Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Feb 15.
Published in final edited form as: Ophthalmic Genet. 2009 Jun;30(2):57. doi: 10.1080/13816810802626399

RPE65: Role in the visual cycle, human retinal disease, and gene therapy

Xue Cai 1, Shannon M Conley 1, Muna I Naash 1,*
PMCID: PMC2821785  NIHMSID: NIHMS168011  PMID: 19373675

Abstract

RPE65 is an isomerohydrolase expressed in retinal pigment epithelium. It is critical for the regeneration of the visual pigment necessary for both rod and cone-mediated vision. Mutations in human RPE65 cause Leber’s congenital amaurosis and other forms of autosomal recessive retinitis pigmentosa which are associated with early-onset blindness. Several RPE65 animal models including two different mouse models and a naturally occurring canine model have been thoroughly characterized to determine the mechanisms that underlie RPE65 associated retinal dystrophies. More recently, substantial effort has gone into designing gene therapies for these diseases. Based on several encouraging reports from animal models, at least three clinical trials are currently underway for the treatment of LCA using modified AAV vectors carrying the RPE65 cDNA and have reported positive preliminary results.

Keywords: RPE65, retinal pigment epithelium, retina, LCA, gene therapy

Introduction

The visual cycle (or retinoid cycle) is the process by which 11-cis-retinal is regenerated from all-trans-retinal after a photoisomerization event. This cycle takes place in the photoreceptors and retinal pigment epithelium (RPE) and involves a series of enzymatic reactions1. RPE65 is a key isomerase in this process and is responsible for converting all-trans-retinyl ester to 11-cis-retinol. Without RPE65, 11-cis-retinal levels are substantially reduced and retinyl esters accumulate in RPE. RPE65 is an evolutionarily conserved 61-kDa membrane associated protein found in the smooth endoplasmic reticulum of the RPE cells25. Mutations in RPE65 are associated with recessive blinding diseases and models of RPE65 deficiency have been a common target for gene therapy studies over the past several years.

Gene cloning and structure

RPE65 was first described in 19936. Initial studies by Hamel et al. using the RPE9 monoclonal antibody suggested that the protein was 65 kDa in size and partitioned in the microsomal membrane fraction6. The same group also first isolated RPE65 cDNA in 19934 (although other groups were also studying the cDNA7) and expression studies in E. coli showed a protein size of 61 kDa. In 1994, the RPE65 gene was mapped to human chromosome 1 (mouse chromosome 3), and was refined to 1p31 by fluorescence in situ hybridization5,8. In 1995, Nicoletti et al. described the human gene structure5. The RPE65 gene contains 14 coding exons5,8 spanning 20 kb and encodes a protein of 533 amino acids5,810. In the coding region, the canine RPE65 gene shares 88–89% nucleotide sequence identity with the human and bovine sequence and 83% identity with the rat9. Sequence analysis combined with the observed tissue-specific expression, high abundance, evolutionary conservation, and developmental regulation, suggest that this protein is functionally important and its expression is tightly regulated5,10. Due to its functional importance in vision (discussed below), several groups have tried to study RPE65 in vitro, but transcripts are rapidly lost from RPE cells established in culture. It has been suggested that this downregulation may occur posttranscriptionally due to regions in the 3′UTR that may target the RNA for rapid degradation in the absence of other environmental cues4,5.

RPE65 function

RPE65 expression is developmentally regulated6. In rat, Rpe65 message is detectable as early as embryonic day (E) 18 and levels peak twice during retinal development. From postnatal day (P) 2 to P4, Rpe65 message stabilizes at levels about 10 fold higher than those seen at E18. The second peak (from P10-P12) is substantially larger, transcript levels are approximately 200 times higher than those seen at P2-410. The detection of RPE65 protein in the RPE (at P4-5) coincides with the earliest appearance of photoreceptor outer segment (OS) membranes11,12, and protein is continuously detectable throughout life10. These observations suggest that developmental expression of RPE65 may be coordinated with expression of some photoreceptor-specific proteins10.

RPE65 has two forms; a soluble form called sRPE65, and a palmitoylated, membrane-bound form known as mRPE65. Although it was proposed in 2004 that the palmitoylation state of RPE65 could serve as a molecular switch mechanism to help regulate retinoid recycling and transport13 (as a result of the differential binding affinity of the two forms of RPE65 for different retinoids), subsequent experiments have shown that this is not likely to be the case14.

It was originally thought that the role of RPE65 was restricted to the binding and mobilization of all-trans-retinyl esters for processing by the isomerohydrolase responsible for the critical conversion of all-trans- to 11-cis- retinoids15. However, more recently, three independent groups simultaneously published evidence that RPE65 is the isomerohydrolase1618. Moiseyev et al. showed that RPE65 protein expression correlates linearly with isomerohydrolase activity and its enzymatic activity can be reconstituted in vitro after transfection with RPE65 cDNA17. The retinal isomerohydrolase activity is iron dependent, with Fe2+ being the critical species, and purified RPE65 does contain iron19. Travis’ group conducted a cDNA expression screen using bovine RPE16. They isolated a single clone with high isomerohydrolase activity which, when sequenced, proved to be RPE6516. The third line of independent evidence came from the Redmond group which reconstituted the visual cycle in vitro. Only when RPE65 was expressed with other visual cycle proteins did retinoid isomerization occur18. The role of iron in this enzymatic process was confirmed by their observation that RPE65 constructs carrying mutations in the residues thought to be necessary for iron coordination (H180, H241, H313, H527) had little or no enzymatic activity.18

Although RPE65 was originally identified only in RPE, it is also expressed in both amphibian and mammalian cone photoreceptors but not rods20. The role of RPE65 in cones is unknown, but this observation (and others) provides support for the hypothesis that mammalian cones may utilize a different retinoid processing cycle in addition to the traditional one used by rods2023. If this is the case, retinoid regeneration in cones is likely still RPE65 dependent. When mice lacking RPE65 were crossed into the “cone-only” Nrl−/− or Rho−/− backgrounds, 11-cis-retinal was undetectable and retinyl esters and total retinoids were increased24,25. Consistent with the absence of pigment, retinal sensitivity dropped by a factor of one thousand and cone OSs degenerated24,25. Taken together, these data suggest that RPE65 is not only essential for rod function but it is also indispensable for cone function.

RPE65 associated diseases

Over 60 different mutations in the RPE65 gene have been associated with a heterogeneous group of inherited retinal dystrophies including Leber’s congenital amaurosis (LCA) and autosomal recessive retinitis pigmentosa (RP). These diseases are usually associated with blindness from birth or early childhood (reviewed in26 and summarized in http://www.retina-international.com/sci-news/rpe65mut.htm). Mutations in the RPE65 gene account for approximately 2% of cases of recessive RP and approximately 16% of cases of LCA in humans27,28.

In vivo assessments of human outer nuclear layer (ONL) thickness by high-resolution optical cross-sectional imaging indicate that in spite of early-onset blindness, the retinas of LCA patients retain normal retinal laminae with a definable photoreceptor layer into adulthood21,29 supporting the possibility that gene replacement therapy may be a viable treatment option. Furthermore, many patients have normal optic nerve diameter, anatomic structure including subcortical structure (in the lateral geniculate nucleus), and both gray and white matter; however, cortical responses are markedly diminished as shown by functional magnetic resonance imaging29.

Recent findings showed that many point mutations in RPE65, including those associated with LCA, have different effects on protein function. For example, in vivo and in vitro experiments have shown that Y144D, P363T, and Y368H RPE65 mutant proteins are significantly less stable than wild-type RPE6530,31. In addition to effects on protein levels, different point mutations cause different deficiencies in RPE65 function, and study of specific mutations has helped explain the heterogeneity observed in disease phenotypes. While Y144D and P363T RPE65 mutant proteins have no detectable isomerohydrolase activity (acting as virtual null alleles)31, other mutations cause a milder phenotype. For example, R91W32, P25L33, and L22P34 (and others) cause a slightly milder form of LCA with later onset deficits in vision than other mutations. These mutant proteins have been shown to retain more functional capacity, measured by 11-cis retinal levels and capacity for rhodopsin regeneration in the case of R91W32 and by in vitro isomerase activity in the case of P25L and L22P33, than mutations resulting in a null allele32.

RPE65 animal models

Several animal models with RPE65 mutations have been characterized. Among them the naturally occurring canine35 and murine (Rpe65rd12)36 models and the genetically engineered Rpe65−/− knockout have been widely used for pathological, biochemical, genetic, structural, functional, and therapeutic studies. Rpe65−/− knockout mice develop a slow retinal degeneration with relatively normal retinal anatomy at 7 weeks with normal OS structure and ONL thickness3,37, but at 15 weeks the OS length starts to decrease and inclusions start to appear in the RPE3. By 12 months, 30% of photoreceptor nuclei are lost, with even more severe photoreceptor loss at 18–24 months37. Rpe65−/− mice lack 11-cis retinal and 11-cis-retinyl esters, and accumulate excessive levels of all-trans retinyl esters in the RPE3 providing further support for the idea that the RPE65 protein is essential for the isomerization of all-trans retinyl esters3,38. Based on microarray analysis, mice lacking RPE65 have altered expression of many genes including some involved in phototransduction, apoptosis, cytoskeletal organization, and extracellular matrix regulation39. The knockout mice have nearly undetectable dark-adapted ERG responses3, and undetectable levels of functional rhodopsin even though the opsin apoprotein is structurally intact3,38. There was some confusion about the status of cone function in these mice3,38, but critical experiments by Seeliger et al. shed some light on this issue23. They crossed Rpe65−/− mice into either the Cnga3−/− background which lacks cone function or the Rho−/− which lacks rod function23, and demonstrated that the small ERG signal detected in Rpe65−/− mice comes from the remaining functional rods and not cones. Furthermore, they suggest that the dramatic decrease in the presence of functional chromophore in Rpe65−/− rods enables them to respond under photopic conditions which would normally saturate rods23. This hypothesis is supported by the observation that the Rpe65−/− has significantly decreased levels of cone-specific phototransduction genes but no changes in rod phototransduction genes, suggesting an early loss of cones3942. This conclusion was confirmed by the structural observation that cone degeneration started as early as 2 weeks of age with massive cone loss at 4 weeks40.

The naturally occurring mutant model, Rpe65rd12, is caused by a nonsense mutation in exon 3 in which a C to T transition creates a premature stop codon, R44X, resulting in loss-of-function due to truncation of the protein and mRNA degradation36. The phenotype is similar to the Rpe65−/− knockout; rod ERG response was profoundly diminished and small lipid-like droplets were deposited in RPE cells at 3 weeks. With time, droplet accumulation persists and is accompanied by slow retinal degeneration; at 6 weeks, the retina appears mostly normal with voids occasionally appearing in the OSs, but by 7 months, the OSs are obviously shorter and there is an approximately 30% reduction in the number of photoreceptor nuclei compared to wild-type36,43. No RPE65 expression, 11-cis retinal, or rhodopsin is detected in the retinas and retinyl esters accumulate in the RPE36,43. The only notable difference between the Rpe65−/− and the Rpe65rd12 models is that under ophthalmoscopic examination the latter model exhibits small white dots spread evenly throughout the retina at 5 to 9 months36.

The final commonly used RPE65 animal model is the Swedish Briard dog. This dog has very poor vision with severely depressed dark- and light-adapted ERG responses and serves as a naturally occurring large-animal model of LCA. Molecular analysis revealed a 4-nucleotide (AAGA) deletion in the RPE65 gene resulting in a frameshift and a premature stop codon which truncates the protein and causes the observed hereditary retinal dystrophy11,44. This deletion in the canine RPE65 sequence (487–490) corresponds to nucleotides 340–343 of human exon 59. Affected dogs have a normal fundus appearance until 3 years of age, however at 5 weeks of age, ERGs are abnormal with barely detectable scotopic responses and very low photopic amplitudes35. The disease is slowly progressive, with advanced age large lipid-like inclusions appear in the RPE, the OSs are disorganized, and rods and cones are lost starting from the peripheral retina and proceeding to the center45.

Gene therapy

Canine and mouse models of LCA caused by defects in the RPE65 gene are very attractive for gene therapy for a variety of reasons. RPE65-associated LCA is monogenic, and is caused by a recessive, loss-of-function mutation giving gene replacement therapy good prospects for attaining measurable improvements in visual function. Furthermore, the Briard dog model has a phenotype strikingly similar to that in human LCA patients46.

Rpe65−/− animal models (dogs and mice) have been treated with gene replacement therapy with positive outcomes47, although one prerequisite for success with these models appears to be the presence of an intact photoreceptor layer21. RPE is readily transduced by a variety of viruses48. Most recently, lentivirus has been used to transfer the mouse cDNA to the five day old Rpe65−/− mouse leading to long-lasting transgene expression in the RPE cells and maintenance of normal ERG function and cone number until at least 4 months49,50.. Untreated mice demonstrated the typical deficits in ERG function and degeneration of almost all cones by that time49,50.

Currently, recombinant adeno-associated virus (rAAV), a nonpathogenic parvovirus which contains no viral coding regions and has low cytotoxicity is the most popular vector used for therapeutic gene delivery. It elicits minimal immune response and induces efficient, long-term transgene expression. rAAV is also the only viral vector to efficiently transduce both RPE and photoreceptors48.

Rpe65−/− and Rpe65rd12 mice have been treated with rAAV vectors containing the RPE65 cDNA at different times including embryonic day 1451, postnatal day 1442,43, postnatal day 1852, 1–2.5 months51,53, and 6 & 13 months52. Treatment at all ages resulted in efficient RPE transduction, with transferred RPE65 protein expression detectable up to 7 months post-injection by immunohistochemistry43,51. Rhodopsin and retinyl-ester levels in treated mice were nearly normal43,51, 11-cis-retinal levels were significantly increased42, and retinal morphology was close to normal43,53. Treated animals exhibited pronounced functional rescue as well; ERG responses were significantly improved (approaching normal) at 4–7 months after treatment43,51,53,54 with consequent improvements in visual acuity53 or visual guided behavior43. It was reported that treatment age affected treatment success; a smaller portion of animals treated at 17–26 months of age exhibited rescue compared to those treated at younger ages21.

Successful RPE65 gene replacement therapy has also been shown using the Swedish Briard dog model. rAAV was subretinally injected at a variety of ages from 1 month to 4 years and resulted in significantly improved visual function (ERG), retinoid content, and visual behavior29,47,53,5558. Functional recovery was evident within 2 weeks of treatment53,57 and peaked at 3 months post injection57,59,60 but substantial rescue persisted for more than 7 years53. Interestingly, improvement in photopic ERG responses was also seen in the contralateral, untreated control eye in the treated dogs58. As with the rodent models, it was reported that functional recovery is better when treatment occurs earlier (8 months) rather than later (30 months)57,60. Although most studies in dogs and rodents report improvements in both scotopic and photopic visual function, some cases observed that improvements were limited to cones42,54. It has been reported that photopic ERG amplitudes were increased, that cone numbers were increased, and that in some cases cone rescue appeared earlier and persisted longer than improvements in rods42,52,59.

Various groups have worked to optimize rAAV vectors for RPE65 gene transfer, expression, and rescue. It was reported that inclusion of a modified Kozak sequence at the translational start site and the incorporation of a lengthy stuffer sequence improve gene expression. Neither varying the rAAV2 serotype nor the promoter had a significant impact on gene expression, but the site of injection proved critical. While subretinal injections led to RPE65 expression in the RPE and functional rescue, neither was observed following intravitreal injection. After subretinal injection, gene expression was usually limited to RPE cells surrounding the injection area but in some cases the expression was more widely distributed55. As a final step before beginning clinical trials, toxicity studies were undertaken in rodents, dogs and primates. Generally treatment was well-tolerated with no significant adverse effects45,53,55. Occasionally inflammation presented soon after treatment, but resolved by three months post-injection in dogs61 and by one week post-injection in monkeys62. Overall, RPE65 gene replacement therapy using rAAV vectors has been successful by multiple standards; the treatment is safe and effective, and rescue is persistent in multiple animal models. This body of work has paved the way for the exciting clinical trials now ongoing for people suffering from LCA2.

Recently, very encouraging reports were published by three different groups who are treating LCA2 patients as part of Phase 1 clinical trials6365. Although the purpose of these initial studies was merely to test the safety of the vector, all three groups do report modest improvements in visual acuity. Maguire et al. subretinally injected rAAV containing human RPE65 cDNA under the control of the ubiquitous chicken β-actin promoter into the eyes of three young adults. They report that 3/3 patients showed evidence of improvement in retinal function based on testing of visual acuity and pupillometry (for papillary light reflex). After treatment, the sensitivity of papillary response to light in the treated eyes was three times higher than the baseline65. Visual acuity improved and the visual field was enlarged 2 weeks after treatment65. The second group headed by Robin Ali used an rAAV vector containing human RPE65 cDNA and the human RPE65 promoter63. Of the three initial patients treated, one showed evidence of improvement in retinal function by microperimetry, dark-adapted perimetry, and visual mobility63. The third group reported increased light sensitivity and expanded visual fields in treated patients as early as PI-3064. Although treatment was limited to the area of the injection, patients reported substantial improvements64. Importantly, none of the groups observed any significant intraocular inflammation or immune responses.

As a result of the groundbreaking positive reports, these three trials (NCT00516477, NCT00643747, NCT00481546 www.clinicaltrials.gov) are ongoing and expanding their patient population to further examine treatment safety and efficacy. The success of RPE65 gene therapy in both animal models and humans provides great promise for the treatment of blinding LCA2. In addition, results from all these studies will be relevant to the larger gene therapy research community in terms of study design and model selection, vector optimization and delivery, and outcome assessment.

References

  • 1.Travis GH, Golczak M, Moise AR, Palczewski K. Diseases caused by defects in the visual cycle: retinoids as potential therapeutic agents. Annual review of pharmacology and toxicology. 2007;47:469. doi: 10.1146/annurev.pharmtox.47.120505.105225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Bavik CO, Busch C, Eriksson U. Characterization of a plasma retinol-binding protein membrane receptor expressed in the retinal pigment epithelium. The Journal of biological chemistry. 1992;267 (32):23035. [PubMed] [Google Scholar]
  • 3.Redmond TM, et al. Rpe65 is necessary for production of 11-cis-vitamin A in the retinal visual cycle. Nature genetics. 1998;20 (4):344. doi: 10.1038/3813. [DOI] [PubMed] [Google Scholar]
  • 4.Hamel CP, et al. Molecular cloning and expression of RPE65, a novel retinal pigment epithelium-specific microsomal protein that is post-transcriptionally regulated in vitro. The Journal of biological chemistry. 1993;268 (21):15751. [PubMed] [Google Scholar]
  • 5.Nicoletti A, et al. Molecular characterization of the human gene encoding an abundant 61 kDa protein specific to the retinal pigment epithelium. Human molecular genetics. 1995;4 (4):641. doi: 10.1093/hmg/4.4.641. [DOI] [PubMed] [Google Scholar]
  • 6.Hamel CP, et al. A developmentally regulated microsomal protein specific for the pigment epithelium of the vertebrate retina. Journal of neuroscience research. 1993;34 (4):414. doi: 10.1002/jnr.490340406. [DOI] [PubMed] [Google Scholar]
  • 7.Bavik CO, et al. The retinal pigment epithelial membrane receptor for plasma retinol-binding protein. Isolation and cDNA cloning of the 63-kDa protein. The Journal of biological chemistry. 1993;268 (27):20540. [PubMed] [Google Scholar]
  • 8.Hamel CP, et al. The gene for the retinal pigment epithelium-specific protein RPE65 is localized to human 1p31 and mouse 3. Genomics. 1994;20 (3):509. doi: 10.1006/geno.1994.1212. [DOI] [PubMed] [Google Scholar]
  • 9.Aguirre GD, et al. Congenital stationary night blindness in the dog: common mutation in the RPE65 gene indicates founder effect. Molecular vision. 1998;4:23. [PubMed] [Google Scholar]
  • 10.Manes G, et al. Rat messenger RNA for the retinal pigment epithelium-specific protein RPE65 gradually accumulates in two weeks from late embryonic days. FEBS letters. 1998;423 (2):133. doi: 10.1016/s0014-5793(98)00081-7. [DOI] [PubMed] [Google Scholar]
  • 11.Gal A, et al. Mutation spectrum of PRE65 in autosomal recessive childhood-onset severe retinal destrophy. Investigative ophthalmology & visual science. 1998;39:S901. [Google Scholar]
  • 12.Galbavy ES, Olson MD. Morphogenesis of rod cells in the retina of the albino rat: a scanning electron microscopic study. The Anatomical record. 1979;195 (4):707. doi: 10.1002/ar.1091950410. [DOI] [PubMed] [Google Scholar]
  • 13.Xue L, et al. A palmitoylation switch mechanism in the regulation of the visual cycle. Cell. 2004;117 (6):761. doi: 10.1016/j.cell.2004.05.016. [DOI] [PubMed] [Google Scholar]
  • 14.Takahashi Y, Moiseyev G, Chen Y, Ma JX. The roles of three palmitoylation sites of RPE65 in its membrane association and isomerohydrolase activity. Investigative ophthalmology & visual science. 2006;47 (12):5191. doi: 10.1167/iovs.06-0614. [DOI] [PubMed] [Google Scholar]
  • 15.Jin M, Yuan Q, Li S, Travis GH. Role of LRAT on the retinoid isomerase activity and membrane association of Rpe65. The Journal of biological chemistry. 2007;282 (29):20915. doi: 10.1074/jbc.M701432200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Jin M, et al. Rpe65 is the retinoid isomerase in bovine retinal pigment epithelium. Cell. 2005;122 (3):449. doi: 10.1016/j.cell.2005.06.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Moiseyev G, et al. RPE65 is the isomerohydrolase in the retinoid visual cycle. Proceedings of the National Academy of Sciences of the United States of America. 2005;102 (35):12413. doi: 10.1073/pnas.0503460102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Redmond TM, et al. Mutation of key residues of RPE65 abolishes its enzymatic role as isomerohydrolase in the visual cycle. Proceedings of the National Academy of Sciences of the United States of America. 2005;102 (38):13658. doi: 10.1073/pnas.0504167102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Moiseyev G, et al. RPE65 is an iron(II)-dependent isomerohydrolase in the retinoid visual cycle. The Journal of biological chemistry. 2006;281 (5):2835. doi: 10.1074/jbc.M508903200. [DOI] [PubMed] [Google Scholar]
  • 20.Znoiko SL, Crouch RK, Moiseyev G, Ma JX. Identification of the RPE65 protein in mammalian cone photoreceptors. Investigative ophthalmology & visual science. 2002;43 (5):1604. [PubMed] [Google Scholar]
  • 21.Jacobson SG, et al. Identifying photoreceptors in blind eyes caused by RPE65 mutations: Prerequisite for human gene therapy success. Proceedings of the National Academy of Sciences of the United States of America. 2005;102 (17):6177. doi: 10.1073/pnas.0500646102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Mata NL, et al. Rpe65 is a retinyl ester binding protein that presents insoluble substrate to the isomerase in retinal pigment epithelial cells. The Journal of biological chemistry. 2004;279 (1):635. doi: 10.1074/jbc.M310042200. [DOI] [PubMed] [Google Scholar]
  • 23.Seeliger MW, et al. New views on RPE65 deficiency: the rod system is the source of vision in a mouse model of Leber congenital amaurosis. Nature genetics. 2001;29 (1):70. doi: 10.1038/ng712. [DOI] [PubMed] [Google Scholar]
  • 24.Feathers KL, et al. Nrl-knockout mice deficient in Rpe65 fail to synthesize 11-cis retinal and cone outer segments. Investigative ophthalmology & visual science. 2008;49 (3):1126. doi: 10.1167/iovs.07-1234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Wenzel A, et al. RPE65 is essential for the function of cone photoreceptors in NRL-deficient mice. Investigative ophthalmology & visual science. 2007;48 (2):534. doi: 10.1167/iovs.06-0652. [DOI] [PubMed] [Google Scholar]
  • 26.Thompson DA, Gal A. Vitamin A metabolism in the retinal pigment epithelium: genes, mutations, and diseases. Prog Retin Eye Res. 2003;22 (5):683. doi: 10.1016/s1350-9462(03)00051-x. [DOI] [PubMed] [Google Scholar]
  • 27.Morimura H, et al. Mutations in the RPE65 gene in patients with autosomal recessive retinitis pigmentosa or leber congenital amaurosis. Proceedings of the National Academy of Sciences of the United States of America. 1998;95 (6):3088. doi: 10.1073/pnas.95.6.3088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gu SM, et al. Mutations in RPE65 cause autosomal recessive childhood-onset severe retinal dystrophy. Nature genetics. 1997;17 (2):194. doi: 10.1038/ng1097-194. [DOI] [PubMed] [Google Scholar]
  • 29.Aguirre GK, et al. Canine and human visual cortex intact and responsive despite early retinal blindness from RPE65 mutation. PLoS medicine. 2007;4 (6):e230. doi: 10.1371/journal.pmed.0040230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Chen Y, Moiseyev G, Takahashi Y, Ma JX. Impacts of two point mutations of RPE65 from Leber’s congenital amaurosis on the stability, subcellular localization and isomerohydrolase activity of RPE65. FEBS letters. 2006;580 (17):4200. doi: 10.1016/j.febslet.2006.06.078. [DOI] [PubMed] [Google Scholar]
  • 31.Takahashi Y, Chen Y, Moiseyev G, Ma JX. Two point mutations of RPE65 from patients with retinal dystrophies decrease the stability of RPE65 protein and abolish its isomerohydrolase activity. The Journal of biological chemistry. 2006;281 (31):21820. doi: 10.1074/jbc.M603725200. [DOI] [PubMed] [Google Scholar]
  • 32.Samardzija M, et al. R91W mutation in Rpe65 leads to milder early-onset retinal dystrophy due to the generation of low levels of 11-cis-retinal. Human molecular genetics. 2008;17 (2):281. doi: 10.1093/hmg/ddm304. [DOI] [PubMed] [Google Scholar]
  • 33.Lorenz B, et al. Investigative ophthalmology & visual science. 2008. A novel RPE65 hypomorph expands the clinical phenotype of RPE65 mutations. A comprehensive clinical and biochemical functional study. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Marlhens F, et al. Autosomal recessive retinal dystrophy associated with two novel mutations in the RPE65 gene. Eur J Hum Genet. 1998;6 (5):527. doi: 10.1038/sj.ejhg.5200205. [DOI] [PubMed] [Google Scholar]
  • 35.Nilsson SE, Wrigstad A, Narfstrom K. Changes in the DC electroretinogram in Briard dogs with hereditary congenital night blindness and partial day blindness. Experimental eye research. 1992;54 (2):291. doi: 10.1016/s0014-4835(05)80218-0. [DOI] [PubMed] [Google Scholar]
  • 36.Pang JJ, et al. Retinal degeneration 12 (rd12): a new, spontaneously arising mouse model for human Leber congenital amaurosis (LCA) Molecular vision. 2005;11:152. [PubMed] [Google Scholar]
  • 37.Rohrer B, et al. Correlation of regenerable opsin with rod ERG signal in Rpe65−/− mice during development and aging. Investigative ophthalmology & visual science. 2003;44 (1):310. doi: 10.1167/iovs.02-0567. [DOI] [PubMed] [Google Scholar]
  • 38.Ablonczy Z, et al. 11-cis-retinal reduces constitutive opsin phosphorylation and improves quantum catch in retinoid-deficient mouse rod photoreceptors. The Journal of biological chemistry. 2002;277 (43):40491. doi: 10.1074/jbc.M205507200. [DOI] [PubMed] [Google Scholar]
  • 39.Cottet S, et al. Biological characterization of gene response in Rpe65−/− mouse model of Leber’s congenital amaurosis during progression of the disease. Faseb J. 2006;20 (12):2036. doi: 10.1096/fj.06-6211com. [DOI] [PubMed] [Google Scholar]
  • 40.Znoiko SL, et al. Downregulation of cone-specific gene expression and degeneration of cone photoreceptors in the Rpe65−/− mouse at early ages. Investigative ophthalmology & visual science. 2005;46 (4):1473. doi: 10.1167/iovs.04-0653. [DOI] [PubMed] [Google Scholar]
  • 41.Ekesten B, Gouras P, Salchow DJ. Ultraviolet and middle wavelength sensitive cone responses in the electroretinogram (ERG) of normal and Rpe65 −/−mice. Vision research. 2001;41 (19):2425. doi: 10.1016/s0042-6989(01)00140-7. [DOI] [PubMed] [Google Scholar]
  • 42.Chen Y, Moiseyev G, Takahashi Y, Ma JX. RPE65 gene delivery restores isomerohydrolase activity and prevents early cone loss in Rpe65−/− mice. Investigative ophthalmology & visual science. 2006;47 (3):1177. doi: 10.1167/iovs.05-0965. [DOI] [PubMed] [Google Scholar]
  • 43.Pang JJ, et al. Gene therapy restores vision-dependent behavior as well as retinal structure and function in a mouse model of RPE65 Leber congenital amaurosis. Mol Ther. 2006;13 (3):565. doi: 10.1016/j.ymthe.2005.09.001. [DOI] [PubMed] [Google Scholar]
  • 44.Veske A, Nilsson SE, Narfstrom K, Gal A. Retinal dystrophy of Swedish briard/briard-beagle dogs is due to a 4-bp deletion in RPE65. Genomics. 1999;57 (1):57. doi: 10.1006/geno.1999.5754. [DOI] [PubMed] [Google Scholar]
  • 45.Narfstrom K, et al. Functional and structural recovery of the retina after gene therapy in the RPE65 null mutation dog. Investigative ophthalmology & visual science. 2003;44 (4):1663. doi: 10.1167/iovs.02-0595. [DOI] [PubMed] [Google Scholar]
  • 46.Wrigstad A. Hereditary dystrophy of the retinna and the retinal pigment epithelium in a strain of Briard dogs: a clinical, morphological and electrophysiological study. Linkoping Univ Med Dissertations. 1994;423:1. [Google Scholar]
  • 47.Acland GM, et al. Gene therapy restores vision in a canine model of childhood blindness. Nature genetics. 2001;28 (1):92. doi: 10.1038/ng0501-92. [DOI] [PubMed] [Google Scholar]
  • 48.Buch PK, Bainbridge JW, Ali RR. AAV-mediated gene therapy for retinal disorders: from mouse to man. Gene therapy. 2008;15 (11):849. doi: 10.1038/gt.2008.66. [DOI] [PubMed] [Google Scholar]
  • 49.Bemelmans AP, et al. Lentiviral gene transfer-mediated cone vision restoration in RPE65 knockout mice. Advances in experimental medicine and biology. 2008;613:89. doi: 10.1007/978-0-387-74904-4_9. [DOI] [PubMed] [Google Scholar]
  • 50.Bemelmans AP, et al. Lentiviral gene transfer of RPE65 rescues survival and function of cones in a mouse model of Leber congenital amaurosis. PLoS medicine. 2006;3 (10):e347. doi: 10.1371/journal.pmed.0030347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Dejneka NS, et al. In utero gene therapy rescues vision in a murine model of congenital blindness. Mol Ther. 2004;9 (2):182. doi: 10.1016/j.ymthe.2003.11.013. [DOI] [PubMed] [Google Scholar]
  • 52.Nusinowitz S, et al. Cortical visual function in the rd12 mouse model of Leber Congenital Amarousis (LCA) after gene replacement therapy to restore retinal function. Vision research. 2006;46 (22):3926. doi: 10.1016/j.visres.2006.05.010. [DOI] [PubMed] [Google Scholar]
  • 53.Bennicelli J, et al. Reversal of blindness in animal models of leber congenital amaurosis using optimized AAV2-mediated gene transfer. Mol Ther. 2008;16 (3):458. doi: 10.1038/sj.mt.6300389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Lai CM, et al. Recombinant adeno-associated virus type 2-mediated gene delivery into the Rpe65−/− knockout mouse eye results in limited rescue. Genetic vaccines and therapy. 2004;2 (1):3. doi: 10.1186/1479-0556-2-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Acland GM, et al. Long-term restoration of rod and cone vision by single dose rAAV-mediated gene transfer to the retina in a canine model of childhood blindness. Mol Ther. 2005;12 (6):1072. doi: 10.1016/j.ymthe.2005.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Jacobs JB, et al. Eye movement recordings as an effectiveness indicator of gene therapy in RPE65-deficient canines: implications for the ocular motor system. Investigative ophthalmology & visual science. 2006;47 (7):2865. doi: 10.1167/iovs.05-1233. [DOI] [PubMed] [Google Scholar]
  • 57.Le Meur G, et al. Restoration of vision in RPE65-deficient Briard dogs using an AAV serotype 4 vector that specifically targets the retinal pigmented epithelium. Gene therapy. 2007;14 (4):292. doi: 10.1038/sj.gt.3302861. [DOI] [PubMed] [Google Scholar]
  • 58.Narfstrom K, et al. In vivo gene therapy in young and adult RPE65−/− dogs produces long-term visual improvement. The Journal of heredity. 2003;94 (1):31. doi: 10.1093/jhered/esg015. [DOI] [PubMed] [Google Scholar]
  • 59.Narfstrom K, et al. Assessment of structure and function over a 3-year period after gene transfer in RPE65−/− dogs. Documenta ophthalmologica. 2005;111 (1):39. doi: 10.1007/s10633-005-3159-0. [DOI] [PubMed] [Google Scholar]
  • 60.Rolling F, et al. Gene therapeutic prospects in early onset of severe retinal dystrophy: restoration of vision in RPE65 Briard dogs using an AAV serotype 4 vector that specifically targets the retinal pigmented epithelium. Bulletin et memoires de l’Academie royale de medecine de Belgique. 2006;161 (10–12):497. [PubMed] [Google Scholar]
  • 61.Jacobson SG, et al. Safety of recombinant adeno-associated virus type 2-RPE65 vector delivered by ocular subretinal injection. Mol Ther. 2006;13 (6):1074. doi: 10.1016/j.ymthe.2006.03.005. [DOI] [PubMed] [Google Scholar]
  • 62.Jacobson SG, et al. Safety in nonhuman primates of ocular AAV2-RPE65, a candidate treatment for blindness in Leber congenital amaurosis. Human gene therapy. 2006;17 (8):845. doi: 10.1089/hum.2006.17.845. [DOI] [PubMed] [Google Scholar]
  • 63.Bainbridge JW, et al. Effect of gene therapy on visual function in Leber’s congenital amaurosis. The New England journal of medicine. 2008;358 (21):2231. doi: 10.1056/NEJMoa0802268. [DOI] [PubMed] [Google Scholar]
  • 64.Cideciyan AV, et al. Human gene therapy for RPE65 isomerase deficiency activates the retinoid cycle of vision but with slow rod kinetics. Proceedings of the National Academy of Sciences of the United States of America. 2008;105 (39):15112. doi: 10.1073/pnas.0807027105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Maguire AM, et al. Safety and efficacy of gene transfer for Leber’s congenital amaurosis. The New England journal of medicine. 2008;358 (21):2240. doi: 10.1056/NEJMoa0802315. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES