UvA-DARE (Digital Academic Repository)

Function and pathology of the human retinal pigment epithelium

Booij, J.C.

Publication date 2010 Document Version Final published version

Link to publication

Citation for published version (APA): Booij, J. C. (2010). Function and pathology of the human retinal pigment epithelium.

General rights It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons).

Disclaimer/Complaints regulations If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You will be contacted as soon as possible.

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

Download date:04 Oct 2021 Function and pathology of the human retinal pigment epithelium Function and pathology of the human retinal pigment epithelium

Judith C. Booij

Judith C. Booij

Booij_Omslag.indd 1 18-06-10 10:24 Function and pathology of the human retinal pigment epithelium

Function and pathology of the human retinal pigment epithelium

ACADEMISCH PROEFSCHRIFT

ter verkrijging van de graad van doctor

aan de Universiteit van Amsterdam

op gezag van de Rector Magnificus

Mw. prof. dr. D.C. van den Boom

ten overstaan van een door het college voor promoties ingestelde

commissie, in het openbaar te verdedigen in de Agnietenkapel

op 9 juli 2010, te 10 uur

door

Judith Cornelia Booij

geboren te Delft Promotiecommissie:

Promotor: Prof. dr. A.A.B. Bergen

Overige leden: Prof. dr. J.M.F.G. Aerts Prof. dr. M. Kamermans Prof. dr. R.O. Schlingemann Dr. C.C.W. Klaver

Faculteit der geneeskunde

The research conducted in this thesis was conducted at the Netherlands Institute for Neuro- science.

Publication of this thesis was financially supported by: Nederlands Instituut voor Neurowetenschappen (NIN) de Landelijke Stichting voor Blinden en Slechtzienden (LSBS) Stichting Blindenhulp Rotterdamse Vereniging blindenbelangen

Cover: Simone Vinke, Ridderprint & J. C. Booij Book design: J. C. Booij Print: Ridderprint

ISBN: 978-90-5335-293-9 Contents

1 Introduction 7

2 Functional annotation of the human retinal pigment epithelium transcriptome 19 BMC Genomics 2009, 10:164-182

3 A new strategy to identify and annotate human RPE-specific expression 45 PlosOne 2010, 5(3) e9341

4 The dynamic nature of Bruch’s membrane 77 Prog Retin Eye Res. 2010 Jan;29(1):1-18

5 On the origin of Bruch’s membrane and drusen 117 manuscript in preparation

6 Identification of mutations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 in patients with juvenile retinitis pigmentosa 139 J Med Genet. 2005 Nov;42(11):e67-e75

7 Course of visual decline in relation to the Best1 genotype in vitelliform macular dystrophy 155 Ophthalmology, 2010 april 8 epub ahead of print

8 Simultaneous mutation detection in 90 retinal disease genes in multiple patients using a custom designed 300 kb retinal resequencing chip. 169 Ophthalmology (2010), in press

9 General discussion 189

10 Summary 201

11 Nederlandse samenvatting 209

12 List of publications 217

13 Dankwoord 221

14 Curriculum vitae 225

15 Color figures 229

Introduction

1 Chapter 1

The retina is a complex multilayered neural structure lining the back of the eye. It consists of the neural retina and the retinal pigment epithelium (RPE) with Bruch’s membrane (BM), the choroid and the sclera as supportive layers underneath. The neural retina contains nine layers and many different cell types including the photoreceptor cells. The RPE has multiple functions, which are described below[1].

When light is projected onto the retina, the different retinal cell types work together to con- vert the light into neuro-chemical and electrical signals that are sent through the optic nerve and to the brain. The first and one of the most crucial components in this system is the pho- totransduction cascade, a biochemical system involved in maintaining the light-sensitivity of the photoreceptor cells. The proteins that play a role in the phototransduction cascade are present in both the RPE and the photoreceptors and some of them are actively transported be- tween the two cell types (see Figure 1)[2-6]. The genes encoding these proteins are expressed at high levels in the RPE and the photoreceptor cells[7,8].

Figure 1. Schematic overview of the genes involved in the phototransduction cascade along with their location in the photoreceptors and RPE.

ROS

ABCA4 RHO RDH12

visual pigments all-trans retinal

IPM 11-cis retinal all-trans retinol

11-cis retinol all-trans retinyl esters

RLBP1 LRAT RPE RGR RPE65 RDH5

ROS: rod outer segment, IPM: interphotoreceptor matrix, RPE: retinal pigment epithelium cell. RHO: rhodopsin, ABCA4: ATP-binding cassette, subfamily A, member 4, RDH12: retinol dehydrogenase 12, RLBP1: retinalde- hyde-binding 1, RGR: retinal G protein-coupled receptor, RDH5: retinol dehydrogenase 5, RPE65: retinal pigment epithelium-specific protein 65kD, LRAT: lecithin retinol acyltransferase. Circle and arrows indicate the visual cycle, letters in square boxes indicate proteins involved[9]. A color version of this figure can be found in chapter 15.

8 Introduction

Throughout the retina there is a delicate homeostasis of incoming and outgoing molecules and electrolytes, in part regulated by genes expressed and proteins present in the RPE and photoreceptors. This homeostasis is easily disturbed by, for example, a mutation in one of the genes expressed in the RPE or photoreceptors, with detrimental effects on visual function.

Embryology of the RPE Recently, the development of RPE and the retina was thoroughly described in an excellent review[10], see also Figure 2.

Figure 2. Embryological development of the retina.

Reprinted with permission from Progress in Retinal and Eye Research, 14,1, Reichenbach A and Pritz-Hohmeier S., Normal and disturbed early development of the eye Anlagen, 1-45, Copyright (1995), with permission from Elsevier. A color version of this figure can be found in chapter 15.

In short, both RPE and photoreceptors are highly dependent upon each other for their de- velopment: the embryonic RPE secretes factors required for the development of the photo- receptors, and the developing photoreceptors contribute to the development of the RPE[10]. Consequently, in case of loss of the RPE, the photoreceptors and even the entire retina do not develop normally[10]. The retina develops from the neuroectoderm of the diencephalon. Around the fourth week of gestation a fold is created in the neuro-epithelium, which forms the future optic cup. In the sixth week, the optic cup invaginates. Two layers of neuroepi- thelial cells now face each other, separated by a remnant of lumen that subsequently forms the interphotoreceptor matrix (IPM) (Figure 2). One of the layers of neuroepithelium is to become the RPE, the other layer the neural retina. The RPE matures under the influence of locally expressed genes, including OTX2, MITF, CRBP, CRABP, IRBP and RPE65[10]. Part of the cells of the neural retina eventually differentiate into photoreceptor cells. Neural crest- derived mesenchyme is deposited around the optic cup, which will form the future choroid. The choroidal vasculature is created through angiogenesis from existing blood vessels from the paraocular mesenchyme[11]. At the same time, the primitive retina, now consisting of an RPE cell layer, an outer nuclear zone and an acellular marginal zone flanked by basal mem-

9 Chapter 1 branes, continues to differentiate. BM is formed in between the RPE basement membrane and the basement membrane of the choroidal endothelium. The RPE basement membrane becomes incorporated in the future BM. The development of BM starts on the side of the RPE, which is capable of synthesizing many of the extracellular matrix components that are present in BM[10]. Our own data show that the choroid is most likely also involved in the development and maintenance of BM (Booij et al (2010), manuscript in preparation). The RPE further differentiates into a polarized structure with the creation of an apical to basolateral polarity and the formation of tight junctions in between the RPE cells. The tight junctions between the RPE cells prevent free transport of substances through the RPE cell layer. Therefore, the RPE subsequently expresses transporters in order to allow specific sub- stances like ions and glucose to selectively pass the RPE. RPE cells also interact, via their microvilli, with the photoreceptor cells as they differentiate. The relationships between the RPE cells and rod photoreceptors on one hand and cone photo- receptors on the other hand are in many ways unique and testify to the close communication between the RPE cells and the photoreceptor cells as they develop[10].

Functional properties of the RPE The RPE is a monolayer of cells that plays an essential role in ocular development, function, maintenance and disease. Figure 3 summarizes a number of processes that take place in the matured RPE: absorption of light, transport of nutrients and ions across the RPE. regulation of the ion balance in the subretinal space, participation in the visual cycle, phagocytosis of photoreceptor waste products and secretion of growth factors.

Figure 3. Schematic overview of important functional properties of the RPE.

From: Strauss 2005, Physiol Rev Physiological review, 85 (845-881) O. Strauss et al. Reprinted with permission from The American Physiological Society. A color version of this figure can be found in chapter 15.

10 Introduction

The RPE is essential for the protection of the neural retina against oxidative damage, caused by light falling onto the retina, by phagocytosis and by degradation of photoreceptor outer segments and, finally, caused by a limited blood supply to the RPE[10]. The melanin in the RPE absorbs light, thereby reducing scattering and increasing the quality of vision. This also helps to decrease photochemiacally induced oxidative stress in the retina. Finally, the RPE contains high concentrations of antioxidants (for instance superoxide dismutase, catalase, lutein and zeaxanthin) which protects the RPE retina from oxidative damage[10]. We used our data driven analyses to assign and/or confirm a number of additional functional properties to the RPE, including ATP synthesis, ribosomal activity, phosphatidylinositol sig- naling and amino sugars metabolism.

The RPE forms the outer blood-retina barrier, preventing uncontrolled transport of substanc- es between the eye and the circulation. In addition, the RPE actively transports excess water and ions like chloride, potassium and sodium away from the subretinal space toward the circulation through a broad range of Cl--, K+-, and Na+- transporters and co-transporters, excellently reviewed by Strauss[10]. It also removes photoreceptor waste products like lactic acid and cholesterol from rod outer segment membranes. In addition, the RPE supplies the photoreceptors with nutrients like glucose by transporting them through the RPE cells from the circulation[10].

The RPE is also involved in the regulation of the ion balance in the subretinal space. When photoreceptors are stimulated by light, sodium flows into the cell and potassium flows out. The subsequent changes in sodium and potassium concentration in the subretinal space are corrected for by the RPE[10,12]. Furthermore the RPE recycles retinal from the photorecep- tor cells, which is necessary for the continuation of the visual cycle[10]. Light stimulation causes 11-cis-retinal to be converted to all-trans-retinal in the photoreceptor and is necessary for the light to be converted to an electrical signal[9] (Figure 1). The RPE secretes a number of growth factors that maintain the structure and cellular differentiation of the adjacent tis- sues, such as FGF, TGF-β, IGFR, CNTF, PDGF and PEDF[10]. In addition, the RPE func- tions as a physical support for the photoreceptors[10].

Finally, the metabolically highly active photoreceptors continuously shed and renew their outer segment membranes under the influence of light. These membranes, filled with oxi- dized and damaged photoreceptor outer segments are phagocytosed and degraded or recycled by the RPE[10].

Pathology of the RPE A disturbance of the normal function of the RPE (described above), caused, for instance by mutations in genes encoding RPE proteins, can lead to malfunction of the retina and to reti- nal degeneration. Below we will describe three important phenotypes (retinitis pigmentosa, Leber congenital amaurosis and vitelliform macular degeneration).

11 Chapter 1

Clinicial characteristics of retinal dystrophies Retinitis pigmentosa Retinitis pigmentosa (RP) is a heterogeneous group of hereditary retinal dystrophies with a worldwide prevalence of approximately 1 in 4,000[13]. The age of onset for RP patients var- ies from early childhood to late in adult life. RP is characterized by night-blindness, progres- sive constriction of the visual field, pale optic discs, narrow retinal vasculature, pigmentary changes in the retina and reduced electroretinogram (ERG) amplitudes[14].

Leber congenital amaurosis Leber congenital amaurosis (LCA) is an autosomal recessive retinal dystrophy with a world- wide prevalence of 1 in 35,000[15]. Severe visual impairment occurs in the first year of life[16,17]. LCA is characterized by a congenital nystagmus, a decreased pupillary response, and although the fundoscopic appearance may initially be normal, pigmentary changes oc- cur later in the disease process in the retina and the ERG is always severely reduced[15,18].

Despite the differences between the RP and LCA phenotypes described above, there is also considerable overlap. Numerous intermediate phenotypes exist including ‘‘early onset severe rod-cone dystrophy’’ or ‘‘early onset retinal degeneration’’[16]. A clear clinical diagnosis is not always easily established, and the diagnostic criteria are frequently not agreed upon between ophthalmologists[19].

Vitelliform macular degeneration Vitelliform macular dystrophy (VMD or Best disease) is an autosomal dominant macular disease with an estimated prevalence of 1 in 10,000[20]. The disease is characterized by the presence of one of six typical macular appearances involving the accumulation and subse- quent degradation of lipofuscin in and below the RPE (see chapter 7). The clinical diagnosis of VMD is generally confirmed by a decreased light peak-dark trough ratio (Arden ratio) on the electro-oculogram (EOG).

Inheritance pattern of retinal dystrophies In up to 29% of patients, RP is part of a syndrome, like Usher syndrome or Bardet-Biedl syndrome[13]. Of the non-syndromal RP cases in the Netherlands 30% is autosomal reces- sive (AR), 22% is autosomal dominant (AD), 10% is X-linked (XL) and 37% is isolated (IRP) or unknown[21]. In addition, several unusual inheritance patterns, such as digenic inheritance[22] and uniparental isodisomy[23], in which a child inherits two copies of a single parental , have been observed. LCA is most frequently inherited in an AR way[15]. LCA was associated with mental retardation in up to 52% of patients in an older study (reviewed by den Hollander et al[24]). Whether syndromic forms of LCA includ- ing mental retardation exist, remains controversial. However, more recent studies including computer tomography (CT) scans often exclude LCA in such patients and come to a differ- ent diagnosis[24]. VMD is most frequently inherited in an autosomal dominant manner[20], although autosomal recessive inheritance has also been described[25].

12 Introduction

Genotypes of retinal dystrophies Currently, 44 genes, when mutated, are known to be capable of causing RP[26]; eighteen (41%) of these genes have high expression levels in the RPE (top 10% of all investigated genes, Booij, Bergen unpublished data, see also chapter 3). For LCA, thirteen genes have been identified so far[26], three (23%) of these genes have high expression levels inthe RPE (Booij, Bergen unpublished data). There is extensive genetic overlap between arRP and LCA, illustrated by the fact that many genes have been implicated in both arRP and LCA: CRB1[27,28], GUCY2D[29,30], RDH12[31,32], RPE65[17,33], and TULP1[34-36]. RDH12 mutations are also found in patients with autosomal dominant and recessive cone-rod dystrophies (CORD)[37]. In the SEMA4A and ABCR genes, mutations may lead to RP and cone rod dystrophy, in addition to Stargardt’s disease[38,39]. Mutations in the IMPDH1 gene have been described in LCA as well as in a dominant form of RP[40]. In a recently published article, mutations were found in the SPATA7 gene in both juvenile RP patients and LCA pa- tients[41]. Moreover, a pedigree has been described where two parents, diagnosed with RP, had a daughter with a homozygous mutation in the RPE65 gene and an LCA phenotype[17].

In contrast to RP and LCA, VMD is caused by mutations in a single gene, Best1 (NM_004183). This gene is transcribed into the bestrophin-1 protein, that is located in the basolateral membrane of RPE cells[42-47]. The bestrophin-1 protein is expressed at high levels in the RPE[7,42] and functions as a calcium-dependent chloride channel. Although only a single gene appears to be involved in VMD, the situation is complicated by the large number of phenotypes that can be caused by mutations in the Best1 gene. With the recent finding of an autosomal dominant form of RP caused by mutations in the Best1 gene[48], there are now five partially overlapping ocular phenotypes associated with Best1 mutations: VMD, adult-onset VMD, autosomal-dominant vitreoretinochoroidopathy (ADVIRC), autosomal- recessive bestrophinopathy (ARB) and dominant RP.

Mutation detection in retinal degeneration In addition to the three retinal phenotypes described above (RP, LCA and VMD), many more diseases affecting the retina and/or the RPE can occur. For a total of thirty different retinal phenotypes, many of which show overlapping features, approximately 150 retinal disease genes have now been identified. On average more than five different genes have to be consid- ered per phenotype. This genetic heterogeneity combined with the complex nature of retinal phenotypes prevents efficient identification of disease-causing mutations[13,49]. The iden- tification of a causative mutation is important for confirmation of the clinical diagnosis and inheritance pattern, for predictions of clinical course of the disease, for genetic counseling and family planning and for future (gene-targeted) treatment.

Many different phenotypes, caused by mutations in genes expressed in the RPE or the pho- toreceptors, have been described[26]. In fact, this thesis showed that the great majority of genes involved in retinal diseases are expressed at high levels in the RPE or the photorecep- tors [8,50] (Figure 4).

13 Chapter 1

Figure 4. Schematic overview of genes involved in the visual cycle that, when mutated, are responsible for causing a retinal phenotype.

ROS Stargardt disease Macular degeneration RP RP CSNB LCA ABCA4 RHO RDH12

visual pigments all-trans retinal

IPM 11-cis retinal all-trans retinol

11-cis retinol all-trans retinyl esters

RLBP1 LRAT RPE RGR RPE65 RPA RDH5 LCA RP Fundus LCA albipunctatus

ROS: rod outer segment, IPM: interphotoreceptor matrix, RPE: retinal pigment epithelium cell. Circle and arrows indicate the visual cycle[9]. Boxed names indicate genes, see Figure 1. RP: retinitis pigmentosa, CSNB: congenital stationary night blindness, LCA: Leber congenital amaurosis, RP: retinitis punctata albicans. A color version of this figure can be found in chapter 15.

Scope of the thesis This thesis describes the functional properties of the healthy RPE (chapters two and three) and its influence on the adjacent Bruch’s membrane (BM). We based our findings on gene expression levels and functional annotations (chapters three and five), including a review of the functional properties of BM (chapter four). We also described a number of retinal degenerations that can be caused by mutations in genes expressed in the RPE: retinitis pig- mentosa (RP), Leber congenital amaurosis (LCA) (chapter six) and Best vitelliform macular degeneration (VMD) (chapter seven). Finally, we described a new large scale screening tool for the simultaneous detection of mutations in multiple retinal disease genes (chapter eight).

14 Introduction

References 1. Lange G (2000) Ophthalmology, a pocket textbook atlas. 2. van Soest S, Westerveld A, de Jong PT, Bleeker-Wagemakers EM, Bergen AA (1999) Retinitis pigmentosa: defined from a molecular point of view. Surv Ophthalmol 43: 321-334. 3. Hageman GS, Zhu XL, Waheed A, Sly WS (1991) Localization of carbonic anhydrase IV in a specific capillary bed of the human eye. Proc Natl Acad Sci U S A 88: 2716- 2720. 4. Sugiura M, Kono K, Liu H, Shimizugawa T, Minekura H et al (2002) Ceramide kinase, a novel lipid kinase. Molecular cloning and functional characterization. J Biol Chem 277: 23294-23300. 5. Surguchov A, Bronson JD, Banerjee P, Knowles JA, Ruiz C et al (1997) The human GCAP1 and GCAP2 genes are arranged in a tail-to-tail array on the short arm of chro- mosome 6 (p21.1). Genomics 39: 312-322. 6. Rice DS, Huang W, Jones HA, Hansen G, Ye GL et al (2004) Severe retinal degenera- tion associated with disruption of semaphorin 4A. Invest Ophthalmol Vis Sci 45: 2767- 2777. 7. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ et al (2009) Function- al annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 10: 164. 8. Booij JC, ten Brink JB, Swagemakers SM, Verkerk AJMH, Essing AHW et al (2010) A New Strategy to Identify and Annotate Human RPE-specific Gene Expression. Plos one in press. 9. Thompson DA, Gal A (2003) Vitamin A metabolism in the retinal pigment epithelium: genes, mutations, and diseases. Prog Retin Eye Res 22: 683-703. 10. Strauss O (2005) The retinal pigment epithelium in visual function. Physiol Rev 85: 845-881. 11. Saint-Geniez M, D’Amore PA (2004) Development and pathology of the hyaloid, cho- roidal and retinal vasculature. Int J Dev Biol 48: 1045-1058. 12. Baylor D (1996) How photons start vision. Proc Natl Acad Sci U S A 93: 560-565. 13. Haim M (2002) Epidemiology of retinitis pigmentosa in Denmark. Acta Ophthalmol Scand Suppl 1-34. 14. Merin S, Auerbach E (1976) Retinitis pigmentosa. Surv Ophthalmol 20: 303-346. 15. Koenekoop RK (2004) An overview of Leber congenital amaurosis: a model to under- stand human retinal development. Surv Ophthalmol 49: 379-398. 16. Lorenz B, Gyurus P, Preising M, Bremser D, Gu S et al (2000) Early-onset severe rod- cone dystrophy in young children with RPE65 mutations. Invest Ophthalmol Vis Sci 41: 2735-2742. 17. Morimura H, Fishman GA, Grover SA, Fulton AB, Berson EL et al (1998) Mutations in the RPE65 gene in patients with autosomal recessive retinitis pigmentosa or leber congenital amaurosis. Proc Natl Acad Sci U S A 95: 3088-3093. 18. Perrault I, Rozet JM, Gerber S, Ghazi I, Leowski C et al (1999) Leber congenital amau- rosis. Mol Genet Metab 68: 200-208.

15 Chapter 1

19. Rivolta C, Sharon D, DeAngelis MM, Dryja TP (2002) Retinitis pigmentosa and allied diseases: numerous diseases, genes, and inheritance patterns. Hum Mol Genet 11: 1219- 1227. 20. Mohler CW, Fine SL (1981) Long-term evaluation of patients with Best’s vitelliform dystrophy. Ophthalmology 88: 688-692. 21. van den Born LI, Bergen AA, Bleeker-Wagemakers EM (1992) A retrospective study of registered retinitis pigmentosa patients in The Netherlands. Ophthalmic Paediatr Genet 13: 227-236. 22. Kajiwara K, Berson EL, Dryja TP (1994) Digenic retinitis pigmentosa due to mutations at the unlinked peripherin/RDS and ROM1 loci. Science 264: 1604-1608. 23. Thompson DA, McHenry CL, Li Y, Richards JE, Othman MI et al (2002) Retinal dys- trophy due to paternal isodisomy for chromosome 1 or chromosome 2, with homoallel- ism for mutations in RPE65 or MERTK, respectively. Am J Hum Genet 70: 224-229. 24. den Hollander AI, Roepman R, Koenekoop RK, Cremers FP (2008) Leber congenital amaurosis: genes, proteins and disease mechanisms. Prog Retin Eye Res 27: 391-419. 25. Boon CJ, Klevering BJ, Leroy BP, Hoyng CB, Keunen JE et al (2009) The spectrum of ocular phenotypes caused by mutations in the BEST1 gene. Prog Retin Eye Res 28: 187-205. 26. Retnet www.sph.uth.tmc.edu/Retnet/ 27. den Hollander AI, ten Brink JB, de Kok YJ, van SS, van den Born LI et al (1999) Muta- tions in a human homologue of Drosophila crumbs cause retinitis pigmentosa (RP12). Nat Genet 23: 217-221. 28. den Hollander AI, Heckenlively JR, van den Born LI, de Kok YJ, van der Velde-Visser SD et al (2001) Leber congenital amaurosis and retinitis pigmentosa with Coats-like ex- udative vasculopathy are associated with mutations in the crumbs homologue 1 (CRB1) gene. Am J Hum Genet 69: 198-203. 29. Perrault I, Rozet JM, Calvas P, Gerber S, Camuzat A et al (1996) Retinal-specific gua- nylate cyclase gene mutations in Leber’s congenital amaurosis. Nat Genet 14: 461-464. 30. Perrault I, Hanein S, Gerber S, Lebail B, Vlajnik P et al (2005) A novel mutation in the GUCY2D gene responsible for an early onset severe RP different from the usual GUCY2D-LCA phenotype. Hum Mutat 25: 222. 31. Perrault I, Hanein S, Gerber S, Barbet F, Ducroq D et al (2004) Retinal dehydrogenase 12 (RDH12) mutations in leber congenital amaurosis. Am J Hum Genet 75: 639-646. 32. Janecke AR, Thompson DA, Utermann G, Becker C, Hubner CA et al (2004) Mutations in RDH12 encoding a photoreceptor cell retinol dehydrogenase cause childhood-onset severe retinal dystrophy. Nat Genet 36: 850-854. 33. Marlhens F, Bareil C, Griffoin JM, Zrenner E, Amalric P et al (1997) Mutations in RPE65 cause Leber’s congenital amaurosis. Nat Genet 17: 139-141. 34. Banerjee P, Kleyn PW, Knowles JA, Lewis CA, Ross BM et al (1998) TULP1 mutation in two extended Dominican kindreds with autosomal recessive retinitis pigmentosa. Nat Genet 18: 177-179. 35. Gu S, Lennon A, Li Y, Lorenz B, Fossarello M et al (1998) Tubby-like protein-1 muta- tions in autosomal recessive retinitis pigmentosa. Lancet 351: 1103-1104.

16 Introduction

36. Hanein S, Perrault I, Gerber S, Tanguy G, Barbet F et al (2004) Leber congenital am- aurosis: comprehensive survey of the genetic heterogeneity, refinement of the clinical definition, and genotype-phenotype correlations as a strategy for molecular diagnosis. Hum Mutat 23: 306-317. 37. Sun W, Gerth C, Maeda A, Lodowski DT, Van Der KL et al (2007) Novel RDH12 muta- tions associated with Leber congenital amaurosis and cone-rod dystrophy: biochemical and clinical evaluations. Vision Res 47: 2055-2066. 38. Abid A, Ismail M, Mehdi SQ, Khaliq S (2006) Identification of novel mutations in the SEMA4A gene associated with retinal degenerative diseases. J Med Genet 43: 378-381. 39. Shroyer NF, Lewis RA, Yatsenko AN, Lupski JR (2001) Null missense ABCR (ABCA4) mutations in a family with stargardt disease and retinitis pigmentosa. Invest Ophthalmol Vis Sci 42: 2757-2761. 40. Bowne SJ, Sullivan LS, Mortimer SE, Hedstrom L, Zhu J et al (2006) Spectrum and frequency of mutations in IMPDH1 associated with autosomal dominant retinitis pig- mentosa and leber congenital amaurosis. Invest Ophthalmol Vis Sci 47: 34-42. 41. Wang H, den Hollander AI, Moayedi Y, Abulimiti A, Li Y et al (2009) Mutations in SPATA7 cause Leber congenital amaurosis and juvenile retinitis pigmentosa. Am J Hum Genet 84: 380-387. 42. Hartzell HC, Qu Z, Yu K, Xiao Q, Chien LT (2008) Molecular physiology of bestro- phins: multifunctional membrane proteins linked to best disease and other retinopathies. Physiol Rev 88: 639-672. 43. Sun H, Tsunenari T, Yau KW, Nathans J (2002) The vitelliform macular dystrophy pro- tein defines a new family of chloride channels. Proc Natl Acad Sci U S A 99: 4008- 4013. 44. Xiao Q, Prussia A, Yu K, Cui YY, Hartzell HC (2008) Regulation of bestrophin Cl chan- nels by calcium: role of the C terminus. J Gen Physiol 132: 681-692. 45. Qu Z, Chien LT, Cui Y, Hartzell HC (2006) The anion-selective pore of the bestrophins, a family of chloride channels associated with retinal degeneration. J Neurosci 26: 5411- 5419. 46. Pusch M (2004) Ca(2+)-activated chloride channels go molecular. J Gen Physiol 123: 323-325. 47. Qu Z, Fischmeister R, Hartzell C (2004) Mouse bestrophin-2 is a bona fide Cl(-) chan- nel: identification of a residue important in anion binding and conduction. J Gen Physiol 123: 327-340. 48. Davidson AE, Millar ID, Urquhart JE, Burgess-Mullan R, Shweikh Y et al (2009) Mis- sense mutations in a retinal pigment epithelium protein, bestrophin-1, cause retinitis pigmentosa. Am J Hum Genet 85: 581-592. 49. Booij JC, Florijn RJ, ten Brink JB, Loves W, Meire F et al (2005) Identification of mu- tations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. J Med Genet 42: e67. 50. van Soest S, de Wit GM, Essing AH, ten Brink JB, Kamphuis W et al (2007) Compari- son of human retinal pigment epithelium gene expression in macula and periphery high- lights potential topographic differences in Bruch’s membrane. Mol Vis 13: 1608-1617.

17

Functional annotation of the hu- man retinal pigment epithelium transcriptome

2

Judith C Booij, Simone van Soest, Sigrid MA Swagemakers, Anke HW Essing, Annemieke JMH Verkerk, Peter J van der Spek, Theo GMF Gorgels and Arthur AB Bergen

BMC Genomics 2009, 10:164-182 Chapter 2

Abstract Background: To determine level, variability and functional annotation of gene expression of the human retinal pigment epithelium (RPE), a key tissue involved in retinal diseases like age-related macular degeneration and retinitis pigmentosa. Macular RPE cells from six se- lected healthy human donor eyes (aged 63-78 years) were laser dissected and used for 22K microarray studies (Agilent technologies). Data were analyzed with Rosetta Resolver, the web tool DAVID and Ingenuity software.

Results: In total, we identified 19,746 array entries with significant expression in the RPE. Gene expression was analyzed according to expression levels, interindividual variability and functionality. A group of highly (n=2,194) expressed RPE genes showed an overrepresenta- tion of genes of the oxidative phosphorylation, ATP synthesis and ribosome pathways. In the group of moderately expressed genes (n=8,776) genes of the phosphatidylinositol signaling system and aminosugars metabolism were overrepresented. As expected, the top 10 percent (n=2,194) of genes with the highest interindividual differences in expression showed func- tional overrepresentation of the complement cascade, essential in inflammation in age-related macular degeneration, and other signaling pathways. Surprisingly, this same category also includes the genes involved in Bruch’s membrane (BM) composition. Among the top 10 per- cent of genes with low interindividual differences, there was an overrepresentation of genes involved in local glycosaminoglycan turnover.

Conclusions: Our study expands current knowledge of the RPE transcriptome by assigning new genes, and adding data about expression level and interindividual variation. Functional annotation suggests that the RPE has high levels of protein synthesis, strong energy demands, and is exposed to high levels of oxidative stress and a variable degree of inflammation. Our data sheds new light on the molecular composition of BM, adjacent to the RPE, and is useful for candidate retinal disease gene identification or gene dose-dependent therapeutic studies.

20 The human retinal pigment epithelium transcriptome

Background The retinal pigment epithelium (RPE) is a multifunctional neural-crest derived cell layer, flanked by the photoreceptor cells on the apical side and the Bruch’s membrane (BM)/cho- roid complex on the basolateral side. Among others, the RPE supplies the photoreceptors with nutrients, regulates the ion balance in the subretinal space and recycles retinal from the photoreceptor cells, which is necessary for the continuation of the visual cycle.1 It also phagocytoses and degrades photoreceptor outer segments and absorbs light that is projected onto the retina[1]. Finally, the RPE secretes a number of growth factors that maintain the structure and cellular differentiation of the adjacent tissues[1]. The importance of the RPE in vision is illustrated by the major involvement of this mono- layer of cells in genetically determined retinal diseases like age related macular degeneration (AMD) and retinitis pigmentosa (RP) [2]. Since the great majority of genes implicated in AMD or RP are expressed in either the RPE or the photoreceptors, the identification of ad- ditional genes highly expressed in the RPE may provide valuable clues in the search for new genes involved in retinal disease[2-6]. Obviously, the functional properties of RPE cells are determined by the genes they express and the proteins they encode. Although the RPE cell is one of the best studied neural cell types[3-12], large scale assignment of expressed genes to the RPE has been largely depen- dent on RNA based studies. Assignment of proteins to the RPE has been hampered by its autofluorescence and melanin content. Large-scale RPE related expression studies were per- formed using cDNA arrays, serial analysis of gene expression (SAGE), expressed sequence tag (EST) analysis, and multiple RT-PCRs. The number of eyes used in these studies ranged from one to fifteen, and the number of genes under investigation from 29 to 30,000[8-12]. While these studies provided valuable information, they were limited in either the number of genes or the number of eyes under investigation, or they lacked specificity due to the tis- sue sampling method used. Moreover, most or all of these studies focused on the mean gene expression profile of all samples together, rather than documenting potential interindividual differences[8-12]. A robust and specific dataset on RPE expression levels from a substantial number of individuals is lacking and a great deal remains unknown with regard to the inter- individual expression differences. A number of biological processes and cellular functions of genes expressed in the RPE were described in three of the above mentioned studies[8,10,12]. All three identified pro- tein metabolism and signal transduction as an important functional class of genes expressed by the RPE[8,10,12]. Similarly, cell structure[8,10], cell proliferation[8,10], gene transcrip- tion[10,11] and energy metabolism were described in two out of three studies. Finally, in- dividual studies also identified overrepresentation of membrane proteins[10], transport or channel proteins[10], heat shock proteins[10] and vitamin A metabolism[11]. In a recent microarray study we compared RPE gene expression in the macula with the retinal periphery and demonstrated, among other things, consistent differential expression of extracellular ma- trix genes corresponding with proteins in BM[13].

21 Chapter 2

The aim of the current study is to describe the gene expression levels and the interindividual variation in gene expression of native human macular RPE cells in a systematic fashion. In addition, we annotate the functions and biological pathways associated with RPE expressed (disease) genes.

To our knowledge this is the first study to present data on (interindividual differences in) hu- man macular RPE gene expression and interindividual differences on a large scale of 22,000 genes, resulting in a further detailed description of the RPE transcriptome.

Results RNA from six selected human macular RPE samples was hybridized to six custom made 22k microarrays enriched for neural transcripts. We functionally annotated and analyzed the data using Rosetta Resolver, the web tool DAVID and Ingenuity software, with regard to gene ex- pression level and variability as well as functional annotation. Furthermore, we specifically looked at the expression levels and variability of retinal disease genes.

Figure 1. Distribution of the mean intensity (μint) of all genes across percentile bins of 10 percent. 690,113 9000 *

8000

) 7000

6000

5000

4000

3000

average expression level (μ int 2000

1000

0 10 20 30 40 50 60 70 80 90 100

percentile Each bin contains 2,194 genes. Note that the mean expression level associated with the 100th percentile bin is not displayed fully in this graph due to the height of the expression exceeding the scope of the graph.

22 The human retinal pigment epithelium transcriptome

Analysis of gene expression levels (μint)

The mean expression intensities (μint) ranged from 73 to 690,113 (arbitrary units), (see Addi- tional file 1: Expression level and interindividual variation in all genes on the custom micro- array (http://www.ncbi.nlm.nih.gov/pmc/articles/PMC2679759/?tool=pubmed)). The distri- bution of μint across percentile bins of 10 percent of all genes is shown in Figure 1. We used th th th th the 90 , 50 and 10 percentile of the μint to categorize our data into groups with high (>90 ), moderate (50th-90th), low (10th-50th) and very low (<10th) expression. We focused our analysis on the biologically most relevant gene groups with high, moderate and low gene expression levels. These categories yielded 2,194 genes with high RPE expression, 8,776 genes with moderate expression and 8,776 genes with low expression. The results of the overrepresen- tation analysis are presented below, and in Table 1. The overrepresentation analysis of all expressed genes, irrespective of their gene expression level (Table1), did not yield additional functional categories apart from ECM-receptor interaction, and is not presented separately.

Table 1. Overrepresented Kegg pathways in macular RPE expressed genes with high, moderate and low expression levels and high or low levels of interindividual variability (coefficient of variation, CV).

expression level all expression high moderate low levels (>90th perc) (50th-90th perc) (10th-50th perc) all CV ecm-receptor interac- oxidative phosphory- phosphatidylinositol neuroactive ligand tion (E) lation (B,E) signaling system (B,E) receptor interaction ribosome (B,E) aminosugars metabo- (B,E) ATP synthesis (B,E) lism (E) long-term depression (E) o-glycan biosynthesis (E) calcium signaling pathway (E) high type I diabetes mel- antigen processing and focal adhesion (E) type I diabetes mel- CV litus (B,E) presentation (B,E) cytokine-cytokine re- litus (E) focal adhesion (B,E) complement and co- ceptor interaction (E) cytokine-cytokine re- agulation cascades (E) ceptor interaction (E) CV complement and co- agulation cascades (E) antigen processing and presentation (E) ecm-receptor interac- tion (E) low - glycosaminoglycan - - CV degradation (E)

Overrepresented pathways were identified with B: a Benjamini-Hochberg corrected p value<0.001, or E: an Ease score p value < 0.001. Perc: percentile, high CV: > 90th percentile, low CV: < 10th percentile.

23 Chapter 2

Genes with high expression levels (μint >90th percentile, n=2,194) We considered the group of highly expressed genes the most biologically relevant, and, con- sequently, for this group bioinformatic analysis was more extensive than for other categories. In addition to a Kegg pathway analysis, we also performed an Ingenuity analysis of the over- lap between our highly expressed genes and those identified in the literature. Kegg pathway analysis revealed oxidative phosphorylation, ribosome and ATP synthesis as significantly overrepresented pathways (Benjamini-Hochberg p value < 0.001) (Table 1). There was an overlap of 1,407 genes between the highly expressed genes of the RPE transcriptome and the genes identified in retina/RPE genes identified in at least two studies in the literature14 Ingenuity analysis of the overlapping genes revealed oxidative phosphorylation as the most significant pathway involved. Comparison of our highly expressed genes to those expressed only in RPE studies (n=17)[14], showed a clustering of genes in the cell-cell signaling and interaction network (Figure 2).

Figure 2. Ingenuity analysis of the cross section of genes previously identified in RPE studies[14] with genes highly expressed in the RPE transcriptome.

The resulting network shows a connectivity chart illustrating biological functions comprising genes, proteins and ligands related to cell-cell signaling and cell-cell interaction. This network contains 13 of the 16 genes entered into the analysis. Filled objects represent the genes entered into the analysis, empty objects are genes introduced by the ingenuity software creating a connection between the entered genes.

24 The human retinal pigment epithelium transcriptome

The thirty most highly expressed RPE genes from our data set are presented in Table 2. Most notably, this list contains two glutamate transporters (SLC1A2 [genbank:AF131756] and SLC17A7 [genbank: NM_020309])[15], one of which is known to be expressed in the RPE (SLC17A7 [genbank: NM_020309])[14] and a gene (CST3 [genbank: NM_000099]), that was previously suggested to have an association with AMD[16,17], with known expression in the RPE [18]. The top thirty list contained three additional genes with known expression in the RPE (PTGDS [genbank: NM_000954][17,19], TTR [genbank: NM_000371][14,20] and HSP90B1 [genbank: NM_003299][21]) and two genes that play a role in the protection against oxidative stress (MT1A [genbank: K01383][22], and TP53 [genbank: NM_000546] [23]). Finally, we identified a number of genes with a relevant cellular function described in other tissues than the retina, like CLU [NM_001831] (complement system) and ACN9 [NM_020186] (gluconeogenesis)[24,25].

Table 2. The thirty most highly expressed genes in macular RPE identified in six different human donors, sorted by intensity in descending order.

mean inten-

gene Genbank sity μint CV symbol ID perc perc gene name relevant function, ref KIAA0241 AA205569 99 98 KIAA0241 AI003379 AI003379 99 98 Transcribed locus ACN9 NM_020186 99 98 ACN9 homolog gluconeogenesis[25] MT1A BG191659 99 98 Metallothionein 1A protection against reactive oxygen species[22] SLC17A7 NM_020309 99 98 Solute carrier family 17 (so- glutamate transporter, ex- dium-dependent inorganic phosphate cotransporter), member 7 pressed in RPE[14,15] TP53 NM_000546 99 98 Tumor protein p53 protection against reactive oxygen species[23] ELF2 NM_006874 99 98 E74-like factor 2 (ets do- main transcription factor) AI272368 AI272368 99 95 cDNA clone AA807363 AA807363 99 79 cDNA clone SERPINB5 NM_002639 99 98 Serpin peptidase inhibitor, clade B, member 5 MCM7 NM_005916 99 97 minichromosome mainte- nance deficient 7 SLC1A2 NM_004171 99 96 Solute carrier family 1 glutamate transporter[1] member 2 HBB NM_000518 99 99 Hemoglobin, beta

25 Chapter 2

mean inten-

gene Genbank sity μint CV symbol ID perc perc gene name relevant function, ref PTGDS NM_000954 99 94 Prostaglandin D2 synthase released from RPE during 21kDa rod phagocytosis[17,19] T26536 T26536 99 75 cDNA clone EEF1A1 NM_001402 99 80 Eukaryotic translation elon- gation factor 1 alpha 1 TTR NM_000371 99 80 Transthyretin (prealbumin, maintains normal levels of retinol and retinol binding amyloidosis type I) proteins in plasma 14,20 BE260168 BE260168 99 90 cDNA clone CST3 NM_000099 99 92 Cystatin C associated with AMD 16,18 ZNF503 NM_032772 99 95 Zinc finger protein 503 BG190000 BG190000 99 94 cDNA clone BE262306 BE262306 99 80 cDNA clone HSP90B1 NM_003299 99 94 Heat shock protein 90kDa 21 beta (Grp94), member 1 GNGT1 NM_021955 99 90 Guanine nucleotide binding protein (G protein), gamma transducing activity poly- peptide 1 RPL3 NM_000967 99 20 Ribosomal protein L3 RPL41 NM_021104 99 23 Ribosomal protein L41 AI857840 AI857840 99 77 cDNA clone PCSK7 NM_004716 99 90 Proprotein convertase subtilisin/kexin type 7 AL521537 AL521537 99 75 cDNA clone CLU NM_001831 99 42 Clusterin member of complement system 24 Among these genes we identified five genes with known expression in the RPE (SLC17A7[14], CST3[18], TTR[14], HSP90B1[21] and PTGDS[17]). The SLC17A7 gene is a glutamate transporter, like the SLC1A2 gene which is also in the top 30 highly expressed genes. The CST3 gene was previously suggested to have an association with AMD[16,17]. The list also contains two genes with a role in the protection against oxidative stress (MT1A[22], TP53[23]). The GNGT1 gene, expressed in photoreceptors [30], suggests the inevitable presence of photoreceptor contamination. Perc: percentile. Ref: reference.

th th Genes with moderate expression levels (μint 50 -90 percentile) Upon analyzing this group of 8,776 genes, we found a statistically significant overrepresen- tation of the Kegg pathways phosphatidylinositol signaling and aminosugars metabolism (Benjamini-Hochberg p value < 0.001) (Table 1).

26 The human retinal pigment epithelium transcriptome

th th Genes with low expression levels (μint 10 -50 percentile) Among the 8,776 genes with low expression levels there was a statistically significant over- representation of the neuroactive ligand-receptor interaction (Benjamini-Hochberg p value 0.001), long-term depression, O-glycan biosynthesis and calcium signaling pathways (Ease score p value < 0.001) (Table 1).

Analysis of gene expression variability (CV) We analyzed the interindividual variability in gene expression (CV) among the 19,746 genes with expression levels in the RPE higher than the 10th percentile, (see Additional file 1: Ex- pression level and interindividual variation in all genes on the custom microarray (http:// www.ncbi.nlm.nih.gov/pmc/articles/PMC2679759/?tool=pubmed)). Aside from the overrep- resented cluster ECM-receptor interaction (Ease score p value < 0.001)(Table 1), this yielded little extra information compared to the CV assignment in subcategories of high, moderate and low expression levels (Table 1 and below), and is not presented in detail here. The thirty genes with the highest interindividual variation in expression levels in our dataset are pre- sented in Table 3.

Table 3. The top thirty genes with the highest interindividual variation in expression levels (CV) between six healthy human donors, sorted descending by coefficient of variation (CV).

mean inten-

gene Genbank sity μint CV symbol ID perc perc gene name HSD17B2 NM_002153 95 99 Hydroxysteroid (17-beta) dehydrogenase 2 MYOC NM_000261 99 99 Myocilin, trabecular meshwork inducible glucocorticoid response OGN NM_014057 92 99 Osteoglycin (osteoinductive factor, mimecan) SFRP4 NM_003014 93 99 Secreted frizzled-related protein 4 AOC2 NM_009590 91 99 Amine oxidase, copper containing 2 (retina-specific) DIO3 NM_001362 85 99 Deiodinase, iodothyronine, type III SLC2A5 NM_003039 62 99 Solute carrier family 2 (facilitated glucose/fructose transporter), member 5 XIST AK025198 99 99 X (inactive)-specific transcript TFPI2 NM_006528 99 99 Tissue factor pathway inhibitor 2 CYR61 NM_001554 98 99 Cysteine-rich, angiogenic inducer, 61 FGFBP2 NM_031950 84 99 Ksp37 protein FBP2 NM_003837 64 99 Fructose-1,6-bisphosphatase 2 EGFL6 NM_015507 74 99 EGF-like-domain, multiple 6 IL8 NM_000584 66 99 Interleukin 8 MFAP4 L38486 99 99 Microfibrillar-associated protein 4

27 Chapter 2

mean inten-

gene Genbank sity μint CV symbol ID perc perc gene name CCL2 NM_002982 90 99 Chemokine (C-C motif) ligand 2 ZIC1 NM_003412 67 99 Zinc family member 1 (odd-paired homolog, Drosophila) COL9A1 NM_001851 88 99 Collagen, type IX, alpha 1 CCL26 NM_006072 79 99 Chemokine (C-C motif) ligand 26 PITX2 NM_000325 89 99 Paired-like homeodomain transcription factor 2 ALDH1A1 NM_000689 78 99 Aldehyde dehydrogenase 1 family, member A1 HBG1 NM_000559 75 99 Hemoglobin, gamma A S100A6 NM_014624 99 99 S100 calcium binding protein A6 IL6 NM_000600 77 99 Interleukin 6 (interferon, beta 2) HBG2 NM_000184 71 99 Hemoglobin, gamma G C13orf33 NM_032849 93 99 Chromosome 13 open reading frame 33 RBM3 NM_006743 96 99 RNA binding motif (RNP1, RRM) protein 3 CFB NM_001710 94 99 Complement factor B EGR1 NM_001964 99 99 Early growth response 1 PTX3 NM_002852 97 99 Pentraxin-related gene, rapidly induced by IL-1 beta Note that although XIST, EGFL6 and RBM3 are x-chromosomal transcripts, their high interindividual variation could not be explained by the gender of the donors (data not shown). Perc: percentile.

Genes with high interindividual variability (CV>90th percentile)

Among the 390 genes with both a high CV and high μint there was an overrepresentation of genes involved in antigen processing as well as the complement and coagulation cascades.

The 824 genes with a high CV and moderate μint showed an overrepresentation of genes in- volved in focal adhesion and cytokine-cytokine receptor interaction, and the 762 genes with high CV and low μint showed an overrepresentation of genes involved in type I diabetes melli- tus. The latter group contains mainly major histocompatibility complex genes and interleukin 1α [genbank: NM_000575].

Table 4. The thirty genes with the least interindividual variation in macular RPE gene expression levels among six healthy human donors, sorted ascending by coefficient of variation (CV).

mean gene sym- Genbank intensity

bol ID μint perc CV perc gene name EXOC3 BC001511 85 <1 Exocyst complex component 3 PDXK AI571369 74 <1 Pyridoxal (pyridoxine, vitamin B6) kinase ACY1 NM_000666 63 <1 Aminoacylase 1

28 The human retinal pigment epithelium transcriptome

mean gene sym- Genbank intensity bol ID μint perc CV perc gene name SS18L1 AB014593 70 <1 Synovial sarcoma translocation gene on chromo- some 18-like 1 FOSL1 NM_005438 86 <1 FOS-like antigen 1 FPRL2 NM_002030 56 <1 Formyl peptide receptor-like 2 CPSF4 NM_006693 62 <1 Cleavage and polyadenylation specific factor 4, 30kDa FAM110B AK023658 40 <1 Chromosome 8 open reading frame 72 RAB20 AW861333 20 <1 Transcribed locus CHD2 AW896069 15 <1 Chromodomain helicase DNA binding protein 2 BI001591 BI001591 20 <1 Transcribed locus FATE1 NM_033085 37 <1 Fetal and adult testis expressed 1 CLEC4E NM_014358 44 <1 C-type lectin domain family 4, member E PARN NM_002582 75 <1 Poly(A)-specific ribonuclease (deadenylation nuclease) KIAA0586 NM_014749 59 <1 KIAA0586 TMEM156 NM_024943 41 <1 Transmembrane protein 156 CSNK2A1 NM_001895 59 <1 Casein kinase 2, alpha 1 polypeptide CGI-96 NM_015703 35 <1 CGI-96 protein LOC442100 BM127012 26 <1 Transcribed locus MYST3 NM_006766 93 <1 MYST histone acetyltransferase (monocytic leukemia) 3 ZMYND8 AF144233 33 <1 Protein kinase C binding protein 1 C20orf11 AK025775 88 <1 Chromosome 20 open reading frame 11 GAK NM_005255 89 <1 Cyclin G associated kinase SOBP NM_018013 67 <1 hypothetical protein FLJ10159 ZNF665 NM_024733 16 <1 Zinc finger protein 665 BG742052 BG742052 57 <1 cDNA clone OR2A7 AF327904 57 <1 Olfactory receptor, family 2, subfamily A, mem- ber 7 PRKCE NM_005400 18 <1 Protein kinase C, epsilon RNF14 AB022663 88 <1 Ring finger protein 14 SELENBP1 NM_003944 93 <1 Selenium binding protein 1

Perc: percentile.

29 Chapter 2

Genes with low interindividual variability (CV<10th percentile) Table 4 shows the thirty genes with the most stable expression in macular RPE. Among the th th expressed genes (μint >10 percentile) with stable RPE gene expression (CV<10 percentile, n=1,972) there were no genes overrepresented in Kegg pathways. One hundred and ninety four of these 1,972 genes had high expression levels, 1,064 had moderate expression levels and 714 had low expression levels. Using the DAVID software, a significant overrepresen- tation of genes in the glycosaminoglycan degradation pathway was found in the group of 194 genes with stable expression and high expression levels (Ease score p value < 0.001) (Table 1).

Table 5. Expression levels and interindividual differences of currently known macular disease genes with RPE expression[26].

mean intensity μint gene symbol Genbank accession (perc) CV (perc) high interindividual variation (CV>10th perc) CFB NM_001710 10,382 (94) 215 (99) C3 NM_000064 9,910 (94) 145 (98) FBLN5 NM_006329 2,823 (76) 136 (98) PRPH2 NM_000322 32,204 (98) 86 (93) GUCA1B NM_002098 2,329 (72) 83 (93) CST3 NM_000099 338 (26) 79 (92) CFH NM_000186 8,508 (92) 77 (91) TIMP3 NM_000362 7,209 (91) 73 (90) intermediate interindividual variation (CV 10th - 90th perc) C2 NM_000063 1,689 (65) 71 (90) BEST1 NM_004183 132,611 (99) 58 (84) HTRA1 NM_002775 45,420 (99) 47 (74) EFEMP1 NM_004105 20,292 (97) 44 (70) C1QTNF5 NM_015645 23,226 (98) 40 (64) TLR4 NM_003266 225 (45) 38 (60) low interindividual variation (CV<90th perc) none - - - Data are grouped by coefficient of variation (CV) in descending order into three groups, high CV (>90th percentile, CV>72), intermediate (CV 10th - 90th percentile) and low CV (<10th percentile, CV<19). Perc: percentile.

30 The human retinal pigment epithelium transcriptome

Gene expression analysis of known retinal disease genes Known macular disease genes We then investigated both the expression levels and interindividual expression differences of 14 macular disease genes in our RPE gene expression dataset (Table 5)26. In terms of expres- sion levels, 63 percent of the macular disease genes were found in the top 10 percent of genes with high macular RPE expression levels. In terms of variability, 50 percent of the macular disease genes were found in the top 10 percent of genes with highly variable macular RPE expression levels. In addition, none of the macular degeneration genes were found in the 10 percent of genes with stable macular RPE expression. A number of genes currently known or suggested to be associated with AMD, showed high (C3 [genbank: NM_000064], CFB [genbank: NM_001710], CFH [genbank: NM_000186], HTRA1 [genbank: NM_002775], and CST3 [genbank: NM_000099]) or moderate (FBLN5 [genbank: NM_006329]) expression levels in the RPE. With the exception of HTRA1 [gen- bank: NM_002775], all these genes also showed high interindividual variation.

Known peripheral retinal disease genes Finally, we analyzed the gene expression levels and interindividual differences in expression of 93 genes known to be involved in diseases of the peripheral retina[26] in our macular RPE expression dataset (Table 6). Of this group, 32 percent were found in the 10 percent of genes with high expression levels in the macular RPE. Eleven percent of the known peripheral disease genes were found in the 10 percent of genes with high interindividual variation in expression in the macular RPE.

Table 6. Expression levels and interindividual differences of currently known peripheral disease genes with RPE expression[26].

mean intensity (μint) gene symbol Genbank accession (percentile) CV (percentile) high interindividual variation (CV>90th percentile) COL9A1 NM_001851 5,628 (88) 225 (100) RBP4 NM_006744 20,291 (97) 196 (100) COL2A1 NM_001844 527 (37) 122 (97) GNAT1 NM_000172 100,271 (100) 80 (92) RDH5 NM_002905 6,716 (90) 78 (92) RLBP1 NM_000326 43,011 (99) 74 (91) intermediate interindividual variation (CV 10th - 90th percentile) PRCD AK054729 46,964 (99) 66 (88) GUCY2D NM_000180 2,672 (75) 64 (87) NPHP3 AI200954 3,266 (79) 63 (86) IMPDH1 NM_000883 12,177 (95) 63 (86) LRAT NM_004744 15,935 (96) 62 (86)

31 Chapter 2

mean intensity (μint) gene symbol Genbank accession (percentile) CV (percentile) LRP5 NM_002335 519 (37) 62 (86) RD3 AV721413 11,010 (95) 61 (86) RGR NM_002921 39,454 (99) 57 (83) TULP1 NM_003322 6,284 (89) 57 (83) AHI1 AL136797 7,168 (91) 56 (82) SEMA4A NM_022367 2,824 (76) 54 (80) TEAD1 AL133574 4,473 (84) 51 (78) PANK2 NM_024960 273 (20) 48 (75) PEX7 NM_000288 1,734 (65) 48 (75) OAT NM_000274 3,762 (81) 45 (71) CDH3 NM_001793 9,589 (93) 45 (71) BBS10 NM_024685 1,556 (63) 45 (71) PRPF8 NM_006445 6,510 (90) 40 (64) TIMM8A NM_004085 913 (50) 39 (63) FZD4 NM_012193 7,964 (92) 39 (62) OPA3 NM_025136 20,540 (97) 39 (62) COL11A1 NM_001854 911 (50) 38 (61) PGK1 NM_000291 12,114 (95) 37 (59) CYP4V2 AK022114 5,411 (87) 37 (58) JAG1 NM_000214 1,394 (60) 35 (53) MYO7A NM_000260 6,691 (90) 34 (53) BBS2 NM_031885 4,446 (84) 34 (51) PAX2 NM_003990 2,946 (77) 33 (50) BBS1 NM_024649 3,750 (81) 32 (46) NYX NM_022567 372 (28) 32 (45) ABCC6 NM_001171 1,555 (63) 28 (35) MERTK NM_006343 4,763 (85) 28 (35) ARL6 BI914103 1,023 (53) 28 (34) MFRP NM_031433 13,200 (96) 27 (33) PXMP3 NM_000318 5,341 (87) 25 (25) ALMS1 AB002326 962 (51) 23 (20) MKKS NM_018848 2,056 (69) 22 (16) NDP NM_000266 1,335 (59) 20 (14) PHYH NM_006214 3,293 (79) 20 (12) WFS1 NM_006005 4,935 (86) 19 (10)

32 The human retinal pigment epithelium transcriptome

mean intensity (μint) gene symbol Genbank accession (percentile) CV (percentile) low interindividual variation (CV<10th percentile) PRPF31 NM_015629 3,187 (78) 17 (06) PEX1 NM_000466 3,008 (77) 15 (04) PRPF3 NM_004698 7,057 (91) 14 (03) TRIM32 NM_012210 2,277 (71) 13 (02) CLN3 NM_000086 2,330 (72) 8 (0) Data are grouped by coefficient of variation (CV) in descending order into three groups, high CV (>90th percentile, CV>72), an intermediate group (CV 10th - 90th percentile) and CV low (<10th percentile, CV<19).

Discussion This study presents the first comprehensive analysis of the macular RPE transcriptome, with a focus on interindividual differences in RPE gene expression levels. We based our analyses on microarray data from six healthy human donor eyes. In addition, we performed a Kegg pathway analysis on genes with high, moderate and low expression levels and on genes with high and low interindividual variation in expression. Only five genes from our top 30 most highly expressed RPE genes were previously known to be expressed in the human RPE in vivo: SLC17A7 [genbank: NM_020309][14], CST3 [gen- bank: NM_000099])[18], PTGDS [genbank: NM_000954][17], TTR [genbank: NM_000371] [14] and HSP90B1 [genbank: NM_003299])[21] illustrating the lack of knowledge on the RPE transcriptome.

Strengths and limitations of the study design A recent statistical review suggested that a microarray study investigating a single tissue type, requires 6 biological replicate samples to draw statistically significant conclusions[27]. Consequently, we used the RPE gene expression from 6 different individuals. Previous RPE gene expression studies were based on less than six eyes, with the exception of a single cDNA microarray study limited to 4,325 genes that was based on 15 individuals[8-12]. Our study design has a number of strong points and limitations, previously described in de- tail[13]. In summary, the strength of our study design comes from our strict selection criteria for the donor eyes (see Figure 3), the use of a laser dissection microscope for high cellular specificity and minimal tissue manipulation, large scale analysis using a 22k microarray and a common reference design for comparison of all samples. Overall, our study was designed to minimize gene expression differences due to sampling methodology (see Figure 3) and technical causes, avoiding unnecessary mechanical handling of the freshly frozen tissue, the use of laser dissection microscopy to isolate homogeneous cell samples, stringent control of RNA quality and amplification procedures[9,11-13]. Initially, we performed dye swap experiments as technical replicates for three of our samples in order to ascertain the potential variability induced by dye bias. We observed a high correlation between the data from our analysis including and excluding the dye swap experiment (data not shown). At the same time, Dobbin (2003), Simon (2003) and others, used a similar study design as we did, and

33 Chapter 2 concluded that in a common reference design it is not necessary to perform dye swaps if the common reference is consistently labeled with the same dye[28,29]. Potential gene-specific dye bias will affect all experimental samples equally, and therefore does not confound the comparisons[27]. Consequently, we decided to perform the remaining three experiments without a dye swap.

Figure 3. Flow diagram of criteria used for the selection of donor eyes. no age 60-80 years X

yes yes eye disease X

no yes cancer X

no no post mortem time < 30 hours X

yes yes abnormalities: X visual inspection no yes abnormalities: histology X no yes drusen > 1 per 10 sections X no yes poor morphology X

no

laser dissection microscopy Donor eyes were required to meet all selection criteria before inclusion in the study. X indicates exclusion from the study. Donors were all between 60 and 80 years old in order to exclude the presence of undetected monogenic disorders. The presence of any known eye disease or malignancy was used as an exclusion criterion since both can alter (RPE) gene expression levels. Post mortem times were required to be less than 30 hours to reduce the effects of RNA degradation. Ocular abnormalities on visual or histological inspection served as exclusion criteria, specifi- cally any signs of early AMD, defined by us as the presence of more than 1 druse per 10 histological sections. Poor morphology of the retina was also an exclusion criterion.

One of the methodological limitations of our study was the limited number of eyes that met our selection criteria. The availability of a larger number of eyes would render more robust results with regard to interindividual variation. Nonetheless, our data give a good first im- pression of variability in gene expression levels in the RPE. An additional limitation is that a small amount of photoreceptor contamination was inevitably present in our RPE sample, see also Table 2[8,13,20]. Furthermore, we cannot distinguish possible transient from permanent gene expression level differences. Our study is also limited by the fact that the measurement

34 The human retinal pigment epithelium transcriptome of gene expression of individual genes by microarray is inevitably influenced by a number of factors, like oligo design and the continuous updates of the sequence. To correct for this last limitation, we focused our analysis on groups of genes with a wide range of expression levels, rather than on individual gene expression levels. Finally, our cut off criteria for high and low expression levels and interindividual differences are arbitrary. While this may indeed have consequences for individual genes, the impact on our functional analysis, which is based on large numbers of genes, will be minimal. Despite these limitations, our data, combined with data from other retinal gene expression studies (which use a range of techniques, like SAGE and RT-PCR, that bear their own limi- tations)[13,31], contributes significantly to the currently expanding knowledge of the RPE transcriptome.

Functional assessment of native gene expression in the macular RPE The notion that the identity of a cell type is determined by the genes it expresses, prompted us to analyze the native macular RPE transcriptome. In the following section we describe the overrepresented functional groups that we identified in the RPE.

Highly expressed RPE genes and oxidative stress Both functional annotation with DAVID and Ingenuity analysis independently indicate a statistically significant overrepresentation of genes associated with oxidative phosphoryla- tion and ATP synthesis in our dataset. This is in line with the fact that the RPE has a high metabolic activity and energy demand. The down side of this high activity is that the RPE has to deal with large amounts of oxidative stress. The oxidative stress in the RPE is further aug- mented by the light projected onto the retina combined with the rich oxygen supply and lipid peroxidation in phagocytosed rod outer segments[1,32]. Given the high level of oxidative stress, the expression of genes contributing to the defense of the RPE cell against oxidative stress is essential for cell survival. Our data confirm this notion, which is highlighted by the expression of the MT1A [genbank: K01383] gene, a metallothionein, and the TP53 [gen- bank: NM_000546] gene in the top 30 most highly expressed RPE genes. Metallothioneins are thought to play a role in protection against oxidative stress; addition of TP53 [genbank: NM_000546] to human cell lines leads to a 50 percent decrease in reactive oxygen spe- cies[22,23].

RPE and the immune system Our data show an overrepresentation of genes with highly variable expression in a number of pathways related to the immune system. We identified the following four pathways, the complement and coagulation cascades (high expression levels), the antigen processing and presentation pathway (high expression levels) and the cytokine-cytokine receptor interaction pathway (moderate expression levels). Both the antigen processing and presentation path- way and the type 1 diabetes mellitus pathway contain MHC genes responsible for antigen presentation. Cytokine production is highly sensitive to inflammation in the RPE[33]. The cytokine-cytokine receptor pathway contains a number of chemokines, small secreted pro- teins involved in the chemotaxic attraction of monocytes and neutrophils. The highly variable

35 Chapter 2 expression of genes involved in the immune system is most likely explained by both genetic differences and a variable degree of subclinical inflammation (local or systemic) among our donors.

RPE genes and the extracellular matrix (Bruch’s membrane) The close interaction of the RPE with Bruch’s membrane (BM) is exemplified by the over- representation of genes in two pathways. The first pathway contains genes involved in extra- cellular matrix (ECM) receptor interaction[13]. The ECM receptor interaction pathway, part of the focal adhesion pathway, contains collagens type I, III and IV, thrombospondin, lam- inin beta 1 [genbank: NM_002291], fibronectin 1 [genbank: NM_002291], reelin [genbank: NM_005045], and cd44 antigen [genbank: NM_000610]. Collagen type IV, laminin and fi- bronectin are all main components of basement membranes, such as BM. Surprisingly, the genes in this group showed highly variable expression, which may indicate that the molecular composition of BM is different among individuals. Alternatively, it has been described that with age, the solubility of collagens in BM decreases significantly[34]. Thus, the high vari- ability in expression levels of collagen genes between our samples can perhaps be explained by differences in the physiological donor age. A second pathway that connects RPE expressed genes to BM is the glycosaminoglycan (GAG) degradation pathway. There was an overrepresentation of genes with stable and high expression in this pathway. GAG synthesis has been shown in cultured RPE and GAG’s are secreted into the extracellular matrix and BM[35]. Interestingly, GAG’s are rapidly turned over in the RPE, and the composition of GAG in BM changes with age[35-37]. Our data sug- gest there is a strict regulation of GAG turnover in the RPE, even in donors of different ages.

Additional RPE gene functions In addition to the involvement of the RPE genes in oxidative stress, BM and the immune sys- tem, analysis of our data revealed the following two functional categories: protein synthesis and glutamate transport. A high level of protein synthesis is essential for the RPE to maintain its multiple functions[1]. This is exemplified by the overrepresentation of genes with high expression in the ribosomal protein activity pathway. Glutamate transport is an important process in the RPE. The top 30 most highly expressed RPE genes contained two glutamate transporters SLC1A2 [genbank: AF131756] and SL- C17A7 [genbank: AF131756]. The latter transporter was already known to be expressed in the human RPE in vivo[14,38]. Glutamate is an important neurotransmitter that is released from the photoreceptors both in a light influenced fashion, and upon apoptosis. Since high concentrations of glutamate are neurotoxic, re-uptake and transport of glutamate are essential for the normal retinal homeostasis[38]. Finally, in the overlap between previous RPE studies[14] and genes with high expression in our RPE transcriptome, we identified the cell-cell signaling and interaction network. This network contains several genes involved in signal transduction, like SLC7A2 [genbank: AL512749] and NCK2 [genbank: BC007195] further emphasizing the important role of the RPE in interaction with other cell types[39,40].

36 The human retinal pigment epithelium transcriptome

Comparison with literature Comparison of our most highly expressed RPE genes to the literature revealed a distinct overlap. Schulz and coworkers recently combined different analyses of the retina/RPE/cho- roid transcriptome, and described 13,000 retina/RPE genes found in at least two studies[14]. Out of these 13,000, we currently assign 7,231 genes to be expressed by the RPE, 1,407 of which are highly expressed. (see Additional file 2: overlap between highly expressed RPE genes and retina/RPE genes in at least two studies (http://www.ncbi.nlm.nih.gov/pmc/ar- ticles/PMC2679759/?tool=pubmed)) In addition, the same review[14] suggested that 246 genes were expressed only in RPE studies. We also assign 137 of these 246 genes to the RPE; 17 out of these 137 genes have high expression levels in our RPE transcriptome analysis (see Additional file 3: overlap between highly expressed RPE genes and genes found only in RPE studies (http://www.ncbi.nlm.nih.gov/pmc/articles/PMC2679759/?tool=pubmed)). Finally, of the genes previously described to be specifically expressed either in the retina or the RPE in individual studies[14], 39 genes are also present in our RPE transcriptome analy- sis. Twenty two of these 39 genes had high expression levels (see Additional file 4: overlap between highly expressed RPE genes and retina/RPE genes in single studies (http://www. ncbi.nlm.nih.gov/pmc/articles/PMC2679759/?tool=pubmed)).

While data on interindividual variation in RPE gene expression are lacking in the litera- ture, functional properties of RPE genes have been investigated previously. The combined functional annotation from three studies resemble our functional annotation in the following areas: gene regulation, transcription, protein metabolism, cell proliferation, survival and sig- naling, energy metabolism, cytoskeleton and inflammation[8,10,11]. The current study adds the following more specific functional categories, oxidative phosphorylation, ATP synthesis, ribosome, phosphatidylinositol signaling and aminosugars metabolism. Among the highly expressed RPE genes we identified an overrepresentation of the complement cascade and genes involved in the composition of BM.

Gene expression analysis of known retinal disease genes In our macular RPE sample we observed that 63 percent of genes involved in macular disor- ders according to the literature[26], had high expression levels. In contrast, only 32 percent of the peripheral retinal disease genes[26] were highly expressed in our sample. These figures may be biased, since the search for candidate genes has been focused on cell-specific highly expressed genes in the first place. The figures probably reflect the fact that RPE gene expres- sion differences exist between the retinal macula and the periphery[13]. However, our data probably also imply that the mean expression level of a gene in the RPE is informative in the search for new candidate disease genes. With respect to the variability in gene expression, we found that the interindividual differ- ences of currently known macular retinal disease genes were somewhat higher than the over- all pattern of variation seen in the entire array. Whether or not this finding is coincidental remains to be elucidated.

37 Chapter 2

Conclusions In conclusion, we present comprehensive data on (interindividual differences of the) gene ex- pression profile of the RPE based on 22,000 genes from six different healthy human donors. This is the first study to describe the interindividual variability in gene expression levels from a microarray analysis of the RPE transcriptome.

There was no correlation between the height of gene expression (μint) and the interindividual variability (CV) (data not shown). We noted a more than hundred fold difference in CV be- tween genes with stable expression and genes with variable expression levels. Our data show that the RPE most likely has high levels of protein synthesis, a high energy demand and is subject to high levels of oxidative stress as well as a variable degree of inflam- mation. Finally, our data show high interindividual variability in expression of ECM genes and indicate a high and constant level of glycosaminoglycan (GAG) turnover, two functions related to BM. The fact that large interindividual differences exist in the expression of a number of known retinal disease genes has not only functional implications, but is also relevant for new can- didate disease gene identification and the development of dose-dependent (gene) therapeutic strategies.

Methods Human donor eyes This study was performed in agreement with the declaration of Helsinki on the use of hu- man material for research. Material used in this study was provided to us by the Corneabank Amsterdam. In order to minimize genetic heterogeneity, we selected six eyes from a total of 200 human donor eyes using strict selection criteria, (see Figure 3). In summary, donors were excluded when their age was not between 60 and 80 years, when they had an eye disease or any form of malignancy and when the time between death and enucleation of the eye was more than 30 hours. Furthermore, eyes were excluded when they showed any abnormalities upon visual or histological examination: more specifically, when more than one druse was seen in 10 histological sections, or when retinal morphology was poor. All donors were Cau- casian, five were male, one was female. The donors died of cardiovascular or cerebrovascular causes or of chronic obstructive pulmonary disease. Donors did not have a known ophthalmic disorder or malignancy. Globes were enucleated between 14 and 27 hours post mortem and frozen several hours later according to a standard protocol. Donors were aged 63 to 78 years at the time of death. We chose old donors in order to minimize the likelihood of the presence of yet undiagnosed monogenic eye diseases. This does not rule out the presence of the most common retinal disease in the old eye, age related macular degeneration (AMD). Therefore the donor retinas were thoroughly screened for early signs of AMD by histological examina- tion (the presence of more than 1 druse in 10 sections). Visual examination and histological examination, including periodic acid Schiff (PAS) staining, indicated no retinal pathology in any of the donor eyes.

38 The human retinal pigment epithelium transcriptome

RPE cell sampling Globes were snap-frozen and stored at -80°C until use. A macular fragment of 16 mm2 with the fovea in its center was cut from each of the retinas, as described previously[13]. In sum- mary, for each eye, 10 cryosections, 8 μm thick, spaced no more than 220 μm apart were stained with periodic-acid Schiff and microscopically examined for abnormalities, such as drusen indicative of early-AMD. Twenty μm sections from the macular areas were used for the isolation of RPE cells. These sections were dehydrated with ethanol and air-dried before microdissection with a Laser Microdissection System (PALM, Bernried, Germany) using a pulsed laser. A total of up to 10,000 RPE cells per eye were microdissected and stored at -80° Celsius.

RNA isolation and (single) amplification Total RNA was isolated and the mRNA component was amplified essentially as described previously[13]. Next, the amplified RNA (aRNA) samples were quantified with a nanodrop (Isogen Life Science B.V., The Netherlands) and the quality was checked on a BioAnalyzer (Agilent Technologies, Amstelveen, The Netherlands). Subsequently, aRNA samples were labeled with either a Cy3 or a Cy5 fluorescent probe.

Microarray handling A common reference design was applied in our microarray hybridizations using the common reference sample described in the study of van Soest et al (2007)[13]. In summary, the com- mon reference sample consists of aRNA from a pool of RPE/choroid isolated from 10 donor eyes (mean age 60 years). aRNA from all six donors and the common reference sample was labeled. Subsequently, labeled aRNA from the donors was hybridized against the common reference sample to six 22k custom arrays. Initially, a dye swap experiment was performed for three of the six donor samples in order to assess potential variability introduced by dye- bias for methodological reasons (see discussion). Dye swaps were disregarded in the final analysis. Arrays were enriched for sequences expressed in RPE, neural retina and brain (Agi- lent Technologies, Amstelveen, The Netherlands), (see Additional file 1: Expression level and interindividual variation in all genes on the custom microarray (http://www.ncbi.nlm.nih. gov/pmc/articles/PMC2679759/?tool=pubmed)). Hybridization, washing and scanning were performed as described previously[13].

Data analysis Scanned images were processed with Feature Extraction software (v 8.5 Agilent). Data from all six hybridizations was analyzed with Rosetta Resolver software (Rosetta Inpharmatics). The signal of each of the six RPE samples was normalized using the common reference sample. This enabled a direct comparison of the six RPE samples (Figure 4). We used six biological replicates in order to draw significant conclusions[27]. For each gene we calcu- lated the mean signal intensity (μint) and standard deviation (σ) of the six biological replicates. While a limited number of genes is present on the array more than once, for the analyses of large groups of genes we regarded the number of entries on the array equal to the number of genes. Genes were grouped according to their mean intensity (μint). We defined μint above the

39 Chapter 2

th th th 90 percentile as high expression, μint between the 90 and 50 percentile as moderate expres- th th sion and μint between the 50 and the 10 percentile as low expression. We considered the genes in these three groups to have potential biological significance. Genes with a intμ below the 10th percentile were considered to have very low expression with a doubtful biological significance.

Figure 4. Study design: A. Experimental setup. RPE sample common microarray reference pool experiment 1 2

1 + experiment 1

2 + experimentexperiment 2 2 common 3 + experimentexperiment 3 3 6 common referencereference 3 4 + experimentexperiment 4 4

5 + experimentexperiment 5 5

6 + experimentexperiment 6 6 5 4 Six RPE samples from 6 different donors were hybridized to six microarrays along with the common reference sample. B. Data analysis. The common reference was used to normalize the RPE expression data from the six ar- rays which enabled comparison of the six individuals.

In order to describe the interindividual differences in gene expression levels between all six eyes systematically, we calculated the coefficient of variation (CV), defined as the standard deviation divided by the mean (σ/ μint), for each gene. We considered genes with a CV above the 90th percentile to have “high” interindividual variation in expression and genes with a CV below the 10th percentile to have “low” interindividual variation, or stable expression. Obviously, the categories for intensity and variability of expression were chosen somewhat arbitrarily, but they were essential to facilitate systematic analysis and to minimize the num- ber of false positive results. A functional analysis of Kegg pathways (Kyoto Encyclopedia of Genes and Genomes) was performed on genes with high, moderate and low expression levels and on genes with high and low interindividual variation using the DAVID online software[41]. Cut off criteria used were a p-value of less than 0.001 using either a Benjamini-Hochberg correction or an Ease score, which is a modified Fisher’s exact test[41,42]. We compared our RPE transcriptome to a compilation of the mammalian retina/RPE tran- scriptome, which is based on multiple independent gene expression studies of combinations of the neural retina/RPE/choroid in the literature[14]. Overlap between the two datasets was analyzed using Ingenuity Pathways Analysis (Ingenuity® Systems) resulting in a connectiv- ity network describing the underlying biology of RPE cells at the genomic and proteomic level[43].

40 The human retinal pigment epithelium transcriptome

References 1. Strauss O. The retinal pigment epithelium in visual function. Physiol Rev 2005; 85: 845-881. 2. van Soest S, Westerveld A, de Jong PT, Bleeker-Wagemakers EM, Bergen AA. Retinitis pigmentosa: defined from a molecular point of view. Surv Ophthalmol 1999; 43: 321- 334. 3. Hageman GS, Zhu XL, Waheed A, Sly WS. Localization of carbonic anhydrase IV in a specific capillary bed of the human eye. Proc Natl Acad Sci U S A 1991; 88: 2716-2720. 4. Sugiura M, Kono K, Liu H et al. Ceramide kinase, a novel lipid kinase. Molecular clon- ing and functional characterization. J Biol Chem 2002; 277: 23294-23300. 5. Surguchov A, Bronson JD, Banerjee P et al. The human GCAP1 and GCAP2 genes are arranged in a tail-to-tail array on the short arm of chromosome 6 (p21.1). Genomics 1997; 39: 312-322. 6. Rice DS, Huang W, Jones HA et al. Severe retinal degeneration associated with disrup- tion of semaphorin 4A. Invest Ophthalmol Vis Sci 2004; 45: 2767-2777. 7. Wang Q, Chen Q, Zhao K et al. Update on the molecular genetics of retinitis pigmen- tosa. Ophthalmic Genet 2001; 22: 133-154. 8. Ishibashi K, Tian J, Handa JT. Similarity of mRNA phenotypes of morphologically nor- mal macular and peripheral retinal pigment epithelial cells in older human eyes. Invest Ophthalmol Vis Sci 2004; 45: 3291-3301. 9. Kociok N, Joussen AM. Varied expression of functionally important genes of RPE and choroid in the macula and in the periphery of normal human eyes. Graefes Arch Clin Exp Ophthalmol 2007; 245: 101-113. 10. Buraczynska M, Mears AJ, Zareparsi S et al. Gene expression profile of native human retinal pigment epithelium. Invest Ophthalmol Vis Sci 2002; 43: 603-607. 11. Sharon D, Blackshaw S, Cepko CL, Dryja TP. Profile of the genes expressed in the hu- man peripheral retina, macula, and retinal pigment epithelium determined through se- rial analysis of gene expression (SAGE). Proc Natl Acad Sci U S A 2002; 99: 315-320. 12. Bowes Rickman C., Ebright JN, Zavodni ZJ et al. Defining the human macula transcrip- tome and candidate retinal disease genes using EyeSAGE. Invest Ophthalmol Vis Sci 2006; 47: 2305-2316. 13. van Soest S, de Wit GM, Essing AH et al. Comparison of human retinal pigment epithe- lium gene expression in macula and periphery highlights potential topographic differ- ences in Bruch’s membrane. Mol Vis 2007; 13: 1608-1617. 14. Schulz HL, Goetz T, Kaschkoetoe J, Weber BH. The Retinome - defining a reference transcriptome of the adult mammalian retina/retinal pigment epithelium. BMC Genom- ics 2004; 5: 50. 15. Bellocchio EE, Reimer RJ, Fremeau RT, Jr., Edwards RH. Uptake of glutamate into synaptic vesicles by an inorganic phosphate transporter. Science 2000; 289: 957-960. 16. Zurdel J, Finckh U, Menzer G, Nitsch RM, Richard G. CST3 genotype associated with exudative age related macular degeneration. Br J Ophthalmol 2002; 86: 214-219.

41 Chapter 2

17. Beuckmann CT, Gordon WC, Kanaoka Y et al. Lipocalin-type prostaglandin D synthase (beta-trace) is located in pigment epithelial cells of rat retina and accumulates within interphotoreceptor matrix. J Neurosci 1996; 16: 6119-6124. 18. Wasselius J, Hakansson K, Johansson K, Abrahamson M, Ehinger B. Identification and localization of retinal cystatin C. Invest Ophthalmol Vis Sci 2001; 42: 1901-1906. 19. Ershov AV, Parkins N, Lukiw WJ, Bazan NG. Modulation of early response gene ex- pression by prostaglandins in cultured rat retinal pigment epithelium cells. Curr Eye Res 2000; 21: 968-974. 20. Episkopou V, Maeda S, Nishiguchi S et al. Disruption of the transthyretin gene results in mice with depressed levels of plasma retinol and thyroid hormone. Proc Natl Acad Sci U S A 1993; 90: 2375-2379. 21. Wu WC, Kao YH, Hu PS, Chen JH. Geldanamycin, a HSP90 inhibitor, attenuates the hypoxia-induced vascular endothelial growth factor expression in retinal pigment epi- thelium cells in vitro. Exp Eye Res 2007; 85: 721-731. 22. Bremner I, Beattie JH. Metallothionein and the trace minerals. Annu Rev Nutr 1990; 10: 63-83. 23. Sablina AA, Budanov AV, Ilyinskaya GV et al. The antioxidant function of the p53 tumor suppressor. Nat Med 2005; 11: 1306-1313. 24. Trougakos IP, Gonos ES. Clusterin/apolipoprotein J in human aging and cancer. Int J Biochem Cell Biol 2002; 34: 1430-1448. 25. http://www.genecards.org/index.shtml 26. www.sph.uth.tmc.edu/Retnet/ 27. Lee NH, Saeed AI. Microarrays: an overview. Methods Mol Biol 2007; 353: 265-300. 28. Dobbin K, Shih JH, Simon R. Statistical design of reverse dye microarrays. Bioinfor- matics 2003; 19: 803-810. 29. Simon RM, Dobbin K. Experimental design of DNA microarray experiments. Biotech- niques 2003; Suppl: 16-21. 30. Peng YW, Robishaw JD, Levine MA, Yau KW. Retinal rods and cones have distinct G protein beta and gamma subunits. Proc Natl Acad Sci U S A 1992; 89: 10882-10886. 31. Skern R, Frost P, Nilsen F. Relative transcript quantification by quantitative PCR: roughly right or precisely wrong? BMC Mol Biol 2005; 6: 10. 32. Cai J, Nelson KC, Wu M, Sternberg P, Jr., Jones DP. Oxidative damage and protection of the RPE. Prog Retin Eye Res 2000; 19: 205-221. 33. Holtkamp GM, Kijlstra A, Peek R, de Vos AF. Retinal pigment epithelium-immune sys- tem interactions: cytokine production and cytokine-induced changes. Prog Retin Eye Res 2001; 20: 29-48. 34. Karwatowski WS, Jeffries TE, Duance VC et al. Preparation of Bruch’s membrane and analysis of the age-related changes in the structural collagens. Br J Ophthalmol 1995; 79: 944-952. 35. Stramm LE, Haskins ME, Aguirre GD. Retinal pigment epithelial glycosaminoglycan metabolism: intracellular versus extracellular pathways. In vitro studies in normal and diseased cells. Invest Ophthalmol Vis Sci 1989; 30: 2118-2131. 36. Kliffen M, Mooy CM, Luider TM et al. Identification of glycosaminoglycans in age- related macular deposits. Arch Ophthalmol 1996; 114: 1009-1014.

42 The human retinal pigment epithelium transcriptome

37. Tate DJ, Jr., Oliver PD, Miceli MV et al. Age-dependent change in the hyaluronic acid content of the human chorioretinal complex. Arch Ophthalmol 1993; 111: 963-967. 38. Miyamoto Y, Del Monte MA. Na(+)-dependent glutamate transporter in human retinal pigment epithelial cells. Invest Ophthalmol Vis Sci 1994; 35: 3589-3598. 39. Hoshide R, Ikeda Y, Karashima S et al. Molecular cloning, tissue distribution, and chro- mosomal localization of human cationic amino acid transporter 2 (HCAT2). Genomics 1996; 38: 174-178. 40. Chen M, She H, Davis EM et al. Identification of Nck family genes, chromosomal lo- calization, expression, and signaling specificity. J Biol Chem 1998; 273: 25171-25178. 41. Dennis G, Jr., Sherman BT, Hosack DA et al. DAVID: Database for Annotation, Visual- ization, and Integrated Discovery. Genome Biol 2003; 4: 3. 42. Hosack DA, Dennis G, Jr., Sherman BT, Lane HC, Lempicki RA. Identifying biological themes within lists of genes with EASE. Genome Biol 2003; 4: R70. 43. www.ingenuity.com

Acknowledgements We thank Dr. L. Pels and co-workers of the Corneabank, Amsterdam, for providing donor eyes. This study was supported by grants from the Foundation Fighting Blindness (T-GE- 0101-0172), The Netherlands Organization for Scientific Research (NWO; project 948-00- 013), The Royal Netherlands Academy of Arts and Sciences (KNAW), de Algemene 27 Nederlandse Vereniging ter Voorkoming van Blindheid (ANVVB) and het Edward en Marianne Blaauwfonds van het Amsterdams Universiteitsfonds.

43

A new strategy to identify and an- notate human RPE-specific gene expression

3

Judith C Booij, Jacoline B ten Brink, Sigrid MA Swagemakers, Annemieke JMH Verkerk, Anke HW Essing, Peter J van der Spek, Arthur AB Bergen.

PlosOne 2010, 5(3) e9341 Chapter 3

Abstract Background: The aim of the study was to identify and functionally annotate cell type-spe- cific gene expression in the human retinal pigment epithelium (RPE), a key tissue involved in age-related macular degeneration and retinitis pigmentosa.

Methodology: RPE, photoreceptor and choroidal cells were isolated from selected freshly frozen healthy human donor eyes using laser microdissection. RNA isolation, amplification and hybridization to 44k microarrays was carried out according to Agilent specifications. Bioinformatics was carried out using Rosetta Resolver, David and Ingenuity software.

Principal Findings: Our previous 22k analysis of the RPE transcriptome showed that the RPE has high levels of protein synthesis, strong energy demands, is exposed to high levels of oxidative stress and a variable degree of inflammation. We currently use a complementary new strategy aimed at the identification and functional annotation of RPE-specific expressed transcripts. This strategy takes advantage of the multilayered cellular structure of the retina and overcomes a number of limitations of previous studies. In triplicate, we compared the transcriptomes of RPE, photoreceptor and choroidal cells and we deduced RPE specific ex- pression. We identified at least 114 entries with RPE-specific gene expression. Thirty-nine of these 114 genes also show high expression in the RPE, comparison with the literature showed that 85% of these 39 were previously identified to be expressed in the RPE. In the group of 114 RPE specific genes there was an overrepresentation of genes involved in (membrane) transport, vision and ophthalmic disease. More fundamentally, we found RPE-specific in- volvement in the RAR-activation, retinol metabolism and GABA receptor signaling path- ways.

Conclusions: In this study we provide a further specification and understanding of the RPE transcriptome by identifying and analyzing genes that are specifically expressed in the RPE.

46 Human RPE-specific gene expression

Introduction The retinal pigment epithelium (RPE) is a monocellular retinal layer that plays a particularly important role in visual function. This is illustrated by its involvement in a large number of severe retinal disorders like age-related macular degeneration and retinitis pigmentosa. The RPE has multiple functions including supplying the photoreceptors with nutrients, recy- cling retinal from the photoreceptors and regulating the ion balance in the subretinal space. The RPE secretes a number of growth factors. Thereby it is involved in the maintenance of the structure and cellular differentiation of adjacent (cell) layers, the photoreceptors on the apical side, and Bruch’s membrane and the choroid on the basolateral side (See Figure 1)[1].

Figure 1. Overview of the (cell) layers surrounding the RPE.

Phot

RPE

Bruch's membrane

Choroid

Choroidal bloodvessel The solid rectangles connecting the RPE cells indicate tight junctions present between the cells. Phot: photorecep- tors. A color version of this figure can be found in chapter 15.

Embryologically, both the RPE and the photoreceptors develop from the neuro-epithelium. A fold in the neuro-epithelium causes two layers of this structure to face each other[1]. One of these layers develops into the RPE, the other into the neural retina. The neural retina fur- ther differentiates and photoreceptors develop in close interaction with the microvilli of the RPE[1]. The development of the choroid from neural crest cells also depends on cues from the mature RPE[2]. The choroidal vasculature is created through angiogenesis from existing blood vessels from the paraocular mesenchyme[2]. Finally, Bruch’s membrane (BM) is cre- ated from the basement membrane of the RPE and the basement membrane of the endotheli- um. In this way, the retina develops into a neatly arranged multi-layered structure (Figure 1).

We recently analyzed the RPE transcriptome[3]. We reported on the expressed genes and correlated molecular pathways in the RPE from cells that were specifically isolated from healthy human donor eyes[3]. Functional annotation of the RPE transcriptome showed that the RPE has high levels of protein synthesis, strong energy demands, is exposed to high

47 Chapter 3 levels of oxidative stress and has a variable degree of inflammation[3]. These data confirmed and expanded our knowledge on functional properties of the RPE, previously identified in a number of studies[4-7]. Nonetheless, due to the study design, our previous study was limited by unavoidable contamination from adjacent cell types[3,5,8].

In the current study we have overcome this limitation by using a new 44k microarray strat- egy that includes both the RPE cell layer and the adjacent cell layers in the experiment. We compared the transcriptomes of the photoreceptor, RPE and choroidal cells using a single platform. We deduced that at least 114 genes are specifically expressed in the RPE and we describe their corresponding pathways.

Results We performed microarray analyses on RNA specifically isolated from RPE, photorecep- tors and choroid cells from healthy human donor eyes. Recently, we defined the RPE tran- scriptome by measuring gene expression levels and identifying functional properties of the RPE[3]. The current study provides a further specification of the RPE transcriptome by iden- tifying genes expressed at much higher levels in the RPE than in either adjacent cell layer, the photoreceptors and the choroid. In addition, we expand our dataset from a 22k microarray platform to a 44k microarray platform, resulting in a more extensive coverage of the human genome[9], and we used more advanced bioinformatics to analyze the data.

RPE gene expression compared to either photoreceptors or choroid Of the 33,712 features present on the microarray (GSE20191, see supplementary file S1, available on request), 1,904 (5.6%) genes had at least 2.5-fold higher expression in the RPE than in the photoreceptors (RPE>phot) on average over three arrays (see Table 1 and supple- mentary file S2, available on request). There was a significant overrepresentation of genes that encode signal proteins, glycoproteins, secreted proteins, membrane proteins, cell adhe- sion proteins, extracellular matrix proteins, proteins involved in the immune response, Ca2+- binding and actin binding (David [10]). Functional analysis of these genes revealed an over- representation of genes in the following pathways: cell adhesion molecules, melanogenesis and type I diabetes mellitus. Furthermore, 1,126 (3.3%) of 33,712 genes on the array had at least 2.5-fold higher expres- sion in the RPE than in the choroid (RPE>chor) on average over three arrays (see Table 2 and supplementary file S3, available on request). In this group of genes there was a significant overrepresentation of genes coding for proteins involved in vision, retinitis pigmentosa and (membrane) transport, as well as genes with a symport function and genes in the olfactory transduction pathway.

48 Human RPE-specific gene expression

Table 1 Top 30 genes with highest expression in RPE compared to photoreceptors [26,29].

gene symbol Genbank ID FC RPE/phot ITGB8 BC042028 44 C1orf168 AK093468 41 COL8A1 AL359062 37 SLC26A4 NM_000441 37 SLC26A7 NM_052832 37 DKFZp761G0122 AL713743 37 CLIC6 NM_053277 34 VASN NM_138440 33 SMOC2 NM_022138 31 SLC16A12 AK124901 29 C6orf105 NM_032744 28 SLC6A13 BC020867 28 PRDM16 NM_022114 28 LGI1 NM_005097 27 PLD5 NM_152666 26 A_32_P114831 A_32_P114831 26 KCNS3 NM_002252 25 BEST1 NM_004183 23 IL8 NM_000584 23 TMEM27 NM_020665 23 NTN4 NM_021229 23 RWDD3 AK126344 23 LRP8 NM_004631 23 SLCO1C1 NM_017435 23 COL8A2 NM_005202 22 MYRIP NM_015460 22 GPNMB NM_002510 20 PLA2G7 NM_005084 20 KCNJ13 NM_002242 19 FAM40B AB032996 19 FC: fold change, average expression level in RPE compared to photoreceptors in three arrays.

49 Chapter 3

Table 2 Top 30 genes with highest expression in RPE compared to choroid[26,29].

gene symbol Genbank ID FC RPE/chor RBP3 NM_002900 37 RPE65 NM_000329 24 MPP4 NM_033066 24 ELOVL4 NM_022726 23 GUCA1C NM_005459 23 PDE6G NM_002602 22 NEUROD1 NM_002500 20 BEST1 NM_004183 20 SLC6A13 BC020867 19 ITGB8 BC042028 19 RP1 NM_006269 19 GUCA1B BX537393 19 PDC NM_002597 19 CNGB3 NM_019098 18 PROM1 NM_006017 18 PRPH2 NM_000322 18 HCN1 AK094523 18 DKFZp761G0122 AL713743 17 HOOK1 AK027250 17 GNAT2 NM_005272 16 RLBP1 NM_000326 16 C6orf105 NM_032744 16 CNGA1 NM_000087 16 ABCA4 NM_000350 16 TMEM27 NM_020665 16 RWDD3 AK126344 16 OPCML BX537377 15 NRL NM_006177 15 C1orf168 AK093468 15 TMEM16B NM_020373 15 Note that cellular contamination of the RPE cells may be present, identified by the ABCA4 transcript, which is truly a photoreceptor-specific transcript[30]. FC: fold change, average expression level in RPE compared to cho- roid in three arrays.

50 Human RPE-specific gene expression

A novel strategy to detect RPE-specific gene expression Until now, all RPE transcriptome studies, including our own[3], were aimed at excluding the cell layers adjacent to the RPE from the experiment in order to prevent cellular contamination and to achieve the highest RPE tissue specificity possible[3,4]. However even isolation of RPE cells by meticulous laser dissection microscopy resulted in unavoidable contamination with adjacent cell types[5,8] (this study). In the current study however, we did not discard the adjacent cell layers as possible contami- nants, but we included them as valuable resources for comparison of gene expression. Our primary objective was to further specify the expression level of genes in the RPE relative to their expression in the photoreceptors and the choroid. In this way, we deduced RPE-specific gene expression.

RPE-specific gene expression As a first analysis we identified all genes with average expression levels at least 2.5-fold higher in the RPE than in both photoreceptors and choroid. This resulted in a list of 458 entries with RPE-specific expression (see supplementary file S4, available on request). Func- tional annotation showed an overrepresentation of genes involved in inositol metabolism, retinol metabolism, genetic disorders and ophthalmic diseases (data not shown). In order to illustrate the usefulness of our approach to identify RPE specific transcripts, we next employed even stricter criteria (i.e. in all six arrays at least 2.5-fold higher expression levels in the RPE than in both photoreceptors and choroid (RPE>phot&chor FC >2.5)). This yielded 114 genes (see Table 3). We deduced from our recently published data set[3] that 39 out of these 114 genes are very highly expressed in the RPE (see Table 4).

Table 3. 114 genes with at least 2.5 fold higher RPE expression in the RPE than in both the photoreceptors and the choroid in all six microarrays, defined as RPE-specific expression.

gene symbol Genbank ID FC gene symbol Genbank ID FC ITGB8 BC042028 31 CCNO NM_021147 8 C1orf168 AK093468 28 RDH10 NM_172037 8 DKFZp761G0122 AL713743 27 PBX4 NM_025245 8 SLC6A13 BC020867 24 SKIP NM_130766 8 C6orf105 NM_032744 22 SLC7A10 NM_019849 8 CLIC6 NM_053277 22 CXCL14 NM_004887 8 BEST1 NM_004183 21 A_24_P234871 A_24_P234871 8 RPE65 NM_000329 20 COL20A1 NM_020882 8 PLD5 NM_152666 19 LHX2 NM_004789 8 TMEM27 NM_020665 19 C1QTNF5 NM_015645 7 RWDD3 AK126344 19 SLC22A8 NM_004254 7 LRP8 NM_004631 19 ROBO2 AK074780 7 LGI1 NM_005097 18 THC1970019 THC1970019 7

51 Chapter 3

gene symbol Genbank ID FC gene symbol Genbank ID FC SLC16A12 AK124901 17 ADAD2 NM_139174 7 FAM40B AB032996 17 OR51E2 NM_030774 7 PRDM16 NM_022114 16 SLC6A20 NM_020208 7 MYRIP NM_015460 16 SLC16A8 NM_013356 7 BMP7 NM_001719 15 THC2004763 THC2004763 7 ERMN AB033015 14 A_24_P109661 A_24_P109661 7 SLC13A3 NM_022829 14 ACOT11 NM_147161 7 SLCO1C1 NM_017435 14 CDH3 NM_001793 6 LRAT NM_004744 13 FRZB NM_001463 6 OPCML BX537377 13 LOC439949 AY007155 6 RLBP1 NM_000326 13 SERPINF1 NM_002615 6 TRPM3 NM_206948 13 GPAM NM_020918 6 KCTD4 NM_198404 13 SPOCK1 NM_004598 6 THC1934449 THC1934449 13 FLJ30594 NM_153011 6 LRP8 NM_017522 12 MUPCDH NM_031264 6 SLC39A12 NM_152725 12 C1orf168 AK125198 6 DUSP6 NM_001946 12 CLDN19 NM_148960 6 ERMN BC026345 11 LMO1 NM_002315 6 SLCO1A2 NM_005075 11 GLDC NM_000170 6 RBP1 NM_002899 11 A_24_P186746 A_24_P186746 5 SLC16A3 NM_004207 11 RDH11 NM_016026 5 CNDP1 NM_032649 11 SFRP5 NM_003015 5 SLCO1A2 NM_134431 11 SGK1 NM_005627 5 THC1839330 THC1839330 11 KRT18 NM_000224 5 SULF1 NM_015170 11 OPHN1 NM_002547 5 SLC16A14 NM_152527 10 TDRD9 NM_153046 5 WFDC1 NM_021197 10 EZR NM_003379 5 SLC6A13 NM_016615 10 FAM40B BC019064 5 CACNB2 BG428517 10 C7orf46 BC042034 5 SLC2A12 NM_145176 10 CTSD AK022293 5 SLC6A12 NM_003044 10 DHCR7 NM_001360 4 KIAA0953 AF131834 10 ITGAV NM_002210 4 ADORA2B NM_000676 10 GALNT11 NM_022087 4 CA14 NM_012113 10 THC1967593 THC1967593 4 PNPLA3 NM_025225 10 LOC650392 BC036550 4

52 Human RPE-specific gene expression

gene symbol Genbank ID FC gene symbol Genbank ID FC RGR NM_002921 10 HPD NM_002150 4 STRA6 NM_022369 9 BCAT1 NM_005504 4 C7orf46 NM_199136 9 A_23_P122650 A_23_P122650 4 KIRREL2 NM_199180 9 A_32_P226525 A_32_P226525 4 RDH5 NM_002905 9 PCP4 NM_006198 4 BMP4 NM_001202 9 A_32_P112546 A_32_P112546 4 TMEM56 NM_152487 9 KIAA1576 NM_020927 4 THC1892753 THC1892753 9 A_23_P73096 A_23_P73096 4 CNKSR3 NM_173515 9 BASP1 NM_006317 3 FC: fold change, average expression level in RPE compared to choroid and photoreceptors in six arrays.

We next hypothesized that these 39 genes, given their very high expression, could easily have been detected by other means in previous studies. We compared our 39 highly expressed RPE-specific genes from the current study to a database containing 13,037 genes described to be expressed specifically in either the retina or the RPE, described by Schulz et al[4] and were able to identify 29 of our genes in their database (75%) (see Table 4). A more thorough manual search of individual studies on the expression of genes in the RPE showed that 85% of the 39 genes were previously identified to be expressed in the RPE (Table 4).

Table 4. 39 RPE-specific genes with high expression levels as determined in our previous study[3].

gene sym- Genbank review bol ID FC RPE expression in literature ref Schulz C6orf105 NM_032744 22 microarray 31 BEST1 NM_004183 21 immunohistochemical staining 32 TMEM27 NM_020665 19 this study, Figure 11 LRP8 NM_004631 19 microarray 31 1 LGI1 NM_005097 18 review, exclusively in RPE studies 4 1 FAM40B AB032996 17 cDNA clones 33 ERMN AB033015 14 this study, Figure 11 1 LRAT NM_004744 13 western blot, northern blot 34 RLBP1 NM_000326 13 fluorescence immunocytochemistry 35 1 DUSP6 NM_001946 12 est database 36 1 RBP1 NM_002899 11 review, genes expressed in retina/RPE 4 1 SLC16A3 NM_004207 11 immunofluorescence 37 1 WFDC1 NM_021197 10 immunocytochemistry, microarray 8,38 1 KIAA0953 AF131834 10 review, genes expressed in retina/RPE 4

53 Chapter 3

gene sym- Genbank review bol ID FC RPE expression in literature ref Schulz CA14 NM_012113 10 immunocytochemistry 39 1 RGR NM_002921 10 microarray 31 1 STRA6 NM_022369 9 immunohistochemistry 40 RDH5 NM_002905 9 northern blot 19 1 BMP4 NM_001202 9 RNAse protection assay 41 1 CXCL14 NM_004887 8 microarray, RT-PCR 38 1 LHX2 NM_004789 8 in situ hybridization 42 1 C1QTNF5 NM_015645 7 cDNA library, RT-PCR 43,44 1 SLC6A20 NM_020208 7 this study, Figure 11 SLC16A8 NM_013356 7 PCR 45 1 CDH3 NM_001793 6 western blot 46 1 FRZB NM_001463 6 review, genes expressed in retina/RPE 4 1 SERPINF1 NM_002615 6 amino acid sequencing 47 1 SPOCK1 NM_004598 6 this study, Figure 11 LMO1 NM_002315 6 this study, Figure 11 RDH11 NM_016026 5 bovine and monkey, immunohistochemistry 20 1 and in situ hybridization SFRP5 NM_003015 5 northern blot 48 1 SGK1 NM_005627 5 this study, Figure 11 1 KRT18 NM_000224 5 CNV RPE RT-PCR 49 1 EZR NM_003379 5 rat immunofluorescence and immunoelectron 50 1 microscopy DHCR7 NM_001360 4 microarray 31 1 ITGAV NM_002210 4 microarray 31 1 GALNT11 NM_022087 4 review, genes expressed in retina/RPE 4 1 PCP4 NM_006198 4 review, genes expressed in retina/RPE 4 1 BASP1 NM_006317 3 microarray 31 1 Column four and five show 33 of the 39 genes (85%) were previously described (RNA or protein level) in indi- vidual studies in the literature using several different techniques to have RPE expression. Column six shows that 29 of the 39 genes (74%) were also present in a study by Schulz et al.[4] reporting on 13,037 genes expressed in the retina/RPE (their supplementary table 2) ref: literature reference, FC: fold change, average expression level in RPE compared to choroid and photoreceptors in six arrays, 1 in column six: present.

For the remaining 15% (6 genes) no data was available in the literature. Using semi quantita- tive QPCR (s-QPCR), we confirmed RPE expression of these genes (Table 4 and Figure 2). Both ERMN (AB033015) and SLC6A20 (NM_020208) appear to be more highly expressed in RPE compared to both choroid or photoreceptors, thereby fully confirming the microar- ray data. TMEM27 (NM_020665) and LMO1 (NM_002315) appear also to more highly

54 Human RPE-specific gene expression expressed in the RPE than in the choroid, but show approximate equal expression in the photoreceptors. SGK1 (NM_005627) is ubiquitously expressed in all three cells examined. The expression of SPOCK1 (NM_004598) in the RPE is rather low compared to the photo- receptors and does apparantly not confirm the microarray data.

Figure 2. Confirmation of microarray results by s-QPCR.

Beta actin, a household gene, was used to normalize gene expression in all cells of the retina. The black bars in- dicate RPE expression levels, the grey bars indicate choroid expression and the white bars indicate photoreceptor expression. For all genes RPE expression was shown in the microarray. See also text and Table 4.

Functional annotation of RPE-specific gene expression We used the list of 114 RPE-specifically expressed genes for further functional annotation of the major pathways specific for the RPE. The online database DAVID did not identify any Kegg pathways in this group of genes. However, among the 95 genes recognized by the Ingenuity software, there was a significant overrepresentation of genes in the following func- tional categories: symport, membrane transport, vision, glycoprotein, transport (p<0.0001). Moreover, using the Ingenuity database we identified three canonical pathways (see Figure 3): RAR-activation (Figure 4), retinol metabolism (Figure 5) and GABA receptor signaling (Figure 6). We found four biological functions with a significant overrepresentation of genes: ophthalmic disease, visual system development and function, genetic disorder and nervous system development and function (see Figure 7). Finally, we identified four networks cor- related with our RPE-specific gene list (see Figure 8-11). Similar functional annotations were found for the group of 39 highly expressed RPE-specific genes (data not shown). Upon closer inspection of this group we also identified eight (almost one in five!) known retinal disease genes (BEST1 [NM_004183], C1QTNF5 [NM_015645], CDH3 [NM_001793], LRAT [NM_004744], RDH5 [NM_002905], RDH11 [NM_016026], RGR [NM_002921], RLBP1 [NM_000326] and STRA6 [NM_022369]). Fifteen entries rep- resented membrane bound or transmembrane genes.

55 Chapter 3

Figure 3. Three canonical pathways identified by Ingenuity software[24] in the group of RPE-specific genes. 3

2.5

2 e) lu va

p- 1.5 (B-H (B-H og -l 1

0.5

0 RAR activation Retinol metabolism GABA receptor signaling

The three bars represent the canonical pathway identified, the x-axis identifies the pathways. The y-axis shows the -log of the Benjamini-Hochberg (B-H) p-value. The dotted line represents the threshold above which there are statistically significantly more genes in a pathway than expected by chance.

56 Human RPE-specific gene expression

Figure 4. The RAR-activation pathway identified by the Ingenuity software.

This is one of the canonical pathways that contains statistically significantly more genes than expected by chance in the group of 114 genes with RPE-specific expression. This figure shows the proteins corresponding to the over- represented genes. Colored fields indicate their presence among the 114 genes with RPE-specific expression, un- colored genes are added by the software to form pathways. Solid lines between molecules indicate direct physical relationships between molecules (such as regulating and interacting protein domains); dotted lines indicate indirect functional relationships (such as co-regulation of expression of both genes in cell lines).

57 Chapter 3

Figure 5. The retinol metabolism pathway identified by the Ingenuity software.

This is one of the canonical pathways that contains statistically significantly more genes than expected by chance in the group of 114 genes with RPE-specific expression. This figure shows the proteins corresponding to the over- represented genes. Colored symbols indicate their presence among the 114 genes with RPE-specific expression, uncolored genes are added by the software to form pathways. Solid lines between molecules indicate direct physi- cal relationships between molecules (such as regulating and interacting protein domains); dotted lines indicate indirect functional relationships (such as co-regulation of expression of both genes in cell lines).

58 Human RPE-specific gene expression

Figure 6. The GABA receptor signaling pathway identified by the Ingenuity software.

This is one of the canonical pathways that contains statistically significantly more genes than expected by chance in the group of 114 genes with RPE-specific expression. This figure shows the proteins corresponding to the over- represented genes. Colored symbols indicate their presence among the 114 genes with RPE-specific expression, uncolored genes are added by the software to form pathways. Solid lines between molecules indicate direct physi- cal relationships between molecules (such as regulating and interacting protein domains); dotted lines indicate indirect functional relationships (such as co-regulation of expression of both genes in cell lines).

59 Chapter 3

Figure 7. Four most significant biological functions identified by Ingenuity software[24] in the group of RPE- specific genes.

4

3.5

3

e) 2.5 lu va

p- 2 B-H B-H

og( 1.5 -l

1

0.5

0 Ophthalmic disease Visual system Genetic disorder Nervous system development and development and function function

The four bars represent the canonical pathway identified, the x-axis identifies the pathways. The y-axis shows the -log of the Benjamini-Hochberg (B-H) p-value. The dotted line represents the threshold above which there are statistically significantly more genes in a pathway than expected by chance.

60 Human RPE-specific gene expression

Figure 8. Most significant molecular network generated by the Ingenuity software.

The network is generated from our dataset with RPE-specific expression (114 entries; see text). Note that the col- ored symbols represent gene entries that occur in our data set, while the transparent entries are molecules from the knowledge database, inserted to connect all relevant molecules in a single network. Solid lines between molecules indicate direct physical relationships between molecules (such as regulating and interacting protein domains); dot- ted lines indicate indirect functional relationships (such as co-regulation of expression of both genes in cell lines). Abbreviation of gene names are according to standard abbreviations used in Genbank[28]. The main functionalities given by Ingenuity for this molecular network are Nervous system development and func- tion, Visual system development and function, organismal development. This network overlaps with the network in Figure 2) Highlights in this network include: (a) the regulating role of the MAPK/ERK pathway, a very complex signal transduction pathway that couples intracellular responses to the binding of growth factors to cell surface recep- tors; (b) platelet-derived growth factor (PDGF BB), known to induce RPE cell proliferation and migration and the development of proliferative vitreoretinopathy (PVR), acts indirectly on multiple molecules in this network, and c) the RPE retinol metabolism is present in the periphery of this molecular network.

61 Chapter 3

Figure 9. Second most significant molecular network generated by the Ingenuity software.

The network is generated from our microarray dataset with RPE-specific expression (114 entries; see text). For explanation of symbols on the diagrams see legend Figure 8. The main functionalities given by Ingenuity for this molecular network are cellular development, hematological system development and function and connective tissue development and function. Highlights in this network include: (a) The dual presence of GABA receptors SLC6A12 [NM_003044] and SL- C6A13 [NM_016615]; the presence and interactions of (b) the insulin-1 (INS1) protein and (c) the hormone progesterone.

62 Human RPE-specific gene expression

Figure 10. Third significant molecular network generated by Ingenuity.

The network is generated from our microarray dataset with RPE-specific expression (114 entries; see text). For explanation of symbols on the diagram see legend Figure 8. The main functionalities given by Ingenuity for this molecular network are cellular movement, nervous system development and function and gene expression. Please note the central signaling role of beta-estradiol; known to affect retinal function and disease.

63 Chapter 3

Figure 11. Fourth significant molecular network generated by Ingenuity.

The network is generated from our microarray dataset with RPE-specific expression (114 entries; see text). For explanation of symbols on the diagrams see legend Figure 8. This network overlaps with the network in Figure 8. The main functionalities given by Ingenuity for this molecular network are lipid metabolism, molecular transport and nucleic acid metabolism. The highlights of this network include: The central roles for (a) the HNF4A transcription factor and (b) the NFkap- pa Beta and (c) Wnt signaling pathways.

Discussion Study design In this study we provide a further specification of the RPE transcriptome, by analyzing the genes expressed in the RPE in reference to their expression in photoreceptors and choroid. We performed functional analyses on our microarray data using the online database DAVID and Ingenuity software in order to categorize the data and identify important functional prop- erties in 114 genes specifically expressed in the RPE. Moreover, we identified 39 genes with

64 Human RPE-specific gene expression

RPE-specific expression and high expression levels in the RPE and in 85% of genes we were able to confirm RPE expression using the literature. For the remaing 15%, sQPCRs were carried out which in most cases confirmed RPE enriched expression. Our current study design was focused on comparative gene expression, genes with high expression levels in two or multiple tissues were not included. As discussed extensively elsewhere[5,8], some degree of cellular contamination by applying LDM procedures to the retina is unavoidable. An important issue in this study was whether we could overcome the cellular contamination problem by comparing the expression profiles of three adjacent cellular monolayers, in other words, did we succeed in identifying RPE specific transcripts? A number of considerations are important. In the first place, obviously, if we would choose our comparative criteria even stricter (for example expression RPE > 20 x than in photoreceptor or choroid, versus current criteria RPE > 2.5 x than in photoreceptor or choroid), we would obtain an even more specific RPE expressed data set. In the second place, in our current dataset of 114 RPE genes, we did not find any obvious contamination of highly expressed (known) photoreceptor transcripts, like opsins. In the third place, the ma- jority of the 114 genes interact with each other via a limited number of molecular pathways and networks, with functional annotations which more or less can can be attributed to the RPE. This further confirms the RPE specific origin of these transcripts. For the majority of completely unknown genes, our sQPCR data confirmed the microarray data. Only one entry (SPOCK1) showed apparent higher photoreceptor expression than the RPE equivalent, and did not confirm the microarray data. The reason for this discrepancy remains to be elucidated. All data taken together, RPE specificity for the majority of 114 transcripts identified is highly likely; better than previously, although some minor degree of contamination can still not be excluded.

Functional properties of the RPE RPE versus photoreceptors only Compared to the photoreceptors, the RPE cells express genes from three different functional categories at significantly higher levels. The first is cell adhesion molecules. The RPE rep- resents the outer blood-retina barrier (BRB) and this group of molecules most likely illus- trated the importance for the RPE to adhere firmly to BM in order to maintain the integrity of the barrier. The second pathway is the melanogenesis pathway uniquely present in the RPE. Indeed, in the heavily pigmented RPE, the pigment granules protect against oxidative stress[11]. The third RPE pathway is the type I diabetes mellitus pathway. Twelve out of 13 genes are members of the major histocompatibility complex (MHC). This pathway was also present among the genes with highly variable expression levels in the RPE in our previous study[3]. The MHC genes are responsible for antigen presentation and are implicated in the RPE specific immune response in both health and disease[12].

In addition to the Kegg pathways described above, there were genes overrepresented in a number of functional categories, related to major functions of the RPE. There were secreted proteins, proteins involved in signaling and Ca2+-binding proteins. It is well known that many

65 Chapter 3

RPE-specific Ca2+ channels are involved in intracellular signaling, cellular signal transduc- tion and the regulation of secretion of various factors[13]. Additional functional categories included membrane proteins, cell adhesion proteins and ex- tracellular matrix (ECM) proteins, that correlate with the structural role of the RPE and its interaction with BM[1,14]. Finally, there was an overrepresentation of glycoproteins. Glycoproteins are crucial for the phagocytosis of photoreceptor outer segments[15].

RPE versus choroid only In the group of genes with expression levels higher in the RPE than in the choroid, there was an overrepresentation of genes in the olfactory transduction pathway. This pathway contains mainly guanylate cyclases and calcium/calmodulin-dependent protein kinases, naturally pres- ent in such highly active cells as the RPE. In addition, there was a high abundance of genes involved in vision and transport. Obviously, the RPE plays a dominant role in the transport of many signaling molecules and in the transport of waste material from the photoreceptor cells.

Genes with RPE-specific expression We functionally annotated the 114 genes with RPE-specific expression using both David and Ingenuity software. As expected, both programs yielded vision, visual and nervous system development and function, ophthalmic disease and genetic disorder as significantly over- represented groups of genes. Indeed, most of the genes in which mutations lead to retinal disorders are genes with high and specific expression in either RPE or photoreceptors[3]. Using David we also found an overrepresentation of genes involved in sym- or transport. This most likely represents one of the major functions of the RPE, which is the transport of biomolecules from the choroid toward the photoreceptors and vice versa. The photorecep- tors heavily depend on nutrients, oxygen, hormones, etc. from the bloodstream. Meanwhile, waste products like oxidized cholesterol, visual cycle intermediates and excess water leave the retina through the RPE[1].

The Ingenuity database also revealed three canonical pathways, RAR-activation, retinol me- tabolism and GABA receptor signaling. Binding of retinoic acid to the retinoic acid receptor (RAR), leads to tissue-specific activation or suppression of downstream transcription[16,17]. In mature RPE, this process is invaluable for maintenance of differentiation and homeosta- sis[16]. The significant presence of this pathway may thus be explained by the need of the RPE to counteract local insults (oxidative stress, lipid digestion, lipofuscin accumulation) and maintain homeostasis. The retinol metabolism pathway contained the RDH5 [NM_002905], RDH10 [NM_172037] and RDH11 [NM_016026] genes, all three genes are involved in the RPE part of the visual cycle[18]. It is well known that the RPE converts all-trans retinoids to 11-cis isomers. More specifically, RDH5 [NM_002905] and RDH11 [NM_016026] convert 11-cis retinol to 11-cis retinal, while RDH10 [NM_172037] converts 11-cis retina to all-trans retinal[18-21].

66 Human RPE-specific gene expression

Finally, GABA is an important inhibitory neurotransmitter from the GABA receptor sig- naling pathway (Grsp) present in both brain and retina[22]. This pathway is involved in the retinal reuptake of GABA from the subretinal space. Interestingly, at least one protein from this pathway (SLC6A12 [NM_003044]) was previously observed in the rat and bullfrog RPE[22].

Comparison of RPE-expressed genes to the literature Previous studies, combined into one review, claimed to reveal 246 genes to be expressed exclusively in the RPE[4]. Strikingly, in the current study, we found that 23 out of these 246 genes (9%) had higher expression levels in the photoreceptors than in the RPE. In addition, 72 of the 246 genes (29%) had higher expression levels in the choroid than the RPE. This indicates that at least a certain level of contamination appears to be present in the RPE signal of previously performed studies. Consequently, care should be taken when interpreting the results from these studies.

Conclusions Our study provides a detailed description of RPE-specific gene expression, as compared to both adjacent cell layers, photoreceptors and the choroid. In addition we provide a detailed functional analysis of the functional properties of the RPE-specific genes. We show the in- volvement of the RPE in RAR-activation, retinol metabolism and GABA receptor signaling. Moreover, for 85% of the genes we call RPE-specific with high expression levels, we could more or less verify our results using the literature. Finally, we added a substantial number of new genes significantly expressed in the RPE.

Methods Human donor eyes This study was performed in agreement with the declaration of Helsinki on the use of human material for research. Material used in this study was provided to us by the Corneabank Am- sterdam. In accordance with Dutch law, the Corneabank ensured none of the donors objected to the use of their eyes for scientific purposes. Approval of the medical ethics committee was not required as data were analyzed anonymously. A detailed description of our methods can be found elsewhere[3]. In brief, we selected five eyes from five human postmortem donors. Globes were enucleated between 16 and 22 hours post mortem and frozen several hours later according to a standard protocol. Donors were aged 63 to 78 years at time of death. We chose older donors in order to minimize the likelihood of the presence of yet undiagnosed monogenic eye diseases. The donors died of cardiovascular or cerebrovascular causes or of chronic obstructive pulmonary disease. Donors did not have a known ophthalmic disorder. Visual examination and histological examination, including periodic acid Schiff (PAS) stain- ing, indicated no retinal pathology. Three eyes were selected for the analysis of RPE vs. choroid, due to limited tissue availability only one of these eyes was also used for the analysis of RPE vs. photoreceptors. For the second and third comparison of RPE vs. photoreceptors, two additional eyes were selected.

67 Chapter 3

Cell sampling Globes were snap-frozen and stored at -80°C until use. A macular fragment of 16 mm2 with the fovea in its centre was cut from each of the retinas, as described previously[8]. For each eye, multiple cryosections were stained with periodic-acid Schiff and microscopically exam- ined for abnormalities. Twenty μm sections from the macular areas were used for the isola- tion of choroid, RPE cells and photoreceptor cells. A Cresyl Violet staining (LCM Staining Kit, Ambion) was applied to the sections intended for the isolation of photoreceptor cells, according to the manufacturer’s protocol. No staining was applied to sections to be used for the isolation of RPE cells or choroid. All sections were dehydrated with ethanol and air-dried before microdissection with a Laser Microdissection System (PALM, Bernried, Germany) (Figure 12). Cells were stored at -80° Celsius.

Figure 12. Frozen sections of the retina before and after laser dissection microscopy to isolate a) and b) RPE cells, c) and d) choroid, and e) and f) photoreceptor cells. a) b)

c) d)

e) f)

Note the relatively poor morphology due to the use of frozen sections. The sections used for the isolation of pho- toreceptor cells were stained with cresyl violet (see methods section), the other sections were unstained. The scale can be found in figures a and b.

68 Human RPE-specific gene expression

RNA isolation and amplification Total RNA was isolated and the mRNA component was amplified[8]. Amplified RNA (aRNA) was quantified with a nanodrop (Isogen Life Science B.V., The Netherlands) and the quality was checked on a BioAnalyzer (Agilent Technologies, Amstelveen, The Netherlands). Sub- sequently, aRNA samples were labeled with either a Cy3 or a Cy5 fluorescent probe.

Microarray design and handling For all hybridizations a 44k microarray was used (Agilent Technologies, Amstelveen, The Netherlands). For three of the donors, photoreceptor RNA was hybridized against RPE RNA. In addition, for three donors, choroidal RNA was hybridized against RPE RNA. Hybridiza- tion, washing and scanning were performed as described previously[8].

Data analysis Scanned images were processed and analyzed with Feature Extraction software (v 8.5 Agi- lent) and Rosetta Resolver software (Rosetta Inpharmatics). All data is MIAME compliant and the raw data has been deposited in the Gene Expression Omnibus(GEO) database. For each gene we calculated either the ratio between the RPE and the photoreceptor signal, or the ratio between the RPE and the choroid signal, depending on the array. This resulted in three ratio’s for the RPE versus photoreceptors and three ratio’s for the RPE versus choroid for each gene. Only when all three ratios for a single gene were greater than 2.5, did we con- sider a gene to have a meaningfully higher gene expression (GE) in one tissue compared to the other. We chose differences in GE of at least two and a half-fold (fold change (FC)>2.5) as cut off criterion for RPE compared to photoreceptor GE (RPE>phot), RPE compared to choroid GE (RPE>chor) and for RPE compared to photoreceptor as well as choroid GE (RPE>phot&chor). The same criteria were applied to photoreceptor compared to RPE GE (phot>RPE) and choroid compared to RPE GE (chor>RPE). A functional analysis of Kegg pathways (Kyoto Encyclopedia of Genes and Genomes) and functional categories was per- formed on all groups of genes with at least FC>2.5 using the David online software[10]. Cut off criteria used were a p-value of less than 0.001 using either a Benjamini-Hochberg cor- rection or an Ease score, a modified Fisher’s exact test[10,23]. More advanced analyses of our RPE-specific genes was performed with the Ingenuity knowledge database[24] (version IPA 7.4, april 2009) yielding biological functions, canonical pathways and gene networks. We searched the literature for proof of RPE expression of our RPE-specific genes with high expression levels using the Genecards website[25] the online Mendelian inheritance in man (OMIM) website[26] and the Pubmed website[27] using the gene name combined with ‘RPE’ or ‘retina’ or ‘retinal pigment epithelium’ or ‘expression’ as search criteria.

Confirmation of microarray results For confirmation of our microarray data sQRT-PCR was used (and not QPCR) since sQRT- PCR is less sensitive for a) relatively poor RNA quality which is unavoidably obtained from human donor eyes, and b) for adjacent cell contamination of the LDM samples. Moreover, our aim was to find an approximation of expression in the RPE, choroid and photoreceptors.

69 Chapter 3 sQRT-PCR was carried out, in triplicate, using exon spanning primers on RNA from LDM derived cell samples of the RPE, choroid and photoreceptors. Primer sequences are available on request. Six genes for which no further literature data on RPE gene expression was avail- able, were tested (table 4). B-actine, a household gene, was used to normalize gene expres- sion in between all cells of the retina.

70 Human RPE-specific gene expression

References 1. Strauss O (2005) The retinal pigment epithelium in visual function. Physiol Rev 85: 845-881. 2. Saint-Geniez M, D’Amore PA (2004) Development and pathology of the hyaloid, cho- roidal and retinal vasculature. Int J Dev Biol 48: 1045-1058. 3. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ et al (2009) Function- al annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 10: 164. 4. Schulz HL, Goetz T, Kaschkoetoe J, Weber BH (2004) The Retinome - defining a ref- erence transcriptome of the adult mammalian retina/retinal pigment epithelium. BMC Genomics 5: 50. 5. Ishibashi K, Tian J, Handa JT (2004) Similarity of mRNA phenotypes of morphologi- cally normal macular and peripheral retinal pigment epithelial cells in older human eyes. Invest Ophthalmol Vis Sci 45: 3291-3301. 6. Buraczynska M, Mears AJ, Zareparsi S, Farjo R, Filippova E et al (2002) Gene expres- sion profile of native human retinal pigment epithelium. Invest Ophthalmol Vis Sci 43: 603-607. 7. Bowes Rickman C., Ebright JN, Zavodni ZJ, Yu L, Wang T et al (2006) Defining the hu- man macula transcriptome and candidate retinal disease genes using EyeSAGE. Invest Ophthalmol Vis Sci 47: 2305-2316. 8. van Soest S, de Wit GM, Essing AH, ten Brink JB, Kamphuis W et al (2007) Compari- son of human retinal pigment epithelium gene expression in macula and periphery high- lights potential topographic differences in Bruch’s membrane. Mol Vis 13: 1608-1617. 9. Agilent (2009) http://www.home.agilent.com/agilent/home.jspx 10. Dennis G, Jr., Sherman BT, Hosack DA, Yang J, Gao W et al (2003) DAVID: Database for Annotation, Visualization, and Integrated Discovery. Genome Biol 4: 3. 11. Schraermeyer U, Heimann K (1999) Current understanding on the role of retinal pig- ment epithelium and its pigmentation. Pigment Cell Res 12: 219-236. 12. Crane IJ, Liversidge J (2008) Mechanisms of leukocyte migration across the blood- retina barrier. Semin Immunopathol 30: 165-177. 13. Rosenthal R, Strauss O (2002) Ca2+-channels in the RPE. Adv Exp Med Biol 514: 225- 235. 14. Del Priore LV, Geng L, Tezel TH, Kaplan HJ (2002) Extracellular matrix ligands pro- mote RPE attachment to inner Bruch’s membrane. Curr Eye Res 25: 79-89. 15. Tien LF, McLaughlin BJ, Cooper NG (1991) Glycoproteins in the retinal pigment epi- thelium of normal and dystrophic rats. Invest Ophthalmol Vis Sci 32: 319-326. 16. Mark M, Ghyselinck NB, Chambon P (2009) Function of retinoic acid receptors during embryonic development. Nucl Recept Signal 7: e002. 17. Rochette-Egly C, Germain P (2009) Dynamic and combinatorial control of gene expres- sion by nuclear retinoic acid receptors (RARs). Nucl Recept Signal 7: e005. 18. Wu BX, Moiseyev G, Chen Y, Rohrer B, Crouch RK et al (2004) Identification of RDH10, an All-trans Retinol Dehydrogenase, in Retinal Muller Cells. Invest Ophthal- mol Vis Sci 45: 3857-3862.

71 Chapter 3

19. Simon A, Hellman U, Wernstedt C, Eriksson U (1995) The retinal pigment epithelial- specific 11-cis retinol dehydrogenase belongs to the family of short chain alcohol dehy- drogenases. J Biol Chem 270: 1107-1112. 20. Haeseleer F, Jang GF, Imanishi Y, Driessen CA, Matsumura M et al (2002) Dual-sub- strate specificity short chain retinol dehydrogenases from the vertebrate retina. J Biol Chem 277: 45537-45546. 21. Wu BX, Chen Y, Chen Y, Fan J, Rohrer B et al (2002) Cloning and characterization of a novel all-trans retinol short-chain dehydrogenase/reductase from the RPE. Invest Ophthalmol Vis Sci 43: 3365-3372. 22. Peterson WM, Miller SS (1995) Identification and functional characterization of a dual GABA/taurine transporter in the bullfrog retinal pigment epithelium. J Gen Physiol 106: 1089-1122. 23. Hosack DA, Dennis G, Jr., Sherman BT, Lane HC, Lempicki RA (2003) Identifying biological themes within lists of genes with EASE. Genome Biol 4: R70. 24. Ingenuity (2009) www.ingenuity.com 25. Genecards (2009) http://www.genecards.org/index.shtml 26. online Mendelian inheritance in man (OMIM) (2009) http://www.ncbi.nlm.nih.gov/ sites/entrez?db=OMIM 27. Pubmed (2009) http://www.ncbi.nlm.nih.gov/sites/entrez?db=PubMed 28. Genbank (2009) http://www.ncbi.nlm.nih.gov/Genbank/ 29. entrez gene http://www.ncbi.nlm.nih.gov/sites/entrez?db=gene&cmd=search&term= 30. Efferth T (2003) Adenosine triphosphate-binding cassette transporter genes in ageing and age-related diseases. Ageing Res Rev 2: 11-24. 31. Cai H, Shin MC, Tezel TH, Kaplan HJ, Del Priore LV (2006) Use of iris pigment epithe- lium to replace retinal pigment epithelium in age-related macular degeneration: a gene expression analysis. Arch Ophthalmol 124: 1276-1285. 32. Marmorstein AD, Marmorstein LY, Rayborn M, Wang X, Hollyfield JG et al (2000) Be- strophin, the product of the Best vitelliform macular dystrophy gene (VMD2), localizes to the basolateral plasma membrane of the retinal pigment epithelium. Proc Natl Acad Sci U S A 97: 12758-12763. 33. Fam40B (2009) http://www.ncbi.nlm.nih.gov/IEB/Research/Acembly/av.cgi?db=human&l=FAM40B 34. Ruiz A, Winston A, Lim YH, Gilbert BA, Rando RR et al (1999) Molecular and bio- chemical characterization of lecithin retinol acyltransferase. J Biol Chem 274: 3834- 3841. 35. Huang J, Possin DE, Saari JC (2009) Localizations of visual cycle components in reti- nal pigment epithelium. Mol Vis 15: 223-234. 36. Est Dusp6 (2009) http://www.ncbi.nlm.nih.gov/nucest/19021976?report=genbank&log $=seqview 37. Deora AA, Philp N, Hu J, Bok D, Rodriguez-Boulan E (2005) Mechanisms regulating tissue-specific polarity of monocarboxylate transporters and their chaperone CD147 in kidney and retinal epithelia. Proc Natl Acad Sci U S A 102: 16245-16250.

72 Human RPE-specific gene expression

38. Radeke MJ, Peterson KE, Johnson LV, Anderson DH (2007) Disease susceptibility of the human macula: differential gene transcription in the retinal pigmented epithelium/ choroid. Exp Eye Res 85: 366-380. 39. Nagelhus EA, Mathiisen TM, Bateman AC, Haug FM, Ottersen OP et al (2005) Car- bonic anhydrase XIV is enriched in specific membrane domains of retinal pigment epi- thelium, Muller cells, and astrocytes. Proc Natl Acad Sci U S A 102: 8030-8035. 40. Kawaguchi R, Yu J, Honda J, Hu J, Whitelegge J et al (2007) A membrane receptor for retinol binding protein mediates cellular uptake of vitamin A. Science 315: 820-825. 41. Vogt RR, Unda R, Yeh LC, Vidro EK, Lee JC et al (2006) Bone morphogenetic pro- tein-4 enhances vascular endothelial growth factor secretion by human retinal pigment epithelial cells. J Cell Biochem 98: 1196-1202. 42. Tetreault N, Champagne MP, Bernier G (2009) The LIM homeobox transcription factor Lhx2 is required to specify the retina field and synergistically cooperates with Pax6 for Six6 trans-activation. Dev Biol 327: 541-550. 43. Agarwal N, Swaroop A, Zheng K, Liou JD, O’Rourke K et al (1995) Expression and chromosomal localization of cDNA clones from an enriched human retinal pigment epithelial (RPE) cell line library: identification of two RPE-specific genes. Cytogenet Cell Genet 69: 71-74. 44. Hayward C, Shu X, Cideciyan AV, Lennon A, Barran P et al (2003) Mutation in a short- chain collagen gene, CTRP5, results in extracellular deposit formation in late-onset retinal degeneration: a genetic model for age-related macular degeneration. Hum Mol Genet 12: 2657-2667. 45. Yoon H, Donoso LA, Philp NJ (1999) Cloning of the human monocarboxylate trans- porter MCT3 gene: localization to chromosome 22q12.3-q13.2. Genomics 60: 366-370. 46. Burke JM, Cao F, Irving PE, Skumatz CM (1999) Expression of E-cadherin by human retinal pigment epithelium: delayed expression in vitro. Invest Ophthalmol Vis Sci 40: 2963-2970. 47. Steele FR, Chader GJ, Johnson LV, Tombran-Tink J (1993) Pigment epithelium-derived factor: neurotrophic activity and identification as a member of the serine protease in- hibitor gene family. Proc Natl Acad Sci U S A 90: 1526-1530. 48. Chang JT, Esumi N, Moore K, Li Y, Zhang S et al (1999) Cloning and characterization of a secreted frizzled-related protein that is expressed by the retinal pigment epithelium. Hum Mol Genet 8: 575-583. 49. Stahl A, Paschek L, Martin G, Feltgen N, Hansen LL et al (2009) Combinatory inhibi- tion of VEGF and FGF2 is superior to solitary VEGF inhibition in an in vitro model of RPE-induced angiogenesis. Graefes Arch Clin Exp Ophthalmol 247: 767-773. 50. Bonilha VL, Finnemann SC, Rodriguez-Boulan E (1999) Ezrin promotes morphogen- esis of apical microvilli and basal infoldings in retinal pigment epithelium. J Cell Biol 147: 1533-1548.

73 Chapter 3

Acknowledgements We thank Dr. L. Pels and co-workers of the Corneabank, Amsterdam, for providing donor eyes. This study was supported by grants from the Foundation Fighting Blindness (T-GE- 0101-0172), The Netherlands Organization for Scientific Research (NWO; project 948-00- 013), The Royal Netherlands Academy of Arts and Sciences (KNAW), de Algemene 27 Nederlandse Vereniging ter Voorkoming van Blindheid (ANVVB) and het Edward en Marianne Blaauwfonds van het Amsterdams Universiteitsfonds.

74 Human RPE-specific gene expression

75

The dynamic nature of Bruch’s membrane

4

Judith C Booij, Dominique C Baas, Juldyz Beisekeeva, Theo GMF Gorgels and Arthur AB Bergen

Prog Retin Eye Res. 2010 Jan;29(1):1-18 Chapter 4

Abstract Bruch’s membrane (BM) is a unique pentalaminar structure, which is strategically located between the retinal pigment epithelium (RPE) and the fenestrated choroidal capillaries of the eye. BM is an elastin- and collagen-rich extracellular matrix that acts as a molecular sieve. BM partly regulates the reciprocal exchange of biomolecules, nutrients, oxygen, fluids and metabolic waste products between the retina and the general circulation. Accumulating evi- dence suggests that the molecular, structural and functional properties of BM are dependent on age, genetic constitution, environmental factors, retinal location and disease state. As a result, part of the properties of BM are unique to each human individual at a given age, and therefore uniquely affect the development of normal vision and ocular disease.

The changes occurring in BM with age include increased calcification of elastic fibres, in- creased cross-linkage of collagen fibres and increased turnover of glycosaminoglycans. In addition, advanced glycation end products (AGEs) and fat accumulate in BM.

These age-related changes may not only influence the normal age-related health of photo- receptor cells, but also the onset and progression of diseases like retinitis pigmentosa (RP) and age-related macular degeneration (AMD). Undoubtedly, BM is the site of drusen devel- opment. Confluent drusen and uncontrolled activation of the complement cascade are most likely the first signs of AMD. Furthermore, the nature of adhesive interactions between the RPE and BM are instrumental in the development of retinal detachments and proliferative retinal disease. Finally, BM is passively or actively involved in a range of other retinal dis- orders such as Pseudoxanthoma elasticum (PXE), Sorsby’s Fundus Dystrophy and Malattia Leventinese.

Here, we review the dynamic nature of Bruch's membrane, from molecule to man, during development, aging and disease. We propose a simple and straightforward nomenclature for BM deposits. Finally, we attempt to correlate recently published mRNA expression profiles of the RPE and choroid with molecular, structural and functional properties of BM. Our review may shed light on the complex involvement of BM in retinal pathology, notably age- related macular degeneration.

78 The dynamic nature of Bruch’s membrane

Introduction In the past, many investigators considered Bruch’s membrane (BM) a relatively boring and simple sheet of extracellular matrix, merely occupying space between the retinal pigment epithelium (RPE) and the choroid. Recently, however, interest in BM has increased exponen- tially; and understandably so, given its strategic location between the retina and the general circulation, and its crucial role in retinal function, aging and ocular disease. The pentalaminar BM structure forms a single functional unit with the RPE and choriocapillaris. It is involved in the essential exchange of numerous biomolecules, oxygen, nutrients and waste products in between these tissues. Given its unique location and structure, BM plays a crucial role in cell-cell communication, cellular differentiation, proliferation or migration, tissue remodel- ling and in shaping pathologic processes[1-4]. The nature of BM is highly dynamic, and depends on genetic factors, environmental burden, the topographic position in the retina, age and disease[5-11]. It has also become clear that BM is the focal point of local and systemic risk factors for initial stages of the most frequent untreatable blinding disorder known to man: age-related macular degeneration (AMD). In addition, BM is primarily or secondarily involved in a number of additional genetically determined ophthalmic disorders such prolif- erative vitreoretinopathy (PVR)[12,13], pseudoxanthoma elasticum (PXE)[14,15], Marfan syndrome[16-18], Sorsby’s fundus dystrophy[19-21] and others. Here, we review the existing knowledge on BM, with a focus on its normal structure and function and the role of BM in normal aging and early AMD pathology. Finally, we provide an outlook on possible AMD disease prevention or treatment.

Bruch’s Membrane: molecular composition, structure and normal physiology Development of BM The retina is a derivative of the neuroectoderm of the diencephalon. Around the 4th week of gestation, a secondary eye bubble, surrounded by ectoderm and mesenchyme, develops, which forms the future optic cup. Upon invagination of the optic cup in the sixth week of development, the future RPE and the undifferentiated neural retina can be distinguished. In the next stage, neural crest-derived mesenchyme, which will form the future choroid, starts to condensate around the optic cup. At the same time, the primitive retina, by now consisting of an RPE cell layer, an outer nuclear zone and an acellular marginal zone flanked by basal membranes, continues to differentiate. The outer RPE basal membrane becomes incorporated in the future BM. Analysis of the developing chick retina showed that, at the tenth week of gestation, collagen fibrils are deposited beneath the basal RPE lamina. The elastic fibre layer can be detected 3-4 weeks later. Full differentiation of the elastic fibre layer to a perforated sheet is achieved by mid-term. In chickens, BM most likely continues to mature over the next weeks, months or even years[2,22]. In normal mice, initially the basal membranes of the RPE and choroid develop, followed by the development of the collagen layers, and, fi- nally the central elastin layer[23]. Relatively little is known about the molecular and cellular events that regulate the early developmental phases of BM in man. However, it is not hard to imagine that deposition of collagen and elastin proteins is preceded by upregulation of the expression of corresponding genes in the adjacent tissues. Once BM formation has started,

79 Chapter 4 the ECM layers may affect cell-cell communication directly, thereby possibly creating the opportunity to further direct their own formation and differentiation. Although it is not ex- actly clear how BM is formed, gene expression data on adult RPE and choroid from our own lab (Figure 1)[11, 24], indicate that both the choroid and RPE cells are, in principle, capable of synthesizing the major components of BM. In conclusion, available evidence suggests that BM is (ultimately) formed or maintained from both the RPE and choroidal side, perhaps in a coordinated or stochastic fashion.

Figure 1. Schematic drawing of proteins present in Bruch’s membrane (see main text) and the corresponding genes expressed in adjacent cells.

Phot

COL4A3 FNDC5 TIMP3 RPE [COL 4A1, 4A2, 4A3, 4A4, 4A5, 5] LAM HSPG bm RPE [COL1, 3] FN inner collagen elastin layer BM [COL 4, 6, 18] ELN TIMP3* [COL1, 3] FN outer collagen [COL 4A1, 4A2, 5] LAM HSPG bm choroid COL 1A1, 1A2, 3A1, 4A1, LAM A2, FN1, 4A2, 5A1, 5A2, 6A1, 6A2, A4, B1, FNDC3A 18A1 C1, C3 Chor

We identified three genes (COL4A3, FNDL5, TIMP3) with higher expression levels in the RPE than in the cho- roid, and 17 genes (remainder) with higher expression levels in the choroid than in the RPE. A qualitative impres- sion of the gene expression is given. Gene expression levels were determined by RNA microarray study comparing gene expression levels from two adjacent tissue types from the same donor. Experiments were performed in tripli- cate (on three different healthy old human donor eyes)[24]. bm RPE: basement membrane of the RPE, bm choroid: basement membrane of the choroid. The abbreviation for basement membrane is not used in the accompanying text. A color version of this figure can be found in chapter 15.

Ultrastructure and protein content of BM layers According to the classification of Hogan in the early 1960’s[25], BM consists of five lay- ers (Figure 1). From the RPE toward the choroid, the following layers can be distinguished histologically: the basement membrane of the RPE, the inner collagenous layer (ICL), the elastin layer (EL), the outer collagenous layer (OCL) and, finally, the basement membrane of the choriocapillaris (Figure 1).

80 The dynamic nature of Bruch’s membrane

The basement membrane of the RPE The basement membrane of the RPE is a continuous BM layer approximately 0.14-0.15 µm in thickness in the young[26]. It resembles in many aspects other basement membranes in the body[27,28]. The RPE basement membrane contains many components similar to the base- ment membrane of the choriocapillaris: collagens type IV[29], laminin[30], fibronectin[31], heparan sulphate and chondroitin/dermatan sulphate[32] (Figure 1). In contrast, collagen type VI is not present in the basement membrane of the RPE.

The inner collagenous layer (ICL) The inner collagenous layer is approximately 1.4 µm in diameter. The ICL consists of 60 nm thick striated fibres of collagen type I, III, and V, organized in a multilayered grid-like structure[33]. The collagen grid is embedded in a mass of interacting biomolecules, such as glycosaminoglycans (chondroitin sulphate, dermatan sulphate and hyaluronic acid)[32] and components of the coagulation and complement system.

The elastin layer (EL) The elastin layer (EL) consists of several stacked layers of linear elastin fibres of varying shapes and sizes. The fibres form a perforated sheet with interfibrillary spaces of about 1 µm. The sheet is about 0.8 mm thick in the young and extends from the edge of the optic nerve to the pars plana of the ciliary body[33]. In addition to elastin fibres, the EL contains collagen type VI, fibronectin and other protein-associated substances. Recently, Chong and co-workers[34] found, by examining 121 human donor eyes, that the EL is three to six times thinner in the macula than in the periphery in all studied age groups. Collagen fibres from the ICL and OCL frequently cross the EL[33].

The outer collagenous layer (OCL) The OCL is less thick than the ICL (0.7 µm in the young) but the structure and components are largely identical to those of the ICL (see above)[33].

The basement membrane of the choriocapillaris The basement membrane of the choriocapillaris is a non-continuous, interrupted BM layer due to the so-called intercapillary columns of the choroid. The basement membrane has an average thickness of 0.14 mm in the young eye and is predominantly composed of laminin, heparan sulphate and collagens type IV, V and VI . Aisenbrey (2006) found that the RPE synthesizes laminins that preferentially adhere BM to the RPE through interaction with inte- grins[30]. Heparan sulphate (HS) is a common glycosaminoglycan in the BM. The (two) HS polysaccharide side chains bind to a variety of protein ligands in BM and regulate a range of biological activities. Roberts and Forrester[22] proposed that collagen type IV in the base- ment membrane of the choriocapillaris may inhibit endothelial cell migration into the BM. Collagen type V is present in most types of connective tissue, particularly in pericellular spaces and near basement membranes, and plays a role in platelet aggregation, epithelial cell migration and binding of interstitial collagen fibrils[35]. Type VI collagen is the major

81 Chapter 4 structural component of microfibrils, and specific for the choroidal basement membrane. It may be involved in anchoring BM to the capillary endothelial cells of the choroid[36]. Col- lagen VI possibly interacts with collagen I[22] which is abundant in both the BM-OCL and the choroidal matrix. A remarkable structural feature of the choriocapillaries adjacent to the choroidal basement membrane is the endothelial fenestrations, or pores, that are permeable to macromolecules[37].

Molecular composition of BM considering gene expression of RPE and choroid Since BM is an acellular layer, it most likely depends on the adjacent RPE and choroidal cells for the production of most of its extracellular matrix constituents[9]. Furthermore, a large number of biomolecules and waste products from the RPE and choroid pass through BM and can get trapped there influencing both the structure and function of BM. Finally, the molecular composition of BM changes with age and there is an extensive turnover of ECM molecules, driven by matrix metallo proteinases (MMPs), during life. We hypothesized that the RPE and choroidal gene expression profiles are potentially relevant to the molecular composition of BM[9] (Figure 1[24]). In this regard, we observe that 1) both the RPE and choroid are capable of producing BM proteins, such as collagens, fibronectin and heparan sulphate containing proteoglycans; and 2) for those proteins known to be pres- ent in BM, the choroidal cells (endothelial cells, fibroblasts, smooth muscle cells) appear to contribute more to BM than the RPE does. In summary, approximately sixteen different pro- teins (subunits) were previously assigned to BM by immunohistochemistry. Two (COL4A3, TIMP3; 12.5%) of these are predominantly synthesized by the RPE, two (FN, HSP; 12.5%) are synthesized both by the RPE and choroidal cells, and the majority (i.e. numerous colla- gens, elastin; 75%) of the known BM proteins is mainly synthesized by choroidal cells (Fig. 1). These data support the hypothesis of Sivaprasad and co-workers who suggested a com- mon origin of BM and the vascular intima[38]. Finally, it is of interest to note that both the RPE and choroidal cells produce mRNA of many more collagen (subunits) and other ECM molecules, that have not (yet) been assigned to BM[9,24].

Structure and molecular composition of BM in the macula and retinal periphery Evidence exists that BM is structurally different in the macular area compared to the retinal periphery. Interpretation of the literature in this respect is difficult since investigators use eyes from both human donors and animal models. Also the ages of the studied eyes vary considerably and frequently retinal punches are used from undefined or various retinal loca- tions. Nevertheless, in sheep, topographical differences in BM thickness were observed as early as 1983[39]. In human donor eyes of all ages, Chong and colleagues[34] found that the EL of BM in the macular area was three to six times thinner, and two to five times more porous than in the reti- nal periphery. RPE microarray studies in human donor eyes of age 17-36 years, performed by van Soest et al[9] provided evidence that RPE gene expression of at least 33 structural BM proteins (collagens, laminins, fibronectins and a number of proteoglycans) was lower in the macular area than in the retinal periphery, while two genes showed an inversed expres- sion pattern (COMP and THBS4). For collagen type IV chains, most fibronectin types, as

82 The dynamic nature of Bruch’s membrane well as elastin, no regional differences in RPE gene expression was observed[9] (Bergen and van Soest, unpublished results). These data suggest that the topographical differences in EL thickness observed by Chong et al[34] may not be due to the higher or lower transcriptional activity of the elastin gene. Taken together, these data may suggest that region-specific structural and functional prop- erties and/or turnover rates of components of BM exist. With the exception of the genes ALDH1A3, cKIT, FLJ36353, NADH dehydrogenase, RTBND2, TIMM17B, and WFDC1 [9,40,41], these gene expression data await further verification from other topographical studies on mRNA and protein level. Confirmation of additional differentially expressed genes will enable definite molecular and bioinformatic modelling of BM, and correlation of observed molecular, structural and functional differences.

BM functions The three primary functions of BM include 1) regulating the diffusion of (bio-) molecules between choroid and RPE, 2) providing physical support for RPE cell adhesion, migration and perhaps differentiation, and 3) acting as a barrier, restricting choroidal and retinal cel- lular migration (Figure 2). Obviously, the functional aspects are closely related to the local structure and molecular composition of BM (in the macula or in the periphery).

Figure 2. Schematic drawing showing the normal structure and functions of Bruch’s membrane.

biomolecules RPE RPE

BMP ICL x x

E

OCL

BMC

biomolecules Choroid

Transmission Electron Microscopic image of BM courtesy of Prof. J. Marshall. BM allows for the transport of biomolecules across the membrane (double arrows), it attaches RPE cells to the membrane (filled circle) and it acts as a physical barrier to prevent the migration of RPE cells and choroidal cells across the membrane (arrow with X). A color version of this figure can be found in chapter 15.

83 Chapter 4

Diffusion properties As BM is located between the RPE and choroid, its passive transport function is obvious. BM acts as a semi-permeable filter for the reciprocal exchange of biomolecules between the retina and the choroid. Given the acellular nature of BM, diffusion is primarily regulated by passive processes. Diffusion across BM depends on its molecular composition, which, in turn, is influenced by several factors like age and location in the retina. Indeed, Marshall and co-workers found a relationship between BM porosity and water flow: The EL showed the greatest pore size between the more or less randomly organized fibres and the largest water conductivity. The ICL had the smallest pores and the lowest conductivity[33]. Diffusion through the BM also depends on hydrostatic pressure on both sides of BM and on rescue and concentration of specific biomolecules and anorganic ions. Biomolecules try- ing to pass through the BM from the choroid to the RPE include nutrients, lipids, pigment precursors, vitamins (vitamin A), oxygen, minerals, antioxidant components, trace elements and (other) serum constituents[33,42,43]. All these molecules bind to BM, or are taken up by the RPE from the bloodstream via BM, since they are needed for optimal function of the photoreceptor RPE complex and also the neural retina.

(Bio-)molecules trying to pass BM from RPE to choroid, include CO2, water, ions, oxidized lipids, oxidized cholesterol, and other waste products cleared by the RPE. The waste prod- ucts consist of metabolic, visual cycle or electrophysiological waste from the photoreceptors or the RPE, as well as waste products from partly digested, partly oxidized, membranous fragments of shed photoreceptor outer segments (POS)[33,42-44]. Obviously, the filtration properties of BM are closely related to its structure and molecular composition and vary from the macula to the periphery. BM’s permeability to water is influenced by age-related collagen cross-linking, and the build up of hydrophobic lipids (lipid wall) and membranous debris in the aging BM[45].

Using Ussing chambers, Moore et al[46], and Statira et al[47,48] measured the movement of water across BM-choroid samples experimentally. They found an exponential decrease in permeability to water with age, measured over time-intervals of 9.5 to 15 years; these find- ings were later confirmed by Hillenkamp et al[49]. The rate of loss of water permeability was largest in the ICL and larger in the macular area than in the retinal periphery[48]. Interest- ingly, most of the loss of hydraulic permeability appears to occur in early life, long before BM debris can be visualized[50]. The latter finding suggests that, like many other biological and pathogenic processes, the molecular changes in BM precede the changes that can be visualized histologically[46,51].

RPE Cell adhesion and differentiation In addition to having filtration properties, BM also provides support and acts as an attach- ment site for the RPE[52,53]. BM may also act in RPE differentiation[54] and wound heal- ing[55,56]. BM from young donors is much more efficient in the attachment of RPE cells than BM from old donors[57]. Not all layers of BM show equally strong adhesion properties. For example, the RPE basal lamina of BM is the layer that RPE cells normally adhere to[52].

84 The dynamic nature of Bruch’s membrane

RPE-BM adhesion is mediated by integrin cell surface receptors. Integrins are a group of (RPE basal) membrane proteins that is capable of binding to a number of extracellular and BM matrix components such as laminin isoforms and type IV collagen in so-called anchoring plaques[58]. Indeed, cultured RPE cells overexpressing integrins show more adhesion to BM than normal RPE cells[59]. The RPE continues to develop until approximately 6 months after birth. After that, the RPE is generally considered to be post-mitotic. However, in the event of mechanical or light-induced damage to the RPE cell layer, the RPE cells can proliferate in a manner similar to that of other epithelial tissues. Small defects are corrected by migration of cells from the edges of the wound, in larger defects, there is also proliferation of cells. Wound healing is more ef- ficient in the presence of the basal lamina of the RPE than when this layer of BM is absent or damaged[60]. In vitro wound healing is impaired upon inhibition of integrins. It is known that BM thickens and calcifies with age, which may also impede the attachment of the RPE to BM. Finally, the process of wound healing appears to be disturbed in AMD patients, possibly due to alterations in the RPE and/or BM. This is exemplified by the fact that viable RPE cells from AMD patients did not grow well in culture. Moreover, dissection of a choroidal neovas- cularization (CNV) membrane in AMD patients was not followed by complete restoration of the RPE cell layer[61].

Division barrier for cell migration The outer blood-retina barrier (oBRB) is formed by RPE cells that are connected to each other by tight junctions. The oBRB prevents transport of molecules larger than 300 kDa into and out of the retina[62]. The barrier function of the RPE is physically supported by BM, that acts as a semi-permeable molecular sieve. The inner blood-retina barrier (iBRB) is comprised of a single layer of non-fenestrated retinal vascular endothelial cells connected by tight junctions[62]. If both leukocytes and endothelial cells are normal, leukocytes do not cross the iBRB[62-64]. Experiments in mice and rats showed that bone marrow-derived cells cannot easily pass the iBRB as these cells could not be detected in the retina one year after injection into the circulation[65,66]. Nonetheless, lymphocytes have been shown to infiltrate the normal retina despite an apparently intact iBRB[62] and may pass BM and the oBRB. If lymphocytes are activated, they are able to initiate a transient breakdown in the BRB, en- abling sampling of the retinal environment and possibly further recruitment of inflammatory cells[62].

The aging Bruch’s membrane and AMD pathology Normal Aging of BM and AMD Pathology The distinction between normal aging and pathology in, for example, AMD, is not clear cut. Interestingly, aging itself is the strongest risk factor for developing AMD. However, features of aging and disease may overlap, may be different for different cell types involved, and may even raise an almost philosophical discussion. For example, some investigators view aging itself as a pathology, that can ultimately be cured, while others do not.

85 Chapter 4

Here, we consider “normal aging” those changes that occur in the majority of individuals continuously from adulthood to old age, without direct clinical consequences. It is clear, however, that these age-related, non-pathogenic changes may affect the overall fitness of cells and tissues, predisposing them to a pathogenic state. We regard “pathogenic changes” as those changes that cause local loss of function, leading to clinical consequences. Since these categories certainly overlap, it may be useful to illustrate our line of thought with a number of examples. We consider the continuously decreasing vitality of RPE, photoreceptors and choroidal cells part of normal aging. This decrease in cellular fitness continues or is accelerated when spe- cific pathological processes set in. This is in line with observations of Curcio in man[67,68]. Interestingly, the rate of spontaneous photoreceptor cell loss with age may be genetically determined itself, since different wild type mouse strains have different rates of spontaneous photoreceptor loss[69]. We also consider the formation of lipofuscin in RPE as part of normal aging. However, we classify excessive lipofuscin accumulation, leading to local loss of function, cell damage and cell death, as (AMD) pathology[70]. Also, the deposition of lipids in BM can be seen as a normal aging phenomenon, until the point where the build up of the lipid wall starts to affect local RPE function. We also think that the formation of basal laminar deposits as well as (subclinical) drusen development, and the controlled involvement of the complement system to clear BM debris is a normal aging phenomenon. In 90 out of 100 apparently healthy Dutch donor eyes, age between 70 and 80, we observed subclinical macular drusen after histology and PAS staining (Bergen and co-workers, unpublished results). Many investigators consider the appearance of drusen a hallmark of AMD (pathogenicity). However, in our view, pathology only sets in when the involvement of the complement system becomes uncontrolled, and abnormal loss of local cellular function, cell damage and cell death occur.

Figure 3. Changes in Bruch’s membrane with age and its relation to AMD pathology. Lipid accumulation BM thickness ↑ Collagen cross-linking ↑ Aging BM changes AGE accumulation Aging BM changes Proteoglycansize ↑ Heparansulphate fraction ↑ EL calcification ↑

Uncontrolled Neovascularisation complement activation

AMD Pathology The upper half of the picture shows the changes that occur in BM with age. The lower half of the picture show how the age related changes can progress into AMD pathology, either through uncontrolled activation of the comple- ment system, through the occurrence of neovascularisation or as an indirect result of the aging of BM.

86 The dynamic nature of Bruch’s membrane

In summary, we support the early views of Marshall and co-workers that, with age, RPE and BM change continuously. Normal BM aging can insidiously change into (AMD) pa- thology[33]. Consequently, the processes underlying normal aging and AMD pathology are difficult to separate, and are discussed in a comprehensive fashion below. Our views are il- lustrated in Figure 3.

Structural and molecular changes of the aging BM The pentalaminar structure of BM, identified at birth, undergoes age-related changes throughout the larger part of life. The molecular composition and physiological aspects of BM change dramatically. In general, there is an overall increase in thickness and a reduced filtration capacity due to molecular modification and reconfiguration[50,71,72]. Remodelling of BM with age occurs at the biochemical and histological level, each causing functional changes (see Figure 4).

Figure 4. Schematic overview of the changes occuring in the aging BM, the functional consequences and the subsequent pathology.

Structural changes Functional effect Pathology

formation of lipid wall lipid accumulation basal deposits, drusen? transport impeded

elasticity ↓ BM thickness ↑ hydraulic permeability ↓ focal deposits, drusen

permeability, elasticity, turnover of BM components collagen cross-linking ↑ flexibility, filtration ↓ ↓ (Ramrattan et al. 1994)

aging, AMD AGE accumulation protein function ↓ (Yamada et al. 2006

AMD (Meri et al 1994) proteoglycan size ↑ anti-inflammatory response ↓ retinal pathology heparan sulphate fraction ↑ (Hewitt et al 1989)

elasticity ↓ neovascular AMD EL calcification ↑ membrane becomes brittle (Spraul et al. 1999)

Increased cross-linkage of collagen With age, increased collagen cross-linking occurs in BM. This has a negative effect on the permeability of BM and changes the nature of the extracellular matrix. Cross-linkage in- creases the strength and density of the collagen network but it decreases its elasticity, flex- ibility and perhaps filtration capabilities. The dense collagen network gradually becomes

87 Chapter 4

(more) inaccessible for the RPE collagenases which results in a less effective turnover of BM components (see Figure 3)[72].

Turnover of BM proteoglycans Proteoglycans (PG) are heavily glycosylated glycoproteins that are “the glue” of extracellu- lar matrices such as BM. PG contain a core protein covalently linked to glycosaminoglycan (GAG) chains. Individual functions of proteoglycans are determined by both the type of core protein and the type of GAG chains. Evidence from RPE cell culture experiments[32] and immuno-electron microscopy stainings[73] suggest that the RPE predominantly synthesizes heparan sulphate, which is incorporated into BM. Indeed, 58% of the PGs in BM are of the heparan sulfate type, which is primary located at the basal lamina of the RPE and cho- roid. Forty-two percent of BM PGs are chondroitin sulphate or dermatan sulphate, which are uniquely associated with collagen fibrils[32,74]. Interestingly, newly synthesized PG’s consist of 25% heparan sulphate and for 75% chon- droitin/dermatan sulphate[32]. Since the ratio between heparan sulphate and chondroitin/der- matan sulphate in BM remains unchanged during life, these data suggested that the turnover of heparan sulphate is slower than that of chondroitin sulphate/dermatan sulphate[32,74]. RPE gene expression data of several individuals suggested that proteoglycan turnover rate is tightly controlled[11]. Nonetheless, after the age of 70 years, there is a slight shift toward larger PG’s, indicating the inability of cells to normally digest the PG core protein[32]. Eyes from a limited number of donors with retinitis pigmentosa and diabetic retinopathy revealed relatively high heparan sulphate levels (55% - 64%) in BM compared with healthy controls (23%)[32,75]. PGs have important structural and filtration properties in BM, and may play a role in the anti-inflammatory response. PG molecules form structural networks through interaction with their side chains. These interactions occur not only among the different types of PGs, but also with collagen fibres in the OCL and ICL, and, most likely, with hyaluronic acid in the interphotoreceptor space. In terms of filtration properties, the negatively charged PG side chains cause PGs to bind wa- ter and positively charged cations such as sodium, potassium and calcium. In addition, they cause PGs to form a barrier for the passage of negatively charged macromolecules[32,76]. Thus, by physical, electrostatic and biochemical means, PGs regulate the movement of mol- ecules through the BM matrix. Evidence also reveals that PGs can affect the activity and sta- bility of proteins and signalling molecules within the matrix[76]. Finally, PGs, and especially heparan sulphate, may have anti-inflammatory properties. Meri et al[77] showed that heparan sulphate interacts with complement factor H, an important regulator of the complement cas- cade, and regulator of the local immune response. Conversely, Kelly and co-workers[78] showed that BM-associated heparan sulphate can modulate complement activity, by inhibition of the cleavage of complement factors B and C3 to, respectively Bb and C3b. Given the major involvement of the complement system and the innate immune system in the development of AMD[79], the interaction between heparan sulphate PG(s) and CFH may be one of the key molecular switches that turn normal RPE aging into AMD pathology.

88 The dynamic nature of Bruch’s membrane

Mineralization of BM Calcification of BM As in other soft tissues in the body, calcium can be deposited in the connective tissue of BM. In post mortem eyes, the deposits can readily be demonstrated, for example with the von Kossa staining method. Van der Schaft et al. (1992) studied 182 human maculae older than 33 years of age, and found calcium deposition in BM in 59% of the samples[71]. The presence and extent of the calcification was positively correlated with age, but not with AMD. How- ever, a later study by Spraul et al. (1999) did demonstrate a significant correlation between BM calcification and AMD[80]. The potential correlation with neovascular AMD agrees well with the notion that calcification of BM renders the membrane more brittle and more suscep- tible to breaks, allowing faster neovascularization. This finding was corroborated by the pathology observed in PXE, an autosomal recessive dis- ease characterized by soft connective tissue calcification[81,82]. For reasons yet unknown, BM is a preferred site for these ectopic calcifications and PXE patients often also develop eye pathology. In these patients, extensive calcification of the elastic fibres of BM makes the membrane brittle, and prone to breaks. The breaks are visible upon funduscopy as angioid streaks that radiate from the macula toward the periphery[33]. Ultimately, the breaks in BM lead to neovascularisation in these patients resulting in loss of visual acuity[83].

The mechanism of soft tissue calcification is currently the focus of intense scientific interest, especially in the cardiovascular system. These studies reveal a multitude of molecules and processes that can influence this process. It has become increasingly clear that control of calcification involves a delicate balance between pro-calcific and anti-calcific mediators[84]. Anti-calcific factors include molecules such as pyrophosphate, and several proteins such as fetuin A, Matrix Gla Protein (MGP) and vitamin K. Pro-calcific factors include high phos- phate levels, damaged extracellular matrix and cell death. Many different environmental and genetic factors may be involved in BM calcification. For PXE, a mouse model was made by disrupting the causative Abcc6 gene. Among other things, the PXE mice develop ectopic calcification in BM 85. Further elucidation of the calcification process holds the promise that soft tissue and BM calcification can perhaps be influenced by drugs or by dietary means[86] (unpublished observations in PXE mice, dietary influences in DCC mice;T. Gorgels[87]).

Iron depositions It is well known that (ab-)normal levels of iron ions contribute to various diseases of aging, including atherosclerosis, Alzheimer’s disease, Parkinson’s disease and retinal degenera- tion[88]. The subject of iron homeostasis and toxicity in retinal degeneration was recently and excellently reviewed elsewhere[89]. In summary, iron is essential for many metabolic processes, but can also cause damage through inducing (local) oxidative stress. An entire net- work of molecules, including metal receptors/transporters and ceruloplasmin, try to maintain (local) iron homeostasis: the balance between benefit and damage. However, with age, iron accumulates in the body. The resulting iron overload in the retina and RPE can cause retinal degeneration. Within the RPE cell, iron and related ferruginous compounds play an impor- tant role in the lysosome mediated build up of lipofuscin, and in general cellular oxidative

89 Chapter 4 stress, aging and apoptosis[90,91]. At the surface of BM, iron and other metal ions may play a role in the oligomerization of CFH molecules, thereby indirectly affecting the inhibition the complement cascade[92].

Zinc depositions In 1988, Newsome and co-workers[93] found a beneficial effect for oral zinc supplementa- tion in AMD. These findings were corroborated by a subsequent 6.5 year follow up study of the AREDS population[94]. In parallel, millimolar amounts of zinc were found in sub-RPE deposits and BM[95,96]. While the true relationship between the presence of zinc and AMD is undoubtedly very complex, Nan and co-workers[92] suggested, on the basis of a series of elegant experiments, that Zn is involved in the oligomerization of CFH, and consequently in AMD related complement regulation.

Advanced Glycation End products (AGEs) in BM Advanced glycation end products (AGEs) are chemically modified glycosylated or oxidized fats and proteins. Outside the body, they are produced by smoking or cooking. The dietary intake of exogenous AGEs may be related to the AGE serum levels, AGE accumulation in tissues, and ultimately total body damage by (per-) oxidative stress. Inside the body, AGEs can be produced by the combined metabolism of fat, proteins and sugar [97,98]. In general, cellular proteolysis of AGE releases AGE by-products into the serum which can be excreted in the urine. However, extracellular matrix proteins in the body are resistant to proteolysis. Consequently, AGEs accumulate preferentially on structural proteins, like col- lagens in BM, where they inhibit protein function and cause age related damage[99,100]. High concentrations of AGEs in serum or tissues activate the AGE receptor (RAGE), which is present on multiple cells in the body. Local activation of RAGEs frequently aggravates diseases like atherosclerosis, diabetic nephropathy, and neurodegeneration through inflam- matory and other mechanisms[101,102]. AGE accumulations containing pentosidine and carboxymethyllysine (CML) form age-pro- moting structures in BM, basal deposits, and choroid[103-106]. Indeed, Yamada et al (2006) found that the RPE showed more intense immuno-staining for the AGE receptors RAGE and AGER1 in areas containing basal deposits than in areas of normal BM[105]. In summary, these data suggests that both specific AGEs and AGE receptors are locally present in BM, basal deposits or drusen, and promote aging and/or the development of AMD [105].

Accumulation of lipids in BM As age increases, there is a progressive accumulation of lipids (phospholipids, triglycerides, fatty acids and free cholesterol) in BM, especially in the macular area[107]. Lipoprotein-like particle (LLPs) composition in BM resembles plasma LDL more than it does photoreceptor membrane composition[108]; moreover, these lipids are mainly derived from the RPE. Only a small proportion is of extracellular choroidal origin[45,109-111]. In young eyes, lipid inclusions, such as LLPs, small granules and membrane-like structures, are associated with fine elastin and collagen filaments in the ICL, EL and OCL. Huang and co-workers (2008) found that, once the EL and ICL were filled with particles, LLP continued

90 The dynamic nature of Bruch’s membrane to accumulate near the RPE, but did not increase in the OCL anymore[111]. Thus, with age, these lipid inclusions filled the interfibrillary spaces of the EL and accumulated in the ICL, forming the so-called lipid wall[26,45]. Holz et al (1994) observed that the macula of the elderly contained seven times higher con- centrations of cholesterol esters than the retinal periphery. However, these findings are not undisputed and perhaps the lipid wall thickness and content may even differ per individu- al[109]. Interestingly, Holz and others[107,109] also observed that the ratio of phospholipids to neutral fats varies per individual, perhaps in part depending on diet. The accumulation of lipids with increasing age, impairs the capacity of Bruch’s membrane to facilitate fluid and macromolecular exchange between the choroid and the RPE or vice versa[50], which is es- sential for normal retinal function. With age, small and large extracellular deposits, such as basal deposits and drusen (discussed below) slowly but surely appear in BM. These deposits contain large amounts of (un-) esteri- fied cholesterol, oxy-cholesterols, and many other lipid-based biomolecules.

BM thickening Throughout life, the ‘normal’ BM almost doubles in size[71]. In a comparative study of 120 human donor eyes, Ramrattan et al (1994) found that the overall thickness of BM shows a positive linear relationship with age. BM thickness increased from 2 µm in the first decade to 4.7 µm 80 years later[72]. The largest part of BM thickening occurs in the ICL, followed by the OCL. This process starts in the retinal periphery, where RPE gene expression of most structural components of BM appears to be higher than in the macular area[9,112]. In general, BM thickening is caused by increased deposition and cross-linking of (less solu- ble) collagen fibres and increased deposition of biomolecules, the majority being (oxidized) waste products of RPE metabolism. There is an age-related accumulation of granular, mem- branous, filamentous and vesicular material eventually resulting in focal deposits and drusen. Obviously, the thickening of BM eventually leads to several functional changes, such as changes in elasticity and hydraulic permeability.

Functional changes of the aging BM Recoil and decreased elasticity The elasticity of the BM-choroid complex decreases with age while the recoil capacity does not [33]. As discussed above, BM may loose much of its elasticity during life because of in- creased collagen cross-linking, calcification of the elastic fibres and AGE mediated oxidative stress damage. Overall, the decrease in elasticity and recoil is not exacerbated in AMD[33].

Decreased hydraulic permeability The normal diffusion properties of BM are discussed above. Age-related changes in BM, such as accumulation of (neutral) lipids, turnover of proteoglycans, as well as calcification most likely alter the biophysical properties of BM. Indeed, the overall water permeability of BM decreases with age primarily due to the changing properties of the inner half of BM[33].

91 Chapter 4

Basal deposits and drusen Nomenclature and classification Basal deposits Basal deposits are accumulations of waste material between the RPE and BM[113]. Two types of basal deposits exist, basal laminar deposits and basal linear deposits. Basal laminar deposits, located internal to the basement membrane of the RPE cells, contain granular ma- terial with collagen type IV, laminin, glycoproteins, glycosaminoglycans (chondroitin- and heparan sulphate), carbohydrates (N-acetylgalactosamine), cholesterol, and apolipoproteins B and E[33,113-115]. Basal linear deposits are located in the ICL and are electron-dense, containing phospholipid granules[115,116]. Basal linear deposits are stronger markers for progression to drusen and AMD than basal laminar deposits[113]. The nomenclature of basal deposits is confusing and authors have used several different terms for numerous deposits in different layers of BM in the past (reviewed by Marshall in 1998)[33]. The frequently used term “sub-RPE deposit”[117] is also unclear, since it does not clarify the exact location of the deposit below the RPE. In our view, the most straightfor- ward and simple classification is to use the term based on the layer(s) in which, or in between which, the deposits are detected: OCL deposits, ICL deposits, ICL-RPE-basement membrane deposits, etc. If the layer containing a deposit cannot be defined, it seems appropriate to sim- ply use the term BM deposit or basal deposit. The presence of basal deposits and subclinical drusen have been reported as early as in the third decade[118].

Drusen Drusen are extracellular deposits that form below the RPE in BM. Drusen initially appear in the macular area, but certainly also occur in the retinal periphery. Several types of drusen exist, they can be defined from a clinical, histological and molecular point of view. Clinically, drusen are defined according to their location, size and shape: ophthalmologists usually make a distinction between macular and peripheral drusen, small and large drusen, or drusen with defined (hard) and less well defined borders (soft and confluent drusen)[118,119]. The presence of soft, confluent drusen is a major risk factor for AMD[118]. When (confluent) drusen become visible by ophthalmoscopy, normal aging of the RPE and BM insidiously progresses to AMD pathology. Histologically, drusen can be described by their size, shape and PAS staining[113,114]. They can be seen as small yellow patches, initially in the macular area under the RPE. In the case of well-defined hard drusen, histological staining usually reveals local atrophy of the photo- receptors over clearly defined mounds beneath the RPE, solidly stained by PAS. In the case of less well defined soft drusen, which may become coalescent (confluent drusen), linear granu- lar bands can be observed locally, with a light PAS staining[113]. Lengyel and co-workers suggested that so called auto-fluorescent drusen are strongly associated with the lateral walls of the intercapillary pillars of the choriocapillaris[120]. The molecular classification of drusen is discussed below.

92 The dynamic nature of Bruch’s membrane

Molecular composition of drusen Drusen contain acute phase proteins, C-reactive protein, complement components, comple- ment inhibitors, apolipoproteins, lipids and many more proteins[113]. They vary in fat and cholesterol content, with a stable ratio between esterified cholesterol (EC) and unesterified cholesterol (UC). Frequently, drusen proteins are post-translationally modified[119]. A number of complement proteins were first immunolocalized to drusen by Hageman and co-workers[121]. Using a proteomics approach, Crabb et al (2002) subsequently identified 129 different proteins in drusen (Table 1)[6]. Sixty-five percent of these proteins were pres- ent in drusen from both non-affected and AMD donor eyes. The most common proteins in drusen of non-affected eyes were tissue metalloproteinase inhibitor 3, clusterin, vitronectin, and serum albumin. The presence of crystallin, and oxidatively modified proteins (TIMP3 and vitronectin) or lipids (docosahexaenoate-containing) in drusen suggested that oxidative stress is critical for drusen formation[6]. Next, a number of additional proteins were assigned to drusen, including the Amyloid Beta protein[122].

Table 1. Origin of drusen proteins: gene expression in the choroid or the RPE and the presence of proteins in serum.

Primary protein gene expression protein Sequence Name Sequence Code drusen chor>RPE RPE>chor serum ACTB NM_001101 + ACTG1 NM_001614 + ACTN1 NM_001102 + ALB NM_000477 + + ALDH1A1 NM_000689 + + + AMBP NM_001633 + ANXA1 NM_000700 + + + ANXA2 NM_004039 + + + ANXA5 NM_001154 + + + ANXA6 NM_001155 + APCS NM_001639 + APOA1 NM_000039 + APOA4 NM_000482 + APOE NM_000041 + APP NM_000484 + + ATP5A1 NM_004046 + + BFSP1 NM_001195 + BFSP2 NM_003571 + BGN NM_001711 + C3 NM_000064 + + +

93 Chapter 4

Primary protein gene expression protein Sequence Name Sequence Code drusen chor>RPE RPE>chor serum C5 NM_001735 + + C6 NM_000065 + C7 NM_000587 + C8B NM_000066 + C9 NM_001737 + CKB NM_001823 + + CLU NM_001831 + + COL1A2 NM_000089 + + + COL6A1 NM_001848 + + + COL6A2 NM_001849 + + + COL8A1 NM_001850 + + + CRYAA NM_000394 + CRYAB NM_001885 + + + CRYBA1 NM_005208 + CRYBA4 NM_001886 + CRYBB1 NM_001887 + CRYBB2 NM_000496 + CRYGB NM_005210 + CRYGC NM_020989 + CRYGD NM_006891 + CRYGS NM_017541 + CTSD NM_001909 + + + DIP2C NM_014974 + EFEMP1 NM_004105 + + + ELN NM_000501 + EPHX2 NM_001979 + FBLN5 NM_006329 + + + FGG NM_021870 + + FN1 NM_054034 + + + FRZB NM_001463 + + + GAPDH NM_002046 + + GPNMB NM_002510 + + H3F3A NM_002107 + HBA1 NM_000558 + + +

94 The dynamic nature of Bruch’s membrane

Primary protein gene expression protein Sequence Name Sequence Code drusen chor>RPE RPE>chor serum HBA2 NM_000517 + + + HIST1H1E NM_005321 + HIST1H2AE NM_021052 + + + HIST1H2BJ NM_021058 + + HIST1H2BL NM_003519 + + HIST1H4H NM_003543 + + HIST2H2AA3 NM_003516 + + + HIST2H2BE NM_003528 + HIST4H4 BC111093.1 + HP NM_005143 + IGHA1 AF067420 + + IGHG1 BC037361 + + + IGHG2 AAH62335 + IGHG3 AAH33178 + + IGHG3 ENST00000319391 + IGKC BC073779.1 + + LAMB2 NM_002292 + + LMNA NM_005572 + + + LTF NM_002343 + LUM NM_002345 + + LYZ NM_000239 + MFAP4 NM_002404 + + + MYH9 NM_002473 + + + MYL6 NM_079425 + OGN NM_033014 + + + ORM1 NM_000607 + PLA2G2A NM_000300 + + PLG NM_000301 + PRDX1 NM_002574 + PRELP NM_002725 + + PSMB5 NM_002797 + RBP3 NM_002900 + + RGR NM_002921 + + + RNASE4 NM_002937 + + +

95 Chapter 4

Primary protein gene expression protein Sequence Name Sequence Code drusen chor>RPE RPE>chor serum RPS14 NM_005617 + S100A7 NM_002963 + S100A8 NM_002964 + + S100A9 NM_002965 + + + SAA1 NM_000331 + + + SEMA3B NM_004636 + + SERPINA1 NM_000295 + SERPINA3 NM_001085 + + + SERPINF1 NM_002615 + + + SMC6 NM_024624 + + SPP2 NM_006944 + SPTAN1 NM_003127 + THBS4 NM_003248 + + TIMP3 NM_000362 + + + TNC NM_002160 + TUBA1A NM_006009 + TUBA1B NM_006082 + TUBA1C NM_032704 + TUBA3C NM_006001 + TUBB NM_178014 + TUBB2C NM_006088 + + TUBB3 NM_006086 + + TYRP1 NM_000550 + + UBB NM_018955 + VIM NM_003380 + + + VTN NM_000638 + Drusen proteins identified by Crabb et al.[6] and confirmed by Ingenuity analysis[126] to have a sequence code, see also Figure 6. Plus: gene expression levels were found to be at least 2.5-fold higher in the choroid compared to the RPE (chor>RPE) of the same donor eye, in triplicate microarray measurements from three older healthy humans[24]. Gene expression levels at least 2.5-fold higher in the RPE than the choroid are found in the column RPE>chor, indicated by a plus[24]. Plus in the serum column indicates proteins in serum identified by Ingenuity analysis.

96 The dynamic nature of Bruch’s membrane

Figure 5. Overlap in protein content of Alzheimer plaques, AMD drusen, and atherosclerotic plaques. 7 (6%) shared proteins: Amyloid (beta, P), APOE, C3, CLU, FGG, VTN

Alzheimer plaque AMD drusen Atherosclerotic plaque

a j

d f k e i b l c g h

24 (20%) shared proteins: 10 (8%) shared proteins: Alpha B-Crystallin, Amyloid (beta, Amyloid (beta, P), APCS, P), APOE, ATP5A1, complement Apolipoprotein (A1, E), comple- (C3, C7), CLU, FGG, HP, IGHG1, ment C3, CLU, FGG, PLG, VTN IGKC, OGN, PRDX1, Serpin (A1, A3) TIMP3, Tubulin (A1A, A1C, A3C, B2C, B3), UBB, VTN

We analyzed 121 proteins based on the article by Crabb et al[6] and additional literature searches. We obtained NM numbers for 85 of the proteins characterized by Crabb, an additional 36 proteins were added based on the recent literature[167]. Note that drusen share 20 % of their protein content with AD plaques and less than 10 % with ath- erosclerotic plaques. Alzheimer plaque picture courtesy of Dr. I. Huitinga, Netherlands Brain Bank, donor number nhb:2006-060,VU:S06/189. Atheroslcerotic plaque picture courtesy of Dr. P Sampaio Gutierrez. a. Amyloid deposits, b. Cell nucleus, c. Enthorinal cortex, d. Retinal Pigment Epithelium, e. Bruch’s Membrane, f. Druse, g. Choroid, h. Choroidal bloodvessel, i. Intima with atheroslerotic plaque, j. Tunica media, k. Fibrous cap, l. Lipid + necrotic core. A color version of this figure can be found in chapter 15.

Given their origin, location, and pathobiological involvement, one could possibly consider drusen a mixture between atherosclerotic plaques[123] and Alzheimer (AD) plaques[124,125]. However, comparison of the known molecular constituents of these three extracellular de- posits (see Figure 5) showed that they only share seven proteins (Amyloid (beta, P), APOE, C3, CLU, FGG and VTN). Drusen and AD plaques have 24 known molecular constituents in common. In contrast, drusen share only 10 known molecular components with athero- sclerotic plaques. Therefore, we hypothesize that drusen resemble AD plaques more than atherosclerotic plaques.

Drusen: where do they come from? A priori, drusen constituents are most likely derived from (modified) fats and proteins pro- duced by the 1) photoreceptor cells, and/or 2) RPE cells, and/or 3) choroidal cells (endo- thelial cells, fibroblasts, smooth muscle cells), or 4) derived from serum constituents. As discussed above, the majority of lipids in drusen, including (oxidized) cholesterol are derived from the RPE and photoreceptors cells, and only a small part from the serum.

97 Chapter 4

To gain insight into the question: “where do drusen proteins come from?”, we compared drusen protein content (modified from Crab et al 2002)[6] with triplicate mRNA expression profiles from human photoreceptor cells, RPE cells and choroidal cells[24] and a proteomics profile from human serum generated by the Ingenuity knowledge database[126] (Bergen un- published results) (Figure 6 and Table 1).

Figure 6. Schematic drawing of proteins in drusen and the corresponding genes expressed in adjacent cell layers.

3 genes

Phot

13 genes RPE

113 proteins

BM

41 genes

Chor

36 proteins

We could annotate 113 genes with genbank codes (using Ingenuity) which correspond to drusen proteins identified by Crabb et al.[6]. Thirty six of these genes had higher expression levels in choroid compared to RPE (chor>RPE), thirteen genes had higher expression levels in RPE than in choroid (RPE>chor) and only three genes had higher expression levels in photoreceptors than RPE (phot>RPE). Gene expression levels were determined by RNA mi- croarray study comparing gene expression levels from two adjacent tissue types from the same donor. Experiments were performed in triplicate (on three different healthy older human donor eyes)[24]. Details of the 113 genes can be found in Table 1. A color version of this figure can be found in chapter 15.

By doing so, we can track, within obvious limitations, the main potential origin of the dru- sen proteins. At least 23 (20 %) of 113 proteins identified in drusen are present in serum[6]. Thirty-six (32 %) drusen proteins are potentially (also) synthesized by local choroidal cells. Thirteen (12 %) drusen proteins are potentially (also) derived from RPE cells and three (3 %) (also) from the photoreceptor cells. For the sake of argument, we assumed - and of course this is an oversimplification- that the amount of mRNA expression in cells adjacent to BM is, in general, linear with the amount of protein produced and subsequently transported to drusen. In that case, it is remarkable that the larger part of drusen proteins appear to be derived from the choroidal cells and or serum, and not from the photoreceptors (Figure 6). In summary, and perhaps surprisingly, human drusen consist of 1) lipids primarily derived from the photoreceptor cells and serum, and 2) proteins apparently primarily derived from choroidal cells and serum. Finally, it must be pointed out that the type and relative amount of molecular constituents of drusen, and patho-biological effects frequently do not have a linear relationship. For example, a minor fraction of serum-derived molecules from the comple-

98 The dynamic nature of Bruch’s membrane ment cascade can have large functional or pathological consequences in terms of function or pathology.

Why do drusen develop preferentially in the macular area? It is currently not known why drusen develop mainly in the macular area. However, a combi- nation of specific structural, molecular and functional properties may predispose the macula to develop drusen. First, the extremely high density of photoreceptors, particularly in the perifoveal ring, may play a role[127]. Local phagocytosis of photoreceptor outer segments (POS) by the RPE causes a highly focussed and localized peak of oxidative stress, and focal build-up of membraneous waste products. In addition, the specific structural properties of BM in the macular area may also play a role. As discussed above, BM in the macula has a thinner elastic layer and a more open maze compared to the periphery. Initially, in the still healthy eye, the macular RPE and BM may get rid of an excess of oxidized molecules and neutral fats (by transporting them toward the bloodstream rapidly. Most likely, however, additional proteins reach BM from the choroidal side. After oxidative modification, a subset of these proteins may get physically or chemical- ly trapped in BM, thereby initiating the first events of drusen formation in the macular area. Finally, local functional macular RPE properties, as annotated by gene expression profiles, according to van Soest (2007) and Booij (2009) may also play a role[9,11]. Most importantly, we compared the data of van Soest[9] and Crabb[6]. Van Soest and coworkers identified 438 genes (out of 22.000), that were significantly differentially expressed in macular RPE com- pared to RPE in the retinal periphery. Crabb et al. identified 129 proteins in drusen using a proteomics approach[6]. Interestingly, the overlap between these two datasets consists of 16 genes, while by chance alone this overlap would be no greater than 2.5 genes (Bergen et al. manuscript in preparation).

Bruch’s membrane and AMD pathology Oxidative stress in the RPE and its effects on BM Multiple studies in man, AMD animal models and (RPE) cell lines point toward an important role of oxidative stress in the development of AMD[119]. Epidemiological studies in man demonstrated that smoking is associated with increased risk of AMD. Smoking is a well known source of oxidative stress[118]. A decreased risk of AMD was found with the use of high dietary antioxidants, such as lutein and zeaxanthin or vitamin C and Beta-carotene[128,129]. Animal models susceptible to oxidative stress, like SOD1-/- [130] and ERCC6-/- knock-out mice[131] show remarkable signs of retinal degeneration, if not AMD. Finally, using RPE cell lines, several investigators showed that oxidative stress is implicated in retinal degeneration[132,133]. In the macula, where photoreceptor density is high and incoming light is focussed, the RPE and BM are highly susceptible to high levels of oxidative stress. Sources of local oxidative stress include the high metabolic rate of photoreceptors required to sustain their normal func- tion and structural renewal, the exposure to light, the high local oxygen pressure and the high metabolic rate of the RPE due to processing of photoreceptor outer segments (POS)[79].

99 Chapter 4

The combination of high levels of oxidative stress and segmental POS digestion in the RPE most likely results in the oxidative modification of lipid-related molecules, such as choles- terol[134] and docosahexanoic fatty acid that accumulate in drusen or are exported to the bloodstream through BM. In addition, a vast number of molecular constituents that can nei- ther be digested, nor exported across the plasma membrane, accumulate inside the RPE cell. A well known example is the accumulation of the bisretinoid A2E, an indigestible remnant of POS, and an important constituent of intracellular lipofuscin[135,136]. Finally, the RPE cells attempt to export unneeded or indigestible residual molecules basolaterally, where they accumulate in BM, BM BLDs, or drusen, or diffuse to the bloodstream[33].

Complement activation, inflammation and the immune response The recent finding of genetic associations between AMD and genes from the complement system (CFH[137-140], C2/FB and C3[79,141,142) established the long suspected role of the innate immune system in AMD[121,143]. The detailed role of the regulation of the com- plement system and all complement factors individually has recently been reviewed else- where[119,144]. Multiple immune-related cells, including macrophages, fibroblasts, and lymphocytes have been implicated in RPE atrophy, the breakdown of BM, and neovascularisation in AMD[143]. In the healthy and balanced situation (i.e. in the absence of AMD), the (alternative) comple- ment system is activated just sufficiently by foreign antigens to clear up debris in BM, while at the same time invoking relatively little RPE cell cellular damage through the membrane attack complex. But what is the trigger that activates this pathway? Zhou et al found that intermediates of li- pofuscin in the RPE, like the bisretinoid pigment A2E, were recognized as non-self antigens, and activated the complement cascade[145]. Alternatively, Johnson et al. co-localized “the Alzheimer protein” amyloid beta (Aβ) which activated complement components in drusen, and suggested that Aβ invokes a local inflammatory response[122]. Recently, the latter data were substantiated by functional studies by Wang and colleagues who showed that Aβ in- teraction with complement factor I activated the complement cascade[146]. Subsequently, Hollyfield and co-workers showed that oxidative modification of docosahexanoic acid,a polyunsaturated fatty acid abundantly present in the retina, resulted in a unique fragment, carboxyethylpyrrole, that can also invoke a local immune response[147]. To further com- plicate the issue, Scholl et al. reported that not only local factors, but also systemic comple- ment serum factors can activate the alternative complement pathway in the RPE/BM[148]. Complement factors, such as CFH, and C3 are expressed in the RPE and occur in serum, while others only occur in serum[148]. In summary, it is likely that multiple non-self antigen triggers can invoke a local complement/ immune response leading to AMD. The local complement response may therefore initially be determined by both the type and the amount of non-self-antigen present. In addition to the type and amount of trigger, the actual activity of the complement cascade is regulated by both genetic variation in, and biochemical interaction between a number of regulatory proteins, such as CFH, MCP (CD46), C3 and Factor I. These proteins contain complement control re- peats (CCPs) that can bind to other complement factors and/or substances like CRP, heparin and heparan sulphate present at the surface of BM. Through more or less effective binding

100 The dynamic nature of Bruch’s membrane of these regulatory protein domains, due to DNA sequence variation, metal ion traces, post translational protein modifications or simply by bio-availability of regulatory proteins, the actual activity of the complement system is controlled[144]. It is of interest to note here that BM’s proteoglycan turnover and content appears to change with age (discussed above) and disease. For example, Landers et al. found an increased hepa- ran sulphate content in retinas of animal models affected by retinal degeneration[149]. These changes may modify the natural binding characteristics and inhibitory complement capacity of CFH or other complement molecules at the surface of BM with age. This may be one of the factors that determine the rate of photoreceptor cell loss during aging and neural degen- eration. The interaction of local and systemic complement factors, their regulatory binding to extracellular matrix components, growth factors, and other molecules at the BM interface is, so far, poorly understood, and currently subject to thorough investigation[147,150].

A breach in the barrier: choroidal neovascularization (CNV) The whole process of choroidal neovascularization (CNV) has recently been reviewed else- where[151]. In summary, CNV is a process whereby new vessels sprout from the choroid and penetrate BM. In many aspects, CNV resembles normal wound healing in the skin[151].

CNV is controlled by local pro- and anti-angiogenic factors. Among these factors is a combi- nation of proteins secreted by the RPE and/or choroidal cells, including well studied growth factors like VEGF-A and PEDF. Another factor is the physical barrier embodied by the RPE/ BM complex. Leukocytes, lymphocytes, macrophages and endothelial cells may directly or indirectly degrade BM through the breakdown of collagen, thereby facilitating neovascu- larization[152]. Nevertheless, new vessels can also penetrate the intact BM[152]. Specific subtypes of macrophages, mononuclear phagocytic series (MPS) cells, have been implicated in the development of new vessels in healthy and AMD eyes[153]. Furthermore, multiple neovascularization studies in normal and genetically modified mouse models support the notion that additional factors, such as the breakdown of BM integrity, are essential for the induction of CNV. To illustrate the complexity of this issue, we present in Table 2 the genes that are expressed at higher levels in the RPE than in the choroid with the functional an- notation ‘angiogenesis’[126]. (Bergen and Booij; unpublished results). Analysis of our data showed that at least 23 genes may be involved in this process. If the local balance between pro- and anti-angiogenic factors is disturbed substantially in favour of VEGF, CNV or en- hanced fibrosis may occur[154]. For further illustration purposes, Table 3 shows the genes that are expressed at higher levels in the choroid than in the RPE with the functional annota- tion ‘angiogenesis’[24].

101 Chapter 4

Table 2. The overlap between genes associated with the term angiogenesis (Ingenuity) and genes identified in triplicate RNA microarray measurements from older healthy human donor eyes[24] with expression levels higher in the RPE than the choroid.

Gene name Sequence code BAI1 NM_001702 EGF NM_001963 EPAS1 NM_001430 EPHA2 NM_004431 FLT1 NM_002019 HGF BC022308 HMMR NM_012484 IGF1R NM_000875 IGHG1 AF035027 IL2 NM_000586 IL18 NM_001562 INSR NM_000208 MAPK8 AK125150 MFGE8 NM_005928 MMP9 NM_004994 NOS1 NM_000620 NOS3 NM_000603 PLG NM_000301 PTHLH M31157 SERPINF1 NM_002615 SOD1 NM_000454 TP73 NM_005427 VEGFA NM_003376

Table 3. The overlap between genes associated with the term angiogenesis (Ingenuity) and genes identified in triplicate RNA microarray measurements from older healthy human donor eyes[24] with expression levels higher in the choroid than the RPE.

Gene name Sequence code ALOX5 NM_000698 ANPEP NM_001150 APOE NM_000041 C3 NM_000064 C5 NM_001735

102 The dynamic nature of Bruch’s membrane

Gene name Sequence code C3AR1 NM_004054 CCL13 NM_005408 CFB NM_001710 COL18A1 NM_030582 CSF2 NM_000758 CXCL12 NM_000609 CYR61 NM_001554 EFNA1 NM_004428 HGF NM_000601 HMMR BC035392 ICAM1 NM_000201 IGF1 NM_000618 IGF2 NM_000612 IGFBP3 NM_000598 IGHG1 BC037361 IL13 NM_002188 INHBA NM_002192 ITGB2 NM_000211 LEP NM_000230 MAPK8 AL137667 MYC NM_002467 NR3C1 NM_000176 NRP1 NM_003873 PLAU NM_002658 PTGS2 NM_000963 S100A4 NM_002961 SCYE1 NM_004757 THBS2 NM_003247 TIMP2 NM_003255 TNF NM_000594 TP53 NM_000546 VCAM1 NM_001078 VEGFC NM_005429

103 Chapter 4

Outlook and perspectives: Toward rational, genomics-driven, molecular thera- pies for AMD Summarizing the events leading up to early AMD The molecular pathology of AMD has recently been reviewed extensively elsewhere[119]. In summary, Ding et al (2009) reviewed clinical, epidemiological, and genetic aspects of AMD, as well as the use of mouse models for potential AMD therapy. In contrast, we here reviewed the central role of BM in normal retinal aging, in drusen formation and in the early stages of AMD. Obviously, the normal function and pathology of BM can only be understood in the con- text of the molecular and cellular events involving the adjacent cell layers (photoreceptor, RPE and choroid) as well as systemic factors (from serum). As discussed, many normal ag- ing processes affect BM, such as thickening of its layers due to fat deposition, calcification of the EL, oxidative stress and drusen formation. Clearly, these normal (subclinical) aging events may predispose BM and the RPE to disease, especially in the macular area. BM is the key acellular tissue involved in the development of age-related macular degeneration. Its extracellular matrix, heavily dominated by heparan sulphate, appears to be the regulatory playground of both local and systemic interactions involving complement activators, proteo- glycans, chemokines, cytokines, growth factors and, above all, toxic waste products. In the healthy, non-AMD, situation these molecular interactions may follow a fixed pattern, which maintains the local homeostasis. This local homeostasis in each individual probably depends on, and is limited by environmental factors, genetic constitution and local anatomy of the neural retina, RPE and BM. However, fuelled by changes due to normal aging, such as prolonged oxidative stress and immune activation, the molecular interactions at the sur- face of BM change. These changes may be accommodated until local homeostasis cannot be maintained anymore, which ultimately leads to the devastating clinical manifestations of AMD.

Prevention and therapy of AMD; is there a role for BM biology? So, with the current state-of-the-art knowledge, can we “prevent” or “cure” AMD? Over the last decades our understanding of environmental risk factors as well as molecular, cellular, and even systemic events underlying AMD has grown tremendously. In summary, environ- mental risk factors now include smoking, diet and perhaps light exposure[118]. Dietary intake of saturated fats increases the risk of AMD[155]. Intake of anti-oxidants (lutein, zeaxanthin, Beta carotene, vitamin C), omega-3 polyunsaturated fatty acids, (nuts, fish) and zinc supple- ments may be beneficial[94]. Genetic association studies were successfully used to implicate several AMD disease genes (APOE, CFH, C3, C2/BF, HTRA1/ARMS2)[148], or to identify potential candidate disease genes (CXCR3, Il-8, ERCC6)[119]. Genetic and functional stud- ies were instrumental to the discovery of functional pathways in AMD. The most important are, as discussed above, fat metabolism, oxidative stress, complement activation, and STAT3/ VEGF induced neovascularization. So far, prevention and therapeutic efforts have largely focussed on these four pathways.

104 The dynamic nature of Bruch’s membrane

By far the best option is, of course, to prevent AMD; that is to aim to postpone its onset or to slow its progression. Most likely, for a large majority of individuals, this can be done by avoiding risk factors for AMD. Although not every individual would benefit equally due to differences in their genetic constitution[156], one should quit smoking, wear sunglasses, and change to a diet that contains sufficient zinc, anti-oxidants and unsaturated fatty acids. Alternatively, is it possible to influence the progression of AMD by drugs? Several negative and positive developments can be noted. Previously, serum, lipid lowering drugs, like statins, were used to treat AMD, but the outcome of clinical trials were variable[157]. In addition, the obvious idea of local manipulation of the complement system in AMD by complement inhibitors may reduce neovascularization[158], but is not without risk: It may turn an essen- tially useful chronic inflammation at the RPE-BM interface into a harmful acute inflamma- tion.On the positive side, a CNTF trial is ongoing that aims to supplement the photoreceptors with small quantities of CNTF over a prolonged period of time[159]. In this case, the pho- toreceptors remain more viable, which may delay disease onset. In addition, for dry AMD, a drug called fenretinide, which halts the accumulation of retinol-related toxins, thought to be involved in vision loss, shows promise[160,161]. Finally, VEGF based treatments have, of course, been relatively successful in treating the wet form of AMD[162].

These new (experimental) therapies focus on cells (photoreceptors, RPE) or on systemic factors (e.g. statins). However, aside from studies by Del Priore et al[163] and Marshall (un- published), curiously enough, so far, little attention has been paid to BM, the prime site of AMD development. As discussed, BM not only plays a key role in normal aging of the pho- toreceptor-RPE-BM-choroid complex, but is also essentially involved in pathogenic effects of fat metabolism, oxidative stress, complement activation, and neovascularization. Even a number of structural BM genes were identified as AMD genes (Fibl-3 (EFEMP-1), Fibl-5, Fibl-6, CTRP5)[79]. For at least two of these genes, a corresponding animal model (CTPR5 -/-)[164], Fibl-3[165] showing AMD like features, is available. At least, in terms of local BM therapy, there are two options: 1) removal of pathogenic or non-self compounds from BM which slowly accumulate during aging and disease; or 2) “medical bioremediation”[166]: the use of microbial enzymes to augment or restore missing, or failing, metabolic functions. However, both approaches have been largely unsuccessful up till now. Further dissection and definition of the molecular events involved in age-related BM changes is therefore warranted to successfully develop drugs: To end with a sentence from the beginning of this article: For too long many investigators have considered BM to be a relatively boring and simple sheet of extracellular matrix, merely occupying space between the retinal pigment epithelium (RPE) and the choroid. This is about to change.

105 Chapter 4

References 1. Takei Y, Ozanics V. Origin and development of Bruch’s membrane in monkey fetuses: an electron microscopic study. Invest Ophthalmol 1975; 14: 903-916. 2. Olson MD. Development of Bruch’s membrane in the chick: an electron microscopic study. Invest Ophthalmol Vis Sci 1979; 18: 329-338. 3. Campochiaro PA, Jerdon JA, Glaser BM. The extracellular matrix of human retinal pig- ment epithelial cells in vivo and its synthesis in vitro. Invest Ophthalmol Vis Sci 1986; 27: 1615-1621. 4. Guymer R, Luthert P, Bird A. Changes in Bruch’s membrane and related structures with age. Prog Retin Eye Res 1999; 18: 59-90. 5. Hogan M, Alvarado J. Studies on the human macula. IV. Aging changes in Bruch’s membrane. Arch Ophthalmol 1967; 77: 410-420. 6. Crabb JW, Miyagi M, Gu X et al. Drusen proteome analysis: an approach to the etiology of age-related macular degeneration. Proc Natl Acad Sci U S A 2002; 99: 14682-14687. 7. Klein R, Peto T, Bird A, Vannewkirk MR. The epidemiology of age-related macular degeneration. Am J Ophthalmol 2004; 137: 486-495. 8. Curcio CA, Presley JB, Malek G et al. Esterified and unesterified cholesterol in drusen and basal deposits of eyes with age-related maculopathy. Exp Eye Res 2005; 81: 731- 741. 9. van Soest SS, de Wit GM, Essing AH et al. Comparison of human retinal pigment epithelium gene expression in macula and periphery highlights potential topographic differences in Bruch’s membrane. Mol Vis 2007; 13: 1608-1617. 10. Bhutto IA, Uno K, Merges C et al. Reduction of endogenous angiogenesis inhibitors in Bruch’s membrane of the submacular region in eyes with age-related macular degenera- tion. Arch Ophthalmol 2008; 126: 670-678. 11. Booij JC, van Soest S, Swagemakers SMA et al. Functional annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 2009; 10: 164. 12. Stone EM, Kimura AE, Folk JC et al. Genetic linkage of autosomal dominant neovas- cular inflammatory vitreoretinopathy to chromosome 11q13. Hum Mol Genet 1992; 1: 685-689. 13. Pastor JC, de la Rua ER, Martin F. Proliferative vitreoretinopathy: risk factors and pathobiology. Prog Retin Eye Res 2002; 21: 127-144. 14. Aessopos A, Stamatelos G, Savvides P et al. Angioid streaks in homozygous beta thal- assemia. Am J Ophthalmol 1989; 108: 356-359. 15. van Soest S, Swart J, Tijmes N et al. A locus for autosomal recessive pseudoxanthoma elasticum, with penetrance of vascular symptoms in carriers, maps to chromosome 16p13.1. Genome Res 1997; 7: 830-834. 16. McKusick VA, Traisman HS, Bianchine JW. More speculation on Marfan syndrome. J Pediatr 1972; 80: 530-531. 17. Appel A, Horwitz AL, Dorfman A. Cell-free synthesis of hyaluronic acid in Marfan syndrome. J Biol Chem 1979; 254: 12199-12203. 18. Beighton P, de Paepe A, Danks D et al. International Nosology of Heritable Disorders of Connective Tissue, Berlin, 1986. Am J Med Genet 1988; 29: 581-594.

106 The dynamic nature of Bruch’s membrane

19. Sorsby A, Mason ME. A fundus dystrophy with unusual features. Br J Ophthalmol 1949; 33: 67-97. 20. Capon MR, Marshall J, Krafft JI et al. Sorsby’s fundus dystrophy. A light and electron microscopic study. Ophthalmology 1989; 96: 1769-1777. 21. Ayyagari R, Griesinger IB, Bingham E et al. Autosomal dominant hemorrhagic macular dystrophy not associated with the TIMP3 gene. Arch Ophthalmol 2000; 118: 85-92. 22. Roberts JM, Forrester JV. Factors affecting the migration and growth of endothelial cells from microvessels of bovine retina. Exp Eye Res 1990; 50: 165-172. 23. Hirabayashi Y, Fujimori O, Shimizu S. Bruch’s membrane of the brachymorphic mouse. Med Electron Microsc 2003; 36: 139-146. 24. Booij JC, ten Brink JB, Swagemakers SMA et al. A new strategy to identify and anno- tate human RPE-specific gene expression. 2009; 25. Hogan M. Ultrastructure of the choroid. Its role in the pathogenesis of chorioretinal dis- ease. Transactions of the Pacific Coast Oto-ophthalmological Society Annual Meeting 1961; 42: 61-87. 26. Guymer R, Bird AC. Bruch’s membrane, drusen and age-related macular degeneration. 1998; 693-703. 27. Schachern PA, Paparella MM, Duvall AJ, III, Choo YB. The human round window membrane. An electron microscopic study. Arch Otolaryngol 1984; 110: 15-21. 28. Lamme EN, de Vries HJ, van VH et al. Extracellular matrix characterization during healing of full-thickness wounds treated with a collagen/elastin dermal substitute shows improved skin regeneration in pigs. J Histochem Cytochem 1996; 44: 1311-1322. 29. Chen C, Wu L, Jiang F, Liang J, Wu DZ. Scotopic sensitivity of central retina in early age-related macular degeneration. Yan Ke Xue Bao 2003; 19: 15-19. 30. Aisenbrey S, Zhang M, Bacher D et al. Retinal pigment epithelial cells synthesize lami- nins, including laminin 5, and adhere to them through alpha3- and alpha6-containing integrins. Invest Ophthalmol Vis Sci 2006; 47: 5537-5544. 31. Pauleikhoff D, Zuels S, Sheraidah GS et al. Correlation between biochemical composi- tion and fluorescein binding of deposits in Bruch’s membrane. Ophthalmology 1992; 99: 1548-1553. 32. Hewitt AT, Nakazawa K, Newsome DA. Analysis of newly synthesized Bruch’s mem- brane proteoglycans. Invest Ophthalmol Vis Sci 1989; 30: 478-486. 33. Marmor MF, Wolfensberger TJ. The retinal pigment epithelium. 1998; 34. Chong NH, Keonin J, Luthert PJ et al. Decreased thickness and integrity of the macular elastic layer of Bruch’s membrane correspond to the distribution of lesions associated with age-related macular degeneration. Am J Pathol 2005; 166: 241-251. 35. Narayanan AS, Page RC. Biosynthesis and regulation of type V collagen in diploid hu- man fibroblasts. J Biol Chem 1983; 258: 11694-11699. 36. Marshall GE, Konstas AG, Reid GG, Edwards JG, Lee WR. Collagens in the aged hu- man macula. Graefes Arch Clin Exp Ophthalmol 1994; 232: 133-140. 37. Altunay H. Fine structure of the retinal pigment epithelium, Bruch’s membrane and choriocapillaris in the camel. Anat Histol Embryol 2007; 36: 116-120.

107 Chapter 4

38. Sivaprasad S, Bailey TA, Chong VN. Bruch’s membrane and the vascular intima: is there a common basis for age-related changes and disease? Clin Experiment Ophthal- mol 2005; 33: 518-523. 39. Braekevelt CR. Fine structure of the choriocapillaris, Bruch’s membrane and retinal epithelium in the sheep. Anat Embryol (Berl) 1983; 166: 415-425. 40. Ishibashi K, Tian J, Handa JT. Similarity of mRNA phenotypes of morphologically nor- mal macular and peripheral retinal pigment epithelial cells in older human eyes. Invest Ophthalmol Vis Sci 2004; 45: 3291-3301. 41. Bowes Rickman C., Ebright JN, Zavodni ZJ et al. Defining the human macula transcrip- tome and candidate retinal disease genes using EyeSAGE. Invest Ophthalmol Vis Sci 2006; 47: 2305-2316. 42. Bok D. The retinal pigment epithelium: a versatile partner in vision. J Cell Sci Suppl 1993; 17: 189-195. 43. Strauss O. The retinal pigment epithelium in visual function. Physiol Rev 2005; 85: 845-881. 44. Wimmers S, Karl MO, Strauss O. Ion channels in the RPE. Prog Retin Eye Res 2007; 45. Huang JD, Presley JB, Chimento MF, Curcio CA, Johnson M. Age-related changes in human macular Bruch’s membrane as seen by quick-freeze/deep-etch. Exp Eye Res 2007; 85: 202-218. 46. Moore DJ, Hussain AA, Marshall J. Age-related variation in the hydraulic conductivity of Bruch’s membrane. Invest Ophthalmol Vis Sci 1995; 36: 1290-1297. 47. Starita C, Hussain AA, Pagliarini S, Marshall J. Hydrodynamics of ageing Bruch’s membrane: implications for macular disease. Exp Eye Res 1996; 62: 565-572. 48. Starita C, Hussain AA, Patmore A, Marshall J. Localization of the site of major resis- tance to fluid transport in Bruch’s membrane. Invest Ophthalmol Vis Sci 1997; 38: 762- 767. 49. Hillenkamp J, Hussain AA, Jackson TL, Cunningham JR, Marshall J. The influence of path length and matrix components on ageing characteristics of transport between the choroid and the outer retina. Invest Ophthalmol Vis Sci 2004; 45: 1493-1498. 50. Moore DJ, Clover GM. The effect of age on the macromolecular permeability of human Bruch’s membrane. Invest Ophthalmol Vis Sci 2001; 42: 2970-2975. 51. Fisher RF. The influence of age on some ocular basement membranes. Eye 1987; 1 ( Pt 2): 184-189. 52. Del Priore LV, Tezel TH. Reattachment rate of human retinal pigment epithelium to lay- ers of human Bruch’s membrane. Arch Ophthalmol 1998; 116: 335-341. 53. Del Priore LV, Geng L, Tezel TH, Kaplan HJ. Extracellular matrix ligands promote RPE attachment to inner Bruch’s membrane. Curr Eye Res 2002; 25: 79-89. 54. Gong J, Sagiv O, Cai H, Tsang SH, Del Priore LV. Effects of extracellular matrix and neighboring cells on induction of human embryonic stem cells into retinal or retinal pigment epithelial progenitors. Exp Eye Res 2008; 86: 957-965. 55. Tezel TH, Del Priore LV. Repopulation of different layers of host human Bruch’s mem- brane by retinal pigment epithelial cell grafts. Invest Ophthalmol Vis Sci 1999; 40: 767- 774.

108 The dynamic nature of Bruch’s membrane

56. Tezel TH, Del Priore LV, Kaplan HJ. Reengineering of aged Bruch’s membrane to en- hance retinal pigment epithelium repopulation. Invest Ophthalmol Vis Sci 2004; 45: 3337-3348. 57. Gullapalli VK, Sugino IK, Van Patten Y, Shah S, Zarbin MA. Impaired RPE survival on aged submacular human Bruch’s membrane. Exp Eye Res 2005; 80: 235-248. 58. Marshall GE, Konstas AG, Reid GG, Edwards JG, Lee WR. Type IV collagen and laminin in Bruch’s membrane and basal linear deposit in the human macula. Br J Oph- thalmol 1992; 76: 607-614. 59. Fang IM, Yang CH, Yang CM, Chen MS. Overexpression of integrin alpha6 and beta4 enhances adhesion and proliferation of human retinal pigment epithelial cells on layers of porcine Bruch’s membrane. Exp Eye Res 2009; 88: 12-21. 60. Wang H, Ninomiya Y, Sugino IK, Zarbin MA. Retinal pigment epithelium wound heal- ing in human Bruch’s membrane explants. Invest Ophthalmol Vis Sci 2003; 44: 2199- 2210. 61. Binder S, Stanzel BV, Krebs I, Glittenberg C. Transplantation of the RPE in AMD. Prog Retin Eye Res 2007; 26: 516-554. 62. Crane IJ, Liversidge J. Mechanisms of leukocyte migration across the blood-retina bar- rier. Semin Immunopathol 2008; 30: 165-177. 63. Xu H, Forrester JV, Liversidge J, Crane IJ. Leukocyte trafficking in experimental auto- immune uveitis: breakdown of blood-retinal barrier and upregulation of cellular adhe- sion molecules. Invest Ophthalmol Vis Sci 2003; 44: 226-234. 64. Xu H, Manivannan A, Goatman KA et al. Reduction in shear stress, activation of the endothelium, and leukocyte priming are all required for leukocyte passage across the blood--retina barrier. J Leukoc Biol 2004; 75: 224-232. 65. Albini TA, Wang RC, Reiser B et al. Microglial stability and repopulation in the retina. Br J Ophthalmol 2005; 89: 901-903. 66. Xu H, Chen M, Mayer EJ, Forrester JV, Dick AD. Turnover of resident retinal microglia in the normal adult mouse. Glia 2007; 55: 1189-1198. 67. Curcio CA, Millican CL, Allen KA, Kalina RE. Aging of the human photoreceptor mosaic: evidence for selective vulnerability of rods in central retina. Invest Ophthalmol Vis Sci 1993; 34: 3278-3296. 68. Jackson GR, Owsley C, Curcio CA. Photoreceptor degeneration and dysfunction in ag- ing and age-related maculopathy. Ageing Res Rev 2002; 1: 381-396. 69. Danciger M, Yang H, Ralston R et al. Quantitative genetics of age-related retinal degen- eration: a second F1 intercross between the A/J and C57BL/6 strains. Mol Vis 2007; 13: 79-85. 70. Beatty S, Koh H, Phil M, Henson D, Boulton M. The role of oxidative stress in the pathogenesis of age-related macular degeneration. Surv Ophthalmol 2000; 45: 115-134. 71. van der Schaft TL, Mooy CM, de Bruijn WC et al. Histologic features of the early stages of age-related macular degeneration. A statistical analysis. Ophthalmology 1992; 99: 278-286. 72. Ramrattan RS, van der Schaft TL, Mooy CM et al. Morphometric analysis of Bruch’s membrane, the choriocapillaris, and the choroid in aging. Invest Ophthalmol Vis Sci 1994; 35: 2857-2864.

109 Chapter 4

73. Call TW, Hollyfield JG. Sulfated proteoglycans in Bruch’s membrane of the human eye: localization and characterization using cupromeronic blue. Exp Eye Res 1990; 51: 451- 462. 74. Inatani M, Tanihara H. Proteoglycans in retina. Prog Retin Eye Res 2002; 21: 429-447. 75. Hewitt AT, Newsome DA. Altered synthesis of Bruch’s membrane proteoglycans as- sociated with dominant retinitis pigmentosa. Curr Eye Res 1985; 4: 169-174. 76. Rops AL, van der Vlag J, Lensen JF et al. Heparan sulfate proteoglycans in glomerular inflammation. Kidney Int 2004; 65: 768-785. 77. Meri S, Pangburn MK. Regulation of alternative pathway complement activation by glycosaminoglycans: specificity of the polyanion binding site on factor H. Biochem Biophys Res Commun 1994; 198: 52-59. 78. Kelly UL, Ding J, Hageman GS et al. Heparan Sulfate in Human Bruch’s Membrane/ Choroid Tissue Increases the Rate of Proteolytic Cleavage of C3b by factors H and I. Arvo anual meeting 2009; 79. Scholl HP, Fleckenstein M, Issa PC et al. An update on the genetics of age-related macular degeneration. Mol Vis 2007; 13: 196-205. 80. Spraul CW, Lang GE, Grossniklaus HE, Lang GK. Histologic and morphometric analy- sis of the choroid, Bruch’s membrane, and retinal pigment epithelium in postmortem eyes with age-related macular degeneration and histologic examination of surgically excised choroidal neovascular membranes. Surv Ophthalmol 1999; 44 Suppl 1: S10- S32. 81. Bergen AA, Plomp AS, Schuurman EJ et al. Mutations in ABCC6 cause pseudoxan- thoma elasticum. Nat Genet 2000; 25: 228-231. 82. Le Saux O., Urban Z, Tschuch C et al. Mutations in a gene encoding an ABC transporter cause pseudoxanthoma elasticum. Nat Genet 2000; 25: 223-227. 83. Hu X, Plomp AS, van Soest S et al. Pseudoxanthoma elasticum: a clinical, histopatho- logical, and molecular update. Surv Ophthalmol 2003; 48: 424-438. 84. Giachelli CM. The emerging role of phosphate in vascular calcification. Kidney Int 2009; 85. Gorgels TG, Hu X, Scheffer GL et al. Disruption of Abcc6 in the mouse: novel insight in the pathogenesis of pseudoxanthoma elasticum. Hum Mol Genet 2005; 14: 1763- 1773. 86. Larusso J, Li Q, Jiang Q, Uitto J. Elevated Dietary Magnesium Prevents Connective Tissue Mineralization in a Mouse Model of Pseudoxanthoma Elasticum (Abcc6(-/-)). J Invest Dermatol 2009; 87. Schoensiegel F, Weichenhan D, Gorgels TGMF et al. Dystrophic cardiac calcification and pseudoxanthoma elasticum: the yin and yang of Abcc6 loss-of-function mutations. 2009; 88. Brewer GJ. Iron and copper toxicity in diseases of aging, particularly atherosclerosis and Alzheimer’s disease. Exp Biol Med (Maywood ) 2007; 232: 323-335. 89. He X, Hahn P, Iacovelli J et al. Iron homeostasis and toxicity in retinal degeneration. Prog Retin Eye Res 2007; 26: 649-673. 90. Kurz T, Terman A, Brunk UT. Autophagy, ageing and apoptosis: the role of oxidative stress and lysosomal iron. Arch Biochem Biophys 2007; 462: 220-230.

110 The dynamic nature of Bruch’s membrane

91. Kurz T, Karlsson M, Brunk UT, Nilsson SE, Frennesson C. ARPE-19 retinal pigment epithelial cells are highly resistant to oxidative stress and exercise strict control over their lysosomal redox-active iron. Autophagy 2009; 5: 494-501. 92. Nan R, Gor J, Lengyel I, Perkins SJ. Uncontrolled zinc- and copper-induced oligom- erisation of the human complement regulator factor H and its possible implications for function and disease. J Mol Biol 2008; 384: 1341-1352. 93. Newsome DA, Swartz M, Leone NC, Elston RC, Miller E. Oral zinc in macular degen- eration. Arch Ophthalmol 1988; 106: 192-198. 94. AREDS annual report. A randomized, placebo-controlled, clinical trial of high-dose supplementation with vitamins C and E, beta carotene, and zinc for age-related macular degeneration and vision loss: AREDS report no. 8. Arch Ophthalmol 2001; 119: 1417- 1436. 95. Galin MA, Nano HD, Hall T. Ocular zinc concentration. Invest Ophthalmol 1962; 1: 142-148. 96. Lengyel I, Flinn JM, Peto T et al. High concentration of zinc in sub-retinal pigment epithelial deposits. Exp Eye Res 2007; 97. Baynes JW. The role of AGEs in aging: causation or correlation. Exp Gerontol 2001; 36: 1527-1537. 98. Smit AJ, Lutgers HL. The clinical relevance of advanced glycation endproducts (AGE) and recent developments in pharmaceutics to reduce AGE accumulation. Curr Med Chem 2004; 11: 2767-2784. 99. Gugliucci A, Mehlhaff K, Kinugasa E et al. Paraoxonase-1 concentrations in end-stage renal disease patients increase after hemodialysis: correlation with low molecular AGE adduct clearance. Clin Chim Acta 2007; 377: 213-220. 100. Glenn JV, Mahaffy H, Wu K et al. Advanced glycation end product (AGE) accumula- tion on Bruch’s membrane: links to age-related RPE dysfunction. Invest Ophthalmol Vis Sci 2009; 50: 441-451. 101. Sourris KC, Forbes JM. Interactions between advanced glycation end-products (AGE) and their receptors in the development and progression of diabetic nephropathy - are these receptors valid therapeutic targets. Curr Drug Targets 2009; 10: 42-50. 102. Pietkiewicz J, Seweryn E, Bartys A, Gamian A. [Receptors for advanced glycation end products and their physiological and clinical significance]. Postepy Hig Med Dosw (Online ) 2008; 62: 511-523. 103. Handa JT, Verzijl N, Matsunaga H et al. Increase in the advanced glycation end product pentosidine in Bruch’s membrane with age. Invest Ophthalmol Vis Sci 1999; 40: 775- 779. 104. Naka Y, Bucciarelli LG, Wendt T et al. RAGE axis: Animal models and novel insights into the vascular complications of diabetes. Arterioscler Thromb Vasc Biol 2004; 24: 1342-1349. 105. Yamada Y, Ishibashi K, Ishibashi K et al. The expression of advanced glycation end- product receptors in rpe cells associated with basal deposits in human maculas. Exp Eye Res 2006; 82: 840-848. 106. Iacobini C, Menini S, Ricci C et al. Advanced lipoxidation end-products mediate lipid- induced glomerular injury: role of receptor-mediated mechanisms. J Pathol 2009;

111 Chapter 4

107. Sheraidah G, Steinmetz R, Maguire J et al. Correlation between lipids extracted from Bruch’s membrane and age. Ophthalmology 1993; 100: 47-51. 108. Wang L, Li CM, Rudolf M et al. Lipoprotein particles of intraocular origin in human Bruch membrane: an unusual lipid profile. Invest Ophthalmol Vis Sci 2009; 50: 870- 877. 109. Holz FG, Sheraidah G, Pauleikhoff D, Bird AC. Analysis of lipid deposits extracted from human macular and peripheral Bruch’s membrane. Arch Ophthalmol 1994; 112: 402-406. 110. Curcio CA, Millican CL, Bailey T, Kruth HS. Accumulation of cholesterol with age in human Bruch’s membrane. Invest Ophthalmol Vis Sci 2001; 42: 265-274. 111. Huang JD, Curcio CA, Johnson M. Morphometric analysis of lipoprotein-like particle accumulation in aging human macular Bruch’s membrane. Invest Ophthalmol Vis Sci 2008; 49: 2721-2727. 112. Newsome DA, Huh W, Green WR. Bruch’s membrane age-related changes vary by region. Curr Eye Res 1987; 6: 1211-1221. 113. Kang SJ, Grossniklaus HE. Histopathology of age-related macular degeneration. 2009; 3-10. 114. van der Schaft TL, de Bruijn WC, Mooy CM, de Jong PT. Basal laminar deposit in the aging peripheral human retina. Graefes Arch Clin Exp Ophthalmol 1993; 231: 470-475. 115. Lommatzsch A, Hermans P, Muller KD et al. Are low inflammatory reactions involved in exudative age-related macular degeneration? Morphological and immunhistochemi- cal analysis of AMD associated with basal deposits. Graefes Arch Clin Exp Ophthalmol 2008; 246: 803-810. 116. Bressler NM, Silva JC, Bressler SB, Fine SL, Green WR. Clinicopathologic correlation of drusen and retinal pigment epithelial abnormalities in age-related macular degenera- tion. Retina 1994; 14: 130-142. 117. Loeffler KU, Lee WR. Terminology of sub-RPE deposits: do we all speak the same language? Br J Ophthalmol 1998; 82: 1104-1105. 118. de Jong PT. Age-related macular degeneration. N Engl J Med 2006; 355: 1474-1485. 119. Ding X, Patel M, Chan CC. Molecular pathology of age-related macular degeneration. Prog Retin Eye Res 2009; 28: 1-18. 120. Lengyel I, Tufail A, Hosaini HA et al. Association of drusen deposition with choroidal intercapillary pillars in the aging human eye. Invest Ophthalmol Vis Sci 2004; 45: 2886- 2892. 121. Hageman GS, Luthert PJ, Victor Chong NH et al. An integrated hypothesis that consid- ers drusen as biomarkers of immune-mediated processes at the RPE-Bruch’s membrane interface in aging and age-related macular degeneration. Prog Retin Eye Res 2001; 20: 705-732. 122. Johnson LV, Leitner WP, Rivest AJ et al. The Alzheimer’s A beta -peptide is deposited at sites of complement activation in pathologic deposits associated with aging and age- related macular degeneration. Proc Natl Acad Sci U S A 2002; 99: 11830-11835.

112 The dynamic nature of Bruch’s membrane

123. Mullins RF, Russell SR, Anderson DH, Hageman GS. Drusen associated with aging and age-related macular degeneration contain proteins common to extracellular deposits as- sociated with atherosclerosis, elastosis, amyloidosis, and dense deposit disease. FASEB J 2000; 14: 835-846. 124. Glenner GG. The pathobiology of Alzheimer’s disease. Annu Rev Med 1989; 40: 45-51. 125. Maccioni RB, Munoz JP, Barbeito L. The molecular bases of Alzheimer’s disease and other neurodegenerative disorders. Arch Med Res 2001; 32: 367-381. 126. www.ingenuity.com 127. Chen C, Wu L, Wu D et al. The local cone and rod system function in early age-related macular degeneration. Doc Ophthalmol 2004; 109: 1-8. 128. The effect of five-year zinc supplementation on serum zinc, serum cholesterol and he- matocrit in persons randomly assigned to treatment group in the age-related eye disease study: AREDS Report No. 7. J Nutr 2002; 132: 697-702. 129. Evans J. Antioxidant supplements to prevent or slow down the progression of AMD: a systematic review and meta-analysis. Eye 2008; 22: 751-760. 130. Dong A, Xie B, Shen J et al. Oxidative stress promotes ocular neovascularization. J Cell Physiol 2009; 219: 544-552. 131. Gorgels TG, van der Pluijm I, Brandt RM et al. Retinal degeneration and ionizing radia- tion hypersensitivity in a mouse model for Cockayne syndrome. Mol Cell Biol 2007; 27: 1433-1441. 132. Cai J, Nelson KC, Wu M, Sternberg P, Jr., Jones DP. Oxidative damage and protection of the RPE. Prog Retin Eye Res 2000; 19: 205-221. 133. Jarrett SG, Lin H, Godley BF, Boulton ME. Mitochondrial DNA damage and its poten- tial role in retinal degeneration. Prog Retin Eye Res 2008; 27: 596-607. 134. Joffre C, Leclere L, Buteau B et al. Oxysterols induced inflammation and oxidation in primary porcine retinal pigment epithelial cells. Curr Eye Res 2007; 32: 271-280. 135. Weng J, Mata NL, Azarian SM et al. Insights into the function of Rim protein in photo- receptors and etiology of Stargardt’s disease from the phenotype in abcr knockout mice. Cell 1999; 98: 13-23. 136. Sparrow JR, Boulton M. RPE lipofuscin and its role in retinal pathobiology. Exp Eye Res 2005; 80: 595-606. 137. Haines JL, Hauser MA, Schmidt S et al. Complement factor H variant increases the risk of age-related macular degeneration. Science 2005; 308: 419-421. 138. Edwards AO, Ritter R, III, Abel KJ et al. Complement factor H polymorphism and age- related macular degeneration. Science 2005; 308: 421-424. 139. Klein RJ, Zeiss C, Chew EY et al. Complement factor H polymorphism in age-related macular degeneration. Science 2005; 308: 385-389. 140. Hageman GS, Anderson DH, Johnson LV et al. A common haplotype in the complement regulatory gene factor H (HF1/CFH) predisposes individuals to age-related macular degeneration. Proc Natl Acad Sci U S A 2005; 102: 7227-7232. 141. Yates JR, Sepp T, Matharu BK et al. Complement C3 variant and the risk of age-related macular degeneration. N Engl J Med 2007; 357: 553-561. 142. Despriet DD, van Duijn CM, Oostra BA et al. Complement component C3 and risk of age-related macular degeneration. Ophthalmology 2009; 116: 474-480.

113 Chapter 4

143. Penfold PL, Killingsworth MC, Sarks SH. Senile macular degeneration: the involve- ment of immunocompetent cells. Graefes Arch Clin Exp Ophthalmol 1985; 223: 69-76. 144. Richards A, Kavanagh D, Atkinson JP. Inherited complement regulatory protein defi- ciency predisposes to human disease in acute injury and chronic inflammatory statesthe examples of vascular damage in atypical hemolytic uremic syndrome and debris accu- mulation in age-related macular degeneration. Adv Immunol 2007; 96: 141-177. 145. Zhou J, Jang YP, Kim SR, Sparrow JR. Complement activation by photooxidation prod- ucts of A2E, a lipofuscin constituent of the retinal pigment epithelium. Proc Natl Acad Sci U S A 2006; 103: 16182-16187. 146. Wang J, Ohno-Matsui K, Yoshida T et al. Altered function of factor I caused by amyloid beta: implication for pathogenesis of age-related macular degeneration from Drusen. J Immunol 2008; 181: 712-720. 147. Hollyfield JG, Bonilha VL, Rayborn ME et al. Oxidative damage-induced inflammation initiates age-related macular degeneration. Nat Med 2008; 14: 194-198. 148. Scholl HP, Charbel IP, Walier M et al. Systemic complement activation in age-related macular degeneration. PLoS ONE 2008; 3: e2593. 149. Landers RA, Rayborn ME, Myers KM, Hollyfield JG. Increased retinal synthesis of heparan sulfate proteoglycan and HNK-1 glycoproteins following photoreceptor degen- eration. J Neurochem 1994; 63: 737-750. 150. Kempen J, Minami M, Lin J et al. Investigation of the Interactions Between Comple- ment Factor H, C3b, and Amyloid ß. 2008; 151. Schlingemann RO. Role of growth factors and the wound healing response in age- related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2004; 242: 91-101. 152. Penfold PL, Provis JM, Billson FA. Age-related macular degeneration: ultrastructural studies of the relationship of leucocytes to angiogenesis. Graefes Arch Clin Exp Oph- thalmol 1987; 225: 70-76. 153. Penfold PL, Provis JM, Madigan MC, van Driel D, Billson FA. Angiogenesis in normal human retinal development: the involvement of astrocytes and macrophages. Graefes Arch Clin Exp Ophthalmol 1990; 228: 255-263. 154. Bhutto IA, McLeod DS, Hasegawa T et al. Pigment epithelium-derived factor (PEDF) and vascular endothelial growth factor (VEGF) in aged human choroid and eyes with age-related macular degeneration. Exp Eye Res 2006; 82: 99-110. 155. Seddon JM, Cote J, Rosner B. Progression of age-related macular degeneration: as- sociation with dietary fat, transunsaturated fat, nuts, and fish intake. Arch Ophthalmol 2003; 121: 1728-1737. 156. Klein R, Knudtson MD, Klein BE et al. Inflammation, complement factor h, and age- related macular degeneration: the Multi-ethnic Study of Atherosclerosis. Ophthalmol- ogy 2008; 115: 1742-1749. 157. Chuo JY, Wiens M, Etminan M, Maberley DA. Use of lipid-lowering agents for the pre- vention of age-related macular degeneration: a meta-analysis of observational studies. Ophthalmic Epidemiol 2007; 14: 367-374. 158. Rohrer B, Long Q, Wilson RB et al. A Targeted Inhibitor of the Alternative Complement Pathway Reduces Angiogenesis in a Mouse Model of Age-related Macular Degenera- tion. Invest Ophthalmol Vis Sci 2009;

114 The dynamic nature of Bruch’s membrane

159. Emerich DF, Thanos CG. NT-501: an ophthalmic implant of polymer-encapsulated cili- ary neurotrophic factor-producing cells. Curr Opin Mol Ther 2008; 10: 506-515. 160. Maeda K, Kambara M, Tian Y, Hofmann AF, Sugiyama Y. Uptake of ursodeoxycholate and its conjugates by human hepatocytes: role of Na(+)-taurocholate cotransporting polypeptide (NTCP), organic anion transporting polypeptide (OATP) 1B1 (OATP-C), and oatp1B3 (OATP8). Mol Pharm 2006; 3: 70-77. 161. Marmor MF, Jain A, Moshfeghi D. Total rod ERG suppression with high dose compas- sionate Fenretinide usage. Doc Ophthalmol 2008; 117: 257-261. 162. Witmer AN, Vrensen GF, Van Noorden CJ, Schlingemann RO. Vascular endothelial growth factors and angiogenesis in eye disease. Prog Retin Eye Res 2003; 22: 1-29. 163. Del Priore LV, Tezel TH, Kaplan HJ. Maculoplasty for age-related macular degener- ation: reengineering Bruch’s membrane and the human macula. Prog Retin Eye Res 2006; 25: 539-562. 164. Hayward C, Shu X, Cideciyan AV et al. Mutation in a short-chain collagen gene, CTRP5, results in extracellular deposit formation in late-onset retinal degeneration: a genetic model for age-related macular degeneration. Hum Mol Genet 2003; 12: 2657- 2667. 165. Marmorstein LY, McLaughlin PJ, Peachey NS, Sasaki T, Marmorstein AD. Formation and progression of sub-retinal pigment epithelium deposits in Efemp1 mutation knock- in mice: a model for the early pathogenic course of macular degeneration. Hum Mol Genet 2007; 16: 2423-2432. 166. Mathieu JM, Schloendorn J, Rittmann BE, Alvarez PJ. Medical bioremediation of age- related diseases. Microb Cell Fact 2009; 8: 21. 167. http://www.ncbi.nlm.nih.gov/sites/entrez?db=PubMed

Acknowledgements The authors thank the ‘Algemene Nederlandse Vereniging ter Voorkoming van Blindheid’ (ANVVB), the Macula Foundation of the Netherlands, the ‘Landelijke Stichting voor Blin- den en Slechtzienden’, the ‘Rotterdamse Blinden Belangen’, the ‘Blindenpenning’, the ‘Gelderse Blinden Vereniging’, and ‘Stichting Uitzicht’ for their financial contributions.

115

On the origin of Bruch’s mem- brane and drusen proteins

5

Judith C Booij, Jacoline B ten Brink, Peter J van der Spek, Arthur AB Bergen

manuscript in preparation Chapter 5

Abstract Purpose: To identify the origin of proteins that make up the acellular Bruch’s membrane (BM) and drusen, both implicated in the initial stages of age-related macular degeneration (AMD).

Methods: We performed microarray analyses of laser-dissected retinal pigment epithelium (RPE), choroidal and photoreceptor cells. We compared the gene expression levels of all three tissues and we studied the expression of the genes that are translated into BM and drusen proteins in all three above mentioned cell layers. A human serum proteome list was generated using the Ingenuity knowledge database.

Results: Initially, we briefly described the identification and functional annotation of genes expressed in the human RPE, choroid and photoreceptor cells. Next, we generated a fourth data set containing the human serum proteome. We subsequently showed that most collagen and laminin subtypes previously identified in BM are expressed at higher levels in choroidal cells than in RPE cells. A similar analysis showed that most genes corresponding to drusen proteins were either expressed at higher levels in choroidal cells than in RPE cells or that these drusen proteins were derived from serum.

Conclusions: In this study we provided insight into the origin of BM and drusen proteins and we showed that the majority is derived from choroidal cells and/or serum. This knowledge may be useful in the search for a (systemic) treatment for retinal disorders, like age-related macular degeneration (AMD).

118 The origin of human Bruch’s membrane proteins

Introduction Bruch’s membrane (BM) is a collagen and elastin-rich acellular layer in the retina. It is lo- cated between the photoreceptor/ retinal pigment epithelium (RPE) cell layers on one side, and the choroidal blood vessels on the other side. The 3 μm thick BM plays an essential role in maintaining the normal function of the outer retina. More specifically, BM regulates the diffusion of (bio-) molecules between photoreceptors/RPE and choroid, it provides physical support for RPE cell adhesion, migration and differentiation, and it acts as a division barrier restricting choroidal and retinal cellular migration[1]. BM is implicated in many retinal dis- orders such as age-related macular degeneration and retinitis pimentosa[2].

BM forms a single functional unit with the adjacent RPE and choroid. Previous studies in chickens and mice showed that, during embryogenesis, both the RPE and choroid contribute to the development of BM[3,4]. Already in 1989, Hewitt and coworkers suggested that hu- man BM is also composed of molecules synthesized by both the RPE and the choroid[5]. Nonetheless, in humans, the extent of -genetic- involvement of the RPE, the choroid or even the photoreceptors in the development and maintenance of BM remains to be elucidated. Our current study addresses this specific question: where do the proteins in BM, an acellular layer incapable of producing proteins itself, come from?

With age, extracellular deposits, called drusen, develop in BM[1]. Several types of drusen ex- ist (reviewed by Ding et al[6]), for example hard and soft drusen. They occur both in healthy eyes and in eyes affected with early AMD. Drusen are not only present in the macular area, but also in the retinal periphery. The presence of soft confluent drusen in the macula is a ma- jor risk factor for the development of age-related macular degeneration (AMD)[7]. In a recent review we described that drusen consist of lipids, cholesterol, fat, complement components and inhibitors, apoplipoproteins, acute phase proteins, C-reactive protein, and many other proteins[1]. Most of these drusen components were identified only recently, and hardly anything is known with regard to where they come from. The most likely sources of drusen components, given their location, are either photoreceptor cells, RPE cells, choroidal cells, serum proteins, or debris from BM turnover.

In another recent publication, we analysed RPE-specific gene expression in the macular area of healthy human donor eyes[8]. We made a direct comparison of expression levels of 44K genes in the macular RPE, the choroid and in the photoreceptor cells with the aim to identify and analyse RPE-specific gene expression[8]. Using a complementary approach, we currently briefly reviewed the gene expression levels of the RPE, and we newly described the transcriptomes of the photoreceptors and the cho- roid. In addition, we generated a list of proteins representing the human serum proteome. Based on these datasets, we determined the most likely origin of proteins known to be present in either BM or drusen.

119 Chapter 5

Results Gene expression of RPE, photoreceptors and choroid RPE gene expression We previously analyzed the human RPE transcriptome extensively[8-10]. In summary, we functionally annotated the RPE transcriptome on the basis of retinal location[9] and on the basis of gene expression levels and/or variability[10]. These studies showed that the RPE transcriptome varies 1-5% depending on retinal location, and that the RPE has strong energy demands, is exposed to high levels of oxidative stress and a variable degree of inflamma- tion[9,10]. Finally, we compared the transcriptomes of the photoreceptor, choroidal and RPE cells to further define a set of RPE expressed genes. In the latter study, we identified 458 genes (1% of all entries examined) that we consider to be specific for the RPE compared to the adjacent cell layers, based on the fact that their expression levels are at least 2.5 fold higher in the RPE than in both choroid and photoreceptors. Functional annotation of the genes specifically expressed in the RPE yielded an overrepresentation of genes involved in inositol metabolism, retinol metabolism, genetic disorders and ophthalmic diseases. When we combine the most highly expressed RPE genes (top 10 percentile) [10] with the 458 RPE-specific expressed genes[8], we end up with a set of genes, which may be considered to represent the most important functions of the RPE.

Photoreceptor gene expression Pure photoreceptor cell isolation from the retina by LDM is relatively easy, given the large size and neat local arrangement of the cells. However, in our dataset we cannot distinguish between rod and cone photoreceptor cells. Some cellular/RNA contamination from Müller- glia cells may also be inevitable, since processes of these cells extend into the outer nuclear layer. Cellular contamination from adjacent, easily distinguishable horizontal or bipolar cells is likely to be minimal. However, our previous studies did show that it is extremely difficult to separate photoreceptors from RPE cells and vice versa[8]. In the macular area, we identified 336 (1%) genes with expression levels 2.5 times higher in the photoreceptor cell layer than in the RPE (phot>RPE) (see supplementary file 1, available on request). This part of the transcriptome is most likely specific for the photoreceptors, if we compare it to RPE cells. For this fraction of the photoreceptor transcriptome, we found an overrepresentation of genes involved in the following functional categories: vision, retinitis pigmentosa, sensory transduction, cone-rod dystrophy, alternative splicing and cGMP bind- ing. The top thirty genes with the highest photoreceptor FC compared to RPE are presented in Table 1.

Table 1 Genes with highest expression in photoreceptors compared to RPE (NCBI entrezgene, NCBI OMIM)

gene symbol Genbank ID FC phot/RPE GUCY2F NM_001522 7 THC1886017 THC1886017 6

120 The origin of human Bruch’s membrane proteins

gene symbol Genbank ID FC phot/RPE A_24_P911171 A_24_P911171 6 AKNA AB075848 5 NM_173498 NM_173498 5 USH2A NM_007123 5 A_32_P88605 A_32_P88605 5 THC1838874 THC1838874 5 PRO1966 AF116677 5 THC1934099 THC1934099 5 THC1848468 THC1848468 4 SLAIN2 AK025264 4 A_24_P885910 A_24_P885910 4 A_23_P119407 A_23_P119407 4 AL359574 AL359574 4 THC1861725 THC1861725 4 LOC441351 BC040180 4 LOC157627 AL832535 4 LOC723809 BC050337 4 LOC389660 XM_372042 4 MPP4 NM_033066 4 THC1814602 THC1814602 4 THC1993549 THC1993549 4 CATSPER2 NM_172095 4 LOC157627 AK091593 4 THC1878776 THC1878776 4 LOC94431 CD110160 4 THC1821072 THC1821072 4 PCDH15 XM_373461 4 RP1 NM_006269 4

Choroidal gene expression Choroidal gene expression is in fact determined by a number of cell types, including smooth muscle cells, endothelial cells, fibroblasts, resident or incoming macrophages, etc. Probably due to this heterogeneity, we found that there were 1,527 (5%) genes with expression levels 2.5 times higher in the choroid than the RPE (chor>RPE) (see supplementary file 2, avail- able on request). This part of the transcriptome is specific for the choroid as compared to the adjacent RPE cells. The top 30 with the highest expression in choroid vs RPE is shown in

121 Chapter 5

Table 2. Among the 1,527 genes, there was an overrepresentation of genes in the following Kegg pathways: cell adhesion molecules (CAMs), complement and coagulation cascades, type I diabetes mellitus, focal adhesion, antigen processing and presentation and extracellular matrix (ECM)-receptor interaction.

Table 2 Genes with highest expression in choroid compared to RPE (NCBI entrezgene, NCBI OMIM)

gene symbol Genbank ID FC chor/RPE SFRP4 NM_003014 26 THC1977129 THC1977129 26 IGL alpha L38562 22 OGN NM_033014 19 FAM150B AY358517 19 KLF6 BC015987 18 MYOC NM_000261 17 VIT NM_053276 16 LEPR U50748 15 IGH alpha AK130614 15 A_23_P435390 A_23_P435390 14 NRP2 AK024680 14 DCN NM_001920 14 ROPN1 NM_017578 14 PITX2 NM_153426 14 EDN3 NM_000114 14 IFI6 BC024289 13 ENST00000328419 ENST00000328419 13 ROPN1B BC015413 13 MYOT NM_006790 13 RSPO2 NM_178565 13 CCL26 NM_006072 13 IGKV3D-15 AF035031 12 FMOD NM_002023 12 IGK alpha BC030813 12 ENST00000331696 ENST00000331696 12 MLANA NM_005511 11 ANXA1 NM_000700 11 PAX3 NM_181458 11 TMEFF2 NM_016192 11

122 The origin of human Bruch’s membrane proteins

Where do Bruch’s membrane proteins come from? In order to elucidate the potential origin of proteins previously identified in Bruch’s mem- brane (Figure 1), we determined their corresponding gene expression in the adjacent cell layers (RPE, choroid and photoreceptors). We focused specifically on the expression levels of the following gene groups: collagens, laminins, fibronectins, elastin, TIMP3 and heparan sulphate proteoglycans (HSPG). In ad- dition, we generated a serum proteome list to define if BM proteins could (also) be derived from serum (Table 3). A substantial number of collagen proteins have been described in the five layers of BM[1,11- 13]. The laminin and HSPG proteins are present in the basement membranes on both the RPE and the choroidal side of BM[14]. The fibronectin proteins are found in the inner and the outer collagen layers of BM[15]. The TIMP3 protein is located in the elastin layer of BM[16]. Elastin is, obviously, located in the elastin layer[1,12](Figure 1).

Figure 1. Known BM proteins and their expression in the photoreceptors, the RPE and the choroid.

Phot

COL4A3 FNDC5 TIMP3

RPE bm RPE [COL 4A1, 4A2, 4A3, 4A4, 4A5, 5] LAM HSPG [COL1, 3] FN inner collagen [COL 4, 6, 18] ELN TIMP3* elastin layer BM [COL1, 3] FN outer collagen [COL 4A1, 4A2, 5] LAM HSPG bm choroid COL 1A1, 1A2, 3A1, 4A1, LAM A2, FN1, 4A2, 5A1, 5A2, 6A1, 6A2, A4, B1, FNDC3A 18A1 C1, C3 Chor

A color version of this figure can be found in chapter 15.

123 Chapter 5

Table 3 Proteins identified in serum according to Ingenuity software analysis[45].

gene symbol Genbank ID gene symbol Genbank ID gene symbol Genbank ID A2M NM_000014 FGFBP2 NM_031950 NEDD4 NM_006154 ACE NM_152831 FGG NM_021870 NPPA NM_006172 ACHE NM_000665 FGL1 NM_004467 NPPB NM_002521 ACTL6B NM_016188 FN1 NM_054034 NPY NM_000905 ADAMTS13 NM_139025 GC NM_000583 ORM1 NM_000607 ADIPOQ NM_004797 GCG NM_002054 OTOR NM_020157 ADM NM_001124 GDF15 NM_004864 OXT NM_000915 AGT NM_000029 GH1 NM_000515 PDIA2 NM_006849 AHSG NM_001622 GHRH NM_021081 PF4 NM_002619 ALB NM_000477 GHRL NM_016362 PGCP NM_016134 ANGPTL4 NM_139314 GP1BA NM_000173 PLA2G7 NM_005084 APLN NM_017413 GPBP1 NM_022913 PLAT NM_000930 APOA1 NM_000039 GPLD1 NM_001503 PLAUR NM_002659 APOA2 NM_001643 GPT NM_005309 PLEC1 NM_201384 APOA4 NM_000482 GPX3 NM_002084 PLG NM_000301 APOA5 NP_001160070.1 GSN NM_000177 PLTP NM_006227 APOB NM_000384 GSR NM_000637 POMC NM_000939 APOC1 NM_001645 GUSB NM_000181 PPBP NM_002704 APOC2 NM_000483 GZMA NM_006144 PPY NM_002722 APOC3 NM_000040 HBA2 NM_000517 PRDX1 NM_002574 APOD NM_001647 HBB NP_000509.1 PRDX2 NM_181738 APOE NM_000041 HP NM_005143 PRL NM_000948 APOH NM_000042 IBSP NM_004967 PRNP X82545 APOL6 NM_030641 ICAM1 NM_000201 PROC NM_000312 APOM NM_019101 IFNG NM_000619 PROS1 NM_000313 APP NM_000484 IGF1 NM_000618 PTH NM_000315 ATP13A1 NM_020410 IGF2 NM_000612 PYY NM_004160 AVP NM_000490 IGFALS NM_004970 RELN NM_005045 AZGP1 NM_001185 IGFBP1 NM_000596 REN NM_000537 B2M NM_004048 IGFBP2 NM_000597 RET NM_020630 BGLAP NM_199173 IGFBP3 NM_000598 RETN NM_020415 C3 NM_000064 IGFBP4 NP_001543.2 RNASE3 NM_002935 CAMP NM_004345 IGH-2 XP_002347520.1 RXRA NM_002957 CCL13 NM_005408 IGHM AK090464 RXRB NM_021976

124 The origin of human Bruch’s membrane proteins

gene symbol Genbank ID gene symbol Genbank ID gene symbol Genbank ID CCL14 NP_116738.1 IGJ NP_653247.1 S100A4 NM_002961 CCL2 NM_002982 IL10 NM_000572 S100A8 NM_002964 CCL22 NP_002981.2 IL12B NM_002187 S100A9 NM_002965 CCL23 NM_005064 IL18 NM_001562 SAA1 NM_000331 CCL26 NM_006072 IL1B NM_000576 SELE NM_000450 CD163 NM_004244 IL1RN NM_173841 SELL NM_000655 CD40LG NM_000074 IL2 NM_000586 SERPINA1 NM_000295 CES1 NM_001266 IL4 NM_000589 SERPINA3 NM_001085 CETP NM_000078 IL5 NM_000879 SERPINA3K NP_001076.2 CFD NM_001928 IL6 NM_000600 SERPINA3M NP_001076.2 CFHR1 NM_002113 IL6R NM_000565 SERPINA6 NP_001747.2 CHGA NM_001275 IL7 NM_000880 SERPINA7 NM_000354 CLEC3B NM_003278 INHA NM_002191 SERPINC1 NM_000488 CLU NM_001831 INHBA NM_002192 SERPINE1 NM_000602 CNDP1 NM_032649 INHBB NM_002193 SERPINF1 NM_002615 COL1A1 NM_000088 INS NM_000207 SERPINF2 NM_000934 CPB2 NM_001872 INS1 NP_000198.1 SERPING1 NM_000062 CRP NM_000567 ITIH3 NM_002217 SHBG NM_001040 CSF2RA NM_172247 ITIH4 NM_002218 SLC17A5 NM_012434 CSH1 NM_022640 ITPA NM_033453 SPP1 NM_000582 CXCL2 NM_002089 KIT NM_000222 TCN1 NM_001062 CXCL3 NM_002090 KNG1 NM_000893 TF NM_001063 DBP NM_001352 LAG3 NM_002286 TFPI NM_006287 ELA2 NM_001972 LBP NP_004130.2 TGFB1 NM_000660 ENPP1 NM_006208 LCAT NM_000229 THPO NM_000460 F10 NM_000504 LCN2 NM_005564 TIMP1 NM_003254 F12 NM_000505 LEP NM_000230 TIMP2 NM_003255 F13A1 NP_000120.2 LEPR NM_002303 TNF NM_000594 F2 NM_000506 LGMN NM_005606 TNFRSF1A NM_001065 F3 NM_001993 LIPC NM_000236 TNFRSF1B NM_001066 F5 NM_000130 LPA NM_005577 TNFRSF8 NM_001243 F7 NM_000131 LPL NM_000237 TNFSF12 NM_153012 F8 NM_000132 LRG1 NM_052972 TOR2A NM_130459 F9 NM_000133 LTF NM_002343 TTR NP_000362.1 FABP3 NM_004102 LYZ NM_000239 TXN NM_003329

125 Chapter 5

gene symbol Genbank ID gene symbol Genbank ID gene symbol Genbank ID FAP NM_004460 MASP1 NM_139125 TXNRD1 NM_003330 FCN2 NM_015839 MASP2 NM_139208 VIP NM_003381 FCN3 NM_003665 MBL2 NM_000242 VTN NM_000638 FGA NM_021871 MMP1 NM_002421 VWF NM_000552 FGB NM_005141 MMP2 NM_004530 XDH NM_000379 FGF1 NM_000800 MMP3 NM_002422 ZFR NM_016107 FGFBP1 NM_005130 MMP9 NM_004994 RPE contribution to BM Of the known BM proteins, one collagen subtype (COL4A3 [NM_000091]) and the TIMP3 genes had higher expression in the RPE than in the choroid (Figure 1). Moreover, two genes corresponding to the collagen subtypes (COL8A1 [NM_001850] and COL20A1 [NM_020882]), which could be present in BM on theoretical grounds (for instance because other subtypes of the protein have been identified in BM, candidate genes), also had higher expression in the RPE than in the choroid (Figure 2).

Figure 2. Candidate proteins not identified in BM, but in terms of function or structure closely related to BM pro- teins, and their expression in the photoreceptors, the RPE and the choroid. COL25A1

Phot

COL8A1 COL20A1 RPE bm RPE [COL 4A1, 4A2, 4A3, 4A4, 4A5, 5] LAM HSPG [COL1, 3] FN inner collagen [COL 4, 6, 18] ELN TIMP3* elastin layer BM [COL1, 3] FN outer collagen [COL 4A1, 4A2, 5] LAM HSPG bm choroid COL 8A2, 10A1, 13A1, 15A1, 16A1, 21A1

Chor

A color version of this figure can be found in chapter 15.

126 The origin of human Bruch’s membrane proteins

Choroidal contribution to BM Altogether, sixteen collagen-, five laminin and two fibronectin gene subtypes were more highly expressed in the choroid than in the RPE. Ten (COL1A1, COL1A2, COL3A1, COL4A1, COL4A2, COL5A1, COL5A2, COL6A1, COL6A2 and COL18A1) of these sixteen collagen subtypes were previously assigned to BM (Figures 1 and 2). Three major cell types in the choroid are smooth muscle cells, endothelial cells en fibroblasts[1]. Using Ingenuity analysis, we compared the known gene expression of these cell types with proteins previously identified in BM. In smooth muscle cells, at least two genes are expressed that have corresponding BM proteins (FN1 and LAMA2). In endo- thelial cells, these genes are also expressed, along with COL4A2, COL15A1 and COL18A1. Finally, in fibroblasts, at least FN1 and three collagen subtypes (COL1A1, COL1A2 and COL14A1) are expressed.

Photoreceptor contribution to BM The COL25A1 gene, for which the protein product is not known to be present in BM or elsewhere in the retina, was the only collagen gene with higher gene expression in the photo- receptors than the RPE (Figure 2).

Serum contribution to BM Analysis of the serum proteome revealed the COL1A1 and FN1 protein entries in BM.

Drusen proteins: where do they come from? Using proteomics, Crabb and coworkers identified 114 proteins in drusen[17]. Subsequently, other investigators found additional drusen proteins like ubiquitin, integrins, TIMP2, ad- vanced glycation end products, beta-amyloid, fibronectin and C1q[17-20]. In total, we re- cently compiled a list of 129 drusen constituents[1]. Using the Ingenuity knowledge database, we identified 23 drusen proteins (20%) in the serum proteome (Figure 3). We also analyzed the RPE, photoreceptor and choroidal gene expression levels corresponding to the drusen proteins. In the following sections we describe the genes expressed in each of the cell types.

RPE proteins in drusen Higher gene expression levels in the RPE than the choroid (RPE>Chor) was observed for twelve drusen proteins (Figure 3). Seven of these proteins are involved in signalling, three are extracellular matrix proteins and five are secreted proteins.

Choroidal proteins in drusen There were 29 genes corresponding to drusen components with higher gene expression levels in the choroid than the RPE (Chor>RPE) (Figure 3). Among these, there was an overrepre- sentation of genes involved in cell-cell communication through interaction with extracellular matrix receptors.

127 Chapter 5

Photoreceptor proteins in drusen Only two crystallin genes (CRYBA1 [NM_005208] and CRYGS [NM_017541]) correspond- ing to drusen proteins were identified with higher genes expression in the photoreceptors than in the RPE (Figure 3). These are heat shock / chaperone proteins which may indicate that the photoreceptor cells are especially equipped to deal with oxidative stress in order to prevent apoptosis.

Figure 3. Known drusen proteins and their expression in the photoreceptors, the RPE and the choroid. See text.

CRYBA1 CRYGS

Phot

SERPINF RGR CTSD SMC6 RPE COL8A1 FRZB TIMP3

114 proteins

BM

COL1A2, COL6A1, COL6A2, ALDH1A1, ANXA1 HIST1H2AE CRYAB, FN1, HBA1, HBA2, IGHA1, ANXA2 HIST1H2BL IGKC, LMNA, LUM, MFAP4, OGN, ANXA5 HIST1H4H C3 PLA2G2A, PRELP, RNASE4, SAA1, HIST2H2AA3 FBLN5 SERPINA3, VIM

Chor

ALB, APOA1, APOA4, APOE, C3, CLU, FGG, FN1, HBA2, HBB, HP, LTF, LYZ, ORM1, PLG, PRDX1, S100A8, S100A9, SAA1, SERPINA1, SERPINA3, SERPINF1, VTN

A color version of this figure can be found in chapter 15.

Discussion We described the functional annotation of RPE gene expression in a previous study[10]. Advantages and limitations of our microarray approach were also discussed extensively pre- viously[9,10]. In summary, high cellular specificity comes from our use of strict selection cri- teria for donor eyes and the use of a laser dissection microscope for the isolation of all three cell types (RPE, choroid and photoreceptors). The current study design, using microarrays on laser microdissected material from RPE, choroid and photoreceptors avoids unnecessary experimental manipulation and minimizes contamination from adjacent cell types[10]. Limi- tations include the limited number of samples in our study and the relatively low molecular

128 The origin of human Bruch’s membrane proteins weight RNA available, inevitably due to the relatively poor quality of postmortem material from human donors. Furthermore, our current study is limited in the fact that we cannot quan- tify absolute gene expression levels of individual cell layers. Rather, we designed the study so that we can compare quantitatively gene expression levels of one cell layer with those of another cell layer.

Functional annotation of choroidal and photoreceptor gene expression Choroidal gene expression Our data-driven functional annotation analyses implicates that the choroid is involved in local immune processes (antigen processing and presentation, complement and coagulation cas- cades). Indeed, physiological and pathological studies have shown that the choroid and RPE are involved in chronic inflammatory processes in healthy eyes and/or eyes that are affected by age-related macular degeneration and perhaps in immune surveillance of the eye[21]. The functional overrepresentation of (focal) cell adhesion and ECM genes expressed in the cho- roid may correspond to the need of this tissue, which consist of a heterogeneous mixture of cell types and blood vessels, to maintain a single firm matrix structure (between the RPE and the sclera). Our data confirm recent findings that the proteins in the BM-choroid complex are involved in cell adhesion, tissue homeostasis, wound healing and blood vessel growth[22].

Photoreceptor gene expression Our photoreceptor gene expression analyses and functional annotation confirm the well known role of these light sensitive cells in vision. Moreover, we found an overrepresentation of retinal disease genes among the highly expressed photoreceptor genes. This suggests that additional genes with high photoreceptor expression, not currently known to be involved in retinal diseases (Table 1 and Supplementary file 1, available on request), are also good can- didate retinal disease genes.

Bruch’s membrane Several previously published studies focused on the molecular composition of Bruch’s mem- brane (BM)[11,16,23,24]. It is known that BM is an acellular layer and therefore, BM con- stituents must be synthesized elsewhere. We hypothesized that the majority of BM compo- nents (collagens, fibronectins, laminins and elastin) are likely to have corresponding gene expression in the RPE, the choroid, or the photoreceptors.

Collagens Interestingly, we found that genes that correspond to the majority of the BM collagens have the highest gene expression levels in the choroidal cells (Figure 2). We also observed prefer- ential choroidal cell expression for candidate BM proteins (COL8A2, COL10A1, COL13A1, COL15A1, COL16A1, COL21A1) (a number of collagenous proteins related to collagen proteins found in BM) (Figure 2). Taken together, these data suggest that the choroid, more than the RPE or the photoreceptors, plays a crucial role in the maintenance of BM. Our data support the view of Chong and coworkers, that BM structurally strongly resembles the in- tima of blood vessel walls[25]. Consequently, malfunction of the cardiovascular system and

129 Chapter 5 in particular the cappilary bed of the choroid could have major implications for the integrity of BM.

With regard to the candidate BM proteins (proteins for which related (sub)types have been shown to be present in BM) expressed in the choroid (COL8A2, COL10A1, COL13A1, COL15A1, COL16A1, COL21A1) and in the photoreceptors, (COL8A1 and COL20A1) we observed the following: collagen 8 consists of two subunits, COL8A1 (expressed at higher levels in the RPE, cornea and lens[26]) and COL8A2 (expressed primarily in the choroid), which indicated that the RPE and choroid both contribute to the presence of collagen type 8. Higher expression in the choroid was also seen for COL10A1 which structurally closely re- sembles COL8A1[27] and for COL15A1 and COL18A1. The latter two proteins are secreted into extracellular matrices and are antiangiogenic. Moreover, collagen type 15 is known to be present in basement membranes and is thought to adhere them to the underlying connective tissue stroma[28]. Its potential presence in BM awaits confirmation. Higher expression in the photoreceptors was seen for COL25A1, a membrane-bound collagen expressed in brain[29]. Proteolytic processing of COL25A1 results in a soluable form of COL25A1 known to bind senile plaques in Alzheimer’s disease[30]. Since the molecular constituents of Alzheimer’s disease senile plaques and drusen overlap[1], COL25A1 could also be present in drusen. If this is the case, the expression of the COL25A1 gene in the photoreceptor cells implies that these cells may also contribute one or more proteins to drusen.

Fibronectins Fibronectins are extracellular matrix glycoproteins with high molecular weight, present in basement membranes. They interact with collagens and are involved in adhesion and migra- tion[31]. Fibronectin is present in both the ICL and OCL of BM[23]. To our knowledge, no specification of the fibronectin subtypes in BM has been published as of yet. Most impor- tantly, the FN1 gene is expressed at higher levels in the choroid and may be involved in the clearance of immune complexes and cellular debris on the BM-choroid interface[32]. The FNDC5 gene, with a comparable function to FN1, is expressed at higher levels in the RPE[1].

Laminins Laminins, together with collagen type IV, are ECM proteins that are major structural compo- nents of basal laminae[27]. Indeed, laminin proteins are found in the apical and basal base- ment membranes of BM[14]. Interestingly, four subtypes of laminins had higher gene ex- pression in the choroid than in the RPE. These are LAMA2, LAMA4, LAMB1 and LAMC1. These data suggest that the choroid is an important supplier of laminins in BM. Changes in the supply of laminins to BM due to choroidal aging and deterioration may have major con- sequences for the structural integrity of BM.

TIMP3 and metalloproteinases Using monoclonal antibodies, Farris and coworkers showed that the TIMP3 protein was located in the endothelial cells of the choriocapillaris and possibly in the elastin layer of BM[16,33,34]. TIMP3 is an inhibitor of matrix metalloproteinases, a group of zinc-binding

130 The origin of human Bruch’s membrane proteins endopeptidases that are involved in the turnover of extracellular matrices (ECM). Conse- quently TIMP3 stabilizes the ECM. In addition, TIMP3 is a potent angiogenesis inhibitor in BM. The TIMP3 gene, expressed in the RPE in mice[35], is mutated in Sorsby’s fundus dystrophy[36], a disease with a strong clinical resemblance to AMD and characterized by choroidal neovascularisation. We confirmed that this gene is expressed at higher levels in the RPE than the choroid.

Drusen Human drusen consist mainly of lipids and proteins partially modified by oxidative stress[1,21] and a number of trace elements. A number of investigators showed that drusen lipids are primarily derived from photoreceptor cells (reviewed in [1]). We recently hypoth- esized that drusen proteins primarily originate from choroidal cells and serum[1]. In the current study study we provided qualitative data that suggest that 23% of all drusen proteins may primarily come from choroidal cells, and 18% of all drusen proteins are derived from the serum. In contrast, only a very small proportion of drusen proteins may be primarily derived from RPE (10%) or photoreceptors (8%) (Figure 3).

Most interestingly, among the drusen proteins that are primarily synthesized in the choroidal cells are: three different annexins, complement factor C3 and fibulin 5 (FBLN5) (Figure 3). Annexins play an important anti-inflammatory role mediated by the inhibition of phospholi- pase A2, which prevents the biosynthesis of two strong inflammatory factors (prostaglandins and leukotrienes)[27]. C3 is one of the key complement factors regulating the local innate immune response[37] and is expressed at higher levels in the choroidal cells than in the RPE. Nonetheless, in the context of AMD pathogenesis, it is of interest to note that not only choroidal and RPE cells synthesize this compound, but that C3 also circulates in the blood. FBLN5 promotes adhesion of endothelial cells to each other via the interaction of integrins and the arginyl-glycyl-aspartic acid (RGD) motif[38]. Integrins also play a very important role in the binding of RPE cells to BM[39]. FBLN5 could be a vascular ligand for integrin receptors and may play a role in vascular development and remodeling and therefore may play a role in prevention of neovascularisation in AMD[38]. These data confirm the intimate relationship between the homeostasis of the RPE, the bal- ance and stability of BM constituents, the development of drusen, the activity of the immune system and the presence of neovascularisation as seen in retinal disorders like AMD[1,7]. Finally, four different members of the histone protein family, present in drusen, apparantly originate from the choroid. Histones are core proteins of the DNA, required for chromosomal organization that are released from cells upon apoptosis. Histones are hydrophobic and are hard to degrade. Whether these findings indicate that waste material in drusen is predomi- nantly derived from the choroid, or the choroid merely has higher transcriptional activity remains to be elucidated. Of the crystallin proteins present in drusen, two subtypes had the higher gene expression in the photoreceptors than in the RPE (CRYBA1, CRYGS). The beta crystallins were previously identified in photoreceptors[40]. One crystallin, CRYAB, a highly immunogenic protein[31], was more highly expressed in the choroid.

131 Chapter 5

Drusen also contain the COL8A1 protein, which in our data set, has highest gene expres- sion levels in the RPE. In addition to COL8A1, drusen also contain COL1A2, COL6A1 and COL6A2, with highest corresponding gene expression in the choroid. Collagen type 6, composed of three subunits, is a component of microfibrillar structures and plays a role in cell migration and differentiation. However, the presence of these collagens in drusen may not be solely explained by the gene expression in adjacent cells. Aside from a direct cellular origin, there is also the possibility that drusen constituents are derived from BM turnover products that were trapped in the heterogeneous molecule mixture of drusen. BM continu- ously renews itself and with age the composition of BM changes slightly. Indeed, BM is a dynamic structure[1,41,42].

Waste products from the choroid can possibly be directed towards the serum. However, waste products of (apoptotic) RPE cells may be trapped in drusen since they cannot pass BM.

Conclusions Our data suggest that BM proteins are synthesized by the RPE and choroid. With regard to collagen proteins present in BM, we showed that COL4A3 was primarily expressed in the RPE, and that ten other collagen chains (COL1A1, COL1A2, COL3A1, COL4A1, COL4A2, COL5A1, COL5A2, COL6A1, COL6A2 and COL18A1) were primarily expressed in the choroid. Expression of subtypes of the laminin genes (LAMA2, LAMA4, LAMB1, LAMC1 and LAMC3) was also primarily observed in the choroid. With regard to drusen, our data suggest that over 40% of drusen proteins primarily come from choroid and/or serum. Since drusen are an important hallmark of age-related macular degeneration, this insight perhaps provides opportunities for systemic treatment of AMD.

Methods Human donor eyes This study was performed in agreement with the declaration of Helsinki on the use of hu- man material for research. Material used in this study was provided to us by the Corneabank Amsterdam. A detailed description of our methods can be found elsewhere[8]. In brief, we selected five eyes without ocular pathology from five human postmortem donors. Globes were enucleated and subjected to visual and histological examination, including periodic acid Schiff (PAS) staining. RPE, photoreceptor and choroidal cells were isolated from macular fragments using a Laser Microdissection System (PALM, Bernried, Germany). Total RNA was isolated and the mRNA component was amplified[9], labeled and hybridized to a 44k microarray (Agilent Technologies, Amstelveen, The Netherlands). Three arrays were used to compare the photoreceptors to the RPE, three additional arrays for the comparison of the cho- roid to the RPE. In each comparison, two tissue samples of a single donor were compared.

132 The origin of human Bruch’s membrane proteins

Data analysis Scanned images were processed and analyzed with Feature Extraction software (v 8.5 Agi- lent) and Rosetta Resolver software (Rosetta Inpharmatics). For each gene we calculated the following ratio’s, RPE vs. choroid, choroid vs. RPE and photoceptor vs. RPE, depending on the array. This resulted in three duplicate measurements for each ratio. Only when the ratio in all three duplicate measurements for a single gene was greater than 2.5, we considered a gene to have a meaningfully higher gene expression (GE) in one tissue compared to the other. These genes with higher expression in a cell layer can be considered specific for this cell layer and important for its characteristic features and functions. In addition, by using this cri- terium we eliminate expression data caused by contamination of our laser disected samples by tissue fragments from adjacent layers. Expression levels due to contamination will always be lower than expression levels in the true tissue of origin. A functional analysis of Kegg pathways (Kyoto Encyclopedia of Genes and Genomes) and functional categories was per- formed on all groups of genes with an average FC>2.5 using the David online software[43]. Cut off criteria used were a p-value of less than 0.001 using either a Benjamini-Hochberg correction or an Ease score, a modified Fisher’s exact test[43,44].

Generation of a serum proteome list A serum proteome list was generated using the Ingenuity knowledge database. The list con- tains 227 entries and is presented in Table 3. This list was used to identify BM proteins originating from serum.

133 Chapter 5

References 1. Booij JC, Baas DC, Beisekeeva J, Gorgels TG, Bergen AA (2010) The dynamic nature of Bruch’s membrane. Prog Retin Eye Res 29: 1-18. 2. Retnet (2009) www.sph.uth.tmc.edu/Retnet/ 3. Olson MD (1979) Development of Bruch’s membrane in the chick: an electron micro- scopic study. Invest Ophthalmol Vis Sci 18: 329-338. 4. Hirabayashi Y, Fujimori O, Shimizu S (2003) Bruch’s membrane of the brachymorphic mouse. Med Electron Microsc 36: 139-146. 5. Hewitt AT, Nakazawa K, Newsome DA (1989) Analysis of newly synthesized Bruch’s membrane proteoglycans. Invest Ophthalmol Vis Sci 30: 478-486. 6. Ding X, Patel M, Chan CC (2009) Molecular pathology of age-related macular degen- eration. Prog Retin Eye Res 28: 1-18. 7. de Jong PT (2006) Age-related macular degeneration. N Engl J Med 355: 1474-1485. 8. Booij JC, ten Brink JB, Swagemakers SM, Verkerk AJMH, Essing AHW et al (2010) A New Strategy to Identify and Annotate Human RPE-specific Gene Expression. Plos one in press. 9. van Soest S, de Wit GM, Essing AH, ten Brink JB, Kamphuis W et al (2007) Compari- son of human retinal pigment epithelium gene expression in macula and periphery high- lights potential topographic differences in Bruch’s membrane. Mol Vis 13: 1608-1617. 10. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ et al (2009) Function- al annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 10: 164. 11. Chen L, Miyamura N, Ninomiya Y, Handa JT (2003) Distribution of the collagen IV isoforms in human Bruch’s membrane. Br J Ophthalmol 87: 212-215. 12. Marmor MF, Wolfensberger TJ (1998) The retinal pigment epithelium. 13. Chong NH, Keonin J, Luthert PJ, Frennesson CI, Weingeist DM et al (2005) Decreased thickness and integrity of the macular elastic layer of Bruch’s membrane correspond to the distribution of lesions associated with age-related macular degeneration. Am J Pathol 166: 241-251. 14. Aisenbrey S, Zhang M, Bacher D, Yee J, Brunken WJ et al (2006) Retinal pigment epithelial cells synthesize laminins, including laminin 5, and adhere to them through alpha3- and alpha6-containing integrins. Invest Ophthalmol Vis Sci 47: 5537-5544. 15. Tezel TH, Del Priore LV, Kaplan HJ (2004) Reengineering of aged Bruch’s membrane to enhance retinal pigment epithelium repopulation. Invest Ophthalmol Vis Sci 45: 3337-3348. 16. Fariss RN, Apte SS, Olsen BR, Iwata K, Milam AH (1997) Tissue inhibitor of metal- loproteinases-3 is a component of Bruch’s membrane of the eye. Am J Pathol 150: 323- 328. 17. Crabb JW, Miyagi M, Gu X, Shadrach K, West KA et al (2002) Drusen proteome analy- sis: an approach to the etiology of age-related macular degeneration. Proc Natl Acad Sci U S A 99: 14682-14687. 18. Kang SJ, Grossniklaus HE (2009) Histopathology of age-related macular degenera- tion.3-10.

134 The origin of human Bruch’s membrane proteins

19. Johnson LV, Leitner WP, Rivest AJ, Staples MK, Radeke MJ et al (2002) The Alzheim- er’s A beta -peptide is deposited at sites of complement activation in pathologic deposits associated with aging and age-related macular degeneration. Proc Natl Acad Sci U S A 99: 11830-11835. 20. Hageman GS, Zhu XL, Waheed A, Sly WS (1991) Localization of carbonic anhydrase IV in a specific capillary bed of the human eye. Proc Natl Acad Sci U S A 88: 2716- 2720. 21. Hageman GS, Luthert PJ, Victor Chong NH, Johnson LV, Anderson DH et al (2001) An integrated hypothesis that considers drusen as biomarkers of immune-mediated pro- cesses at the RPE-Bruch’s membrane interface in aging and age-related macular degen- eration. Prog Retin Eye Res 20: 705-732. 22. Yuan X, Gu X, Crabb JS, Yue X, Shadrach K et al (2010) Quantitative Proteomic Com- parison of the Macular Bruch’s Membrane/Choroid Complex from Age-related Macular Degeneration and Normal Eyes. Mol Cell Proteomics . 23. Das A, Frank RN, Zhang NL, Turczyn TJ (1990) Ultrastructural localization of ex- tracellular matrix components in human retinal vessels and Bruch’s membrane. Arch Ophthalmol 108: 421-429. 24. Del Priore LV, Tezel TH, Kaplan HJ (2006) Maculoplasty for age-related macular de- generation: reengineering Bruch’s membrane and the human macula. Prog Retin Eye Res 25: 539-562. 25. Sivaprasad s, Bailey ta, Chong VNH (2005) Bruch’s membrane and the vascular intima: is there a common basis for age-related changes and disease? clininical and expiremtal ophthalmology 33: 518. 26. Muragaki Y, Shiota C, Inoue M, Ooshima A, Olsen BR et al (1992) alpha 1(VIII)- collagen gene transcripts encode a short-chain collagen polypeptide and are expressed by various epithelial, endothelial and mesenchymal cells in newborn mouse tissues. Eur J Biochem 207: 895-902. 27. online Mendelian inheritance in man (OMIM) (2009) http://www.ncbi.nlm.nih.gov/ sites/entrez?db=OMIM 28. Myers JC, Kivirikko S, Gordon MK, Dion AS, Pihlajaniemi T (1992) Identification of a previously unknown human collagen chain, alpha 1(XV), characterized by extensive interruptions in the triple-helical region. Proc Natl Acad Sci U S A 89: 10144-10148. 29. Hashimoto T, Wakabayashi T, Watanabe A, Kowa H, Hosoda R et al (2002) CLAC: a novel Alzheimer amyloid plaque component derived from a transmembrane precursor, CLAC-P/collagen type XXV. EMBO J 21: 1524-1534. 30. Osada Y, Hashimoto T, Nishimura A, Matsuo Y, Wakabayashi T et al (2005) CLAC binds to amyloid beta peptides through the positively charged amino acid cluster within the collagenous domain 1 and inhibits formation of amyloid fibrils. J Biol Chem 280: 8596-8605. 31. Muro AF, Chauhan AK, Gajovic S, Iaconcig A, Porro F et al (2003) Regulated splic- ing of the fibronectin EDA exon is essential for proper skin wound healing and normal lifespan. J Cell Biol 162: 149-160. 32. Bing DH, Almeda S, Isliker H, Lahav J, Hynes RO (1982) Fibronectin binds to the C1q component of complement. Proc Natl Acad Sci U S A 79: 4198-4201.

135 Chapter 5

33. Farris RN, Apte SS, Luthert PJ, Bird AC, Milam AR (1998) Accumulation of tissue inhibitorof metalloproteinases-3 in human eyes with Sorsby’s fundus ystrophy or retini- tispigmentosa. Br J Ophthalmol 82: 1329-1334. 34. Apte SS, Mattei MG, Olsen BR (1994) Cloning of the cDNA encoding human tissue inhibitor of metalloproteinases-3 (TIMP-3) and mapping of the TIMP3 gene to chromo- some 22. Genomics 19: 86-90. 35. Della NG, Campochiaro PA, Zack DJ (1996) Localization of TIMP-3 mRNA expres- sion to the retinal pigment epithelium. Invest Ophthalm Vis Sci 37: 1921-1924 36. Weber BH, Vogt G, Pruett RC, Stohr H, Felbor U (1994) Mutations in the tissue inhibi- tor of metalloproteinases-3 (TIMP3) in patients with Sorsby’s fundus dystrophy. Nat Genet 8: 352-356. 37. Villiers CL, Cretin F, Lefebvre N, Marche PN, Villiers MB (2008) A new role for com- plement C3: regulation of antigen processing through an inhibitory activity. Mol Im- munol 45: 3509-3516. 38. Nakamura T, Ruiz-Lozano P, Lindner V, Yabe D, Taniwaki M et al (1999) DANCE, a novel secreted RGD protein expressed in developing, atherosclerotic, and balloon- injured arteries. J Biol Chem 274: 22476-22483. 39. Afshari FT, Fawcett JW (2009) Improving RPE adhesion to Bruch’s membrane. Eye (Lond) 23: 1890-1893. 40. Graw J (2009) Genetics of crystallins: cataract and beyond. Exp Eye Res 88: 173-189. 41. Pauleikhoff D, Harper CA, Marshall J, Bird AC (1990) Aging changes in Bruch’s mem- brane. A histochemical and morphologic study. Ophthalmology 97: 171-178. 42. Sheraidah G, Steinmetz R, Maguire J, Pauleikhoff D, Marshall J et al (1993) Correlation between lipids extracted from Bruch’s membrane and age. Ophthalmology 100: 47-51. 43. Dennis G, Jr., Sherman BT, Hosack DA, Yang J, Gao W et al (2003) DAVID: Database for Annotation, Visualization, and Integrated Discovery. Genome Biol 4: 3. 44. Hosack DA, Dennis G, Jr., Sherman BT, Lane HC, Lempicki RA (2003) Identifying biological themes within lists of genes with EASE. Genome Biol 4: R70. 45. Ingenuity (2009) www.ingenuity.com

Acknowledgements We thank Dr. L. Pels and co-workers of the Corneabank, Amsterdam, for providing donor eyes. This study was supported by grants from the Foundation Fighting Blindness (T-GE- 0101-0172), The Netherlands Organization for Scientific Research (NWO; project 948-00- 013), The Royal Netherlands Academy of Arts and Sciences (KNAW) and de Algemene Nederlandse Vereniging ter Voorkoming van Blindheid (ANVVB).

136 The origin of human Bruch’s membrane proteins

137

Identification of mutations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa 6

Judith C Booij, Ralph J Florijn, Jacoline B ten Brink, Willem Loves, Francoise Meire, Mary J van Schooneveld, Paulus TVM de Jong and Arthur AB Bergen

J Med Genet. 2005 Nov;42(11):e67-e75 Chapter 6

Abstract Purpose: To identify mutations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa (RP).

Methods: In a group of 35 unrelated patients with juvenile autosomal recessive retinitis pigmentosa (arRP), Leber’s congenital amaurosis (LCA), or juvenile isolated RP (IRP), mu- tation analysis was carried out by Denaturing High Performance Liquid Chromatography (DHPLC) analysis, followed by direct sequencing.

Results: Our arRP, LCA and IRP patients showed typical combinations of eye signs associ- ated with RP: pale optic discs, narrow arterioles, pigmentary changes, as well as nystagmus. Mutations were found in 34% of patients: in CRB1 (11%), GUCY2D (11%), RPE65 (6%), RPGRIP1 (6%). We report nine mutations, including a new combination of two mutations in CRB1, and new mutations in GUCY2D and RPGRIP1. The new GUCY2D mutation (c.3283delC, p.Pro1069ArgfsX37) is the first pathological sequence change reported in the intracellular C-terminal domain of GUCY2D, and did not lead to the commonly associated LCA, but to a juvenile RP phenotype. We established the polymorphic nature of three previ- ously described (pathological) sequence changes in AIPL1, CRB1, and RPGRIP1. Finally, seven new polymorphic changes, useful for further association studies, were found.

Conclusions: We detected new and previously described sequence changes in RP patients in CRB1, GUCY2D, and RPGRIP1; and in LCA patients in CRB1, GUCY2D, and RPE65. Our data, combined with those of the literature suggest that LCA and juvenile arRP are closely related and belong to a continuous spectrum of juvenile RP.

140 Mutations in LCA genes in juvenile retinitis pigmentosa patients

Introduction Retinitis pigmentosa (RP) denotes a group of hereditary retinal dystrophies with a worldwide prevalence of approximately 1 in 4,000. The disease is clinically and genetically very hetero- geneous[1]. The phenotype of the juvenile form of RP shows overlap with Leber’s congenital amaurosis (LCA), an autosomal recessive retinal dystrophy with a worldwide prevalence of 1 in 35,000[2], although differences exist as well. Juvenile RP patients are considered to have good central vision during the first decade of life, whereas the definition of age of onset of severe visual impairment in LCA ranges from birth to the first year of life[3,4]. RP is char- acterized by night-blindness, progressive constriction of the visual field, pale discs, retinal vascular narrowing and pigmentary changes in the retina and reduced ERG amplitudes[5]. LCA is characterized by a congenital nystagmus, a decreased pupillary response, although the fundoscopic appearance is initially normal, pigmentary changes in the retina, as well as a reduced ERG[2,6]. Various intermediate phenotypes between LCA and RP are known and are sometimes described as ‘early onset severe rod-cone dystrophy’ or ‘early onset retinal degeneration’[3]. The clinical distinction between the subtypes of RP is not always clear, and the diagnostic criteria are frequently not agreed upon between ophthalmologists[7]. In 29% of patients, RP is part of a syndrome, such as Usher’s or Bardet-Biedl’s[1]. Of the non-syn- dromal RP cases in the Netherlands, 37% is isolated (IRP); of the remainder, the transmission mode is in 30% autosomal recessive (AR), in 22% autosomal dominant (AD), and in 10% X-linked (XL)[8]. In addition, several unusual inheritance patterns, such as digenic inheri- tance[9] and uniparental isodisomy[10], in which a child inherits two copies of a single pa- rental chromosome, have been observed. LCA is most frequently inherited in an AR way[2].

The molecular genetic diversity within the different RP phenotypes is also well documented in the literature[3,4,11]. At least 17 genes are currently known to be involved in arRP, while five additional loci have been identified[12]. So far, mutations can only be identified in ap- proximately 50% of arRP and LCA patients. Eight LCA genes have been cloned so far: AIPL1[13], CRB1[14], CRX[15], GUCY2D[16], RDH12[17], RPE65[18], RPGRIP1[19], and TULP1[20,21]. The genetic overlap between arRP and LCA is illustrated by the fact that all but three of these genes (AIPL1, CRX and RPGRIP1) were previously implicated in arRP (CRB1[14,22], GUCY2D[16,23], RDH12[24,25], RPE65[4,18], and TULP1[26-28]). To illustrate the diversity and complexity of genotype-phenotype correlations in RP and LCA, two examples are mentioned here. Previously, a correlation was suggested between mutations in the RPE65 gene and a phenotype consisting of severe visual impairment within the first year of life, with a measurable visual acuity at the age of six to ten years, and a con- genital nystagmus in three out of four patients. However, no such correlation was observed in another study[3,4]. Similarly, two parents, diagnosed with RP, had a daughter who carried a homozygous RPE65 mutation and who was reported to have LCA (absent or severely di- minished vision within the first year of life)[4]. In summary, there appears to be significant overlap between juvenile arRP and LCA in clini- cal and genetic sense. This makes the clinical distinction between juvenile arRP and LCA difficult, if not impossible[7,12]. We hypothesized that mutations in the five most common LCA genes could also be responsible for juvenile arRP[7,12,28,29]. Therefore we tested our

141 Chapter 6 arRP, IRP and LCA patients for mutations in the following “LCA” disease genes: AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1.

Methods Patients This study was performed in agreement with the declaration of Helsinki. Patients included in this study or their legal representatives gave permission for blood to be drawn for mutation detection. All patients diagnosed with arRP and IRP with a disease onset before age 20 years, as well as all patients with an LCA diagnosis, presented to the Netherlands Ophthalmic Re- search Institute (NORI) were included in this study. Most patients were Dutch, four patients were Belgian, one patient originated from Turkey, one patient had a Mexican and a Dutch parent and one had an Indonesian and a Dutch parent. The diagnosis of RP and LCA were based on a combination of the following criteria. For RP, the diagnostic criteria were: (subjective) night blindness, a pale optic disc, narrow arteri- oles, pigmentary changes, ERG amplitude reduction (>50%), and visual field abnormalities, compatible with RP1. Patients were considered to have juvenile RP when their first reported symptoms occurred before the age of 20. For LCA, the criteria were: signs of blindness or severe visual impairment from birth or within the first year of life, an ERG reduction of more than 50%, and a congenital nystagmus[1]. Patients were diagnosed with AR inheritance when they had two unaffected parents and one or more affected siblings or consanguineous parents. Patients were diagnosed with IRP when no affected relatives were known, and there was no parental consanguinity, or no inheritance pattern could be established. Patients were excluded when their family tree showed evidence of AD or XL inheritance[1], or when their RP or LCA was part of a syndrome.

Table 1. Characteristics of cloned genes implicated in recessive LCA according to the literature.

gene chromosome described in found in (%) of ref * primer ref † location patients AIPL1 17p13.2 LCA 5-10 13 this study CRB1 1q31.3 LCA, arRP 9-13 14 14 CRX 19q13.32 LCA 1-3 15 GUCY2D 17p13.1 LCA, arRP 10-20 16 16 RDH12 14q24.1 LCA, arRP 4 17 RPE65 1q31.2 LCA, arRP 6-16 18 4 RPGRIP1 14q11.2 LCA 4-6 19 31 TULP1 6p21.31 LCA,arRP several families 20 21 arRP: autosomal recessive retinitis pigmentosa, IRP: isolated retinitis pigmentosa, LCA: Leber’s congenital am- aurosis. *references first describing the cloning of the respective genes, † references describing the primers we used.

142 Mutations in LCA genes in juvenile retinitis pigmentosa patients

Mutation detection The DNA isolated from peripheral blood was stored at 4ºC until use[30]. The following prim- ers were used for the amplification of AIPL1: exon 1 forward (1F) cctggtcccctgtcttcttt, exon 1 reverse (1R) tgttgaaagctgctgtgggg, 2F ggccttgaacagtgtgtcta, 2R gagcccagaaaagactagtc, 3F ggccttttatggcccaccta, 3R ctgtccctctccagtgctgg, 4F tggggtccctgcctcactga, 4R ccccagagtcagc- gccactt, 5F aagtggcgctgactctgggg, 5R tgtctccgtggccctgggct, 6F cctctgaggctgggaaggga, 6R gaccaggccacttgctccct. Primers used for CRB1, GUCY2D, RPE65, and RPGRIP1 along with their characteristics are referred to in Table 1.

The length of the PCR product was checked on a 2% agarose gel, after which it was mixed with an equal volume of PCR product from a healthy control person in order to detect ho- mozygous sequence changes. This was followed by a heteroduplex formation step before samples were analyzed for mutations using Denaturing High Performance Liquid Chroma- tography analysis (DHPLC) (Transgenomic) at two or three different temperatures. Exons showing a DHPLC pattern that differed from a healthy control, and the healthy control itself were re-amplified and purified using a QIA-quick PCR purification kit (Qiagen) for direct sequencing. A sequence reaction (25 cycles consisting of 10 seconds at 96 degrees, 5 seconds at 50 degrees and 4 minutes at 60 degrees Celsius) was performed using forward and reverse primers separately (Table 1). Sequence products were precipitated with 75% isopro- panol and dissolved in template suppression reagent after which the products were denatured at 94 degrees Celsius for 2 minutes. Samples were subsequently analyzed using the ABI- 310 (Applied Biosystems). We described mutations using the following Genbank reference cDNA sequences[32]: AIPL1: gi6716706; CRB1: NM_012076.2; GUCY2D: NM_000180.1; RPGRIP1: NM_020366.2; RPE65: gi3077820; choosing the first A of the ATG as nucleotide +1, and the ATG as codon +1, except for CRB1 where 135 nucleotides before the first A of the ATG was considered as nucleotide +1[33].

Definition of pathogenicity We considered a sequence variation to be pathogenic when we found it exclusively in pa- tients and not in at least 120 from 60 control persons. In addition, we made a literature based prediction of the effect of a new mutation, using the description of mutations in or the function of similar domains. We considered sequence variations polymorphic when their allele frequency was not significantly higher in patients than in control persons, or when they did not co-segregate with the disease phenotype in a family. Our patient group consisted of 35 patients, our control group of at least 60 healthy persons (120 chromosomes).

143 Chapter 6

Results Patients Our mutation screening panel consisted of 17 arRP, 9 LCA, and 9 IRP patients. Table 2 shows the clinical characteristics of the 12 patients in whom we found mutations.

Table 2. Clinical characteristics of juvenile arRP, LCA and IRP patients in whom mutations were identified in the present study.

fundoscopic picture changes † - patient diagnosis age at last examination (years)* age of onset (default: years, m: months) * pale optic disc arteri narrow oles pigment changes‡ >= ERG reduction 50% † visual field † nystagmus †

8432 arRP 67 <20 1 1 1 - 1 0 25378 arRP 14 5 1 1 1 1 0 0 25855 arRP 67 20 1 1 1 1 1 0 9359 LCA 30 4 m 1 1 1 1 1 1 22802 LCA 43 0 1 1 1 1 - 1 24670 LCA 35 6 m 1 1 1 1 1 1 26669 LCA 3 0 0 0 1 1 - 1 25140 IRP 15 1.5 0 0 0 1 1 0 25474 IRP 32 <20 0 0 0 - - - 25556 IRP 13 1.5 0 0 0 1 1 1 25718 IRP 24 14 1 0 1 1 1 0 26425 IRP 9 2 1 1 1 1 1 - arRP: autosomal recessive retinitis pigmentosa, LCA: Leber’s congenital amaurosis, IRP: isolated retinitis pig- mentosa. * age in years or months (m). † 0 = no changes, 1= changes reported or measured, - = missing data or not measured. ‡ pigmentary changes could be present either peripherally or throughout the entire retina, and could have the appearance of either typical bone spicules or aspecific pigmentary abnormalities.

Visual field changes corresponding with RP were seen in arRP patients, IRP patients, and LCA patients. As expected, an ERG reduction of at least 50% was reported in nearly all RP and LCA patients (data were missing on two patients). Nystagmus, corresponding with an LCA phenotype, was seen in all LCA patients and in one IRP patient. The RP and LCA pa- tients, in whom no mutations were found, showed similar overlapping clinical features (data not shown).

144 Mutations in LCA genes in juvenile retinitis pigmentosa patients

Mutations We identified nine different mutations in 12 patients (34%), in the CRB1 (11%), GUCY2D (11%), RPE65 (6%), and RPGRIP1 (6%) genes. In three of these patients two mutations were found (9%), one patient had a homozygous mutation, two patients had a compound heterozygous mutation. In nine patients we detected a mutation in a single allele only. We found a mutation in three arRP patients (17%), in four LCA patients (44%), and in five IRP patients (56%). Two of the mutations identified in this study are new, and we describe a new combination of two mutations. All mutations are presented in Table 3.

Table 3. Mutations in patients with juvenile arRP, LCA or IRP.

sequence gene variation 39 predicted effect 39 patient disease reference† CRB1 [c.2536A>T]+ [p.Lys801X]+ 25378 arRP 14,22, this study‡ [c.2978G>A] [p.Cys948Tyr] [c.2978G>A] +[?] [p.Cys948Tyr]+[?] 22802 LCA 14 [c.2978G>A] +[?] [p.Cys948Tyr] +[?] 25718 IRP 14 [c.2978G>A] +[?] [p.Cys948Tyr] +[?] 26425 IRP 14 GUCY2D [c.3283delC]+[?] [p.Pro1069ArgfsX37]+[?] 25556 IRP this study [c.2176C>T]+[?] [p.Pro701Ser]+[?] 25855 arRP 34 [c.2176C>T]+[?] [p.Pro701Ser]+[?] 8432 arRP 34 [c.2377C>T]+[?] [p.Arg768Trp]+[?] 26669 LCA 40 RPE65 [c.1102T>C]+ [p. Tyr368His]+ 24670 LCA 36,41 [c.11+5G>A]* [splice site] [c.1102T>C]+ [p. Tyr368His]+ 9359 LCA 3 [ c.1102T>C] [p. Tyr368His] RPGRIP1 [c.1614_1623del]+[?] [p.Glu538Glufs2]+[?] 25140 IRP this study [c.3341A>G]+[?] [p.Asp1114Gly]+[?] 25474 IRP 31 * traditional name of RPE65:c.11+5G>A is RPE65:IVS1+5G>A. † The first description of the mutation in the literature. ‡ This combination of mutations has not been described in the literature. arRP: autosomal recessive retinitis pigmentosa, LCA: Leber’s congenital amaurosis, IRP: isolated retinitis pig- mentosa.

The first new mutation identified in this study was a heterozygous deletion resulting in a frameshift in GUCY2D:c.3283delC, p.Pro1069ArgfsX37 in IRP patient 25556. This male patient had a disease onset at age 1.5 years. At age 13 he had a visual acuity of 0.1, a severely

145 Chapter 6 reduced ERG, a limited visual field and nystagmus. Co-segregation of the deletion in the family could not be determined due to the unavailability of parental DNA. However, the deletion was not found in our control population. We also identified GUCY2D mutations in arRP patients 25855 and 8432 (p.Pro701Ser)[34] and in LCA patient 26669 (p.Arg768Trp) [35]. The second new mutation was a deletion causing a frameshift in RPGRIP1:c1614_1623del, p.Glu538Glufs2 in IRP patient 25140. Visual problems in this patient were noted in the sec- ond year of life, she had a visual acuity of 0.1 at the age of 15, and experienced night blind- ness and color vision impairments. Fundoscopy revealed an atrophic macular area, and the ERG was severely reduced. We did not detect this deletion in our control population. In ad- dition, we found one known RPGRIP1 mutation in IRP patient 25474 (p.Asp1114Gly)[31]. We found a new combination of two mutations in CRB1 in arRP patient 25378: [c.2536A>T, p.Cys948Tyr][14]+[c.2978G>A, p.Lys801X][22]. The age of onset was five years, at age 14, the visual acuity was finger counting, and subjective night blindness was reported. Earlier data on visual acuity and night blindness were missing. The ERG was severely reduced and the fundoscopic picture showed narrow arterioles, a pale optic disc and pigmentary changes in the periphery. CRB1 mutations were also found in LCA patient 22802 and in IRP patients 25718 and 26425 (p.Cys948Tyr)[14]. In the RPE65 gene we found mutations in LCA patient 9359 ([p.Tyr368His]+[c.11+5G>A]) [35,36] and in LCA patient 24670 (p.Tyr368His homozygously). We confirm the pathogenic nature of the RPE65 c.11+5G>A sequence change. It is located in a donor splice site where the G nucleotide is predicted to be conserved in 84% of primate splice sites[35-37]. In addi- tion, we do not find it in our controls[38].

Polymorphic sequence changes We observed seven new and 15 known polymorphisms in all five genes. In addition, we established or confirmed the polymorphic nature of three sequence changes previously de- scribed as pathogenic mutations. Table 4 summarizes the polymorphisms along with their allele frequency in our patient population.

Table 4. Frequency of known and new polymorphisms in arRP, LCA and IRP patients in this study.

Allele frequen- exon/ sequence varia- predicted cy in patients gene intron tion[39] effect[39] (%) reference AIPL1 ex 3 c.286G>A p.Val96Ile 1.3 13, this study * CRB1 IVS3 c.283-35T>C 2.6 this study IVS4 c.330+34C>T 5.3 this study IVS4 c.330-54C>A 2.6 this study ex 2 c.614T>C p.Ile205Thr 1.3 44, this study *

146 Mutations in LCA genes in juvenile retinitis pigmentosa patients

Allele frequen- exon/ sequence varia- predicted cy in patients gene intron tion[39] effect[39] (%) reference IVS1 c.24-12A>T 7.9 44 GUCY2D ex 3 c.814C>T p.His247His 19.7 this study ex 9 c.2011C>T p.Leu646Leu 1.3 this study ex 2 c.134T>C p.Trp21Arg 10.5 46 ex 2 c.227G>T p.Ala52Ser 9.2 16 ex 10 c.2182G>A p.Ala703Ala 15.8 47 ex 12 c.2419C>A p.Leu782His 15.8 47 ex 20 c.3528C>T 39.5 47 IVS13 c.859+37G>T 14.5 47 RPE65 ex 10 c.1056G>A p.Glu352Glu 7.9 4 IVS12 c.446+20A>C 3.9 4 RPGRIP1 IVS6 c.302-16_-15insATA 6.6 this study IVS6 c.907-153_-154delGG 10.5 this study ex 4 c.525A>G p.Pro175Pro 5.3 19 ex 4 c.574A>G p.Lys192Glu 9.2 19 ex 13 c.1639G>T p.Ala547Ser 14.5 19,45, this study * ex 14 c.1797G>A p.Pro599Pro 18.4 19 ex 15 c.2292G>A p.Ala764Ala 2.6 31 ex 18 c.3097G>C p.Gln1033Glu 23.7 19 ex 22 c.3546C>T p.Asp1182Asp 6.6 19 Polymorphisms described in bold have either not been described in the literature or have been described as disease causing missense mutations (*)

We identified seven new polymorphic sequence changes, which are described below. Three new intronic polymorphisms were located in the CRB1 gene (c.283-35T>C, c.330+34C>T, and c.330-54C>A), and two new silent polymorphisms were found in GUCY2D (c.814C>T, p.His247His and c.2011C>T, p.Leu646Leu). In RPGRIP1 a new insertion (c.302-16_-15in- sATA), and a new deletion (c.907-153_-154delGG) was found. Bioinformatic analysis of all intronic sequence changes did not lead to changes in the splice sites. Analysis of the silent polymorphisms did not reveal predicted changes in exonic splicing enhancers (ESE)[42,43].

147 Chapter 6

We identified three sequence changes of uncertain pathogenic nature. In the literature, these are tentatively described as polymorphisms. The first sequence change wasAIPL1 :c.286G>A, p.Val96Ile[13] with an allele frequency in patients of 1%, and in matched control persons of 7% (n=122 chromosomes). The second sequence change, CRB1:c.614T>C, p.Ile205Thr[44], was found only once in our patient population (allele frequency 1%). We did not find it in our controls. However, we found no co-segregation with the disease phenotype in arRP fam- ily 21724 (Figure 1). The third sequence change we identified was RPGRIP1:c.1639G>T, p.Ala547Ser[45] with similar allele frequencies of 14% in patients and 13% in controls. Interestingly, in three patients with that Ala547Ser variation, we detected an additional het- erozygous pathogenic mutation: (GUCY2D:p.Pro701Ser (arRP patient 25855), RPGRIP1:p. Asp1114Gly (IRP patient 25474). IRP patient 25140, as well as her unaffected father, had the RPGRIP1:p.Ala547Ser variation in combination with the RPGRIP1:p.Glu538Glufs2 mu- tation. The latter indicates that the p.Ala547Ser sequence change in combination with the p.Glu538Glufs2 mutation is not sufficient to lead to an RP phenotype.

Figure 1. Family tree of a large consanguineous family with arRP with three different CRB1 sequence changes.

I:1 I:2 I:3 I:4

II:1 II:2 II:3 II:4 II:5 II:6 II:7 II:8

III:1 III:2 III:3 III:4 III:5 III:6 III:7 III:8 III:9 III:10

IV:1 IV:2 IV:3 IV:4 IV:5 IV:6 IV:7 IV:8

V:1 V:2 V:3 V:4

VI:1 VI:2 VI:3 VI:4 VI:5 VI:6 VI:7 VI:8 VI:9 VI:10 VI:11 VI:12 VI:13 VI:14 Pathogenic mutations Ser403X ------Arg764Cys ------

Unknown pathogenicity - Ile205Thr - - - Ile205Thr Ile205Thr - Ile205Thr - - Ile205Thr - -

Patient VI:I, with RP with PPRPE (preservation of the para-arteriolar retinal pigment epithelium, or RP12), had two mutations in the CRB1 gene (p.Ser403X and p.Arg764Cys), both identified in a previous study.14 The p.Ile205Thr sequence change, previously reported as a disease causing mutation and as a polymorphism,44,54 was found in patients VI:6, and VI:12, and in unaffected siblings VI:2, VI:7 and VI:9. However, the sequence change was not found in patient VI:4, which means that two yet unidentified pathogenic mutations are present in this patient. Unless one of these was a spontaneous mutation in patient VI:4, one of the unaffected parents (unavailable for mutation screening) carried two disease causing mutations, making him or her affected. In summary, even though we did not detect this sequence change in our control population, the lack of co-segregation with the disease phe- notype in this family leads us to the conclusion that the p.Ile205Thr sequence change is most likely polymorphic. Solid squares: RP patients.

Discussion Both clinically and genetically, there was a considerable overlap between our RP patients and our LCA patients. Clinically, visual field changes, as seen in RP, were present in RP pa- tients and in two LCA patients. A nystagmus, frequently seen in LCA, was observed in LCA

148 Mutations in LCA genes in juvenile retinitis pigmentosa patients patients as well as an IRP patient. In addition, fundoscopic changes were seen in arRP, IRP and LCA patients. In summary, CRB1 and GUCY2D mutations were found in RP patients as well as LCA pa- tients, RPE65 mutations were found only in LCA patients, and RPGRIP1 mutations, previ- ously only described in LCA, were newly identified in two RP patients.

Mutations: frequency and distribution In this study we found nine mutations in 34% of our 35 patients. We identified mutations in 56% of our LCA patients, which corresponds well to the literature (50%)[29,29], although in contrast to the current study, most studies did not screen for all known genes at once. In 27% of our arRP patients we found one or two mutations, which is less than what is reported in the literature[12]. The latter may be due to the fact that we screened our patients for LCA gene mutations, and not for all the known RP gene mutations. Screening our arRP patients for additional, known RP gene mutations will most likely lead to the detection of several more mutations. In spite of the recessive nature of the disease, a second disease-causing mutation was not found in nine out of twelve patients. The design of our DHPLC analysis (sensitivity close to 99 %) makes it unlikely that a second exonic mutation in one of the screened genes was left undetected[48,49]. Obviously, it remains possible that promoter or regulatory sequences of the screened genes contain yet unidentified mutations. The spectrum of mutations in our study (CRB1 (11%), GUCY2D (11%), RPE65 (6%), and RPGRIP1 (6%)) shows only slight differences compared to those reported in the literature. This indicates that the gene distribution in our cohort is similar to juvenile RP and LCA popu- lations worldwide[13,14,16,18,19].

New mutations Two new mutations were identified in this study. The first is the first the mutation reported in the intracellular C-terminal domain of GUCY2D: c.3283delC, p.Pro1069ArgfsX37, in IRP patient 25556. The deletion creates a stop codon at position 1106 of the cDNA, which results in a larger, most likely unstable, mRNA or a dysfunctional protein. GUCY2D is a well con- served protein, containing an extracellular ANF-receptor domain, a transmembrane domain, an intracellular protein kinase homology domain and an intracellular guanylate cyclase do- main[50]. Mutations have been described in all these domains[16,34], except for the intracel- lular C-terminal domain. The function of the latter domain is unknown. GUCY2D is one of the enzymes indirectly responsible for the regulation of phototransduction in the photorecep- tor cells through a regulation of cGMP levels[51]. An altered GUCY2D protein could disturb this process leading to an RP phenotype, as seen in this IRP patient. Patient 25556 had an age of onset and visual field changes compatible with an RP phenotype. However, the lack of fun- doscopic changes and the presence of a nystagmus indicate a phenotypic overlap with LCA. The second new mutation identified in this study isRPGRIP1 :c.1614_1623del, p.Glu538Glufs2 in IRP patient 25140. A stop codon is introduced at position 540 of the cDNA, which will prevent the transcription of the protein kinase C conserved region 2 (C2) and the RPGR binding domain[52]. The RPGRIP1 protein is localized in the connecting cilia of human

149 Chapter 6 cone and rod photoreceptors that connect inner and outer segments, where it binds RPGR to the cilium[52]. RPGR plays a role in the maintenance of photoreceptor viability[53]. RPGR mutations were previously found in X-linked RP, and RPGRIP1 mutations are known to cause LCA[19,31]. Therefore we consider this RPGRIP1 deletion likely to be involved in RP, even though a second mutation was not found in this patient (25140). Patient 25140 had a constricted visual field, a visual acuity at age 15 of 0.1, and no nystagmus, which pleads against the diagnosis of LCA. However, the absence of fundoscopic changes, the relatively early onset (1.5 years) and the presence of color vision defects are more typical for an LCA- related phenotype. This patient is a typical example of the phenotypic overlap that exists between juvenile RP and LCA[7].

Polymorphic sequence changes We established or confirmed the polymorphic nature of three sequence changes previously reported as pathogenic mutations. The first (AIPL1:c.286G>A, p.Val96Ile) was previously described as a disease causing mutation[13]. We consider it to be a rare neutral polymorphic change because of its high allele frequency (7%) in our control persons. The second sequence change we report to be polymorphic (CRB1: c.614T>C, p.Ile205Thr) was initially described as a disease causing mutation in a Spanish arRP family[44]. More recently it was reported in a Dutch study as a polymorphic sequence change[54] We find it heterozygously in a single family where is does not co-segregate with the disease phenotype (Figure 1). Moreover, the change is absent in 192 control chromosomes, which lead us to suggest that Ile205Thr is a rare neutral sequence change. The pathogenic nature of the third sequence change (RP- GRIP1: c.1639G>T, p.Ala547Ser) is unclear in the literature. It was described as a polymor- phism[19], and a disease causing mutation[45]. The high frequency of this sequence change in our control group (13%) suggests RPGRIP1:p.Ala547Ser is indeed a neutral polymor- phism. We found it heterozygously along with a homozygous disease causing RPE65 muta- tion in LCA patient 9359. Furthermore, the presence of Ala547Ser in the unaffected parents of LCA patient 9359 and IRP patient 25140, along with other heterozygous mutations, sug- gests it does not cause a disease phenotype.

In conclusion, we observed a large overlap in the mutation spectrum and the clinical signs in the juvenile arRP, LCA, and juvenile IRP patients both in our study and in the litera- ture[4,7,14,22,34,36,55]. Our clinical data, combined with the mutation spectrum found in our patients, do not provide hard evidence that a clear distinction between juvenile RP and LCA can be made. We expect that eventually, more clarity will come from molecular genetic analysis of the underlying gene defects.

150 Mutations in LCA genes in juvenile retinitis pigmentosa patients

References 1. Haim M. Epidemiology of retinitis pigmentosa in Denmark. Acta Ophthalmol Scand Suppl 2002; 1-34. 2. Koenekoop RK. An overview of Leber congenital amaurosis: a model to understand human retinal development. Surv Ophthalmol 2004; 49: 379-398. 3. Lorenz B, Gyurus P, Preising M et al. Early-onset severe rod-cone dystrophy in young children with RPE65 mutations. Invest Ophthalmol Vis Sci 2000; 41: 2735-2742. 4. Morimura H, Fishman GA, Grover SA et al. Mutations in the RPE65 gene in patients with autosomal recessive retinitis pigmentosa or leber congenital amaurosis. Proc Natl Acad Sci U S A 1998; 95: 3088-3093. 5. Merin S, Auerbach E. Retinitis pigmentosa. Surv Ophthalmol 1976; 20: 303-346. 6. Perrault I, Rozet JM, Gerber S et al. Leber congenital amaurosis. Mol Genet Metab 1999; 68: 200-208. 7. Rivolta C, Sharon D, DeAngelis MM, Dryja TP. Retinitis pigmentosa and allied dis- eases: numerous diseases, genes, and inheritance patterns. Hum Mol Genet 2002; 11: 1219-1227. 8. van den Born LI, Bergen AA, Bleeker-Wagemakers EM. A retrospective study of reg- istered retinitis pigmentosa patients in The Netherlands. Ophthalmic Paediatr Genet 1992; 13: 227-236. 9. Kajiwara K, Berson EL, Dryja TP. Digenic retinitis pigmentosa due to mutations at the unlinked peripherin/RDS and ROM1 loci. Science 1994; 264: 1604-1608. 10. Thompson DA, McHenry CL, Li Y et al. Retinal dystrophy due to paternal isodisomy for chromosome 1 or chromosome 2, with homoallelism for mutations in RPE65 or MERTK, respectively. Am J Hum Genet 2002; 70: 224-229. 11. Phelan JK, Bok D. A brief review of retinitis pigmentosa and the identified retinitis pigmentosa genes. Mol Vis 2000; 6: 116-124. 12. (www.sph.uth.tmc.edu/retnet/disease.htm) 13. Sohocki MM, Perrault I, Leroy BP et al. Prevalence of AIPL1 mutations in inherited retinal degenerative disease. Mol Genet Metab 2000; 70: 142-150. 14. den Hollander AI, ten Brink JB, de Kok YJ et al. Mutations in a human homologue of Drosophila crumbs cause retinitis pigmentosa (RP12). Nat Genet 1999; 23: 217-221. 15. Freund CL, Gregory-Evans CY, Furukawa T et al. Cone-rod dystrophy due to mutations in a novel photoreceptor-specific homeobox gene (CRX) essential for maintenance of the photoreceptor. Cell 1997; 91: 543-553. 16. Perrault I, Rozet JM, Calvas P et al. Retinal-specific guanylate cyclase gene mutations in Leber’s congenital amaurosis. Nat Genet 1996; 14: 461-464. 17. Haeseleer F, Jang GF, Imanishi Y et al. Dual-substrate specificity short chain retinol dehydrogenases from the vertebrate retina. J Biol Chem 2002; 277: 45537-45546. 18. Marlhens F, Bareil C, Griffoin JM et al. Mutations in RPE65 cause Leber’s congenital amaurosis. Nat Genet 1997; 17: 139-141. 19. Dryja TP, Adams SM, Grimsby JL et al. Null RPGRIP1 alleles in patients with Leber congenital amaurosis. Am J Hum Genet 2001; 68: 1295-1298.

151 Chapter 6

20. North MA, Naggert JK, Yan Y, Noben-Trauth K, Nishina PM. Molecular characteriza- tion of TUB, TULP1, and TULP2, members of the novel tubby gene family and their possible relation to ocular diseases. Proc Natl Acad Sci U S A 1997; 94: 3128-3133. 21. Hagstrom SA, North MA, Nishina PL, Berson EL, Dryja TP. Recessive mutations in the gene encoding the tubby-like protein TULP1 in patients with retinitis pigmentosa. Nat Genet 1998; 18: 174-176. 22. den Hollander AI, Heckenlively JR, van den Born LI et al. Leber congenital amaurosis and retinitis pigmentosa with Coats-like exudative vasculopathy are associated with mutations in the crumbs homologue 1 (CRB1) gene. Am J Hum Genet 2001; 69: 198- 203. 23. Perrault I, Hanein S, Gerber S et al. A novel mutation in the GUCY2D gene responsible for an early onset severe RP different from the usual GUCY2D-LCA phenotype. Hum Mutat 2005; 25: 222. 24. Perrault I, Hanein S, Gerber S et al. Retinal dehydrogenase 12 (RDH12) mutations in leber congenital amaurosis. Am J Hum Genet 2004; 75: 639-646. 25. Janecke AR, Thompson DA, Utermann G et al. Mutations in RDH12 encoding a photo- receptor cell retinol dehydrogenase cause childhood-onset severe retinal dystrophy. Nat Genet 2004; 36: 850-854. 26. Banerjee P, Kleyn PW, Knowles JA et al. TULP1 mutation in two extended Dominican kindreds with autosomal recessive retinitis pigmentosa. Nat Genet 1998; 18: 177-179. 27. Gu S, Lennon A, Li Y et al. Tubby-like protein-1 mutations in autosomal recessive reti- nitis pigmentosa. Lancet 1998; 351: 1103-1104. 28. Hanein S, Perrault I, Gerber S et al. Leber congenital amaurosis: comprehensive survey of the genetic heterogeneity, refinement of the clinical definition, and genotype-pheno- type correlations as a strategy for molecular diagnosis. Hum Mutat 2004; 23: 306-317. 29. Cremers FP, van den Hurk JA, den Hollander AI. Molecular genetics of Leber congeni- tal amaurosis. Hum Mol Genet 2002; 11: 1169-1176. 30. Miller SA, Dykes DD, Polesky HF. A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Res 1988; 16: 1215. 31. Gerber S, Perrault I, Hanein S et al. Complete exon-intron structure of the RPGR-inter- acting protein (RPGRIP1) gene allows the identification of mutations underlying Leber congenital amaurosis. Eur J Hum Genet 2001; 9: 561-571. 32. http://www.retina-international.org/sci-news/mutation.htm 33. den Dunnen JT, Antonarakis SE. Mutation nomenclature extensions and suggestions to describe complex mutations: a discussion. Hum Mutat 2000; 15: 7-12. 34. Dharmaraj SR, Silva ER, Pina AL et al. Mutational analysis and clinical correlation in Leber congenital amaurosis. Ophthalmic Genet 2000; 21: 135-150. 35. Lotery AJ, Namperumalsamy P, Jacobson SG et al. Mutation analysis of 3 genes in patients with Leber congenital amaurosis. Arch Ophthalmol 2000; 118: 538-543. 36. Thompson DA, Gyurus P, Fleischer LL et al. Genetics and phenotypes of RPE65 muta- tions in inherited retinal degeneration. Invest Ophthalmol Vis Sci 2000; 41: 4293-4299. 37. Perrault I, Rozet JM, Ghazi I et al. Different functional outcome of RetGC1 and RPE65 gene mutations in Leber congenital amaurosis. Am J Hum Genet 1999; 64: 1225-1228.

152 Mutations in LCA genes in juvenile retinitis pigmentosa patients

38. Shapiro MB, Senapathy P. RNA splice junctions of different classes of eukaryotes: se- quence statistics and functional implications in gene expression. Nucleic Acids Res 1987; 15: 7155-7174. 39. http://www.genomic.unimelb.edu.au/mdi/mutnomen/ 40. Lotery AJ, Namperumalsamy P, Jacobson SG et al. Mutation analysis of 3 genes in patients with Leber congenital amaurosis. Arch Ophthalmol 2000; 118: 538-543. 41. Lotery AJ, Namperumalsamy P, Jacobson SG et al. Mutation analysis of 3 genes in patients with Leber congenital amaurosis. Arch Ophthalmol 2000; 118: 538-543. 42. http://www.genet.sickkids.on.ca/~ali/splicesitefinder.html 43. http://genes.mit.edu/burgelab/rescue-ese/ 44. Bernal S, Calaf M, Garcia-Hoyos M et al. Study of the involvement of the RGR, CRPB1, and CRB1 genes in the pathogenesis of autosomal recessive retinitis pigmentosa. J Med Genet 2003; 40: e89. 45. Hameed A, Abid A, Aziz A et al. Evidence of RPGRIP1 gene mutations associated with recessive cone-rod dystrophy. J Med Genet 2003; 40: 616-619. 46. Rozet JM, Perrault I, Gerber S et al. Functional Analyses Of RetGC1 Missense Muta- tions Identified In Leber’s Congenital Amaurosis. Invest Ophthal Vis Sci 2000; S533. 47. Loyer M, Pina AL, El Hilali E et al. Novel Mutations In The Retinal Guanylyl Cy- clase Gene In New Populations Of Patients With Lebers’s Congenital Amaurosis. Invest Ophthalmol Vis Sci 1998; 39: S295. 48. Yamanoshita O, Kubota T, Hou J et al. DHPLC is superior to SSCP in screening p53 mutations in esophageal cancer tissues. Int J Cancer 2005; 114: 74-79. 49. Xiao W, Oefner PJ. Denaturing high-performance liquid chromatography: A review. Hum Mutat 2001; 17: 439-474. 50. http://www.sanger.ac.uk/cgi-bin/Pfam/swisspfamget.pl?name=Q02846 51. Perrault I, Rozet JM, Gerber S et al. Spectrum of retGC1 mutations in Leber’s congeni- tal amaurosis. Eur J Hum Genet 2000; 8: 578-582. 52. Hong DH, Yue G, Adamian M, Li T. Retinitis pigmentosa GTPase regulator (RPGRr)- interacting protein is stably associated with the photoreceptor ciliary axoneme and an- chors RPGR to the connecting cilium. J Biol Chem 2001; 276: 12091-12099. 53. Hong DH, Pawlyk BS, Shang J et al. A retinitis pigmentosa GTPase regulator (RPGR)- deficient mouse model for X-linked retinitis pigmentosa (RP3). Proc Natl Acad Sci U S A 2000; 97: 3649-3654. 54. den Hollander AI, Davis J, van der Velde-Visser SD et al. CRB1 mutation spectrum in inherited retinal dystrophies. Hum Mutat 2004; 24: 355-369. 55. Lotery AJ, Jacobson SG, Fishman GA et al. Mutations in the CRB1 gene cause Leber congenital amaurosis. Arch Ophthalmol 2001; 119: 415-420.

ACKNOWLEDGEMENTS We thank all patients for their participation in this study.

153

Course of visual decline in rela- tion to the Best1 genotype in vitel- liform macular dystrophy 7

Judith C Booij, Camiel JF Boon, Mary J van Schooneveld, Jacoline B ten Brink, Arne Bak- ker, Paulus TVM de Jong, Carel B Hoyng, Arthur AB Bergen, Caroline CW Klaver

Ophthalmology, 2010 april 8 epub ahead of print Chapter 7

Abstract Purpose: To describe the disease course in vitelliform macular dystrophy (VMD) patients with a Best1 mutation, and to determine the association between Best1 genotype and visual prognosis.

Design: Consecutive case series. Participants: Fifty-three VMD patients with Best1 muta- tions from 27 Dutch families, aged 11 to 87 years.

Methods: Best-corrected visual acuity (VA), fundus appearance, and Arden ratio on the electro-oculogram (EOG) during clinical follow up were assessed from medical records. Mutation analysis of the Best1 gene was performed on DNA samples using denaturing high pressure liquid chromatography (dHPLC) and direct sequencing.

Main Outcome Measures: Cumulative life-time risk of visual decline below 0.5, 0.3, and 0.1, for the entire group and stratified for genotype.

Results: Median age of onset of visual complaints was 33 years (range: 2-78). The cumula- tive risk of VA below 0.5 (20/40) was 50% at 55 years and 75% at 66 years. The cumulative risk of decline <0.3 (20/63) was 50% by age 66 and 75% by age 74. Two patients progressed to VA<0.1 (20/200). Fourteen different mutations were found. Most patients (96%) had mis- sense mutations; the Thr6Pro, Ala10Val and Tyr227Asn mutations were most common. Vi- sual decline was significantly faster in patients with an Ala10Val mutation than either the Thr6Pro or the Tyr227Asn mutation (p=0.001).

Conclusions: Age of onset of visual complaints varies greatly among VMD patients. All patients show a gradual decrease in VA and most progress to visual impairment at a relatively late age. Our data suggest a phenotype-genotype correlation, as the Ala10Val mutation has a more rapid disease progression than other common mutations.

156 Genotype-phenotype correlation in vitelliform macular dystrophy

Introduction Vitelliform macular dystrophy (VMD) is an autosomal dominant macular disease with an estimated prevalence of 1 in 10,000[1]. The diagnosis is generally established when a vitel- liform or egg yolk-like lesion is present in the macula. This lesion consists of lipofuscin located within and below the retinal pigment epithelium (RPE). The egg yolk stage is usually followed by a scrambled egg appearance, which is caused by break down of the deposits. In later stages, the macula may become atrophic or show scar formation as a result of cell death of the RPE and photoreceptors. The clinical diagnosis of VMD is generally confirmed by a decreased light peak-dark trough ratio (Arden ratio) on the electro-oculogram (EOG). VMD is caused by mutations in a single gene, Best1 (NM_004183), located on chromosome 11q13[2-5]. This gene, formerly known as the VMD2 gene, is transcribed into a 585 amino acid protein, bestrophin-1, and belongs to a family of four bestrophins that contain a highly conserved tripeptide motif of arginine (R), phenylalanine (F), and proline (P)[4-8]. The be- strophin-1 protein is located in the basolateral membrane of RPE cells, has four transmem- brane domains in the N-terminal part of the protein[9], and forms tetrameric or pentameric Cl- channels that are activated by alterations in the intracellular Ca2+ concentration[7,10-15]. Studies in mice show drastically altered Cl- conductances in mice with mBest2 mutations[7]. The bestrophin-1 protein is expressed at high levels in the RPE[7]. The number of disease- causing mutations is high; approximately 100 mutations at over 80 different positions have been reported[17]. The majority of these are missense mutations that cluster near the trans- membrane portions and the NH2 terminus. The specific functional consequences of these mutations remain to be elucidated, but the current notion is that they alter Cl- conductance and/or the Ca2+ flux through voltage-dependent channels in the RPE[18]. As a consequence, the electrical current in the RPE is reduced, and epithelial transport is disrupted, ultimately resulting in accumulation of lipofuscin[10]. Previously, it was thought that congenital Best disease, adult-onset vitelliform macular dys- trophy, and autosomal dominant vitreoretinochoroidopathy (ADVIRC) were different dis- ease entities. The identification of mutations in the Best1 gene altered this notion, and re- vealed that most individuals with these phenotypes linked to the same gene[8]. Since then, it is clear that the expression and clinical consequences of a mutation in Best1 are highly vari- able[19,20]. Whether the onset and course of disease is determined by the type and location of the mutation, however, is still unresolved. In the current study, we investigated the genotype-phenotype correlation in VMD patients who were examined in three specialized centers in the Netherlands between 2003 and 2007. Here we report the distribution of Best1 mutations within this group, and the association be- tween onset and course of the macular dystrophy and visual acuity (VA) over time.

Patients and methods Study population Patients with a clinical and molecular diagnosis of VMD who visited three specialized oph- thalmogenetic clinics in the Netherlands between 2003 and 2007 were considered eligible for inclusion in the study. The clinical and molecular diagnosis included a VMD lesion in the

157 Chapter 7 retina (see below) and the presence of a disease-causing mutation in the Best1 gene. Exclu- sion criteria were: uncertainty with regard to the clinical diagnosis or the absence of a disease causing mutation. Twenty-nine probands and 24 affected relatives met these criteria and were subsequently invited to participate in the study. The study was approved by the Medical Ethical Committee of Erasmus University, and conformed to the tenets of the declaration of Helsinki. All subjects signed a written informed consent prior to inclusion.

Clinical parameters Age of onset, fundus appearance, best-corrected visual acuity (VA), and results of electro- physiological testing were retrieved from medical records. We evaluated all available param- eters at first presentation, and at all subsequent visits. In addition, subjects were invited for an additional clinical examination including fundus photography. The VMD lesions were evaluated on fundus photographs and classified based on a six stage grading scheme developed by Mohler and Fine[1]. In short, VMD stage 1 represents no changes or subtle RPE pigment changes; stage 2 is an egg yolk-like or vitelliform lesion; stage three a pseudohypopyon; stage 4 a scrambled egg appearance; stage 5 atrophy of the RPE; and stage 6 a fibrous scar with or without choroidal neovascularisation. We recorded the first occurrence of VA<0.5. This level is the cut-off point for visual disability according to US definition, and reflects the legal inability to drive a car[21]. We also recorded the first measurement of VA<0.3. This is the cut-off point for low vision according to WHO criteria, and is associated with increasing difficulties in reading without visual aids[22]. EOG recordings were generally carried out at initial diagnosis, and performed according to standard ISCEV criteria[23]. Arden ratio’s were registered and stratified in normal (≥ 1.85), mildly reduced (< 1.85 and ≥ 1.3), and severely reduced (<1.3).

Mutation analysis DNA was extracted from peripheral blood leukocytes by standard methods[24]. DNA was amplified using intronic primers described by Petrukhin et al[5]. For exon 4, we developed the following alternative primers: forward strand cgctcgcagcagaaagct, reverse strand ctc- cacccatcttccattc[17]. After heteroduplex formation, we screened samples for mutations using denaturing high pressure liquid chromatography (dHPLC, Transgenomic). Variations were analyzed by direct sequencing (ABI-310, Applied Biosystems). Sequence changes were an- notated according to Genbank accession NM_004183.1. The first A of the ATG was chosen as nucleotide +1, and the ATG as codon +1.

Statistical Analysis Correlation of VMD stage and VA between right and left eye was calculated with Kappa statistics. Cumulative risks of fundus abnormalities and visual loss were studied with Ka- plan-Meier product-limit survival analysis. For fundus abnormalities, first appearance of ab- normalities (stage 1 or greater vs. stage 0) and first appearance of atrophy or scar formation (stage 5 or 6 vs. stage 0-4) were analyzed. For visual function, onset of mild (VA<0.5 vs. VA≥0.5) and severe visual impairment (VA< 0.3 vs. VA≥0.3) were determined. These analy- ses were also performed stratified for the three most frequent mutations and stratified for age

158 Genotype-phenotype correlation in vitelliform macular dystrophy of onset in childhood (<10 years) or in adulthood (>18 years). Differences in cumulative risks between Best1 genotypes were compared using the Breslow test[25].

Table 1. Demographic and clinical characteristics of vitelliform macular dystrophy at first presentation and during follow up.

At presentation At presentation Variables (all patients) (pts with follow up) Follow up patients, n 53 40 40 pedigrees, n 29 23 23 Gender 22 males, n 29 22 females, n 25 18 18 Age at first presentation 34 (2-78) 35 (2-78) (range), yrs VMD stage n (%) better eye worse better eye worse better eye worse 0 7 (13) 6 (11) 5 (13) 4 (10) 4 (11) 3 (8) 1 10 (19) 5 (9) 6 (15) 4 (10) 1 (3) 0 (0) 2 9 (17) 9 (17) 8 (20) 7 (18) 3 (8) 3 (8) 3 7 (13) 8 (15) 4 (10) 5 (13) 1 (3) 0 (0) 4 7 (13) 8 (15) 6 (15) 7 (18) 5 (14) 3 (8) 5 10 (19) 12 (23) 8 (20) 9 (23) 17 (46) 14 (38) 6 3 (6) 5 (9) 3 (8) 4 (10) 6 (16) 14 (38) EOG recordings n (%) 41 (76) normal (≥1.85) 0 (0) reduction mild (1.3-1.85) 9 (22) reduction severe (<1.3) 32 (78) VA n (%) better eye worse better eye worse better eye worse ≥0.5 38 (72) 22 (42) 26 (65) 16 (40) 13 (33) 4 (10) 0.5-0.3 13 (25) 10 (19) 12 (30) 7 (18) 13 (33) 5 (13) 0.3-0.1 2 (4) 18 (35) 2 (5) 14 (35) 13 (33) 23 (59) <0.1 0 (0) 2 (4) 0 (0) 3 (8) 0 (0) 7 (18) VMD: vitelliform macular dystrophy, EOG: electro oculogram, VA: visual acuity.

Results Clinical characteristics at first presentation Demographic data of all 53 patients are presented in Table 1. The patients were genealogi- cally linked to 27 pedigrees; 16 individuals were sporadic patients and 37 had at least one relative with a VMD phenotype. The study population consisted of 29 males and 24 females. The age at the first visit to an ophthalmologist ranged from age 2 to 78 years, with a median of 33 years. Seventeen patients had an early onset (age 18 yrs). At first presentation, all stages were observed (Table 1); the correlation between the left and right eye was high (Kappa 0.76). The VA was below 0.5 in the better eye in 15 (29%) patients and in the worse eye in 30

159 Chapter 7

(59%) patients. VA at first presentation showed little correlation between the two eyes (VA left vs. right eye: Kappa 0.24). The majority of patients (78%) had a severely reduced EOG, with an Arden ratio < 1.3. The remainder of the patients had a mildly reduced Arden ratio (< 1.85).

Mutations in the Best1 gene We identified 14 different mutations in the 27 pedigrees (Table 2). Single missense mutations were most common (11/14). The most frequent mutation was the Tyr227Asn in seven mem- bers of six pedigrees (22% of probands), followed by Thr6Pro mutation in 18 members of five different pedigrees (19%), and Ala10Val in seven members of two pedigrees (7%). Two patients from one pedigree carried a compound heterozygous missense mutation (Ala195Val and Leu134Val). In addition, we identified an in-frame deletion and an in-frame insertion in two probands.

Table 2. Mutations identified in the 53 patients with vitelliform macular dystrophy in theBest1 gene.

mutation effect no. of patients no. of pedigrees ref [c.16A>C] [p.Thr6Pro] 18 5 5 [c.29C>T] [p.Ala10Val] 7 2 27 [c.47C>A] [p.Ser16Tyr] 1 1 28 [c.288G>C] [p.Gln96His] 5 1 27 [ c.583_584insTGG] [p.Lys194_Ala195insVal] 1 1 28 [c.584C>T] [c. 400C>G] [p.Ala195Val] [p.Leu134Val] 2 1 28 [ c.584C>T] [p.Ala195Val] 1 1 29 [c.653G>A] [p.Arg218His] 3 3 29 [c.679 T>A] [p.Tyr227Asn] 7 6 5 [c.877 C>A] [p.Gln293Lys] 3 2 27 [c.893T>C] [p.Phe298Ser] 1 1 30 [c.896G>C] [p.Gly299Ala] 1 1 31 [c.905A>C] [p.Asp302Ala] 2 1 32 [c.904_912delGATGATGAT] [p.Asp302_Asp304del] 1 1 28 total 53 27 The A of the ATG-translation initiation codon is described as nucleotide 1, according to NM_004183.2.

Figure 1 shows the location of the mutations in the protein[26]. Three mutations were located in the N-terminal intracellular part of the protein (Thr6Pro, Ala10Val and Ser16Tyr); seven mutations were located in the intracellular loop (Gln96His, Leu134Val, Lys194_Ala195ins- Val, Ala195Val, Arg218His, Tyr227Asn and Gly229Ala); and four mutations were located in the C-terminal intracellular part (Gln293Lys, Phe298Ser, Asp302Ala and Asp302_Asp- 304del). The Leu134Val mutation, the Lys194_Ala195insVal mutation and the Ala195Val

160 Genotype-phenotype correlation in vitelliform macular dystrophy mutation were located in the third and fourth putative transmembrane domain. No mutations were found in the extracellular loops.

Figure 1. Localization of mutations identified in our patient population. 50 68 253 269

extracellular

intracellular 28 90 231 291 96 227 * 293 218 298 299 16 130 201 302 3 10 * 302 1 134 6* 1 195 2 194

149 179 Model adapted from Milenkovic et al.[18]. * indicates the three most common mutations in our study group, used for further analysis, 1 indicates these two mutations occured together, 2 indicates an insertion, 3 indicates a dele- tion.

Clinical course and genotype-phenotype correlation Data on multiple visits were available in 40 patients with a mean follow up of 15.3 years. All but two patients showed progression in stage of disease. Fourteen patients progressed to VMD stage 5 (atrophy of the RPE; 38%); 14 other patients progressed to stage 6 (fibrous scar, 38%). Figure 2a shows the manifestation of VMD stage 5 or 6 as a function of age. Half of the patients who developed atrophy or a scar did so by age 54 years, 75% by 63 years. We assessed the cumulative risk of VMD stage 5 or 6 for the three most common muta- tions in our population, and found statistically significant differences (P=0.002) (Thr6Pro, Ala10Val and Tyr227Asn; Fig 2b). Of the patients with the Ala10Val mutation, 50% reached stage 5 or 6 at age 30 years compared to 50% of the patients with the Thr6Pro and Tyr227Asn mutations at age 45 and 58 respectively.

Visual decline was present in all but three patients. One of these patients showed no deterio- ration during a follow up of only one year, and two patients from a single pedigree remained stable over a 2 and 5 year follow up, respectively, despite subtle changes in the macula, a severely decreased EOG and the presence of a disease-causing mutation. Fifty percent of pa- tients had VA<0.5 at age 55; 75% at age 66 years. For VA<0.3, these ages were 66 years and 74 years, respectively (Figure 3a and 3c). Two patients (5%) progressed to legal blindness (VA ≤ 0.1) after age 60 years. The cumulative risk of visual decline stratified for Thr6Pro, Ala10Val, and Tyr227Asn is shown in Figure 3b and 3d. Patients with the Ala10Val mutation

161 Chapter 7 had a significantly faster visual decline than patients with either Thr6Pro or Tyr227Asn. Fifty percent of patients with Ala10Val deteriorated to VA <0.5 at age 29 years, while 50% patients with Thr6Pro and Tyr227Asn did so at ages 58 years and 46 years, respectively (P<0.03). For deterioration to VA<0.3, these ages were 34 years for Ala10Val, and 53 years for Tyr227Asn (P=0.001). For Thr6Pro the majority did not decline below VA 0.3; only 30% did so beyond age 44.

Figure 2. Age of occurence of fundus stage 5 or 6 in the patient group.

1.0 a) 1.0 b) 6 0.8 0.8

0.6 0.6

0.4 0.4 Cumulative Risk F5 0.2 0.2

0.0 0.0 0 20 40 60 80 0 20 40 60 80 Age (years)

Occurrence of fundus stage 5 or 6 (F56) (atrophy or a fibrous scar with or without choroidal neovascularisation) a) in all subjects, b) in the three most common mutations Thr6Pro (dashed), Ala10Val (solid) and Tyr227Asn (dash-dot). Progression is significantly faster in patients with theAla10Val mutation (p=0.002) Squares indicate censored case.

Figure 3. Age of occurence of moderately decreased and severely decreased visual acuity in the patient group. 1.0 a) 1.0 b)

0.8 0.8

0.6 0.6

0.4 0.4 Cumulative Risk VA<0.5 0.2 0.2

0.0 0.0 0 20 40 60 80 0 20 40 60 80

1.0 c) 1.0 d)

0.8 0.8

0.6 0.6

0.4 0.4 Cumulative Risk VA<0.3 0.2 0.2

0.0 0.0 0 20 40 60 80 0 20 40 60 80 Age (years) Age of occurrence of a) moderately decreased visual acuity (VA<0.5) in all patients; and b) stratified for the three most common mutations Thr6Pro (dashed), Ala10Val (solid) and Tyr227Asn (dash-dot); c) severely decreased visual acuity (VA<0.3) in all patients; and d) stratified for the three most common mutations. Visual decline is significantly faster in patients with the Ala10Val mutation (VA<0.5 p<0.03; VA<0.3 p=0.001 ). Squares indicate censored cases.

162 Genotype-phenotype correlation in vitelliform macular dystrophy

Case presentation of phenotypic variability Two pedigrees with the Thr6Pro mutation showed remarkable variability in the manifestation of the disease. The first pedigree included three relatives with the Thr6Pro mutation (Figure 4a). A female proband presented with visual loss in both eyes at age 27 years. Her right eye had VA 0.2 and a macular scar (VMD stage 6); her left eye had VA 0.6 and showed only subtle macular changes (stage 1). Her EOG showed an Arden ratio of 1.5 in both eyes. Strik- ingly, her 55 year old mother had no visual complaints but showed very subtle accumulation of lipofuscin near the temporal vascular arcades upon examination (VA 1.0 OU; no macular changes; Arden ratio 2.0) despite the presence of the mutation. The maternal uncle had been diagnosed with VMD at age 51 years (OD: VA 0.8, VMD stage 2; OS: VA 1.2, VMD stage 1; OU: Arden ratio 1.0). His right eye progressed to a severe VMD phenotype in 6 years (VA <0.01, stage 5), while his left eye remained stable.

Figure 4. Two pedigrees with the Thr6Pro mutation. a b

I:1 I:2 I:1 I:2 Thr6Pro

II:1 II:2 II:3 Thr6Pro Thr6Pro

II:1 Thr6Pro III:1 Thr6Pro

a) Presentation of the first pedigree. All three individuals (III:1, II:2, and II:3) carried the Thr6Pro mutation. At age 27, the proband (arrow; III:1) had bilateral visual complaints (visual acuity (VA) right eye (OD) 0.2, left eye (OS) 0.6; fundoscopically OD macular scar (stage 6), OS subtle changes (stage 1)) with an Arden ratio of 1.5 in both eyes. Her 55 year old mother (II:2) had no visual complaints (VA 1.0 both eyes (OU); no macular changes; Arden ratio 2.0). The maternal uncle (II:3) was diagnosed at age 51 and progressed to a severe bilateral phenotype (VA<0.01, stage 5) in 6 years. Fundoscopic pictures of individual III:1 at age 27 years, and of individual II:2 at age 55 years. b) Presentation of the second pedigree. Both individuals (II:1 and I:2) carried the Thr6Pro mutation. At age 31, the proband (arrow, II:1) had VA 0.8 and vitelliform macular dystrophy (VMD) stage 2 in the left eye; his right eye had VA 1.0 and no macular changes, Arden ratio’s OU were 1.0. See lower two black-and-white fundus photographs. At age 47 years (upper two fundus photographs), his left eye had progressed to VA 0.13 and VMD stage 6, while his right eye remained completely normal. The patients’ mother had no clinical signs of VMD apart from an Arden ratio of 1.5 at the age of 57 years (see fundus photographs). Squares indicate males, circles indicate females, filled symbols indicate affected individuals. A color version of this figure can be found in chapter 15.

163 Chapter 7

The second pedigree consisted of a male proband with a unilateral presentation and his moth- er, both carrying the Thr6Pro mutation (Figure 4b). At age 31 years, the proband’s left eye had VA 0.8 and VMD stage 2; his right eye had VA 1.0 and no macular changes, while the Arden ratio’s OU were 1.0. At age 47 years, his left eye had progressed to VA 0.13 and VMD stage 6, whereas his right eye remained completely normal. The patients’ mother had no clinical signs of VMD apart from an Arden ratio of 1.5 at the age of 57 years.

Statistical considerations Our study design may have led to ascertainment bias. Therefore we performed multiple anal- yses to exclude this effect in our data (data not shown). In summary, we found no differences between the age when an end stage phenotype was reached for patients with an age of onset in childhood, compared to an age of onset in adulthood. Moreover, our data do not show any differences in follow up for patients with relatively good visual acuity at the last visit com- pared to patients with poor VA. Finally, we show that the group of people with follow up and the group without follow up do not differ significantly.

Discussion We show that VMD, caused by mutations in the Best1 gene, is a disease with a variable age of onset and clinical course, ultimately leading to visual decline. Our survival analyses indicate that the variability may in part be explained by differences in genotype. The majority of pa- tients deteriorated to such extent that driving was prohibited at middle age, and that reading was generally compromized after age 70. However, development of legal blindness was rare, even at higher ages. The Ala10Val mutation showed a significantly faster decline causing severe visual impairment in the majority of patients by 44 years of age. Some methodological considerations should be addressed before we elaborate on our find- ings. Although we are not aware of larger studies describing the course of VMD, our study was still limited in sample size. This hampered precise estimates in our subgroup analyses. Nevertheless, we were able to detect significant differences between the genotypes over time. Another limitation was our retrospective study design. We were obliged to rely on clinical observations from various ophthalmologists and clinics without data acquisition according to standard protocols and time intervals. We do not think this distorted our results significantly, since our participants underwent regular standard ophthalmologic examinations, and our out- come parameters were derived from standard clinical care. Finally, as specified in the results section, our data was not subject to high levels of ascertainment bias. Thus far, descriptions of the clinical course of VMD were mainly based on reports of small families. In the current study, we calculated the mean rate of visual decline in a large series of VMD cases. On average, a VA<0.5, was reached by age 55 years. A VA<0.3 was reached by age 66. In our study, the functional endpoint was most often visual impairment; only two (5%) patients progressed to legal blindness. Despite the general visual decline observed in VMD patients, clinicians are well aware of the high variability of the phenotype. This is illustrated by our study, as left and right eyes of individual patients showed little correlation at any given time point, and even unilateral expression occurred. Moreover, severely affected probands who initially appeared to be spo-

164 Genotype-phenotype correlation in vitelliform macular dystrophy radic, revealed, upon family screening, relatives with subtle manifestations of the phenotype. Mutation analysis in these family members showed the presence of the same disease-causing Best1 mutation as the proband. Therefore we advocate genetic testing in first degree relatives of ‘sporadic’ probands even in the absence of signs or symptoms of disease. This may reveal unknown carriers, confirm the dominant inheritance pattern, and inevitably improve genetic counseling. Congenital Best disease and adult-onset vitelliform macular dystrophy were previously thought to be two distinct phenotypes. The former was considered to have an early age of onset and mildly to severely decreased EOG while the latter supposedly had an adult onset, a lack of distinct VMD stages and a mildly decreased to normal EOG[15,27]. Our study sug- gests that a continuous spectrum of the disease is more likely, since our patients showed high variability in age of onset, VMD stages and EOG values. We hypothesize that congenital Best disease and adult-onset vitelliform macular dystrophy caused by mutations in the Best1 gene are two extremes of a single etiology. Why is the clinical course for Ala10Val more severe than for Thr6Pro or Tyr227Asn? This is not immediately clear for a number of reasons: first, both Ala10Val and Thr6Pro lie in the first intracellular domain. Second, in the Ala10Val mutation both amino acids are hydro- phobic, whereas Thr6Pro changes the amino acid from hydrophilic to hydrophobic. Third, Tyr227Asn is located closer to one of the transmembrane domains that form the channel than Ala10Val, which would suggest that the former mutation has the greatest effect on the protein. Finally, the Tyr227Asn mutation is located in a more conserved region than the Ala- 10Val. Hence, we can not explain the faster progression for patients with an Ala10Val muta- tion, and the underlying biological mechanism still remains to be elucidated. Apart from the dominant negative effect, either single nucleotide polymorphisms or expres- sion variability in the wild type allele may influence the severity of the disease. This is not uncommon in ophthalmic disorders and has also been shown in autosomal dominant retinitis pigmentosa (RP11 and RP1)[28]. Two contradictory publications exist on the protein structure and the location of the trans- membrane domains and intra- or extracellular loops of the Best1 protein[26,29]. Both models describe six hydrophobic domains and an intracellular loop containing either the third[29] or the third and fourth[26] hydrophobic loop. In 2003, Tsunenari et al. described a model in which all but the third hydrophobic domain are transmembrane domains. More recently however, Milenkovic et al. (2007) showed that only the first two and the last two hydrophobic domains are transmembrane domains. In the current publication we based our conclusions of the new theory of Milenkovic et al[26]. Nevertheless, the actual folding of the protein into a three dimensional structure could be different from these models and may explain the differ- ences in the clinical course observed in this study. In conclusion, our study shows that the highly variable clinical course in the VMD phenotype may be explained, at least in part, by variation in genotype. However, virtually all patients show a gradual decline in VA without reaching legal blindness. It is intriguing that mutations do not necessarily lead to a phenotype. Insight into these mechanisms may provide clues for future clinical care.

165 Chapter 7

References 1. Mohler CW, Fine SL. Long-term evaluation of patients with Best’s vitelliform dystro- phy. Ophthalmology 1981;88:688-92. 2. Forsman K, Graff C, Nordstrom S, et al. The gene for Best’s macular dystrophy is lo- cated at 11q13 in a Swedish family. Clin Genet 1992;42:156-9. 3. Stone EM, Nichols BE, Streb LM, et al. Genetic linkage of vitelliform macular degen- eration (Best’s disease) to chromosome 11q13. Nat Genet 1992;1:246-50. 4. Marquardt A, Stohr H, Passmore LA, et al. Mutations in a novel gene, VMD2, encod- ing a protein of unknown properties cause juvenile-onset vitelliform macular dystrophy (Best’s disease). Hum Mol Genet 1998;7:1517-25. 5. Petrukhin K, Koisti MJ, Bakall B, et al. Identification of the gene responsible for Best macular dystrophy. Nat Genet 1998;19:241-7. 6. Stohr H, Milenkowic V, Weber BH. VMD2 and its role in Best’s disease and other reti- nopathies [in German]. Ophthalmologe 2005;102:116-21. 7. Hartzell HC, Qu Z, Yu K, et al. Molecular physiology of bestrophins: multifunc- tional membrane proteins linked to Best disease and other retinopathies. Physiol Rev 2008;88:639-72. 8. Boon CJ, Klevering BJ, Leroy BP, et al. The spectrum of ocular phenotypes caused by mutations in the BEST1 gene. Prog Retin Eye Res 2009;28:187-205. 9. Marmorstein AD, Marmorstein LY, Rayborn M, et al. Bestrophin, the product of the Best vitelliform macular dystrophy gene (VMD2), localizes to the basolateral plasma membrane of the retinal pigment epithelium. Proc Natl Acad Sci U S A 2000;97:12758- 63. 10. Sun H, Tsunenari T, Yau KW, Nathans J. The vitelliform macular dystrophy protein defines a new family of chloride channels. Proc Natl Acad Sci U S A 2002;99:4008-13. 11. Xiao Q, Prussia A, Yu K, et al. Regulation of bestrophin Cl channels by calcium: role of the C terminus. J Gen Physiol 2008;132:681-92. 12. Qu Z, Chien LT, Cui Y, Hartzell HC. The anion-selective pore of the bestrophins, a fam- ily of chloride channels associated with retinal degeneration. J Neurosci 2006;26:5411- 9. 13. Pusch M. Ca(2+)-activated chloride channels go molecular. J Gen Physiol 2004;123:323- 5. 14. Qu Z, Fischmeister R, Hartzell C. Mouse bestrophin-2 is a bona fide Cl(-) channel: identification of a residue important in anion binding and conduction. J Gen Physiol 2004;123:327-40. 15. Kramer F, White K, Pauleikhoff D, et al. Mutations in the VMD2 gene are associ- ated with juvenile-onset vitelliform macular dystrophy (Best disease) and adult vitel- liform macular dystrophy but not age-related macular degeneration. Eur J Hum Genet 2000;8:286-92. 16. Booij JC, van Soest S, Swagemakers SM, et al. Functional annotation of the human retinal pigment epithelium transcriptome. BMC Genomics [serial online] 2009;10:164. Available at: http://www.biomedcentral.com/1471-2164/10/164. Accessed November 8, 2009.

166 Genotype-phenotype correlation in vitelliform macular dystrophy

17. Retina International Mutation Databases [database online]. Updated September 18, 2009. Available at: http://www.retina-international.org/sci-news/mutation.htm. Ac- cessed November 8, 2009. 18. Milenkovic VM, Soria RB, Aldehni F, et al. Functional assembly and purinergic activa- tion of bestrophins. Pflugers Arch 2009;458:431-41. 19. Seddon JM, Sharma S, Chong S, et al. Phenotype and genotype correlations in two Best families. Ophthalmology 2003;110:1724-31. 20. Renner AB, Tillack H, Kraus H, et al. Late onset is common in Best macular dystrophy associated with VMD2 gene mutations. Ophthalmology 2005;112:586-92. 21. Leat SJ, Legge GE, Bullimore MA. What is low vision? A re-evaluation of definitions. Optom Vis Sci 1999;76:198-211. 22. World Health Organization. Prevention of Blindness and Visual Impairment: Change the definition of blindness. Available at: http://www.who.int/blindness/Change%20 the%20Definition%20of%20Blindness.pdf. Accessed November 8, 2009. 23. Brown M, Marmor M, Vaegan, et al. ISCEV Standard for Clinical Electro-oculography (EOG) 2006. Doc Ophthalmol 2006;113:205-12. 24. Miller SA, Dykes DD, Polesky HF. A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Res 1988;16:1215. 25. Crowley J, Breslow N. Statistical analysis of survival data. Annu Rev Public Health 1984;5:385-411. 26. Milenkovic VM, Rivera A, Horling F, Weber BH. Insertion and topology of nor- mal and mutant bestrophin-1 in the endoplasmic reticulum membrane. J Biol Chem 2007;282:1313-21. 27. Musarella MA. Molecular genetics of macular degeneration. Doc Ophthalmol 2001;102:165-77. 28. Daiger SP, Shankar SP, Schindler AB, et al. Genetic factors modifying clinical expres- sion of autosomal dominant RP. Adv Exp Med Biol 2006;572:3-8. 29. Tsunenari T, Sun H, Williams J, et al. Structure-function analysis of the bestrophin fam- ily of anion channels. J Biol Chem 2003;278:41114-25. 30. Kramer F, Mohr N, Kellner U, et al. Ten novel mutations in VMD2 associated with Best macular dystrophy (BMD). Hum Mutat 2003;22:418. 31. Boon CJ, Jeroen KB, Keunen JE, et al. Fundus autofluorescence imaging of retinal dystrophies. Vision Res 2008;48:2569-77. 32. Chung MM, Oh KT, Streb LM, et al. Visual outcome following subretinal hemorrhage in Best disease. Retina 2001;21:575-80.

ACKNOWLEDGEMENTS We thank all patients for their participation in this study.

167

Simultaneous mutation detection in 90 retinal disease genes in mul- tiple patients using a custom de- signed 300 kb retinal resequencing chip. 8

Judith C Booij, Arne Bakker, Jamilia Kulumbetova, Youssef Moutaoukil, Bert Smeets, Joke Verheij, Hester Y Kroes, Caroline CW Klaver, Mary van Schooneveld, Arthur AB Bergen, Ralph J Florijn

Ophthalmology (2010), in press Chapter 8

Abstract Purpose: To develop a high-throughput, cost-effective diagnostic strategy for the identifica- tion of known and new mutations in 90 retinal disease genes.

Design: Evidence based study.

Participants: Sixty patients with a variety of retinal disorders, including Leber congenital amaurosis, ocular albinism, pseudoxanthoma elasticum, retinitis pigmentosa, Stargardt dis- ease.

Methods: We designed a custom 300 kb resequencing chip. Polymerase chain reaction (PCR) amplification, DNA fragmentation and chip hybridization were performed according to Affymetrix recommendations. Hybridization signals were analyzed using Sequence pilot module seq-C mutation detection software (2009). This resequencing approach was validated by Sanger sequence technology.

Main outcome measures: (disease causing) sequence changes.

Results: We developed a retinal resequencing chip that covers all exons of 90 retinal disease genes. We developed and tested multiplex primer sets for 1,445 amplicons representing the genes included on the chip. We validated our approach by screening 87 exons from 25 retinal disease genes containing 87 known sequence changes previously identified in our patient group using Sanger sequencing. Call rates for successfully hybridized amplicons was 98% to 100%. Of the known single nucleotide changes, 99% could be detected on the chip. As expected, deletions could not be detected reliably.

Conclusions: We designed a custom resequencing chip that can detect known and new se- quence changes in 90 retinal disease genes, using a new high throughput strategy with a high sensitivity and specificity for one tenth of the cost of conventional direct sequencing. The developed amplification strategy allows for the pooling of multiple patients with non- overlapping phenotypes, enabling many patients to be analyzed simultaneously in a fast and cost-effective manner.

170 The resequencing chip

Introduction

Numerous potential disease genes interact specifically in molecular mechanisms that affect photoreceptor and RPE function, making these two cell types the main cellular contributors to genetically determined retinal disease. Among the retinal diseases there are phenotypes affecting only the eye, like retinitis pig- mentosa (RP), Leber congenital amaurosis (LCA), cone-rod dystrophy, and Best vitelliform macular dystrophy, and there are syndromic phenotypes like Bardett-Biedl syndrome and Usher syndrome[1]. These phenotypes are caused by mutations in individual genes[1]. All inheritance patterns can be observed among the different retinal phenotypes, from autosomal recessive (LCA), autosomal dominant (optic atrophy), and X-linked (congenital stationary night blindness) to multifactorial diseases like age-related macular degeneration (AMD)[1]. Multiple inheritance types can even be observed within a single phenotype, for instance RP can be inherited in an autosomal dominant, autosomal recessive or X-linked manner.

The identification of a causative mutation is important for confirmation of the clinical di- agnosis and inheritance pattern, for predictions of clinical course of the disease, for genetic counseling and family planning and for future gene-targeted treatment. However, the com- plex nature of retinal phenotypes and genetic heterogeneity prohibits efficient identification of disease-causing mutations. Many phenotypes are not well delineated and may show over- lapping features with other phenotypes[2,3]. In these cases, it is difficult to predict what the causative gene is. Furthermore, approximately 150 retinal disease genes have now been identified for as many as 30 different retinal phenotypes (see Table 1)[1]. For a disease like arRP, at least 25 possible causative genes exist[1] and have to be screened. On average, more than five different genes have to be considered per phenotype.

Table 1. Genes present on the custom resequencing chip, sorted by the disease they represent[1,17].

Retinal phenotype Genes on the array Aniridia PAX6 Bardet-Biedl syndrome, AR ARL6, BBS1, BBS2, BBS4, BBS5, BBS7, BBS9, BBS10, MKKS,TTC8, TRIM32 Cone or cone-rod dystrophy, AD AIPL1, CRX, GUCA1A, GUCY2D, RIMS1, SEMA4A, UNC119 Cone or cone-rod dystrophy, AR ABCA4,CNGB3, RDH5 Cone or cone-rod dystrophy, XL RPGR Congenital stationary night blindness, AD GNAT1, PDE6B, RHO Congenital stationary night blindness, AR CABP4, GRM6, RDH5, SAG Congenital stationary night blindness, XL CACNA1F, NYX Deafness alone or syndromic, AR CDH23, MYO7A, PCDH15, USH1C Open-angle glaucoma, juvenile-onset MYOC Open-angle glaucoma OPTN

171 Chapter 8

Leber congenital amaurosis, AD CRX, IMPDH1 Leber congenital amaurosis, AR AIPL1, CEP290, CRB1, CRX, GUCY2D, LRAT, RDH12, RPE65, RPGRIP1, TULP1 Leber hereditary optic neuropathy MTND1*, MTND4*, MTND6* Macular degeneration, AD BEST1, EFEMP1, ELOVL4, PRPH2, TIMP3 Macular degeneration, AR ABCA4, CFH Macular degeneration, XL RPGR Macular degeneration, age related ABCA4, C2*, CFB*, CFH*, LOC387715* Oculocutaneous albinism, type 2 OCA2 Optic atrophy, AD OPA1 Retinitis pigmentosa, AD CA4, CRX, IMPDH1, NR2E3, NRL, PRPF3, PRPF8, PRPF31, RHO, ROM1, RP1, RP9, SEMA4A, Retinitis pigmentosa, AR ABCA4, CERKL, CNGA1, CNGB1, CRB1, LRAT, MERTK, NR2E3, NRL, PDE6A, PDE6B, RGR, RHO, RLBP1, RP1, RPE65, SAG, TULP1, USH2A Retinitis pigmentosa, XL RP2, RPGR Syndromic/systemic diseases with retinopa- ABCC6 thy, dom, pseudoxanthoma elasticum Syndromic/systemic diseases with retinopa- ABCC6, ALMS1, CEP290, OPA3, TTPA thy, AR Usher syndrome, AR CDH23, GPR98, MYO7A, PCDH15, USH1C, USH1G, USH2A Waardenburg syndrome MITF, PAX3, Other retinopathy, AD BEST1, CRB1,OPN1SW, Other retinopathy, AR BEST1, CNGA3, CNGB3, NR2E3, RLBP1 Other retinopathy, XL NDP, OPN1LW, RS1

AR autosomal recessive, AD autosomal dominant, XL X-linked. * one or more SNPs associated with the disease

Currently, a number of mutation screening strategies are available, each with their own ad- vantages and limitations. Prescreening tools like single strand chain polymophism (SSCP) and denaturing high pressure liquid chromatography (dHPLC), two techniques capable of detecting but not identifying changes in DNA fragments relative to a reference DNA sample, are relatively inexpensive but cannot identify a mutation without the use of additional tools like direct sequencing. Direct sequencing methods can identify mutations, but they can only screen a single amplicon (i.e. one exon plus flanking intronic sequence) at a time. This makes the technique time-consuming and expensive when used for multiple genes. More recent- ly developed tools that screen multiple genes simultaneously, like genotyping microarrays (Asper)[4], can only detect a fixed number of known mutations. Second generation sequenc- ing (ABI[5], Illumina[6], Roche[7]), whereby 20 Gb can be sequenced simultaneously may provide a solution in the future, but currently this technique is very expensive, and needs to be validated yet for diagnostic purposes.

172 The resequencing chip

In summary, in ophthalmogenetics there is great need for a tool that can identify both known and new mutations in multiple genes in a fast and cost-effective manner, since any new pa- tient may require screening of multiple genes related to more than one phenotype before a certain molecular diagnosis can be established.

In this study, we developed and validated a new custom-designed retinal resequencing chip. This high throughput sequencing tool can detect both known and new sequence changes in 265,000 nucleotides from 90 different retinal disease genes. Our unique amplification and pooling strategy allows for cost-effective testing of multiple patients in a single experiment.

Methods For this study we designed a custom 300kb Affymetrix resequencing chip[8] capable of se- quencing both forward and reverse strands of 90 retinal disease genes. For validation of the chip we included DNA from patients with retinal diseases containing 87 known sequence changes in 25 of the genes on the chip. The study was performed in agreement with the dec- laration of Helsinki. Approval of the medical ethics commitee of the AMC Amsterdam for ophthalmic examination and DNA screening of patients was obtained. Patients with a clinical diagnosis of a retinal disease, whose blood was sent to the Netherlands Institute for Neuro- science for molecular diagnostic confirmation of the diagnosis, in whom a disease-causing mutation was identified, were included in this study.

Experimental procedure DNA was isolated from peripheral blood using standard methods [9]. Primers were devel- oped using the ELXRdb website[10]. When necessary, additional primers were developed using the Primer3 software (http://frodo.wi.mit.edu/primer3/). Multiplex amplification of all amplicons in sets of four to five amplicons was performed with a Qiagen multiplex PCR kit (Qiagen, Venlo, The Netherlands) with 10 ng of input DNA per reaction in 10 µl reactions. A positive hybridization control was included in the whole procedure by way of a plasmid vector containing synthetic (non-naturally occurring) DNA. Amplification consisted of 15 seconds at 94ºC, followed by 45 cycles of 94ºC for 30 seconds, 58ºC for 90 seconds, and 72ºC 90 seconds, the protocol was concluded by 72ºC for 10 minutes. All amplified DNA samples were checked on a 2% MetaPhor agarose gel (Lonza Benelux BV, Breda, The Neth- erlands). Samples were quantified with PicoGreen® quantification (Invitrogen, Breda, The Netherlands) and an average of 441 ng (in 1.3 µl) of each multiplex reaction was pooled. Pooled samples were purified using a DNA Amplification Clean-Up Kit (Clontech, Saint- Germain-en-Laye, France). For each pooled sample, a test fragmentation was performed in order to establish the activity of the fragmentation reagent in combination with the sample. Based on the results of this test, samples were fragmented and checked for fragment lengths on a 3% agarose gel. All fragmented samples were labeled and stored at -20°C until they were shipped to ServiceXS[11], where they were hybridized to the chip, washed, stained and scanned according to Affymetrix protocols[8].

173 Chapter 8

Data analysis Scanned images were extracted from the chips with the GCOS software (Affymetrix UK Ltd, United Kingdom). Data was subsequently imported into the Sequence module seq-C (JSI medical systems GmbH, Kippenheim, Germany) that automatically performs statistical cor- rection and normalization of the data (Figure 1). In short: for each of the 265,000 bases, both forward and reverse strands, the signal strengths were recorded for the A, C, T and G nucleo- tide and compared to an average signal, including standard deviation (sd), obtained from a larger set of previously hybridized chips (normalization set). For each strand, nucleotides were called (A, C, T, G, homozygous or heterozygous) based on their signal strength com- pared to the average and sd of the normalization set. We imported the reference sequences of all genes into the Sequence pilot module seq-C software using Ensembl (www.ensembl.org/ index.html) along with nearly all known pathogenic and non-pathogenic sequence changes. Nucleotide calls (manual or by the Sequence pilot module seq-C software) were compared to the imported reference sequence.

Figure 1. Analytical workflow.

Mutation Statistical analysis

Call rate Number of base 99% pairs screened

No hybridization because amplicon was not amplied

All hybridization results were imported into and analyzed in the JSI software. For each amplicon the program reports call rates and the number of changes identified. Both forward and reverse strands are visualized and statis- tics are given for the signal (A, C, T, or G) for each base position compared to the normalization set of previously analyzed chips. A color version of this figure can be found in chapter 15.

174 The resequencing chip

Results Custom resequencing chip design We designed a custom 300kb Affymetrix resequencing chip, allowing for the direct sequenc- ing of 265,000 double-stranded bases of retinal disease genes. For 90 retinal disease genes, the array covers all coding exons and 12 bases of flanking intronic sequence on either side of the exons. For seven genes several SNPs are included (Table 1). Array specifications required that 12 additional bases are added on the 5’ and 3’ side of each amplicon, these 24 bases are not included in the sequence analysis. Sequence-specific oligonucleotide probes were de- veloped, containing eight unique 25-mer probes per base position, according to Affymetrix resequencing array tiling strategies[8]. Four variants of each probe were included, with each possible nucleotide at the 13th position, A, G, C, or T, allowing for the detection of both known and novel sequence changes[8].

Primer design and multiplex amplification of target sequences We developed a flexible, reliable, efficient and high-throughput method for target-specific amplification (Figure 2). For the 1,445 amplicons and additional SNPs, we designed ampli- con-specific primers, each containing a universal common poly-nucleotide tail. We tested and amplified all amplicons using a multiplex PCR strategy with an average of 4 amplicons per multiplex reaction (see Table 2).Thus, in total 352 multiplex PCR reactions covered all sequences available on the chip. Common tail-specific amplifications were tested and used to avoid otherwise unavoidable inter-sample differences in multiplex target am- plification[12]. The amplification of single exons compared to for example long range PCR fragments [13] resulted in increased flexibility; i.e. exons were amplified per gene or per individual, or both. All primer-sequences are available upon request.

Figure 2. Technical work flow. multiplex genomic DNA select amplicons amplification and pooling fragmentation

exon 1 exon 1 exon 2 exon 2 exon 3 exon 3

Image extraction hybridization labeling

chip * * * * * * *

Starting with genomic DNA, amplicons are selected for multiplex amplification. Subsequently, all samples are pooled, fragmented and labeled before hybridization to the chip. Images are extracted from the chip using the GCOS software (Affymetrix UK Ltd, United Kingdom). Data was subsequently imported into the JSI SeqC soft- ware (JSI medical systems GmbH, Kippenheim, Germany) and analyzed as shown in Figure 1.

175 Chapter 8

Table 2. Number of amplicons and multiplexes per gene on the resequencing chip.

no of multiplex gene NM_no. PCR reactions No. of amplicons ABCA4 NM_000350 11 49 ABCC6 AF076622 6 31 AIPL1 NM_014336 2 7 ALMS1 NM_015120 10 55 ARL6 NM_032146 6 7 BBS1 NM_024649 3 17 BBS2 NM_031885 4 17 BBS4 NM_033028 4 16 BBS5 NM_152384 2 11 BBS7 NM_018190 5 19 BBS9 NM_198428 6 22 BBS10 NM_024685 2 6 BEST1 NM_004183 3 11 C2* NM_000063 1 2 CA4 NM_000717 2 7 CABP4 NM_145200 2 7 CACNA1F NM_005183 10 47 CDH23 NM_022124 12 70 CEP290 NM_025114 14 55 CERKL NM_201548 7 14 CFB* NM_001710 1 2 CFH* NM_000186 1 3 CNGA1 NM_000087 2 10 CNGA3 NM_001298 2 10 CNGB1 NM_001297 7 32 CNGB3 NM_019098 4 19 CRB1 NM_012076 4 17 CRX NM_000554 1 5 EFEMP1 NM_004105 2 11 ELOVL4 NM_022726 2 6 GNAT1 NM_000172 2 8 GRM6 NM_000843 3 12 GUCA1A NM_000409 1 4 GUCY2D NM_000180 4 18

176 The resequencing chip

no of multiplex gene NM_no. PCR reactions No. of amplicons IMPDH1 NM_000883 3 14 LOC387715* NM_001099667 1 1 LRAT NM_004744 2 4 GPR98 NM_032119 18 102 MERTK NM_006343 6 21 MITF NM_000248 2 11 MKKS NM_170784 5 8 MTND1 YP_003024026 1 1 MTND4 YP_003024035 1 1 MTND6 YP_003024037 1 1 MYO7A NM_000260 8 47 MYOC NM_000261 2 6 NDP NM_000266 1 2 NR2E3 NM_014249 4 8 NRL NM_006177 2 8 NYX NM_022567 2 6 OCA2 NM_000275 5 23 OPA1 NM_130837 6 31 OPA3 NM_001017989 1 3 OPN1LW NM_020061 2 8 OPN1SW NM_001708 2 6 OPTN NM_021980 3 14 PAX3 NM_181458 2 10 PAX6 NM_001604 2 11 PCDH15 NM_033056 9 36 PDE6A NM_000440 5 23 PDE6B NM_000283 4 24 PRPF3 NM_004698 3 14 PRPF8 NM_006445 9 42 PRPF31 NM_015629 3 11 RDH12 NM_152443 4 7 RDH5 NM_002905 1 4 PRPH2 NM_000322 1 3 RGR NM_002921 3 7 RHO NM_000539 1 5

177 Chapter 8

no of multiplex gene NM_no. PCR reactions No. of amplicons RIMS1 NM_014989 6 31 RLBP1 NM_000326 1 6 ROM1 NM_000327 1 4 RP1 NM_006269 7 32 RP2 NM_006915 2 6 RP9 NM_203288 1 5 RPE65 NM_000329 3 14 RPGR NM_000328 4 19 RPGRIP1 NM_020366 5 25 RS1 NM_000330 2 6 SAG NM_000541 3 14 SEMA4A NM_022367 3 15 TIMP3 NM_000362 2 5 TRIM32 NM_012210 6 7 TTC8 NM_144596 4 15 TTPA NM_000370 1 5 TULP1 NM_003322 3 12 UNC119 NM_005148 1 4 USH1C NM_005709 5 21 USH1G NM_173477 2 6 USH2A NM_206933 17 77 Note that due to their size a number of amplicons were amplified together in a single amplicon and a number of exons were split into multiple amplicons. * indicates individual SNPs associated with age related macular degen- eration

Patient and sample selection In order to validate our mutation detection strategy, we selected 60 patients with an inherited retinal disease with one or more mutations in one of the 90 retinal disease genes present on the chip. Among the retinal diseases were glaucoma, Leber congenital amaurosis, optic atrophy, ocular albinism, pseudoxanthoma elasticum, dominant and recessive retinitis pig- mentosa, retinoschisis and Stargardt disease. The presence of mutations in the DNA of these patients was initially proven by direct sequencing of both forward and reverse strands in the relevant disease genes, while the presence of additional apparently non-pathogenic SNPs was detected by direct sequencing of a single strand. As a result, we identified 87 known patho- genic and polymorphic sequence changes in our study population.

178 The resequencing chip

For validation of the chip we devised a patient selection strategy enabling us to test all 87 sequence changes on a single chip (see Figure 3). For each patient we amplified only those exons containing a known sequence change. Following multiplex amplification, all ampli- cons were pooled and further processed together. This enabled us to analyze all 87 known sequence changes in a single chip experiment (see Figure 3). Pooled, fragmented samples were labeled and hybridized to the chip, washed, stained and scanned according to Affyme- trix protocols (see Figure 2)[11]. All experiments were performed in duplicate, in order to ascertain the reproducibility of the experiment.

Figure 3. Patient selection strategy. multiplex amplification pooling hybridization final results

gene A

gene B patient A gene A gene C phenotype gene B arRP gene D gene C gene D patient A 2 mutations in ARRP gene A gene ... gene E chip gene F patient B 1 mutation in ADRP gene F gene E gene ... patient B patient ... mutation in ... gene ... phenotype gene F gene ... adRP gene ... gene ... gene ... patient ... gene .... phenotype ... gene ... Patients with non-overlapping phenotypes are selected, the genes corresponding to their phenotype are amplified, then all amplified amplicons are pooled and samples are hybridized to a single chip. Mutations identified can be linked to the patients based on the gene in question.

Validation of the chip Thirty-two of the 1,445 amplicons (2%) failed to amplify in the multiplex procedure, there- fore hybridization was not successful and these exons were excluded from further analysis. Twelve of these 32 amplicons were GC-rich first exons. Average call rates for successfully hybridized amplicons was 98-100% for both chips (Affymetrix specifications >90%). The difference in call rate for all amplicons between the two duplicate chips ranged from 0% to 2% (Affymetrix specifications for reproducibility >99.9%). The 60 patients included in this study carried 87 sequence changes in 25 of the 90 retinal disease genes present on the array (Table 3). These 87 sequence changes included 84 single nucleotide changes, eight of which were intronic, and three small deletions (up to four bases). Five of the sequence changes were homozygous; the remainder was heterozygous. Sixty- one sequence changes were missense mutations, twelve were nonsense mutations, and seven were silent sequence changes.

179 Chapter 8

Of the 84 point mutations 83 (99%) were identified in both chips, and one was not detected in either chip (RGR c.454C>A p.H152N) (1%). Seventy-six point mutations (90%) were correctly called by the Sequence pilot module seq-C software (version and learning curve august 2009) using the default settings (Table 3). One (1%) sequence change was called by the Sequence pilot module seq-C software in one chip and detected manually in the second chip. Six additional mutations (7%) were detected manually on both chips. As expected, none of the three deletions could be detected by the Sequence pilot module seq-C software or manually.

Table 3. Positive control samples, patients, mutations and genes.

pheno- rs no./ Present Gene patient type sequence change reference on chip Nucleotide Protein 1 2 ABCA4 P28529/1 STGD c.1268A>G p.H423R rs.3112831 1 1 ABCA4 P28529/1 STGD c.1269C>T p.H423R rs.4147831 1 1 ABCA4 P28547/1 STGD c.6122G>A p.G2041D 18 1 1 ABCA4 P28547/1 STGD c.5682G>C p.L1894L rs1801574 1 1 ABCA4 P28547/1 STGD c.5814A>G p.L1938L rs4147857 1 1 ABCA4 P28549/1 STGD c.872C>T p.P291L current study 1 1 ABCA4 P28549/1 STGD c.4297G>A p.V1433I rs56357060 1 1 ABCA4 P28563/1 STGD c.1150A>G p.K384E current study 1 1 ABCA4 P28563/1 STGD c.319C>T p.Arg107X current study m m ABCA4 P28587/1 STGD c.2828G>A p.R943Q rs1801581 1 1 ABCA4 P28592/1 STGD c.635G>A p.R212H rs6657239 1 1 ABCA4 P28592/1 STGD c.6445C>T p.R2149X rs61750654 1 1 ABCA4 P28592/1 STGD c.6285T>C p.D2095D rs1801555 1 1 ABCC6 P26095/7 PXE c.3188T>G p.L1063R 19 1 1 ABCC6 P26095/7 PXE c.1077A>G p.S359S 20 1 1 ABCC6 P26095/7 PXE c.1132C>T p.Q378X 21 1 1 ABCC6 P26095/7 PXE c.1141T>C p.L381L 21 1 1 ABCC6 P26095/7 PXE c.1338+7c>G hom Intron rs9940089 1 1 ABCC6 P28586/1 PXE c.527T>C p.L176P current study 1 1 ABCC6 P28590/1 PXE c.373G>A p.E125K rs3853814 m m ABCC6 P28590/1 PXE c.473C>T p.A158V rs2606921 1 1 ABCC6 P28590/1 PXE c.1841T>C p.V614A rs12931472 1 1 ABCC6 P28590/1 PXE c.3803G>A p.R1268Q rs2238472 1 1 ABCC6 P28596/1 PXE c.3421C>T hom p.R1141X 22 1 1 ABCC6 P28599/1 PXE c.2071-1G>A Intron current study 1 1

180 The resequencing chip

pheno- rs no./ Present Gene patient type sequence change reference on chip Nucleotide Protein 1 2 ABCC6 P28599/1 PXE c.232G>A p.A78T rs2856597 1 m ABCC6 P28618/1 PXE c.1552C>T p.R518X 23 1 1 BBS5 P28651/1 BB c.551A>G p.N184S 24 1 1 CEP290 P25415/1 LCA c. 4723A>T p.K1575X 25 1 1 CEP290 P26523/1 LCA c.4961del AA p.Q1654Qfs1658X 25 0 0 CEP290 P26523/1 LCA c.2991+1655A>G cryptic splice site 26 1 1 CNGA1 P28218/1 arRP c.947C>T p.S316F 27 1 1 CRB1 26425/1 arRP c.3122T>C p.M1041T rs62635656 1 1 CRB1 P27949-1 arRP c.614T>C p.I205T 28 1 1 CRB1 P28155/1 arRP c.3992G>A p.R1331H rs62636285 1 1 CRB1 P28454/1 arRP c.2234C>T p.T745M rs28939720 1 1 CRB1 P28538/1 arRP c.2842+5G>A Intron current study 1 1 CRB1 P28538/1 arRP c.2843G>A p.C948Y 29 1 1 ELOVL4 P27743/1 STGD c.800T>C p.I267T 30 1 1 NR2E3 P28068/1 ADRP c.932G>A p.R311Q rs.28937873 1 1 GUCY2D P25556/1 LCA c.3283delC p.P1069RfsX37 current study 0 0 GUCY2D P8432/1 LCA c.2176C>T p.P701S 31 1 1 BEST1 P25632/2 VMD c.16A>C p.T6P 32 1 1 BEST1 P25632/2 VMD c.1608T>C p.T536T 32 1 1 BEST1 P28560/1 VMD c.653G>A p.R218H 33 1 1 NRL P27903/1 ADRP c.152C>T p.P51L 34 1 1 OCA2 P28559/1 XOA c.1465A>G p.N489D 35 1 1 OCA2 P28559/1 XOA c.1327G>A p.V443I 36 1 1 OPA1 P25750/1 DOA c.2109T>C p.A703A 37 1 1 OPA1 P28510/1 DOA c.43C>A p.Q15K current study 1 1 OPA1 P28518/1 DOA c.623A>G p.K208R current study 1 1 OPA1 P28518/19 DOA c.148A>G p.T50A current study 1 1 OPA1 P28558/1 DOA c.1035+4T>C hom Intron 37 1 1 OPA1 P28558/1 DOA c.983_984+3delAGGTA p.K328SfsX4 38 0 0 OPTN P28493/1 glaucoma c.964A>G hom p.K322E rs523747 1 1 OPTN P28493/1 glaucoma c.553-5C>T Intron rs2244380 1 1 PDE6A P21130/1 arRP c.2549G>T p.G850V 39 m m PDE6A P28179/1 arRP c.304C>A p.R102S 39 1 1

181 Chapter 8

pheno- rs no./ Present Gene patient type sequence change reference on chip Nucleotide Protein 1 2 PDE6A P28509/1 arRP c.647A>G p.N216S 39 1 1 PDE6B 25806/1 arRP c.892C>T p.Q298X 40 m m PDE6B P28093/1 arRP c.655T>C p.Y219H 40 1 1 PRPF3 P2101/5 ADRP c.1477C>T p.P493S 41 1 1 RGR 25378/2 LCA c.454C>A p.H152N 42 0 0 RGR P28195/1 LCA c.756+5A>G Intron 43 1 1 RP1 P24701/1 ADRP c.1118C>T p.T373I 44 1 1 RP1 P28516/1 ADRP c.2953A>T p.N985Y rs2293869 1 1 RPE65 P28089/1 arRP c.715T>G hom p.Y239D rs61752896 1 1 RPGRIP1 P25474/1 LCA c.3341A>G p.D1114G 45 1 1 RS1 P21940/1 XLRS c.625C>T p.R209C 46 1 1 RS1 P28553/1 XLRS c.214G>A p.E72K 47 1 1 TIGR P28493/1 glaucoma c.227G>A p.R76K rs2234936 1 1 TYR P28456/1 XOA c.547G>A p.V183M current study m m TYR P28461/1 XOA c.1118C>A p.T373K 48 1 1 USH2A P25857/1 arRP c.1036A>C p.N346H 49 1 1 USH2A P27930/1 arRP c.5975A>G p.Y1992C rs41303287 1 1 USH2A P28235/1 arRP c.1434G>C p.E478D 50 1 1 USH2A P28311/1 arRP c.13274C>T p.T4425M 51 m m USH2A P28337/1 arRP c.1256G>T p.C419F 52 1 1 USH2A P28337/1 arRP c.4106C>T p.S1369L 53 1 1 USH2A P28380/1 arRP c.2276G>T p.C759F 49 1 1 USH2A P28383/1 arRP p.2137G>C p.G713R 49 1 1 USH2A P28383/1 arRP c.1663C>G p.L555V 54 1 1 USH2A P28500/1 arRP c.4560C>T p.I1520I 55 1 1 USH2A P28509/1 arRP c.6587G>C p.S2196T 56 1 1 USH2A P28509/1 arRP c.15377T>C p.I5126T 56 1 1 USH2A P28542/1 arRP c.6240G>T p.K2080N 56 1 1 USH2A P8432/1 arRP c.11864G>A p.W3955X 57 1 1 Sequence changes known to be present in our 77 subjects. Unless otherwise indicated all changes are present in heterozygous form. hom: homozygous sequence change. 1 called by the JSI software, m called manually, 0 not called. STGD Stargardt disease, PXE pseudoxanthoma elasticum, BB Bardet Biedl syndrome, LCA Leber con- genital amaurosis, arRP autosomal recessive retinitis pigmentosa, ADRP autosomal dominant retinitis pigmentosa, VMD vitelliform macular dystrophy, XOA X-linked optic atrophy, DOA dominant optic atrophy, XLRS X-linked retinoschisis.

182 The resequencing chip

Additional sequence changes In addition to the 87 known sequence changes, the Sequence seq-C software called 360 ad- ditional sequence changes present on both chips spread over 38 genes (on average 9 new sequence changes per gene, one per four amplicons). We selected 79 amplicons to verify the presence of new sequence changes detected by the chip-technology. Direct sequencing of these amplicons confirmed the presence of new sequence variants in all cases. Two additional new sequence variants were only called on a single chip (not in duplicate). Their presence in the DNA was confirmed by direct sequencing. Finally, a number of ad- ditional new sequence changes (8 on the first chip and 7 on the duplicate) were identified in only one strand on one chip. Direct sequencing of the relevant amplicons revealed that these changes were not present in the DNA. Therefore, analysis based on only a single strand may lead to false positive results.

Chip costs and time required The costs of the resequencing chip including design and purchase of the chip, primers, am- plification of the target sequences, labeling reagents, hybridization and validation amount to €1,500 per chip. This is five to ten times cheaper than the current cost of conventional direct sequencing, depending on the number of patients/amplicons used on a single chip (personal communication Ralph Florijn)(see table 4). Total time required for the processing and analysis of a single sample will be between four and six weeks. However, many patient samples and several chips can be processed in parallel.

Table 4 Overview of costs of the retinal resequencing chip.

Category Costs per chip (euro’s) Chip design 500 Physical chip 340 Amplification (including primer design) 200 Labeling 50 Hybridization 250 Analysis 10/sequence change Compared to the total costs of conventional sequencing, €10 per amplicon for 1,445 amplicons, the resequencing chip is up to ten times cheaper.

Discussion In this study we present a high-throughput diagnostic tool for the detection of known and new mutations in patients with inherited retinal diseases. We describe the design, validation and use of a new 300kb retinal resequencing chip that allows for the direct sequencing of all coding exons and flanking introns of 90 retinal disease genes. We devised a unique pooling strategy that allows for the screening of multiple patients and/or genes per chip, which will reduce the costs per screened individual considerably. We were able to identify 99% of the

183 Chapter 8 single nucleotide changes on the chip. All 86 sequence changes identified in the first chip, were confirmed in the duplicate experiment.

Custom resequencing chip design Genes on the chip The chip can screen 90 retinal disease genes and individual SNPs in seven additional genes implicated in 30 retinal phenotypes. All coding exons and splice site regions of the genes are screened on both forward and reverse strand[1]. For each patient, multiple genes, related to a single phenotype, can be screened simultaneously, which is essential given the extensive clinical and genetic heterogeneity of most ophthalmogenetic disorders.

Multiplex PCR By using a multiplex PCR design, we were able to combine on average four amplicons of different sizes in a single PCR reaction, thereby reducing the number of reactions from 1,445 to 352, ensuring a significant time and cost reduction.

Pooling strategy For validation of our chip, we devised a pooling strategy that allowed us to screen 87 se- quence changes from many different genes in a single chip experiment. For each patient we amplified only the exons that contained a known sequence change. In contrast to applying a single patient’s DNA to each chip, our strategy allowed for many sequence changes to be tested in a single experiment (Figure 3). For future diagnostic purposes we will use a slight modification of this pooling strategy. We will screen a patient’s DNA for all exons of all the relevant genes present on the chip that are implicated in a single phenotype (Table 1). Simultaneously, on the same chip and in the same hybridization experiment we will also screen DNA from patients with other phenotypes. For example, for a patient with autosomal recessive congenital stationary night blindness we will screen all coding exons and flanking intronic sequences of the following genes: CABP4, GRM6, RDH5 and SAG. Subsequently for a patient with autosomal recessive Bardet-Biedl syndrome we will add the ARL6, BBS1, BBS2, BBS4, BBS5, BBS7, BBS9, BBS10, MKKS, TTC8 and TRIM32 genes on the same chip. We will continue this process, until a large pro- portion of the genes on the chip has been assigned to individual patients (Figure 3).

Validation of known sequence changes on the chip Call rates for nucleotides for the majority of genes and amplicons was high (98% to 100%). Only thirty-two (2%) amplicons were poorly hybridized to the chip and therefore had low call rates. These amplicons were either very GC-rich or otherwise require amplification using a separate protocol before being added to the pool. For future diagnostic purposes these am- plicons will be amplified individually using optimal circumstances before they will be added to the pool and further analysed. The detection rate for point mutations was high, 99% of the known sequence changes could be identified manually, in only 1 case (1%) we could not identify a missense mutation on either chip, despite its presence in the pool.

184 The resequencing chip

Ninety percent of sequence changes were identified by the current version of the Sequence seq-C software. The Sequence seq-C software analysis is based on signal strength and stan- dard deviation from previously analyzed chips. This implies that the software has a learning curve and therefore the number of automatically called sequence changes will increase as the number of analyzed chips increases. Overall, our results indicated a very high sensitivity for the detection of point mutations. Small deletions could not be detected using our approach. Approximately 24% of identified mutations are small insertions or deletions[14]. This is a limitation of resequencing chips. Therefore, we propose to use the retinal resequencing chip as a prescreening tool for diag- nostics. By performing our experiment in duplicate, we showed a high reproducibility of the results (100%).

Additional sequence changes On average, the chip detected one new sequence change per four amplicons. Since all human beings carry unique polymorphic sequence changes, these new sequence changes will also be detected in our resequencing chip. The absence of false positives in our chips show that these changes are all genuine. Obviously, the pathogenic nature of new sequence changes found with the chip (similar to conventional sequencing) needs to be established by screening a large number of ethnically matched controls.

Strengths and limitations The retinal resequencing chip has a number of advantages compared to currently available diagnostic tools. It can detect both known and new sequence changes, in contrast to the cur- rently used mutation chips[4] that detect only known sequence changes. The resequencing chip is five to ten times cheaper than conventional sequencing, even with an occupancy of no more than 50%. An additional cost reduction was achieved by our ability to reduce the amplification volume to 10 µl. Analysis time of the resequencing chip is comparable to direct sequencing methodology.

The retinal resequencing chip shares a number of limitations with most currently used diag- nostic screening techniques. Intronic sequences, with the exception of the splice site regions, cannot be screened since they are absent from the chip and small deletions cannot be de- tected. Although deletions are far less common than point mutations in, for example, retinitis pigmentosa patients[1], not finding a mutation cannot completely exclude the causative role of a gene in the phenotype. Second generation sequencing techniques or larger resequencing chips will enable the analysis of intronic sequences, although these techniques are still expen- sive and have not been validated yet for diagnostic purposes. Unfortunately, newly identified disease genes cannot be added to an existing chip in a later stage and designing an entirely new chip is a costly matter. Moreover, like in direct sequencing, the analysis of amplicons containing GC stretches or repetitive sequences is a challenge. Therefore such amplicons will need to be amplified individually under optimal circumstances, not in a multiplex reaction. Subsequent addition to the pool will yield high call rates for all amplicons. Finally, genes

185 Chapter 8 containing homologous sequences should ideally not be combined on a single chip, due to the chances of cross hybridization.

In conclusion, clinical or genetic diagnosis and classification of genetic retinal disorders suf- fers from extensive genetic and clinical heterogeneity, multiple genetic inheritance modes. There is great need for a flexible and cheap diagnostic tool capable of screening multiple genes in a single experiment. Here, we describe such a flexible, sensitive, high-throughput method for the detection of known and new mutations, the retinal resequencing chip. Our 300 kb retinal resequencing chip, in combination with a unique patient pooling strategy allows for cost-effective direct sequencing of DNA from multiple patients for multiple genes, with very high detection rates (99%), high reproducibility (100%) and low false positive rates (0%) in the tested amplicons. We intend to use this method as a prescreening tool be- fore formal DNA diagnostics. Any sequence changes identified with the resequencing chip will be confirmed using established methods like direct Sanger sequencing of the amplicon in question. For availability see our website (http://www.nin.knaw.nl/research_groups/ber- gen_group/diagnostics/, accessed December 17, 2009).

186 The resequencing chip

References 1. Retnet http://www.sph.uth.tmc.edu/Retnet/ 2. Haim M (2002) Epidemiology of retinitis pigmentosa in Denmark. Acta Ophthalmol Scand Suppl 1-34. 3. Booij JC, Florijn RJ, ten Brink JB, Loves W, Meire F et al (2005) Identification of mu- tations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. J Med Genet 42: e67. 4. Asperbio http://www.asperbio.com/index_testing.htm 5. ABI solid sequencing http://www3.appliedbiosystems.com/AB_Home/application- stechnologies/SOLiDSystemSequencing/index.htm 6. Illumina genome resequencing http://www.illumina.com/pages.ilmn?ID=203 7. Roche 454-sequencing http://www.roche.com/products/product-list.htm?type=researchers&id=4 8. Affymetrix resequencing chip http://www.affymetrix.com/products_services/arrays/ specific/custom_seq.affx#1_1 9. Miller SA, Dykes DD, Polesky HF (1988) A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Res 16: 1215. 10. ELXRdb http:/elxr.swmed.edu/elxrdb_query.html 11. ServiceXS http://www.servicexs.com/ 12. Beiboer SH, Wieringa-Jelsma T, Maaskant-van Wijk PA, van der Schoot CE, van ZR et al (2005) Rapid genotyping of blood group antigens by multiplex polymerase chain reaction and DNA microarray hybridization. Transfusion 45: 667-679. 13. Leveque M, Marlin S, Jonard L, Procaccio V, Reynier P et al (2007) Whole mitochon- drial genome screening in maternally inherited non-syndromic hearing impairment us- ing a microarray resequencing mitochondrial DNA chip. Eur J Hum Genet 15: 1145- 1155. 14. HGMD database http://www.hgmd.cf.ac.uk/ac/hahaha.php

Acknowledgements We thank Prof dr. M.H. Breuning, Dr. A. Plomp, Dr. L. Prick, dr F. Meire and Dr M. van Genderen and other ophthalmologists and clinical geneticists for sending in DNA of patients. We thank the ‘Algemene Nederlandse Vereniging ter Voorkoming van Blindheid’ (ANVVB), the ‘Landelijke Stichting voor Blinden en Slechtzienden’ (LSBS), the ‘Rotterdamse Verenig- ing Blindenbelangen ‘ (RVB), the ‘Stichting Blindenpenning’, the ‘Stichting Oogfonds Ned- erland’ and the ‘Gelderse Blindenstichting’ for their financial support.

187

General discussion

9 Chapter 9

This thesis focused on the molecular, functional, phenotypical and pathological aspects of the retinal pigment epithelium (RPE). We described the molecular and functional properties of the RPE and adjacent structures, including the photoreceptors, the choroid and Bruch’s membrane; we described retinal diseases caused by mutations in genes that are expressed in the RPE; and we described a new mutation screening strategy for RPE-related phenotypes. Chapter one contains the introduction to the subsequent chapters. In chapters two and three we identified genes with high expression in the RPE and genes with tissue specificity in the RPE compared to the photoreceptors and the choroid, based on microarray experiments of 22,000 and 44,000 features respectively. In chapter four we reviewed the composition of Bruch’s membrane and elaborated on the processes of aging and pathology in this membrane. Chapter five described the contribution of the RPE, the choroid and the photoreceptors to the maintenance of an important adjacent structure, Bruch’s membrane. In chapter six we showed the clinical and genetic heterogeneity that exists among two recessive phenotypes, retinitis pigmentosa and Leber congenital amaurosis, caused by mutations in numerous genes expressed in the RPE (and photoreceptors). In chapter seven we described a dominant retinal dystrophy, Best vitelliform macular degeneration (VMD), caused by mutations in a single gene with RPE-specific expression and tried to explain the differences in phenotype by the different causative mutations. Finally, in chapter eight we described the development a high- throughput tool for the detection of mutations in 87 different RPE and photoreceptor disease genes.

Chapter one: Introduction In chapter one we described the anatomy, embryology and function of the retinal pigment epithelium (RPE). We also described a number of important phenotypes related to the RPE.

Chapter two: The human RPE, gene expression and function A widely appreciated problem in the molecular diagnostics of retinal diseases, despite all the available methods of mutation detection, is that many disease causing genes have not yet been identified. In chapter two we characterized the RPE genetically and described the RPE transcriptome. This way we gained insight into the molecular machinery of the RPE and we facilitated future studies toward the identification of additional candidate genes for retinal disease phenotypes. We also gained insight into the processes associated with aging of the RPE. We studied gene expression in the RPE using microarrays. A unique feature of our study of the RPE transcriptome was the inclusion of data on interindividual differences in RPE gene expression levels. This enabled us to not only describe genes expressed in the RPE, but also give an estimate of the differences in expression levels between individuals.

Our data provided a valuable addition to the knowledge on RPE gene expression: only 17% of the top 30 most highly expressed RPE genes were previously known to be expressed in the human RPE in vivo.

190 General discussion

Our findings showed a distinct overlap with previous gene expression studies. Of the 13,000 retina/RPE genes found in at least two studies[1] according to a recent review, we assigned 56% to the RPE. Moreover we showed that 11% have high expression levels in the RPE. Of the 246 genes described to be exclusively expressed in the RPE[1] we confirmed the RPE expression of 56% of these genes and showed high RPE expression levels for 7%.

Our functional analyses of the RPE confirmed the presence of a number of previously known functional pathways (gene regulation, gene transcription, protein metabolism, cell prolifera- tion, survival and signaling, energy metabolism, cytoskeleton and inflammation)[2-4]. It also added several new ones (oxidative phosphorylation, ATP synthesis, ribosome, phosphati- dylinositol signaling, aminosugars metabolism, complement cascade and genes involved in the composition of BM), thereby adding valuable information to the existing knowledge of the molecular machinary of the RPE. We showed that the RPE has a high metabolic activity and energy demand, with high levels of oxidative stress as a result. Our data indicated that RPE-expressed genes contribute significantly to the defense of the RPE cell against oxidative stress.

In chapter two we considered not only genes with high expression levels, but also genes with moderate or relatively low expression in the RPE. We showed that many of the known retinal disease genes are expressed at high levels in the RPE. This may be subject to a bias because genes with high expression levels are more easily detected and are most likely to be identified first. We hypothesized that genes with lower expression levels are also likely to be important for the function and dysfunction of tissues and therefore most likely contain large amounts of undiscovered information.

Chapter two highlights a relatively underappreciated area, namely that of variable gene ex- pression. All studies to date disregarded or minimized the differences in gene expression levels between samples[5]. In our study we also minimized the variability in gene expression levels due to handling of the samples. We further hypothesized that not all gene expression variability is background noise. We showed that certain specific functional groups of genes have high interindividual variability in their expression levels. We found that the highest variability occurs among genes involved in the immune system. This can most likely be ex- plained by both genetic differences and a variable degree of subclinical inflammation (local or systemic) among our donors as well as possible differences in the level of contamination from resident or migrated cell types involved in immunity. We also found somewhat higher interindividual variability in expression among the genes known to be involved in macular retinal diseases. The reason for this is not clear, but these expression differences may, in part, explain the phenotypicial heterogeneity of RPE phenotypes.

Finally, we also observed that 63% of the known macular disease genes were expressed at high levels in our macular RPE sample. In contrast, only 32% of the peripheral retinal disease genes showed high expression in our macular sample. This further suggests that there are topographical differences in RPE gene expression and that functional consequences of muta- tions may in part be dependent on retinal location. A second conclusion that can be drawn

191 Chapter 9 from these data is that information on macular or peripheral gene expression profiles (expres- sion levels) may be indicative in the search for new macular or peripheral disease genes.

In chapter two we gave a first description of the close interaction of the RPE with Bruch’s membrane by way of two functional pathways. We identified an overrepresentation of genes in the RPE encoding extracellular matrix (ECM) proteins and genes encoding proteins in- volved in the degradation of glycosaminoglycans (GAG). ECM proteins, including GAGs are important constituents of BM, as we described in chapter seven.

In spite of the high cellular specificity obtained by using laser dissection microscopy for the isolation of RPE cells, our study was still limited by a small amount of photoreceptor contamination, inevitably present in our RPE sample[2,6,7]. In order to resolve this issue we designed a new strategy and we performed additional experiments described in chapter three.

Chapter three: Human RPE-specific gene expression In chapter three we provided a further specification of the RPE transcriptome, by analyzing the genes expressed in the RPE in reference to their expression in photoreceptors and cho- roid, adding to the knowledge on tissue-specific gene expression. We identified 458 genes specifically expressed in the RPE, 114 genes using stricter criteria, and 39 genes with RPE- specific expression and high expression in the RPE. Fifty percent of the latter has been identi- fied previously in RPE studies[5].

Our analyses of the RPE in chapter three added tissue specificity when compared to the analyses described in chapter two. We described the following functional pathways specific to the RPE when compared to the photoreceptors: cell adhesion, melanogenesis and type one diabetes. Functional categories specific to the RPE when compared to the choroid are the olfactory transduction pathway and vision and transport.

Functional analyses of RPE-specific genes showed an overrepresentation of genes involved in transport and symport, membrane and glycoprotein, visual and nervous system develop- ment and function, ophthalmic disease and genetic disorder. Our analyses also yielded three canonical pathways, RAR-activation, retinol metabolism and GABA receptor signaling. Quite expectedly, these pathways are different from the ones identified in chapter four. By adding tissue specificity we eliminated a number of genes and gene groups from our analyses that are important but not unique to the RPE.

Another important finding from our study regarded the levels of contamination present in studies performed to date. Previous studies described 246 RPE specific genes[1]. We showed that 23 of the 246 genes (9%), had higher expression levels in the photoreceptors than in the RPE and 72 genes (29%) had higher expression levels in the choroid than the RPE. Thus the subject of mRNA contamination from adjacent cell types has not received the attention it deserves in previous gene expression studies on the RPE. However, it is crucial to assign

192 General discussion molecular events to a specific cellular layer and it should be taken into account when inter- preting results. The combination of the highly expressed genes described in chapter two and the 458 genes with specific expression from the current chapter represent, on a molecular level, the most important functional properties of the RPE.

Chapter four: Review of human Bruch’s membrane The RPE and Bruch’s membrane (BM) form one functional unit. BM is an acellular structure, consisting of five layers. It is located between the RPE and the choroid and plays an essen- tial role in many processes involved in vision. When the integrity of BM is compromised, retinal disease like age-related macular degeneration (AMD) can occur. In chapter four we provided a systematic review of Bruch’s membrane. We reviewed the molecular, structural and functional properties of BM and how they are dependent on age, genetic constitution, environmental factors, retinal location and disease state. We also described how some of the properties of BM are unique to each human individual at a given age, and therefore uniquely affect the development of normal vision and ocular disease. BM composition changes with age, including increased calcification of elastic fibres, in- creased cross-linkage of collagen fibres, increased turnover of glycosaminoglycans and ac- cumulation of advanced glycation end products (AGEs) and fat. We described how age-related changes in BM can influence the onset and progression of diseases like retinitis pigmentosa (RP) and age-related macular degeneration (AMD). BM is passively or actively involved in a range of other retinal disorders such as Pseudoxanthoma elasticum (PXE), Sorsby’s Fundus Dystrophy and Malattia Leventinese. The nature of ad- hesive interactions between the RPE and BM is instrumental in the development of retinal detachments and proliferative retinal disease. Undoubtedly, BM is the site of drusen devel- opment. Confluent drusen and uncontrolled activation of the complement cascade are most likely the first sign of AMD. We provided initial evidence, using microarray analysis of RPE cells, photoreceptors and choroid, that multiple structures adjacent to BM may play a major role in the maintenance of BM and may be involved in the onset and development of retinal pathology.

Further characterisation of the proteins in BM and in drusen in healthy eyes and in eyes af- fected by AMD, and the identification of their genetic origin, will enhance the understanding of the processes leading to AMD. Analyzing the changes that occur in eyes affected by AMD compared to healthy eyes may lead to the identification of proteins, genes and/or entire path- ways that trigger the onset of diseases like AMD. This may eventually lead to the develop- ment of gene or protein targeted therapy for AMD.

Chapter five: The origin of human Bruch’s membrane proteins In chapter five we used microarray data derived from the RPE, the photoreceptors and the choroid to identify the origin of proteins that make up the acellular BM.

193 Chapter 9

We showed that most collagen and laminin subtypes known to be present in BM are ex- pressed at higher levels in the choroid than in the RPE. Fibronectin also appears to have higher expression in the choroid, indicating the importance of the choroid in the maintenance of BM compared to the RPE. Nevertheless, the RPE also plays a major role in the maintenance of BM, as seen for instance in the TIMP3 gene, that is highly expressed in the RPE.

For proteins contained in drusen, we showed that most corresponding genes are either ex- pressed at higher levels in the choroid than the RPE (in particular genes involved in extra- cellular matrix receptor interaction and cell-cell communication) or that these proteins were derived from serum. We provided insight into the origin of BM and drusen components and show that the major- ity is derived from choroidal cells and/or serum. This knowledge is useful in the search for a treatment for retinal disorders, including age-related macular degeneration (AMD).

The fact that drusen proteins appear to come from choroid more than from the RPE, implies that the search for ways to prevent the formation of drusen should also be sought in the cho- roid. Serum proteins also appear to contribute substantially to drusen formation. This indi- cates that drusen are most likely formed by both systemic and local factors, and that potential treatment strategies for AMD should take this into account.

Chapter six: mutation analysis in early-onset peripheral retinal dystrophies Retinitis pigmentosa (RP) and Leber congenital amaurosis (LCA) are two retinal dystrophies with a wide spectrum of clinical features, but also with extensive overlap. The phenotypic heterogeneity can perhaps in part be explained by the genetic heterogeneity that exists among both phenotypes. autosomal recessive RP (arRP) can be caused by mutations in 17 genes, LCA by mutations in 13 genes[8]. Moreover, 8 of these genes have been reported in both phenotypes[8]. The genes responsible for RP and LCA are expressed mainly in the retinal pigment epithe- lium (RPE) and in the photoreceptors. We showed in chapter four that many retinal disease genes have high expression levels in the RPE. This knowledge can be very useful for the identification of new disease genes: the expression level of genes in the RPE can be used as a criterion for identifying new candidate disease genes.

In order to describe the genetic overlap, we tested juvenile RP patients as well as LCA pa- tients for mutations in LCA genes. In chapter six, we described 26 patients with juvenile RP and nine patients with LCA. We identified mutations in 34% of our 35 patients: in 44% of our LCA patients and in 31% of our RP patients. The former is in accordance with the literature, the latter is less than what is reported, due to the fact that we screened these RP patients for mutations in LCA genes[9]. In more than half the LCA and RP patients, a causative mutation cannot be found in spite of the recessive nature of the two phenotypes (our study and literature)[10]. This is due to a number of factors. First, due to the phenotypical heterogeneity, a certain diagnosis cannot

194 General discussion always be made, or with progression of disease, the clinical diagnosis is changed. Even when a certain clinical diagnosis can be established, there are many possible causative genes, not all of which are usually screened in diagnostics, due to the high costs of screening multiple genes one by one by direct sequencing. Secondly, genes commonly associated with one phe- notype can, in individual cases, be responsible for another phenotype. It is difficult if not im- possible to determine which genes should be screened based on the phenotype. Thirdly, direct sequencing of individual genes, although it allows us to find new and known mutations, is limited with respect to either the identification of intronic sequence changes or the identi- fication of deletions or insertions. In chapter eight we described a new screening method developed to reduce cost and to increase speed of screening. Most currently used techniques are not capable of identifying intronic promoter or regulatory sequences that may contain yet unidentified mutations. Finally, one or more genes could be involved in a phenotype, making the disease digenic or polygenic.

In chapter six, we established or confirmed the polymorphic nature of three sequence changes previously reported as pathogenic mutations. This indicates that care should be taken when interpreting a sequence change in a patient. This is especially relevant with the introduction of high throughput sequencing technology, where many sequence changes will be found while only few are pathogenic.

In many patients the disease causing sequence changes can still not be identified. In part, this can be explained by the fact that a great number of disease genes has not yet been identified. However, it is also very likely that known disease genes contain pathogenic changes that have not been identified, for instance, deletions or insertions that cannot be found with most currently used techniques, or changes in intronic sequences that are currently overlooked. It is also possible that sequence changes, found to be polymorphic in certain populations, will prove to be disease causing in others. In order to resolve these issues, is seems imperative that large numbers of healthy controls are screened for new sequence changes, using methods that are internationally agreed upon. Large scale screening methods like whole genome se- quencing may prove to be a good strategy. Creating databases with information from multiple nationalities will provide valuable information that can be used when new sequence changes are identified in patients.

Large scale studies of patients with arRP and LCA including phenotype and genotype charac- terization, should provide insights into the further delineation of the phenotypes.

Chapter seven: Genotype-phenotype correlation in vitelliform macular dystro- phy Vitelliform macular degeneration (VMD) is primarily a dominant retinal degeneration caused by mutations in the Best1 gene, a gene expressed at high levels in the RPE. The age of onset and clinical course of VMD are highly variable[11].

195 Chapter 9

In chapter seven, we described the clinical course of 53 VMD patients. We showed the dis- ease has a variable age of onset and clinical course, ultimately leading to visual decline, but rarely to blindness. The high variability of the phenotype exists between and within pedigrees and there was little correlation between the left and right eyes of individual patients. Even unilateral expression ocurred.

Two subtypes of VMD, congenital Best disease and adult-onset vitelliform macular dystro- phy were previously thought to be two distinct phenotypes [12,13]. The phenotypical data in chapter seven suggest that it is more likely that the two are part of a continuous spectrum.

Unlike RP and LCA, VMD is caused by mutations in a single gene, which is uniquely ex- pressed in the RPE. Nevertheless, our survival analyses indicate that the variability in phe- notype may still, in part, be explained by differences in genotype. More specifically, the Ala10Val mutation showed a significantly faster decline causing severe visual impairment (VA<0.3) in fifty percent of patients at age 34 years, compared to 53 years for another fre- quently observed mutation. For a third frequently observed mutation, fifty percent did not even reach a VA<0.3 during the study period. Additional factors influencing the variable VMD phenotype may be polymorphic sequence changes in the mutated allele or the healthy allele, that influence the severity of the disease. This is not uncommon in ophthalmic disor- ders and has been shown in autosomal dominant retinitis pigmentosa (RP11 and RP1) [14]. Moreover, the actual folding of the protein into a three dimensional structure could be dif- ferent from the currently existing models. This may cause mutations to be in intracellular stretches instead of extracellular stretches or vice versa, which may explain the differences in the clinical course observed in chapter two.

Long term follow up of large groups of VMD patients with different mutations in the Best1 gene will provide useful prognostic information. Moreover, further insights into the exact location of the intracellular and extracellular domains of the protein and into the function of the channel formed by the protein, may help identify regions of the gene that give, when mutated, a better or a worse prognosis.

An important indication that VMD may not be purely determined by genotype, comes from a VMD patient with poor vision in one eye and very good vision in the other eye (see chapter 7). Determining the factors that influence the onset of VMD is crucial in any attempt to slow down or halt the phenotype. Identifying whether a patient with a unilateral presentation of VMD suffers from mosaicism, has differences in gene expression levels, has differences in protein levels or turnover rates, or is subject to different environmental factors in his left and right eye will provide important clues.

Chapter eight: High throughput mutation analysis In chapter eight we described a new tool that was developed to enable the screening of multiple retinal disease genes in a single cost-effective experiment: the retinal resequencing chip. Recently, efforts have been made to facilitate the screening of multiple genes at once,

196 General discussion like the mutation chips by Asper[15]. Unfortunately this technique is limited to the detec- tion of known sequence changes. Second generation sequencing (Applied Biosystems [16], Illumina[17], Roche[18]), whereby 20 Gb can be sequenced simultaneously may provide a solution in the future, but currently this technique is very expensive, and is not yet validated for diagnostic purposes. Therefore, we developed a high throughput method for the detection of new and known se- quence changes in 87 retinal disease genes. This retinal resequencing chip is capable of direct sequencing both forward and reverse strands of all coding exons of 87 retinal disease genes with high sensitivity and specificity. Moreover, it is capable of detecting both known and new mutations corresponding to 30 retinal phenotypes [9]. Our unique pooling strategy al- lows for the screening of multiple patients and multiple genes per chip, which will reduce the costs per screened individual considerably. The chip has a sensitivity of 99%, when manually analyzing the data. The sensitivity of the JSI software is currently 90%. However, adding new chips, to be used for normalization, will increase the sensitivity of the software. The re- producibility of the chip is 100% and the specificity is 100% for the detection of single base changes. Therefore, we believe this chip is a valuable addition to the tools currently available for screening patients with retinal diseases.

Advantages of our retinal resequencing chip include the use of a multiplex pcr design, with an average of four amplicons of different sizes per pcr reaction, thereby ensuring a signifi- cant time reduction. In addition, for diagnostic purposes, we will be able to include multiple patients in a single chip experiment, thereby significantly reducing costs. Only amplicons for genes related to the phenotype of the patient will be amplified and hybridized to a chip, along with the amplicons of different patients relevant for their phenotype.

Our retinal resequencing chip is limited in its inability to identify small deletions. In addition, two percent of all amplicons were poorly hybridized to the chip. Like in conventional direct sequencing, these amplicons will require individual amplification using a specific protocol. This will slightly reduce the efficiency of the resequencing protocol. Unfortunately, newly identified disease genes cannot be added to an existing chip in a later stage and designing an entirely new chip is a costly matter. In the future however, large scale sequencing of entire genes will become more affordable, thereby reducing the need for chips capable of screening only fixed sets of genes. Combining our knowledge of genes highly expressed in the RPE with the ability to sequence large numbers of genes in a single experiment, could prove to be very useful. Screening large numbers of genes with high RPE expression in patients with a retinal disease, may lead to the identification of new disease genes. Moreover, combining this strategy with a homozygosity mapping approach in consanguineous families with retinal disease, will greatly reduce the number of genes to be screened and will increase the chances of finding a disease causing mutation.

Identification of mutations in new genes in patients with retinal diseases may lead to novel in- sights into the etiology and improvement of further clinical classification of the phenotypes.

197 Chapter 9

References 1. Schulz HL, Goetz T, Kaschkoetoe J, Weber BH (2004) The Retinome - defining a ref- erence transcriptome of the adult mammalian retina/retinal pigment epithelium. BMC Genomics 5: 50. 2. Ishibashi K, Tian J, Handa JT (2004) Similarity of mRNA phenotypes of morphologi- cally normal macular and peripheral retinal pigment epithelial cells in older human eyes. Invest Ophthalmol Vis Sci 45: 3291-3301. 3. Buraczynska M, Mears AJ, Zareparsi S, Farjo R, Filippova E et al (2002) Gene expres- sion profile of native human retinal pigment epithelium. Invest Ophthalmol Vis Sci 43: 603-607. 4. Sharon D, Blackshaw S, Cepko CL, Dryja TP (2002) Profile of the genes expressed in the human peripheral retina, macula, and retinal pigment epithelium determined through serial analysis of gene expression (SAGE). Proc Natl Acad Sci U S A 99: 315-320. 5. Booij JC, ten Brink JB, Swagemakers SM, Verkerk AJMH, Essing AHW et al (2010) A New Strategy to Identify and Annotate Human RPE-specific Gene Expression. Plos one in press. 6. van Soest S, de Wit GM, Essing AH, ten Brink JB, Kamphuis W et al (2007) Compari- son of human retinal pigment epithelium gene expression in macula and periphery high- lights potential topographic differences in Bruch’s membrane. Mol Vis 13: 1608-1617. 7. Peng YW, Robishaw JD, Levine MA, Yau KW (1992) Retinal rods and cones have dis- tinct G protein beta and gamma subunits. Proc Natl Acad Sci U S A 89: 10882-10886. 8. Retnet www.sph.uth.tmc.edu/Retnet/ 9. www.sph.uth.tmc.edu/retnet/disease.htm 10. Booij JC, Florijn RJ, ten Brink JB, Loves W, Meire F et al (2005) Identification of mu- tations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. J Med Genet 42: e67. 11. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ et al (2009) Function- al annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 10: 164. 12. Kramer F, White K, Pauleikhoff D, Gehrig A, Passmore L et al (2000) Mutations in the VMD2 gene are associated with juvenile-onset vitelliform macular dystrophy (Best dis- ease) and adult vitelliform macular dystrophy but not age-related macular degeneration. Eur J Hum Genet 8: 286-292. 13. Musarella MA (2001) Molecular genetics of macular degeneration. Doc Ophthalmol 102: 165-177. 14. Daiger SP, Shankar SP, Schindler AB, Sullivan LS, Bowne SJ et al (2006) Genetic fac- tors modifying clinical expression of autosomal dominant RP. Adv Exp Med Biol 572: 3-8. 15. Asper (2010) http://www.asperbio.com/index.php?nid=6&pid=5 16. Applied Biosystems (2010) http://www3.appliedbiosystems.com/AB_Home/applica- tionstechnologies/SOLiD-System-Sequencing-B/index.htm 17. Illumina (2010) http://www.illumina.com/technology/sequencing_technology.ilmn 18. Roche (2010) http://www.roche.com/media/media_releases/med_dia_2007-01-16.htm

198 General discussion

199

Summary

10 Chapter 10

Chapter one: Introduction Chapter one describes the anatomy, embryology and function of the retinal pigment epithelium (RPE). It also describes a number of important phenotypes related to the RPE.

Chapter two: The human RPE, gene expression and function The RPE is one of the most important cell layers of the retina, involved in vision. Therefore it not surprising that genes expressed in the RPE are also involved in macular and peripheral retinal diseases like age-related macular degeneration (AMD) and retinitis pigmentosa (RP). We determined level and variability of gene expression in the RPE and we performed bioin- formatic analyses in order to gain insight into the functional properties of the RPE. To this end we isolated macular RPE cells from healthy human donor eyes using laser dissection mi- croscopy and we performed microarray analyses using 22k microarrays. Our bioinformatic analyses were carried out using the web tool David and Ingenuity software.

The top 10% of genes with the highest expression levels in the RPE showed an overrepre- sentation of genes involved in oxidative phosphorylation, ATP synthesis and the function of the ribosome. The top 10% of genes with the highest interindividual variation in expression included genes involved in AMD, in signaling and in the complement cascade. Surprisingly, genes involved in Bruch’s membrane (BM) composition also showed very highly variable expression levels.

In conclusion, we showed in chapter two that the RPE has strong energy demands, high levels of protein synthesis and is exposed to high levels of oxidative stress and to a variable degree of inflammation. These data confirm that the results from a large number of functional studies in animal models can also be applied to humans. In addition, they can be useful for the identification of candidate retinal disease genes.

Chapter three: Human RPE-specific gene expression Large scale studies into RPE gene expression are often limited in specificity due to poor sample selection, cellular manipulation or contamination from adjacent cell types. In chapter three we analyzed the molecular characteristics of the RPE with a specific focus on elimi- nating contamination from cell types adjacent to the RPE, photoreceptors and choroid. We isolated RPE, photoreceptors and choroidal cells from healthy human donor eyes using la- ser microdissection in order to identify RPE-specific expressed transcripts. In addition, we expanded our dataset from 22,000 features (chapter two) to 44,000 features, covering the entire human genome. We made triplicate comparisons of the transcriptomes of the RPE, the photoreceptors and the choroid and we deduced RPE-specific expression.

We identified 114 genes with RPE-specific expression (defined as at least 2.5 times higher expression in the RPE than in the adjacent photoreceptor and choroidal cells in all compari- sons). Thirty-nine of these 114 genes also showed high expression in the RPE, as determined

202 Summary in our experiments from chapter two[1]. In order to find support for our findings, we per- formed a literature search. Indeed, 85% of these 39 genes were previously identified to be expressed in the RPE. The remaining 15% were not identified previously. However, we con- firmed their expression in the RPE by sRT-PCR. In the group of 114 RPE-specific genes there was an overrepresentation of genes involved in ophthalmic disease and genes involved in (membrane) transport and vision. More fundamentally, we found RPE-specific involvement in the retinoic acid receptor (RAR)-activation, the retinol metabolism and gamma aminobu- tyric acid (GABA)-receptor signaling pathways. In conclusion, in chapter five we provided a further specification and understanding of the RPE transcriptome by identifying and analyz- ing genes that are specifically expressed in the RPE.

Chapter four: Review of human Bruch’s membrane Bruch’s membrane (BM) is a unique pentalaminar structure, strategically located between the RPE and the choroid. It plays an essential role in maintaining the function of the retina. This acellular structure regulates the diffusion of molecules between the RPE/retina and the general circulation. More specifically, together with the RPE it regulates the exchange of nutrients, oxygen, fluids and metabolic waste between the choroid and the photoreceptors. It also provides physical support for RPE cell adhesion, migration and possibly differentiation, and it acts as a barrier restricting choroidal and retinal cellular migration[2]. With age, many functional properties of BM change, leading not only to normal age-related changes in the RPE and photoreceptor cells, but possibly also influencing the onset and/or progression of diseases like RP and AMD. In the aging BM, white-yellow deposits, drusen, can be seen[3]. Soft confluent drusen are a first manifestation of AMD[2], indicating the in- volvement of BM in the onset and development of this disease.

In chapter four we reviewed the current literature on the normal structure and function of BM and the changes that occur with age and placed them in the context of our new RPE expression data. We described the formation and composition of drusen and we discussed the changes that occur in pathological processes like AMD.

Chapter five: The origin of human Bruch’s membrane proteins Studies in chickens and mice have shown that both the RPE and the choroid contribute to the development of BM, and it has been suggested that BM is composed of molecules synthe- sized both in the RPE and the choroid[4,5]. Indeed, the acellular nature of BM explains the necessity of involvement of adjacent tissues in the development, maintenance and turnover of BM. However, the extent of involvement of the RPE, the choroid or even the photorecep- tors has not been thoroughly examined in humans.

In chapter five we studied the human cellular expression levels of RPE and choroidal genes coding for BM proteins, in order to gain insight into the origin of BM components. In a simi- lar fashion we studied the genes encoding drusen proteins in order to determine the origin of proteins present in drusen.

203 Chapter 10

We showed that most collagen and laminin subtypes known to be present in BM are ex- pressed at higher levels in the choroid than in the RPE. For drusen proteins we showed that most corresponding genes are either expressed at higher levels in the choroid than in the RPE or are derived from serum. The drusen proteins with higher gene expression in the choroid are involved in extracellular matrix receptor interaction and cell-cell communication. This suggests the choroid and serum play an important role in the formation of drusen. The ma- jority of BM and drusen proteins appear to be derived from the choroid. This knowledge on the origin of BM and drusen proteins may have implications for the research into treatments for retinal disorders like AMD. The knowledge on the origin of BM proteins may also have implications for the age of onset of retinal disorders.

Chapter six: Mutation analysis in early-onset peripheral retinal dystrophies Autosomal recessive retinitis pigmentosa (arRP) and Leber congenital amaurosis (LCA) are two retinal dystrophies, characterized by RPE and photoreceptor cell death, primarily affect- ing the peripheral retina. During the course of the disease, in both cases the central retina may become involved. RP is characterized by night blindness, progressive constriction of the vi- sual field and typical fundus abnormalities. The disease is very heterogeneous and is thought to have overlapping features with LCA. LCA is an autosomal recessive retinal dystrophy that leads to severe visual impairment in the first year of life. LCA is characterized by a con- genital nystagmus and a decreased pupillary response. Onset of severe visual complaints for LCA patients is usually earlier in life than for juvenile RP patients. Overlap between the two phenotypes consists of pigmentary changes in the retina and a reduced electro-retinogram (ERG). Many intermediate phenotypes have been described, both concerning age of onset, symptoms and retinal signs[6].

Genetically, at least 17 genes are currently known to be involved in arRP. Thirteen LCA genes have been cloned; five of these genes can cause both arRP and LCA[7]. In chapter six we screened 35 unrelated patients with juvenile arRP, juvenile isolated RP (IRP, RP with no known inheritance pattern, for example due to the fact that the patient is the only affected person in a family), or LCA, for mutations in five LCA genes, AIPL1, CRB1, GUCY2D, RPE65 and RPGRIP1. We found mutations in 34% of our patients, in 17% of arRP patients, in 56% of IRP patients and in 44% of LCA patients. Mutations were found in CRB1 (11% of all patients), GUCY2D (11% of all patients), RPE65 (6% of all patients), and RPGRIP1 (6% of all patients). Among the mutations identified, there was a new combination of two muta- tions in CRB1, and new mutations in GUCY2D and RPGRIP1. The new GUCY2D mutation did not lead to the commonly associated LCA, but to a juvenile RP phenotype. RP patients carried mutations in three genes associated with LCA (CRB1, GUCY2D, and RPGRIP1) and LCA patients carried mutations in CRB1, GUCY2D, and RPE65. These data, combined with previous reports, suggest that LCA and juvenile RP are closely related and belong to a continuous spectrum of juvenile retinitis pigmentosa. The identification of mutations remains important for the selection of patients for (future) clinical trials and (gene) therapy.

204 Summary

Chapter seven: Genotype-phenotype correlation in vitelliform macular dystro- phy Vitelliform macular degeneration (VMD) is a retinal dystrophy primarily affecting the macu- lar area. In VMD the RPE cells and photoreceptors eventually become damaged. The VMD phenotype can also lead to significant vision loss. VMD predominantly has an autosomal dominant inheritance pattern, although autosomal recessive inheritance is also oberved, and has a prevalence of approximately one in 10,000[8]. The clinical diagnosis is based on a decreased visual acuity (VA), the presence of one of six typical VMD stages in the macular area and a decreased light peak-dark trough ratio on the electro-oculogram (EOG). VMD is caused by a mutation in the Best1 gene, coding for a protein with four transmembrane do- mains, the transcribed protein is located in the basolateral membrane of the RPE. The onset and clinical course of VMD are highly variable, it is uncertain whether this is caused by the type and/or the location of the mutation.

In chapter seven we described the phenotype-genotype correlation in a group of 53 patients with VMD, proven by mutation analysis, with an mean follow up time of 15.3 years for 40 patients. Virtually all patients showed a gradual decline in VA without reaching legal blind- ness. However, we also found that the phenotype was highly variable, and in some cases depended strongly on the genotype: patients with an Ala10Val mutation had significantly faster progression compared to patients with a Thr6Pro or a Tyr227Asn mutation. In addition, we described two pedigrees with the same mutation, but with highly variable phenotypes. Finally, our data demonstrated there are great differences in clinical consequences associ- ated with the presence of a disease-causing mutation. We showed a proband with a disease- causing sequence change and a pronounced phenotype; we also showed related carriers of the same sequence change with no pathology. Insight into the underlying mechanisms and further studies in a larger number of pedigrees may provide clues for future clinical care.

Chapter eight: High throughput mutation analysis The RPE and the photoreceptors express the great majority of retinal disease genes at high levels. Currently, approximately 150 retinal disease genes have been identified for as many as 30 different retinal diseases. The identification of a causative mutation is crucial for con- firmation of the clinical diagnosis, for patient counseling and for future gene-targeted treat- ment. However, many phenotypes are poorly delineated and may show overlap with other phenotypes. In addition, the large number of genes that can be associated with the disease pose a diagnostic challenge. The costs to screen these genes one by one in every patient are prohibitively high. A screening tool that can identify both known and new mutations in a large number of potential disease genes in a cost-effective manner is therefore essential.

In chapter eight we described the development and validation of a new 300 kb retinal rese- quencing chip, capable of screening 90 different retinal disease genes in a single experiment. We developed and tested multiplex primers for 1,445 amplicons corresponding to the genes on the chip. In order to validate the chip, we performed the following duplicate experiment: 87 exons from 25 retinal disease genes containing 87 known sequence changes, were simul-

205 Chapter 10 taneously amplified from DNA samples of 60 patients with a variety of retinal disorders. We found call rates for successfully hybridized amplicons of 98% to 100%. The sensitivity of the chip for point mutations was 99%, while, as expected, deletions could not be detected. The specificity of the chip was 100%. The large number of genes on the chip enables us, for future diagnostic purposes, to analyse multiple patient with phenotypes corresponding to mutations in non-overlapping genes together on a single chip.

206 Summary

References 1. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ et al (2009) Function- al annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 10: 164. 2. Booij JC, Baas DC, Beisekeeva J, Gorgels TG, Bergen AA (2010) The dynamic nature of Bruch’s membrane. Prog Retin Eye Res 29: 1-18. 3. de Jong PT (2006) Age-related macular degeneration. N Engl J Med 355: 1474-1485. 4. Olson MD (1979) Development of Bruch’s membrane in the chick: an electron micro- scopic study. Invest Ophthalmol Vis Sci 18: 329-338. 5. Hirabayashi Y, Fujimori O, Shimizu S (2003) Bruch’s membrane of the brachymorphic mouse. Med Electron Microsc 36: 139-146. 6. Booij JC, Florijn RJ, ten Brink JB, Loves W, Meire F et al (2005) Identification of mu- tations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. J Med Genet 42: e67. 7. Retnet www.sph.uth.tmc.edu/Retnet/ 8. Boon CJ, Klevering BJ, Leroy BP, Hoyng CB, Keunen JE et al (2009) The spectrum of ocular phenotypes caused by mutations in the BEST1 gene. Prog Retin Eye Res 28: 187-205.

207

Nederlandse samenvatting

11 Chapter 11

Functie en dysfunctie van het retinaal pigment epitheel

Hoofdstuk één: Introductie Hoofdstuk 1 beschrijft de anatomie, embryologie en de functie van het retinaal pigment epi- theel (RPE). Daarnaast beschrijft het enkele belangrijke ziektes (retinitis pigmentosa, Leber congenitale amaurosis en Best vitelliforme macula degeneratie) die betrekking hebben op het RPE.

Hoofdstuk twee: Het menselijke RPE, gen expressie en functie Het RPE is één van de belangrijkste cellagen van het netvlies, betrokken bij het zien. Het is daarom niet verwonderlijk dat genen die in het RPE tot expressie komen ook betrokken zijn bij maculaire en perifere retinale aandoeningen zoals macula degeneratie en retinitis pigmen- tosa (RP). We hebben bepaald welke genen in het RPE tot expressie komen en het niveau en de variabiliteit van deze genexpressie bepaald. Vervolgens hebben we met behulp van bioinformatica inzicht verkregen in de functionele eigenschappen van het RPE. Hiertoe iso- leerden we maculaire RPE cellen uit gezonde humane donorogen met behulp van laser dis- sectie microsopie. We voerden microarray analyses uit gebruik makend van 22k microarrays. functionele karakterisatie van het RPE hebben we uitgevoerd met een internet-gebaseerd programma David en Ingenuity software.

De top 10% genen met de hoogste expressie niveaus in het RPE lieten een oververtegenwoor- diging zien van genen die betrokken zijn bij de oxidatieve fosforilering, ATP synthese en het functioneren van het ribosoom. De 10% genen met de grootste interindividuele variatie in ex- pressie bevatten genen die betrokken zijn bij ouderdoms macula degeneratie (AMD), genen betrokken bij signaal overdracht in de cel en genen uit de complement cascade. Verrassend genoeg lieten de genen betrokken bij de continue opbouw en afbraak van Bruch’s membraan (BM) ook zeer hoog variabele expressie zien.

Concluderend hebben we in hoofdstuk twee laten zien dat het RPE een grote energie behoefte heeft, een hoge mate van eiwit synthese heeft, en dat het blootstaat aan hoge niveaus van oxidatieve stress en een variabele mate van ontsteking. De data uit dit hoofdstuk bevestigen dat een groot aantal resultaten uit functionele studies in diermodellen ook op de mens van toepassing is. Daarnaast kunnen ze nuttig zijn voor het identificeren van nieuwe kandidaat- genen voor retinale ziektes.

Hoofdstuk drie: Humane RPE-specifieke gen expressie Grootschalige studies naar menselijke RPE genexpressie hebben vaak een beperkte specifici- teit als gevolg van beperkte selectie en karakterisatie van het donorweefsel, door celmanipu- latie of door contaminatie door naastgelegen cel types. In hoofdstuk drie hebben we de mo- leculaire karakteristieken van het RPE geanalyseerd aan de hand van een nieuwe strategie. Essentieel onderdeel van deze strategie was dat we de cellulaire en moleculaire contaminatie

210 Nederlandse samenvatting door analyse van beide naastgelegen cel types, de fotoreceptoren en het choroid, hebben geëlimineerd. Hiertoe isoleerden we eerst RPE, fotoreceptoren en choroïd cellen uit gezonde humane donorogen met behulp van laser dissectie microscopie om genen te identificeren die specifiek in het RPE tot expressie komen. We maakten, in drievoud, vergelijkingen van het transcriptoom van het RPE, de fotoreceptoren en het choroïd, en leidden daaruit RPE- specifieke expressie af. Bovendien breidden we onze dataset uit van 22.000 genen (hoofdstuk twee) naar 44.000 genen, daarmee het hele humane genoom dekkend.

We hebben 114 genen met RPE-specifieke expressie geïdentificeerd (gedefinieerd als ten minste 2,5 keer hogere expressie in het RPE dan in de naastgelegen fotoreceptoren en cho- roidale cellen in alle vergelijkingen). Negenendertig van deze 114 genen hadden ook hoge expressie in het RPE, zoals gedefiniëerd in onze experimenten uit hoofdstuk twee[1]. Om onze bevindingen te bevestigen, hebben we een literatuuronderzoek uitgevoerd. In de li- teratuur bleek van 85% van deze 39 genen reeds beschreven te zijn dat ze in het RPE tot expressie komen. Van de overige 15% is de expressie in het RPE nog niet onderzocht danwel beschreven in de literatuur, maar door ons met sRT-PCR bevestigd. Binnen de groep van 114 RPE-specifieke genen was er een overrepresentatie van genen betrokken bij oogheelkundige aandoeningen en genen betrokken bij (membraan) transport, en visus. Meer fundamenteel vonden we RPE-specifieke betrokkenheid in de retinoic acid receptor (RAR)-activatie, het retinol metabolisme en de gamma aminobutyric acid (GABA) receptor signalering. Conclu- derend, hebben we in hoofdstuk drie een verdere specificatie en een beter begrip van het RPE transcriptoom gegeven, door het identificeren en analyseren van genen die specifiek in het RPE tot expressie komen.

Hoofdstuk vier: Bespreking van Bruch’s membraan Bruch’s membraan (BM) is een unieke pentalaminaire structuur, strategisch gelokaliseerd tussen het RPE en het choroïd. Het speelt een essentiële rol in het behouden van de functie van het RPE en de neurale retina. Deze acellulaire structuur reguleert mede de diffusie van moleculen tussen de retina en de bloedcirculatie in de rest van het lichaam. Meer specifiek reguleert het samen met het RPE de uitwisseling van voedingsstoffen, zuurstof, vloeistoffen en metabole afvalstoffen tussen het choroïd en de fotoreceptoren. BM biedt ook fysieke steun aan het RPE, faciliteert cel-celadhesie, speelt mogelijk een rol bij de differentiatie en fungeert bovendien als barriëre tegen de migratie van choroïdale en retinale cellen[2]. Met de leeftijd veranderen veel functies van het BM die belangrijk zijn voor het normale functioneren van de retina maar mogelijk ook medebepalend zijn voor het aanvangstijdstip en/of het verloop van aandoeningen als RP en AMD. Op latere leeftijd kunnen in BM wit- gele ophopingen, drusen, worden waargenomen[3]. Zachte drusen kunnen in BM voorkomen en zijn de eerste manifestatie van AMD[3]. Dit geeft de betrokkenheid van BM bij het ont- staan van deze aandoening aan.

211 Chapter 11

In hoofdstuk vier geven we een overzicht van de huidige literatuur over de normale struc- tuur en functie van BM en de veranderingen die met de leeftijd optreden. We beschrijven de vorming en bestanddelen van drusen en we bespreken de veranderingen die optreden tijdens pathologische processen zoals AMD.

Hoofdstuk vijf: Oorsprong van Bruch’s membraan en drusen componenten Studies in kippen en muizen hebben aangetoond dat zowel het RPE als het choroïd bijdragen aan de ontwikkeling van BM[4-6]. De noodzaak van de betrokkenheid van naastgelegen weefsels in de ontwikkeling, het onderhoud en de continue afbraak en heropbouw van BM kan worden verklaard door het acellulaire karakter van BM. De mate van betrokkenheid van het RPE, het choroïd of zelfs de fotoreceptoren was nog nooit grondig bestudeerd in mense- lijke netvliezen.

In hoofdstuk vijf bestudeerden we de expressie niveaus van genen die coderen voor BM eiwitten in het RPE, het choroïd en de fotoreceptoren van gezonde humane donorogen, om inzicht te verkrijgen in de oorsprong van BM componenten. Gebruik makend van dezelfde methodes bestudeerden we ook de genen die coderen voor drusen eiwitten om zo de oor- sprong van eiwitten in drusen te achterhalen.

We toonden aan dat de meeste collageen en laminine subtypes, waarvan bekend is dat ze in BM aanwezig zijn, een hogere corresponderende genexpressie hebben in het choroïd dan in het RPE. Voor drusen eiwitten toonden we aan dat de meeste corresponderende genen hoger tot expressie komen in het choroid dan in het RPE of dat ze afkomstig zijn uit het serum. De drusen eiwitten met hogere genexpressie in het choroid zijn betrokken bij extracellulaire matrix-receptor-interactie en cel-cel communicatie. De meerderheid van de BM en drusen eiwitten lijkt dus afkomstig te zijn uit het choroid. Deze kennis over de afkomst van BM en drusen eiwitten kan implicaties hebben voor het onderzoek naar behandelingen van retinale aandoeningen zoals AMD. De kennis over BM eiwitten kan eveneens van belang zijn voor de leeftijd waarop monogene retinale aandoeningen ontstaan.

Hoofdstuk zes: mutatie analyse in perifere retinale dystrofieën met een vroege ontstaansleeftijd Autosomaal recessieve retinitis pigmentosa (arRP) en Leber congenitale amaurosis (LCA) zijn twee retinale dystrofieën, die gekarakteriseerd worden door celsterfte in het RPE en de fotoreceptoren, beginnend in de perifere retina. In de loop van het ziektebeeld kan bij beide aandoeningen de centrale retina ook aangedaan raken. RP wordt gekenmerkt door nacht- blindheid, progressieve cirkelvormige beperking van het visuele blikveld en typische pig- mentvlekken in de fundus. De aandoening is zeer heterogeen en lijkt genetische en klinische overlap te vertonen met LCA. LCA is een autosomaal recessieve retinale dystrofie die leidt tot ernstige visuele beperkingen in het eerste levensjaar. LCA wordt gekarakteriseerd door een congenitale nystagmus en een verminderde pupilreactie. Ernstige visuele klachten ont- staan bij LCA patiënten meestal eerder dan bij juveniele RP patiënten. Overlap tussen de

212 Nederlandse samenvatting twee fenotypes bestaat uit pigmentveranderingen in de retina en een verlaagd signaal op het electro-retinogram (ERG). Er bestaan veel fenotypische tussenvormen, zowel wat betreft ontstaansleeftijd, symptomen als retinale veranderingen[7].

Genetisch zijn er momenteel minstens 17 genen bekend die arRP kunnen veroorzaken, terwijl 13 LCA genen momenteel gekloneerd zijn; mutaties in vijf van deze genen kunnen zowel arRP als LCA veroorzaken[8]. In hoofdstuk zes screenden we 35 ongerelateerde patiënten met juveniele arRP, juvenile geïsoleerde RP (IRP, RP waarbij een overervingspatroon niet is vastgesteld of kan worden vastgesteld vanwege het feit dat er slechts 1 patiënt is), of LCA op mutaties in vijf LCA genen, AIPL1, CRB1, GUCY2D, RPE65 en RPGRIP1. We vonden mutaties in 34% van alle patiënten; in 17% van de arRP patiënten, in 56% van de IRP pati- ënten en in 44% van de LCA patiënten. Mutaties werden gevonden in CRB1 (11% van alle patiënten), GUCY2D (11% van alle patiënten), RPE65 (6% van alle patiënten) en RPGRIP1 (6% van alle patiënten). Tussen de geïdentificeerde mutaties bevond zich een combinatie van twee verschillende mutaties in CRB1 en twee nieuwe mutaties in GUCY2D en RPGRIP1. De nieuwe GUCY2D mutatie leidde niet tot LCA, waar mutaties in dit gen doorgaans mee geas- socieerd worden, maar tot een juveniel RP phenotype. RP patiënten droegen mutaties in drie genen geassocieerd met LCA (CRB1, GUCY2D en RPGRIP1) en LCA patiënten droegen mutaties in CRB1, GUCY2D en RPE65. Deze data, gecombineerd met eerdere publicaties, suggereren dat LCA en juveniele RP nauw aan elkaar gerelateerd zijn en behoren tot één continu spectrum van juvenile retinitis pigmentosa. Het identificeren van mutaties blijft be- langrijk voor de selectie van patiënten voor (toekomstige) klinische trials en (gen) therapie.

Hoofdstuk zeven: Fenotype-genotype correlatie in vitelliforme macula degen- reratie Vitelliforme macula degeneratie (VMD) is een retinale dystrofie die met name betrekking heeft op de gele vlek (macula). Bij VMD raken de centrale RPE cellen en secundair de foto- receptoren beschadigd. Het VMD fenotype kan ook leiden tot significant visus verlies. VMD erft voornamelijk autosomaal dominant en soms autosomaal recessief over en heeft een pre- valentie van één op 10.000[9]. De klinische diagnose is gebaseerd op een achteruitgang van de gezichtsscherpte, de aanwezigheid van één van zes typerende VMD stadia in de gele vlek en verlaagde electrische potentialen op het electro-oculogram (EOG). VMD wordt veroor- zaakt door een mutatie in het Best1 gen. Dit gen codeert voor een eiwit met vier transmem- braan domeinen dat zich bevindt in het basolaterale membraan van het RPE. Het ontstaan en het klinisch beloop van VMD zijn zeer variabel, het is onduidelijk of dit veroorzaakt wordt door het type mutatie en/of de lokatie van de mutatie in de sequentie.

In hoofdstuk zeven hebben we de fenotype-genotype correlatie in een groep van 53 patiënten met VMD beschreven. Alle geïncludeerde patiënten droegen een ziekte-veroorzakende mu- tatie in het Best1 gen. We hebben 40 VMD patiënten over een periode van gemiddeld 15,3 jaar gevolgd. Nagenoeg alle patiënten lieten een geleidelijke achteruitgang van de visus zien zonder volledig blind te worden. Aan de andere kant hebben we ook geobserveerd dat de uiting van de ziekte (het fenotype) zeer variabel was en in sommige gevallen sterk afhanke-

213 Chapter 11 lijk was van de mutatie in het Best1 gen (het genotype). Patiënten met een Ala10Val mutatie hadden een significant snellere progressie van de ziekte in vergelijking met patiënten met een Thr6Pro of een Tyr227Asn mutatie. Bovendien hebben we twee families beschreven die dezelfde mutatie droegen, maar zeer variabele fenotypes toonden. Tot slot lieten onze data zien dat er grote verschillen zijn in de klinische consequenties van de aanwezigheid van een ziekte-veroorzakende mutatie. We hebben een proband beschreven met een ziekteveroorza- kende Best1 mutatie en een uitgesproken VMD fenotype. Familieleden van deze proband hadden dezelfde mutatie maar geen enkel verschijnsel van de ziekte. Verder onderzoek naar de onderliggende mechanismen en onderzoek in grotere aantallen patiënten en families kan mogelijk aanwijzingen leveren voor toekomstige klinische zorg.

Hoofdstuk acht: Grootschalige mutatie analyse Het RPE en de fotoreceptoren brengen het overgrote deel van de retinale ziektegenen hoog tot expressie. Op dit moment zijn er ongeveer 150 retinale ziektegenen geïdentificeerd voor 30 verschilende retinale ziektes. De identificatie van een ziekteveroorzakende mutatie is cru- ciaal voor de bevestiging van de klinische diagnose, voor counseling van patiënten en voor toekomstige gentherapie. Helaas zijn er veel retinale ziektes waarvan het fenotype niet dui- delijk omschreven is, ook kunnen de fenotypes veel overlap vertonen met andere fenotypes. Bovendien vorm het grote aantal genen dat geassocieerd kan zijn met één ziekte een diagnos- tische uitdaging. De kosten verbonden aan het één voor één screenen van deze genen zijn zo hoog dat dit praktisch niet uitvoerbaar is. Een screening hulpmiddel dat zowel bekende als nieuwe mutaties in een groot aantal potentiële ziektegenen op een kosten-effectieve manier kan identificeren is daarom essentieel.

In hoofdstuk acht hebben we de ontwikkeling en validatie beschreven van een nieuwe 300 kb retinale resequencing chip, waarmee wij in staat waren om 87 verschillende retinale ziekte genen in één enkel experiment te screenen. We ontwikkelden en testten multiplex primers voor 1.445 amplicons corresponderend met de 87 ziektegenen op de chip. Ter validatie van de chip voerden we het volgende experiment in duplo uit: 87 exonen van 25 retinale ziek- tegenen waarin 87 bekende sequentie veranderingen aanwezig waren, werden simultaan ge- amplificeerd vanuit DNA monsters van 60 patiënten met een breed scala aan retinale aandoe- ningen. We vonden call rates voor succesvol gehybridiseerde amplicons van 98% tot 100%. De sensitiviteit van de chip voor puntmutaties was 99%, terwijl, zoals verwacht, deleties niet konden worden gedetecteerd. De specificiteit van de chip was 100%. Door het grote aantal genen op de chip, kunnen voor toekomstige diagnostische doeleinden meerdere patiënten met fenotypes gekenmerkt door mutaties in niet-overlappende genen, samen op één chip geanalyseerd worden, hierdoor wordt grootschalige gen screening kosten-effectief.

214 Nederlandse samenvatting

Referenties

1. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ et al (2009) Functio- nal annotation of the human retinal pigment epithelium transcriptome. BMC Genomics 10: 164. 2. Booij JC, Baas DC, Beisekeeva J, Gorgels TG, Bergen AA (2010) The dynamic nature of Bruch’s membrane. Prog Retin Eye Res 29: 1-18. 3. de Jong PT (2006) Age-related macular degeneration. N Engl J Med 355: 1474-1485. 4. Olson MD (1979) Development of Bruch’s membrane in the chick: an electron micro- scopic study. Invest Ophthalmol Vis Sci 18: 329-338. 5. Hirabayashi Y, Fujimori O, Shimizu S (2003) Bruch’s membrane of the brachymorphic mouse. Med Electron Microsc 36: 139-146. 6. Hewitt AT, Nakazawa K, Newsome DA (1989) Analysis of newly synthesized Bruch’s membrane proteoglycans. Invest Ophthalmol Vis Sci 30: 478-486. 7. Booij JC, Florijn RJ, ten Brink JB, Loves W, Meire F et al (2005) Identification of mu- tations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. J Med Genet 42: e67. 8. Retnet www.sph.uth.tmc.edu/Retnet/ 9. Boon CJ, Klevering BJ, Leroy BP, Hoyng CB, Keunen JE et al (2009) The spectrum of ocular phenotypes caused by mutations in the BEST1 gene. Prog Retin Eye Res 28: 187-205.

215

List of publications

12 Chapter 12

218 List of publications

RPE enriched gene expression and the origin of Bruch’s membrane and drusen. Booij JC, Bergen AAB manuscript in preparation

Simultaneous mutation detection in 90 retinal disease genes in multiple patients using a 300kb retinal resequencing chip. Booij JC, Bakker A, Kulumbetova J, Moutaoukil Y, Florijn RJ, Bergen AAB Opthalmology (2010), in press

A new strategy to identify and annotate human RPE-specific gene expression. Booij JC, ten Brink JB, Swagemakers SMA, Verkerk AJMH, Essing AHW, van der Spek PJ, Bergen AAB PlosOne 2010, 5(3) e9341

Course of visual decline in relation to the Best1 genotype in vitelliform macular dystro- phy. Booij JC, Boon CJF, van Schooneveld MJ, ten Brink JB, Bakker A, de Jong PTVM, Hoyng CB, Bergen AAB, Klaver CCW Opthalmology, 2010 april 8 epub ahead of print

The Dynamic Nature of Bruch’s Membrane. Booij JC, Baas DC, Beisekeeva J, Gorgels TGMF, Bergen AAB Progress in Retinal and Eye Research. 2010 Jan;29(1):1-18

Functional annotation of the human retinal pigment epithelium transcriptome. Booij JC, van Soest S, Swagemakers SM, Essing AH, Verkerk AJ, van der Spek PJ, Gorgels TG, Bergen AA. BioMed Central Genomics. 2009 Apr;10(1):164-182

Vitamine A therapie bij patiënten met retinitis pigmentosa. Booij JC, Klaver CCW Uitzicht, kwartaalblad van Retina Nederland. 2006; 23(3):22-24

Identification of mutations in the AIPL1, CRB1, GUCY2D, RPE65, and RPGRIP1 genes in patients with juvenile retinitis pigmentosa. Booij JC, Florijn RJ, ten Brink JB, Loves W, Meire F, van Schooneveld MJ, de Jong PT, Bergen AA. Journal of Medical Genetics. 2005 Nov;42(11):e67-e75

219

Dankwoord

13 Chapter 13

Ik wil graag iedereen bedanken die heeft bijgedragen aan het tot stand komen van dit proef- schrift. Allereerst bedankt aan alle patiënten die belangeloos meewerkten aan het onderzoek en zonder wie dit proefschrift nooit geschreven had kunnen worden.

Bedankt leden van de commissie, Prof. Aerts, Prof. Kamermans, Prof. Schlingemann en Dr. Klaver voor het het lezen van mijn proefschrift en het plaatsnemen in de commissie.

Bedankt aan de collega’s die inhoudelijk hebben bijgedragen, mijn promotor Arthur Bergen, mijn oud-collega’s van het NIN, Arne Bakker, Jacoline ten Brink, Frederike Dijk, Anke Es- sing, Ralph Florijn, Theo Gorgels, Joop van Heerikhuize, Paulus de Jong, Willem Kamphuis, Willem Loves, Youssef Moutaouakil, Cecile Ottenheim, Astrid Plomp, Mary van Schoone- veld, Simone van Soest, Inge Versteeg en Nataliya Yeremenko. Ook bedankt aan de mensen die vanuit andere instituten meehielpen, Yavyz Ariyurek, Michael van Es, Ruben van ‘t Slot. Bedankt ook Helene Thygesen, voor de vele analyses die je voor me uitvoerde en voor het meedenken.

Bedankt ook aan alle co-auteurs van mijn artikels, Camiel Boon, Carel Hoyng, Francoise Meire, Peter van der Spek, Sigrid Swagemakers en Annemieke Verkerk.

Helma de Bruin, bedankt voor de bank, hij staat er nog steeds! Bedankt Sylvia Tol, voor het ongeremde enthousiasme waarmee je aan ons project werkte. Marcia Tempelers, bedankt voor al je interesse en de gezellige momenten aan je receptie. Als ik mijn computerwerk even zat was, kon ik altijd bij jou terecht voor een praatje. Thanks also to Jamilia Kulumbetova and Juldyz Beisekeeva for helping me with my project. Aukje van Tol, bedankt voor de gezel- ligheid. Wendy Aartsen, bedankt voor de goede gesprekken en de gezellige etentjes. Wendy Hettema, bedankt voor de gezellige tijd en je hulp op het werk. Bedankt beide Wendys dat jullie ook na het NIN goede vriendinnen zijn gebleven.

Dominique Baas, bedankt voor je gezelligheid, je steun, je hulp en voor de vele milimeters die we van onze toetsenborden hebben gesleten met onze mails. Jij begreep hoe het was om oio te zijn. Bedankt dat je mijn paranymf wilde zijn.

Leden van de PartySquad: Nathalie Koning, Jinte Middeldorp, Kasper Roet, Alexander Hei- mel, JP Sommeijer, Sridhara Chakravarty, Frederike Dijk, Colleen Shields, bedankt voor de leuke tijd. Het NIN was niet alleen maar werk.

Nathalie Koning, dank je voor al je uitleg over Indesign, en het beantwoorden van mijn ein- deloze stroom vragen over de layout van mijn boekje.

Thanks Sridhara Chakravarthy, you’ve always known :)

Bedankt aan de patiëntvereniging Retina Nederland en in het bijzonder Yvonne van de Wiel voor het vele goede werk dat jullie doen. Ook bedankt dat ik op de familiedag een verhaal mocht vertellen, dat jullie mij de Retina Nederland prijs hebben uitgereikt en dat ik een hele

222 Dankwoord dag met jullie mee mocht door de Amsterdamse wateren, waar ik een kans kreeg om iets te horen van hoe mensen met een ernstige oogaandoening hier in het dagelijks leven mee omgaan. Bedankt ook aan de Nederlandse Vereniging van Blinden en Slechtzienden. Van de vele organisaties die bestaan voor slechtziende en blinde mensen in Nederland heb ik met deze twee te maken gehad en iets kunnen zien van het belangrijke werk dat jullie doen.

Bedankt, Gre Luyten voor de begeleiding toen ik student was en voor het mij interesseren in het onderzoek. Bedankt Caroline Klaver voor je begeleiding tijdens mijn oudste co-schap, voor je introductie bij het IOI, en voor de gezellige werketentjes bij jou thuis. Je volhardend- heid en enthousiasme heeft ervoor gezorgd dat we een mooi artikel konden schrijven.

Pap en Mam, bedankt dat jullie me het gevoel gaven dat jullie, of ik nou tijd had of niet, er altijd voor me waren. Bedankt Pap voor het kritisch lezen van het hele proefschrift, het moet een hele kluif zijn geweest. Je hebt zelfs uit de bijschriften van de figuren de fouten gehaald.

Martijn, jij begreep als geen ander hoe het is om een proefschrift te schrijven en wist daarom precies wanneer je me moest motiveren en wanneer je me moest afremmen. Bedankt voor al je geduld, je opbeurende woorden en je eindeloze vertrouwen in mij en mijn boekje. Jij gaf me de energie om mijn boekje af te maken, ik hoop dat ik voor jou hetzelfde mag doen. Dankjewel dat je mijn maatje bent en mijn paranymf wilde zijn. Vooral ook bedankt dat je op het IOI werkte en microarrays moest doen, zodat ik je kon ontmoeten. Ook jij hebt mijn boekje kritisch gelezen, dankjewel!

Bedankt ook aan de mensen die mijn boekje van voor naar achter hebben gelezen, jullie zijn zeldzaam.

223

Curriculum vitae

14 Chapter 14

226 Curriculum vitae

Curriculum Vitae Judith Booij werd geboren op 29 juli 1980 in Delft. Haar middelbare school opleiding volgde ze aan het Christelijk Lyceum in Delft, waarna ze in 1998 geneeskunde ging studeren aan de Erasmus Universiteit in Rotterdam. In 2004 slaagde ze voor haar artsexamen, waarna ze aan haar promotie onderzoek begon in het toenmalig Interuniversitair Oogheelkundig Instituut (IOI, momenteel Nederlands Instituut voor Neurowetenschappen). Momenteel werkt ze bij het Nederlands instituut voor onderzoek van de gezondheidszorg (NIVEL) in Utrecht.

227

Color figures

15 Chapter 15

Chapter 1, Figure 1. Schematic overview of the genes involved in the phototransduction cascade along with their location in the photoreceptors and RPE.

ROS

ABCA4 RHO RDH12

visual pigments all-trans retinal

IPM 11-cis retinal all-trans retinol

11-cis retinol all-trans retinyl esters

RLBP1 LRAT RPE RGR RPE65 RDH5

ROS: rod outer segment, IPM: interphotoreceptor matrix, RPE: retinal pigment epithelium cell. Circle and arrows indicate the visual cycle, letters in square boxes indicate proteins involved[9].

Chapter 1, Figure 2. Embryological development of the retina.

Reprinted with permission from Progress in Retinal and Eye Research, 14,1, Reichenbach A and Pritz-Hohmeier S., Normal and disturbed early development of the eye Anlagen, 1-45, Copyright (1995), with permission from Elsevier

230 Color figures

Chapter 1, Figure 3. Schematic overview of the functional properties of the RPE.

From: Strauss 2005, Physiol Rev. Physiological review, 85 (845-881) O. Strauss et al. Reprinted with permission from The American Physiological Society

Chapter 1, Figure 4. Schematic overview of genes involved in the visual cycle that, when mutated, are responsible for causing a retinal phenotype.

ROS Stargardt disease Macular degeneration RP RP CSNB LCA ABCA4 RHO RDH12

visual pigments all-trans retinal

IPM 11-cis retinal all-trans retinol

11-cis retinol all-trans retinyl esters

RLBP1 LRAT RPE RGR RPE65 RPA RDH5 LCA RP Fundus LCA albipunctatus

ROS: rod outer segment, IPM: interphotoreceptor matrix, RPE: retinal pigment epithelium cell. Circle and arrows indicate the visual cycle[9].

231 Chapter 15

Chapter 3, Figure 1. Overview of the (cell) layers surrounding the RPE.

Phot

RPE

Bruch's membrane

Choroid

Choroidal bloodvessel The dark brown rectangles connecting the RPE cells indicate tight junctions present between the cells. Phot: pho- toreceptors.

Chapter 4, Figure 1. Schematic drawing of proteins present in Bruch’s membrane (see main text) and the correspond- ing genes expressed in adjacent cells.

Phot

COL4A3 FNDC5 TIMP3 RPE [COL 4A1, 4A2, 4A3, 4A4, 4A5, 5] LAM HSPG bm RPE [COL1, 3] FN inner collagen elastin layer BM [COL 4, 6, 18] ELN TIMP3* [COL1, 3] FN outer collagen [COL 4A1, 4A2, 5] LAM HSPG bm choroid COL 1A1, 1A2, 3A1, 4A1, LAM A2, FN1, 4A2, 5A1, 5A2, 6A1, 6A2, A4, B1, FNDC3A 18A1 C1, C3 Chor

We identified three genes (COL4A3, FNDL5, TIMP3) with higher expression levels in the RPE than in the choroid, and 17 genes (remainder) with higher expression levels in the choroid than in the RPE. A qualitative impression of the gene expression is given. Gene expression levels were determined by RNA microarray study comparing gene expression levels from two adjacent tissue types from the same donor. Experiments were performed in triplicate (on three different healthy old human donor eyes)[24]. bm RPE: basement membrane of the RPE, bm choroid: basement membrane of the choroid. The abbreviation for basement membrane is not used in the accompanying text.

232 Color figures

Chapter 4, Figure 2. Schematic drawing showing the normal structure and functions of Bruch’s membrane.

biomolecules RPE RPE

BMP ICL x x

E

OCL

BMC

biomolecules Choroid

Transmission Electron Microscopic image of BM courtesy of Prof. J. Marshall. BM allows for the transport of bio- molecules across the membrane (double arrows), it attaches RPE cells to the membrane (filled circles) and it acts as a physical barrier to prevent the migration of RPE cells and choroidal cells across the membrane (arrow with X).

233 Chapter 15

Chapter 4, Figure 5. Overlap in protein content of Alzheimer plaques, AMD drusen, and atherosclerotic plaques. 7 (6%) shared proteins: Amyloid (beta, P), APOE, C3, CLU, FGG, VTN

Alzheimer plaque AMD drusen Atherosclerotic plaque a j

d f k e i b l c g h

24 (20%) shared proteins: 10 (8%) shared proteins: Alpha B-Crystallin, Amyloid (beta, Amyloid (beta, P), APCS, P), APOE, ATP5A1, complement Apolipoprotein (A1, E), comple- (C3, C7), CLU, FGG, HP, IGHG1, ment C3, CLU, FGG, PLG, VTN IGKC, OGN, PRDX1, Serpin (A1, A3) TIMP3, Tubulin (A1A, A1C, A3C, B2C, B3), UBB, VTN We analyzed 121 proteins based on the article by Crabb et al[6] and additional literature searches. We obtained NM numbers for 85 of the proteins characterized by Crabb, an additional 36 proteins were added based on the recent literature[167]. Note that drusen share 20 % of their protein content with AD plaques and less than 10 % with ath- erosclerotic plaques. Alzheimer plaque picture courtesy of Dr. I. Huitinga, Netherlands Brain Bank, donor number nhb:2006-060,VU:S06/189. Atheroslcerotic plaque picture courtesy of Dr. P Sampaio Gutierrez. a. Amyloid deposits, b. Cell nucleus, c. Enthorinal cortex, d. Retinal Pigment Epithelium, e. Bruch’s Membrane, f. Druse, g. Choroid, h. Choroidal bloodvessel, i. Intima with atheroslerotic plaque, j. Tunica media, k. Fibrous cap, l. Lipid + necrotic core.

234 Color figures

Chapter 4, Figure 6. Schematic drawing of proteins in drusen and the corresponding genes expressed in adjacent cell layers.

3 genes

Phot

13 genes RPE

113 proteins

BM

41 genes

Chor

36 proteins

We could annotate 113 genes with genbank codes (using Ingenuity) which correspond to drusen proteins identified by Crabb et al.[6]. Thirty six of these genes had higher expression levels in choroid compared to RPE (chor>RPE), thirteen genes had higher expression levels in RPE than in choroid (RPE>chor) and only three genes had higher expression levels in photoreceptors than RPE (phot>RPE). Gene expression levels were determined by RNA mi- croarray study comparing gene expression levels from two adjacent tissue types from the same donor. Experiments were performed in triplicate (on three different healthy older human donor eyes)[24]. Details of the 113 genes can be found in Table A.

235 Chapter 15

Chapter 5, Figure 1 Known BM proteins and their expression in the photoreceptors, the RPE and the choroid.

Phot

COL4A3 FNDC5 TIMP3

RPE bm RPE [COL 4A1, 4A2, 4A3, 4A4, 4A5, 5] LAM HSPG [COL1, 3] FN inner collagen [COL 4, 6, 18] ELN TIMP3* elastin layer BM [COL1, 3] FN outer collagen [COL 4A1, 4A2, 5] LAM HSPG bm choroid COL 1A1, 1A2, 3A1, 4A1, LAM A2, FN1, 4A2, 5A1, 5A2, 6A1, 6A2, A4, B1, FNDC3A 18A1 C1, C3 Chor

Chapter 5, Figure 2 Candidate proteins not identified in BM, but in terms of function or structure closely related to BM proteins, and their expression in the photoreceptors, the RPE and the choroid. COL25A1

Phot

COL8A1 COL20A1 RPE bm RPE [COL 4A1, 4A2, 4A3, 4A4, 4A5, 5] LAM HSPG [COL1, 3] FN inner collagen [COL 4, 6, 18] ELN TIMP3* elastin layer BM [COL1, 3] FN outer collagen [COL 4A1, 4A2, 5] LAM HSPG bm choroid COL 8A2, 10A1, 13A1, 15A1, 16A1, 21A1

Chor

236 Color figures

Chapter 5, Figure 3

CRYBA1 CRYGS

Phot

SERPINF RGR CTSD SMC6 RPE COL8A1 FRZB TIMP3

114 proteins

BM

COL1A2, COL6A1, COL6A2, ALDH1A1, ANXA1 HIST1H2AE CRYAB, FN1, HBA1, HBA2, IGHA1, ANXA2 HIST1H2BL IGKC, LMNA, LUM, MFAP4, OGN, ANXA5 HIST1H4H C3 PLA2G2A, PRELP, RNASE4, SAA1, HIST2H2AA3 FBLN5 SERPINA3, VIM

Chor

ALB, APOA1, APOA4, APOE, C3, CLU, FGG, FN1, HBA2, HBB, HP, LTF, LYZ, ORM1, PLG, PRDX1, S100A8, S100A9, SAA1, SERPINA1, SERPINA3, SERPINF1, VTN

237 Chapter 15

Chapter 7, Figure 4. Two pedigrees with the Thr6Pro mutation. a b

I:1 I:2 I:1 I:2 Thr6Pro

II:1 II:2 II:3 Thr6Pro Thr6Pro

II:1 Thr6Pro III:1 Thr6Pro

a) Presentation of the first pedigree. All three individuals (III:1, II:2, and II:3) carried the Thr6Pro mutation. At age 27, the proband (arrow; III:1) had bilateral visual complaints (visual acuity (VA) right eye (OD) 0.2, left eye (OS) 0.6; fundoscopically OD macular scar (stage 6), OS subtle changes (stage 1)) with an Arden ratio of 1.5 in both eyes. Her 55 year old mother (II:2) had no visual complaints (VA 1.0 both eyes (OU); no macular changes; Arden ratio 2.0). The maternal uncle (II:3) was diagnosed at age 51 and progressed to a severe bilateral phenotype (VA<0.01, stage 5) in 6 years. Fundoscopic pictures of individual III:1 at age 27 years, and of individual II:2 at age 55 years. b) Presentation of the second pedigree. Both individuals (II:1 and I:2) carried the Thr6Pro mutation. At age 31, the proband (arrow, II:1) had VA 0.8 and vitelliform macular dystrophy (VMD) stage 2 in the left eye; his right eye had VA 1.0 and no macular changes, Arden ratio’s OU were 1.0. See lower two black-and-white fundus photographs. At age 47 years (upper two fundus photographs), his left eye had progressed to VA 0.13 and VMD stage 6, while his right eye remained completely normal. The patients’ mother had no clinical signs of VMD apart from an Arden ratio of 1.5 at the age of 57 years (see fundus photographs). Squares indicate males, circles indicate females, filled symbols indicate affected individuals.

238 Color figures

Chapter 8, Figure 1. Analytical workflow.

Mutation Statistical analysis

Call rate Number of base 99% pairs screened

No hybridization because amplicon was not amplied

All hybridization results were imported into and analyzed in the JSI software. For each amplicon the program reports call rates and the number of changes identified. Both forward and reverse strands are visualized and statis- tics are given for the signal (A, C, T, or G) for each base position compared to the normalization set of previously analyzed chips

239