito, K., Park, K.K., Liu, K., Hu, Y., Smith, P.D., Wang, C., Cai, B., Xu, B., Connolly, L., Kramvis, I.,
Yonemura, S., Eiraku, M., Sasai, Y., 2012. Self-formation of optic cups and storable Sahin, M., He, Z., 2008. Promoting axon regeneration in the adult CNS by modulation
stratified neural retina from human ESCs. Cell Stem Cell 10, 771–785. of the PTEN/mTOR pathway. Science 322, 963–966.
Nasonkin, I.O., Merbs, S.L., Lazo, K., Oliver, V.F., Brooks, M., Patel, K., Enke, R.A., Park, S.E., Georgescu, A., Huh, D., 2019. Organoids-on-a-chip. Science 364, 960–965.
Nellissery, J., Jamrich, M., Le, Y.Z., Bharti, K., Fariss, R.N., Rachel, R.A., Zack, D.J., Pearson, R.A., Gonzalez-Cordero, A., West, E.L., Ribeiro, J.R., Aghaizu, N., Goh, D.,
Rodriguez-Boulan, E.J., Swaroop, A., 2013. Conditional knockdown of DNA me- Sampson, R.D., Georgiadis, A., Waldron, P.V., Duran, Y., Naeem, A., Kloc, M.,
thyltransferase 1 reveals a key role of retinal pigment epithelium integrity in pho- Cristante, E., Kruczek, K., Warre-Cornish, K., Sowden, J.C., Smith, A.J., Ali, R.R.,
toreceptor outer segment morphogenesis. Development 140, 1330–1341. 2016. Donor and host photoreceptors engage in material transfer following trans-
Nelson, B.R., Gumuscu, B., Hartman, B.H., Reh, T.A., 2006. Notch activity is down- plantation of post-mitotic photoreceptor precursors. Nat. Commun. 7, 13029.
regulated just prior to retinal ganglion cell differentiation. Dev. Neurosci. 28, Pera, E.M., Wessely, O., Li, S.Y., De Robertis, E.M., 2001. Neural and head induction by
128–141. insulin-like growth factor signals. Dev. Cell 1, 655–665.
Neumann, C.J., Nuesslein-Volhard, C., 2000. Patterning of the zebrafish retina by a wave Pera, M.F., Andrade, J., Houssami, S., Reubinoff, B., Trounson, A., Stanley, E.G., Ward-
of sonic hedgehog activity. Science 289, 2137–2139. van Oostwaard, D., Mummery, C., 2004. Regulation of human embryonic stem cell
Ng, L., Hurley, J.B., Dierks, B., Srinivas, M., Salto, C., Vennstrom, B., Reh, T.A., Forrest, differentiation by BMP-2 and its antagonist noggin. J. Cell Sci. 117, 1269–1280.
D., 2001. A thyroid hormone receptor that is required for the development of green Phillips, M.J., Perez, E.T., Martin, J.M., Reshel, S.T., Wallace, K.A., Capowski, E.E., Singh,
cone photoreceptors. Nat. Genet. 27, 94–98. R., Wright, L.S., Clark, E.M., Barney, P.M., Stewart, R., Dickerson, S.J., Miller, M.J.,
Nickerson, P.E.B., Ortin-Martinez, A., Wallace, V.A., 2018. Material exchange in photo- Percin, E.F., Thomson, J.A., Gamm, D.M., 2014. Modeling human retinal develop-
receptor transplantation: updating our understanding of donor/host communication ment with patient-specific induced pluripotent stem cells reveals multiple roles for
and the future of cell engraftment science. Front. Neural Circuits 12, 17. visual system homeobox 2. Stem Cells 32, 1480–1492.
Nishida, A., 2005. [Mechanisms of retinal photoreceptor cell fate determination]. Nippon. Pittler, S.J., Zhang, Y., Chen, S., Mears, A.J., Zack, D.J., Ren, Z., Swain, P.K., Yao, S.,
Ganka Gakkai Zasshi 109, 708–716. Swaroop, A., White, J.B., 2004. Functional analysis of the rod photoreceptor cGMP
Nishida, A., Furukawa, A., Koike, C., Tano, Y., Aizawa, S., Matsuo, I., Furukawa, T., 2003. phosphodiesterase alpha-subunit gene promoter: nrl and Crx are required for full
Otx2 homeobox gene controls retinal photoreceptor cell fate and pineal gland de- transcriptional activity. J. Biol. Chem. 279, 19800–19807.
velopment. Nat. Neurosci. 6, 1255–1263. Pollak, J., Wilken, M.S., Ueki, Y., Cox, K.E., Sullivan, J.M., Taylor, R.J., Levine, E.M., Reh,
Nishino, J., Kim, I., Chada, K., Morrison, S.J., 2008. Hmga2 promotes neural stem cell T.A., 2013. ASCL1 reprograms mouse Muller glia into neurogenic retinal progenitors.
self-renewal in young but not old mice by reducing p16Ink4a and p19Arf Expression. Development 140, 2619–2631.
Cell 135, 227–239. Porter, F.D., Drago, J., Xu, Y., Cheema, S.S., Wassif, C., Huang, S.P., Lee, E., Grinberg, A.,
Noctor, S.C., Flint, A.C., Weissman, T.A., Dammerman, R.S., Kriegstein, A.R., 2001. Massalas, J.S., Bodine, D., Alt, F., Westphal, H., 1997. Lhx2, a LIM homeobox gene, is
Neurons derived from radial glial cells establish radial units in neocortex. Nature 409, required for eye, forebrain, and definitive erythrocyte development. Development
714–720. 124, 2935–2944.
Oh, E.C., Cheng, H., Hao, H., Jia, L., Khan, N.W., Swaroop, A., 2008. Rod differentiation Quadrato, G., Nguyen, T., Macosko, E.Z., Sherwood, J.L., Min Yang, S., Berger, D.R.,
factor NRL activates the expression of nuclear receptor NR2E3 to suppress the de- Maria, N., Scholvin, J., Goldman, M., Kinney, J.P., Boyden, E.S., Lichtman, J.W.,
velopment of cone photoreceptors. Brain Res. 1236, 16–29. Williams, Z.M., McCarroll, S.A., Arlotta, P., 2017. Cell diversity and network dy-
Oh, Y.M., Mahar, M., Ewan, E.E., Leahy, K.M., Zhao, G., Cavalli, V., 2018. Epigenetic namics in photosensitive human brain organoids. Nature 545, 48–53.

28
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

Quinn, P.M., Buck, T.M., Mulder, A.A., Ohonin, C., Alves, C.H., Vos, R.M., Bialecka, M., regulate the trafficking of ciliary tip kinesins. Hum. Mol. Genet. 26, 2480–2492.
van Herwaarden, T., van Dijk, E.H.C., Talib, M., Freund, C., Mikkers, H.M.M., Sharma, T.P., Wiley, L.A., Whitmore, S.S., Anfinson, K.R., Cranston, C.M., Oppedal, D.J.,
Hoeben, R.C., Goumans, M.J., Boon, C.J.F., Koster, A.J., Chuva de Sousa Lopes, S.M., Daggett, H.T., Mullins, R.F., Tucker, B.A., Stone, E.M., 2017. Patient-specific induced
Jost, C.R., Wijnholds, J., 2019. Human iPSC-derived retinas recapitulate the fetal pluripotent stem cells to evaluate the pathophysiology of TRNT1-associated Retinitis
CRB1 CRB2 complex formation and demonstrate that photoreceptors and muller glia pigmentosa. Stem Cell Res. 21, 58–70.
are targets of AAV5. Stem Cell Reports 12, 906–919. Shi, Y., Inoue, H., Wu, J.C., Yamanaka, S., 2017. Induced pluripotent stem cell tech-
Quiring, R., Walldorf, U., Kloter, U., Gehring, W.J., 1994. Homology of the eyeless gene of nology: a decade of progress. Nat. Rev. Drug Discov. 16, 115–130.
Drosophila to the Small eye gene in mice and Aniridia in humans. Science 265, Shimada, H., Lu, Q., Insinna-Kettenhofen, C., Nagashima, K., English, M.A., Semler, E.M.,
785–789. Mahgerefteh, J., Cideciyan, A.V., Li, T., Brooks, B.P., Gunay-Aygun, M., Jacobson,
Qureshi, I.A., Gokhan, S., Mehler, M.F., 2010. REST and CoREST are transcriptional and S.G., Cogliati, T., Westlake, C.J., Swaroop, A., 2017. In vitro modeling using cilio-
epigenetic regulators of seminal neural fate decisions. Cell Cycle 9, 4477–4486. pathy-patient-derived cells reveals distinct cilia dysfunctions caused by CEP290
Ramachandran, R., Fausett, B.V., Goldman, D., 2010a. Ascl1a regulates Muller glia mutations. Cell Rep. 20, 384–396.
dedifferentiation and retinal regeneration through a Lin-28-dependent, let-7 Shindou, H., Koso, H., Sasaki, J., Nakanishi, H., Sagara, H., Nakagawa, K.M., Takahashi,
microRNA signalling pathway. Nat. Cell Biol. 12, 1101–1107. Y., Hishikawa, D., Iizuka-Hishikawa, Y., Tokumasu, F., Noguchi, H., Watanabe, S.,
Ramachandran, R., Reifler, A., Parent, J.M., Goldman, D., 2010b. Conditional gene ex- Sasaki, T., Shimizu, T., 2017. Docosahexaenoic acid preserves visual function by
pression and lineage tracing of tuba1a expressing cells during zebrafish development maintaining correct disc morphology in retinal photoreceptor cells. J. Biol. Chem.
and retina regeneration. J. Comp. Neurol. 518, 4196–4212. 292, 12054–12064.
Ramachandran, R., Zhao, X.F., Goldman, D., 2011. Ascl1a/Dkk/beta-catenin signaling Silverman, S.M., Wong, W.T., 2018. Microglia in the retina: roles in development, ma-
pathway is necessary and glycogen synthase kinase-3beta inhibition is sufficient for turity, and disease. Annu Rev Vis Sci 4, 45–77.
zebrafish retina regeneration. Proc. Natl. Acad. Sci. U. S. A. 108, 15858–15863. Singh, M.S., Balmer, J., Barnard, A.R., Aslam, S.A., Moralli, D., Green, C.M., Barnea-
Rapaport, D.H., Wong, L.L., Wood, E.D., Yasumura, D., LaVail, M.M., 2004. Timing and Cramer, A., Duncan, I., MacLaren, R.E., 2016. Transplanted photoreceptor precursors
topography of cell genesis in the rat retina. J. Comp. Neurol. 474, 304–324. transfer proteins to host photoreceptors by a mechanism of cytoplasmic fusion. Nat.
Rathnasamy, G., Foulds, W.S., Ling, E.A., Kaur, C., 2019. Retinal microglia - a key player Commun. 7, 13537.
in healthy and diseased retina. Prog. Neurobiol. 173, 18–40. Sinn, R., Wittbrodt, J., 2013. An eye on eye development. Mech. Dev. 130, 347–358.
Raymond A., P., Reifler J., M., Rivlin K., P., et al., 1988. Regeneration of goldfish retina: Sison, S.L., Vermilyea, S.C., Emborg, M.E., Ebert, A.D., 2018. Using patient-derived in-
rod precursors are a likely source of regenerated cells. J. Neurobiol. 19, 431–463. duced pluripotent stem cells to identify Parkinson's disease-relevant phenotypes.
Reh, T.A., Tully, T., 1986. Regulation of tyrosine hydroxylase-containing amacrine cell Curr. Neurol. Neurosci. Rep. 18, 84.
number in larval frog retina. Dev. Biol. 114, 463–469. Skowronska-Krawczyk, D., Zhao, L., Zhu, J., Weinreb, R.N., Cao, G., Luo, J., Flagg, K.,
Rehfeld, F., Rohde, A.M., Nguyen, D.T., Wulczyn, F.G., 2015. Lin28 and let-7: ancient Patel, S., Wen, C., Krupa, M., Luo, H., Ouyang, H., Lin, D., Wang, W., Li, G., Xu, Y., Li,
milestones on the road from pluripotency to neurogenesis. Cell Tissue Res. 359, O., Chung, C., Yeh, E., Jafari, M., Ai, M., Zhong, Z., Shi, W., Zheng, L., Krawczyk, M.,
145–160. Chen, D., Shi, C., Zin, C., Zhu, J., Mellon, P.L., Gao, W., Abagyan, R., Zhang, L., Sun,
Reichenbach, A., Bringmann, A., 2013. New functions of Muller cells. Glia 61, 651–678. X., Zhong, S., Zhuo, Y., Rosenfeld, M.G., Liu, Y., Zhang, K., 2015. P16INK4a upre-
Reinhardt, P., Schmid, B., Burbulla, L.F., Schondorf, D.C., Wagner, L., Glatza, M., Hoing, gulation mediated by SIX6 defines retinal ganglion cell pathogenesis in glaucoma.
S., Hargus, G., Heck, S.A., Dhingra, A., Wu, G., Muller, S., Brockmann, K., Kluba, T., Mol. Cell 59, 931–940.
Maisel, M., Kruger, R., Berg, D., Tsytsyura, Y., Thiel, C.S., Psathaki, O.E., Klingauf, J., Sluch, V.M., Chamling, X., Liu, M.M., Berlinicke, C.A., Cheng, J., Mitchell, K.L., Welsbie,
Kuhlmann, T., Klewin, M., Muller, H., Gasser, T., Scholer, H.R., Sterneckert, J., 2013. D.S., Zack, D.J., 2017. Enhanced stem cell differentiation and immunopurification of
Genetic correction of a LRRK2 mutation in human iPSCs links parkinsonian neuro- genome engineered human retinal ganglion cells. Stem Cells Transl Med 6,
degeneration to ERK-dependent changes in gene expression. Cell Stem Cell 12, 1972–1986.
354–367. Sluch, V.M., Davis, C.H., Ranganathan, V., Kerr, J.M., Krick, K., Martin, R., Berlinicke,
Reubinoff, B.E., Itsykson, P., Turetsky, T., Pera, M.F., Reinhartz, E., Itzik, A., Ben-Hur, T., C.A., Marsh-Armstrong, N., Diamond, J.S., Mao, H.Q., Zack, D.J., 2015.
2001. Neural progenitors from human embryonic stem cells. Nat. Biotechnol. 19, Differentiation of human ESCs to retinal ganglion cells using a CRISPR engineered
1134–1140. reporter cell line. Sci. Rep. 5, 16595.
Reyes-Aguirre, L.I., Ferraro, S., Quintero, H., Sanchez-Serrano, S.L., Gomez-Montalvo, A., Smith, P.D., Sun, F., Park, K.K., Cai, B., Wang, C., Kuwako, K., Martinez-Carrasco, I.,
Lamas, M., 2013. Glutamate-induced epigenetic and morphological changes allow rat Connolly, L., He, Z., 2009. SOCS3 deletion promotes optic nerve regeneration in vivo.
Muller cell dedifferentiation but not further acquisition of a photoreceptor pheno- Neuron 64, 617–623.
type. Neuroscience 254, 347–360. Smukler, S.R., Runciman, S.B., Xu, S., van der Kooy, D., 2006. Embryonic stem cells as-
Rezaie, T., Child, A., Hitchings, R., Brice, G., Miller, L., Coca-Prados, M., Heon, E., Krupin, sume a primitive neural stem cell fate in the absence of extrinsic influences. J. Cell
T., Ritch, R., Kreutzer, D., Crick, R.P., Sarfarazi, M., 2002. Adult-onset primary open- Biol. 172, 79–90.
angle glaucoma caused by mutations in optineurin. Science 295, 1077–1079. Spalding, K.L., Rush, R.A., Harvey, A.R., 2004. Target-derived and locally derived neu-
Riesenberg, A.N., Le, T.T., Willardsen, M.I., Blackburn, D.C., Vetter, M.L., Brown, N.L., rotrophins support retinal ganglion cell survival in the neonatal rat retina. J.
2009a. Pax6 regulation of Math5 during mouse retinal neurogenesis. Genesis 47, Neurobiol. 60, 319–327.
175–187. Srinivas, M., Ng, L., Liu, H., Jia, L., Forrest, D., 2006. Activation of the blue opsin gene in
Riesenberg, A.N., Liu, Z., Kopan, R., Brown, N.L., 2009b. Rbpj cell autonomous regulation cone photoreceptor development by retinoid-related orphan receptor beta. Mol.
of retinal ganglion cell and cone photoreceptor fates in the mouse retina. J. Neurosci. Endocrinol. 20, 1728–1741.
29, 12865–12877. Stappert, L., Roese-Koerner, B., Brustle, O., 2015. The role of microRNAs in human neural
Riordan, J.R., Rommens, J.M., Kerem, B., Alon, N., Rozmahel, R., Grzelczak, Z., Zielenski, stem cells, neuronal differentiation and subtype specification. Cell Tissue Res. 359,
J., Lok, S., Plavsic, N., Chou, J.L., et al., 1989. Identification of the cystic fibrosis 47–64.
gene: cloning and characterization of complementary DNA. Science 245, 1066–1073. Stavridis, M.P., Lunn, J.S., Collins, B.J., Storey, K.G., 2007. A discrete period of FGF-
Rong, S.S., Lu, S.Y., Matsushita, K., Huang, C., Leung, C.K.S., Kawashima, R., Usui, S., induced Erk1/2 signalling is required for vertebrate neural specification.
Tam, P.O.S., Young, A.L., Tsujikawa, M., Zhang, M., Nishida, K., Wiggs, J.L., Tham, Development 134, 2889–2894.
C.C., Pang, C.P., Chen, L.J., 2019. Association of the SIX6 locus with primary open Stern, C.D., 2005. Neural induction: old problem, new findings, yet more questions.
angle glaucoma in southern Chinese and Japanese. Exp. Eye Res. 180, 129–136. Development 132, 2007–2021.
Rowan, S., Chen, C.M., Young, T.L., Fisher, D.E., Cepko, C.L., 2004. Transdifferentiation Strauss, O., 2005. The retinal pigment epithelium in visual function. Physiol. Rev. 85,
of the retina into pigmented cells in ocular retardation mice defines a new function of 845–881.
the homeodomain gene Chx10. Development 131, 5139–5152. Sun, Y., Smith, L.E.H., 2018. Retinal vasculature in development and diseases. Annu Rev
Rowe, R.G., Daley, G.Q., 2019. Induced pluripotent stem cells in disease modelling and Vis Sci 4, 101–122.
drug discovery. Nat. Rev. Genet. 20, 377–388. Swaroop, A., Kim, D., Forrest, D., 2010. Transcriptional regulation of photoreceptor de-
Roy, A., de Melo, J., Chaturvedi, D., Thein, T., Cabrera-Socorro, A., Houart, C., Meyer, G., velopment and homeostasis in the mammalian retina. Nat. Rev. Neurosci. 11,
Blackshaw, S., Tole, S., 2013. LHX2 is necessary for the maintenance of optic identity 563–576.
and for the progression of optic morphogenesis. J. Neurosci. 33, 6877–6884. Swaroop, A., Xu, J.Z., Pawar, H., Jackson, A., Skolnick, C., Agarwal, N., 1992. A con-
Sahlender, D.A., Roberts, R.C., Arden, S.D., Spudich, G., Taylor, M.J., Luzio, J.P., served retina-specific gene encodes a basic motif/leucine zipper domain. Proc. Natl.
Kendrick-Jones, J., Buss, F., 2005. Optineurin links myosin VI to the Golgi complex Acad. Sci. U. S. A. 89, 266–270.
and is involved in Golgi organization and exocytosis. J. Cell Biol. 169, 285–295. Takahashi, K., Yamanaka, S., 2006. Induction of pluripotent stem cells from mouse em-
Santos-Ferreira, T., Llonch, S., Borsch, O., Postel, K., Haas, J., Ader, M., 2016a. Retinal bryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676.
transplantation of photoreceptors results in donor-host cytoplasmic exchange. Nat. Takeda, M., Takamiya, A., Jiao, J.W., Cho, K.S., Trevino, S.G., Matsuda, T., Chen, D.F.,
Commun. 7, 13028. 2008. alpha-Aminoadipate induces progenitor cell properties of Muller glia in adult
Santos-Ferreira, T., Volkner, M., Borsch, O., Haas, J., Cimalla, P., Vasudevan, P., mice. Investig. Ophthalmol. Vis. Sci. 49, 1142–1150.
Carmeliet, P., Corbeil, D., Michalakis, S., Koch, E., Karl, M.O., Ader, M., 2016b. Stem Teotia, P., Chopra, D.A., Dravid, S.M., Van Hook, M.J., Qiu, F., Morrison, J., Rizzino, A.,
cell-derived photoreceptor transplants differentially integrate into mouse models of Ahmad, I., 2017a. Generation of functional human retinal ganglion cells with target
cone-rod dystrophy. Investig. Ophthalmol. Vis. Sci. 57, 3509–3520. specificity from pluripotent stem cells by chemically defined recapitulation of de-
Sasai, Y., 2013. Cytosystems dynamics in self-organization of tissue architecture. Nature velopmental mechanism. Stem Cells 35, 572–585.
493, 318–326. Teotia, P., Van Hook, M.J., Fischer, D., Ahmad, I., 2019. Human retinal ganglion cell axon
Sasai, Y., Eiraku, M., Suga, H., 2012. In vitro organogenesis in three dimensions: self- regeneration by recapitulating developmental mechanisms: effects of recruitment of
organising stem cells. Development 139, 4111–4121. the mTOR pathway. Development 146.
Schwarz, N., Lane, A., Jovanovic, K., Parfitt, D.A., Aguila, M., Thompson, C.L., da Cruz, L., Teotia, P., Van Hook, M.J., Wichman, C.S., Allingham, R.R., Hauser, M.A., Ahmad, I.,
Coffey, P.J., Chapple, J.P., Hardcastle, A.J., Cheetham, M.E., 2017. Arl3 and RP2 2017b. Modeling glaucoma: retinal ganglion cells generated from induced

29
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

pluripotent stem cells of patients with SIX6 risk allele show developmental ab- Watanabe, Y., Mizuseki, K., Sasai, Y., 2005. Directed differentiation of telencephalic
normalities. Stem Cells 35, 2239–2252. precursors from embryonic stem cells. Nat. Neurosci. 8, 288–296.
Tetreault, N., Champagne, M.P., Bernier, G., 2009. The LIM homeobox transcription Watanabe, T., Raff, M.C., 1992. Diffusible rod-promoting signals in the developing rat
factor Lhx2 is required to specify the retina field and synergistically cooperates with retina. Development 114, 899–906.
Pax6 for Six6 trans-activation. Dev. Biol. 327, 541–550. Westenskow, P., Piccolo, S., Fuhrmann, S., 2009. Beta-catenin controls differentiation of
Tomita, K., Ishibashi, M., Nakahara, K., Ang, S.L., Nakanishi, S., Guillemot, F., Kageyama, the retinal pigment epithelium in the mouse optic cup by regulating Mitf and Otx2
R., 1996. Mammalian hairy and Enhancer of split homolog 1 regulates differentiation expression. Development 136, 2505–2510.
of retinal neurons and is essential for eye morphogenesis. Neuron 16, 723–734. Wetts, R., Fraser, S.E., 1988. Multipotent precursors can give rise to all major cell types of
Tropepe, V., Hitoshi, S., Sirard, C., Mak, T.W., Rossant, J., van der Kooy, D., 2001. Direct the frog retina. Science 239, 1142–1145.
neural fate specification from embryonic stem cells: a primitive mammalian neural Wiley, L.A., Burnight, E.R., DeLuca, A.P., Anfinson, K.R., Cranston, C.M., Kaalberg, E.E.,
stem cell stage acquired through a default mechanism. Neuron 30, 65–78. Penticoff, J.A., Affatigato, L.M., Mullins, R.F., Stone, E.M., Tucker, B.A., 2016. cGMP
Tucker, B.A., Anfinson, K.R., Mullins, R.F., Stone, E.M., Young, M.J., 2013a. Use of a production of patient-specific iPSCs and photoreceptor precursor cells to treat retinal
synthetic xeno-free culture substrate for induced pluripotent stem cell induction and degenerative blindness. Sci. Rep. 6, 30742.
retinal differentiation. Stem Cells Transl Med 2, 16–24. Wilson, H.V., 1907. A new method by which sponges may Be artificially reared. Science
Tucker, B.A., Mullins, R.F., Streb, L.M., Anfinson, K., Eyestone, M.E., Kaalberg, E., Riker, 25, 912–915.
M.J., Drack, A.V., Braun, T.A., Stone, E.M., 2013b. Patient-specific iPSC-derived Wilson, S.W., Houart, C., 2004. Early steps in the development of the forebrain. Dev. Cell
photoreceptor precursor cells as a means to investigate retinitis pigmentosa. Elife 2, 6, 167–181.
e00824. Wohl, S.G., Hooper, M.J., Reh, T.A., 2019. MicroRNAs miR-25, let-7 and miR-124 reg-
Tucker, B.A., Park, I.H., Qi, S.D., Klassen, H.J., Jiang, C., Yao, J., Redenti, S., Daley, G.Q., ulate the neurogenic potential of Muller glia in mice. Development 146.
Young, M.J., 2011a. Transplantation of adult mouse iPS cell-derived photoreceptor Wohl, S.G., Reh, T.A., 2016. miR-124-9-9* potentiates Ascl1-induced reprogramming of
precursors restores retinal structure and function in degenerative mice. PLoS One 6, cultured Muller glia. Glia 64, 743–762.
e18992. Wong, R.C.B., Lim, S.Y., Hung, S.S.C., Jackson, S., Khan, S., Van Bergen, N.J., De Smit, E.,
Tucker, B.A., Scheetz, T.E., Mullins, R.F., DeLuca, A.P., Hoffmann, J.M., Johnston, R.M., Liang, H.H., Kearns, L.S., Clarke, L., Mackey, D.A., Hewitt, A.W., Trounce, I.A.,
Jacobson, S.G., Sheffield, V.C., Stone, E.M., 2011b. Exome sequencing and analysis of Pebay, A., 2017. Mitochondrial replacement in an iPSC model of Leber's hereditary
induced pluripotent stem cells identify the cilia-related gene male germ cell-asso- optic neuropathy. Aging (Albany NY) 9, 1341–1350.
ciated kinase (MAK) as a cause of retinitis pigmentosa. Proc. Natl. Acad. Sci. U. S. A. Wu, S., Chang, K.C., Nahmou, M., Goldberg, J.L., 2018. Induced pluripotent stem cells
108, E569–E576. promote retinal ganglion cell survival after transplant. Investig. Ophthalmol. Vis. Sci.
Tucker, B.A., Solivan-Timpe, F., Roos, B.R., Anfinson, K.R., Robin, A.L., Wiley, L.A., 59, 1571–1576.
Mullins, R.F., Fingert, J.H., 2014. Duplication of TBK1 stimulates autophagy in iPSC- Xia, X., Ahmad, I., 2016a. let-7 microRNA regulates neurogliogenesis in the mammalian
derived retinal cells from a patient with normal tension glaucoma. J. Stem Cell Res. retina through Hmga2. Dev. Biol. 410, 70–85.
Ther. 3, 161. Xia, X., Ahmad, I., 2016b. Unlocking the neurogenic potential of mammalian muller glia.
Turner, D.L., Cepko, C.L., 1987. A common progenitor for neurons and glia persists in rat Int J Stem Cells 9, 169–175.
retina late in development. Nature 328, 131–136. Xia, X., Teotia, P., Ahmad, I., 2018. Lin28a regulates neurogliogenesis in mammalian
Turner, D.L., Snyder, E.Y., Cepko, C.L., 1990. Lineage-independent determination of cell retina through the Igf signaling. Dev. Biol. 440, 113–128.
type in the embryonic mouse retina. Neuron 4, 833–845. Xia, X., Teotia, P., Ahmad, I., 2019. miR-29c regulates neurogliogenesis in the mamma-
Ueki, Y., Wilken, M.S., Cox, K.E., Chipman, L., Jorstad, N., Sternhagen, K., Simic, M., lian retina through REST. Dev. Biol. 450, 90–100.
Ullom, K., Nakafuku, M., Reh, T.A., 2015. Transgenic expression of the proneural Xiang, Y., Tanaka, Y., Cakir, B., Patterson, B., Kim, K.Y., Sun, P., Kang, Y.J., Zhong, M.,
transcription factor Ascl1 in Muller glia stimulates retinal regeneration in young Liu, X., Patra, P., Lee, S.H., Weissman, S.M., Park, I.H., 2019. hESC-derived thalamic
mice. Proc. Natl. Acad. Sci. U. S. A. 112, 13717–13722. organoids form reciprocal projections when fused with cortical organoids. Cell Stem
VanderWall, K.B., Vij, R., Ohlemacher, S.K., Sridhar, A., Fligor, C.M., Feder, E.M., Edler, Cell 24, 487–497 e487.
M.C., Baucum 2nd, A.J., Cummins, T.R., Meyer, J.S., 2019. Astrocytes regulate the Yang, X.J., 2004. Roles of cell-extrinsic growth factors in vertebrate eye pattern formation
development and maturation of retinal ganglion cells derived from human plur- and retinogenesis. Semin. Cell Dev. Biol. 15, 91–103.
ipotent stem cells. Stem Cell Reports 12, 201–212. Yang, Z., Ding, K., Pan, L., Deng, M., Gan, L., 2003. Math5 determines the competence
Vecino, E., Rodriguez, F.D., Ruzafa, N., Pereiro, X., Sharma, S.C., 2016. Glia-neuron in- state of retinal ganglion cell progenitors. Dev. Biol. 264, 240–254.
teractions in the mammalian retina. Prog. Retin. Eye Res. 51, 1–40. Yao, K., Qiu, S., Tian, L., Snider, W.D., Flannery, J.G., Schaffer, D.V., Chen, B., 2016. Wnt
Venugopalan, P., Wang, Y., Nguyen, T., Huang, A., Muller, K.J., Goldberg, J.L., 2016. regulates proliferation and neurogenic potential of muller glial cells via a lin28/let-7
Transplanted neurons integrate into adult retinas and respond to light. Nat. Commun. miRNA-dependent pathway in adult mammalian retinas. Cell Rep. 17, 165–178.
7, 10472. Yao, K., Qiu, S., Wang, Y.V., Park, S.J.H., Mohns, E.J., Mehta, B., Liu, X., Chang, B.,
Visvanathan, J., Lee, S., Lee, B., Lee, J.W., Lee, S.K., 2007. The microRNA miR-124 an- Zenisek, D., Crair, M.C., Demb, J.B., Chen, B., 2018. Restoration of vision after de
tagonizes the anti-neural REST/SCP1 pathway during embryonic CNS development. novo genesis of rod photoreceptors in mammalian retinas. Nature 560, 484–488.
Genes Dev. 21, 744–749. Yaron, O., Farhy, C., Marquardt, T., Applebury, M., Ashery-Padan, R., 2006. Notch1
Vitorino, M., Jusuf, P.R., Maurus, D., Kimura, Y., Higashijima, S., Harris, W.A., 2009. functions to suppress cone-photoreceptor fate specification in the developing mouse
Vsx2 in the zebrafish retina: restricted lineages through derepression. Neural Dev. retina. Development 133, 1367–1378.
4, 14. Yasumoto, K., Takeda, K., Saito, H., Watanabe, K., Takahashi, K., Shibahara, S., 2002.
Volkner, M., Zschatzsch, M., Rostovskaya, M., Overall, R.W., Busskamp, V., Anastassiadis, Microphthalmia-associated transcription factor interacts with LEF-1, a mediator of
K., Karl, M.O., 2016. Retinal organoids from pluripotent stem cells efficiently re- Wnt signaling. EMBO J. 21, 2703–2714.
capitulate retinogenesis. Stem Cell Reports 6, 525–538. Ying, Q.L., Stavridis, M., Griffiths, D., Li, M., Smith, A., 2003. Conversion of embryonic
Voronina, V.A., Kozhemyakina, E.A., O'Kernick, C.M., Kahn, N.D., Wenger, S.L., Linberg, stem cells into neuroectodermal precursors in adherent monoculture. Nat.
J.V., Schneider, A.S., Mathers, P.H., 2004. Mutations in the human RAX homeobox Biotechnol. 21, 183–186.
gene in a patient with anophthalmia and sclerocornea. Hum. Mol. Genet. 13, Yoshida, S., Mears, A.J., Friedman, J.S., Carter, T., He, S., Oh, E., Jing, Y., Farjo, R.,
315–322. Fleury, G., Barlow, C., Hero, A.O., Swaroop, A., 2004. Expression profiling of the
Walker, A., Su, H., Conti, M.A., Harb, N., Adelstein, R.S., Sato, N., 2010. Non-muscle developing and mature Nrl-/- mouse retina: identification of retinal disease candi-
myosin II regulates survival threshold of pluripotent stem cells. Nat. Commun. 1, 71. dates and transcriptional regulatory targets of Nrl. Hum. Mol. Genet. 13, 1487–1503.
Wall, D.S., Mears, A.J., McNeill, B., Mazerolle, C., Thurig, S., Wang, Y., Kageyama, R., Yoshida, T., Ozawa, Y., Suzuki, K., Yuki, K., Ohyama, M., Akamatsu, W., Matsuzaki, Y.,
Wallace, V.A., 2009. Progenitor cell proliferation in the retina is dependent on Notch- Shimmura, S., Mitani, K., Tsubota, K., Okano, H., 2014. The use of induced plur-
independent Sonic hedgehog/Hes1 activity. J. Cell Biol. 184, 101–112. ipotent stem cells to reveal pathogenic gene mutations and explore treatments for
Wan, J., Zheng, H., Chen, Z.L., Xiao, H.L., Shen, Z.J., Zhou, G.M., 2008. Preferential retinitis pigmentosa. Mol. Brain 7, 45.
regeneration of photoreceptor from Muller glia after retinal degeneration in adult rat. Yoshikawa, M., Nakanishi, H., Yamashiro, K., Miyake, M., Akagi, T., Gotoh, N., Ikeda,
Vis. Res. 48, 223–234. H.O., Suda, K., Yamada, H., Hasegawa, T., Iida, Y., Yamada, R., Matsuda, F.,
Wan, J., Zheng, H., Xiao, H.L., She, Z.J., Zhou, G.M., 2007. Sonic hedgehog promotes Yoshimura, N., Nagahama Study, G., 2017. Association of glaucoma-susceptible
stem-cell potential of Muller glia in the mammalian retina. Biochem. Biophys. Res. genes to regional circumpapillary retinal nerve fiber layer thickness and visual field
Commun. 363, 347–354. defects. Investig. Ophthalmol. Vis. Sci. 58, 2510–2519.
Wang, J., Deretic, D., 2014. Molecular complexes that direct rhodopsin transport to Young, R.W., 1985. Cell differentiation in the retina of the mouse. Anat. Rec. 212,
primary cilia. Prog. Retin. Eye Res. 38, 1–19. 199–205.
Wang, L., Hiler, D., Xu, B., AlDiri, I., Chen, X., Zhou, X., Griffiths, L., Valentine, M., Young, T.L., Cepko, C.L., 2004. A role for ligand-gated ion channels in rod photoreceptor
Shirinifard, A., Sablauer, A., Thiagarajan, S., Barabas, M.E., Zhang, J., Johnson, D., development. Neuron 41, 867–879.
Frase, S., Dyer, M.A., 2018a. Retinal cell type DNA methylation and histone mod- Yun, S., Saijoh, Y., Hirokawa, K.E., Kopinke, D., Murtaugh, L.C., Monuki, E.S., Levine,
ifications predict reprogramming efficiency and retinogenesis in 3D organoid cul- E.M., 2009. Lhx2 links the intrinsic and extrinsic factors that control optic cup for-
tures. Cell Rep. 22, 2601–2614. mation. Development 136, 3895–3906.
Wang, S.W., Kim, B.S., Ding, K., Wang, H., Sun, D., Johnson, R.L., Klein, W.H., Gan, L., Zaghloul, N.A., Yan, B., Moody, S.A., 2005. Step-wise specification of retinal stem cells
2001. Requirement for math5 in the development of retinal ganglion cells. Genes during normal embryogenesis. Biol. Cell 97, 321–337.
Dev. 15, 24–29. Zagozewski, J.L., Zhang, Q., Eisenstat, D.D., 2014. Genetic regulation of vertebrate eye
Wang, X.W., Li, Q., Liu, C.M., Hall, P.A., Jiang, J.J., Katchis, C.D., Kang, S., Dong, B.C., Li, development. Clin. Genet. 86, 453–460.
S., Zhou, F.Q., 2018b. Lin28 signaling supports mammalian PNS and CNS axon re- Zerti, D., Dorgau, B., Felemban, M., Ghareeb, A.E., Yu, M., Ding, Y., Krasnogor, N., Lako,
generation. Cell Rep. 24, 2540–2552 e2546. M., et al., 2019. Developing a simple method to enhance the generation of cone and
Watanabe, K., Kamiya, D., Nishiyama, A., Katayama, T., Nozaki, S., Kawasaki, H., rod photoreceptors in pluripotent stem cell-derived retinal organoids. Stem Cells.

30
I. Ahmad, et al. Progress in Retinal and Eye Research xxx (xxxx) xxxx

https://doi.org/10.1002/stem.3082. Zhou, S., Flamier, A., Abdouh, M., Tetreault, N., Barabino, A., Wadhwa, S., Bernier, G.,
Zhang, S.C., Wernig, M., Duncan, I.D., Brustle, O., Thomson, J.A., 2001. In vitro differ- 2015. Differentiation of human embryonic stem cells into cone photoreceptors
entiation of transplantable neural precursors from human embryonic stem cells. Nat. through simultaneous inhibition of BMP, TGFbeta and Wnt signaling. Development
Biotechnol. 19, 1129–1133. 142, 3294–3306.
Zhang, X.M., Yang, X.J., 2001. Regulation of retinal ganglion cell production by Sonic Zhu, J., Cifuentes, H., Reynolds, J., Lamba, D.A., 2017. Immunosuppression via loss of
hedgehog. Development 128, 943–957. IL2rgamma enhances long-term functional integration of hESC-derived photo-
Zhang, Y., Williams, P.R., Jacobi, A., Wang, C., Goel, A., Hirano, A.A., Brecha, N.C., receptors in the mouse retina. Cell Stem Cell 20, 374–384 e375.
Kerschensteiner, D., He, Z., 2019. Elevating growth factor responsiveness and axon Zhu, J., Reynolds, J., Garcia, T., Cifuentes, H., Chew, S., Zeng, X., Lamba, D.A., 2018.
regeneration by modulating presynaptic inputs. Neuron 103, 39–51 e35. Generation of transplantable retinal photoreceptors from a current good manu-
Zhao, X., Liu, J., Ahmad, I., 2002. Differentiation of embryonic stem cells into retinal facturing practice-manufactured human induced pluripotent stem cell line. Stem
neurons. Biochem. Biophys. Res. Commun. 297, 177–184. Cells Transl Med 7, 210–219.
Zhao, X.F., Wan, J., Powell, C., Ramachandran, R., Myers Jr., M.G., Goldman, D., 2014. Zuber, M.E., Gestri, G., Viczian, A.S., Barsacchi, G., Harris, W.A., 2003. Specification of
Leptin and IL-6 family cytokines synergize to stimulate Muller glia reprogramming the vertebrate eye by a network of eye field transcription factors. Development 130,
and retina regeneration. Cell Rep. 9, 272–284. 5155–5167.
Zhong, X., Gutierrez, C., Xue, T., Hampton, C., Vergara, M.N., Cao, L.H., Peters, A., Park, Zuber, M.E., Perron, M., Philpott, A., Bang, A., Harris, W.A., 1999. Giant eyes in Xenopus
T.S., Zambidis, E.T., Meyer, J.S., Gamm, D.M., Yau, K.W., Canto-Soler, M.V., 2014. laevis by overexpression of XOptx2. Cell 98, 341–352.
Generation of three-dimensional retinal tissue with functional photoreceptors from Zuchero, J.B., Barres, B.A., 2015. Glia in mammalian development and disease.
human iPSCs. Nat. Commun. 5, 4047. Development 142, 3805–3809.

31

You might also like