Skip to main content
  • Research article
  • Open access
  • Published:

Expanded progenitor populations, vitreo-retinal abnormalities, and Müller glial reactivity in the zebrafish leprechaun/patched2 retina

Abstract

Background

The roles of the Hedgehog (Hh) pathway in controlling vertebrate retinal development have been studied extensively; however, species- and context-dependent findings have provided differing conclusions. Hh signaling has been shown to control both population size and cell cycle kinetics of proliferating retinal progenitors, and to modulate differentiation within the retina by regulating the timing of cell cycle exit. While cell cycle exit has in turn been shown to control cell fate decisions within the retina, a direct role for the Hh pathway in retinal cell fate decisions has yet to be established in vivo.

Results

To gain further insight into Hh pathway function in the retina, we have analyzed retinal development in leprechaun/patched2 mutant zebrafish. While lep/ptc2 mutants possessed more cells in their retinas, all cell types, except for Müller glia, were present at identical ratios as those observed in wild-type siblings. lep/ptc2 mutants possessed a localized upregulation of GFAP, a marker for 'reactive' glia, as well as morphological abnormalities at the vitreo-retinal interface, where Müller glial endfeet terminate. In addition, analysis of the over-proliferation phenotype at the ciliary marginal zone (CMZ) revealed that the number of proliferating progenitors, but not the rate of proliferation, was increased in lep/ptc2 mutants.

Conclusion

Our results indicate that Patched2-dependent Hh signaling does not likely play an integral role in neuronal cell fate decisions in the zebrafish retina. ptc2 deficiency in zebrafish results in defects at the vitreo-retinal interface and Müller glial reactivity. These phenotypes are similar to the ocular abnormalities observed in human patients suffering from Basal Cell Naevus Syndrome (BCNS), a disorder that has been linked to mutations in the human PTCH gene (the orthologue of the zebrafish ptc2), and point to the utility of the lep/ptc2 mutant line as a model for the study of BCNS-related ocular pathologies. Our findings regarding CMZ progenitor proliferation suggest that, in the zebrafish retina, Hh pathway activity may not affect cell cycle kinetics; rather, it likely regulates the size of the retinal progenitor pool in the CMZ.

Background

During retinal development, proliferation and differentiation must be tightly coordinated in order to produce a tissue of the proper size and containing the correct cell types [1]. The Hh pathway has been shown to play a critical role in controlling these two seemingly opposite processes [2]. Early in retinal development, the optic vesicle is composed of a population of proliferating neuroepithelial cells that will ultimately give rise to the mature retina [3]. Differentiation of the retinal neuroepithelium occurs in a succession of temporally overlapping waves [4]. In the zebrafish, the first cells to exit the cell cycle become retinal ganglion cells (RGCs), which differentiate in a Sonic Hh (Shh)-dependent wave [5]. A second Hh-dependent wave of differentiation in the inner nuclear layer (INL) occurs almost simultaneously with the first wave, and is responsible for the differentiation of the multiple cell types of the INL (horizontal, amacrine, and bipolar cells, and Müller glia) and the rod and cone photoreceptors of the outer nuclear layer (ONL) [6]. In addition, extra-retinal Hh signaling originating in the retinal pigment epithelium (RPE) has been suggested to direct photoreceptor differentiation [7]. While the role of the Hh pathway in cell cycle exit and differentiation of retinal progenitors is well described, comparatively less is known about its possible influence on cell fate decisions. In Xenopus, over-expression of p27Xic1, which promotes cell cycle exit, results in increased numbers of early-born cell types (RGCs), while the over-expression of cyclin E1, which inhibits cell cycle exit, biases cell fate towards late-born cell types (e.g. Müller glia) [8]. Similarly, Shh has been shown to promote early cell cycle exit in the Xenopus retina [9]; however, a direct role of the Hh pathway in dictating retinal cell fate has yet to be established in vivo.

While the cells of the central retina of the zebrafish exit the cell cycle by 60 hours post fertilization (hpf) [10], a population of retinal progenitors at the CMZ is maintained and continues to proliferate throughout the animal's lifetime [11, 12]. The spatial pattern of cells within the CMZ, with retinal stem cells at the most peripheral region followed by proliferative retinal progenitors and finally differentiating progenitors more centrally, is thought to recapitulate the temporal sequence of early retinal development [13]. Some of the factors that control early retinal development, such as notch, rx1, pax6a, and vsx2/chx10, are expressed in the zebrafish CMZ [11]. In Xenopus, expression of gli2, gli3, and smoothened, is found at the retinal margin, suggesting a role for the Hh pathway in the stem cell/progenitor population in the CMZ [14]. Shh over-expression studies in chick support a role for the Hh pathway as a mitogen in the CMZ [15], consistent with its described mitogenic effects on retinal progenitors in early zebrafish and mouse retinal development [16, 17].

Invaluable knowledge regarding Hh function in the developing retina has been gained from the analysis of Hh pathway mutants in zebrafish. For example, the zebrafish mutants syu (shh) and smu (smoothened) have helped elucidate the complex mechanisms of Hh-dependent neural differentiation and proliferation [7, 16]. However, retinal differentiation is severely defective or altogether absent in these mutants due to defective Hh signaling, making it difficult to arrive at definitive conclusions about the possible role of the Hh pathway in cell fate decisions. To address this issue, we analyzed retinal development in the zebrafish lep/ptc2 mutant, in which the Hh pathway is 'over-active'. lep/ptc2 mutants possess a non-sense mutation in the exon encoding the sixth trans-membrane domain of Patched2 [18], which is predicted to abolish its function [18, 19]. Loss of Patched function results in an 'over-active' Hh pathway due de-repression of Smoothened [20]. Normally, in the absence of Hh ligand, the Patched protein inhibits the activity of Smoothened. Binding of the Hh ligand to its Patched receptor relieves this inhibition, ultimately resulting in increased transcription of Hh target genes [21, 22]. A non-functional Patched would therefore result in loss of inhibition on, and constitutive activity of, Smoothened in ptc2 deficient cells. lep/ptc2 mutants possess increased proliferation in multiple tissues including the retinal CMZ [18]. Utilizing this mutant as a tool, we sought to gain further insight into Hh function with respect to retinal cell fate decisions and to determine how Hh activity influences progenitor proliferation at the CMZ. Our results revealed no significant change in neuronal composition within the retina, but we did note a reduction in differentiated Müller glia that was accompanied by localized Müller glial reactivity and abnormalities at the vitreo-retinal interface. In addition, the results of cell cycle analyses suggest that the over-proliferation at the CMZ likely results from an expansion of the progenitor cell population, and not from direct effects of the mutation on cell cycle kinetics within progenitor cells.

Results

lep/ptc2 mutants possess enlarged retinas and overgrown irises, while lens and RPE morphology are largely normal

lep/ptc2 mutants were isolated based on expanded PCNA staining in multiple tissues including the retinal CMZ [18]. While mutant embryos appear to be of normal size at 5 days post fertilization (dpf) (Figure 1A), their retinas are enlarged (Figure 1D, E) and they possess reduced pupil size ([18] and Figure 1B, C). To characterize ocular tissues in lep/ptc2 mutants, we performed histological analyses at 33 hpf, around the time of the onset of retinal differentiation, and at 5dpf, when the laminar organization of the embryonic retina is complete [10, 23]. lep/ptc2 retinas appear largely normal at 33dpf and do not possess increases in retinal cell number relative to phenotypically wild-type siblings (Figure 2A, B, E). At 5dpf, however, lep/ptc2 retinal cells are more tightly packed and mutant retinas contain a larger number of cells as compared to their phenotypically wild-type siblings (Figure 2C, D, E). Normal retinal lamination of differentiated neurons has been shown to be dependent on Shh signaling in vitro [24]. Lamination is largely normal in lep/ptc2 mutants and all major cell types are present (Figure 2C, D). Loss-of-function experiments in both frog and mouse have suggested a role for the Hh pathway in the development of the RPE [14, 25]. In lep/ptc2 mutants the RPE is of normal thickness and morphology. The Hh pathway has also been implicated in lens development [26, 27], and the reduced pupil size present in lep/ptc2 mutants was previously attributed to potential degeneration of the lens [28]. No gross morphological defects are evident in the lenses of lep/ptc2 mutants at either 33hpf or 5dpf, however. Additionally, histological analyses revealed an abnormality in the anterior segment of lep/ptc2 eyes where the iris was over-grown and extended further over the lens than in sibling retinas (Figure 2D). The enlarged iris and the overgrowth of the CMZ are the likely causes of the decreased pupil size, rather than defects in lens morphology.

Figure 1
figure 1

lep/ptc2 mutants possess enlarged retinas and reduced pupil size. Whole embryo lateral view shows that lep/ptc2 embryos are largely normal compared to their siblings (A). High magnification images of sibling (B) and lep/ptc2 (C) eyes reveal reduced pupil size in lep/ptc2 (arrows in B, C). Dorsal view images illustrate the overgrowth of retinal tissue as the retina is thicker and appears to engulf the lens (arrows in D, E).

Figure 2
figure 2

lep/ptc2 mutant retinas possess proportional increases in cell numbers in all three nuclear layers. Histology of sibling (A, C) and lep/ptc2 mutant (B, D) retinas at 33hpf (A, B) and 5dpf (C, D). E) Quantification of total number of retinal cells at 33hpf and 5dpf, and of cells in each nuclear layer at 5dpf (18.0%, 17.0%, and 17.8% increases in GCL, INL, and ONL, respectively. ***p < 0.0001). F) Graph of the proportion of each nuclear layer as a percentage of total retinal cells at 5dpf. Scale Bars = 20 um in A, B and 100 um in C, D.

Neuronal composition of the lep/ptc2 retina is normal, while Müller glia are under-represented

Hh signaling has been shown to be required for the differentiation of multiple cell types in the vertebrate retina [6, 7, 29] and is thought to promote cell fate decisions by modulating the timing of cell cycle exit [30]. Shh gain-of-function experiments in Xenopus link a faster cell cycle with early cell cycle exit [9], which in turn results in an over-production of early-born retinal cell types at the expense of late-born cell types [8]. We therefore reasoned that in lep/ptc2 retinas, where the Hh pathway is over-active, early-born cell types would be over-represented while late-born cell types would be reduced in number. Cell counts from histological sections of lep/ptc2 mutants revealed a statistically significant increase in cell number in all three retinal nuclear layers (GCL, INL, and ONL), as well as in total retinal cells, when compared to those from phenotypically wild-type siblings (Figure 2E). However, no significant change in the proportion of each layer was observed when the number of cells per layer was calculated as a percentage of total retinal cells (Figure 2F). This suggests that all three retinal cell layers are proportionally increased in lep/ptc2 retinas.

It is possible that in lep/ptc2 mutants, an overactive Hh pathway could alter neuronal composition within each layer. To test this hypothesis, immunohistochemistry using markers for each specific cell type in the INL and ONL was performed on retinal sections at 5dpf, when all cell types are normally differentiated and the embryonic retina is fully developed. Since the GCL layer is almost exclusively composed of retinal ganglion cells (RGCs), no further analysis was performed on this layer. In order to analyze cell types within the INL, immunohistochemistry was performed using antibodies that mark amacrine cells (5E11), bipolar cells (PKC) and Müller glia (Glutamine Synthetase, GS), while horizontal cells were detected according to their stereotypical morphology in histological sections. This analysis revealed that while the proportion of each neuronal cell type in the INL was not significantly changed in lep/ptc2 mutants, there was a statistically significant decrease in the proportion of differentiated Müller glia (Figure 3). For the ONL, immunohistochemistry was performed to detect rods (Zpr3) and each three cone opsins (blue, red/green, and UV) [31]. Due to the diffuse nature of the staining of some of the markers used, precise cell counts were not performed on these cell types. However, immunohistochemisty did show that these cell types are present in lep/ptc2 and qualitatively appear unaffected (Figure 4). In all, lep/ptc2 mutants possess more cells in their retinas and each cell type, except for Müller glia, was present at identical ratios to those observed in phenotypically wild-type siblings.

Figure 3
figure 3

All neuronal cell types in the lep/ptc2 INL are increased proportionately, while Müller glia are reduced. Immunohistochemistry of sibling (A, C, E) and lep/ptc2 (B, D, F) retinas for Amacrine cells (A, B), Bipolar cells (C, D), and Müller glia (E, F). G) Cell counts for each INL cell type. No significant change is cell number in all neuronal cell types when calculated as a percentage of total INL cells. Müller glia, however, were reduced in lep/ptc2 retinas by 29% relative to sibling retinas (**p = 0.0013).

Figure 4
figure 4

Photoreceptors appear unaffected in lep/ptc2 retinas. Immunohistochemical analysis of rods (A, B), blue cones (C, D), red/green cones (E, F), and UV cones (G, I) in sibling (A, C, E, G) and lep/ptc2 retinas (B, D, F, I). All photoreceptor types were present and appear normal.

lep/ptc2 mutants possess abnormalities at the vitreo-retinal boundary and localized Müller glial reactivity

Müller glial endfeet terminate at, and contribute to, the inner limiting membrane (ILM), which separates the retina from the vitreous [32]. Interestingly, our immunohistochemical analysis of differentiated Müller glia revealed disruptions in the integrity of the ILM (Figure 5A, B). Since humans who possess mutations in the PTCH gene (orthologue of the zebrafish ptc2 [33]) have been previously shown to possess abnormalities in the vitreo-retinal interface [34], we sought to further investigate this phenotype. High magnification examination of GS-stained retinas revealed disruptions in the integrity of the ILM in 40% of the lep/ptc2 retinas examined. While Müller glial endfeet form a tight and continuous boundary at the vitreo-retinal boundary in sibling retinas, the ILM in lep/ptc2 retinas was often discontinuous and disorganized, and endfeet did not appear to properly terminate in the ILM (Figure 5A, B). In Ptc1+/- mice, ocular abnormalities are accompanied by the upregulation of GFAP, a marker of reactive Müller glia, and ectopic proliferation of Müller glia in the central retina [34].

Figure 5
figure 5

lep/ptc2 mutants display Müller glial reactivity and morphological abnormalities in the ILM. Immunohistochemical analysis using glutamine synthetase (GS) antibody, which marks differentiated Müller glia and their endfeet at the ILM, highlights disruptions in the ILM. (A) In sibling retinas, the ILM is tight and continuous (inset). (B) In lep/ptc2 retinas the ILM is discontinuous and Müller glial endfeet do not terminate properly at the ILM (inset). Glial fibrillary acidic protein (GFAP) antibody staining reveals elevated immuno-reactivity in the inner retina, adjacent to the optic nerve of lep/ptc2 (D and arrows in D') mutant retinas as compared to siblings (C, C'). Approximately 40% of analyzed mutants displayed significant ILM disruptions and elevated GFAP levels.

Immunohistochemical analysis of GFAP revealed that its levels in Müller glia were indeed higher in lep/ptc2 mutant retinas as compared to their phenotypically wild-type siblings (Figure 5C, C', D, D'). Higher levels of GFAP were often localized to Müller glial endfeet in the central retina, adjacent to the optic nerve. However, BrdU incorporation assays in lep/ptc2 mutants at 5dpf did not reveal ectopic proliferation in the central retina (see Additional file 1), suggesting that the upregulation of GFAP is not coupled with Müller glial proliferation in this context. In addition, the localized nature of Müller glial reactivity in not due to abnormal cell death in the lep/ptc2 retina and/or optic nerve since no apoptotic nuclei were detected by TUNEL assays (see Additional file 1).

patched2 (ptc2) is expressed in the progenitor/stem cell populations of the CMZ

We next sought to better understand the nature of the over-proliferation at the CMZ by defining the retinal cell populations that express the ptc2 transcript. At 56hpf, prior to the time that the overgrowth phenotype becomes detectable, ptc2 expression was visible at the retinal margin, corresponding to the undifferentiated population of retinal progenitors of the CMZ (Figure 6A). Expression continued to be restricted to the CMZ at 72hpf (Figure 6B), consistent with previously reported expression of Hh target genes in the Xenopus CMZ [14]. Closer inspection of ptc2 expression in the CMZ at 72hpf revealed that low-level ptc2 expression was detected throughout most of the CMZ, consistent with previous reports [35]. Staining was more robust in two distinct CMZ cell populations; the first was immediately apposed to the RPE and the second was a patch of cells at the edge of the CMZ, bordering the differentiated GCL (arrows in Figure 6E). Hh ligands have been shown to be expressed in both the RPE [14] and the GCL [5, 24] during retinal development, and Patched genes are expressed in a Hh-dependent manner [33]. Thus, these CMZ regions may reflect areas of active Hh signaling. Finally, the ptc2 transcript is expressed at the iridio-corneal angle, where the iris and cornea converge [36] (arrows in Figure 6B).

Figure 6
figure 6

ptc2 is expressed at the zebrafish CMZ. In-situ hybridization for the ptc2 transcript shows expression at the retinal margin at 56hpf (arrows in A). At 72hpf, transcript is detected throughout the CMZ (B) and at the iridio-corneal angle (green arrows), with high magnification of CMZ region (E) revealing more robust staining adjacent to the RPE and in a region just apposed to the GCL (red arrows). Whole mount images from a dorsal view of sense (C) and anti-sense (D) probes in-situ hybridization show that staining is specific.

The number of proliferating progenitors, but not the rate of progenitor proliferation, is increased in the lep/ptc2 CMZ

Ptc1+/- mice possess increased numbers of progenitor cells at all stages of retinal development, suggesting a role for the Hh pathway in controlling the size of retinal progenitor populations in the ciliary margin, and these mice also possess a persistent progenitor population in the retinal margin as adults [37]. Studies in post-hatch chicks suggest that Shh regulates the proliferation of CMZ progenitors [15]. In addition, Shh over-expression studies in Xenopus have revealed a role for the Hh pathway in regulating the length of the cell cycle in the progenitor populations of the CMZ [9]. Given the differing roles for Hh pathway activity in these contexts, we wanted to determine whether the over-proliferation in lep/ptc2 retinas was due to an increase in the number of proliferating cells and/or an increase in the proliferation rate of the progenitor cell population in the CMZ. To address these possibilities, we first performed a 'pulse-chase' BrdU analysis on lep/ptc2 and sibling embryos starting at 64hpf, when a mild retinal phenotype is first detectable. Embryos were sorted according to phenotype and exposed to a short pulse (15 min) of BrdU. Half of the embryos were immediately fixed for sectioning, while the other half were transferred to non-BrdU containing fish water and fixed 8 hours later (72hpf). This enabled us to determine the rate of progenitor proliferation in both lep/ptc2 and sibling retinas during the 8 hour chase period. The results of these assays revealed that, on average, lep/ptc2 retinal sections contained 67% more BrdU-positive cells in the CMZs when compared to phenotypically wild-type sibling retinas at 64hpf (Figure 7A, B, E). At 72hpf, after the 8 hour 'chase', the average number of BrdU-positive cells in lep/ptc2 and phenotypically wild-type sibling retinal sections increased by almost identical ratios (2.29-fold changes in lep/ptc2 vs. 2.31 fold change in sibling sections, Figure 7C, D, E). From these data, we can therefore infer that, on average, each proliferating cell gave rise to 2.3 daughter cells during the 8 hour 'chase' period. To further analyze proliferation, we assayed the expression of phosphohistone H3 (PH3) and these analyses revealed that the proportion of CMZ progenitors occupying the M phase of the cell cycle increased by approximately 34% in lep/ptc2 retinas at 64hpf, when compared to their phenotypically wild-type siblings (Figure 7E, F, H). The degree of this increase was almost identical to that observed in progenitor cells occupying the S phase using BrdU as a marker (Figure 7A, B, H), suggesting that the length of the S and M phases was not changed relative to the other. These results support a model in which the expanded CMZ of lep/ptc2 mutants arises from an expansion of the progenitor cell population therein, and not from changes in the length of the cell cycle in these progenitor cells.

Figure 7
figure 7

The number of proliferating CMZ progenitors is increased in lep/ptc2 , while proliferation rate is unaffected. (A, B) BrdU 'pulse' in sibling (A) and lep/ptc2 mutants (B). BrdU+ cells are observed throughout the CMZ. (C, D) After an 8 hour 'chase', BrdU+ cells are observed at the CMZ, as well as within the retina of sibling (C) and lep/ptc2 mutants (D). E, F) PH3 labeling in sibling (E) and lep/ptc2 (F) at 64hpf. G) Quantification of BrdU+ cells in sibling and lep/ptc2 after 'pulse-chase' experiment. lep/ptc2 retinas contained 67% more BrdU+ cells than sibling retinas at 64hpf. After an 8 hour 'chase', the number of BrdU+ cells increased by almost identical ratios in siblings (2.31-fold increase) and lep/ptc2 mutants (2.29-fold increase, ***p < 0.0001). H) Quantification of BrdU- and PH3-positive cells as a percentage of total retinal cells in lep/ptc2 and wild-type siblings at 64hpf. The proportion of both BrdU+ and PH3+ cells increased by similar ratios (1.34-fold increase, **p < 0.001 and 1.36-fold increase, *p < 0.01, respectively).

Discussion

Retinal patterning is normal in lep/ptc2 mutants

Characterization of lep/ptc2 retinas with respect to cell type composition revealed that over-activity of the Hh pathway did not affect retinal patterning. Cell count analysis of all major retinal neuronal cell types revealed no change in the proportion of each cell type. This finding was surprising in light of the well established role of the Hh pathway in patterning and cell fate decisions throughout the CNS in a number of model systems [20, 38, 39], as well as studies in Xenopus which show that the over-expression of Shh results in early cell cycle exit [9], and that can, in turn, influence cell fate decisions [8]. While mouse Ptc1+/- (the orthologue of the zebrafish ptc2 [33]) mutants were found to possess no defects in retinal cell fate specification, Ptc1-/- mice could not be examined since they die in utero [37]. Our results indicate that Ptc2-dependent Hh signaling does not likely play an integral role in neuronal cell fate decisions in the zebrafish retina. These findings raise the possibility that the second zebrafish Patched protein (Patched1) might compensate for the lack of a functional Patched2. While patched1 (blowout) mutant retinas contain all major retinal cell types, the lack of a retinal phenotype might be due to the hypomorphic nature of the mutation [40, 41]. However, this seems unlikely since patched1 transcripts are not expressed at detectable levels in retinal tissue throughout development (data not shown and [35]). The reduced number of differentiated Müller glia, a late born cell type [10], could feasibly be due to a bias towards early born cell types in lep/ptc2 mutants; however, this potential bias does not affect neuronal composition of the retina at 5dpf. It is possible that a transient bias towards early neuronal cell types could occur early in retinal development, but any differences could masked by neurons added from the CMZ later in development. Conversely, any bias could also be present in neuronal populations that arise at the CMZ, resulting in an unnoticeable change in neuronal composition of the 5dpf lep/ptc2 retina. With this in mind, it will be interesting to examine the proportion of retinal cell types in juvenile lep/ptc2 mutants, as some homozygous fish reach two months of age [18]. Finally, the Hh pathway could play a role in neuronal cell fate decisions in the retina through Hh ligand interactions with other Hh receptors. The Ihog Hh receptors have recently been suggested to compete with Patched for Hh binding [42] and act at levels at, or upstream of, Patched [43]. Indeed, the vertebrate Ihog homologue Cdo is expressed in the developing mouse retina [44]. In the future, it would be interesting to test whether certain aspects of retinal development, such as cell fate decisions, might be Ihog-dependent.

lep/ptc2 mutants as a possible model for the study of BCNS-related ocular pathologies

In BCNS patients, abnormalities at the interface between the retina and the vitreous, known as epiretinal membranes (ERMs), are associated with disruptions of the ILM and are thought to result from the overproliferation and ectopic presence of multiple cell types, including glia [45]. BCNS ocular pathologies have been linked to mutations in the human PTCH gene [34, 46, 47], and have been shown to be associated with reactive Müller glia [34]. Müller glial reactivity is often marked by upregulation of GFAP, and lep/ptc2 mutants possessed regions of elevated GFAP levels. Elevated GFAP levels were consistently localized in the inner retina, where Müller glial endfeet terminate and contribute to the ILM. Staining with GS, which labels the Müller glial endfeet, revealed disruptions in ILM integrity in ~40% of mutant retinas assayed. In these cases, Müller glial endfeet did not terminate properly in the ILM, and the ILM was discontinuous. In some human patients, ERMs are found adjacent to retinal arteries [34]; interestingly, we consistently detected the presence of reactive glial endfeet in localized retinal regions adjacent to the optic nerve and embryonic retinal vasculature. Importantly, no increases in cell death were detected in the retina or optic nerve of lep/ptc2 mutants, or in Müller glia themselves (see Additional file 1), ruling out reactivity due to apoptosis. Finally, in other contexts, Müller glia reactivity is often coupled with ectopic proliferation of these cells during 'reactive gliosis' [48]. BrdU analysis did not detect ectopic proliferation in the central retinas of lep/ptc2 mutants (Figure 7 and Additional file 1), indicating that while a subset of these cells was reactive, they did not undergo a proliferative response. While reactive Müller glia have been shown to be associated with ERMs, it is unclear whether reactive Müller glia are the cause of these pathologies or, conversely, whether Müller glia become reactive in response to ERM formation. Further studies will be required to answer this question and to shed light on the cellular and molecular causes of ERM formation.

The retinal progenitor population of the CMZ is expanded in lep/ptc2 mutants

Exact roles for Hh pathway activity in proliferating retinal progenitor cells remain unclear. Over-expression studies in Xenopus suggest a direct role for the Hh pathway in regulating proliferation by influencing the length of the cell cycle in CMZ progenitors [9]. In Ptc1+/- mice, more retinal progenitors are allocated to the CMZ throughout development and into adulthood [37]. BrdU pulse-chase analysis of the progenitor population at the lep/ptc2 CMZ suggests that Patched-dependent Hh signaling controls the number of retinal progenitors in the zebrafish CMZ, and does not directly affect the length of the cell cycle in these cells. Importantly, cell counts at an earlier time-point, before the formation of the CMZ, revealed no increase in cell number in lep/ptc2 retinas as compared to their wild-type siblings (Figure 2E), suggesting that the increase in proliferating progenitors at the CMZ is not simply due to an earlier proliferative event. Shh has been shown to control stem cell maintenance in multiple organs, including the adult brain (reviewed in [49]). Indeed, the Hh pathway genes smo, gli2, and gli3 are expressed in the putative stem cell region of the Xenopus CMZ [14]. In light of our findings, it is possible that the increase in retinal progenitors in the lep/ptc2 CMZ is an indirect result of mis-regulation of the stem cell population rather than a direct effect on progenitor proliferation.

Conclusion

The possible role for the Hh pathway in retinal cell fate decisions has yet to be established in vivo. Our results indicate that Patched2-dependent Hh signaling is not likely to play an integral role in dictating neuronal cell fate decisions in the zebrafish retina. In addition, ocular phenotypes in lep/ptc2 mutants that are similar to those found in human BCNS patients point to the utility of the lep/ptc2 mutant line as a model for the study of BCNS ocular pathologies. With regards to progenitor proliferation, our data support a role for Patched2-dependent Hh signaling in the control of the size of progenitor populations at the retinal CMZ, and not in regulating their rate of proliferation, during zebrafish eye development.

Methods

Zebrafish maintenance and strains

Zebrafish (Danio rerio) were maintained at 28.5°C on a 14 h light/10 h dark cycle. Embryos were obtained from the natural spawning of heterozygous carriers setup in pairwise crosses. Embryos were collected and raised at 28.5°C [50] and were staged according to [51]. ptc2tj222outcrosses were obtained from the Zebrafish International Resource Center and were propagated by repeated outcrosses to AB fish. All animals were treated in accordance with provisions established at the University of Texas at Austin governing animal use and care.

Histology

Histology was performed as described in [52]. Briefly, mutant and wild-type sibling embryos were collected and fixed overnight at 4°C in a solution of 1% (w/v) paraformaldehyde (PFA), 2.5% glytaraldehyde and 3% sucrose in phosphate in a 2% OsO4 solution, washed 3 × 5 min in PBS at room temperature and further dehydrated 2 × 10 min in propylene oxide and infiltrated 1-2 h in a 50% propylene oxide/50% Epon/Araldite mixture (Polysciences, Inc.). Embryos were then incubated overnight at RT in 100% Epon/Araldite resin with caps open to allow for propylene oxide evaporation and resin infiltration, embedded and baked at 60°C for 2-3 days. Sections 1-1.25 μm were cut, mounted on glass slides and stained in a 1% methylene blue/1% borax solution. Sections were mounted in DPX (Electron Microscopy Sciences) and photographed on a Leica DMRB microscope mounted with a DFC320 digital camera. For 5dpf cell counts, the number nuclei in each retinal layer were counted in ten eye sections obtained from different embryos. Horizontal cells were identified by location and morphology. Averages were analyzed and compared using a Student's t-test.

Immunohistochemistry

Immunohistochemistry was performed as described in [53]. Briefly, embryos were collected and fixed overnight at 4°C in a 4% PFA solution in PBS. Embryos were then washed 3 × 10 min in PBS and incubated in a 25% sucrose/PBS solution for 1 h followed by 35% sucrose/PBS for 1 h. Embryos were then mounted in Tissue Freezing Medium (Triangle Biomedical Sciences, Inc.) and immediately transferred to -80°C freezer for storage. Frozen blocks were sectioned at 12 μm, mounted on gelatin coated slides, and let dry for 2 h. Slides were then re-hydrated in PBTD [0.1% Tween, 1% DMSO in 1× PBS] and then blocked using 5%NGS/PBTD for 2 h. Slides were incubated in primary antibody diluted in block in a humid chamber over-night at 4°C. When necessary, nuclear staining was obtained by applying Sytox-Green (Molecular Probes) diluted 1:10,000 in block immediately after removal of primary antibody. Slides were washed 3 × 10 min with PBTD and then applied with secondary antibody in block for 1 h. After 3 × 10 min washes in PBTD, slides were mounted with Vectashield (Vector Laboratories, Inc.) and coverslipped. Samples were stored at 4°C for up to one week and imaged on a Zeiss LSM5 Pascal laser scanning confocal microscope. The following primary antibodies were used: bipolar cells (PKC, 1:200, Santa Cruz Biotechnology), Müller glia (GFAP-zrf1, 1:100, ZIRC and GS, 1:500, Millipore), rods (zpr3, 1:100, ZIRC), green/red cones (zpr1, 1:100, ZIRC), PH3 (1:200, Upstate). Blue opsin (1:500) and UV opsin (1:1,000) antibodies were provided by David Hyde (Univerity of Notre Dame). Amacrine cell antibody (5E11, 1:100) was the gift of Jim Fadool. For cell counts, positively stained cells and nuclei were counted in four eye sections obtained from different embryos. For total cell counts at 33hpf, Sytox-Green staining was used. Averages were analyzed and compared using a Student's t-test.

BrdU incorporation assays

Embryos were dechorionated and incubated in fish water with 10 μM 5-Bromo-2-deoxyuridine (BrdU, Sigma) and 15% DMSO at 4°C for mentioned time periods and either immediately sacrificed or washed three times in fish water and grown at 28.5°C prior to sacrifice after [54]. Embryos were processed for immunohistochemistry as in [40]. Mouse anti-BrdU was used at a 1:50 dilution and Cy3 anti-mouse secondary at a 1:200 dilution. Nuclei were counterstained with Sytox:Green (1:10,000, Molecular Probes). For cell counts, Brdu-positive cells and nuclei were counted in four eye sections obtained from different embryos. Averages were analyzed and compared using a Student's t-test.

In-situ hybridization

Hybridizations were performed as described in [55] using digoxigenin labeled antisense RNA probes. ptc2 probe synthesis construct was provided by Brian Perkins (Texas A&M University).

References

  1. Neumann CJ: Hedgehogs as negative regulators of the cell cycle. Cell Cycle. 2005, 4: 1139-40.

    Article  CAS  PubMed  Google Scholar 

  2. Marti E, Bovolenta P: Sonic hedgehog in CNS development: one signal, multiple outputs. Trends Neurosci. 2002, 25: 89-96. 10.1016/S0166-2236(02)02062-3.

    Article  CAS  PubMed  Google Scholar 

  3. Li Z, Hu M, Ochocinska MJ, Joseph NM, Easter SS: Modulation of cell proliferation in the embryonic retina of zebrafish (Danio rerio). Dev Dyn. 2000, 219: 391-401. 10.1002/1097-0177(2000)9999:9999<::AID-DVDY1063>3.0.CO;2-G.

    Article  CAS  PubMed  Google Scholar 

  4. Amato MA, Boy S, Perron M: Hedgehog signaling in vertebrate eye development: a growing puzzle. Cell Mol Life Sci. 2004, 61: 899-910. 10.1007/s00018-003-3370-7.

    Article  CAS  PubMed  Google Scholar 

  5. Neumann CJ, Nuesslein-Volhard C: Patterning of the zebrafish retina by a wave of sonic hedgehog activity. Science. 2000, 289: 2137-9. 10.1126/science.289.5487.2137.

    Article  CAS  PubMed  Google Scholar 

  6. Shkumatava A, Fischer S, Muller F, Strahle U, Neumann CJ: Sonic hedgehog, secreted by amacrine cells, acts as a short-range signal to direct differentiation and lamination in the zebrafish retina. Development. 2004, 131: 3849-58. 10.1242/dev.01247.

    Article  CAS  PubMed  Google Scholar 

  7. Stenkamp DL, Frey RA: Extraretinal and retinal hedgehog signaling sequentially regulate retinal differentiation in zebrafish. Dev Biol. 2003, 258: 349-63. 10.1016/S0012-1606(03)00121-0.

    Article  CAS  PubMed  Google Scholar 

  8. Ohnuma S, Hopper S, Wang KC, Philpott A, Harris WA: Co-ordinating retinal histogenesis: early cell cycle exit enhances early cell fate determination in the Xenopus retina. Development. 2002, 129: 2435-46.

    CAS  PubMed  Google Scholar 

  9. Locker M, Agathocleous M, Amato MA, Parain K, Harris WA, Perron M: Hedgehog signaling and the retina: insights into the mechanisms controlling the proliferative properties of neural precursors. Genes Dev. 2006, 20: 3036-48. 10.1101/gad.391106.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  10. Schmitt EA, Dowling JE: Early retinal development in the zebrafish, Danio rerio: light and electron microscopic analyses. J Comp Neurol. 1999, 404: 515-36. 10.1002/(SICI)1096-9861(19990222)404:4<515::AID-CNE8>3.0.CO;2-A.

    Article  CAS  PubMed  Google Scholar 

  11. Raymond PA, Barthel LK, Bernardos RL, Perkowski JJ: Molecular characterization of retinal stem cells and their niches in adult zebrafish. BMC Dev Biol. 2006, 6: 36-10.1186/1471-213X-6-36.

    Article  PubMed Central  PubMed  Google Scholar 

  12. Wetts R, Serbedzija GN, Fraser SE: Cell lineage analysis reveals multipotent precursors in the ciliary margin of the frog retina. Dev Biol. 1989, 136: 254-63. 10.1016/0012-1606(89)90146-2.

    Article  CAS  PubMed  Google Scholar 

  13. Perron M, Kanekar S, Vetter ML, Harris WA: The genetic sequence of retinal development in the ciliary margin of the Xenopus eye. Dev Biol. 1998, 199: 185-200. 10.1006/dbio.1998.8939.

    Article  CAS  PubMed  Google Scholar 

  14. Perron M, Boy S, Amato MA, Viczian A, Koebernick K, Pieler T, Harris WA: A novel function for Hedgehog signalling in retinal pigment epithelium differentiation. Development. 2003, 130: 1565-77. 10.1242/dev.00391.

    Article  CAS  PubMed  Google Scholar 

  15. Moshiri A, McGuire CR, Reh TA: Sonic hedgehog regulates proliferation of the retinal ciliary marginal zone in posthatch chicks. Dev Dyn. 2005, 233: 66-75. 10.1002/dvdy.20299.

    Article  CAS  PubMed  Google Scholar 

  16. Stenkamp DL, Frey RA, Mallory DE, Shupe EE: Embryonic retinal gene expression in sonic-you mutant zebrafish. Dev Dyn. 2002, 225: 344-50. 10.1002/dvdy.10165.

    Article  CAS  PubMed  Google Scholar 

  17. Jensen AM, Wallace VA: Expression of Sonic hedgehog and its putative role as a precursor cell mitogen in the developing mouse retina. Development. 1997, 124: 363-71.

    CAS  PubMed  Google Scholar 

  18. Koudijs MJ, den Broeder MJ, Keijser A, Wienholds E, Houwing S, van Rooijen EM, Geisler R, van Eeden FJ: The zebrafish mutants dre, uki, and lep encode negative regulators of the hedgehog signaling pathway. PLoS Genet. 2005, 1: e19-10.1371/journal.pgen.0010019.

    Article  PubMed Central  PubMed  Google Scholar 

  19. Johnson RL, Milenkovic L, Scott MP: In vivo functions of the patched protein: requirement of the C terminus for target gene inactivation but not Hedgehog sequestration. Mol Cell. 2000, 6: 467-78. 10.1016/S1097-2765(00)00045-9.

    Article  CAS  PubMed  Google Scholar 

  20. Goodrich LV, Milenkovic L, Higgins KM, Scott MP: Altered neural cell fates and medulloblastoma in mouse patched mutants. Science. 1997, 277: 1109-13. 10.1126/science.277.5329.1109.

    Article  CAS  PubMed  Google Scholar 

  21. Murone M, Rosenthal A, de Sauvage FJ: Sonic hedgehog signaling by the patched-smoothened receptor complex. Curr Biol. 1999, 9: 76-84. 10.1016/S0960-9822(99)80018-9.

    Article  CAS  PubMed  Google Scholar 

  22. Taipale J, Cooper MK, Maiti T, Beachy PA: Patched acts catalytically to suppress the activity of Smoothened. Nature. 2002, 418: 892-7. 10.1038/nature00989.

    Article  CAS  PubMed  Google Scholar 

  23. Gross JM, Perkins BD, Amsterdam A, Egana A, Darland T, Matsui JI, Sciascia S, Hopkins N, Dowling JE: Identification of zebrafish insertional mutants with defects in visual system development and function. Genetics. 2005, 170: 245-61. 10.1534/genetics.104.039727.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  24. Wang YP, Dakubo G, Howley P, Campsall KD, Mazarolle CJ, Shiga SA, Lewis PM, McMahon AP, Wallace VA: Development of normal retinal organization depends on Sonic hedgehog signaling from ganglion cells. Nat Neurosci. 2002, 5: 831-2. 10.1038/nn918.

    Article  CAS  PubMed  Google Scholar 

  25. Dakubo GD, Mazerolle C, Furimsky M, Yu C, St-Jacques B, McMahon AP, Wallace VA: Indian hedgehog signaling from endothelial cells is required for sclera and retinal pigment epithelium development in the mouse eye. Dev Biol. 2008, 320: 242-55. 10.1016/j.ydbio.2008.05.528.

    Article  CAS  PubMed  Google Scholar 

  26. Karlstrom RO, Talbot WS, Schier AF: Comparative synteny cloning of zebrafish you-too: mutations in the Hedgehog target gli2 affect ventral forebrain patterning. Genes Dev. 1999, 13: 388-93. 10.1101/gad.13.4.388.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  27. Swindell EC, Zilinski CA, Hashimoto R, Shah R, Lane ME, Jamrich M: Regulation and function of foxe3 during early zebrafish development. Genesis. 2008, 46: 177-83. 10.1002/dvg.20380.

    Article  CAS  PubMed  Google Scholar 

  28. Heisenberg CP, Brand M, Jiang YJ, Warga RM, Beuchle D, van Eeden FJ, Furutani-Seiki M, Granato M, Haffter P, Hammerschmidt M, et al: Genes involved in forebrain development in the zebrafish, Danio rerio. Development. 1996, 123: 191-203.

    CAS  PubMed  Google Scholar 

  29. Zhang XM, Yang XJ: Regulation of retinal ganglion cell production by Sonic hedgehog. Development. 2001, 128: 943-57.

    CAS  PubMed  Google Scholar 

  30. Wallace VA: Proliferative and cell fate effects of Hedgehog signaling in the vertebrate retina. Brain Res. 2008, 1192: 61-75. 10.1016/j.brainres.2007.06.018.

    Article  CAS  PubMed  Google Scholar 

  31. Vihtelic TS, Doro CJ, Hyde DR: Cloning and characterization of six zebrafish photoreceptor opsin cDNAs and immunolocalization of their corresponding proteins. Vis Neurosci. 1999, 16: 571-85. 10.1017/S0952523899163168.

    Article  CAS  PubMed  Google Scholar 

  32. Gabriel R, Wilhelm M, Straznicky C: Morphology and distribution of Muller cells in the retina of the toad Bufo marinus. Cell Tissue Res. 1993, 272: 183-92. 10.1007/BF00323585.

    Article  CAS  PubMed  Google Scholar 

  33. Lewis KE, Concordet JP, Ingham PW: Characterisation of a second patched gene in the zebrafish Danio rerio and the differential response of patched genes to Hedgehog signalling. Dev Biol. 1999, 208: 14-29. 10.1006/dbio.1998.9169.

    Article  CAS  PubMed  Google Scholar 

  34. Black GC, Mazerolle CJ, Wang Y, Campsall KD, Petrin D, Leonard BC, Damji KF, Evans DG, McLeod D, Wallace VA: Abnormalities of the vitreoretinal interface caused by dysregulated Hedgehog signaling during retinal development. Hum Mol Genet. 2003, 12: 3269-76. 10.1093/hmg/ddg356.

    Article  CAS  PubMed  Google Scholar 

  35. Stenkamp DL, Frey RA, Prabhudesai SN, Raymond PA: Function for Hedgehog genes in zebrafish retinal development. Dev Biol. 2000, 220: 238-52. 10.1006/dbio.2000.9629.

    Article  CAS  PubMed  Google Scholar 

  36. Soules KA, Link BA: Morphogenesis of the anterior segment in the zebrafish eye. BMC Dev Biol. 2005, 5: 12-10.1186/1471-213X-5-12.

    Article  PubMed Central  PubMed  Google Scholar 

  37. Moshiri A, Reh TA: Persistent progenitors at the retinal margin of ptc+/- mice. J Neurosci. 2004, 24: 229-37. 10.1523/JNEUROSCI.2980-03.2004.

    Article  CAS  PubMed  Google Scholar 

  38. Ekker SC, Ungar AR, Greenstein P, von Kessler DP, Porter JA, Moon RT, Beachy PA: Patterning activities of vertebrate hedgehog proteins in the developing eye and brain. Curr Biol. 1995, 5: 944-55. 10.1016/S0960-9822(95)00185-0.

    Article  CAS  PubMed  Google Scholar 

  39. Ericson J, Rashbass P, Schedl A, Brenner-Morton S, Kawakami A, van Heyningen V, Jessell TM, Briscoe J: Pax6 controls progenitor cell identity and neuronal fate in response to graded Shh signaling. Cell. 1997, 90: 169-80. 10.1016/S0092-8674(00)80323-2.

    Article  CAS  PubMed  Google Scholar 

  40. Lee J, Willer JR, Willer GB, Smith K, Gregg RG, Gross JM: Zebrafish blowout provides genetic evidence for Patched1-mediated negative regulation of Hedgehog signaling within the proximal optic vesicle of the vertebrate eye. Dev Biol. 2008, 319: 10-22. 10.1016/j.ydbio.2008.03.035.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  41. Koudijs MJ, den Broeder MJ, Groot E, van Eeden FJ: Genetic analysis of the two zebrafish patched homologues identifies novel roles for the hedgehog signaling pathway. BMC Dev Biol. 2008, 8: 15-10.1186/1471-213X-8-15.

    Article  PubMed Central  PubMed  Google Scholar 

  42. McLellan JS, Zheng X, Hauk G, Ghirlando R, Beachy PA, Leahy DJ: The mode of Hedgehog binding to Ihog homologues is not conserved across different phyla. Nature. 2008, 455: 979-83. 10.1038/nature07358.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  43. McLellan JS, Yao S, Zheng X, Geisbrecht BV, Ghirlando R, Beachy PA, Leahy DJ: Structure of a heparin-dependent complex of Hedgehog and Ihog. Proc Natl Acad Sci USA. 2006, 103: 17208-13. 10.1073/pnas.0606738103.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  44. Mulieri PJ, Okada A, Sassoon DA, McConnell SK, Krauss RS: Developmental expression pattern of the cdo gene. Dev Dyn. 2000, 219: 40-9. 10.1002/1097-0177(2000)9999:9999<::AID-DVDY1032>3.0.CO;2-M.

    Article  CAS  PubMed  Google Scholar 

  45. Harada C, Mitamura Y, Harada T: The role of cytokines and trophic factors in epiretinal membranes: involvement of signal transduction in glial cells. Prog Retin Eye Res. 2006, 25: 149-64. 10.1016/j.preteyeres.2005.09.001.

    Article  CAS  PubMed  Google Scholar 

  46. Hahn H, Wicking C, Zaphiropoulous PG, Gailani MR, Shanley S, Chidambaram A, Vorechovsky I, Holmberg E, Unden AB, Gillies S, et al: Mutations of the human homolog of Drosophila patched in the nevoid basal cell carcinoma syndrome. Cell. 1996, 85: 841-51. 10.1016/S0092-8674(00)81268-4.

    Article  CAS  PubMed  Google Scholar 

  47. Johnson RL, Rothman AL, Xie J, Goodrich LV, Bare JW, Bonifas JM, Quinn AG, Myers RM, Cox DR, Epstein EH, et al: Human homolog of patched, a candidate gene for the basal cell nevus syndrome. Science. 1996, 272: 1668-71. 10.1126/science.272.5268.1668.

    Article  CAS  PubMed  Google Scholar 

  48. Dyer MA, Cepko CL: Control of Muller glial cell proliferation and activation following retinal injury. Nat Neurosci. 2000, 3: 873-80. 10.1038/78774.

    Article  CAS  PubMed  Google Scholar 

  49. Fuccillo M, Joyner AL, Fishell G: Morphogen to mitogen: the multiple roles of hedgehog signalling in vertebrate neural development. Nat Rev Neurosci. 2006, 7: 772-83. 10.1038/nrn1990.

    Article  CAS  PubMed  Google Scholar 

  50. Westerfield M: The zebrafish book. A guide for the laboratory use of zebrafish (Danio rerio). 2000, Eugene: University of Oregon Press, 4

    Google Scholar 

  51. Kimmel CB, Ballard WW, Kimmel SR, Ullmann B, Schilling TF: Stages of embryonic development of the zebrafish. Dev Dyn. 1995, 203: 253-310.

    Article  CAS  PubMed  Google Scholar 

  52. Nuckels RJ, Gross JM: Histological preparation of embryonic and adult zebrafish eyes. Cold Spring Harbor Protocols . 2007, 2-

    Google Scholar 

  53. Uribe RA, Gross JM: Immunohistochemistry on cryosections from embryonic and adult zebrafish eyes. Cold Spring Harbor Protocols. 2007, 2-

    Google Scholar 

  54. Masai I, Yamaguchi M, Tonou-Fujimori N, Komori A, Okamoto H: The hedgehog-PKA pathway regulates two distinct steps of the differentiation of retinal ganglion cells: the cell-cycle exit of retinoblasts and their neuronal maturation. Development. 2005, 132: 1539-53. 10.1242/dev.01714.

    Article  CAS  PubMed  Google Scholar 

  55. Jowett T, Lettice L: Whole-mount in situ hybridizations on zebrafish embryos using a mixture of digoxigenin- and fluorescein-labelled probes. Trends Genet. 1994, 10: 73-4. 10.1016/0168-9525(94)90220-8.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was funded by NIH RO1-EY18005 to J.M.G. We are grateful to David Hyde, Jim Fadool and Brian Perkins for providing reagents, to Ian Mackaye for technical assistance and to members of the Gross lab for helpful suggestions on this work, and comments on this manuscript. cDNA constructs and antisera were obtained from the Zebrafish International Resource Center, supported by NIH-NCRR grant P40 RR012546.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jeffrey M Gross.

Additional information

Authors' contributions

JB carried out all experimental procedures. JB and JMG conceived of the study, participated in its design and coordination, and drafted the manuscript. All authors read and approved the final manuscript.

Electronic supplementary material

12861_2009_456_MOESM1_ESM.TIFF

Additional file 1: lep/ptc2 mutants do not possess ectopic cell death or proliferation in the central retina and/or optic stalk. (A, B) BrdU incorporation and (C, D) TUNEL assays in 5dpf (A, C) phenotypically wild-type siblings and (B, D) lep/ptc2 mutants. (TIFF 1 MB)

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Bibliowicz, J., Gross, J.M. Expanded progenitor populations, vitreo-retinal abnormalities, and Müller glial reactivity in the zebrafish leprechaun/patched2 retina. BMC Dev Biol 9, 52 (2009). https://doi.org/10.1186/1471-213X-9-52

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-213X-9-52

Keywords