Molecular Vision 2016; 22:319-331 <http://www.molvis.org/molvis/v22/319>
Received 08 February 2016 | Accepted 07 April 2016 | Published 09 April 2016

Molecular studies of phenotype variation in canine RPGR-XLPRA1

Tatyana Appelbaum, Doreen Becker, Evelyn Santana, Gustavo D. Aguirre

Section of Ophthalmology, School of Veterinary Medicine, University of Pennsylvania, Philadelphia, PA

Correspondence to: Gustavo D. Aguirre, School of Veterinary Medicine, University of Pennsylvania, 3900 Delancey Street, Philadelphia, PA, 19104; Phone: (215) 898–9426; FAX: (215) 573–6050; email: gda@vet.upenn.edu

Abstract

Purpose: Canine X-linked progressive retinal atrophy 1 (XLPRA1) caused by a mutation in retinitis pigmentosa (RP) GTPase regulator (RPGR) exon ORF15 showed significant variability in disease onset in a colony of dogs that all inherited the same mutant X chromosome. Defective protein trafficking has been detected in XLPRA1 before any discernible degeneration of the photoreceptors. We hypothesized that the severity of the photoreceptor degeneration in affected dogs may be associated with defects in genes involved in ciliary trafficking. To this end, we examined six genes as potential disease modifiers. We also examined the expression levels of 24 genes involved in ciliary trafficking (seven), visual pathway (five), neuronal maintenance genes (six), and cellular stress response (six) to evaluate their possible involvement in early stages of the disease.

Methods: Samples from a pedigree derived from a single XLPRA1-affected male dog outcrossed to unrelated healthy mix-bred or purebred females were used for immunohistochemistry (IHC), western blot, mutational and haplotype analysis, and gene expression (GE). Cell-specific markers were used to examine retinal remodeling in the disease. Single nucleotide polymorphisms (SNPs) spanning the entire RPGR interacting and protein trafficking genes (RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B) were genotyped in the pedigree. Quantitative real-time PCR (qRT-PCR) was used to examine the expression of a total of 24 genes, including the six genes listed.

Results: Examination of cryosections from XLPRA1-affected animals of similar age (3–4 years) with different disease severity phenotype revealed mislocalization of opsins and upregulation of the Müller cell gliosis marker GFAP. Four to ten haplotypes per gene were identified in RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B for further assessment as potential genetic modifiers of XLPRA1. No correlation was found between the haplotypes and disease severity. During mutational analysis, several new variants, including a single intronic mutation in RAB8A and three mutations in exon 3 of DFNB31 were described (c.970G>A (V324I), c.978T>C (G326=), and c.985G>A (A329T)). Expression analysis of stress response genes in 16-week-old predisease XLPRA1 retinas revealed upregulation of GFAP but not HSPA5, DDIT3, HSPA4, HSP90B1, or HIF1A. Western blot analysis confirmed GFAP upregulation. In the same predisease group, no significant differences were found in the expression of 18 selected genes (RHO, OPN1LW, OPN1MW, RLBP1, RPGRORF15, RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, RAB11B, CRX, RCVRN, PVALB, CALB1, FGFR1, NTRK2, and NTRK3) involved in neuronal function.

Conclusions: Lack of association between haplotypes of RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B and the disease phenotype suggests that these genes are not genetic modifiers of XLPRA1. Upregulation of GFAP, an established indicator of the Müller cell gliosis, manifests as an important early feature of the disease.

Introduction

Retinitis pigmentosa (RP) is the clinical rubric for a large, heterogeneous group of inherited retinal disorders characterized by progressive photoreceptor degenerative disease leading to vision loss [1]. Of these disorders, X-linked RP (XLRP) comprises some of the most severe forms of RP and accounts for approximately 10% to 20% of all RP cases [2,3]. To date, six disease loci (RP2, RP3, RP6, RP23, RP24, and RP34) on the X chromosome have been mapped (RetNet), and RP3 accounts for approximately 75% of XLRP cases [4,5]. Retinitis pigmentosa GTPase regulator (RPGR) is the disease gene of the RP3 locus (RPGR-XLRP) [6,7].

The RPGR gene produces multiple alternatively spliced transcripts, all of which encode an N-terminal RCC1-like domain that is structurally similar to the RCC1 protein [6], a guanine nucleotide exchange factor for the Ran GTPase. One major constitutive isoform spans exons 1 through 19 and carries a C-terminal isoprenylation site. The other major variant contains exons 1 to 14 and terminates with a large, alternative ORF15 exon (RPGRorf15) [6-9]. The RPGRorf15 isoform is expressed predominantly in photoreceptor connecting cilia and basal bodies [10,11] and appears to be play a critical role in retinal function as multiple disease-causing mutations have been identified in humans, dogs, and mice [6,12,13]. Furthermore, gene augmentation therapy with RPGRorf15 preserves function and prevents degeneration in these diseases [14,15].

To date, the X-linked progressive retinal atrophy (XLPRA) dog is the only known naturally occurring large animal model of RPGR mutations [12]. The canine XLPRA phenotype has been linked to RPGR, and homology of canine XLPRA and human RP3 has been established [16,17]. Two disease-causing microdeletion mutations in exon ORF15 have been identified in canine RPGR [12]. One produces a premature stop (del1028–1032) resulting in a C-terminal truncation of 230 residues in XLPRA1. The second is a 2 nt deletion (del1084–1085) that causes a frameshift with the inclusion of 34 basic amino acids and truncation of the terminal 161 residues and is causal for XLPRA2. The phenotype associated with the frameshift mutation in XLPRA2 is severe and consistent, and manifests during retinal development; the phenotype resulting from XLPRA1 is expressed only after normal photoreceptor morphogenesis has been completed and has more gradual progression [12].

Extensive phenotypic diversity is observed in patients with RPGR mutations, between patients with different mutations, and between patients within families who have the same mutation [18-20]. Dogs with XLPRA1 also show remarkable phenotypic variability in clinical cases and, more surprisingly, within a closed research colony where all dogs are maintained in a constant environment, exposed to the same light intensities and cycles, and fed a uniform diet [21,22]. Such phenotypic variability cannot be simply explained by heterogeneity at the primary locus since all affected dogs inherited the same single mutant X chromosome, and the mutation is stably present through multiple generations [12,21]. Since genetic modifiers can alter the course of diseases, including RPGR-XLRP [18,23-25], we hypothesized that a modifier gene or genes contribute to the phenotypic variability observed in XLPRA1. Previously, we analyzed six genes (RPGRIP1, RANBP2, NPM1, PDE6D, NPHP5, and ABCA4) as genetic modifiers of XLPRA1 but excluded all [22]. As studies on RPGR-interacting proteins and those in the RPGRKO model suggest that RPGR functions in ciliary trafficking [9,26], here we focused on modifier genes that might also affect ciliary transport, thus potentiating the opsin mislocalization that contributes to XLPRA1 [27]. Candidate gene modifiers of XLPRA1 were chosen based on their contribution to ciliary formation, vesicular transport, and rod outer segment disc formation [28].

In the present study, we characterized retinas from XLPRA1 dogs with different degrees of disease for the expression of opsins and the Müller cell marker glial fibrillary acidic protein (GFAP), which responds to outer retinal stress and is one of the earliest genes upregulated in several non-allelic forms of inherited retinal degeneration [29,30]. We expanded the analysis of genetic modifiers of XLPRA1 to include genes that encode RPGR-interacting proteins (RPGRIP1L [25,31], DFNB31 [32], and RAB8A [33]) and proteins essential for cilia formation, ciliary trafficking, and cargo delivery (CEP290 [9], RAB11B [34], and CC2D2A [35]). We also examined the expression of 24 genes involved in ciliary trafficking, visual pathway, neuronal maintenance, and cellular stress response to evaluate their possible involvement in early stages of XPLRA1 disease.

Methods

Ethics statement

The research was conducted in full compliance and strict accordance with the Association for Research in Vision and Ophthalmology (ARVO) Resolution on the Use of Animals in Ophthalmic and Vision Research. The protocol was approved by the University of Pennsylvania Institutional Animal Care and Use Committee (IACUC).

Pedigree resources and determination of the phenotypic status

Details of the origin and composition of the colony have been previously published [21,22,36]. Briefly, the colony was established by outcrossing a single XLPRA1-affected male Siberian husky to unrelated healthy female beagles shown to be free from inherited retinal degeneration based on test breeding to known homozygous affected dogs with other autosomal recessive diseases [37]. The carrier progeny were subsequently mated with mixed breed or purebred dogs of varied genetic background to produce informative hemizygous or homozygous affected males or females, respectively, and heterozygous females. All dogs were maintained under specific and standard conditions where all animals have the same exposure to cyclic light (12 h:12 h light-dark cycle), receive the same diet, and have the same medical procedures and vaccinations. A subset of the colony, consisting of 43 dogs (affected with the RPGR mutation (n = 24), carriers (n = 7), and healthy (n = 12)), was selected for the studies (Appendix 1). The dogs in the study were the same used in a prior analysis of potential candidate gene disease modifiers [22]. Dogs were included based on the results of serial clinical assessment of retinal disease status using indirect ophthalmoscopy and electroretinography (ERG). Morphologic criteria were used to establish grades of disease severity taking into account the animal’s age and the degree and extent of disease [21]. Three grades were defined: Mild degeneration present only in periphery after 1.5 years of age or later, Moderate degeneration present only degeneration develops between 11 and 15 months of age, and Severe (photoreceptor degeneration stage 2 or more advanced, present centrally and peripherally) [21]. Of the 24 affected dogs (22 hemizygous males and two homozygous females), 14 were classified as Severe, nine as Moderate, and one as Mild. The skewed phenotype distribution results from selective breeding to produce affected dogs that could be diagnosed at an early age to be ascertained for the linkage mapping and gene and mutation identification studies [12,17]. As the group of dogs with Mild disease contained only one dog, in the disease-association studies only those with Moderate and Severe disease were used for analysis.

Sample collection, DNA/RNA extraction, and cDNA synthesis

Retinas were collected from enucleated eyes under sterile and RNase-free conditions, frozen by immersion in liquid nitrogen, and stored at -70 °C until used. For terminal procedures, the dogs were anesthetized by intravenous injection of pentobarbital sodium (65 mg/ml solution was prepared under sterile conditions by the Penn Medicine Investigational Drug Service), the eyes enucleated, and the dogs immediately euthanatized with intravenous Euthasol (Virbac, Ft. Worth, TX). Total RNA was isolated from canine tissues using a modified TRIzol and single chloroform extraction protocol. Briefly, 1 ml of TRIzol (Invitrogen, Carlsbad, CA) was added to the retina, and 0.2 ml of chloroform (Sigma-Aldrich, Allentown, PA) was added after tissue homogenization; the tissue-Trizol-chloroform mix then was centrifuged at 1,484 ×g at 4 °C for 4 min. The clear aqueous upper phase was transferred to a fresh microfuge tube and mixed to 70% ethanol (1:1, V/V). Next, the sample was loaded on an RNeasy Mini spin column, and the RNA was further purified using an RNeasy kit (Qiagen, Valencia, CA) following the manufacturer’s directions. First-strand cDNA was synthesized in 20 μl reactions using the High Capacity RNA-to-cDNA kit (Applied Biosystems, Foster City, CA) following the manufacturer’s recommendations. Genomic DNA was isolated from blood samples using the QIAamp DNA kit (Qiagen) following the manufacturer’s directions.

Genotype and haplotype analysis within candidate genes

Six genes were chosen for analysis (RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B). Polymorphic sites consisting of SNPs were identified in selected genes based on the Broad Institute SNP database, the Dog Genome SNP Database, and SNP database or based on our own gene sequencing analyses. Genotyping was performed using direct sequencing of PCR products. PHASE software (v. 2.1.1) was used for haplotype reconstruction. Fisher’s exact test (two-sided) was used for statistical analysis (p≤0.05). For this analysis, we posited that a candidate gene modifier would have a homozygous haplotype in dogs with Severe phenotype and heterozygous in those with a Moderate phenotype.

PCR amplification and sequencing

Primer sequences for the PCR experiments are shown in Appendix 2. PCR reactions were performed on 50 ng genomic DNA in a final volume of 25 μl containing 2x PCR master Mix (New England Biolabs, Ipswich, MA), 0.7 μM forward and reverse primers, and PCR-grade water. Cycling conditions were 2 min initial denaturation at 95 °C followed by 35 cycles of 95 °C for 20 s (denaturation), 52 °C for 30 s (annealing), and 68 °C for 30 s (elongation). For sequencing purposes, the PCR products were analyzed on a 2% agarose gel, extracted with the NucleoTrap Gel Extraction Kit (Clontech, Mountain View, CA), and directly sequenced.

Relative quantification (ddCt) assay

Real-time PCR was performed in a total volume of 20 μl in 96-well microwell plates on the Applied Biosystems 7500 Real-Time PCR System. All PCRs were performed using cDNA generated from 15 ng DNase-treated RNA from 16-week-old healthy retinas (n = 3) and XLPRA1-affected dogs (n = 6). The SYBR Green platform was used for gene expression analysis of 24 genes using a primer concentration of 0.2 μM. The primer sequences for real-time PCR are listed in Appendix 3. The TBP gene expression level was used to normalize the cDNA templates. Amplification data were analyzed with the 7500 Software version 2.0.1 (Applied Biosystems). Genes included for this analysis coded for proteins essential for cilia formation, ciliary trafficking, and cargo delivery, and genes involved in ciliary formation, visual pathway, neuronal maintenance, and cellular stress response; the genes are listed in Table 1.

Fluorescent immunohistochemistry

Retina from three XLPRA1-affected dogs and a healthy dog was used for immunohistochemistry (IHC). These dogs are not part of the genetic modifier pedigree. The procedures used for tissue collection, preparation, and sectioning have been previously described [38]. Cryosections were washed and treated with the primary antibodies in PBS solution (1X; 137 mM NaCl, 2.7 mM KCl, 8 mM Na2HPO4, 1.5 mM KH2PO4, pH 7.4), 3% normal horse serum (NHS, Vector Laboratories, Burlingame, CA), 1% bovine serum albumin (BSA, Sigma-Aldrich) and 0.3% Triton X-100 (Sigma-Aldrich) overnight followed by incubation with appropriate fluorescent secondary antibodies (Alexa Fluor Dyes, 1:300; Molecular Probes, Eugene, OR). The following primary antibodies were used in the study: mouse anti-rhodopsin at 1:1,000 (MAB5316, Chemicon, Temecula, CA), rabbit anti-L/M opsin at 1:1,000 (AB5405, Millipore, Billerica, MA), goat anti-hCAR (human cone arrestin) at 1:1,000 (a gift from Dr. W. Beltran, University of Pennsylvania), and rabbit anti-GFAP at 1:1,000 (Z0334, Dako, Carpinteria, CA). 4’, 6-diamidino-2-phenylindole (DAPI, Sigma-Aldrich) stain was used to label the cell nuclei. Slides were mounted with fluoromount G mounting media (catalog No. 01100-01; Southern Biotech, Birmingham, AL), coverslipped, and examined with epifluorescence microscopy with a Zeiss Axioplan microscope (Carl Zeiss Meditech, Oberkochen, Germany). Images were digitally captured (Spot 4.0 camera; Diagnostic Instruments, Inc., Sterling Heights, MI) and imported into a graphics program (Photoshop; Adobe, Mountain View, CA) for display.

Western blot analysis

Western blotting was performed with multiplex infrared detection using secondary antibodies conjugated to IRDye® fluorescent dyes. Briefly, total protein extracts in equal amounts (50 µg) as determined with the BCA Protein Assay Kit (Thermo Fisher Scientific, Rockford, IL) were resolved using 4–15% acrylamide gradient sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS–PAGE gel; Bio-Rad Laboratories, Hercules, CA) in Tris/glycine/SDS buffer (BioRad). Resolved proteins were immunoblotted into nitrocellulose membranes (LI-COR Biosciences, Lincoln, NE) using the Wet/Tank Blotting System (Bio-Rad) and blocked for 1 h at room temperature or overnight at 4 °C with 1× Odyssey Blocking Buffer (LI-COR). Immunoblots were probed overnight at 4 °C with unconjugated primary antibodies diluted in 1× Blocking Buffer. Membranes were then washed 3X with PBST (PBS containing 0.1% Tween 20 (Sigma-Aldrich) and incubated for 1 h at room temperature in goat anti-rabbit IRDye680RD- and goat anti-mouse IRDye800CW-conjugated immunoglobulin G (IgG) secondary antibodies (LI-COR) diluted 1/10,000 each in 1× Blocking Buffer containing 0.1% Tween-20. Immunoblots were washed three times in PBST, once in PBS, and scanned on the Li-COR Odyssey Fc Dual-Mode Imaging System with 700- and 800-nm channels. Normalization to ACTB and analyses were done using Image Studio Software provided by LI-COR. The primary antibodies used were rabbit anti-GFAP at 1:5,000 (Z0334, Dako) and anti-mouse ACTB at 1:10,000 (MAB1501, Millipore). Quantification of the proteins on western blot was performed with Li-COR Odyssey software.

Results

Characterization of XLPRA1-affected dogs with different severity phenotype

Examination of hematoxylin and eosin (H&E)-stained cryosections from XLPRA1-affected dogs of comparable ages (3–4 years) showed morphological alterations compatible with an ongoing disease process (Figure 1A–D). Disease was accompanied by decreased rod density and outer nuclear layer (ONL) thickness. Significant photoreceptor loss (ONL = 1 row) was observed in the youngest animal with Severe phenotype. To evaluate the localization of photoreceptor-specific proteins, double immunolabeling and DAPI nuclear staining were performed in the affected dogs. Immunolabeling with rod opsin showed mislocalization in Mild disease and loss of rods in Moderate and Severe disease. GFAP was upregulated in all disease phenotypes and was mostly present in the inner plexiform layer (IPL) and to a lesser extent in the outer plexiform layer (OPL; Figure 1E–H); increased GFAP expression was more prominent in Moderate and Severe disease. Double immunolabeling with hCAR and M/L opsin antibodies showed cone outer segment disorganization and mislocalization of M/L cone opsin (Figure 1I–L), and disorganization and loss of the outer segments (OS) were observed in the remaining rods and cones.

Genotype-phenotype analysis in XLPRA1 pedigree

As demonstrated previously [27] and in the present study (Figure 1), XLPRA1 is characterized by defective protein trafficking that leads to mislocalization of opsins to the inner segment, nuclear, and synaptic layers. Thus, the presence of mutations that perturb protein trafficking could have severe functional consequences that could result in photoreceptor cell death. To this end, six genes (RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B) were investigated as candidate disease modifiers. The proteins encoded by these genes are involved in common and distinct pathways required for effective protein trafficking in cilia and cargo delivery, and some (RPGRIP1L, DFNB31, and RAB8A) directly interact with RPGR.

Profiling sequence changes in RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B genes

Polymorphic markers were identified by sequence analysis of randomly selected regions within the genes of interest in the XLPRA1-affected study dogs. In addition, details of SNPs from several public SNP databases used for genotype and haplotype analysis are presented in Appendix 2. As two mutations in RPGRIP1L (A229T and R744Q) have been previously shown to contribute to phenotypic variability in retinal degeneration [18,25], we first analyzed the affected RPGR mutant dogs for the presence of A229T and R744Q. Human and canine RPGRIP1L is predicted to have 90.0% amino acid sequencing homology (compare NM_015272 and XP_013964947); therefore, A229 and R744 are in the same position in the human and canine RPGRIP1L sequence. None of the dogs analyzed harbored either mutation.

Sequence analysis of other genes identified four new variants: a single intronic change in RAB8A and three in exon 3 of DFNB31 in affected dogs. Sequence changes in DFNB31 exon 3 include two missense [(c.970G>A (p.V324I) and (c.985G>A (p.A329T))] and one synonymous (c.978T>C (p.G326=)) change (Figure 2). We performed alignment of the DFNB31 orthologous sequences to see whether the missense mutations affected conservative amino acid residues (Figure 3). Analysis of the corresponding protein region in multiple species has identified valine and isoleucine in position 324 of DFNB31; thus, it is unlikely that the V324I missense change observed in canine DFNB31 is consequential. However, the canine sequence has alanine at position 329 of DFNB31 while all other analyzed orthologous sequences contain proline. Although alanine and proline are hydrophobic neutral amino acids, the A329T change found in DFNB31 replaces alanine with threonine, a polar and hydrophilic residue that could potentially affect the structure and function of the protein. However, A329T was not included in the genotype and haplotype analysis as A329T was present only in the founder (H2) that was a carrier for the mutation.

The DFNB31 variants (V324I and G326=) were frequently present in the XLPRA1-affected dogs of the pedigree, but neither was associated with disease phenotype (Appendix 4). Despite a high prevalence of V324I (69.6% of XLPRA1-affected dogs), this variant appeared to have been artificially accumulated in the pedigree as the founder (H2), and most females used for breeding harbored V324I and passed it on to their offspring. Therefore, we concluded that the V324I sequence change is unlikely to have any functional consequences.

Genotype distribution

Genotype frequencies for the SNPs abundant in the XLPRA1-affected dogs are shown in Appendix 4 and were compared for the Moderate and Severe phenotypes. The genotype distributions of all analyzed SNPs were not statistically different between these two groups.

Haplotype distribution

The polymorphic SNPs in Appendix 2 provide the basis for gene-specific haplotype association studies. These results are presented in Table 2 and Table 3. The number of haplotypes per gene varied from four (CEP290) to ten (DFNB31; Table 2). Next, we compared the frequencies of haplotypes in the Moderate and Severe phenotypes of XLPRA1-affected dogs (Table 3). The haplotype frequencies for the genes RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B were not statistically different between the two groups analyzed in this study, thus excluding all six genes as candidate modifiers of the disease phenotype observed in the XLPRA1 pedigree.

Gene expression of selected genes in the study model

Variations in gene expression can also be associated with disease [39,40], and the lack of gene expression data for components of the RPGR molecular network hinders our understanding of its role in phenotypic expression of XLPRA1. In this study, we examined the expression of RPGRORF15, RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B in six XLPRA1-affected predisease (16 week old) retinas to determine whether expression changes in gene expression might contribute to disease progression. To gain more in-depth understanding of the disease-induced changes in gene expression, we also examined the expression of selected genes involved in the visual pathway (RHO, OPN1LW, OPN1MW, and RLBP1), neuronal cells (CRX, RCVRN, PVALB, CALB1, FGFR1, NTRK2, and NTRK3), and cellular stress response (GFAP, HSPA5, DDIT3, HSPA4, HSP90B1, and HIF1).

Out of the total 24 genes examined, only GFAP gene expression was upregulated (6.7-fold increase) in all XLPRA1 animals compared to the control group (p<0.01, Table 1). The GFAP level in XLPRA1 in the predisease samples was highly variable and increased in expression early in other retinal degeneration disorders in dogs [29,30]. The estimated variance in the XLPRA1 group was 9.96 compared to variance of only 0.07 in the control group. Upregulation of the GFAP protein in the 16-week-old XLPRA1 retinas was confirmed with western blot analysis (Figure 4).

Discussion

RPGR is an important component of the ciliary protein network, but details of the molecular function of RPGR in photoreceptors cilia are still incomplete. The protein complexes in which RPGR participates in the ciliary compartment play key roles in the function and maintenance of photoreceptor cells. For example, RPGR is associated through NPHP1, NPHP5, RPGRIP1, RPGRIP1L, RAB8A, CEP290, and SMC1/3 with the nephrocystin protein network and ciliary transport [9,41-44], through DFNB31 to the Usher protein network [32], and through NPM1 to the centrosomal protein network [45]. Defects of proteins in these complexes lead to photoreceptor dysfunction, cell death, and retinal degeneration.

The mislocalization of opsin is thought to contribute to the pathophysiology of photoreceptor degeneration. Changes in opsin localization have been reported for many diseases and injuries where photoreceptor cell death occurs, including human and animal models of retinal degeneration [13,27,46,47]. Opsin mislocalization is a feature of XLPRA1 and XLPRA2 and other canine retinal degeneration disorders (Figure 1 and [27,30,48]). Similar to several other RPGR disease models [13,49], opsin mislocalization was detected in XLPRA1 before any discernible photoreceptor degeneration. Since mutations in ciliary trafficking genes can contribute to ciliary transport defects, six genes (RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B) were selected based on their interaction with RPGR or involvement in ciliary transport and investigated as candidate genetic modifiers of XLPRA1.

The XLPRA1 pedigree is small, and its inbred affected population is not ideal for linkage disequilibrium (LD) analysis. Because of these limitations, we compared gene-specific genotypes and haplotypes frequencies in Severe versus Moderate affected dogs to determine whether there was a putative correlation between genotype or haplotype and disease phenotype. By carrying out this strategy, we found that moderately and severely affected dogs had no differences in genotypes and haplotype frequencies for RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B. Furthermore, these six genes were not variably expressed in predisease XLPRA1 retinas suggesting that the gene products do not serve as modifiers of disease phenotype.

During mutational analysis, three novel sequence variants were discovered in DFNB31 (V324I, G326, and A329T). All three were located in exon 3, which encodes part of the PDZ2 domain responsible for binding to the C-terminal of RPGRorf15 [32]. V324I was present in 69.6% of XLPRA1-affected dogs, suggesting at first that it could be associated with disease. However, on closer examination we found that this mutation artificially accumulated in the pedigree. Additional analysis of DFNB31 orthologous sequences excluded a putative harmful effect of V324I.

In addition to the seven genes involved in protein trafficking (RPGRORF15, RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B), we analyzed the expression of a subset of genes necessary for the visual pathway (RHO, OPN1LW, OPN1MW, and RLBP1), neuronal cell maintenance (CRX, RCVRN, PVALB, CALB1, FGFR1, NTRK2, and NTRK3), molecular chaperons (HSPA4, HSP90B1), unfolded protein response (HSPA5, DDIT3), cellular response to hypoxia (HIF1), and Müller cell gliosis (GFAP) in early XLPRA1. Although opsins are mislocalized in early XLPRA1 [27], we found that their gene expression was not impaired in 16-week-old predegenerate retinas. Aside from GFAP, other genes were not differentially expressed. Notably, GFAP expression was already upregulated at 16 weeks of age, when the XLPRA1 retinas are morphologically normal [21]. GFAP expression in Müller cells in the mammalian retina is normally low but is upregulated in a variety of degenerative conditions, including retinal trauma, diabetic retinopathy [50], choroidal neovascularization [51], retinal detachment [52], glaucoma [53], and age-related macular degeneration [54].

The results obtained in this study clearly demonstrate that even when mutated retinas appear morphologically normal, the Müller cells respond with a dramatic increase in GFAP expression in the early phase of XLPRA1. We also identified GFAP as highly variable in expression across predisease XLPRA1 retinas. Could GFAP gene expression variability lead to or be a marker of phenotypic diversity of XLPRA1? It is possible. Recent publications describe a complex interplay between Müller cells and retinal microglia in pathological conditions [55,56] that mediate adaptive responses within the retina and may be relevant to amplifying and coordinating an inflammatory response. The severity of the photoreceptor degeneration in XLPRA1-affected dogs may be associated with increased retinal inflammation. Therefore, an inflammatory response in the XLPRA1 retina should be further investigated.

GFAP typically increases with disease progression but decreases in some models in advanced stages of degeneration [27,57]. When assessing retinal remodeling events in XLPRA1-affected dogs with different severity phenotype, we observed markedly higher levels of GFAP in all disease stages in comparison to the healthy retina (Figure 1E–H). In addition, GFAP immunolabeling showed Müller glia reactivity increased with severity of the disease. This increase could be a sign of a sustained disturbance in retinal homeostasis in the disease.

In conclusion, this study reports a lack of association between the RAB8A, RPGRIP1L, CEP290, CC2D2A, DFNB31, and RAB11B genes and disease severity phenotype in canine XLPRA1. Selected genes responsible for neuronal cell maintenance and cellular stress response were not differentially expressed in predisease retinas except GFAP, which appears to be as an important early feature of the disease.

Current perspectives on the genetic analysis of canine XLPRA1

The search for potential genetic modifier(s) of the XLPRA1 phenotype will continue. As the next step, we will conduct a genome-wide scan in the XLPRA1-affected pedigree to search for loci associated with the disease phenotype.

Appendix 1. XLPRA1 pedigree.

Appendix 2. Characteristics of the SNP marker panel and associated PCR primers sequences

Appendix 3. Real-time PCR primers used in the study.

Appendix 4. Genotype frequencies in XLPRA1 affected dogs with different severity phenotypes.

Acknowledgments

The authors thank Drs. Jacob G. Appelbaum, Lesley King and Keiko Miyadera for helpful discussions and critical comments, Dr. William Beltran and Svetlana Savina for tissue slides and antibodies, Jacqueline C. Wivel for providing supplementary information related to XLPRA1 dogs. This research was supported by Foundation Fighting Blindness, R01EY06855, R01EY17549, P30EY-001583, Macula Vision Research Foundation, Hope for Vision and Van Sloun Fund. Dr. Tatyana Appelbaum (tatyanak@vet.upenn.edu) and Dr. Gustavo D. Aguirre (gda@vet.upenn.edu) are co-corresponding authors for this study.

References

  1. Sahel J, Bonnel S, Mrejen S, Paques M. Retinitis pigmentosa and other dystrophies. Dev Ophthalmol. 2010; 47:160-7. [PMID: 20703049]
  2. Shu X, Black GC, Rice JM, Hart-Holden N, Jones A, O’Grady A, Ramsden S, Wright AF. RPGR mutation analysis and disease: An update. Hum Mutat. 2007; 28:322-8. [PMID: 17195164]
  3. Breuer DK, Yashar BM, Filippova E, Hiriyanna S, Lyons RH, Mears AJ, Asaye B, Acar C, Vervoort R, Wright AF, Musarella MA, Wheeler P, MacDonald I, Iannaccone A, Birch D, Hoffman DR, Fishman GA, Heckenlively JR, Jacobson SG, Sieving PA, Swaroop A. A comprehensive mutation analysis of RP2 and RPGR in a north american cohort of families with X-linked retinitis pigmentosa. Am J Hum Genet. 2002; 70:1545-54. [PMID: 11992260]
  4. Ott J, Bhattacharya S, Chen JD, Denton MJ, Donald J, Dubay C, Farrar GJ, Fishman GA, Frey D, Gal A. Localizing multiple X chromosome-linked retinitis pigmentosa loci using multilocus homogeneity tests. Proc Natl Acad Sci USA. 1990; 87:701-4. [PMID: 2300556]
  5. Teague PW, Aldred MA, Jay M, Dempster M, Harrison C, Carothers AD, Hardwick LJ, Evans HJ, Strain L, Brock DJ. Heterogeneity analysis in 40 X-linked retinitis pigmentosa families. Am J Hum Genet. 1994; 55:105-11. [PMID: 8023838]
  6. Meindl A, Dry K, Herrmann K, Manson F, Ciccodicola A, Edgar A, Carvalho MR, Achatz H, Hellebrand H, Lennon A, Migliaccio C, Porter K, Zrenner E, Bird A, Jay M, Lorenz B, Wittwer B, D’Urso M, Meitinger T, Wright A. A gene (RPGR) with homology to the RCC1 guanine nucleotide exchange factor is mutated in X-linked retinitis pigmentosa (RP3). Nat Genet. 1996; 13:35-42. [PMID: 8673101]
  7. Vervoort R, Lennon A, Bird AC, Tulloch B, Axton R, Miano MG, Meindl A, Meitinger T, Ciccodicola A, Wright AF. Mutational hot spot within a new RPGR exon in X-linked retinitis pigmentosa. Nat Genet. 2000; 25:462-6. [PMID: 10932196]
  8. Wright AF, Shu X. Focus on molecules: RPGR. Exp Eye Res. 2007; 85:1-2. [PMID: 16769054]
  9. Rachel RA, Li T, Swaroop A. Photoreceptor sensory cilia and ciliopathies: focus on CEP290, RPGR and their interacting proteins. Cilia. 2012; 1:22 [PMID: 23351659]
  10. Hong DH, Pawlyk B, Sokolov M, Strissel KJ, Yang J, Tulloch B, Wright AF, Arshavsky VY, Li T. RPGR isoforms in photoreceptor connecting cilia and the transitional zone of motile cilia. Invest Ophthalmol Vis Sci. 2003; 44:2413-21. [PMID: 12766038]
  11. Hong DH, Pawlyk BS, Adamian M, Sandberg MA, Li T. A single, abbreviated RPGR-ORF15 variant reconstitutes RPGR function in vivo. Invest Ophthalmol Vis Sci. 2005; 46:435-41. [PMID: 15671266]
  12. Zhang Q, Acland GM, Wu WX, Johnson JL, Pearce-Kelling S, Tulloch B, Vervoort R, Wright AF, Aguirre GD. Different RPGR exon ORF15 mutations in canids provide insights into photoreceptor cell degeneration. Hum Mol Genet. 2002; 11:993-1003. [PMID: 11978759]
  13. Thompson DA, Khan NW, Othman MI, Chang B, Jia L, Grahek G, Wu Z, Hiriyanna S, Nellissery J, Li T, Khanna H, Colosi P, Swaroop A, Heckenlively JR. Rd9 is a naturally occurring mouse model of a common form of retinitis pigmentosa caused by mutations in RPGR-ORF15. PLoS One. 2012; 7:e35865 [PMID: 22563472]
  14. Wu Z, Hiriyanna S, Qian H, Mookherjee S, Campos MM, Gao C, Fariss R, Sieving PA, Li T, Colosi P, Swaroop A. A long-term efficacy study of gene replacement therapy for RPGR-associated retinal degeneration. Hum Mol Genet. 2015; 24:3956-70. [PMID: 25877300]
  15. Beltran WA, Cideciyan AV, Iwabe S, Swider M, Kosyk MS, McDaid K, Martynyuk I, Ying GS, Shaffer J, Deng WT, Boye SL, Lewin AS, Hauswirth WW, Jacobson SG, Aguirre GD. Successful arrest of photoreceptor and vision loss expands the therapeutic window of retinal gene therapy to later stages of disease. Proc Natl Acad Sci USA. 2015; 112:E5844-53. [PMID: 26460017]
  16. Zeiss CJ, Ray K, Acland GM, Aguirre GD. Mapping of X-linked progressive retinal atrophy (XLPRA), the canine homolog of retinitis pigmentosa 3 (RP3). Hum Mol Genet. 2000; 9:531-7. [PMID: 10699176]
  17. Zhang Q, Acland GM, Zangerl B, Johnson JL, Mao Z, Zeiss CJ, Ostrander EA, Aguirre GD. Fine mapping of canine XLPRA establishes homology of the human and canine RP3 intervals. Invest Ophthalmol Vis Sci. 2001; 42:2466-71. [PMID: 11581184]
  18. Fahim AT, Bowne SJ, Sullivan LS, Webb KD, Williams JT, Wheaton DK, Birch DG, Daiger SP. Allelic heterogeneity and genetic modifier loci contribute to clinical variation in males with X-linked retinitis pigmentosa due to RPGR mutations. PLoS One. 2011; 6:e23021 [PMID: 21857984]
  19. Sharon D, Sandberg MA, Rabe VW, Stillberger M, Dryja TP, Berson EL. RP2 and RPGR mutations and clinical correlations in patients with X-linked retinitis pigmentosa. Am J Hum Genet. 2003; 73:1131-46. [PMID: 14564670]
  20. Pelletier V, Jambou M, Delphin N, Zinovieva E, Stum M, Gigarel N, Dollfus H, Hamel C, Toutain A, Dufier JL, Roche O, Munnich A, Bonnefont JP, Kaplan J, Rozet JM. Comprehensive survey of mutations in RP2 and RPGR in patients affected with distinct retinal dystrophies: Genotype-phenotype correlations and impact on genetic counseling. Hum Mutat. 2007; 28:81-91. [PMID: 16969763]
  21. Zeiss CJ, Acland GM, Aguirre GD. Retinal pathology of canine X-linked progressive retinal atrophy, the locus homologue of RP3. Invest Ophthalmol Vis Sci. 1999; 40:3292-304. [PMID: 10586956]
  22. Guyon R, Pearce-Kelling SE, Zeiss CJ, Acland GM, Aguirre GD. Analysis of six candidate genes as potential modifiers of disease expression in canine XLPRA1, a model for human X-linked retinitis pigmentosa 3. Mol Vis. 2007; 13:1094-105. [PMID: 17653054]
  23. Cutting GR. Modifier genes in mendelian disorders: The example of cystic fibrosis. Ann N Y Acad Sci. 2010; 1214:57-69. [PMID: 21175684]
  24. Slavotinek A, Biesecker LG. Genetic modifiers in human development and malformation syndromes, including chaperone proteins. Hum Mol Genet. 2003; 12:R45-50. [PMID: 12668596]
  25. Khanna H, Davis EE, Murga-Zamalloa CA, Estrada-Cuzcano A, Lopez I, den Hollander AI, Zonneveld MN, Othman MI, Waseem N, Chakarova CF, Maubaret C, Diaz-Font A, MacDonald I, Muzny DM, Wheeler DA, Morgan M, Lewis LR, Logan CV, Tan PL, Beer MA, Inglehearn CF, Lewis RA, Jacobson SG, Bergmann C, Beales PL, Attie-Bitach T, Johnson CA, Otto EA, Bhattacharya SS, Hildebrandt F, Gibbs RA, Koenekoop RK, Swaroop A, Katsanis N. A common allele in RPGRIP1L is a modifier of retinal degeneration in ciliopathies. Nat Genet. 2009; 41:739-45. [PMID: 19430481]
  26. Rao KN, Li L, Anand M, Khanna H. Ablation of retinal ciliopathy protein RPGR results in altered photoreceptor ciliary composition. Sci Rep. 2015; 5:11137 [PMID: 26068394]
  27. Beltran WA, Acland GM, Aguirre GD. Age-dependent disease expression determines remodeling of the retinal mosaic in carriers of RPGR exon ORF15 mutations. Invest Ophthalmol Vis Sci. 2009; 50:3985-95. [PMID: 19255154]
  28. Hollingsworth TJ, Gross AK. Defective trafficking of rhodopsin and its role in retinal degenerations. Int Rev Cell Mol Biol. 2012; 293:1-44. [PMID: 22251557]
  29. Genini S, Zangerl B, Slavik J, Acland GM, Beltran WA, Aguirre GD. Transcriptional profile analysis of RPGRORF15 frameshift mutation identifies novel genes associated with retinal degeneration. Invest Ophthalmol Vis Sci. 2010; 51:6038-50. [PMID: 20574030]
  30. Genini S, Beltran WA, Aguirre GD. Up-regulation of tumor necrosis factor superfamily genes in early phases of photoreceptor degeneration. PLoS One. 2013; 8:e85408 [PMID: 24367709]
  31. Remans K, Burger M, Vetter IR, Wittinghofer A. C2 domains as protein-protein interaction modules in the ciliary transition zone. Cell Reports. 2014; 8:1-9. [PMID: 24981858]
  32. Wright RN, Hong DH, Perkins B. RpgrORF15 connects to the usher protein network through direct interactions with multiple whirlin isoforms. Invest Ophthalmol Vis Sci. 2012; 53:1519-29. [PMID: 22323458]
  33. Murga-Zamalloa CA, Atkins SJ, Peranen J, Swaroop A, Khanna H. Interaction of retinitis pigmentosa GTPase regulator (RPGR) with RAB8A GTPase: Implications for cilia dysfunction and photoreceptor degeneration. Hum Mol Genet. 2010; 19:3591-8. [PMID: 20631154]
  34. Deretic D, Wang J. Molecular assemblies that control rhodopsin transport to the cilia. Vision Res. 2012; 75:5-10. [PMID: 22892112]
  35. Bachmann-Gagescu R, Phelps IG, Stearns G, Link BA, Brockerhoff SE, Moens CB, Doherty D. The ciliopathy gene cc2d2a controls zebrafish photoreceptor outer segment development through a role in Rab8-dependent vesicle trafficking. Hum Mol Genet. 2011; 20:4041-55. [PMID: 21816947]
  36. Acland GM, Blanton SH, Hershfield B, Aguirre GD. XLPRA: A canine retinal degeneration inherited as an X-linked trait. Am J Med Genet. 1994; 52:27-33. [PMID: 7977457]
  37. Aguirre GD, Acland GM. Models, mutants and man: searching for unique phenotypes and genes in the dog model of inherited retinal degeneration. In: Ostrander EA, Giger U, LindbladToh K. (eds) The Dog and its Genome. 2006; 291–325.
  38. Beltran WA, Hammond P, Acland GM, Aguirre GD. A frameshift mutation in RPGR exon ORF15 causes photoreceptor degeneration and inner retina remodeling in a model of X-linked retinitis pigmentosa. Invest Ophthalmol Vis Sci. 2006; 47:1669-81. [PMID: 16565408]
  39. Chen Y, Zhu J, Lum PY, Yang X, Pinto S, MacNeil DJ, Zhang C, Lamb J, Edwards S, Sieberts SK, Leonardson A, Castellini LW, Wang S, Champy MF, Zhang B, Emilsson V, Doss S, Ghazalpour A, Horvath S, Drake TA, Lusis AJ, Schadt EE. Variations in DNA elucidate molecular networks that cause disease. Nature. 2008; 452:429-35. [PMID: 18344982]
  40. Papatheodorou I, Oellrich A, Smedley D. Linking gene expression to phenotypes via pathway information. J Biomed Semantics. 2015; 6:17 [PMID: 25901272]
  41. Khanna H, Hurd TW, Lillo C, Shu X, Parapuram SK, He S, Akimoto M, Wright AF, Margolis B, Williams DS, Swaroop A. RPGR-ORF15, which is mutated in retinitis pigmentosa, associates with SMC1, SMC3, and microtubule transport proteins. J Biol Chem. 2005; 280:33580-7. [PMID: 16043481]
  42. Murga-Zamalloa CA, Desai NJ, Hildebrandt F, Khanna H. Interaction of ciliary disease protein retinitis pigmentosa GTPase regulator with nephronophthisis-associated proteins in mammalian retinas. Mol Vis. 2010; 16:1373-81. [PMID: 20664800]
  43. Patnaik SR, Raghupathy RK, Zhang X, Mansfield D, Shu X. The Role of RPGR and Its Interacting Proteins in Ciliopathies. J Ophthalmol. 2015; 2015:414781 [PMID: 26124960]
  44. Roepman R, Wolfrum U. Protein networks and complexes in photoreceptor cilia. Subcell Biochem. 2007; 43:209-35. [PMID: 17953396]
  45. Shu X, Fry AM, Tulloch B, Manson FD, Crabb JW, Khanna H, Faragher AJ, Lennon A, He S, Trojan P, Giessl A, Wolfrum U, Vervoort R, Swaroop A, Wright AF. RPGR ORF15 isoform co-localizes with RPGRIP1 at centrioles and basal bodies and interacts with nucleophosmin. Hum Mol Genet. 2005; 14:1183-97. [PMID: 15772089]
  46. Price BA, Sandoval IM, Chan F, Simons DL, Wu SM, Wensel TG, Wilson JH. Mislocalization and degradation of human P23H-rhodopsin-GFP in a knockin mouse model of retinitis pigmentosa. Invest Ophthalmol Vis Sci. 2011; 52:9728-36. [PMID: 22110080]
  47. Li ZY, Kljavin IJ, Milam AH. Rod photoreceptor neurite sprouting in retinitis pigmentosa. J Neurosci. 1995; 15:5429-38. [PMID: 7643192]
  48. Berta AI, Boesze-Battaglia K, Genini S, Goldstein O, O’Brien PJ, Szel A, Acland GM, Beltran WA, Aguirre GD. Photoreceptor cell death, proliferation and formation of hybrid rod/S-cone photoreceptors in the degenerating STK38L mutant retina. PLoS One. 2011; 6:e24074 [PMID: 21980341]
  49. Hong DH, Pawlyk BS, Shang J, Sandberg MA, Berson EL, Li T. A retinitis pigmentosa GTPase regulator (RPGR)-deficient mouse model for X-linked retinitis pigmentosa (RP3). Proc Natl Acad Sci USA. 2000; 97:3649-54. [PMID: 10725384]
  50. Mizutani M, Gerhardinger C, Lorenzi M. Muller cell changes in human diabetic retinopathy. Diabetes. 1998; 47:445-9. [PMID: 9519752]
  51. Caicedo A, Espinosa-Heidmann DG, Pina Y, Hernandez EP, Cousins SW. Blood-derived macrophages infiltrate the retina and activate muller glial cells under experimental choroidal neovascularization. Exp Eye Res. 2005; 81:38-47. [PMID: 15978253]
  52. Sethi CS, Lewis GP, Fisher SK, Leitner WP, Mann DL, Luthert PJ, Charteris DG. Glial remodeling and neural plasticity in human retinal detachment with proliferative vitreoretinopathy. Invest Ophthalmol Vis Sci. 2005; 46:329-42. [PMID: 15623793]
  53. Shi Z, Rudzinski M, Meerovitch K, Lebrun-Julien F, Birman E, Di Polo A, Saragovi HU. Alpha2-macroglobulin is a mediator of retinal ganglion cell death in glaucoma. J Biol Chem. 2008; 283:29156-65. [PMID: 18701465]
  54. Sarthy V. Focus on molecules: Glial fibrillary acidic protein (GFAP). Exp Eye Res. 2007; 84:381-2. [PMID: 16563382]
  55. Wang M, Ma W, Zhao L, Fariss RN, Wong WT. Adaptive Müller cell responses to microglial activation mediate neuroprotection and coordinate inflammation in the retina. J Neuroinflammation. 2011; 8:173 [PMID: 22152278]
  56. Wang M, Wong WT. Microglia-muller cell interactions in the retina. Adv Exp Med Biol. 2014; 801:333-8. [PMID: 24664715]
  57. Hippert C, Graca AB, Barber AC, West EL, Smith AJ, Ali RR, Pearson RA. Muller glia activation in response to inherited retinal degeneration is highly varied and disease-specific. PLoS One. 2015; 10:e0120415 [PMID: 25793273]