Nephrin-associated molecules in human glomerular diseases

Pekka Ihalmo

Department of Bacteriology and Immunology Haartman Institute Faculty of Medicine

Division of Biochemistry Department of Biological and Environmental Sciences Faculty of Biosciences

University of Helsinki Finland

ACADEMIC DISSERTATION

To be publicly discussed with the permission of the Faculty of Biosciences, University of Helsinki, in Small Hall, University main building, Fabianinkatu 33 (4th fl oor), Helsinki, on September 7th, 2007, at 12 noon.

Helsinki 2007 Supervised by Docent Harry Holthöfer, M.D., Ph.D. Department of Bacteriology and Immunology University of Helsinki Finland

Reviewed by Professor Markku Laakso, M.D., Ph.D. Department of Medicine University of Kuopio Finland

and

Docent Eero Lehtonen, M.D., Ph.D. Department of Pathology University of Helsinki Finland

Offi cial opponent Docent Marjo Kestilä, Ph.D. Department of Molecular Medicine National Public Health Institute Helsinki

ISBN 978-952-92-2472-2 (paperback) ISBN 978-952-10-4089-4 (PDF) http://ethesis.helsinki.fi

Helsinki 2007

Contents

LIST OF ORIGINAL PUBLICATIONS ...... 6

ABBREVIATIONS ...... 7

1. ABSTRACT ...... 9

2. REVIEW OF THE LITERATURE...... 10 2.1. Glomerular capillary wall as the fi ltration barrier in the glomerulus ...... 10 2.2. The slit diaphragm is an adhesion and signalling complex in the ...... 12

Nephrin...... 13

Nephrin-like NEPH1, NEPH2 and fi ltrin ...... 14

Cadherin superfamily proteins...... 15

Intracellular scaffolding and signalling proteins...... 16

2.3. Development of the slit diaphragm from a tight junction . . . . 17

2.4. Proteinuria is a hallmark of glomerular injury...... 18

Minimal change disease ...... 18

Focal segmental glomerulosclerosis...... 19

Membranous glomerulopathy ...... 19

Hypertensive nephropathy ...... 20

Diabetic nephropathy ...... 20

2.5. Genetic basis of proteinuria ...... 21

3. AIMS OF THE STUDY ...... 25

4. MATERIALS AND METHODS ...... 26

4.1. Study subjects and clinical samples ...... 26

4.2. Animals...... 28

4.3. Cell culture ...... 28

4.4. Molecular cloning and mRNA expression analysis ...... 28

Complementary DNA clones...... 28

RNA isolation ...... 28

Reverse transcriptase-polymerase chain reaction (RT-PCR) . . . . 29

Northern Blotting ...... 30

4 RNA dot blot assay ...... 31

4.5. Immunochemical methods ...... 31

Antibodies ...... 31

Immunoblotting ...... 32

Immunofl uorescence microscopy ...... 32

Immunoelectron microscopy ...... 32

4.6. Genotyping ...... 33

4.7. Databases, bioinformatics tools, and statistical methods . . . . . 33

5. RESULTS ...... 35 5.1. Filtrin, a novel nephrin-like , is expressed in glomerular podocytes (Studies I and III) ...... 35 5.2. Nphs1as mRNA is transcribed from the nephrin locus in a reverse orientation in the brain and lymphoid tissue (Study II) ...... 37 5.3. Glomerular expression of fi ltrin in patients with proteinuric kidney diseases (Study III) ...... 40 5.4. Genetic association analysis of LRRC7, KIRREL, NPHS2, KIRREL2, and ACTN4 with diabetic nephropathy (Study IV) . . . . 43

6. DISCUSSION ...... 46

Filtrin emerges into the picture of the podocytes ...... 46

Function of fi ltrin in other tissues...... 47

Expression of nephrin and novel regulatory aspects ...... 47 Chromosomal organisation of KIRREL2 and NPHS1 in the ...... 49

Patients with proteinuria (Studies III and IV) ...... 50 Is fi ltrin merely a by-stander in the or is it involved in podocyte injury? ...... 50 Can the genes behind monogenic proteinuric diseases contribute to the understanding of diabetic nephropathy? ...... 52

Concluding remarks ...... 55

7. ACKNOWLEDGEMENTS...... 57

8. APPENDIX...... 59 Members of the European Renal cDNA Bank at the time of the study...... 59

9. REFERENCES ...... 60

ORIGINAL PUBLICATIONS ...... 77

5 LIST OF ORIGINAL PUBLICATIONS

This thesis is based on the following original articles, which are referred to by their Roman numerals in the text.

I Ihalmo, P., Palmén, T., Ahola, H., Valtonen, E., Holthöfer, H. Filtrin is a novel member of nephrin-like proteins. Bio- chemical Biophysical Research Communications. 2003, 300:364-70. II Ihalmo, P., Rinta-Valkama, J., Mai, P., Åstrom, E., Palmén, T., Pham, T.T., Floss, T., and Holthöfer, H. Molecular cloning and characterization of an endogenous antisense transcript of Nphs1. Genomics. 2004, 83:1134-40. III Ihalmo, P., Schmid, H., Rastaldi, M.P., Mattinzoli, D., Lang- ham, R.G., Luimula, P., Kilpikari, R., Lassila, M., Gilbert, R.E., Kerjaschki, D., Kretzler, M., and Holthöfer, H. Expression of fi ltrin in human glomerular diseases. Nephrology Dialysis Transplantation. 2007, 22:1903-9. IV Ihalmo, P., Wessman, M., Kaunisto, M., Kilpikari, R., Parkko- nen, M., Forsblom, C., Holthöfer, H., and Groop, P.-H. for the FinnDiane Study Group. Association analysis of podocyte slit diaphragm genes as candidates for diabetic nephropathy (submitted).

The original articles have been reprinted with the permission of the pub- lishers.

6 ABBREVIATIONS

ACTN4 α-actinin-4 AER albumin excretion rate AMV avian myeloblastosis virus ANOVA analysis of variance BLAST basic local alignment search tool CD2AP CD2-associated protein CDH3 cadherin 3 (P-cadherin) CDH5 cadherin 5 (VE-cadherin) cDNA complementrary DNA CEPH Centre d’Etude du Polymorphisme Humain CI confi dence interval CNF congenital nephrotic syndrome of the Finnish type COL4A3, -A4, -A5 α3-, α4-, α5(IV)collagen CON control group CpG CpG dinucleotide dCTP 2’-deoxycytidine 5’-triphosphate DN diabetic nephropathy Ena/VASP Enabled/vasodilator-stimulated phosphoprotein ERCB European Renal cDNA Bank ESRD end-stage renal disease EST expressed sequence tag FAT FAT tumor suppressor 1 (FAT1) FSGS focal segmental glomerulosclerosis GAPDH glyseraldehyde-3-phosphate dehydrogenase GBM glomerular basement membrane

HbA1c glycosylated hemoglobin HT hypertensive nephropathy HUGO Human Genome Organisation HWE Hardy-Weinberg equilibrium KIRREL kin of IRRE like (NEPH1) KIRREL2 kin of IRRE like 2 (fi ltrin or NEPH3) KIRREL3 kin of IRRE like 3 (NEPH2) LAMB2 laminin β2 LD linkage disequilibrium LMX1B LIM homeobox transcription factor 1β LRRC7 leucine rich repeat containing 7 (densin-180) Macro macroalbuminuria MAF minor allele frequency MCD minimal change disease

7 MGN membranous glomerulopathy M-MLV moloney murine leukemia virus MYH9 myosin, heavy chain 9, non-muscle NCBI National Center for Biotechnology Information NEPH1-3 nephrin-like protein 1-3 Normo normoalbuminuria NPHS1 nephrosis 1 (nephrin) Nphs1as nephrosis 1 antisense NPHS2 nephrosis 2 () OR odds ratio PDZ PSD-95/disc large/ZO-1 PLCE1 phospholipase C ε1 RACE rapid amplifi cation of cDNA ends SH2 Src homology 2 SNP single nucleotide polymorphism TJP1 tight junction protein 1 (ZO-1) TRPC6 transient receptor potential cation channel, subfamily C, member 6 US urinary space WT1 Wilm’s tumor 1 ZO-1 zonula occludens 1

8 1. ABSTRACT

The glomerular epithelial cells and their intercellular junctions, termed slit diaphragms, are essential components of the fi ltration barrier in the kidney glomerulus. Nephrin is a transmembrane adhesion protein of the slit diaphragm and a signalling molecule regulating podocyte physiol- ogy. In congenital nephrotic syndrome of the Finnish type, mutation of nephrin leads to disruption of the permeability barrier and leakage of plasma proteins into the urine. This doctoral thesis hypothesises that novel nephrin-associated molecules are involved in the function of the fi ltration barrier in health and disease. Bioinformatics tools were utilized to identify nephrin-like molecules in genomic databases, and their distribution in the kidney and other tis- sues was investigated. Filtrin, a novel nephrin homologue, is expressed in the glomerular podocytes and, according to immunoelectron micros- copy, localizes at the slit diaphragm. Interestingly, the nephrin and fi ltrin genes, NPHS1 and KIRREL2, locate in a head-to-head orientation on chro- mosome 19q13.12. Another nephrin-like molecule, Nphs1as was cloned in the mouse, however, no expression was detected in the kidney but instead in the brain and lymphoid tissue. Notably, Nphs1as is transcribed from the nephrin locus in an antisense orientation. The glomerular mRNA and protein levels of fi ltrin were measured in kidney biopsies of patients with proteinuric diseases, and marked reduction of fi ltrin mRNA levels was detected in the proteinuric samples as compared to controls. In ad- dition, altered distribution of fi ltrin in injured glomeruli was observed, with the most prominent decrease of the expression in focal segmental glomerulosclerosis. The role of the slit diaphragm-associated genes for the development of diabetic nephropathy was investigated by analys- ing single nucleotide polymorphisms. The genes encoding fi ltrin, densin- 180, NEPH1, podocin, and α-actinin-4 were analysed, and polymorphisms at the α-actinin-4 gene were associated with diabetic nephropathy in a gender-dependent manner. Filtrin is a novel podocyte-expressed protein with localization at the slit diaphragm, and the downregulation of fi ltrin seems to be characteristic for human proteinuric diseases. In the context of the crucial role of nephrin for the glomerular fi lter, fi ltrin appears to be a potential candidate mol- ecule for proteinuria. Although not expressed in the kidney, the nephrin antisense Nphs1as may regulate the expression of nephrin in extrarenal tissues. The genetic association analysis suggested that the α-actinin-4 gene, encoding an actin-fi lament cross-linking protein of the podocytes, may contribute to susceptibility for diabetic nephropathy.

1 Abstract 9 2. REVIEW OF THE LITERATURE

2.1. Glomerular capillary wall as the fi ltration barrier in the kidney glomerulus

The kidney is a vital organ responsible for secreting metabolic end-prod- ucts into the urine. In addition, the kidney maintains body homeostasis by regulating the fl uid, acid-base and electrolyte balance. The nephron is the basic functional unit in the kidney, and primary urine is produced in the fi rst part of the nephron, the glomerulus, which is a capillary tuft surrounded by a urinary space and the Bowman’s capsule [1]. The mechanism of glomerular fi ltration has remained an enigma of nephrology for decades. Notably, about 180 liters of primary urine is fi l- tered through the glomerular capillary wall every day, which is essential for the effi cient excretion of metabolic end-products from the circula- tion. At the same time, this fi lter can retain albumin and other crucial plasma proteins in the circulation, but the fi lter itself does not clog in spite of the enormous load of 15 kg of protein that the kidneys handle every day [1]. The prevailing view is that the composite effect of all layers of the glomerular capillary wall contributes to the charge- and size selec- tive properties of the fi ltration barrier (Figure 1). Blood enters the capillary tuft through the afferent arteriole, which branches into smaller capillaries [1]. The high capillary pressure (55 mmHg) forces water and small molecules through the capillary wall. It has been es- timated that approximately 20% of the plasma volume is fi ltered through the capillary wall, the rest exits the glomerulus via the efferent arteriole [1]. The capillary endothelial cells are thin cells lining the luminal side of the glomerular basement membrane (GBM) (Figure 1) and they represent the initial barrier to the blood constituents. These cells have a porous structure with round to oval fenestrae of 70-100 nm in diameter [1]. The fenestrae and the whole luminal side of the cell are covered by a special glycocalyx with a negative charge due to sialic acid containing glycoproteins and lip- ids [2, 3]. The negative charge of the endothelial cells may account for the charge selectivity of the permeability barrier [4]. The mesangial cells support the glomerular structure on the endocap- illary side [5]. Similar to the smooth muscle cells, the mesangial cells con- tract in response to vasoactive agents and have thus been proposed to regulate glomerular fi ltration. These cells also have phagocytotic proper- ties and secrete bioactive substances such as prostaglandins [6, 7].

10 2 Review of the literature Capillary lumen

Urinary space

Podocyte

Basement membrane

Endothelium

Mesangial cells

Figure 1. Schematic structure of the glomerulus. Primary urine is produced in the glomerulus, which is a capillary tuft surrounded by a urinary space and the Bowman’s capsule. Modifi ed from [107].

The GBM separates the endothelial cells from the podocytes (Figure 1). This membrane is a meshwork of glycoproteins (type IV collagen [α3α4α5], laminin-11 [α5β2γ1], and nidogen/entactin) and proteoglycans (agrin and perlecan) with a high water content (reviewed in [8, 9]). The GBM car- ries an anionic charge due to negatively charged glycosaminoglycan side chains of perlecan and agrin. Podocytes (Greek ‘pod’, foot) cover the GBM from the urinary side with cellular projections known as foot processes (Figure 1; reviewed in [10]). The podocytes are terminally differentiated epithelial cells with an arborized structure similar to neurons. These cells are highly polarized with their apical side and the cell body facing the urinary side of the Bowman’s space, and foot processes are anchored to the GBM on the basal side in an integrin- and dystroglycan-dependent fashion [11-13]. Bundles of actin fi laments on the apical pole and the cortical actin net- work support the structure of the foot processes [14]. The interdigitating foot processes are separated by a slit of 30-40 nm and bridged by a unique cell junction termed the ‘slit diaphragm’ [15]. In the pioneering work by Rodewald and Karnovsky, the slit diaphragm was described as a zipper-like structure with a pore size corresponding to the dimensions of albumin [15]. Recent electron-tomographic data ac-

2 Review of the literature 11 cords with this view, revealing winding strands that form connections at the midline of the slit [16]. Conceptually, the slit diaphragm represents the border between the basal and apical domains of the podocyte [17, 18] and has been proposed to share properties of adherens and tight junctions [17, 19]. In the 1960s, tracer studies were launched to decipher the molecular barrier of the glomerular capillary wall. Experiments using ferritin and dextrans of different molecular weights as molecular markers indicated that the molecules of the size of albumin or larger cannot penetrate the GBM (reviewed in [20]). However, this approach neglected the infl uence of charge of the molecular components of the barrier on fi ltration. More conclusive results were obtained by using ferritins with anionic, neutral and cationic charge, and these experiments suggested that the GBM and endothelial cells prevent the passage of anionic molecules, whereas the fi ltration slit diaphragm appeared to retard the passage of molecules with positive charge [21]. However, this issue has remained a matter of debate. A gel permeation/diffusion hypothesis suggests that GBM is the size-selective barrier and the slit diaphragm only provides resistance to water fl ow [22], but completely contrasting views have also been pro- posed [23, 24]. A set of discoveries have brought the slit diaphragm into focus. In 1988, it was demonstrated that an injection of the monoclonal antibody 5-1-6, directed against the slit diaphragm, induced proteinuria in rats [25]. Later, it was ascertained that the antibody reacted with the slit di- aphragm protein nephrin [26]. The discovery of the causative nephrin gene (NPHS1) for congenital nephrotic syndrome of the Finnish type in 1998 [27], and the elucidation of the genetic defects in other inherited nephropathies (reviewed in chapter 2.5) have launched a new era in ne- phrology.

2.2. The slit diaphragm is an adhesion and signalling complex in the podocytes

The slit diaphragm has been described as a ‘modifi ed adherens junction’ by Reiser and co-workers [19]. This cell-cell contact shares characteristics of the adherens junction (expression of P-cadherin [19]) and tight junc- tion (zonula occludens-1 [17]), and the expression of slit diaphragm-spe- cifi c proteins refl ects its unique properties (e.g. podocin [28]). The mo- lecular composition of the slit diaphragm is reviewed in this chapter with special reference to the functions attributed to these proteins. Figure 2 summarizes the principal components of the slit diaphragm and the as- sociated cytoplasmic adhesion complex. The slit diaphragm can be considered an epithelial cell adhesion junc- tion and a fence between the apical and basal domains of the podocyte

12 2 Review of the literature Cadherin Catenins superfamily ZO-1 CD2AP Immunoglobulin superfamily Podocin

Podocyte

GBM Endothelium

Figure 2. The slit diaphragm is a unique cell-cell junction between the podocyte foot processes. Proteins of the immunoglobulin and cadherin superfamilies comprise the extracellular adhesion complex, while several adaptor and scaffolding proteins are located on the intracellular side. GBM, glomerular basement membrane. Modifi ed from [61].

foot processes [17, 18]. Tight junctions have usually been acknowledged for their function as a selective fi lter, but as illustrated by the permeabil- ity assays (reviewed in chapter 2.1) and the multitude of slit diaphragm encoding genes behind hereditary proteinuria (chapter 2.5), the slit dia- phragm also seems to serve a fi ltering function. Recently, the functions ascribed to the slit diaphragm have extended from a mere structural component to the regulator of transcriptional activity, apoptosis, and podocyte cytoskeleton (reviewed in [29-31]).

Nephrin Nephrin (the offi cial gene name by HUGO nomenclature NPHS1) is an integral adhesion molecule and a structural component of the slit dia- phragm [32-34]. This protein belongs to the immunoglobulin superfamily with eight extracellular immunoglobulin-like domains followed by a fi bronectin type III domain, a single-span transmembrane region, and a C-terminal intracellular domain. A hypothetical model of the homophilic interaction of nephrin molecules was introduced already in the fi rst re- ports of nephrin [34], suggesting that nephrin molecules from adjacent foot processes would interact via their extracellular immunoglobulin-like domains and thus contribute to the fi lamentous diaphragm observed be- tween foot processes. Subsequent experimental studies have provided support for this view [35-37].

2 Review of the literature 13 Nephrin resides in lipid rafts on the plasma membrane, and tyrosine residues of nephrin are phosphorylated upon foot process effacement [38, 39]. Furthermore, nephrin associates with and is phosphorylated by Src family kinase Fyn which stimulates intracellular signalling [40, 41]. Emphasizing the role of nephrin as a signaling protein, the intracellular tyrosine phosphorylated residues serve as potential docking sites for p85, the regulatory subunit of phosphoinositide 3-OH kinase [42], and for the adaptor protein Nck [39, 43], and these interactions appear to regulate the actin cytoskeleton and apoptosis in podocytes [39, 42, 43]. In addi- tion, phosphorylation of nephrin seems to regulate gene expression in the podocytes, and this is facilitated by the interaction with another slit diaphragm protein, podocin [41, 44]. Similar to other plasma membrane receptors, nephrin is endocytosed by β-arrestin in a phosphorylation de- pendent manner [45]. Regulation of nephrin in acquired proteinuric diseases has been under intensive study, and altered protein and mRNA levels are found both in experimental animal models and human diseases (e.g. [46-48]). In par- ticular, downregulation of nephrin in diabetic nephropathy is well es- tablished and, at least partially, mediated by angiotensin II and glycated albumin [49-51]. However, repression of nephrin expression might also be partially explained by the reduced number of podocytes per glomeru- lus [52].

Nephrin-like proteins NEPH1, NEPH2 and fi ltrin The cloning of NEPH1 (KIRREL) in 2001 led to the establishment of a pro- tein family including two other closely related proteins, NEPH2 (KIRREL3) and fi ltrin (KIRREL2) [53-56]. Filtrin is also known as NEPH3. These pro- teins share a common domain architecture with fi ve extracellular immu- noglobulin-like domains and a short intracellular domain of around 200 amino acids [57]. Within podocytes, NEPH1 and NEPH2 localize at the podocyte slit diaphragm [37, 54, 58] (Figure 2). In NEPH1-defi cient mice, the podocyte foot processes are effaced along the GBM, and the mice exhibit severe proteinuria [53]. Similarly to nephrin-defi cient mice, the pups die during the early postnatal period [53]. Soon after the cloning of nephrin-like proteins, the hypothesis of their potential interaction with nephrin was presented. Indeed, the extracel- lular domains of NEPH1 and NEPH2 appear to interact with nephrin, and the interactions are mediated by several extracellular immunoglobulin- like domains [35, 37, 54, 58]. NEPH2 has also been reported to form ho- modimers [54], whereas the data concerning the possible homodimerisa- tion of NEPH1 remains controversial [35, 37, 58]. Similarly to nephrin, NEPH1 resides in lipid rafts [37, 54]. The extracellular domain of NEPH2 is shedded from the plasma membrane by matrix metalloproteinases [54, 59], and seems to leak into the urine of both healthy individuals and proteinuric patients.

14 2 Review of the literature The physiological importance of NEPH1-nephrin interaction has been clarifi ed in vivo by injecting a combination of anti-NEPH1 and anti- nephrin antibodies into rats. The antibody mixture induced a transient pro teinuria in mice, whereas injection of anti-NEPH1 or anti-nephrin an- tibodies alone did not cause any changes. Even though the podocyte foot processes and slit diaphragms remained intact during this experi- ment, the expression of ZO-1, an intracellular binding partner of NEPH1, was considerably reduced [58]. In their intracellular part, NEPH1, NEPH2 and fi ltrin share a motif of nine conserved amino acids surrounding an SH2-binding site of growth factor receptor-bound protein 2 (Grb2) [57]. Sellin and colleagues hypoth- esized that this domain of nine amino acids could mediate interaction between the NEPH1 and podocin. They ascertained that both a single point mutation in the Grb2-binding site and tyrosine phosphorylation of this motif by Src kinases inhibited the interaction between the C-terminal domain of NEPH1 and podocin [57]. NEPH1 interaction via its intracellu- lar domain with podocin was also verifi ed by immunoprecipitation [57]. Furthermore, co-expression of NEPH1 and the IL2-inducible T-cell kinase resulted in a marked activation of the transcription factor AP-1, arguing for a role of NEPH1 as a signalling molecule [57]. Since the C-terminal end of nephrin-like proteins contains a putative PDZ-binding motif, the interaction between them and a PDZ-domain- bearing protein ZO-1 has been assessed [60]. The interaction occurs both in vitro and in vivo and has been mapped to the fi rst PDZ domain of ZO-1 and the last three amino acids of the intracellular domain of NEPH1. This conclusion accords with the data presented by Liu and colleagues [58]. The interaction substantially increased signal transduction as measured by AP-1 activity and induced tyrosine phosphorylation of NEPH1 [60].

Cadherin superfamily proteins To date, only a few proteins in addition to nephrin and nephrin-like pro- teins have been localized to the extracellular adhesion junction of the slit diaphragm (Figure 2). These proteins include P-cadherin (CDH3) [19] and a recently reported VE-cadherin (CDH5) [61], originally attributed to its function in the vascular endothelium [62]. In a podocyte cell line, P-cadherin also recruites several catenins to the intracellular side [19]. Even though reduced levels of P-cadherin have been reported in dia- betic nephropathy [63], the signifi cance of P-cadherin for maintaining the glomerular fi lter remains elusive, since the respective P-cadherin-de- fi cient mice show no kidney dysfunction [64]. In addition to classical cadherins, a large protocadherin FAT1 (FAT) with 34 extracellular cadherin repeats seems to be an essential component of the slit diaphragm, as refl ected by the lethal phenotype of Fat1-defi cient mice [65, 66]. These mice suffer from perinatal lethality and their foot processes are fused according to an electron microscopic examination

2 Review of the literature 15 [66]. FAT1 is acknowledged for its function as a regulator of actin polym- erization (via its interaction with Ena/VASP proteins) and, in addition, in the establishment of cell polarity [67].

Intracellular scaffolding and signalling proteins A number of peripheral membrane proteins locate on the intracellular side of the slit diaphragm, where they link the immunoglobulin super- family members and cadherins to the cortical actin cytoskeleton [68, 69]. Before nephrin, ZO-1 (TJP1) was the only identifi ed slit diaphragm-asso- ciated protein in the podocytes [17]. It is actually a typical tight junction protein, and the ZO-1 isoform without the motif-alpha is expressed in the slit diaphragms [70]. ZO-1 is localized on the intracellular side of the lateral foot process membranes, where it binds the nephrin-like proteins via its PDZ-domain [60]. Nephrin also provides a nexus for several binding partners on the intracellular side (Figure 2). Proteins including CD2-associated protein (CD2AP) [71, 72] and podocin [73] are located on the lateral foot process membrane and they are crucial for the proper fi ltration function. Podocin (NPHS2) is a member of the prohibitin-domain proteins and expressed al- most exclusively in the podocytes [28, 73]. Podocin forms homo-oligomers and is a constituent of lipid rafts [28, 74], where it binds cholesterol via its prohibitin-domain [75]. Podocin also functionally interacts with the slit diaphragm protein TRPC6 [76, 77], a non-selective cation channel, and regulates the activity of this channel in a cholesterol dependent fashion [75]. It has been speculated that podocin and TRPC6 could act as sensors for the intraglomerular pressure and/or glomerular fi ltration in the kid- ney [75]. Podocin associates via its C-terminal domain with nephrin and CD2AP [28, 44] and escorts nephrin to the lipid rafts [74]. CD2AP is an adaptor protein, originally recognized for its function in the immunological synapse [78]. CD2AP is an actin-associated protein in the podocytes [79, 80], but immunoelectron microscopic data and the reported interaction with nephrin suggest localization also at the slit diaphragm [72, 81]. In the close vicinity of the slit diaphragm resides an actin-bundling protein, α-actinin-4 (ACTN4) [82, 83]. Transgenic animal models of Actn4 have pinpointed a role for this molecule in cell motility and actin cytoskeleton regulation [84, 85] and also in the adhesion of po- docytes to the GBM [86]. Densin-180 (LRRC7) is a founding member of a protein family containing several leucine-rich repeats and PDZ domains, and it had been originally identifi ed from the postsynaptic densities in the brain [87]. Recently, it has been reported as a novel nephrin-asso- ciated protein at the slit diaphragm [88] and as a potential interaction partner for α-actinin-4 [89].

16 2 Review of the literature 2.3. Development of the slit diaphragm from a tight junction

Development of the metanephric mammalian kidney is a process that in- volves reciprocal interaction of two tissue components, the mesenchymal blastema and the ureteric bud, originally the mesoderm-derived Wolf- fi an duct [90]. The glomerular development involves a series of events that initiates with the condensation of metanephric mesenchyme cells next to the tip of the ureteric bud [90]. These cells undergo mesenchyme- to-epithelium transition and form an epithelial vesicle, which gives rise to all the segments of the nephron, including the glomerulus and the proximal and distal tubules [90]. Glomerular anlagen are fi rst visible at the S-shape body stage, which has developed from the vesicle [91] (Figure 3). At this stage, the precursor cells of the endothelium and mesangium invade the proximal part of the S-shaped body, and the entity eventually forms a vascularized glomerulus [91, 92]. Visceral epithelial cells lining the presumptive Bowman’s space rep- resent the future podocytes (Figure 3). At the early S-shape body stage, the cells are joined by ZO-1-positive apical tight junctions [17] and are fi rmly attached to the nascent basement membrane, a fence between the epithelial and capillary compartments [91]. During podocyte differ- entiation, the junctional complexes between the cell membranes migrate down towards the basal side. Eventually, a continuous intercellular space becomes visible between the cells, and ladder-like junctions are occasion- ally seen between the cells [91]. The layer of the future podocytes forms a cup around the developing capillary loops, and the formation and in- terdigitation of the foot processes begins (Figure 3). At the capillary loop stage, nephrin and ZO-1 localize at the cell-cell contacts between the developing podocytes [17, 93]. The foot processes surround the capillar- ies, and the tight junctions are gradually replaced with the mature slit diaphragms (Figure 3). [91]

S-shaped body Mature glomerulus

Figure 3. Development of the slit diaphragm. During the development the apical junctional complexes migrate towards the basal slit and fi nally mature into slit diaphragms. Drawn after [235].

2 Review of the literature 17 In experimental nephrosis and human proteinuric diseases, foot proc- ess effacement is often observed [94, 95]. It has been postulated that the loss of the foot process architecture and appearance of tight junctions or ladder-like structures replacing the slit diaphragms corresponds to a re- verse of the podocyte into more primitive developmental stages [91]. For example, in the puromycin aminonucleoside nephrosis of rats, tight junc- tions or ladder-like structures reappear in the intercellular space [94], and this condition may be reversible with the restoration of the foot process architecture. The elucidation of the morphological basis of proteinuria has lead to the view that podocytes are involved in the progression of proteinuric diseases (reviewed in [96]).

2.4. Proteinuria is a hallmark of glomerular injury

Minimal change disease (MCD), focal segmental glomerulosclerosis (FSGS), and membranous glomerulopathy (MGN) are primary glomerular diseases manifesting as nephrotic syndrome [97]. These diseases are clini- cally characterized by heavy proteinuria (i.e. leakage of high molecular weight proteins into the blood fi ltrate and urine), hypoalbuminemia, hyperlipidemia, and oedema [97]. Diabetic nephropathy and hyperten- sive nephropathy are secondary disorders that affect the glomerulus and present a more insidious pathogenetic process. The clinical sequelae of proteinuric diseases are well-described, but the pathogenetic mecha- nisms at the molecular level have remained elusive. The main question is, why the glomerulus becomes leaky under pathological conditions. It has become evident that a common denominator for these diseases is podocyte injury (reviewed in [98]).

Minimal change disease Minimal change disease is characterized by the nephrotic syndrome without glomerular lesions observed in light microscopy (fi rst described by Munk in 1916 [99]). However, in electron microscopic examination, effacement of the podocyte foot processes is frequently encountered, while the structure of GBM seems unaltered [95]. This suggests that po- docyte injury is involved in the pathogenesis of MCD [95]. MCD is the most common primary nephrotic syndrome in children, and it accounts for about 20% of primary glomerular disease in adults [100]. MCD can often be classifi ed as ‘primary’, but secondary forms also exist in association with tumours, toxic or allergic reactions, infections, and autoimmune diseases [97]. Immune-complex deposition or comple- ment activation in the glomerulus of the patients with MCD has not been reported. Instead, it has been postulated that MCD is a disorder of T cell function, where abnormal secretion of cytokines affecting the glomeru- lar capillary permeability may be involved [101, 102]. MCD responds ef-

18 2 Review of the literature fectively to treatment with corticosteroids (prednisone) in children; in adults, steroid resistance and relapses are, on the other hand, common [103]. Gradual transition from MCD into focal segmental glomeruloscle- rosis has been reported making these diseases both diagnostically and therapeutically challenging [104]. The expression ratio of podocin to synaptopodin, two podocyte-expressed genes, has recently been sug- gested for disease stratifi cation [47].

Focal segmental glomerulosclerosis The name ‘focal segmental glomerulosclerosis’ describes the lesions observed in the kidney in this disease: Scarring involves only some of the glomeruli (focal), and in the affected glomeruli, only a part of the glomerulus appears sclerotic (segmental) (fi rst description published in [105]). Several variants of FSGS (e.g. collapsing, glomerular tip, or cellular variant) can be separated in histological examination [97]. Podocyte in- jury and dedifferentiation seems to be a primary cause of the disease [98, 106], and formation of scars may be a sign of a reparative process after podocyte injury [107]. FSGS is the cause of primary glomerular disease in about 15% of adults patients, but is less common in children [100]. FSGS can be considered either primary or secondary if associated with HIV, toxic compounds or drugs [97]. In some patients with primary FSGS, it has been speculated that a circulating factor underlies the glomerular permeability changes [108]. Interestingly, several familial forms of FSGS have been discovered with mutations in the podocyte-expressed genes (reviewed in chapter 2.5). FSGS responds poorly to corticosteroid treat- ment, and often progresses to chronic renal failure [103]. Anti-prolifera- tive agents (e.g. cyclophosphamide) may be effective for the treatment of the steroid-resistant cases [103]. Notably, the prognosis of patients car- rying mutations in genes linked to FSGS is poor.

Membranous glomerulopathy Membranous glomerulopathy is characterized by formation of subepi- thelial immune deposits of IgG and a diffusively thickened GBM [97]. Po- docyte foot process effacement is observed at the ultrastructural level [97, 109]. Even though MGN progresses slowly, it is a common cause of primary glomerular disease accounting for about 40% of cases in adults, but it is less common in children [100]. The formation of immune complexes against the podocyte-membrane proteins, with yet unidentifi ed antigens, seems to trigger the membrane thickening (reviewed in [109]). In rare cases of neonatal membranous ne- phropathy, neutral endopeptidase in the podocyte membrane has been identifi ed as a target for circulating antibodies [110]. In Heymann nephri- tis, a rat model of MGN, the antigen is the podocyte membrane protein megalin, which is a member of the LDL-receptor subfamily [111, 112]. Mechanistically, the immune complexes lead to complement activation

2 Review of the literature 19 and generation of C5b-9 membrane-attack complexes on the podocyte membranes, and formation of oxygen radicals by the injured podocytes seems to promote the glomerular damage [110, 113, 114]. Corticoster- oids alone have been reported to be ineffective in the therapy [103]; The combination of anti-proliferative agents with corticosteroids is consid- ered a better approach [103].

Hypertensive nephropathy Hypertensive nephropathy is histologically characterized by the accumu- lation of hyaline in the walls of arteries and arterioles, collapse of the glomerular tuft, and glomerulosclerosis [115]. It has been estimated that hypertensive male patients suffer from manifold higher relative risk of developing renal failure than subjects with optimal blood pressure [116]. Malignant hypertension is recognized as a distinct cause of renal failure, while the link between essential hypertension and nephropathy is less well determined [117]. The podocytes are sensitive to injury, and animal models of hyper- tension suggest that podocyte failure in hypertension may promote the formation of glomerulosclerosis [118, 119]. It has been postulated that mechanical strain due to increased intraglomerular pressure is the un- derlying cause for podocyte injury [98, 120, 121]. Glomerular ischemia and/or hyperfi ltration have also been suggested to promote glomerular injury. Notably, systemic hypertension and/or elevated intraglomerular pressure associate with other types of nephropathy, e.g. diabetic neph- ropathy [115]. Control of hypertension with antihypertensive agents is essential for the treatment of the disease.

Diabetic nephropathy Diabetic nephropathy is the most common cause of nephrotic proteinu- ria [122]. It has been estimated that approximately one-third of patients with diabetes develop diabetic nephropathy [123, 124], which can be considered the most severe complication of diabetes due to the associ- ated several fold increased risk of cardiovascular disease and mortality [125, 126]. During the course of diabetes, the incidence of diabetic ne- phropathy is not linear, but it increases until 15 years after the onset of diabetes and declines thereafter [123]. Microalbuminuria is the fi rst sign of glomerular dysfunction in type 1 diabetes and is considered a powerful predictor of progression to overt nephropathy [127]. Microalbuminuria refl ects the alterations in the glomerular capillary wall at the early stage of the disease: Increase in podocyte foot process width and reduced number of podocytes per glomerulus have been reported [128-130]. At later stages of the disease, nodular glomerulosclerosis with expansion of the mesangial matrix and thickened GBM can be observed [131]. It should be noted that diabetic nephropathy is part of a generalized microvascular damage occurring af-

20 2 Review of the literature ter prolonged hyperglycaemia [132]. During this process, involvement of the renin-angiotensin system, advanced glycosylation end-products and reactive oxygen species have been indicated as pathogenetic mechanisms (reviewed in [133]). Importantly, angiotensin II seems to have direct ef- fects on the podocytes, causing e.g. podocyte apoptosis and cytoskeleton rearrangement [120, 134]. Hyperglycaemia is the most important risk factor for diabetic neph- ropathy both in type 1 and 2 diabetes [135, 136]. However, patients with poor glycaemic control do not necessarily develop nephropathy, indicat- ing that other factors are involved. Additional risk factors include smok- ing [137], hyperlipidemia [138], and insulin resistance even in patients with type 1 diabetes [139]. The prevailing view is that the combination of environmental and genetic risk factors (reviewed in chapter 2.5) is the driving force for the development of diabetic nephropathy. Elimination of classical risk factors, improvement of glycaemic control, aggressive treatment of hypertension, and cessation of smoking are es- sential for the prevention of the development of diabetic nephropathy. Currently used therapies for diabetic nephropathy focus on blocking the renin-angiotensin system with angiotensin-converting enzyme inhibitors and angiotensin receptor antagonists [140, 141].

2.5. Genetic basis of proteinuria

Milestone discoveries of the genetic basis of proteinuria have shed light on the molecular structure of glomerular capillary wall and have dis- closed mutations in genes encoding adhesion proteins, components of cell cytoskeleton, and mediators of signal transduction (reviewed in [142- 144]). Notably, most of these proteins were located in the glomerular podocytes. Therefore, renal research has in recent years adopted a new concept of a ‘podocentric view of nephrology’ [145]. According to the classifi cation by Wilhelm Kriz [142], hereditary dis- eases affecting the podocytes can be divided into two groups based on their age of onset. First, genetic defects manifesting at birth or soon af- ter birth (‘early onset’) are characterized by the lack of or deformed slit diaphragms and/or nephrotic range proteinuria [142] (Table 1). Second, proteinuric disorders occurring from childhood to late adulthood can be regarded as a consequence of the reduced ability of the podocyte to re- spond to physiological challenge, accounting for the later age of onset [142]. Typically, these diseases segregate in an autosomal-dominant pat- tern in families with variable penetrance (Table 1). The gene encoding the slitdiaphragm protein nephrin, NPHS1, is mu- tated in congenital nephropathy of the Finnish type (CNF) [27]. This disor- der was fi rst characterized by Hallman and co-workers in 1956 [146], and it belongs to the Finnish disease heritage [147], a term used to describe

2 Review of the literature 21 Table 1. Hereditary proteinuric diseases classifi ed according to age at onset. Modifi ed from [142] and [143].

Chromosomal Mode of Disease Gene (protein) region inheritance References

Early onset

Congenital nephrotic syndrome of NPHS1 (nephrin) 19q13.12 recessive [27] the Finnish type

Denys-Drash syndrome WT1 (WT1) 11p13 dominant [161]

Pierson’s syndrome LAMB2 (laminin β2 3p21.31 recessive [152] chain)

Nephrotic syndrome type 3 (NPHS3) PLCE1 (phospholipase 10q23.33 recessive [151] C ε1

Steroid-resistant nephrotic syndrome NPHS2 (podocin) 1q25.2 recessive [73]

Late onset

Alport syndrome COL4A3 (α3(IV) 2q36.3 recessive / [228] [229] collagen) dominant [230]

COL4A4 (α4(IV) 2q36.3 collagen)

COL4A5 (α5(IV) Xq22.3 X-linked [231] collagen)

Epstein and Fechtner syndromes MYH9 (non-muscle 22q12.3 dominant [155] myosin heavy chain 9)

Focal segmental glomerulosclerosis ACTN4 (α-actinin-4) 19q13.2 dominant [83]

Focal segmental glomerulosclerosis TRPC6 (TRPC6) 11q22.1 dominant [77] [76]

Frasier’s syndrome WT1 (WT1) 11p13 dominant [162]

Nail-patella syndrome LMX1B (LMX1B) 9q33.3 dominant [160]

The chromosomal regions are derived from the UCSC Genome Browser (March 2006 Assembly).

22 2 Review of the literature diseases that occur more frequently in Finland than elsewhere in the Eu- rope. In CNF, the podocyte foot processes are fl attened and effaced, and the slit diaphragms are missing [148, 149]. Affected babies suffer from massive treatment-resistant proteinuria already in utero [146, 150]. The NPHS2 gene encoding podocin is mutated in steroid-resistant ne- phrotic syndrome [73]. Recently, a third nephrosis gene, namely phos- pholipase C ε1 (PLCE1), was discovered, with expression in the developing and mature podocytes [151]. Interestingly, some patients with mutations in PLCE1 were steroid-sensitive, indicating that patients carrying certain mutations in PLCE1 could be cured. Laminin β2-chain is a component of mature glomerular basement membranes, and certain mutations in the human LAMB2 gene account for Pierson’s syndrome [152]. Pierson’s syn- drome is characterized by congenital nephrosis with mesangial sclerosis and ocular defects [152]. Familial forms of focal segmental glomerulosclerosis (FSGS) with de- fects in the non-selective cation channel TRPC6 [76, 77] and the actin-fi la- ment cross-linking protein α-actinin-4 (ACTN4) [83] have been character- ized. Interestingly, mutations of NPHS2 with later onset have also been reported in familial FSGS [153], and NPHS2 has been suggested to associ- ate with microalbuminuria in the general population [154]. The impor- tance of the podocyte cytoskeleton was further emphasized by the May- Hegglin anomaly, and Sebastian, Epstein and Fechtner syndromes that involve mutations in the gene encoding the non-muscle myosin heavy chain 9 (MYH9) expressed in the podocytes [155]. In these syndromes, the patients suffer from macrothrombocytopenia and variably from other symptoms including proteinuria [156]. Alport syndrome involves muta- tions in the genes encoding the constituents of type IV collagen (α3α4α5) of the GBM and presents with hematuria and progressive glomerular dis- ease, hearing loss and ocular lesions (reviewed in [157]). Transcription factors LMX1B and WT1 regulate several podocyte-ex- pressed genes and are involved in the early phases of the kidney and podocyte differentiation [158, 159]. Mutations in the LMX1B gene have been mapped in patients with nail-patella syndrome, a dominantly in- herited disorder with skeletal malformations and, in some patients, renal dysfunction due to glomerular lesions [160]. Variants in the Wilms tumor 1 gene WT1 have variable phenotypes: The mutation spectrum of Denys- Drash syndrome is broad covering the whole gene locus resulting in an early onset of nephropathy [161]. Patients with Denys-Drash syndrome suffer from generalized urogenital developmental abnormalities [161]. In Frasier syndrome, characterized by male pseudohermaphrodism and progressive glomerular disease, proteinuria develops later due to mu- tations in the donor splice site of intron 9 of the WT1 gene [162]. Kriz classifi es nail-patella syndrome and Frasier syndrome in the late-onset group based on their clinical features, but emphasizes that the disease mechanisms are not well understood [142].

2 Review of the literature 23 In diabetic nephropathy, familial clustering has been reported in several studies indicating the existence of a genetic component in the λ disease [163-166]. In the Finnish population, the sibling recurrent risk s is 2.3, which indicates the risk that diabetic siblings of a nephropathic proband have compared to siblings of a proband without diabetic neph- ropahty [163]. Furthermore, the albumin excretion rate and the extent of glomerular lesions seem to be partially heritable and genetically deter- mined in families with patients with diabetes [167, 168]. Diabetic nephropathy can be considered a typical complex disease with multiple genetic loci that contribute to the disease development in combination with other biological effects (e.g. hyperglycaemia) [169]. Genome-wide screens using sib-pairs or case-control setting have re- vealed potential susceptibility loci for diabetic nephropathy on chromo- somes 10q and 3q [170-172]. Candidate gene studies have focused on the known biological pathways involved in the pathogenesis of diabetic nephropathy. These pathways include the genes of the renin-angiotensin system, and a recent meta-analysis of several studies support the involve- ment of angiotensin I converting enzyme insertion/deletion polymor- phism in the development of diabetic nephropathy [173]. However, most attempts to determine the effect of a genetic variant in a particular gene on the development of diabetic nephropathy have suffered from the lack of suffi cient statistical power (discussed in [174]). Transgenic animal models have provided new insights in the genet- ics of proteinuria. Defi ciency of the slit diaphragm proteins NEPH1, a nephrin-homologue, and FAT1, a member of the cadherin superfamily, was discovered to cause massive proteinuria and lethality in mouse [53, 66]. Inactivation of the gene encoding CD2-associated protein (Cd2ap) in mice results in congenital nephrotic syndrome and death at 6 to 7 weeks of age [71]. Interestingly, haploinsuffi ciency of CD2AP may also be in- volved in human glomerular disease [175]. To my knowledge, no genetic defect has been mapped to the human homologues of NEPH1 and FAT1 in proteinuria.

24 2 Review of the literature 3. AIMS OF THE STUDY

Since the discovery of the causative gene for the congenital nephrotic syndrome of the Finnish type, NPHS1, and the identifi cation of the gene product nephrin as a constituent of the slit diaphragm [27, 32, 34], the prevailing view has been that the slit diaphragms are of pivotal impor- tance for the molecular sieve in the kidney glomerulus. This thesis hy- pothesized that additional nephrin-associated molecules participate in maintaining the function of the fi ltration barrier in health and disease.

The specifi c aims of this thesis were to evaluate the following: 1. To determine the expression and localization of fi ltrin, a novel ne- phrin-like protein, in the kidney.

2. To clone and characterize the nephrin antisense transcript Nphs1as in mouse.

3. To elucidate the expression pattern and distribution of fi ltrin in the kidneys of patients with proteinuric diseases.

4. To investigate the genetic association of the nephrin-associated genes densin-180 (LRRC7), NEPH1 (KIRREL), podocin (NPHS2), fi ltrin (KIR- REL2), and α-actinin-4 (ACTN4) with diabetic nephropathy in type 1 diabetes.

3 Aims of the study 25 4. MATERIALS AND METHODS

4.1. Study subjects and clinical samples

Samples of normal human kidney were obtained from cadaver donors inappropriate for transplantation due to vascular anatomic reasons (De- partment of Surgery, Helsinki University Central Hospital), and glomeruli were extracted from the cortical kidney tissue using the graded sieving method [176] (Study I). Primary cultures of human podocytes (chapter 4.3) were established from kidney specimens obtained from renal carci- noma patients undergoing nephrectomy. Two sets of patients were selected for quantitative RT-PCR analysis of fi ltrin (KIRREL2), nephrin (NPHS1), and podocin (NPHS2) mRNAs (Study III). The fi rst set of 69 kidney biopsies was collected within the European Renal cDNA Consortium (members listed in Appendix I), and the biopsies were divided into the following groups according to histological diagno- sis (Table 2). In these samples, the glomerular compartment was microdis- sected and processed for RNA isolation as described previously [177]. The second set of samples was collected at the St. Vincent’s Hospital (Fitzroy, Australia; Table 2), and total RNA was extracted from whole biopsies. Nonaffected parts of tumor nephrectomies served as a control group. For immunohistochemical examination of fi ltrin in injured glomeruli, a total of 15 biopsies obtained for routine diagnostic purposes at the San Carlo Borromeo Hospital (Milan, Italy) were used (Study III). The diagnos- tic groups of the biopsies were as follows: minimal change disease (n = 4), membranous glomerulopathy (n = 6), and focal segmental glomeru- losclerosis (n = 5). In addition, renal samples of the tumor-free part of nephrectomy specimen were used as controls (n = 3). Two groups of Finnish patients with type 1 diabetes with and without diabetic nephropathy were selected for the genetic association analysis of densin-180 (LRRC7), NEPH1 (KIRREL), podocin (NPHS2), fi ltrin (KIRREL2), and α-actinin-4 (ACTN4). The patients were part of the Finnish Diabetic Nephropathy (FinnDiane) Study (Study IV). First, the extensive screening of single nucleotide polymorphisms (SNPs) was performed in 1103 pa- tients. Second, an additional set of 1025 patients was analysed in order to confi rm the positive fi ndings (Table 2). The renal status of the patients was based on the albumin excretion rate (AER). Normoalbuminuria was defi ned as AER below 30 mg/24h (< 20 μg/min in an overnight urine col- lection), microalbuminuria as AER of 30-300 mg/24h (20-200 μg/min), and

26 4 Materials and methods Table 2. Clinical characteristics of analyzed subjects.

Age Serum creatinine Diagnostic group n (years) (µmol/l) Proteinuria (g/24h)

European Renal cDNA Consortium

Control subjects 7 64±6 NA NA

Minimal change disease 13 36±16 101±44 5.7 (0.2 - 14.8

Hypertensive nephropathy 16 56±10 172±151 1.0 (0 - 2.6)

Membranous glomerulopathy 31 55±23 130±108 4.9 (0.4 - 17.0)

Focal segmental glomerulosclerosis 9 47±16 128±45 2.3 (0.2 - 10.4)

St. Vincent’s Hospital

Control subjects 6 NA NA NA

Membranous glomerulonephropathy 4 65±4 125±50 4.1 (1.9 - 8.0)

Diabetic nephropathy 7 51±19 173±80 1.2 (0.6 - 21.2)

Minimal change disease 5 31±12 86±21 8.2 (6.3 - 14)

Focal segmental glomerulosclerosis 11 56±19 145±42 4.0 (1.0 - 14.0)

FinnDiane - 1st set of patients AER (mg/24h)

Normoalbuminuria 459 43±10 85±15 7 (1 - 85)

Microalbuminuria 276 37±11 90±19 59 (2 - 613)

Macroalbuminuria 368 39±9 165±125 588 (4 - 8348)

FinnDiane - 2nd set of patients

Normoalbuminuria 607 40±12 83±17 8 (1 - 101)

Macroalbuminuria 158 45±10 163±114 426 (11 - 4609)

End-stage renal disease 260 44±8 - -

Data are summarized as mean ± standard deviation or median (range). AER, albumin excretion rate; NA, not available.

4 Materials and methods 27 macroalbuminuria as AER over 300 mg/24h (> 200 μg/min) in two out of three consecutive urine collections [178]. Patients with normoalbuminu- ria were required to have duration of diabetes over 15 years. Patients on dialysis or with a kidney transplant were considered to have an end-stage renal disease (ESRD). Informed consent was obtained from all subjects. The study protocols followed the principles of the Declaration of Helsinki and were approved by local ethics committees.

4.2. Animals

Normal rat kidneys (Sprague-Dawley) were used to explore the expres- sion and localization of fi ltrin (Studies I and III). Glomeruli were isolated from the kidney tissue using the sieving method as previously described for the rat tissue [176]. Tissues (brain, kidney, pancreas, mesenteric and peripheral lymph nodes, spleen, and thymus) of adult male mice (strain NMRI) were prepared for the expression analyses of the Nphs1 and Nphs1as mRNAs (Study II). In addition, femoral and tibial bone marrow were harvested for preparing of dendritic cells [179]. Macrophages were obtained from peritoneal cavity after stimulation with intraperitoneally injected fetal calf serum. Macrophages were collected from the perito- neal cavity with RPMI 1640 medium (BioWhittaker Europe, Verviers, Bel- gium). The local research ethics committees approved all experiments.

4.3. Cell culture

Various cell lines and primary cell cultures were employed to investigate the expression pattern of fi ltrin, nephrin and Nphs1as (Studies I, II and III; Table 3)

4.4. Molecular cloning and mRNA expression analysis

Complementary DNA clones The cDNA clones UI-M-BH2.1-apt-g-03-0-U (mouse Nphs1as cDNA) and DKFZp564A1164 (human fi ltrin cDNA) were purchased from Research Genetics (Invitrogen, Paisley, UK) and RZPD Resource Center (Berlin, Ger- many) for further experiments.

RNA isolation Total RNA was isolated from tissues and cultured cells using chloroform- phenol extraction in the presence of Trizol reagent (Life Technologies,

28 4 Materials and methods Table 3. Summary of cell culture experiments.

Cell culture Culture media Reference Study

Human podocytes DMEM containing 10% fetal calf serum, 2.5 mM [232] I glutamine, 0.1 mM sodium pyruvate, 5 mM HEPES buff er, insulin, transferrin, and 5 mM sodium selenite.

Mouse dendritic cells RPMI 1640 medium with 10% fetal calf serum [179] II supplemented with 10 ng/ml murine recombinant GM-CSF.

Mouse embryonic stem cell DMEM with 20% fetal calf serum containing 1000 [233] II line 129/Sv units/ml of leukemia inhibitory factor.

Mouse epithelial cortical 1:1 mix of Ham F12 medium and DMEM containing [81] II collecting duct cell line M-1 5% fetal bovine serum and 5 μM dexamethasone.

Mouse insulinoma DMEM with GlutaMax-1 supplemented with [234] II cell line MIN6 15% fetal calf serum, 1 mM sodium pyruvate, and 0.05 mM β-mercaptoethanol.

Gibco BRL, Paisley, UK) or with an RNA isolation kit (Atlas Glass Total RNA Isolation kit [Clontech, Carlsbad, CA, USA] and RNeasy kit [Qiagen, Hilden, Germany]) according to the manufacturers’ instructions. To re- move traces of genomic DNA, isolated RNA was incubated with RNase- free DNaseI (Promega, Madison, WI, USA) for 30 minutes together with human placental RNase inhibitor (Promega).

Reverse transcriptase-polymerase chain reaction (RT-PCR) Several applications of reverse transcriptase-polymerase chain reaction (RT-PCR) were employed in gene expression analysis at the mRNA level. The methods included qualitative, (semi)quantitative, and orientation- specifi c RT-PCR analysis, and, for mRNA cloning, rapid amplifi cation of cDNA ends (RACE). Complementary DNA (cDNA) was synthesized from RNA using the M- MLV (Promega) or AMV (Roche Diagnostic GmbH, Mannheim, Germany)

reverse transcriptase enzymes using either oligo-dT15 primers or random hexamers (Roche Diagnostic GmbH). In orientation-specifi c RT-PCR anal- ysis of the Nphs1 and Nphs1 as mRNAs, gene-specifi c oligonucleotides were used for priming the cDNA synthesis (Study II). Reactions without the reverse transcriptase enzyme served as negative controls. Thereafter, the cDNA was amplifi ed using gene-specifi c primer pairs (Studies I and II) with AmpliTaq Gold DNA polymerase (Applied Biosys- tems, Foster City, CA). In semiquantitative RT-PCR of Nphs1 and Nphs1as

4 Materials and methods 29 (Study II), serial dilutions (1:5 - 1:125) of cDNA samples were amplifi ed within the linear range of the PCR reaction, and the mRNA levels were normalized to the house-keeping gene cyclophilin A (Ppia, peptidylprolyl isomerase A). Cyclophilin A can be considered a suitable internal refer- ence gene for comparing the expression of a target gene between dif- ferent tissues, since the expression of cyclophilin A is not restricted to a specifi c tissue or cell type [180]. The RT-PCR products were analyzed electrophoretically and visualized using ethidium bromide. Appropriate controls without the cDNA template were included in each experiment. The amplifi ed PCR products were confi rmed with DNA sequencing using an ABI 373 sequencer (Applied Biosystems, Foster City, CA, USA) and ABI Prism Dye Terminator kit (Applied Biosystems). For quantitative real-time RT-PCR analysis of RNA isolated from hu- man kidney biopsies, the gene expression assays were ordered from Applied Biosystems. Pre-developed assay reagents were used for fi ltrin (Assay ID: Hs00375638_m1) and glyceraldehyde-3-phosphate-dehydro- genase (GAPDH; Hs99999905_m1). According to a paper reporting the validation of internal reference genes for microdissected biopsies [181], the cyclophilin A mRNA was expressed at low levels in isolated glomeruli and was thus not suitable as a reference gene. Instead, GAPDH can be regarded as a potential housekeeping gene in microdissected glomeruli [181, 182]. For nephrin and podocin, previously developed forward and reverse primers and probes were used [47]. The gene expression measure- ment was based on the TaqMan technology which utilizes the 5’-nuclease activity of the AmpliTaq Gold DNA polymerase and fl uorescently labeled gene-specifi c probe containing a reporter dye at the 5’-end of the probe and a quencher at the 3’-end of the probe. Each sample was measured in duplicate, and the PCR amplifi cation was performed with the ABI Prism 7700 Sequence Detection System (Applied Biosystems) equipped with in- strumentation for fl uorescence detection. The normalized expression of ΔCt Δ the analysed gene was calculated with the 2 method [177], where Ct was defi ned as Ct(gene) - Ct(GAPDH), Ct denoting the threshold cycle for each measurement. For cloning the antisense transcript of the mouse nephrin gene (Nphs1as), RACE [183] was performed with the GeneRacer kit (Invitrogen) according to the manufacturer’s instructions using the gene-specifi c 5’- and 3’-RACE primers (Study II). The nested PCR products were subcloned into a pCR4-TOPO vector (Invitrogen) and sequenced as described above.

Northern Blotting The molecular weight of the Nphs1as mRNA was determined using the Northern blotting analysis [184]. RNA isolated from ES-cells (strain 129/ Sv) was separated with gel-electrophoresis and blotted onto a nylon membrane. The fi lter was hybridized with a [α−32P]UTP-labeled ribo- probe. This strand-specifi c probe was synthesized with a T7 polymerase

30 4 Materials and methods (Promega) from the pCR4-TOPO vector containing the 3’-RACE fragment of the Nphs1as cDNA.

RNA dot blot assay In order to get insight into the fi ltrin mRNA expression pattern in human tissues, an RNA dot blot technique was chosen, and Human Multiple Tis- sue Expression (MTE™) Array (Clontech Laboratories) containing poly(A) mRNAs from 76 tissues was selected for the expression profi ling. Briefl y, [α−32P]dCTP labeled cDNA probe was synthesized with Rediprime II ran- dom prime labeling system (Amersham Pharmacia Biotech, Uppsala, Swe- den) using the insert of nucleotides 496-2273 from the DKFZp564A1164 clone as a template. The MTE blot was hybridized overnight with the purifi ed (NICK column; Amersham Pharmacia Biotech) probe in a hybridi- zation oven, and analyzed after an overnight exposure to a phosphoim- ager plate at -75°C with a Bio-imaging analyzer (Fuji Photo Film co, Kana- kawa, Japan).

4.5. Immunochemical methods

Antibodies Polyclonal anti-fi ltrin antibodies were generated in rabbits, since fi ltrin was a novel and previously uncharacterized protein. Two hydrophilic oli- gopeptides were selected from the fi ltrin protein sequence: The extracel- lular peptide GPS PHF LQQ PED LVV LLG EE(C) represented the amino acids 21-40 of fi ltrin and was highly conserved in the rat and mouse. An additional cysteine (C) was added for coupling the peptide to a carrier protein. The second peptide consists of the amino acids 533-548, CWR HSK ASA SFS EQK N, in the intracellular domain of fi ltrin. The peptides were synthesized, purifi ed, and coupled to the carrier protein keyhole limpet hemocyanin by Alpha Diagnostics International Inc. (San Antonio, TX, USA). Two rabbits (NZW strain) were immunized with each peptide according to standard protocols. Prior to immunizations, appropriate control serum samples were collected. The antiserum against the extra- cellular peptide was immuno-purifi ed with the corresponding peptide coupled to Epoxy-Activated Sepharose CL-6B Resin (Sigma-Aldrich, St. Louis, MO, USA) to yield peptide-specifi c fractions, and antiserum against the intracellular peptide with the protein-A-sepharose (Amersham Bio- sciences), respectively, following the manufacturers’ instructions. The polyclonal anti-nephrin antibody was raised in rabbits immunized with a recombinant intracellular part of rat nephrin [88]. The antiserum was purifi ed with a protein-A-column to isolate IgG fractions. The mono- clonal anti-synaptopodin antibody from Progen (Goettingen, Germany) was used as a podocyte-specifi c marker.

4 Materials and methods 31 Immunoblotting The molecular weight and expression of fi ltrin and nephrin were de- termined by immunoblotting in the kidney and other tissues. Tissue sam- ples were homogenized in RIPA buffer (50 mM Tris-HCl, 150 mM NaCl, 1% NP-40, 0.1% SDS and 0.5% sodiumdeoxycholate, pH 8.0) in the pres- ence of a protease inhibitor cocktail (Complete Mini, EDTA-free, Roche Diagnostics). The detergent extracts were suspended in the Laemmli sample buffer [185], boiled for 5 minutes, separated with SDS-PAGE, and electrotransferred to nitrocellulose membranes (Amersham Life Sciences, Buckinghamshire, UK). Proteins were detected either with the intrac- ellular anti-fi ltrin or with the anti-nephrin antibody followed by HRP- conjugated goat anti-rabbit IgG (Jackson ImmunoResearch Laboratories, Inc., West Grove, PA, USA) using the ECL method (SuperSignal West Pico Chemiluminescent Substrate kit; Pierce, Rockford, IL, USA).

Immunofl uorescence microscopy The distribution of fi ltrin was investigated in the cortical kidney tissue from cadavers and in kidney biopsies from patients with proteinuric dis- eases using immunofl uorescence microscopy (Studies I and III). Five μm thick frozen sections were fi xed with either ice cold-acetone or 3.5% paraformaldehyde in PBS followed by permeabilization with 0.1% Tri- ton X-100 in PBS, and detected with the indirect immunofl uorescence method. Briefl y, the intra- or extracellular anti-fi ltrin antibodies were in- cubated on the sections followed by appropriate fl uorescently-labelled secondary antibodies (FITC goat anti-rabbit IgG (H+L) [Jackson Immu- noResearch Laboratories] or Alexa Fluor 546 goat anti-rabbit IgG [Molec- ular Probes, Invitrogen]). In dual labeling experiments, the sections were incubated with the anti-synaptopodin antibody and Alexa Fluor 488 goat anti-mouse IgG (Molecular Probes) as a secondary antibody prior to staining with the anti-fi ltrin antibody. The sections were embedded in an anti-fading aqueous medium and examined under a fl uorescence microscope equipped with appropriate fi lters. As a secondary antibody control, the primary antibodies were omitted or replaced with irrelevant IgGs. Additionally, the intracellular anti-fi ltrin was immunodepleted by preincubation with the antigenic peptide to confi rm the specifi city in im- munofl uorescence analysis.

Immunoelectron microscopy Subcellular localization of fi ltrin in podocytes was explored with indirect immunoelectron microscopy (Study III). Rat kidneys were perfusion fi xed with 4% freshly prepared paraformaldehyde in PBS. Ultrathin sections were incubated with the extracellular anti-fi ltrin antibody, followed by incubation with 10 nm gold-conjugated goat anti-rabbit IgG (1:50; Am- ersham Biosciences).

32 4 Materials and methods 4.6. Genotyping

In Study IV, the genetic association of LRRC7, KIRREL, NPHS2, KIRREL2, and ACTN4 with diabetic nephropathy was investigated using a compre- hensive set of bi-allelic single nucleotide polymorphisms (SNPs). The SNPs were selected from public SNP databases (Table 4). The Tagger program [186] was employed to select tag SNPs for ACTN4 and NPHS2 with pair- wise r2 > 0.8. In LRRC7, KIRREL, KIRREL2 evenly distributed SNPs covering the gene loci were chosen. DNA was extracted from peripheral blood col- lected from the patients. Genotyping was performed using the Homog- enous MassExtend MassArray System (Sequenom, San Diego, CA, USA) based on the MALDI-TOF determination of primer extension products [187] or with the ABI Prism 7900 Sequence Detection System (Applied Biosystems, Foster City, CA, USA) utilizing allele-specifi c TaqMan probes [188]. To assure genotyping accuracy, internal controls were included in each analysis.

4.7. Databases, bioinformatics tools, and statistical methods

Publicly available databases and bioinformatics search engines were utilized in searching for novel nephrin homologues (including fi ltrin [KIRREL2] and Nphs1as) and in gaining information about their gene structure and genomic localization, expression profi les and protein ar- chitecture (Table 4). Genomic databases were employed for the selection of SNPs, and appropriate tools were selected to estimate the power of statistical tests and haplotype and linkage disequilibrium (LD) values of the analyzed genes (Table 4). Continuous variables (clinical variables and gene expression data) are summarized as mean value ± standard deviation or median value (range). Differences between multiple groups for normally distributed variables were evaluated with ANOVA or the Dunnett 2-sided test, non-normally distributed variables were either log10 transformed prior to analysis or assessed with the non-parametric Kruskal-Wallis test. Pearson correla- tion coeffi cients were calculated and their signifi cances were tested to evaluate correlation between two variables. Hardy-Weinberg calcula- tions were performed to test genotype equilibrium for each SNP, and the genotype data were compared with the χ2-test. Logistic regression model with the nephropathy status as the dependent variable and risk factors as covariates was used to estimate the independent association of SNPs. Statistical analysis was performed with SPSS 12.0.1 for Windows software (SPSS Inc., Chicago, IL, USA).

4 Materials and methods 33 Table 4. Summary of bioinformatic databases and computational tools.

Database / Tool Description URL Study

AUG_EVALUATOR Prediction of translation start codons http://www.itba.mi.cnr.it/webgene/ I

BLAST Basic local alignment search tool http://www.ncbi.nlm.nih.gov/BLAST/ I, II

ClustalW Multiple sequence alignment tool http://www.ebi.ac.uk/clustalw/ I

CpGPlot Detection of http://www.ebi.ac.uk/emboss/cpgplot I, II CpG-rich regions in genomic DNA dbSNP Database of single nucleotide http://www.ncbi.nlm.nih.gov/SNP/ IV polymorphisms

GenBank Database of genetic sequences http://www.ncbi.nlm.nih.gov/Genbank/ I, II

Genetic Power Power analysis of genetic http://pngu.mgh.harvard.edu/~purcell/gpc/ IV Calculator association tests

Genome Browser Database of genetic information http://genome.ucsc.edu/ IV

Genscan Prediction of gene structures in http://genes.mit.edu/GENSCAN.html I genomic DNA

Haploview Tool for haplotype and linkage http://www.broad.mit.edu/mpg/haploview/ IV disequilibrium analysis

HapMap Database of human genetic variation http://www.hapmap.org/ IV

HCtata Prediction of TATA-signals in http://www.itba.mi.cnr.it/webgene/ II eukaryotic genes

InterPro Database of protein families, http://www.ebi.ac.uk/interpro/ I domains and functional sites

NetOGlyc Prediction of mucin type GalNAc http://www.cbs.dtu.dk/services/NetOGlyc/ I O-glycosylation sites in proteins

Primer3 Tool for designing primers for http://frodo.wi.mit.edu/cgi-bin/primer3/ I, II exprssion analysis primer3_www.cgi

Proscan Prediction of promoter regions http://bimas.dcrt.nih.gov/molbio/proscan/ I

Prosite Database of protein families, http://www.expasy.ch/tools/scanprosite/ I domains and functional sites

SignalP Prediction of signal peptide cleavage http://www.cbs.dtu.dk/services/SignalP/ I sites in proteins

Tagger Tool for the selection and evaluation http://www.broad.mit.edu/mpg/tagger/ IV of tag SNPs from genotype data

TMHMM Prediction of transmembrane helices http://www.cbs.dtu.dk/services/TMHMM- I in proteins 2.0/

UniGene Database of gene transtcripts http://www.ncbi.nlm.nih.gov/entrez/ I, II

34 4 Materials and methods 5. RESULTS

5.1. Filtrin, a novel nephrin-like protein, is expressed in glomerular podocytes (Studies I and III)

The NCBI BLAST algorithm was used to search for uncharacterized cDNA clones and expressed sequence tags (ESTs) homologous to human ne- phrin. The cDNA clone DKFZp564A1164, originally isolated from the human fetal brain, was chosen as the candidate molecule for detailed analysis. This molecule was named ‘fi ltrin’, and the offi cial gene name provided by the HUGO Gene Nomenclature Committee is ‘kin of irre like 2’ (KIRREL2) due to its similarity with D. melanogaster irre. The sequence analysis revealed that fi ltrin is a type I transmembrane protein of the immunoglobulin superfamily of adhesion proteins. It is composed of a cleavable N-terminal signal sequence, fi ve extracellular immunoglobulin-like domains followed by an acid alpha-helical trans- membrane region, and a cytoplasmic carboxy-terminal domain with a proline-rich region (Study I; Figure 4). Several putative N- and O-glyco- sylation sites were detected at the extracellular side, and the cytoplasmic domain has potential phosphorylation sites at serine, threonine or tyro- sine residues. The full-length form of fi ltrin consists of 708 amino acids. In addition to sequence homology to human nephrin (30%), fi ltrin shows considerable homology and structural similarity to the nephrin-like pro- teins NEPH1 (44%) and NEPH2 (41%) (Table 5).

TM PDZK1

SP

lg Grb2-SH2

Figure 4. Schematic domain composition of the nephrin-like proteins. The proteins consist of fi ve extracellular immunoglobulin-like domains (Ig) followed by a transmem- brane domain (TM) and a cytoplasmic domain with the Grb2-SH2 and PDZK1 binding sites. The nephrin-like proteins interact with their cytoplasmic Grb2-SH2 and PDZK1 binding sites with podocin and ZO-1, respectively. SP, signal peptide. Drawn after [57].

5 Results 35 Table 5. Nephrin and the nephrin-like proteins, the genes encoding them and molecular charac- teristics.

Chromosomal location Protein

Protein Gene Human Mouse Size (aa) Molecular weigth (kDa)

Nephrin NPHS1 19q13.12 7qB1 1241 180

NEPH1 KIRREL 1q23.1 3qF1 757 90-110

NEPH2 KIRREL3 11q24.2 9qA4 778 100-125

Filtrin/NEPH3 KIRREL2 19q13.12 7qB1 708 107

The size of the protein is given in the number amino acids, and the apparent molecular weight represents the full-length isoform. The chromosomal regions are from the UCSC Genome Browser (March 2006 Assembly for human; February 2006 for mouse). aa, amino acid; kDa, kilo Dalton.

The human fi ltrin gene KIRREL2 resides on 19q13.12 ad- jacent to NPHS1 encoding nephrin (Study I). It is of particular note that NPHS1 and KIRREL2 are transcribed in the opposite directions, and the distance between the transcription start sites is approximately 5 kb in hu- mans and 3 kb in the mouse and rat. This chromosomal arrangement is conserved among the species (Figure 5). In addition to the full-length form of fi ltrin, alternatively spliced variants of the fi ltrin mRNA were discov- ered. The alpha-form lacks exon 4, potentially deleting a part of the extra- cellular immunoglobulin-like domains of the translated protein, and the beta-forms represents a variant consisting only of the sequence coding for the signal peptide fused to the carboxy-terminal intracellular part of the protein (Study I). As reported by Sun and colleagues [56], the spectrum of alternatively spliced variants was broadened with the cloning of a mRNA variant, which lacks exon 2. Two independent methods were employed to detect the expression of fi ltrin in the kidney. First, the fi ltrin mRNA was successfully amplifi ed from the isolated human and rat glomeruli and from the cultured hu- man podocytes with RT-PCR (Study I). Second, anti-fi ltrin antibodies were raised in rabbits to explore fi ltrin at the protein level, and the antibodies were affi nity-purifi ed to reach maximal specifi city. The immunoblotting studies of the glomerular lysates with the extracellular anti-fi ltrin anti- body showed the presence of a protein band with an apparent molecular weight of 107 kDa (Study I). In addition, the cultured human podocytes expressed protein isoforms of varying molecular weight. The difference between the theoretical molecular weight of 75.1 kDa and the experi- mentally verifi ed size of 107 kDa may be explained by post-translational modifi cation by N- or O-glycosylation or other covalent modifi cations.

36 5 Results ~5 kb

NPHS1 KIRREL2

Figure 5. The genomic organization and exon-intron structure of the human NPHS1 and KIRREL2 genes in 19q13.12. Their transcription initiation sites are separated by a stretch of 5 kb, possibly sharing common regulatory elements. The arrows indicate the 5’-3’ direction of the mRNA transcription (Study I).

In order to obtain an insight into the global expression pattern of fi ltrin in human tissues, both bioinformatics and experimental methods were utilized. First, the expression of the fi ltrin-specifi c mRNA was ex- plored with the EST sequence analysis: The UniGene cluster Hs.145729 contained EST sequences isolated from the lung, eye, lymphatic tissue, pancreas, brain, kidney, and pituitary gland (Study I). Second, the expres- sion pattern was deciphered with hybridizing the radioactively labelled fi ltrin-specifi c probe to the nylon fi lter containing poly(A) mRNAs from 72 human tissues. According to the RNA dot blot analysis, fi ltrin mRNA is primarily expressed in the lymph nodes and the pancreas (Study I). The distribution of fi ltrin in the kidney was examined with immuno- histological methods. Immunofl uorescence staining with the extracel- lular anti-fi ltrin antibody revealed expression in the glomerulus and to some extent in proximal and distal tubules (Study I). According to the staining with the intracellular antibody, fi ltrin is expressed in an epi- thelial cell-like pattern in the glomerulus (Figure 6A; Study III). Most of the fi ltrin expression within the glomerulus was confi ned to podocytes as indicated by immunofl uorescent double labeling with the podocyte marker synaptopodin (Study III). The expression of fi ltrin was also ex- amined at the ultrastructural level in podocytes, and the results indicate that fi ltrin localizes at the slit diaphragm of the podocyte foot processes (Figure 7; Study III).

5.2. Nphs1as mRNA is transcribed from the nephrin locus in a reverse orientation in the brain and lymphoid tissue (Study II)

The characterization of the nephrin-defi cient mouse model generated by random insertional mutagenesis revealed that the gene-trap vector PT1βgeo, containing an Engrailed splice-acceptor site fused in-frame to the LacZ/neomycin fusion gene [189], was integrated in an inverted orientation in exon 8 of the Nphs1 gene [190] (Figure 8B). Notably, the

5 Results 37 AB

Figure 6. Distribution of fi ltrin in the kidney glomerulus as shown by immunofl uo- rescence microscopy. (A) Filtrin is expressed in an epithelial cell-like pattern in the normal glomerulus. (B) In primary FSGS, a decrease in the fi ltrin staining can be observed, ranging from segmental to global.

Figure 7. (A and B) On immunoelectron microscopical examination, gold particles were detected at the slit diaphragm area suggesting that fi ltrin localizes at the slit diaphragm. US, urinary space; GBM, glomerular basement membrane (Study III).

38 5 Results trapped ES-cells were successfully screened by the selection marker gene and 5’-RACE, indicating the presence of an actively transcribed gene in an antisense orientation to nephrin. To elucidate this issue, Nphs1-like ex- pressed sequence tags (ESTs) were searched from dbEST using the BLAST tool at NCBI. Among several positive hits, the clone UI-M-BH2.1-apt-g- 03-0-UI, originally isolated and sequenced from the mouse brain (strain C57BL/6J), showed exact complementarity, however, in an antisense ori- entation (Figure 8A; Study II). An orientation-specifi c RT-PCR technique was chosen to exclude pos- sible sequence artifacts and to confi rm that the clone represents a true mRNA transcribed from the Nphs1 locus. Several primer pairs were de- signed with the help of the cDNA clone UI-M-BH2.1-apt-g-03-0-UI and the Nphs1 cDNA. This analysis confi rmed the presence of the Nphs1as mRNA in the mouse brain (strain NMRI) and embryonic stem (ES) cells (strain 129/Sv). No PCR product could be amplifi ed from cDNAs prepared from the kidney and pancreas, suggesting that these tissues do not ex- press Nphs1as. All tissues examined and the ES cells expressed the Nphs1 mRNA (Study II). Notably, the amplifi cation of Nphs1as was only success- ful with the primer pair crossing the intron 7 of Nphs1 (Figure 8A), which is in agreement with the alignment of the EST clone UI-M-BH2.1-apt-g- 03-0-UI with the genomic sequence. More extensive screening of mouse lymphoid tissues indicated that the Nphs1as mRNA is transcribed in the thymus, peripheral lymph nodes and dendritic cells, whereas no expres- sion was observed in the mesenteric lymph nodes, spleen and macro- phages (Study II).

Figure 8. Nphs1as is transcribed from the nephrin encoding locus in a reverse orienta- tion. (A) Genomic location and exon–intron structure of Nphs1 and Nphs1as mRNAs. Locations of the primers 2s/2as and 2Bs/2Bas used to amplify Nphs1as in Study II are indicated. (B) Targeting vector PT1βgeo disrupts exon 8 of Nphs1 in the nephrin- defi cient mouse model. Notably, the vector is integrated in an inverted orientation. CpG, CpG island (Study II).

5 Results 39 Mounting evidence of the existence of the nephrin antisense tran- script encouraged us to proceed to cloning the full-length cDNA. A tech- nique known as the Rapid amplifi cation of cDNA ends (RACE) [183] was utilized to clone both the 5’- and 3’-ends of the transcript from the wild type murine brain RNA. The assembly of the isolated RACE-fragments with the sequence of the clone UI-M-BH2.1-apt-g-03-0-UI revealed that Nphs1as is a continuous 1339 bp mRNA (Study II). Northern blot analysis was used to further examine the molecular weight of Nphs1as mRNA. An RNA blot from embryonic stem cells was hybridized with a strand-specifi c 3’-RACE riboprobe. The size of the single transcript of approximately 1.5 kb detected in the analysis accords with the length of the Nphs1as cDNA isolated with RACE (Study II). Based on the information of the Nphs1 exon-intron structure [191], the transcription of Nphs1as initiates from the Nphs1 exon 12 and ex- tends until intron 6 (Figure 8A). Nphs1as is composed of three exons with consensus GT-AG splice sequences: the nucleotides 1-143 form exon 1, 144-558 exon 2, and 558-1339 exon 3. It is likely that the isolated se- quence contains a genuine 3’end since poly(A)-tail was located at the end and multiple poly(A) signal sites (nucleotide positions 1125-1137) were present near the A-rich region. These sites were not completely identical with the conserved AAUAAA or AUUAAA eukaryotic signal sequences [192], possibly reducing the accuracy or effi ciency of poly(A) tail forma- tion. A translated protein is not to be expected, because the sequence analysis did not revealed an open reading frame. To throw light on the possible correlation between the Nphs1 and Nphs1as levels and nephrin protein expression, semiquantitative RT-PCR of the sense and antisense transcripts and immunoblotting with nephrin- specifi c antibody were employed. Immunoblotting analysis identifi ed a double band of 185 kDa and 165 kDa in the kidney lysates, and a protein of 165 kDa in the lysates from the pancreas and brain (Study II). The intensity of amplifi ed RT-PCR products in dilution series of the cDNAs (from 1:5 to 1:125) was evaluated visually after gel-electrophoresis. In tissues expressing the sense and antisense transcripts, the level of the Nphs1as mRNA was remarkably higher in the brain than in peripheral lymph nodes and thymus. The Nphs1 mRNA was present in similar levels in the brain and peripheral lymph nodes, whereas the thymus expressed only a fraction of the other two (Study II).

5.3. Glomerular expression of fi ltrin in patients with proteinuric kidney diseases (Study III)

The expression of fi ltrin was examined in patients with proteinuric kidney diseases. First, fi ltrin mRNA expression was analyzed in isolated glomeruli from patients with focal segmental glomerulosclerosis (FSGS), minimal

40 5 Results A B

0.9 3.0 0.8 2.5 0.7 0.6 2.0 0.5 1.5 0.4 0.3 1.0

Filtrin/GAPDH 0.2 Filtrin/Nephrin 0.5 0.1 0.0 0.0 CON MCDHT MGN FSGS CON MCDHT MGN FSGS

Figure 9. (A) Expression of the fi ltrin mRNA in glomerular diseases in the ERCB samples. The results are normalized to the GAPDH gene. (B) An alternative normali- zation method was applied, where fi ltrin mRNA levels were adjusted to the nephrin mRNA levels in each sample. CON, control group (non-affected parts of tumour nephrectomies); MCD, minimal change disease; HT, hypertensive nephropathy; MGN, membranous glomerulopathy; FSGS, focal segmental glomerulosclerosis (Study III).

change disease (MCD), hypertensive nephropathy (HT), and membranous glomerulonephropathy (MGN). Control samples were obtained from the healthy kidney poles of individuals who underwent tumor nephrectomies. By real-time quantitative RT-PCR, the ratio of fi ltrin mRNA to GAPDH mRNA was determined. In addition, the levels of nephrin and podocin mRNA were measured from the same samples. Compared with control kidneys (n = 7), statistically signifi cant under-expression of fi ltrin was observed in all diagnostic groups, with the highest repression in patients with FSGS (n = 9, p = 4.5·10-6) and MGN (n = 31, p = 0.3·10-6; Figure 9A). Second, expression ratios of fi ltrin/nephrin and fi ltrin/podocin were calculated in order to adjust the levels of fi ltrin to the podocyte marker genes (Figure 9B). This confi rmed the reduced expression of fi ltrin mRNA in relation to the podocyte markers in proteinuric diseases, thus over- coming the possibility that the observed decrease was due to a reduced number of podocytes per glomerulus. In addition, a positive correlation was observed in the glomerular fi ltrin/GAPDH and nephrin/GAPDH val- ues (n = 63, r = 0.72, p = 2.9·10-11). Third, replication of the mRNA results was performed in a second set of samples, where the cDNAs were prepared from whole biopsies. The results were in agreement with the previous analysis in the case of FSGS (n = 11, p = 1.6·10-3) and also revealed a decrease in patients with diabetic nephropathy (n = 7, p = 5.1·10-3) compared to controls (n = 6). In MGN and MCD, the decrease in fi ltrin mRNA did not reach statistical signifi - cance. In spite of the general decrease in fi ltrin mRNA levels in proteinu- ric diseases, no correlation was observed with the expression of fi ltrin and the levels of proteinuria or serum creatinine at the time of biopsy.

5 Results 41 Figure 10. The exon-intron structure of the ACTN4 gene, with the chromosomal location of the genotyped SNPs. The marker-marker linkage disequilibrium (LD) plots are presented with r2 values (white = 0, 0 < grey < 1, black =1). Haplotype blocks of strong LD are indicated. LD, linkage disequilibrium (Study IV).

Collectively, these results indicate that the mRNA expression of fi ltrin is decreased in kidneys of patients with proteinuric kidney diseases. A moderate staining intensity of fi ltrin was observed in the glomeru- lus of patients with MCD (n = 4) and MGN (n = 6) as compared to controls (n = 3) in immunohistochemical examination. In MCD, fi ltrin was also detected in the mesangium, whereas in MGN, the glomerular staining of fi ltrin appeared granular. A marked decrease in the glomerular staining of fi ltrin characterized patients with FSGS (n = 5), ranging from segmen- tal to global (Figure 6).

42 5 Results 5.4. Genetic association analysis of LRRC7, KIRREL, NPHS2, KIRREL2, and ACTN4 with diabetic nephropathy (Study IV)

To investigate the genetic association of selected nephrin-associated genes with nephropathy in type 1 diabetes, a genetic association analysis was performed in a cross-sectional case-control study setting. A set of 1103 Finnish type 1 diabetic patients was included into the fi rst analysis, and the patients were classifi ed as having normoalbuminuria (n = 459), microalbuminuria (n = 276), or macroalbuminuria (n = 368). In order to confi rm the results of the fi rst analysis, an analysis of patients with nor- moalbuminuria (n = 607) and macroalbuminuria (n = 158) was performed for the selected SNPs at the second stage of analysis. In addition, a set of patients with ESRD (n = 260) was included. In the fi rst screening, a total of 45 SNPs in the LRRC7, KIRREL, NPHS2, KIRREL2, and ACTN4 genes were genotyped for the association analysis (Study IV). Genotype distributions were in Hardy-Weinberg equilibrium (HWE) in all SNPs except in the LRRC7 SNP rs984950 (HWE p-value in pa- tients with normoalbuminuria 2.2·10-6), indicating that the results of this SNP should be interpreted with caution. Statistical analysis revealed that no differences were evident in genotype, allele, or haplotype frequen- cies between patients with normal albumin excretion rate and diabetic nephropathy. The independent association of the SNPs was evaluated by adjusting the genetic effect with duration of type 1 diabetes, systolic and diastolic blood pressure, and HbA1c. However, this did not change the results. Neither were the polymorphisms associated with the albumin excretion rate, serum creatinine level nor time from the onset of type 1 diabetes to the development of diabetic nephropathy, when these vari- ables were used as quantitative variables in the analysis instead of the categorical nephropathy status. Notably, this analysis included also pa- tients with microalbuminuria. Since male gender has been considered as a risk factor for diabetic nephropathy in type 1 diabetes [124], males and females were also ana- lyzed separately. The gender-specifi c analysis yielded modest genotype and allelic associations with diabetic nephropathy for the SNPs rs979972 (p = 0.053), rs6508813 (p = 0.025), rs749701 (p = 0.008) and rs1060186 (p = 0.025) in the ACTN4 gene in females. Figure 10 describes the exon-intron structure of the ACTN4 SNPs with the location of the genotyped SNPs. In- terestingly, high pair-wise linkage disequilibrium (r2 > 0.8) was measured for the ACTN4 SNPs rs6508813-rs1060186. The four SNPs in the ACTN4 gene reaching the statistical signifi cance level 0.05 were selected for the second stage of analysis. At the second stage, a joint analysis of the genotypes from the fi rst and second set of patients was performed. The effect of stratifi cation by gender was investigated, since gender-specifi c effects were observed

5 Results 43 already in the fi rst sample set. The SNPs rs979972 and rs749701 were as- sociated with diabetic nephropathy and ESRD in females (Table 6). In this analysis, the genetic effect was adjusted for duration of type 1 diabetes, systolic and diastolic blood pressure, and HbA1c. For the SNPs rs979972 and rs749701, p-values below 0.05 were detected in the Hardy-Weinberg equilibrium test in macroalbuminuric patients (Table 6). Interestingly, in the joint analysis of patients, the SNPs rs979972 and rs749701 yielded nominally signifi cant association with ESRD also in males, when the pa- tients with normoalbuminuria and ESRD were compared (Table 6).

44 5 Results Table 6. Four ACTN4 SNPs with evidence of an association with diabetic nephropathy. The patients of the fi rst and second stage of analysis were combined, minor allele frequencies are presented sepa- rately for male and female patients with the corresponding adjusted odds ratios (OR) and p-values.

SNP HWE MAFa Diabetic nephropathy ESRD

p-value ORb (95% CI) p-value ORb (95% CI) p-value

Males rs979972 Normo 0.82 0.52 0.74 (0.48 - 1.15) 0.183 0.42 (0.21 - 0.85) 0.015 Macro 0.96 0.51 ESRD 0.09 0.47

rs6508813 Normo 0.55 0.27 0.64 (0.34 - 1.18) 0.153 0.60 (0.23 - 1.45) 0.244 Macro 0.70 0.24 ESRD 0.33 0.27

rs749701 Normo 0.44 0.44 0.65 (0.41 - 1.01) 0.057 0.43 (0.21 - 0.88) 0.021 Macro 0.69 0.40 ESRD 0.11 0.40

rs1060186 Normo 0.46 0.26 0.63 (0.34 - 1.19) 0.157 0.59 (0.24 - 1.47) 0.258 Macro 0.60 0.24 ESRD 0.45 0.26

Females rs979972 Normo 0.56 0.51 1.81 (1.18 - 2.79) 0.007 2.26 (1.14 - 4.48) 0.020 Macro 0.03 0.56 ESRD 0.08 0.60

rs6508813 Normo 0.91 0.27 1.70 (0.98 - 2.94) 0.058 0.98 (0.37 - 2.59) 0.961 Macro 0.35 0.34 ESRD 0.47 0.25

rs749701 Normo 0.67 0.42 1.93 (1.25 - 2.96) 0.003 2.12 (1.05 - 4.28) 0.036 Macro 0.02 0.49 ESRD 0.63 0.48

rs1060186 Normo 0.54 0.26 1.73 (1.00 - 3.00) 0.052 0.86 (0.31 - 2.39) 0.766 Macro 0.33 0.34 ESRD 0.72 0.25

aMinor allele is defi ned as in Study IV. bThe ORs were adjusted for duration of type 1 diabetes, systolic and diastolic blood pressure, and HbA1c. Abbreviations: CI, confi dence interval; ESRD, end- stage renal disease; HWE, Hardy-Weinberg equilibrium; Macro, macroalbuminuria; MAF, minor allele frequency; Normo, normoalbuminuria; OR, odds ratio.

5 Results 45 6. DISCUSSION

The hypothesis of this thesis stated that novel, yet unknown, nephrin- associated molecules participate in maintaining the slit diaphragm func- tion with nephrin. This has been proved to be at least partially true. Novel nephrin homologues NEPH1 [53], NEPH2 [54], and fi ltrin (Studies I and III) have been identifi ed and localized at the slit diaphragm. In addi- tion, an endogenous nephrin antisense transcript Nphs1as was cloned in mouse (Study II), but to date no expression of Nphs1as has been detected in the kidney.

Filtrin emerges into the picture of the podocytes

Filtrin is a novel nephrin-like protein of the immunoglobulin superfamily (Study I) [56, 57]. This is the fi rst study to determine the endogenous fi ltrin at the protein level in the kidney. The colocalization of fi ltrin with synaptopodin, an established podocyte marker, in dual immunofl uores- cence microscopy of human kidney cortex together with the immunob- lotting and RT-PCR data of isolated glomeruli support the expression of fi ltrin in the podocytes (Studies I and III). Nephrin is a crucial protein of the slit diaphragm [32-34]. Therefore, the question arises, which data support the association of fi ltrin with the slit diaphragm complex? At the ultrastructural level, immunoelectron microscopy using fi ltrin-specifi c antibodies indicates that fi ltrin localizes at the slit diaphragm area (Study III). In addition, fi ltrin, similar to the other nephrin-like molecules, interacts with the slit diaphragm proteins podocin and ZO-1 [57, 60]. Collectively, these data suggest that fi ltrin may belong, together with nephrin, to the cell adhesion junction be- tween the podocyte foot processes. The nephrin-like proteins NEPH1, NEPH2 and fi ltrin all localize in the slit diaphragm, and experimental data suggest that NEPH1 and NEPH2 form heterodimers with nephrin [35, 37, 54]. However, the dimensions of the extracellular domain of fi ve tandem immunoglobulin like repeats of the nephrin-like proteins might not allow direct interactions extending over the slit between the lateral foot process membranes. Instead, they may interact with the proximal part of the nephrin and, thus, stabilize the slit diaphragm structure as suggested by Tryggvason and colleagues [143]. Two other slit diaphragm proteins, nephrin and FAT1, are likely to be the ones that extend and form trans-interactions [34, 65]. Importantly, information on the signalling function for NEPH1 is emerging, involving

46 6 Discussion dynamic interactions with podocin and ZO-1, and NEPH1 seems to regu- late podocyte gene expression [29].

Function of fi ltrin in other tissues

The expression profi ling of human tissues suggested that fi ltrin-specifi c mRNA is predominantly expressed in the pancreas and lymph nodes (Study I). The expression in the pancreas has been replicated in later studies, and immunohistochemical data suggest localization in the β-cells of the islets of Langerhans [56, 193]. Recently published data on patients with type 1 diabetes indicates that anti-fi ltrin autoantibodies can be detected in the serum in approximately 10% of the patients during a follow-up time of 10 years after the diagnosis of type 1 diabetes [194]. This suggests that fi ltrin is involved in the autoimmune destruction of pancreatic β-cell. In the case of nephrin and NEPH1, intravenous injection of the antibodies in experimental models results in proteinuria [25, 58], and it is tempting to speculate that fi ltrin in the podocytes could act as an autoantigen in patients with type 1 diabetes and ultimately predispose to diabetic neph- ropathy. However, no evidence on the association of anti-fi ltrin autoan- tibodies with diabetic nephropathy was observed in the clinical study by Rinta-Valkama and colleagues [194]. In the brain, fi ltrin (NEPH3) appears transiently in the post-mitotic neural precursor cells, where it accumulates at the cellular processes and ZO-1 positive adherens junctions [195]. Confi rming the observation, fi ltrin localizes at the cell contacts in an adherens junction model in vitro. No- tably, fi ltrin proteins were shown to form homophilic trans-interactions with each other [195]. In olfactory sensory neurons, fi ltrin and its homo- logue NEPH2 act as homophilic adhesive molecules, and, interestingly, these proteins are expressed in a mutually exclusive pattern in the bun- dle of axons [196]. In these cells, fi ltrin presumably participates in axon guidance [196]. The podocytes have been described as neuron-like cells due to their characteristic morphology, and these cell types share some similarities at the molecular level [10]. However, it should be noted that the properties, especially the gap of 25 nm [197], of a typical adherens junction are not similar to the slit diaphragm. Thus the above mentioned results cannot be directly extrapolated to the mature slit diaphragm, where the lateral juxtaposed foot process membranes are separated by a distance of 30-40 nm [15].

Expression of nephrin and novel regulatory aspects

Expression of nephrin outside the kidney has been reported. In humans, nephrin has been associated with the islets of Langerhans in the pancreas

6 Discussion 47 [193, 198] and the lymphoid tissue [199], but the fi ndings remain contra- dictory (as reported in [200]) and may suffer from methodological limita- tions. In the mouse, the expression pattern is also promiscuous including the pancreas and distinct parts of the brain [201]. In line with the expres- sion profi le in humans, the nephrin mRNA was also successfully amplifi ed in the lymphoid tissue in mice (Study II). Distinct promoter regions and alternatively used upstream exons may partly explain the tissue- and cell type-specifi c expression of nephrin [202-205]. Mechanistically, the transcription factors WT1, PPAR-gamma, and retinoic acid receptors participate in the regulation of nephrin at the promoter level [206, 207]. In addition, an alternatively spliced variant of NPHS1 mRNA has been isolated from the kidney [32]. The cloning of the endogenous antisense transcript of nephrin, Nphs1as, revealed yet another regulatory aspect of nephrin. Nphs1as is expressed in the brain, lymphoid tissue, and embryonic stem cells in the mouse (Study II), but, in contrast to the hypothesis, Nphs1as does not seem to be transcribed in the kidney and thus does not participate in the regulation of the slit dia- phragm in the podocytes. It is reasonable to question, if Nphs1as mRNA is truly transcribed from the nephrin encoding locus in the mouse. How- ever, sequence alignments of the isolated cDNA with the genomic DNA sequence (Study II) and the incorporation of the TRAP vector in exon 8 in the nephrin-defi cient mouse model [190] indicate that Nphs1as is tran- scribed from the nephrin locus. The expression of nephrin mRNA in em- bryonic stem cells has recently been confi rmed by independent research groups [208, 209]. The results presented here do not allow direct conclusions about the function of Nphs1as, but speculations can be presented based on previous knowledge of endogenous antisense transcripts: A number of functions, including transcriptional interference, RNA masking and double-stranded RNA-dependent mechanisms, has been attributed to endogenously ex- pressed antisense transcripts (reviewed in [210]). As discussed by Lapidot and colleagues [210], the relationship between the expression pattern of Nphs1as and its target Nphs1 might give certain indications regard- ing the mode of action. Nephrin is expressed in the brain in a cell-type specifi c manner [191, 201], and both the Nphs1as and Nphs1 mRNA are expressed in reasonable amounts (Study II). Thus, it can be proposed that the function of Nphs1as in the brain is to stabilize the Nphs1 mRNA tran- script levels, but, of course, other mechanisms cannot be excluded. To my knowledge, the human counterpart of the nephrin antisense transcript has not been isolated.

48 6 Discussion Chromosomal organisation of KIRREL2 and NPHS1 in the human genome

The human fi ltrin gene KIRREL2 resides on chromosome 19q13.12 ad- jacent to NPHS1 which encodes nephrin, and the distance between the transcription initiation sites is approximately 5 kb in human (Study I; Ta- ble 5) and 3 kb in mouse and rat (data not shown). The inserted repeti- tive DNA elements in the human DNA sequence may explain the differ- ence in the distance of the intervening region. It has been demonstrated that the regulatory elements required for the endogenous Nphs1 expres- sion pattern in the mouse are located 5.4 kb or even 8.3 kb upstream of the transcription initiation site [202]. Moreover, the region 1.25 kb upstream is shown to direct podocyte-specifi c expression both in human and mouse [203, 204]. Overlapping expression sites of nephrin and fi ltrin exist, particularly in the podocytes and the pancreatic beta-cells. Thus, the regulatory regions of NPHS1 and KIRREL2, although read in opposite directions, may overlap, suggesting the possibility of transcriptional co- regulation. The head-to-head orientation of genes is a rather common phenome- non, but the mechanisms of regulation of bi-directional genes are poorly known [211]. Systematic analysis of gene expression databases has re- vealed that many bi-directional genes are coregulated but some, instead, are antiregulated [211]. In addition, transient transfection assays for a randomly chosen set of genes suggested that most bi-directional pro- moters share gene regulatory elements that control both genes [211]. The TATA-element of the proximal promoter region is suggested to con- tribute to the directionality of transcription and, therefore, bi-directional promoters contain very rarely TATA-boxes [211]. Interestingly, neither the NPHS1 nor KIRREL2 promoter has a TATA-box in their proximal promoter area (Study I). Gene pairs coding for homologous proteins are uncommon. An exam- ple of such a case is provided by the genes SAFB1 and SAFB2, which is a homolous gene pair arranged in a bi-directional divergent confi guration in 19p13.3 [212]. The genes are separated by a rather short intervening region of 500 bp and share similar functions as oestrogen receptors. They are widely co-expressed in human tissues, but the translated proteins have somewhat differing subcellular localizations [212]. On human , a conserved gene arrangement of orthol- ogous genes in mouse and human has been described [213, 214]. Nota- bly, one third of genes on chromosome 19 has been observed to belong to clustered gene families, possibly due to duplication of ancestral genes [214]. It appears that the nephrin and the nephrin-like genes are not confi ned to a single mammalian species, suggesting that chromosomal arrangements or duplications have occurred prior to the divergence of mammalian ancestors.

6 Discussion 49 Patients with proteinuria (Studies III and IV)

This study utilized several patient groups in elucidating the pathogenetic mechanism in proteinuria. Two sample sets were collected using a multi- central approach. First, the Finnish FinnDiane cohort focusing on com- plications in type 1 diabetes, particularly diabetic nephropathy, was uti- lized for the genetic association analysis of the nephrin-associated genes (Study IV). Second, the European Renal cDNA bank, a multinational ini- tiative to investigate the functional genetics of proteinuric renal diseases (participating centres listed in Appendix) was used for gene expression analysis of fi ltrin and the podocyte marker genes (Study III). The other two sample collections were derived from a single hospital in Fitzroy, Australia and in Milan, Italy, and were examined in Study III. All clinical data were analysed in a cross-sectional setting (studies III and IV), which is a powerful study setting, when the phenotypes are well- defi ned and the number of samples provides suffi cient statistical power. Ideally, the study setting would be a prospective one having follow-up data available after the time of collecting the samples. It is also note- worthy that kidney biopsies usually represent a rather established stage of the disease, and may thus provide limited information on the early phases of the disease. Therefore, experimental animal models will be crucial to gain insight into the time perspective of the variation in the expression of fi ltrin and in order to exclude the effect of different envi- ronmental effects.

Is fi ltrin merely a by-stander in the podocyte or is it involved in podocyte injury?

Mutations in the genes encoding nephrin (NPHS1) and NEPH1 (KIRREL) seem to trigger congenital nephrosis, and these molecules are indispensa- ble for the maintenance of the slit diaphragm. We tested the hypothesis, whether fi ltrin is involved in the primary glomerular diseases including minimal change disease, focal segmental glomerulosclerosis and mem- branous glomerulopathy (Study III). The mRNA levels of fi ltrin, nephrin, podocin and GAPDH were measured in kidney biopsies. GAPDH was used for standard normalization and nephrin and podocin were considered established podocyte marker genes. Two set of patients (ERCB and the St. Vincent Hospital) were analyzed. Collectively, the results indicate that fi ltrin is down-regulated in patients with primary glomerular disease in relation to GAPDH and also to the podocyte marker gene nephrin and podocin. The sample size was limited but still suffi cient to detect a statis- tically signifi cant difference in expression values in all diagnostic groups in the ERCB samples, whereas in the second set of samples from the St. Vincent’s Hospital the low number of patients per diagnostic group and

50 6 Discussion high variability in the expression values did not allow drawing defi nitive conclusions in all disease categories (Study III). A statistically signifi cant, positive correlation (p ≈ 2.9·10-11) was observed in the expression values of fi ltrin and nephrin. This result gives additional support for the view that the gene expression in the podocyte is orchestrated potentially via common regulatory pathways. In addition, it is particularly interesting due to the intriguing chromosomal arrangement of the nephrin and fi l- trin genes NPHS1 and KIRREL2. Altered distribution of fi ltrin in the glomerulus was observed in pa- tients with glomerular diseases, especially in FSGS (Study III), which sup- ports the results of the gene expression analysis. Can the nephrin-like proteins mutually compensate each other’s function? It is known that the nephrin-like proteins are only homologous in their extracellular domains but not in the intracellular domain (except the podocin-interacting do- main) [57], and these structural differences may indicate that these pro- teins do not have identical signal transductions mechanisms. The measurement of mRNA levels in kidney samples does not perfectly represent the gene expression activity. It rather refl ects the steady-state level of mRNA turnover, where degradation, modifi cations and transcrip- tion have all their own effects. As for the podocytes, altered podocyte number due to their detachment from the basement membrane or pos- sibly via apoptotic mechanisms may reduce the proportion of podocytes per glomerulus as reported in FSGS and diabetic nephropathy [129, 215]. Another concern is the number of glomeruli per one kidney biopsy. These facts may cause a bias in the expression analysis. In this study, the concept of ‘in silico’ normalization according to the podocyte marker genes nephrin and podocin, originally introduced by the group of Dr. Matthias Kretzler [47], was successfully adapted for the expression analysis of fi ltrin (Study III). The results indicated that fi ltrin mRNA is down-regulated beyond the regulation of nephrin and podocin mRNAs, consistent with an mRNA repression per podocyte and not only a consequence of podocyte loss. This allowed us to overcome - at least partially - the concerns mentioned above. Furthermore, the effect of dis- secting the kidney biopsies into distinct glomerular and tubulointerstitial compartments (described in [177]) was also evaluated in this study. The expression of fi ltrin in microdissected glomeruli was compared to that measured from whole biopsies (Study III). It appeared that a trend of decreased expression was observed in both sample sets, irrespective of the mode of the preparation of the samples (Study III). Based on the knowledge of the pathogenesis of ‘podocyte diseases’, additional means of normalization should be taken into use (including the ‘in silico’ nor- malization) beyond the normalization in the house-keeping genes like GAPDH, which has been a golden standard in renal research.

6 Discussion 51 Can the genes behind monogenic proteinuric diseases contribute to the understanding of diabetic nephropathy?

Most of the current understanding of common diseases derives from the elucidation of the molecular background of rare familial forms of these common traits (discussed in [216]). The genes underlying the he- reditary proteinuric diseases are expressed in the podocytes and include several slit diaphragm associated genes (chapter 2.5). In this study, we tested the hypothesis, whether the genes encoding fi ltrin, NEPH1, po- docin, α-actinin-4 and densin-180 act as susceptibility genes for diabetic nephropathy (Study IV). The kidney complication of diabetes manifests fi rst with microalbuminuria [127], which suggests that the glomerular capillary wall, responsible for retaining albumin within circulation, is af- fected. The selection of the population is critical for the validity of the re- sults of association analysis in complex diseases. The homogeneity and structure of the Finnish population is defi nitely an asset for performing genetic studies (reviewed in [217]). The Finnish population has a charac- teristic disease heritage, and many of these diseases have a Mendelian pattern of inheritance. Several causative genes have been successfully identifi ed, including NPHS1 [27]. The population has also proved useful for complex disease genetic mapping, as illustrated by the mapping of susceptibility gene for asthma [218]. A recent population based study supports the involvement of a genetic component in diabetic nephropa- thy in the Finnish population [163]. In addition, Finland has the highest incidence rate of type 1 diabetes in the world [219]. For the genes ACTN4 and NPHS2, tag SNPs were selected based on the HapMap data in the CEPH (Centre d’Etude Polymorphism Humain) popu- lation. This approach assumes that the information content of the tag SNPs in the CEPH population is transferable to the Finnish population. The issue of tag transferability was recently examined by de Bakker and colleagues [220]. They estimated the correlation (r2) in the genotype data of a set of candidate genes between the CEPH population and the Botnia sample from Finland, and the results indicate that it is justifi ed to utilize the HapMap data for selection of tag SNPs for genotyping in the Finnish population without compromising the statistical power. A spacing ap- proach (with an interval of 6-14 kb) was employed in the selection of the genotyped SNPs for KIRREL and KIRREL2. The HapMap data were utilized also in this selection, but not all tag SNPs were included in genotyping due to the large number of SNPs that the complete covering of these relative small genes would have required. In the case of LRRC7, only a few SNPs were genotyped, so it can only be considered as a pilot study. The results of the genetic association analysis suggest that poly- morphisms in the ACTN4 gene, encoding α-actinin-4, may be involved

52 6 Discussion in susceptibility to diabetic nephropathy in type 1 diabetes (Study IV). However, the link between ACTN4 and diabetic nephropathy was not straightforward, but a genetic interaction of the gender and the ACTN4 polymorphisms was observed. The results were supported by the fi nding that the same variants also seemed to associate with cardiovascular dis- ease (Study IV). Furthermore, the genotype distribution of the polymor- phisms showed a deviation from the Hardy-Weinberg equilibrium in the female patients with macroalbuminuria (Study IV), which may indicate a true genetic association as proposed by Nielsen and colleagues [221]. α-actinin-4 is an actin-fi lament cross-linking protein located in the proximity of the slit diaphragm in podocyte foot processes [82, 83]. The structure and function of the cytoskeleton are of pivotal importance for maintaining the cell shape of the podocytes, and disturbances of the po- docyte cytoskeleton have been detected in the presence of high glucose both in vitro and in vivo [222]. Interestingly, α-actinin-4 is down-regu- lated in kidney biopsies of diabetic patients with microalbuminuria as compared to controls [223], and its expression is also disturbed in cul- tured cells by diabetes-related pathogenetic factors including advanced glycation end-products [224]. The connection between ACTN4 and nephropathy was fi rst discov- ered in patients with familial focal segmental glomerulosclerosis (FSGS) carrying mutations in the ACTN4 gene [83]. This disease belongs to the ‘late-onset’ group according to the classifi cation by Kriz [142]. Regarding diabetic nephropathy, the complication develops gradually and the inci- dence is at its highest after about 15 years’ duration of type 1 diabetes [124]. The results support the hypothesis that a gene previously linked to a rare monogenic type of nephropathy is also a good candidate for diabetic nephropathy, a more common type of nephropathy. The statistical power was estimated for the genetic association analy- sis [225]. While the fi rst set of patients alone remained underpowered to detect association with markers with low minor allele frequencies, the additional genotyping and the joint analysis of the patients at the sec- ond stage of analysis increased the statistical power remarkably (Study IV). Low minor allele frequency (MAF) for certain genotyped SNPs may limit the interpretation of the results. Especially in KIRREL2, all SNPs had MAF under 0.05 in patients with a normal albumin excretion rate (Study IV). Therefore, direct sequencing of the KIRREL2 locus on chromosome 19q13.12 to detect population-specifi c polymorphisms may be required. Furthermore, we cannot exclude the existence of rare variants in some cases of diabetic nephropathy. It can be speculated that the mutation phenotype of the genes of mo- nogenic proteinuria with a congenital onset of the disease, e.g. NPHS1 and NPHS2, would not perfectly fi t with the insidious and slow patho- genesis of diabetic nephropathy. On the other hand, the negative re- sults of the association analysis might also arise from to the fact that the

6 Discussion 53 genotyped SNPs covered only the coding region and the close vicinity of the genes under analysis. This approach assumed that causative muta- tions locate in the coding, splicing site and/or regulatory regions in the proximity of the analysed gene. The causative variants may also reside over long distances from the coding region as exemplifi ed by the case of adult-type hypolactasia [226], and this possibility has to be taken into account when interpreting these results. Even if the genotyped SNPs in the fi ltrin gene KIRREL2 were not associ- ated with diabetic nephropathy, it appears that fi ltrin is down-regulated at the mRNA level in kidney biopsies of patients with diabetic nephropa- thy (Study III). This seems to be the case with the nephrin gene NPHS1: NPHS1 has been excluded as a susceptibility gene for diabetic nephropa- thy [227], but nephrin still seems to contribute to the pathogenesis of diabetic nephropathy. Decreased levels of nephrin are often observed in the glomerulus, and this seems to be mediated in an angiotensin II de- pendent fashion [49, 51]. In addition, a similar down-regulation of fi ltrin mRNA was detected in another secondary glomerular disease hyper- tensive nephropathy (Study III). Interestingly, diabetic and hypertensive nephropathy may have common predisposing factors and pathogenetic mechanisms [98].

54 6 Discussion Concluding remarks

Since the cloning of the nephrin gene NPHS1, novel nephrin-associated molecules of pivotal importance for the slit diaphragm have been discov- ered. This thesis adds two potentially interesting molecules to this list, and addresses their involvement in human glomerular diseases. It is of note that the term ‘nephrin-associated’ was interpreted in a quite broad sense in this thesis, including both nephrin-like and nephrin-interacting molecules. This thesis employed a bioinformatics approach for fi shing novel homologous genes and proceeded to assess the signifi cance of the molecules in pathological conditions by analysing human sample collec- tions. This approach proved successful. Two novel candidates molecules, fi ltrin and Nphs1as, were discovered, and moreover, they are both tran- scribed from the nephrin locus or its proximity. Filtrin is a novel podocyte-expressed protein with localization at the slit diaphragm, and the expression of fi ltrin is down-regulated in human proteinuric diseases. However, several aspects of fi ltrin still deserve to be explored. Molecules of the immunoglobulin superfamily tend to form both homo- and heterodimers, and, therefore, it will be interesting to defi ne the molecular interactions of fi ltrin at the slit diaphragm - es- pecially with nephrin and the nephrin-like proteins NEPH1 and NEPH2. Regarding the fi ltrin gene KIRREL2, an intriguing question is also the potential transcriptional co-regulation of KIRREL2 and NPHS1 in physi- ological and pathological conditions. Whether or not these genes share common regulatory elements in their intergenic region remains to be the elucidated. This thesis reports the discovery of Nphs1as, an endogenous antisense transcript of the nephrin gene, but this line of research was not continued due to the apparent lack of expression in the mouse kidney. Nevertheless, this does not rule out the possibility that Nphs1as could be biologically important. The possible human orthologue has not been cloned, and it is also tempting to speculate that de novo expression of the antisense tran- script would emerge in the kidney under proteinuric conditions, which would provide an unprecedented view in the role of nephrin in glomeru- lar diseases. It will be important to determine at which level (transcrip- tion, pre-mRNA splicing, or translation) the possible interaction between Nphs1 and Nphs1as mRNAs occurs. These experiments may indicate novel regulatory functions of Nphs1as. A candidate gene approach was selected to evaluate the genetic fac- tors in diabetic nephropathy. It seems that common variants in ACTN4 contribute to genetic susceptibility for diabetic nephropathy in type 1 di- abetes. The results suggest that a gene previously linked to a monogenic focal segmental glomerulosclerosis (FSGS) may confer susceptibility to a secondary form of nephropathy, namely diabetic nephropathy. Replica- tion of the fi nding in other populations and fi ne mapping of the gene

6 Discussion 55 locus will be required. It is also worth examining the role of two other nephrin-associated genes TRPC6 and CD2AP in diabetic nephropathy. No- tably, these genes have previously been linked to FSGS. In conclusion, it seems that novel nephrin-like and nephrin-associated molecules are crucial for maintaining the function of the glomerular fi ltration barrier. The fi ndings presented in this thesis provide novel in- sights into the molecular basis of human glomerular diseases and extend the understanding of the molecular anatomy of the slit diaphragm.

56 6 Discussion 7. ACKNOWLEDGEMENTS

This work was carried out at the Department of Bacteriology and Im- munology, University of Helsinki and at the Folkhälsan Research Center in Biomedicum Helsinki during the years 2001 – 2007. I am grateful to the previous and present heads of both the institutes, Professors Risto Renkonen, Seppo Meri, Per-Henrik Groop and Anna-Elina Lehesjoki, for providing excellent research facilities. I wish to thank Professor Carl G. Gahmberg for guiding me through the fi nal metres of my Ph.D. studies and for accepting the role of custos in my defence. Professor Markku Laakso and Docent Eero Lehtonen are warmly acknowledged for reviewing this thesis and for their constructive criticism. I want to express my gratitude to Docent Harry Holthöfer for intro- ducing me to the fascinating world of podocyte research. His encourage- ment during the thesis work and never-ending enthusiasm will not be forgotten. I am most grateful to Docent Per-Henrik Groop. He has taught me the essence of science and created a pleasant and supportive atmos- phere to work at. My special thanks belong to Docent Maija Wessman. I am grateful to Maija for all for her time and effort that she has unself- ishly devoted to my research and this Ph.D. thesis. My collaborators and co-authors have been numerous. It has been a pleasure to work with Tuula Palmén, Heikki Ahola, Johanna Rinta- Valkama, Petra Mai, Eva Åström, Thu Trang Pham, Thomas Floss, Holger Schmid, Maria Pia Rastaldi, Deborah Mattinzoli, Robyn G. Langham, Pauliina Luimula, Riika Kilpikari, Markus Lassila, Richard E. Gilbert, Dontscho Kerjaschki, Matthias Kretzler, Mari Kaunisto, Riika Kilpikari, Maija Parkkonen, and Carol Forsblom. Their contribution has made the publications and this thesis possible. Thank you! It has been a privilege to participate in the work in two research groups. All the previous and present members of the Glomerular Biology Study Group and the FinnDiane Study Group are acknowledged for their help and support. We have shared the excitement of discovery and some- times depressing moments together. Especially, I thank three excellent scientists, Docent Markus Lassila, Sanna Lehtonen, and Carol Forsblom, for always fi nding time for inspiring discussions about science and for their valuable advice regarding practical problems. This work would not have been possible without the FinnDiane pa- tients and other people who have participated in the various studies. Their involvement in the projects – without expecting any personal gain – is acknowledged.

7 Acknowledgements 57 Sami Ruotsalainen is thanked for his skilful help with the image processing of this thesis. I want to thank all my friends for all their support and for keeping me in touch with the world surrounding Biomedicum. I owe my warm- est thanks my family, my parents Tuula and Erkki Ihalmo, my sister Teija Sirko, brother Matti Ihalmo, and their families, for their loving care and endless support. This work was fi nancially supported by the European Commission and numerous private foundations. The personal grants from the Finnish Cultural Foundation, the Federation of European Biochemical Societies (FEBS), the Kyllikki and Uolevi Lehikoinen Foundation, and the Finnish Kidney Foundation are gratefully acknowledged.

Helsinki, July 2007 Pekka Ihalmo

58 7 Acknowledgements 8. APPENDIX

Members of the European Renal cDNA Bank at the time of the study:

C. Cohen, H. Schmid, M. Kretzler, and D. Schlöndorff, Munich, Germany; P. Ronco and J.D. Sraer, Paris, France; M.P. Rastaldi and G. D’Amico, Mi- lan, Italy; F. Mampaso, Madrid, Spain; P. Doran and H.R. Brady, Dublin, Ireland; D. Mönks and C. Wanner, Würzburg, Germany; A.J. Rees and P. Brown, Aberdeen, Scotland; F. Strutz and G. Müller, Göttingen, Germany; P. Mertens and J. Floege, Aachen, Germany; N. Braun and T. Risler, Tü- bingen, Germany; L. Gesualdo and F.P. Schena, Bari, Italy; J. Gerth and G. Wolf, Jena, Germany; R. Oberbauer and D. Kerjaschki, Vienna, Austria; B. Banas and B.K. Krämer, Regensburg, Germany; W. Samtleben, Munich, Germany; H. Peters and H.H. Neumayer, Berlin, Germany; K. Ivens and B. Grabensee, Düsseldorf, Germany; and V. Tesar, Prague, Italy.

8 Appendix 59 9. REFERENCES

1. Tisher CC, Madsen KM. Anatomy of the Kidney. In: Brenner BM, Rector FC, editors. The Kidney. Philadelphia, PA, USA. W. B. Sounders Company, 1991. p. 3-75. 2. Hjalmarsson C, Johansson BR, Haraldsson B. Electron microscopic evaluation of the endothelial surface layer of glomerular capillaries. Microvasc Res 67:9-17, 2004. 3. Kerjaschki D, Sharkey DJ, Farquhar MG. Identifi cation and characterization of podocalyxin--the major sialoprotein of the renal glomerular epithelial cell. J Cell Biol 98:1591-1596, 1984. 4. Rostgaard J, Qvortrup K. Electron microscopic demonstrations of fi lamentous molecular sieve plugs in capillary fenestrae. Microvasc Res 53:1-13, 1997. 5. Kriz W, Elger M, Lemley K, Sakai T. Structure of the glomerular mesangium: a biomechanical interpretation. Kidney Int Suppl 30:S2-9, 1990. 6. Schlondorff D. Roles of the mesangium in glomerular function. Kidney Int 49:1583-1585, 1996. 7. Mene P, Cinotti GA, Pugliese F. Signal transduction in mesangial cells. J Am Soc Nephrol 2:S100-106, 1992. 8. Miner JH. Renal basement membrane components. Kidney Int 56:2016-2024, 1999. 9. Groffen AJ, Veerkamp JH, Monnens LA, van den Heuvel LP. Recent insights into the structure and functions of heparan sulfate proteoglycans in the hu- man glomerular basement membrane. Nephrol Dial Transplant 14:2119-2129, 1999. 10. Pavenstadt H, Kriz W, Kretzler M. Cell biology of the glomerular podocyte. Physiol Rev 83:253-307., 2003. 11. Adler S, Chen X. Anti-Fx1A antibody recognizes a beta 1-integrin on glomeru- lar epithelial cells and inhibits adhesion and growth. Am J Physiol 262:F770- 776, 1992. 12. Korhonen M, Ylanne J, Laitinen L, Virtanen I. The alpha 1-alpha 6 subunits of integrins are characteristically expressed in distinct segments of developing and adult human nephron. J Cell Biol 111:1245-1254, 1990. 13. Raats CJ, van den Born J, Bakker MA, Oppers-Walgreen B, Pisa BJ, Dijkman HB, Assmann KJ, Berden JH. Expression of agrin, dystroglycan, and utrophin in normal renal tissue and in experimental glomerulopathies. Am J Pathol 156:1749-1765, 2000. 14. Ichimura K, Kurihara H, Sakai T. Actin fi lament organization of foot processes in rat podocytes. J Histochem Cytochem 51:1589-1600, 2003. 15. Rodewald R, Karnovsky MJ. Porous substructure of the glomerular slit diaphragm in the rat and mouse. J Cell Biol 60:423-433, 1974.

60 9 References 16. Wartiovaara J, Ofverstedt LG, Khoshnoodi J, Zhang J, Makela E, Sandin S, Ruotsalainen V, Cheng RH, Jalanko H, Skoglund U, Tryggvason K. Nephrin strands contribute to a porous slit diaphragm scaffold as revealed by electron tomography. J Clin Invest 114:1475-1483, 2004. 17. Schnabel E, Anderson JM, Farquhar MG. The tight junction protein ZO-1 is concentrated along slit diaphragms of the glomerular epithelium. J Cell Biol 111:1255-1263, 1990. 18. Kerjaschki D. Caught fl at-footed: podocyte damage and the molecular bases of focal glomerulosclerosis. J Clin Invest 108:1583-1587., 2001. 19. Reiser J, Kriz W, Kretzler M, Mundel P. The glomerular slit diaphragm is a modifi ed adherens junction. J Am Soc Nephrol 11:1-8, 2000. 20. Farquhar MG. The glomerular basement membrane: not gone, just forgotten. J Clin Invest 116:2090-2093, 2006. 21. Rennke HG, Cotran RS, Venkatachalam MA. Role of molecular charge in glomerular permeability. Tracer studies with cationized ferritins. J Cell Biol 67:638-646, 1975. 22. Smithies O. Why the kidney glomerulus does not clog: a gel permeation/ diffusion hypothesis of renal function. Proc Natl Acad Sci U S A 100:4108-4113, 2003. 23. Deen WM, Lazzara MJ, Myers BD. Structural determinants of glomerular permeability. Am J Physiol Renal Physiol 281:F579-596., 2001. 24. Deen WM. What determines glomerular capillary permeability? J Clin Invest 114:1412-1414, 2004. 25. Orikasa M, Matsui K, Oite T, Shimizu F. Massive proteinuria induced in rats by a single intravenous injection of a monoclonal antibody. J Immunol 141:807- 814, 1988. 26. Topham PS, Kawachi H, Haydar SA, Chugh S, Addona TA, Charron KB, Holz- man LB, Shia M, Shimizu F, Salant DJ. Nephritogenic mAb 5-1-6 is directed at the extracellular domain of rat nephrin. J Clin Invest 104:1559-1566, 1999. 27. Kestila M, Lenkkeri U, Mannikko M, Lamerdin J, McCready P, Putaala H, Ruot- salainen V, Morita T, Nissinen M, Herva R, Kashtan CE, Peltonen L, Holmberg C, Olsen A, Tryggvason K. Positionally cloned gene for a novel glomerular protein--nephrin--is mutated in congenital nephrotic syndrome. Mol Cell 1:575-582, 1998. 28. Schwarz K, Simons M, Reiser J, Saleem MA, Faul C, Kriz W, Shaw AS, Holzman LB, Mundel P. Podocin, a raft-associated component of the glomerular slit diaphragm, interacts with CD2AP and nephrin. J Clin Invest 108:1621-1629, 2001. 29. Benzing T. Signaling at the slit diaphragm. J Am Soc Nephrol 15:1382-1391, 2004. 30. Moeller MJ. Dynamics at the slit diaphragm--is nephrin actin’? Nephrol Dial Transplant 22:37-39, 2007. 31. Tryggvason K, Pikkarainen T, Patrakka J. Nck links nephrin to actin in kidney podocytes. Cell 125:221-224, 2006. 32. Holthofer H, Ahola H, Solin ML, Wang S, Palmen T, Luimula P, Miettinen

9 References 61 A, Kerjaschki D. Nephrin localizes at the podocyte fi ltration slit area and is characteristically spliced in the human kidney. Am J Pathol 155:1681-1687, 1999. 33. Holzman LB, St John PL, Kovari IA, Verma R, Holthofer H, Abrahamson DR. Nephrin localizes to the slit pore of the glomerular epithelial cell. Kidney Int 56:1481-1491, 1999. 34. Ruotsalainen V, Ljungberg P, Wartiovaara J, Lenkkeri U, Kestila M, Jalanko H, Holmberg C, Tryggvason K. Nephrin is specifi cally located at the slit diaphragm of glomerular podocytes. Proc Natl Acad Sci U S A 96:7962-7967, 1999. 35. Gerke P, Huber TB, Sellin L, Benzing T, Walz G. Homodimerization and heterodimerization of the glomerular podocyte proteins nephrin and NEPH1. J Am Soc Nephrol 14:918-926, 2003. 36. Khoshnoodi J, Sigmundsson K, Ofverstedt LG, Skoglund U, Obrink B, Wartiovaara J, Tryggvason K. Nephrin promotes cell-cell adhesion through homophilic interactions. Am J Pathol 163:2337-2346, 2003. 37. Barletta GM, Kovari IA, Verma RK, Kerjaschki D, Holzman LB. Nephrin and Neph1 co-localize at the podocyte foot process intercellular junction and form cis hetero-oligomers. J Biol Chem 278:19266-19271, 2003. 38. Simons M, Schwarz K, Kriz W, Miettinen A, Reiser J, Mundel P, Holthofer H. Involvement of lipid rafts in nephrin phosphorylation and organization of the glomerular slit diaphragm. Am J Pathol 159:1069-1077, 2001. 39. Verma R, Kovari I, Soofi A, Nihalani D, Patrie K, Holzman LB. Nephrin ectodo- main engagement results in Src kinase activation, nephrin phosphorylation, Nck recruitment, and actin polymerization. J Clin Invest 116:1346-1359, 2006. 40. Verma R, Wharram B, Kovari I, Kunkel R, Nihalani D, Wary KK, Wiggins RC, Kil- len P, Holzman LB. Fyn binds to and phosphorylates the kidney slit diaphragm component Nephrin. J Biol Chem 278:20716-20723, 2003. 41. Li H, Lemay S, Aoudjit L, Kawachi H, Takano T. SRC-family kinase Fyn phospho- rylates the cytoplasmic domain of nephrin and modulates its interaction with podocin. J Am Soc Nephrol 15:3006-3015, 2004. 42. Huber TB, Hartleben B, Kim J, Schmidts M, Schermer B, Keil A, Egger L, Lecha RL, Borner C, Pavenstadt H, Shaw AS, Walz G, Benzing T. Nephrin and CD2AP associate with phosphoinositide 3-OH kinase and stimulate AKT-dependent signaling. Mol Cell Biol 23:4917-4928, 2003. 43. Jones N, Blasutig IM, Eremina V, Ruston JM, Bladt F, Li H, Huang H, Larose L, Li SS, Takano T, Quaggin SE, Pawson T. Nck adaptor proteins link nephrin to the actin cytoskeleton of kidney podocytes. Nature, 2006. 44. Huber TB, Kottgen M, Schilling B, Walz G, Benzing T. Interaction with podocin facilitates nephrin signaling. J Biol Chem 276:41543-41546, 2001. 45. Quack I, Rump LC, Gerke P, Walther I, Vinke T, Vonend O, Grunwald T, Sellin L. beta-Arrestin2 mediates nephrin endocytosis and impairs slit diaphragm integrity. Proc Natl Acad Sci U S A 103:14110-14115, 2006. 46. Koop K, Eikmans M, Baelde HJ, Kawachi H, De Heer E, Paul LC, Bruijn JA. Ex- pression of podocyte-associated molecules in acquired human kidney diseases. J Am Soc Nephrol 14:2063-2071, 2003.

62 9 References 47. Schmid H, Henger A, Cohen CD, Frach K, Grone HJ, Schlondorff D, Kretzler M. Gene expression profi les of podocyte-associated molecules as diagnostic mark- ers in acquired proteinuric diseases. J Am Soc Nephrol 14:2958-2966, 2003. 48. Luimula P, Ahola H, Wang SX, Solin ML, Aaltonen P, Tikkanen I, Kerjaschki D, Holthofer H. Nephrin in experimental glomerular disease. Kidney Int 58:1461- 1468., 2000. 49. Doublier S, Salvidio G, Lupia E, Ruotsalainen V, Verzola D, Deferrari G, Camussi G. Nephrin expression is reduced in human diabetic nephropathy: evidence for a distinct role for glycated albumin and angiotensin II. Diabetes 52:1023-1030, 2003. 50. Kelly DJ, Aaltonen P, Cox AJ, Rumble JR, Langham R, Panagiotopoulos S, Jerums G, Holthofer H, Gilbert RE. Expression of the slit-diaphragm protein, nephrin, in experimental diabetic nephropathy: differing effects of anti- proteinuric therapies. Nephrol Dial Transplant 17:1327-1332, 2002. 51. Langham RG, Kelly DJ, Cox AJ, Thomson NM, Holthofer H, Zaoui P, Pinel N, Cordonnier DJ, Gilbert RE. Proteinuria and the expression of the podocyte slit diaphragm protein, nephrin, in diabetic nephropathy: effects of angiotensin converting enzyme inhibition. Diabetologia 45:1572-1576, 2002. 52. Susztak K, Raff AC, Schiffer M, Bottinger EP. Glucose-induced reactive oxygen species cause apoptosis of podocytes and podocyte depletion at the onset of diabetic nephropathy. Diabetes 55:225-233, 2006. 53. Donoviel DB, Freed DD, Vogel H, Potter DG, Hawkins E, Barrish JP, Mathur BN, Turner CA, Geske R, Montgomery CA, Starbuck M, Brandt M, Gupta A, Ramirez-Solis R, Zambrowicz BP, Powell DR. Proteinuria and perinatal lethality in mice lacking NEPH1, a novel protein with homology to NEPHRIN. Mol Cell Biol 21:4829-4836, 2001. 54. Gerke P, Sellin L, Kretz O, Petraschka D, Zentgraf H, Benzing T, Walz G. NEPH2 is located at the glomerular slit diaphragm, interacts with nephrin and is cleaved from podocytes by metalloproteinases. J Am Soc Nephrol 16:1693- 1702, 2005. 55. Ihalmo P, Palmen T, Ahola H, Valtonen E, Holthofer H. Filtrin is a novel member of nephrin-like proteins. Biochem Biophys Res Commun 300:364-370, 2003. 56. Sun C, Kilburn D, Lukashin A, Crowell T, Gardner H, Brundiers R, Diefenbach B, Carulli JP. Kirrel2, a novel immunoglobulin superfamily gene expressed primarily in beta cells of the pancreatic islets. Genomics 82:130-142, 2003. 57. Sellin L, Huber TB, Gerke P, Quack I, Pavenstadt H, Walz G. NEPH1 defi nes a novel family of podocin interacting proteins. Faseb J 17:115-117, 2003. 58. Liu G, Kaw B, Kurfi s J, Rahmanuddin S, Kanwar YS, Chugh SS. Neph1 and nephrin interaction in the slit diaphragm is an important determinant of glomerular permeability. J Clin Invest 112:209-221, 2003. 59. Ueno H, Sakita-Ishikawa M, Morikawa Y, Nakano T, Kitamura T, Saito M. A stromal cell-derived membrane protein that supports hematopoietic stem cells. Nat Immunol 4:457-463, 2003. 60. Huber TB, Schmidts M, Gerke P, Schermer B, Zahn A, Hartleben B, Sellin L,

9 References 63 Walz G, Benzing T. The carboxyl terminus of Neph family members binds to the PDZ domain protein zonula occludens-1. J Biol Chem 278:13417-13421, 2003. 61. Cohen CD, Klingenhoff A, Boucherot A, Nitsche A, Henger A, Brunner B, Schmid H, Merkle M, Saleem MA, Koller KP, Werner T, Grone HJ, Nelson PJ, Kretzler M. Comparative promoter analysis allows de novo identifi cation of specialized cell junction-associated proteins. Proc Natl Acad Sci U S A, 2006. 62. Vincent PA, Xiao K, Buckley KM, Kowalczyk AP. VE-cadherin: adhesion at arm’s length. Am J Physiol Cell Physiol 286:C987-997, 2004. 63. Xu ZG, Ryu DR, Yoo TH, Jung DS, Kim JJ, Kim HJ, Choi HY, Kim JS, Adler SG, Natarajan R, Han DS, Kang SW. P-Cadherin is decreased in diabetic glomeruli and in glucose-stimulated podocytes in vivo and in vitro studies. Nephrol Dial Transplant 20:524-531, 2005. 64. Radice GL, Ferreira-Cornwell MC, Robinson SD, Rayburn H, Chodosh LA, Takeichi M, Hynes RO. Precocious mammary gland development in P-cadherin- defi cient mice. J Cell Biol 139:1025-1032, 1997. 65. Inoue T, Yaoita E, Kurihara H, Shimizu F, Sakai T, Kobayashi T, Ohshiro K, Kawachi H, Okada H, Suzuki H, Kihara I, Yamamoto T. FAT is a component of glomerular slit diaphragms. Kidney Int 59:1003-1012, 2001. 66. Ciani L, Patel A, Allen ND, ffrench-Constant C. Mice lacking the giant pro- tocadherin mFAT1 exhibit renal slit junction abnormalities and a partially penetrant cyclopia and anophthalmia phenotype. Mol Cell Biol 23:3575-3582, 2003. 67. Moeller MJ, Soofi A, Braun GS, Li X, Watzl C, Kriz W, Holzman LB. Protocad- herin FAT1 binds Ena/VASP proteins and is necessary for actin dynamics and cell polarization. Embo J 23:3769-3779, 2004. 68. Yuan H, Takeuchi E, Salant DJ. Podocyte slit-diaphragm protein nephrin is linked to the actin cytoskeleton. Am J Physiol Renal Physiol 282:F585-591, 2002. 69. Saleem MA, Ni L, Witherden I, Tryggvason K, Ruotsalainen V, Mundel P, Math- ieson PW. Co-localization of nephrin, podocin, and the actin cytoskeleton: evidence for a role in podocyte foot process formation. Am J Pathol 161:1459- 1466, 2002. 70. Kurihara H, Anderson JM, Farquhar MG. Diversity among tight junctions in rat kidney: glomerular slit diaphragms and endothelial junctions express only one isoform of the tight junction protein ZO-1. Proc Natl Acad Sci U S A 89:7075- 7079, 1992. 71. Shih NY, Li J, Karpitskii V, Nguyen A, Dustin ML, Kanagawa O, Miner JH, Shaw AS. Congenital nephrotic syndrome in mice lacking CD2-associated protein. Science 286:312-315, 1999. 72. Shih NY, Li J, Cotran R, Mundel P, Miner JH, Shaw AS. CD2AP localizes to the slit diaphragm and binds to nephrin via a novel C-terminal domain. Am J Pathol 159:2303-2308, 2001. 73. Boute N, Gribouval O, Roselli S, Benessy F, Lee H, Fuchshuber A, Dahan K, Gubler MC, Niaudet P, Antignac C. NPHS2, encoding the glomerular protein

64 9 References podocin, is mutated in autosomal recessive steroid-resistant nephrotic syndrome. Nat Genet 24:349-354, 2000. 74. Huber TB, Simons M, Hartleben B, Sernetz L, Schmidts M, Gundlach E, Saleem MA, Walz G, Benzing T. Molecular basis of the functional podocin-nephrin complex: mutations in the NPHS2 gene disrupt nephrin targeting to lipid raft microdomains. Hum Mol Genet 12:3397-3405, 2003. 75. Huber TB, Schermer B, Muller RU, Hohne M, Bartram M, Calixto A, Hagmann H, Reinhardt C, Koos F, Kunzelmann K, Shirokova E, Krautwurst D, Harteneck C, Simons M, Pavenstadt H, Kerjaschki D, Thiele C, Walz G, Chalfi e M, Benzing T. Podocin and MEC-2 bind cholesterol to regulate the activity of associated ion channels. Proc Natl Acad Sci U S A 103:17079-17086, 2006. 76. Reiser J, Polu KR, Moller CC, Kenlan P, Altintas MM, Wei C, Faul C, Herbert S, Villegas I, Avila-Casado C, McGee M, Sugimoto H, Brown D, Kalluri R, Mundel P, Smith PL, Clapham DE, Pollak MR. TRPC6 is a glomerular slit diaphragm- associated channel required for normal renal function. Nat Genet 37:739-744, 2005. 77. Winn MP, Conlon PJ, Lynn KL, Farrington MK, Creazzo T, Hawkins AF, Daska- lakis N, Kwan SY, Ebersviller S, Burchette JL, Pericak-Vance MA, Howell DN, Vance JM, Rosenberg PB. A mutation in the TRPC6 cation channel causes familial focal segmental glomerulosclerosis. Science 308:1801-1804, 2005. 78. Dustin ML, Olszowy MW, Holdorf AD, Li J, Bromley S, Desai N, Widder P, Rosenberger F, van der Merwe PA, Allen PM, Shaw AS. A novel adaptor protein orchestrates receptor patterning and cytoskeletal polarity in T-cell contacts. Cell 94:667-677, 1998. 79. Lehtonen S, Zhao F, Lehtonen E. CD2-associated protein directly interacts with the actin cytoskeleton. Am J Physiol Renal Physiol 283:F734-743, 2002. 80. Welsch T, Endlich N, Kriz W, Endlich K. CD2AP and p130Cas localize to differ- ent F-actin structures in podocytes. Am J Physiol Renal Physiol 281:F769-777, 2001. 81. Palmen T, Lehtonen S, Ora A, Kerjaschki D, Antignac C, Lehtonen E, Holthofer H. Interaction of Endogenous Nephrin and CD2-Associated Protein in Mouse Epithelial M-1 Cell Line. J Am Soc Nephrol 13:1766-1772, 2002. 82. Honda K, Yamada T, Endo R, Ino Y, Gotoh M, Tsuda H, Yamada Y, Chiba H, Hirohashi S. Actinin-4, a novel actin-bundling protein associated with cell motility and cancer invasion. J Cell Biol 140:1383-1393, 1998. 83. Kaplan JM, Kim SH, North KN, Rennke H, Correia LA, Tong HQ, Mathis BJ, Rodriguez-Perez JC, Allen PG, Beggs AH, Pollak MR. Mutations in ACTN4, encoding alpha-actinin-4, cause familial focal segmental glomerulosclerosis. Nat Genet 24:251-256, 2000. 84. Kos CH, Le TC, Sinha S, Henderson JM, Kim SH, Sugimoto H, Kalluri R, Gerszten RE, Pollak MR. Mice defi cient in alpha-actinin-4 have severe glomerular disease. J Clin Invest 111:1683-1690, 2003. 85. Michaud JL, Lemieux LI, Dube M, Vanderhyden BC, Robertson SJ, Kennedy CR. Focal and segmental glomerulosclerosis in mice with podocyte-specifi c expression of mutant alpha-actinin-4. J Am Soc Nephrol 14:1200-1211, 2003.

9 References 65 86. Dandapani SV, Sugimoto H, Matthews BD, Kolb RJ, Sinha S, Gerszten RE, Zhou J, Ingber DE, Kalluri R, Pollak MR. Alpha-actinin-4 is required for normal podocyte adhesion. J Biol Chem 282:467-477, 2007. 87. Apperson ML, Moon IS, Kennedy MB. Characterization of densin-180, a new brain-specifi c synaptic protein of the O-sialoglycoprotein family. J Neurosci 16:6839-6852, 1996. 88. Ahola H, Heikkila E, Astrom E, Inagaki M, Izawa I, Pavenstadt H, Kerjaschki D, Holthofer H. A novel protein, densin, expressed by glomerular podocytes. J Am Soc Nephrol 14:1731-1737, 2003. 89. Walikonis RS, Oguni A, Khorosheva EM, Jeng CJ, Asuncion FJ, Kennedy MB. Densin-180 forms a ternary complex with the (alpha)-subunit of Ca2+/calmod- ulin-dependent protein kinase II and (alpha)-actinin. J Neurosci 21:423-433, 2001. 90. Saxen L. Organogenesis of the Kidney. Cambridge, UK. Cambridge University Press, 1987. 91. Reeves W, Caulfi eld JP, Farquhar MG. Differentiation of epithelial foot processes and fi ltration slits: sequential appearance of occluding junctions, epithelial polyanion, and slit membranes in developing glomeruli. Lab Invest 39:90-100, 1978. 92. Kreidberg JA. Podocyte differentiation and glomerulogenesis. J Am Soc Nephrol 14:806-814, 2003. 93. Ruotsalainen V, Patrakka J, Tissari P, Reponen P, Hess M, Kestila M, Holmberg C, Salonen R, Heikinheimo M, Wartiovaara J, Tryggvason K, Jalanko H. Role of nephrin in cell junction formation in human nephrogenesis. Am J Pathol 157:1905-1916, 2000. 94. Ryan GB, Rodewald R, Karnovsky MJ. An ultrastructural study of the glomeru- lar slit diaphragm in aminonucleoside nephrosis. Lab Invest 33:461-468, 1975. 95. Farquhar MG, Vernier RL, Good RA. An electron microscope study of the glomerulus in nephrosis, glomerulonephritis, and lupus erythematosus. J Exp Med 106:649-660, 1957. 96. Kriz W, LeHir M. Pathways to nephron loss starting from glomerular diseases- insights from animal models. Kidney Int 67:404-419, 2005. 97. Glassock RJ, Adler SG, Ward HJ, Cohen AH. Primary Glomerular Diseases. In: Brenner BM, Rector FC, editors. The Kidney. Philadelphia, PA, USA. W.B. Saunders Company, 1991. p. 1182-1279. 98. Shankland SJ. The podocyte’s response to injury: role in proteinuria and glomerulosclerosis. Kidney Int 69:2131-2147, 2006. 99. Munk F. Die Nephrosen. Med. Klin. 12:1019, 1916. 100. Lewis EJ. Management of nephrotic syndrome in adults. In: Cameron JS, Glassock RJ, editors. The nephrotic syndrome. New York: Marcel Dekker; 1988. p. 461-521. 101. Shalhoub RJ. Pathogenesis of lipoid nephrosis: a disorder of T-cell function. Lancet 2:556-560, 1974. 102. Mathieson PW. Immune dysregulation in minimal change nephropathy. Nephrol Dial Transplant 18 Suppl 6:vi26-29, 2003.

66 9 References 103. Chadban SJ, Atkins RC. Glomerulonephritis. Lancet 365:1797-1806, 2005. 104. Tejani A. Morphological transition in minimal change nephrotic syndrome. Nephron 39:157-159, 1985. 105. Rich AR. A hitherto undescribed vulnerability of the juxtamedullary glomeruli in lipoid nephrosis. Bull Johns Hopkins Hosp 100:173-186, 1957. 106. Barisoni L, Kriz W, Mundel P, D’Agati V. The dysregulated podocyte phe- notype: a novel concept in the pathogenesis of collapsing idiopathic focal segmental glomerulosclerosis and HIV-associated nephropathy. J Am Soc Nephrol 10:51-61, 1999. 107. Kriz W, Gretz N, Lemley KV. Progression of glomerular diseases: is the podocyte the culprit? Kidney Int 54:687-697, 1998. 108. Savin VJ, Sharma R, Sharma M, McCarthy ET, Swan SK, Ellis E, Lovell H, Warady B, Gunwar S, Chonko AM, Artero M, Vincenti F. Circulating factor associated with increased glomerular permeability to albumin in recurrent focal segmen- tal glomerulosclerosis. N Engl J Med 334:878-883, 1996. 109. Wasserstein AG. Membranous glomerulonephritis. J Am Soc Nephrol 8:664- 674, 1997. 110. Debiec H, Nauta J, Coulet F, van der Burg M, Guigonis V, Schurmans T, de Heer E, Soubrier F, Janssen F, Ronco P. Role of truncating mutations in MME gene in fetomaternal alloimmunisation and antenatal glomerulopathies. Lancet 364:1252-1259, 2004. 111. Heymann W, Lund HZ, Hackel DB. The nephrotic syndrome in rats; with special reference to the progression of the glomerular lesion and to the use of nephrotoxic sera obtained from ducks. J Lab Clin Med 39:218-224, 1952. 112. Kerjaschki D, Farquhar MG. The pathogenic antigen of Heymann nephritis is a membrane glycoprotein of the renal proximal tubule brush border. Proc Natl Acad Sci U S A 79:5557-5561, 1982. 113. Adler S, Baker PJ, Pritzl P, Couser WG. Detection of terminal complement components in experimental immune glomerular injury. Kidney Int 26:830- 837, 1984. 114. Shah SV. Evidence suggesting a role for hydroxyl radical in passive Heymann nephritis in rats. Am J Physiol 254:F337-344, 1988. 115. Zucchelli P, Zuccala A. Progression of renal failure and hypertensive nephro- sclerosis. Kidney Int Suppl 68:S55-59, 1998. 116. Klag MJ, Whelton PK, Randall BL, Neaton JD, Brancati FL, Ford CE, Shulman NB, Stamler J. Blood pressure and end-stage renal disease in men. N Engl J Med 334:13-18, 1996. 117. Zucchelli P, Zuccala A. Can we accurately diagnose nephrosclerosis? Nephrol Dial Transplant 10 Suppl 6:2-5, 1995. 118. Kretzler M, Koeppen-Hagemann I, Kriz W. Podocyte damage is a critical step in the development of glomerulosclerosis in the uninephrectomised- desoxycorticosterone hypertensive rat. Virchows Arch 425:181-193, 1994. 119. Kriz W, Hosser H, Hahnel B, Simons JL, Provoost AP. Development of vascular pole-associated glomerulosclerosis in the Fawn-hooded rat. J Am Soc Nephrol 9:381-396, 1998.

9 References 67 120. Durvasula RV, Petermann AT, Hiromura K, Blonski M, Pippin J, Mundel P, Pichler R, Griffi n S, Couser WG, Shankland SJ. Activation of a local tissue angiotensin system in podocytes by mechanical strain. Kidney Int 65:30-39, 2004. 121. Endlich N, Kress KR, Reiser J, Uttenweiler D, Kriz W, Mundel P, Endlich K. Podocytes respond to mechanical stress in vitro. J Am Soc Nephrol 12:413-422, 2001. 122. USRDS 2005 Annual Data Report: Atlas of End-Stage Renal Disease in the United States. In: US Renal Data System, National Institutes of Health, National Institute of Diabetes and Digestive and Kidney Diseases, 2005. 123. Krolewski AS, Warram JH, Christlieb AR, Busick EJ, Kahn CR. The changing natural history of nephropathy in type I diabetes. Am J Med 78:785-794, 1985. 124. Andersen AR, Christiansen JS, Andersen JK, Kreiner S, Deckert T. Diabetic nephropathy in Type 1 (insulin-dependent) diabetes: an epidemiological study. Diabetologia 25:496-501, 1983. 125. Borch-Johnsen K, Andersen PK, Deckert T. The effect of proteinuria on relative mortality in type 1 (insulin-dependent) diabetes mellitus. Diabetologia 28:590- 596, 1985. 126. Gall MA, Borch-Johnsen K, Hougaard P, Nielsen FS, Parving HH. Albuminuria and poor glycemic control predict mortality in NIDDM. Diabetes 44:1303-1309, 1995. 127. Viberti GC, Hill RD, Jarrett RJ, Argyropoulos A, Mahmud U, Keen H. Microalbu- minuria as a predictor of clinical nephropathy in insulin-dependent diabetes mellitus. Lancet 1:1430-1432, 1982. 128. Berg UB, Torbjornsdotter TB, Jaremko G, Thalme B. Kidney morphological changes in relation to long-term renal function and metabolic control in adolescents with IDDM. Diabetologia 41:1047-1056, 1998. 129. Steffes MW, Schmidt D, McCrery R, Basgen JM. Glomerular cell number in normal subjects and in type 1 diabetic patients. Kidney Int 59:2104-2113, 2001. 130. Pagtalunan ME, Miller PL, Jumping-Eagle S, Nelson RG, Myers BD, Rennke HG, Coplon NS, Sun L, Meyer TW. Podocyte loss and progressive glomerular injury in type II diabetes. J Clin Invest 99:342-348, 1997. 131. Kimmelstiel P, Wilson C. Intercapillary lesions in glomeruli of kidney. Am J Pathol 12:83-97, 1936. 132. Deckert T, Feldt-Rasmussen B, Borch-Johnsen K, Jensen T, Kofoed-Enevoldsen A. Albuminuria refl ects widespread vascular damage. The Steno hypothesis. Diabetologia 32:219-226, 1989. 133. Wolf G, Chen S, Ziyadeh FN. From the periphery of the glomerular capillary wall toward the center of disease: podocyte injury comes of age in diabetic nephropathy. Diabetes 54:1626-1634, 2005. 134. Ding G, Reddy K, Kapasi AA, Franki N, Gibbons N, Kasinath BS, Singhal PC. Angiotensin II induces apoptosis in rat glomerular epithelial cells. Am J Physiol Renal Physiol 283:F173-180, 2002. 135. The effect of intensive treatment of diabetes on the development and pro- gression of long-term complications in insulin-dependent diabetes mellitus.

68 9 References The Diabetes Control and Complications Trial Research Group. N Engl J Med 329:977-986, 1993. 136. Stratton IM, Adler AI, Neil HA, Matthews DR, Manley SE, Cull CA, Hadden D, Turner RC, Holman RR. Association of glycaemia with macrovascular and microvascular complications of type 2 diabetes (UKPDS 35): prospective observational study. Bmj 321:405-412, 2000. 137. Sawicki PT, Didjurgeit U, Muhlhauser I, Bender R, Heinemann L, Berger M. Smoking is associated with progression of diabetic nephropathy. Diabetes Care 17:126-131, 1994. 138. Mulec H, Johnsen SA, Wiklund O, Bjorck S. Cholesterol: a renal risk factor in diabetic nephropathy? Am J Kidney Dis 22:196-201, 1993. 139. Ekstrand AV, Groop PH, Gronhagen-Riska C. Insulin resistance precedes micro- albuminuria in patients with insulin-dependent diabetes mellitus. Nephrol Dial Transplant 13:3079-3083, 1998. 140. Brenner BM, Cooper ME, de Zeeuw D, Keane WF, Mitch WE, Parving HH, Remuzzi G, Snapinn SM, Zhang Z, Shahinfar S. Effects of losartan on renal and cardiovascular outcomes in patients with type 2 diabetes and nephropathy. N Engl J Med 345:861-869, 2001. 141. Lewis EJ, Hunsicker LG, Clarke WR, Berl T, Pohl MA, Lewis JB, Ritz E, Atkins RC, Rohde R, Raz I. Renoprotective effect of the angiotensin-receptor antagonist irbesartan in patients with nephropathy due to type 2 diabetes. N Engl J Med 345:851-860, 2001. 142. Kriz W. TRPC6 - a new podocyte gene involved in focal segmental glomerulo- sclerosis. Trends Mol Med 11:527-530, 2005. 143. Tryggvason K, Patrakka J, Wartiovaara J. Hereditary proteinuria syndromes and mechanisms of proteinuria. N Engl J Med 354:1387-1401, 2006. 144. Gubler MC. Podocyte differentiation and hereditary proteinuria/nephrotic syndromes. J Am Soc Nephrol 14 Suppl 1:S22-26, 2003. 145. Ly J, Alexander M, Quaggin SE. A podocentric view of nephrology. Curr Opin Nephrol Hypertens 13:299-305, 2004. 146. Hallman N, Hjelt L, Ahvenainen E. Nephrotic syndrome in newborn and young infants. Annals Pediatrica Fennica 2:227-241, 1956. 147. Norio R. Heredity in the congenital nephrotic syndrome. A genetic study of 57 Finnish families with a review of reported cases. Ann Paediatr Fenn 12:Suppl 27:21-94, 1966. 148. Rapola J, Sariola H, Ekblom P. Pathology of fetal congenital nephrosis: im- munohistochemical and ultrastructural studies. Kidney Int 25:701-707, 1984. 149. Autio-Harmainen H. Renal pathology of fetuses with congenital nephrotic syndrome of the Finnish type. 2. A qualitative and quantitative electron microscopic study. Acta Pathol Microbiol Scand [A] 89:215-222, 1981. 150. Huttunen NP. Congenital nephrotic syndrome of Finnish type. Study of 75 patients. Arch Dis Child 51:344-348, 1976. 151. Hinkes B, Wiggins RC, Gbadegesin R, Vlangos CN, Seelow D, Nurnberg G, Garg P, Verma R, Chaib H, Hoskins BE, Ashraf S, Becker C, Hennies HC, Goyal M, Wharram BL, Schachter AD, Mudumana S, Drummond I, Kerjaschki D, Wald-

9 References 69 herr R, Dietrich A, Ozaltin F, Bakkaloglu A, Cleper R, Basel-Vanagaite L, Pohl M, Griebel M, Tsygin AN, Soylu A, Muller D, Sorli CS, Bunney TD, Katan M, Liu J, Attanasio M, O’Toole J F, Hasselbacher K, Mucha B, Otto EA, Airik R, Kispert A, Kelley GG, Smrcka AV, Gudermann T, Holzman LB, Nurnberg P, Hildebrandt F. Positional cloning uncovers mutations in PLCE1 responsible for a nephrotic syndrome variant that may be reversible. Nat Genet 38:1397-1405, 2006. 152. Zenker M, Aigner T, Wendler O, Tralau T, Muntefering H, Fenski R, Pitz S, Schumacher V, Royer-Pokora B, Wuhl E, Cochat P, Bouvier R, Kraus C, Mark K, Madlon H, Dotsch J, Rascher W, Maruniak-Chudek I, Lennert T, Neumann LM, Reis A. Human laminin beta2 defi ciency causes congenital nephrosis with mesangial sclerosis and distinct eye abnormalities. Hum Mol Genet 13:2625- 2632, 2004. 153. Tsukaguchi H, Sudhakar A, Le TC, Nguyen T, Yao J, Schwimmer JA, Schachter AD, Poch E, Abreu PF, Appel GB, Pereira AB, Kalluri R, Pollak MR. NPHS2 muta- tions in late-onset focal segmental glomerulosclerosis: R229Q is a common disease-associated allele. J Clin Invest 110:1659-1666, 2002. 154. Pereira AC, Pereira AB, Mota GF, Cunha RS, Herkenhoff FL, Pollak MR, Mill JG, Krieger JE. NPHS2 R229Q functional variant is associated with microalbuminu- ria in the general population. Kidney Int 65:1026-1030, 2004. 155. Arrondel C, Vodovar N, Knebelmann B, Grunfeld JP, Gubler MC, Antignac C, Heidet L. Expression of the nonmuscle myosin heavy chain IIA in the human kidney and screening for MYH9 mutations in Epstein and Fechtner syndromes. J Am Soc Nephrol 13:65-74, 2002. 156. Seri M, Pecci A, Di Bari F, Cusano R, Savino M, Panza E, Nigro A, Noris P, Gangarossa S, Rocca B, Gresele P, Bizzaro N, Malatesta P, Koivisto PA, Longo I, Musso R, Pecoraro C, Iolascon A, Magrini U, Rodriguez Soriano J, Renieri A, Ghiggeri GM, Ravazzolo R, Balduini CL, Savoia A. MYH9-related disease: May-Hegglin anomaly, Sebastian syndrome, Fechtner syndrome, and Epstein syndrome are not distinct entities but represent a variable expression of a single illness. Medicine (Baltimore) 82:203-215, 2003. 157. Hudson BG, Tryggvason K, Sundaramoorthy M, Neilson EG. Alport’s syndrome, Goodpasture’s syndrome, and type IV collagen. N Engl J Med 348:2543-2556, 2003. 158. Guo JK, Menke AL, Gubler MC, Clarke AR, Harrison D, Hammes A, Hastie ND, Schedl A. WT1 is a key regulator of podocyte function: reduced expression levels cause crescentic glomerulonephritis and mesangial sclerosis. Hum Mol Genet 11:651-659, 2002. 159. Rohr C, Prestel J, Heidet L, Hosser H, Kriz W, Johnson RL, Antignac C, Witzgall R. The LIM-homeodomain transcription factor Lmx1b plays a crucial role in podocytes. J Clin Invest 109:1073-1082, 2002. 160. Dreyer SD, Zhou G, Baldini A, Winterpacht A, Zabel B, Cole W, Johnson RL, Lee B. Mutations in LMX1B cause abnormal skeletal patterning and renal dysplasia in nail patella syndrome. Nat Genet 19:47-50, 1998. 161. Pelletier J, Bruening W, Kashtan CE, Mauer SM, Manivel JC, Striegel JE, Houghton DC, Junien C, Habib R, Fouser L, et al. Germline mutations in the

70 9 References Wilms’ tumor suppressor gene are associated with abnormal urogenital development in Denys-Drash syndrome. Cell 67:437-447, 1991. 162. Barbaux S, Niaudet P, Gubler MC, Grunfeld JP, Jaubert F, Kuttenn F, Fekete CN, Souleyreau-Therville N, Thibaud E, Fellous M, McElreavey K. Donor splice-site mutations in WT1 are responsible for Frasier syndrome. Nat Genet 17:467-470, 1997. 163. Harjutsalo V, Katoh S, Sarti C, Tajima N, Tuomilehto J. Population-based assessment of familial clustering of diabetic nephropathy in type 1 diabetes. Diabetes 53:2449-2454, 2004. 164. Seaquist ER, Goetz FC, Rich S, Barbosa J. Familial clustering of diabetic kidney disease. Evidence for genetic susceptibility to diabetic nephropathy. N Engl J Med 320:1161-1165, 1989. 165. Quinn M, Angelico MC, Warram JH, Krolewski AS. Familial factors determine the development of diabetic nephropathy in patients with IDDM. Diabetolo- gia 39:940-945, 1996. 166. Pettitt DJ, Saad MF, Bennett PH, Nelson RG, Knowler WC. Familial predisposi- tion to renal disease in two generations of Pima Indians with type 2 (non- insulin-dependent) diabetes mellitus. Diabetologia 33:438-443, 1990. 167. Fogarty DG, Rich SS, Hanna L, Warram JH, Krolewski AS. Urinary albumin excretion in families with type 2 diabetes is heritable and genetically cor- related to blood pressure. Kidney Int 57:250-257, 2000. 168. Fioretto P, Steffes MW, Barbosa J, Rich SS, Miller ME, Mauer M. Is diabetic nephropathy inherited? Studies of glomerular structure in type 1 diabetic sibling pairs. Diabetes 48:865-869, 1999. 169. Krolewski AS. Genetics of diabetic nephropathy: evidence for major and minor gene effects. Kidney Int 55:1582-1596, 1999. 170. Osterholm AM, He B, Pitkaniemi J, Albinsson L, Berg T, Sarti C, Tuomilehto J, Tryggvason K. Genome-wide scan for type 1 diabetic nephropathy in the Finnish population reveals suggestive linkage to a single locus on chromosome 3q. Kidney Int 71:140-145, 2007. 171. Moczulski DK, Rogus JJ, Antonellis A, Warram JH, Krolewski AS. Major suscep- tibility locus for nephropathy in type 1 diabetes on chromosome 3q: results of novel discordant sib-pair analysis. Diabetes 47:1164-1169, 1998. 172. McKnight AJ, Maxwell AP, Sawcer S, Compston A, Setakis E, Patterson CC, Brady HR, Savage DA. A genome-wide DNA microsatellite association screen to identify chromosomal regions harboring candidate genes in diabetic nephropathy. J Am Soc Nephrol 17:831-836, 2006. 173. Ng DP, Tai BC, Koh D, Tan KW, Chia KS. Angiotensin-I converting enzyme inser- tion/deletion polymorphism and its association with diabetic nephropathy: a meta-analysis of studies reported between 1994 and 2004 and comprising 14,727 subjects. Diabetologia 48:1008-1016, 2005. 174. Conway BR, Savage DA, Maxwell AP. Identifying genes for diabetic neph- ropathy--current diffi culties and future directions. Nephrol Dial Transplant 21:3012-3017, 2006. 175. Kim JM, Wu H, Green G, Winkler CA, Kopp JB, Miner JH, Unanue ER, Shaw

9 References 71 AS. CD2-associated protein haploinsuffi ciency is linked to glomerular disease susceptibility. Science 300:1298-1300, 2003. 176. Holthofer H, Reivinen J, Miettinen A. Nephron segment and cell-type specifi c expression of gangliosides in the developing and adult kidney. Kidney Int 45:123-130, 1994. 177. Cohen CD, Frach K, Schlondorff D, Kretzler M. Quantitative gene expression analysis in renal biopsies: a novel protocol for a high-throughput multicenter application. Kidney Int 61:133-140, 2002. 178. Mogensen CE, Chachati A, Christensen CK, Close CF, Deckert T, Hommel E, Kastrup J, Lefebvre P, Mathiesen ER, Feldt-Rasmussen B, et al. Microalbuminu- ria: an early marker of renal involvement in diabetes. Uremia Invest 9:85-95, 1985. 179. Dallman MJ, Lamb JR, editors. Haematopoietic and Lymphoid Cell Culture: Cambridge University Press; 2000. 180. Ryffel B, Woerly G, Greiner B, Haendler B, Mihatsch MJ, Foxwell BM. Distribu- tion of the cyclosporine binding protein cyclophilin in human tissues. Immu- nology 72:399-404, 1991. 181. Schmid H, Cohen CD, Henger A, Irrgang S, Schlondorff D, Kretzler M. Valida- tion of endogenous controls for gene expression analysis in microdissected human renal biopsies. Kidney Int 64:356-360, 2003. 182. Ceol M, Del Prete D, Tosetto E, Graziotto R, Gambaro G, D’Angelo A, Anglani F. GAPDH as housekeeping gene at renal level. Kidney Int 65:1972-1973; author reply 1973-1974, 2004. 183. Schaefer BC. Revolutions in rapid amplifi cation of cDNA ends: new strategies for polymerase chain reaction cloning of full-length cDNA ends. Anal Biochem 227:255-273, 1995. 184. Goda SK, Minton NP. A simple procedure for gel electrophoresis and northern blotting of RNA. Nucleic Acids Res 23:3357-3358, 1995. 185. Laemmli UK. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 227:680-685, 1970. 186. de Bakker PI, Yelensky R, Pe’er I, Gabriel SB, Daly MJ, Altshuler D. Effi ciency and power in genetic association studies. Nat Genet 37:1217-1223, 2005. 187. Storm N, Darnhofer-Patel B, van den Boom D, Rodi CP. MALDI-TOF mass spectrometry-based SNP genotyping. Methods Mol Biol 212:241-262, 2003. 188. Livak KJ. SNP genotyping by the 5’-nuclease reaction. Methods Mol Biol 212:129-147, 2003. 189. Wiles MV, Vauti F, Otte J, Fuchtbauer EM, Ruiz P, Fuchtbauer A, Arnold HH, Lehrach H, Metz T, von Melchner H, Wurst W. Establishment of a gene-trap sequence tag library to generate mutant mice from embryonic stem cells. Nat Genet 24:13-14, 2000. 190. Rantanen M, Palmen T, Patari A, Ahola H, Lehtonen S, Astrom E, Floss T, Vauti F, Wurst W, Ruiz P, Kerjaschki D, Holthofer H. Nephrin TRAP Mice Lack Slit Diaphragms and Show Fibrotic Glomeruli and Cystic Tubular Lesions. J Am Soc Nephrol 13:1586-1594., 2002. 191. Putaala H, Sainio K, Sariola H, Tryggvason K. Primary structure of mouse and

72 9 References rat nephrin cDNA and structure and expression of the mouse gene. J Am Soc Nephrol 11:991-1001., 2000. 192. Nevins JR. The pathway of eukaryotic mRNA formation. Annu Rev Biochem 52:441-466, 1983. 193. Rinta-Valkama J, Palmen T, Lassila M, Holthofer H. Podocyte-associated proteins FAT, alpha-actinin-4 and fi ltrin are expressed in Langerhans islets of the pancreas. Mol Cell Biochem 294:117-125, 2007. 194. Rinta-Valkama J, Aaltonen P, Lassila M, Palmen T, Tossavainen P, Knip M, Holthofer H. Densin and fi ltrin in the pancreas and in the kidney, targets for humoral autoimmunity in patients with type 1 diabetes. Diabetes Metab Res Rev 23:119-126, 2007. 195. Minaki Y, Mizuhara E, Morimoto K, Nakatani T, Sakamoto Y, Inoue Y, Satoh K, Imai T, Takai Y, Ono Y. Migrating postmitotic neural precursor cells in the ventricular zone extend apical processes and form adherens junctions near the ventricle in the developing spinal cord. Neurosci Res 52:250-262, 2005. 196. Serizawa S, Miyamichi K, Takeuchi H, Yamagishi Y, Suzuki M, Sakano H. A neu- ronal identity code for the odorant receptor-specifi c and activity-dependent axon sorting. Cell 127:1057-1069, 2006. 197. Farquhar MG, Palade GE. Junctional complexes in various epithelia. J Cell Biol 17:375-412, 1963. 198. Zanone MM, Favaro E, Doublier S, Lozanoska-Ochser B, Deregibus MC, Green- ing J, Huang GC, Klein N, Cavallo Perin P, Peakman M, Camussi G. Expression of nephrin by human pancreatic islet endothelial cells. Diabetologia 48:1789- 1797, 2005. 199. Astrom E, Rinta-Valkama J, Gylling M, Ahola H, Miettinen A, Timonen T, Holthofer H. Nephrin in human lymphoid tissues. Cell Mol Life Sci 63:498-504, 2006. 200. Kuusniemi AM, Kestila M, Patrakka J, Lahdenkari AT, Ruotsalainen V, Holm- berg C, Karikoski R, Salonen R, Tryggvason K, Jalanko H. Tissue expression of nephrin in human and pig. Pediatr Res 55:774-781, 2004. 201. Putaala H, Soininen R, Kilpelainen P, Wartiovaara J, Tryggvason K. The murine nephrin gene is specifi cally expressed in kidney, brain and pancreas: inactiva- tion of the gene leads to massive proteinuria and neonatal death. Hum Mol Genet 10:1-8., 2001. 202. Moeller MJ, Kovari IA, Holzman LB. Evaluation of a new tool for exploring podocyte biology: mouse Nphs1 5’ fl anking region drives LacZ expression in podocytes. J Am Soc Nephrol 11:2306-2314, 2000. 203. Moeller MJ, Sanden SK, Soofi A, Wiggins RC, Holzman LB. Two Gene Frag- ments that Direct Podocyte-Specifi c Expression in Transgenic Mice. J Am Soc Nephrol 13:1561-1567, 2002. 204. Wong MA, Cui S, Quaggin SE. Identifi cation and characterization of a glomerular-specifi c promoter from the human nephrin gene. Am J Physiol Renal Physiol 279:F1027-1032, 2000. 205. Beltcheva O, Kontusaari S, Fetissov S, Putaala H, Kilpelainen P, Hokfelt T, Tryggvason K. Alternatively used promoters and distinct elements direct

9 References 73 tissue-specifi c expression of nephrin. J Am Soc Nephrol 14:352-358, 2003. 206. Guo G, Morrison DJ, Licht JD, Quaggin SE. WT1 activates a glomerular-specifi c enhancer identifi ed from the human nephrin gene. J Am Soc Nephrol 15:2851- 2856, 2004. 207. Benigni A, Zoja C, Tomasoni S, Campana M, Corna D, Zanchi C, Gagliardini E, Garofano E, Rottoli D, Ito T, Remuzzi G. Transcriptional regulation of nephrin gene by peroxisome proliferator-activated receptor-gamma agonist: molecular mechanism of the antiproteinuric effect of pioglitazone. J Am Soc Nephrol 17:1624-1632, 2006. 208. Kramer J, Steinhoff J, Klinger M, Fricke L, Rohwedel J. Cells differentiated from mouse embryonic stem cells via embryoid bodies express renal marker molecules. Differentiation 74:91-104, 2006. 209. Bruce SJ, Rea RW, Steptoe AL, Busslinger M, Bertram JF, Perkins AC. In vitro differentiation of murine embryonic stem cells toward a renal lineage. Dif- ferentiation, 2007. 210. Lapidot M, Pilpel Y. Genome-wide natural antisense transcription: coupling its regulation to its different regulatory mechanisms. EMBO Rep 7:1216-1222, 2006. 211. Trinklein ND, Aldred SF, Hartman SJ, Schroeder DI, Otillar RP, Myers RM. An abundance of bidirectional promoters in the human genome. Genome Res 14:62-66, 2004. 212. Townson SM, Dobrzycka KM, Lee AV, Air M, Deng W, Kang K, Jiang S, Kioka N, Michaelis K, Oesterreich S. SAFB2, a new scaffold attachment factor homolog and estrogen receptor corepressor. J Biol Chem 278:20059-20068, 2003. 213. Grimwood J, Gordon LA, Olsen A, Terry A, Schmutz J, Lamerdin J, Hellsten U, Goodstein D, Couronne O, Tran-Gyamfi M, Aerts A, Altherr M, Ashworth L, Bajorek E, Black S, Branscomb E, Caenepeel S, Carrano A, Caoile C, Chan YM, Christensen M, Cleland CA, Copeland A, Dalin E, Dehal P, Denys M, Detter JC, Escobar J, Flowers D, Fotopulos D, Garcia C, Georgescu AM, Glavina T, Gomez M, Gonzales E, Groza M, Hammon N, Hawkins T, Haydu L, Ho I, Huang W, Is- rani S, Jett J, Kadner K, Kimball H, Kobayashi A, Larionov V, Leem SH, Lopez F, Lou Y, Lowry S, Malfatti S, Martinez D, McCready P, Medina C, Morgan J, Nel- son K, Nolan M, Ovcharenko I, Pitluck S, Pollard M, Popkie AP, Predki P, Quan G, Ramirez L, Rash S, Retterer J, Rodriguez A, Rogers S, Salamov A, Salazar A, She X, Smith D, Slezak T, Solovyev V, Thayer N, Tice H, Tsai M, Ustaszewska A, Vo N, Wagner M, Wheeler J, Wu K, Xie G, Yang J, Dubchak I, Furey TS, DeJong P, Dickson M, Gordon D, Eichler EE, Pennacchio LA, Richardson P, Stubbs L, Rokhsar DS, Myers RM, Rubin EM, Lucas SM. The DNA sequence and biology of human chromosome 19. Nature 428:529-535, 2004. 214. Dehal P, Predki P, Olsen AS, Kobayashi A, Folta P, Lucas S, Land M, Terry A, Ecale Zhou CL, Rash S, Zhang Q, Gordon L, Kim J, Elkin C, Pollard MJ, Richard- son P, Rokhsar D, Uberbacher E, Hawkins T, Branscomb E, Stubbs L. Human chromosome 19 and related regions in mouse: conservative and lineage- specifi c evolution. Science 293:104-111, 2001. 215. Fogo AB. Animal models of FSGS: lessons for pathogenesis and treatment.

74 9 References Semin Nephrol 23:161-171, 2003. 216. Peltonen L, Perola M, Naukkarinen J, Palotie A. Lessons from studying monogenic disease for common disease. Hum Mol Genet 15 Spec No 1:R67-74, 2006. 217. de la Chapelle A, Wright FA. Linkage disequilibrium mapping in isolated populations: the example of Finland revisited. Proc Natl Acad Sci U S A 95:12416-12423, 1998. 218. Laitinen T, Polvi A, Rydman P, Vendelin J, Pulkkinen V, Salmikangas P, Makela S, Rehn M, Pirskanen A, Rautanen A, Zucchelli M, Gullsten H, Leino M, Alenius H, Petays T, Haahtela T, Laitinen A, Laprise C, Hudson TJ, Laitinen LA, Kere J. Characterization of a common susceptibility locus for asthma-related traits. Science 304:300-304, 2004. 219. Tuomilehto J, Karvonen M, Pitkaniemi J, Virtala E, Kohtamaki K, Toivanen L, Tuomilehto-Wolf E. Record-high incidence of Type I (insulin-dependent) diabetes mellitus in Finnish children. The Finnish Childhood Type I Diabetes Registry Group. Diabetologia 42:655-660, 1999. 220. de Bakker PI, Burtt NP, Graham RR, Guiducci C, Yelensky R, Drake JA, Bersaglieri T, Penney KL, Butler J, Young S, Onofrio RC, Lyon HN, Stram DO, Haiman CA, Freedman ML, Zhu X, Cooper R, Groop L, Kolonel LN, Henderson BE, Daly MJ, Hirschhorn JN, Altshuler D. Transferability of tag SNPs in genetic association studies in multiple populations. Nat Genet 38:1298-1303, 2006. 221. Nielsen DM, Ehm MG, Weir BS. Detecting marker-disease association by testing for Hardy-Weinberg disequilibrium at a marker locus. Am J Hum Genet 63:1531-1540, 1998. 222. Zhou X, Hurst RD, Templeton D, Whiteside CI. High glucose alters actin assembly in glomerular mesangial and epithelial cells. Lab Invest 73:372-383, 1995. 223. Nosadini R, Velussi M, Brocco E, Abaterusso C, Piarulli F, Morgia G, Satta A, Faedda R, Abhyankar A, Luthman H, Tonolo G. Altered transcapillary escape of albumin and microalbuminuria refl ects two different pathogenetic mecha- nisms. Diabetes 54:228-233, 2005. 224. Ha TS. High glucose and advanced glycosylated end-products affect the ex- pression of alpha-actinin-4 in glomerular epithelial cells. Nephrology (Carlton) 11:435-441, 2006. 225. Purcell S, Cherny SS, Sham PC. Genetic Power Calculator: design of linkage and association genetic mapping studies of complex traits. Bioinformatics 19:149- 150, 2003. 226. Enattah NS, Sahi T, Savilahti E, Terwilliger JD, Peltonen L, Jarvela I. Identifi ca- tion of a variant associated with adult-type hypolactasia. Nat Genet 30:233- 237, 2002. 227. Pettersson-Fernholm K, Forsblom C, Perola M, Groop PH. Polymorphisms in the nephrin gene and diabetic nephropathy in type 1 diabetic patients. Kidney Int 63:1205-1210, 2003. 228. Longo I, Porcedda P, Mari F, Giachino D, Meloni I, Deplano C, Brusco A, Bosio M, Massella L, Lavoratti G, Roccatello D, Frasca G, Mazzucco G, Muda

9 References 75 AO, Conti M, Fasciolo F, Arrondel C, Heidet L, Renieri A, De Marchi M. COL4A3/COL4A4 mutations: from familial hematuria to autosomal-dominant or recessive Alport syndrome. Kidney Int 61:1947-1956, 2002. 229. Lemmink HH, Mochizuki T, van den Heuvel LP, Schroder CH, Barrientos A, Monnens LA, van Oost BA, Brunner HG, Reeders ST, Smeets HJ. Mutations in the type IV collagen alpha 3 (COL4A3) gene in autosomal recessive Alport syndrome. Hum Mol Genet 3:1269-1273, 1994. 230. Mochizuki T, Lemmink HH, Mariyama M, Antignac C, Gubler MC, Pirson Y, Verellen-Dumou- lin C, Chan B, Schroder CH, Smeets HJ, et al. Identifi cation of mutations in the alpha 3(IV) and alpha 4(IV) collagen genes in autosomal recessive Alport syndrome. Nat Genet 8:77-81, 1994. 231. Barker DF, Hostikka SL, Zhou J, Chow LT, Oliphant AR, Gerken SC, Gregory MC, Skolnick MH, Atkin CL, Tryggvason K. Identifi cation of mutations in the COL4A5 collagen gene in Alport syndrome. Science 248:1224-1227, 1990. 232. Greiber S, Munzel T, Kastner S, Muller B, Schollmeyer P, Pavenstadt H. NAD(P)H oxidase activity in cultured human podocytes: effects of adenosine triphosphate. Kidney Int 53:654- 663, 1998. 233. Hill DP, Wurst W. Screening for novel pattern formation genes using gene trap approaches. Methods Enzymol 225:664-681, 1993. 234. Miyazaki J, Araki K, Yamato E, Ikegami H, Asano T, Shibasaki Y, Oka Y, Yamamura K. Establishment of a pancreatic beta cell line that retains glucose-inducible insulin secretion: special reference to expression of glucose transporter isoforms. Endocrinology 127:126-132, 1990. 235. Patrakka J. Nephrin. Role in human glomerulogenesis and nephrotic disorders. University of Helsinki, Helsinki, Finland, 2001.

76 9 References