Skip to main content
Journal of the American Society of Nephrology : JASN logoLink to Journal of the American Society of Nephrology : JASN
. 2008 Nov;19(11):2159–2169. doi: 10.1681/ASN.2008030312

Podocyte-Selective Deletion of Dicer Induces Proteinuria and Glomerulosclerosis

Shaolin Shi *, Liping Yu *, Celine Chiu *, Yezhou Sun *, Jin Chen *, Greg Khitrov *, Matthias Merkenschlager , Lawrence B Holzman , Weijia Zhang *, Peter Mundel *, Erwin P Bottinger *,§
PMCID: PMC2573016  PMID: 18776119

Abstract

Dicer is an enzyme that generates microRNA (miRNA), which are small, noncoding RNA that function as important regulators of gene and protein expression. For exploration of the functional roles of miRNA in glomerular biology, Dicer was inactivated selectively in mouse podocytes. Mutant mice developed proteinuria 4 to 5 weeks after birth and died several weeks later, presumably from kidney failure. Multiple abnormalities were observed in glomeruli of mutant mice, including foot process effacement, irregular and split areas of the glomerular basement membrane, podocyte apoptosis and depletion, mesangial expansion, capillary dilation, and glomerulosclerosis. Gene profiling revealed upregulation of 190 genes in glomeruli isolated from mutant mice at the onset of proteinuria compared with control littermates. Target sequences for 16 miRNA were significantly enriched in the 3′-untranslated regions of the 190 upregulated genes. Further suggesting validity of the in silico analysis, six of the eight top-candidate miRNA were identified in miRNA libraries generated from podocyte cultures; these included four members of the mir-30 miRNA family, which are known to degrade target transcripts directly. Among 15 upregulated target genes of the mir-30 miRNA, four genes known to be expressed and/or functional in podocytes were identified, including receptor for advanced glycation end product, vimentin, heat-shock protein 20, and immediate early response 3. Receptor for advanced glycation end product and immediate early response 3 are known to mediate podocyte apoptosis, whereas vimentin and heat-shock protein-20 are involved in cytoskeletal structure. Taken together, these results provide a knowledge base for ongoing investigations to validate functional roles for the mir-30 miRNA family in podocyte homeostasis and podocytopathies.


MicroRNA (miRNA) are a group of noncoding small RNA that are present in lower through higher organisms and function to regulate gene expression by translation repression or transcript degradation of target genes.1 A miRNA gene is transcribed by RNA polymerase II2 or III3 to form a primary transcript. This primary transcript is then cleaved by Drosha,4 an RNase III enzyme, in the nucleus to form an approximately 70-nt-long product that is characteristic of a stem-loop structure and is termed precursor miRNA (pre-miRNA). Pre-miRNA is transported to the cytoplasm by exportin 5, a nuclear membrane transporter.5 In the cytoplasm, all pre-miRNA are further processed by Dicer, another RNase III enzyme, to form approximately 22-nt mature double-stranded miRNA.6

The importance of miRNA in development and tissue homeostasis is now well documented. Abrogating Dicer to prevent production of functional miRNA in mice resulted in embryonic lethality at approximately embryonic day 7.5.7 Studies of conditional ablation of Dicer in immune cells,8,9 developing oocytes,10,11 limb buds,12 Purkinje cells,13 chondrocytes,14 and heart15 all have suggested critical roles of miRNA in multiple biologic processes in those target cells, including proliferation, differentiation, and maintenance of cell structure and function. In addition to Dicer-deficient models of global miRNA inactivation, single miRNA have been shown to be required in regulating a variety of cellular processes, for example, inactivation of miR-208 and miR-155 in mice has revealed their roles in mediating cardiomyocyte hypertrophy in response to stress16 and in regulating T helper cell differentiation and the germinal center reaction,17,18 respectively. Besides, miRNA are implicated in other biologic processes.1922 Whereas recent observations suggested a role for mouse miR-192 in glomerular TGF-β and hyperglycemia-induced expression of collagen I α2 in glomeruli,23 the roles of miRNA in kidney remain largely unknown.

Here, we report functional, morphologic, and molecular alterations induced by conditional inactivation of Dicer selectively in podocytes of mice. The mutant mice manifest a sequelae of podocyte and glomerular abnormalities, including foot process (FP) effacement, glomerular basement membrane (GBM) alterations, proteinuria, podocyte apoptosis, podocyte depletion, mesangial expansion, and glomerulosclerosis consistent with the podocyte depletion paradigm of progressive glomerulosclerosis.24,25 Thus, miRNA control expression of genes and gene products that are essential for structural homeostasis and survival of mature podocytes in mice. The identification and characterization of specific miRNA and their target genes that are critical for podocyte maintenance are ongoing and are expected to produce novel insights into mechanisms of glomerular diseases.

RESULTS

NPHS2-Cre Efficiently Deletes Dicer and Eliminates miRNA

To confirm that podocin promoter-controlled NPHS2-Cre transgene efficiently deleted loxP-flanked sequences in podocytes, we crossed Nphs2-Cre transgenic and Rosa26 reporter mice followed by β-gal staining of kidney sections. NPHS2-Cre mediated successful recombination at the Rosa26 reporter locus as demonstrated by positive β-gal staining in glomeruli of Rosa26/+:NPHS2-Cre mice at 2 wk of age (Figure 1, A and B). A schematic representation depicts the Dicer alleles before and after Cre-mediated recombination (Figure 1C). PCR of genomic DNA obtained from isolated glomeruli confirmed Cre-mediated recombination of Dicer alleles in podocytes of glomeruli from conditional knockout (mutant) mice (DicerF/F:NPHS2-Cre) with and without proteinuria (Figure 1D). To confirm that Dicer deletion resulted in loss of miRNA in podocytes, we examined the abundance of mature form of miR-30d and miR-23b, two miRNA that are known to be expressed in podocytes (smiRNAdb [http://www.mirz.unibas.ch/smiRNAdb/]). Compared with control (DicerF/+:NPHS2-Cre), levels of both miRNA were significantly reduced in glomeruli of mutant mice and the miRNA reduction was inversely correlated with the absence or presence of proteinuria (Figure 1E). Residual levels of miRNA present in glomeruli of mutant mice were likely due to variation of efficiency of Dicer deletion in different podocytes/glomeruli and/or to expression of the miRNA in endocapillary cells. To demonstrate further that Dicer was inactive in precursor miRNA processing, we conducted quantitative real-time PCR to examine whether precursor miR-30a, 30c-2, and 30d accumulated in mutant podocytes. The result showed the abundances of these precursor miRNA in mutant podocytes were indeed significantly higher than that in control podocytes (Figure 1F).

Figure 1.

Figure 1.

NPHS2-Cre transgene efficiently deleted Dicer floxed region, leading to downregulation of mature miRNA and accumulation of precursor miRNA in podocytes of conditional knockout (mutant) mice. (A) A glomerulus of a control mouse after β-gal staining. (B) Two glomeruli from a 2-wk-old Rosa26F/+:NPHS2-Cre mouse after β-gal staining. The arrows point to the blue spots depicting podocytes with deletion at Rosa26 allele locus in a single glomerulus. (C) Schematic representation of Dicer alleles: Wild-type, loxP floxed, and deleted. D1 and D3 are the primers for PCR. E20 and E21 depict exons 20 and 21 of Dicer gene. loxP sites are denoted with open arrowheads. (D) Electrophoretic agarose gel analysis of PCR products from glomerular genomic DNA of mice. Lane 1, DicerF/+ (control); lane 2, DicerF/F:NPHS2-Cre (mutant) without proteinuria; lane 3, DicerF/F:NPHS2-Cre (mutant) with proteinuria; lane 4, DicerF/+:NPHS2-Cre; lane 5, no DNA (negative control). Molecular weight (MW) 1 kb+ DNA ladder (Invitrogen). (E) Histogram shows the abundances of miR-30d and miR-23b in mutant mice without (−PU) and with (+PU) proteinuria relative to that in control mice. The bars represent average ± SD (n = 3). (F) Histogram shows the abundances of premiR-30a, 30c-2, and 30d in mutant without (−PU) and with (+PU) proteinuria relative to that in control mice. Bars represent average ± SD (n = 3).

Conditional Knockout (Mutant:DicerF/F:NPHS2-Cre) Mice Developed Progressive Proteinuria, Wasting, and Death

The distribution of observed genotype frequencies among offspring of heterozygotic crosses was consistent with the expected genotype distribution on the basis of Mendelian ratio (data not shown). The mutant mice were indistinguishable from wild-type littermates within the first 5 wk of life, suggesting that Dicer deletion driven by NPHS2-Cre did not cause embryonic lethality or overt postnatal developmental phenotypes.

Progress of proteinuria was monitored periodically in cohorts of mutant (DicerF/F:NPHS2-Cre) and control (DicerF/+:NPHS2-Cre, DicerF/F, and DicerF/+) littermate mice by urine dipsticks. The average age at proteinuria onset was 4.7 ± 2.0 wk (SD), ranging from approximately 2 wk to 10 mo of age (n = 50). Once proteinuria was detectable, it progressed subsequently to high-grade proteinuria and the mice died within several weeks. Proteinuria progression was associated with weight loss and runted and pale appearance followed by death. The average lifespan of mutant mice was 7.89 ± 1.97 wk (SD), ranging from 4 wk to 1 yr (n = 26). There was no gender dimorphism in lifespan and any of the observed phenotype manifestations (15 males, 11 females; data not shown). Mice heterozygous for Dicer deletion in podocytes (DicerF/+:NPHS2-Cre) were indistinguishable from wild-type mice with regard to growth, urinary protein excretion, appearance, and lifespan (n = 13).

Dicer Mutant Mice Developed Progressive Glomerulosclerosis and Tubulointerstitial Fibrosis

Kidneys of mutant mice (DicerF/F:NPHS2-Cre) with severe proteinuria appeared smaller, paler, and denser when compared with control kidneys (Figure 2A). Histopathology of kidney in mutant mice was assessed by examination of periodic acid-Schiff–stained sections. Low-power microscopy revealed protein casts in tubular lumens of degenerated nephrons, the frequency of which was proportional to the level of proteinuria determined by dipstick (Figure 2B). In glomeruli, progressive sclerosis was observed with increased extracellular matrix deposition, mesangial expansion, reduced cellularity, and capillary dilations (Figure 2, D through F).

Figure 2.

Figure 2.

Dicer deficiency in podocytes resulted in progressive glomerular defects. (A) Kidneys from a DicerF/+:NPHS2-Cre mouse (control, left) and a DicerF/F:NPHS2-Cre mouse (mutant, right) that had severe proteinuria and was dying. (B) Periodic acid-Schiff (PAS) staining of a kidney section from a mutant mouse with massive proteinuria. Protein casts were present in most of tubular lumens. (C through F) Glomeruli from PAS-stained kidney sections of control (C) and mutant (D through F) mice. Glomeruli of mutant mice were sclerotic (D through F) with mesangial expansion (white arrow), extracellular deposition (white arrowhead), capillary dilation (▪), and reduced cellularity. In addition, tubulointerstitial fibrosis (black arrow), tubular dilation, and protein casts (⋄) are seen in kidney section of mutant mice (F). (G and H) Electronic microscopy examination showing FP effacement (white arrowhead), GBM irregularity (white arrow), and splitting (red arrow) of mutant mice (6-wk-old with proteinuria of 300 mg/L). Magnifications: ×2.5 in B; ×63 in C through E; ×20 in F.

Ultrastructural analysis of glomeruli by electron microscopy demonstrated normal podocyte FP and GBM morphology in mutant mice without proteinuria (n = 4) and control mice (n = 5). In contrast, FP effacement, splitting of GBM, and irregular GBM were observed in glomeruli of mice with proteinuria (Figure 2, G and H). FP effacement was first detectable in glomeruli of mutant mice with new onset of proteinuria, whereas GBM alterations were observed in mutant mice with advanced proteinuria. These results indicate that FP effacement was the earliest morphologic defect in mutant glomeruli.

Immunofluorescence labeling for podocyte differentiation markers synaptopodin and podocin expression was comparable in control and mutant glomeruli without proteinuria, but both markers were reduced in mutant mice with onset of proteinuria (Figure 3). In contrast, components of GBM, including collagen IV, laminin, and nidogen/entactin,2630 were indistinguishable between control and mutant mice, irrespective of the absence or presence of proteinuria (Supplemental Figure 1S). Together, these observations suggest that podocyte dedifferentiation coincides with onset of FP effacement and proteinuria in mutant mice without detectable difference of GBM components.

Figure 3.

Figure 3.

Synaptopodin and podocin expressions both were downregulated in podocytes of mutant mice with proteinuria (+PU). Different degrees of synaptopodin and podocin downregulations are shown in the middle panels (milder) and right panels (more severe). Magnification, ×40.

Dicer Deletion in Podocyte Leads to Podocyte Apoptosis and Depletion

Podocyte apoptosis has been identified as one cause of podocyte depletion in experimental kidney diseases in mice.3133 Rates of apoptotic podocytes were determined by triple-immunofluorescence labeling including terminal deoxynucleotidyl transferase–mediated digoxigenin-deoxyuridine nick-end labeling (TUNEL) assay, WT1, and DAPI (Figure 4, A through D). Apoptotic podocytes were rare or absent in glomerular sections of control mice or mutant mice without proteinuria, respectively. In contrast, rates of apoptotic podocytes per 100 glomerular sections were increased in mutant mice with low-grade proteinuria or high-grade proteinuria, respectively (Figure 4E). In addition, WT1-positive cells were detectable in tubular lumens of mutant mice with proteinuria but not in control or mutant mice without proteinuria (data not shown). Rates of apoptosis per 100 glomerular sections of WT1-negative cells were also increased in mutant mice with low-grade and high-grade proteinuria, respectively. Apoptosis was detectable in tubular epithelial cells, in particular in mutant mice with high-grade proteinuria (data not shown). Taken together, these results demonstrate that apoptosis of podocytes was closely correlated with the degree of proteinuria in Dicer mutant mice.

Figure 4.

Figure 4.

Mutant podocyte underwent apoptosis, and podocyte number of mutant mice decreased as proteinuria progressed. (A through D) TUNEL assay showing an apoptotic podocyte in the glomerulus of a mutant mouse. WT1, green; TUNEL, red; and DAPI, blue. (E) Rates of apoptotic nuclei in four groups of mice: (1) control mice (n = 9), one apoptotic podocyte found in 930 glomeruli examined; (2) mutant mice without proteinuria (−; n = 4), no apoptotic podocyte in 379 glomeruli; (3) mutant mice with low-grade proteinuria (+; n = 6), three apoptotic podocytes in 590 glomeruli; (4) mutant mice with high-grade proteinuria (++; n = 7), 19 apoptotic podocytes in 874 glomeruli. (F) Podocyte number among four groups of mice as described in E shows podocyte numbers of mutant mice with proteinuria were reduced significantly. **P < 0.01.

WT1 immunostaining was performed to examine whether podocyte number was altered in mutant mice. WT1 and DAPI double-positive cells were counted as podocytes as described previously.32 Five-week-old mutant mice were divided into three groups according to severities of proteinuria: (1) None, (2) low-grade proteinuria (approximately 100 μg/ml), and (3) high-grade proteinuria (≥300 μg/ml). Podocyte counts per glomerular section were comparable between mutant mice without proteinuria (n = 4) and age-matched control mice (n = 11; Figure 4F). In contrast, podocyte counts per glomerular section were significantly decreased (P < 0.01) in mutant mice with low-grade proteinuria (n = 5) and in mutant mice with high-grade proteinuria (n = 5; Figure 4F).

Global Gene Expression Profiling of Glomeruli of Mutant Mice

Inactivation of Dicer/miRNA in podocytes was expected to alter expression of specific miRNA target genes. To identify these genes, we performed microarray analyses on isolated glomeruli from three groups of 4- to 5-wk-old mice: (1) Control (DicerF/+:NPHS2-Cre, 5 mice), (2) mutant without proteinuria (5 mice), and (3) mutant at proteinuria onset (10 mice with proteinuria: 100 to 300 mg/L). Kidneys of mutant mice with new-onset proteinuria were characterized by tubular protein casts and evidence of segmental glomerulosclerosis and mesangial expansion in a few glomeruli, whereas the majority of glomeruli appeared normal by light microscopy. The results from microarray analyses showed that gene expression profiles of control and mutant mice without proteinuria were indistinguishable (data not shown). In contrast, 190 transcripts were significantly upregulated (q < 10%) in mutant mice with new-onset proteinuria compared with control (Supplemental Table 1S). Among those genes, 68 were >1.5-fold elevated (Table 1). A total of 121 genes were downregulated >1.5-fold (Supplemental Table 2S).

Table 1.

68 upregulated transcripts with >1.5-fold change in mutant podocytes

Gene Symbol Gene Name q (%) Log Ratio Fold Change
AU067747 0.00 2.96 7.79
Plod2 Procollagen lysine, 2-oxoglutarate 5-dioxygenase 2 0.00 2.25 4.77
Hspb6 Heat shock protein, alpha-crystallin-related, B6 0.00 2.18 4.54
Mm0.116781 Transcribed locus 0.00 2.09 4.27
Mm0.375137 MRNA similar to angiopoietin-like factor 0.00 2.02 4.06
Dusp14 Dual specificity phosphatase 14 0.00 1.92 3.78
Scel Sciellin 0.00 1.83 3.56
C030009O12Rik RIKEN cDNA C030009O12 gene 0.00 1.73 3.32
Tagln Transgelin 5.71 1.62 3.08
Tnfrsf12a TNF receptor superfamily, member 12a 7.87 1.60 3.04
Gabrb1 γ-Aminobutyric acid (GABA-A) receptor, subunit β1 0.00 1.56 2.95
Spp1 Secreted phosphoprotein 1 3.77 1.55 2.93
AI790298 Expressed sequence AI790298 0.00 1.50 2.84
Mt2 Metallothionein 2 9.69 1.46 2.75
D5Buc30e DNA segment, Chr 5, Bucan 30 expressed 0.00 1.40 2.63
Fgfbp1 Fibroblast growth factor binding protein 1 0.00 1.36 2.57
Krt1–18 Keratin complex 1, acidic, gene 18 5.36 1.33 2.51
F3 Coagulation factor III 2.86 1.31 2.48
Gadd45b Growth arrest and DNA-damage-inducible 45 beta 0.00 1.29 2.45
Serpinb6b Serine (or cysteine) proteinase inhibitor, clade B, 6b 0.00 1.26 2.40
S100a6 S100 calcium binding protein A6 (calcyclin) 9.32 1.21 2.31
2810417M05Rik RIKEN cDNA 2810417M05 gene 4.01 1.10 2.14
Maff v-maf musculoaponeurotic fibrosarcoma oncogene family, F 5.56 1.03 2.05
Mm0.71633 0 day neonate skin cDNA, RIKEN clone:4632424N07 0.00 1.03 2.04
D0H4S114 DNA segment, human D4S114 3.26 1.03 2.04
Masp1 Mannan-binding lectin serine protease 1 0.00 0.99 1.98
Sbk1 SH3-binding kinase 1 0.00 0.99 1.98
Pp11r Placental protein 11 related 0.00 0.97 1.97
BB369191 4.34 0.92 1.89
2900041A09Rik RIKEN cDNA 2900041A09 gene 8.4 0.92 1.89
Ifi30 IFN-γ inducible protein 30 0.00 0.89 1.86
AI323359 4.83 0.88 1.84
Ier3 Immediate early response 3 4.34 0.88 1.84
Ager Advanced glycosylation end product-specific receptor 1.93 0.87 1.83
Bhlhb2 Basic helix-loop-helix domain containing, class B2 7.87 0.87 1.83
Prnp Prion protein 4.91 0.86 1.81
Gmpr Guanosine monophosphate reductase 0.00 0.85 1.80
Acvr2b Activin receptor IIB 0.00 0.84 1.79
Stx11 Syntaxin 11 5.10 0.84 1.79
Tspan4 Tetraspanin 4 2.62 0.82 1.77
Ppm1l Protein phosphatase 1 (formerly 2C)-like 2.95 0.81 1.75
Npepl1 Aminopeptidase-like 1 3.41 0.80 1.74
Sema3b Sema domain, immunoglobulin domain, (semaphorin) 3B 2.62 0.80 1.74
Cldn12 Claudin 12 0.00 0.79 1.72
MGI:2446326 Suprabasin 5.36 0.76 1.69
B3galt2 UDP-Gal:βGlcNAc β1,3-galactosyltransferase, polypeptide 2 4.34 0.76 1.69
9530018I07Rik RIKEN cDNA 9530018I07 gene 6.55 0.75 1.68
Tsrc1 Thrombospondin repeat containing 1 3.26 0.73 1.66
Ppp1r3c Protein phosphatase 1, regulatory (inhibitor) subunit 3C 6.30 0.73 1.66
Efemp1 EGF-containing fibulin-like extracellular matrix protein 1 2.99 0.73 1.66
Mthfd1l Methylenetetrahydrofolate dehydrogenase 1-like 3.32 0.70 1.63
AI506321 6.89 0.70 1.62
Hmgb3 High mobility group box 3 0.00 0.69 1.62
E030024M05Rik RIKEN cDNA E030024M05 gene 3.26 0.69 1.62
5730499H23Rik RIKEN cDNA 5730499H23 gene 4.25 0.68 1.60
Sulf2 Sulfatase 2 5.10 0.68 1.60
Eno1 Enolase 1, α non-neuron 2.39 0.68 1.60
Chst12 Carbohydrate sulfotransferase 12 3.77 0.67 1.59
0610010O12Rik RIKEN cDNA 0610010O12 gene 0.00 0.66 1.58
MGI:1917275 Brain acyl-CoA hydrolase 3.32 0.65 1.57
Mvp Major vault protein 0.00 0.65 1.57
Hlf Hepatic leukemia factor 8.40 0.64 1.56
Aplp1 Amyloid β (A4) precursor-like protein 1 7.87 0.64 1.55
Mustn1 Musculoskeletal, embryonic nuclear protein 1 9.69 0.63 1.55
Vim Vimentin 0.00 0.63 1.55
Gclc Glutamate-cysteine ligase, catalytic subunit 8.40 0.61 1.53
Mm0.361409 Transcribed locus 2.79 0.60 1.51
NM_011937 2.94 0.59 1.51

We selected 14 of the 68 upregulated genes for validation of microarray measurements by quantitative real-time PCR. The upregulation of 13 of 14 genes was confirmed (Supplemental Figure 2S). To confirm that the upregulation of gene expression was specifically in podocytes but not in other types of glomerular cells, we conducted immunofluorescence staining using available antibody against SM22α (Transgelin), a smooth muscle differentiation marker, and found that SM22α signal co-localized with synaptopodin (Supplemental Figure 3S), indicating that the upregulation of SM22α was observed in podocytes of Dicer mutant mice with proteinuria.

Identification of miR-30 Family miRNA and Their 15 Upregulated Target Genes as Candidate miRNA–Target Gene Pairs in Dicer Mutant Podocytes

miRNA have been shown to degrade transcripts directly, resulting in negative regulation of steady-state mRNA levels. Thus, we hypothesized that genes with increased mRNA levels in mutant glomeruli may be direct targets for miRNA-mediated RNA degradation.

If some of the 190 upregulated transcripts of mutant podocytes are indeed miRNA degradation targets, then select miRNA target sequences would be expected to be overrepresented in the 3′ untranslated regions (UTR) of the 190 transcripts. To examine this hypothesis, we used miRBase (http://microrna.sanger.ac.uk) to search for miRNA target sequences in the 3′ UTR of the 190 transcripts, and we found target sequences for 294 miRNA. The target sequence number of individual miRNA ranges from 1 to 19. Next, we performed permutation analyses on the occurrence of miRNA–target gene pairs to determine eight miRNA whose target sequences were significantly overrepresented (false discovery rate <5%) in the 190 upregulated transcripts (Table 2). Among them, six were previously shown to be represented in miRNA library cloning from podocyte culture according to smiRNAdb (http://www.mirz.unibas.ch/smiRNAdb/). These included miR-28; miR-34a; and four members of the miR-30 family, miR-30c-1, miR-30b, miR-30d, and miR-30c-2. Among 68 upregulated genes with >1.5-fold change, 15 contained miR-30 family target sequences, including 0610010O12Rik, Ager (receptor for advanced glycation end products [RAGE]), Aplp1, Chst12, Gabrb1, Gclc, Hspb6, Ier3, Ifi30, Mthfd1l, Npepl1, Scel, Sulf2, Tnfrsf12a, Tspan4, and Vim (Supplemental Table 4S). The miR-30 family members fulfilled all criteria for candidate miRNA. First, miR-30d was regulated in mutant glomeruli. Second, all four miR-30 miRNA are normally present in podocytes. Third, miR-30 miRNA are known RNA target–degrading miRNA. Fourth, miR-30 target sequences were highly enriched in upregulated genes in mutant glomeruli.

Table 2.

miRNA with target sequence significantly enriched in 190 upregulated transcripts

miRNA Hits FDR Podocyte Expressiona
mmu-mir-30c-1 13 0.0297 x
mmu-mir-30b 19 0.0359 x
mmu-mir-28 16 0.0359 x
mmu-mir-34a 14 0.0359 x
mmu-mir-30d 13 0.0359 x
mmu-mir-762 13 0.0359
mmu-mir-30c-2 12 0.0359 x
mmu-mir-711 11 0.0359
a

The miRNA present in podocytes according to smiRNAdb (http://www.mirz.unibas.ch/smiRNAdb/) based on limited podocyte miRNA library sequencing (only 888 clones sequenced) are marked with “x.’

DISCUSSION

Mice mutant for Dicer in podocytes manifested variable expression of progressive proteinuria, structural glomerular defects and glomerulosclerosis consistent with the podocyte depletion paradigm,24,25 and tubulointerstitial fibrosis, demonstrating that miRNA exert essential genomic controls in podocytes. In an attempt to delineate the temporal sequence of the observed endophenotypes, we obtained evidence for podocyte dedifferentiation and FP effacement with GBM defects as the earliest detectable defects, followed by podocyte apoptosis, podocyte depletion, mesangial expansion, and segmental or global glomerulosclerosis. The age of onset and the extent of the observed defects were highly variable among mutant mice. This is likely attributable to variation in expression of transgenic Cre recombinase and Cre-dependent inactivation of Dicer in podocytes in different glomeruli within an individual mouse and/or between different mice, as has been previously observed with the NPHS2-Cre transgenic model.34 Interestingly, histologic and ultrastructural glomerular defects detectable in all mutant mice were coincident only with the onset and progression of proteinuria but not before proteinuria onset.

In an unbiased approach to narrow the spectrum of molecular targets and cognate miRNA responsible for these defects, microarray analysis of glomerular gene expression profiles combined with in silico interrogation of genes with significant upregulation in mutant glomeruli identified candidate molecular mechanisms that may underlie the observed defects. The validity of our applied microarray and in silico approach to identify target-degrading miRNA for which Dicer deficiency-induced loss of expression would be expected to result directly in increased expression of their respective target transcripts is supported by the following observations. First, six of the eight most significant candidate target-degrading miRNA are known to be present in podocytes, on the basis of limited sequencing information available in the smiRNAdb database (http://www.mirz.unibas.ch/smiRNAdb/), on miRNA libraries established from murine podocytes. Second, the identification of four known target mRNA–degrading miRNA of the miR-30 family,35 all of which are expressed in podocytes, among the six most significant candidate podocyte miRNA is highly significant. For example, the observed frequency of hits among the 190 upregulated genes for miR-30 family miRNA indicated between 12 and 19 targets per miR-30 miRNA, compared with the expected frequency of five hits for random miRNA among the 190 upregulated transcripts. Moreover, only a few target-degrading miRNA have been identified (let-7b,36,37 miR-125b,36 miR-1,38 miR-124,38 miR-122,39 miR-16,40 miR-98,41 and miR-106b42) to date in addition to the miR-30 family.

Among the list of 15 target genes of the mir-30 family in our study, several are known to be expressed in injured podocytes in experimental models and human kidney disease and have functional roles in apoptosis. For example, RAGE is barely detectable in podocytes in normal control mice but is strongly expressed in the streptozotocin murine models of diabetic kidney disease and of adriamycin-induced nondiabetic podocyte injury associated with proteinuria and glomerulosclerosis.43,44 Furthermore, RAGE is increased in podocytes of patients with diabetic nephropathy and several nondiabetic glomerular diseases.45 Intriguingly, the known RAGE ligand S100 calcium-binding protein A6 (calcyclin) was among the top 20 upregulated genes in glomeruli of mutant mice with proteinuria (see Table 1). S100A6 binding with RAGE is known to induce apoptosis in epithelial cells through activation of Jun N-terminal kinase.46 Thus, aberrant coincident upregulation of RAGE and its proapoptotic cognate ligand S100A6 may be caused by loss of podocyte mmu-mir-30b and possibly mmu-mir-431, respectively, and may directly contribute to podocyte injury and/or podocyte apoptosis observed in proteinuric mutant mice. The Ier3 is a predicted target of both mmu-mir-30c and mmu-mir-431. Ier3 also functions as positive regulator of apoptosis in keratinocytes47 and has been identified in renal cDNA libraries as protein that interacts with modulators of apoptosis.48 Podocyte apoptosis is an early glomerular manifestation leading to podocyte depletion in TGF-β1 transgenic mice and CD2AP−/− and diabetic mouse models with albuminuria and glomerulosclerosis.3133 Observations in this study provide a rationale for a role of select miRNA–target gene pairs, such as miR-30 and miR-431 and RAGE and S100A6, in progressive glomerular disease.

Podocytes possess major processes containing microtubules and intermediate filaments with vimentin, and FP containing actin filaments as core cytoskeletal elements.49 Heat-shock protein 20 (Hsp20) and the intermediate filament protein vimentin both are shared targets of mmu-mir-30c-1 and 30c-2 and interact with cytoskeletal components. The intermediate filament vimentin is upregulated in podocytes of nephrotic glomeruli in puromycin aminonucleoside model in rat.50 Hsp20 binds to actin in vitro and in vivo, and the association with filamentous actin is dependent on the phosphorylation state of Hsp20.51 Interestingly, phosphorylated Hsp20 has a direct role in smooth muscle contractility and relaxation.52 In addition, Tagln (SM22α and transgelin) is an actin cross-linking protein.53,54 The de novo expression of SM22α in podocytes of mutant glomeruli observed in our studies may disrupt the integrity of the actin cytoskeleton in podocytes and lead to FP effacement. In addition, Plod2 (lysyl hydroxylase 2),55 is one of the three lysyl hydroxylases that mediate lysine hydroxylation of collagens, including collagen IV, which is essential for their proper assembly and cross-linking of collagen fibrils.56 Upregulation of lysyl hydroxylase 2 may therefore underlie alterations of the collagen IV network and thus GBM defects in Dicer mutant mice. Pathway enrichment analysis based on gene expression data identified activation of integrin signaling as top pathway, consistent with the observed disruption of podocyte adhesion to GBM and GBM alterations. Integrin α-3 (Itga3) is the major α subunit of β1 integrin receptors expressed by podocytes, and podocytes in kidneys of integrin α-3 knockout (Itga3−/−) mice are unable to maintain normal podocyte structure, including the elaboration of mature FP along the GBM.57 Indeed, Itga3 mRNA was increased in mutant mice, consistent with integrin signaling activation. Functional validation of the discussed candidate miR-30–target gene pairs requires an expansive experimental approach and is ongoing in our laboratory.

In conclusion, our results demonstrate that miRNA are important genomic regulators of molecular podocyte homeostasis. Disruption of miRNA-controlled gene networks results in progressive proteinuria and glomerulosclerosis. These findings warrant further studies to examine whether the dysregulation of miRNA-controlled gene networks may contribute to the pathogenesis of glomerular diseases. The identification of candidate miRNA–target gene pairs in our study may provide a starting point for further study to validate their role in glomerular diseases.

CONCISE METHODS

Mouse Strain

The use of the mice in this study adhered to the National Institutes of Health Guide for the Care and Use of Laboratory Animals. The mouse line carrying a Dicer allele with two lox P sites that flank RNase III domain was created as described previously.8 NPHS2-Cre was a gift from Dr. Holzman.34 Rosa26 reporter mice were purchased from Jackson Laboratory (Bar Harbor, ME).

Antibodies

The antibodies used in this study were as follows: WT1 (Santa Cruz Biotechnology, Santa Cruz, CA), SM22α (Abcam, Cambridge, MA), laminin (Sigma, St. Louis, MO), entactin/nidogen (Chemicon Int., Temecula, CA), collagen IV (Southern Biotech, Birmingham, AL), and podocin (Santa Cruz, Santa Cruz, CA). Monoclonal synaptopodin antibody was prepared from hybridoma cells. Fluor-conjugated second antibodies, Alexa Fluor 488 goat anti-rabbit IgG (H+L) and Alexa Fluor 568 goat anti-mouse Ig G1, both were purchased from Invitrogen (Carlsbad, CA).

β-Galactosidase Staining of Kidney

The kidneys were excised from the mice, imbedded in OCT and snap-frozen. Five-micrometer sections were cut and stained for the expression of β-galactosidase using β-galactosidase Staining Kit (Specialty Media, Phillipsburg, NJ) following the manufacturer's instructions.

PCR for Genotyping

PCR to detect the deletion product in the podocytes was conducted using the primers D1 (5′-AGTAATGTGAGCAATAGTCCCAG-3′) and D3 (5′-CTGGTGGCTTGAGGACAAGAC-3′), which locate outside the loxP floxed region. PCR reaction was performed with standard protocol using TaqDNA polymerase (Invitrogen). The thermal cycling condition was 94°C for 2 min, followed by 35 cycles of 94°C for 15 s, 57°C for 15 s, and 72°C for 30 s. The predicted size of the PCR product derived from the allele undergone deletion is 309 bp.

Proteinuria Measurement

Strips of Albustix (Bayer Corp., Mishawaka, IN) were used to measure the proteinuria of mice at approximately 6 p.m. on the day of measurement.

Histology

Mice were perfused with 4% paraformaldehyde followed by 18% sucrose PBS solution. The kidneys were cut into two halves, one imbedded with OCT and snap-frozen, another fixed with 10% formalin overnight for paraffin embedding. Four-micrometer sections were cut for both frozen and paraffin-embedded kidneys. Periodic acid-Schiff staining was performed following the protocol recommended by the Animal Models of Diabetic Complications Consortium (http://www.amdcc.org). Immunostaining was conducted briefly as follows: OCT-embedded frozen sections were cut (4 μm), air-dried, and incubated with blocking solution at room temperature for 30 min. Then the sections were incubated with the first antibody for 1 h at room temperature. After PBS wash for three times, second antibody was added and incubated at room temperature for 1 h. The sections were then washed, stained with DAPI, and sealed with Fluoromount-G (Southern Biotech).

Podocyte Number Counting

OCT-embedded frozen kidneys were used. Four-micrometer-thick sections were cut, and rabbit polyclonal WT1 antibody (C-19; Santa Cruz) and the second antibody goat anti-rabbit IgG (H+L) Alex Fluor 488 (Invitrogen, Carlsbad, CA) were used to stain podocytes, followed by DAPI staining. WT1 and DAPI double-positive cells were counted as podocytes. All WT1 and DAPI double-positive cells across the whole section were counted, added up, and then divided by the total number of glomeruli on the section, resulting in the average podocyte number per glomerular cross-section.

TUNEL Assay

The apoptotic cells were detected using ApopTag Red In Situ Apoptosis Detection Kit (Chemicon Int., Temecula, CA) on frozen kidney sections. For identification of apoptotic podocytes, the kidney sections were co-stained with WT1 antibody and DAPI. The triple-positive cells, WT1 (green), TUNEL (red), and DAPI (blue), were considered to be apoptotic podocytes. The ones positive for WT1 and DAPI but negative for TUNEL were nonapoptotic podocytes.

Electron Microscopy

A small piece of kidney from cortex was collected, sliced into 1-mm3 cubes, and fixed in 2% glutaraldehyde solution. The kidney samples were then processed following standard protocol.

Isolation of Glomeruli

The procedure for isolation of glomeruli was based on a published method with modifications.58 Briefly, mice were perfused with 2.5 mg/ml iron oxide solution in PBS. Then the two kidneys from a mouse were diced into 1-mm3 pieces. A total of 100 μl of collagenase A solution (10 mg/ml) and 100 μl of DNase I (1000 U/ml) were added to the tissues, and the mix was incubated at 37°C for 30 min with rotation. After passing the disrupted tissues through a 100-μm cell culture strainer, several times of magnetic concentration of glomeruli were performed to obtain purified glomeruli.

Preparation of Total RNA and miRNA-Enriched Small RNA from Glomeruli

Glomerular total RNA and small RNA (200 bp) fraction with miRNA enriched were prepared using mirVana miRNA isolation kit (Ambion, Austin, TX) according to the manufacturer's instructions.

Microarray

Affymetrix Genechip Mouse Genome 430 2.0 Array containing approximately 39,000 gene probes was used for gene expression profiling of the glomeruli of mutant and control mice using total RNA from purified glomeruli. The whole procedure followed the manual instruction of the product. The Affymetrix Genechip data analysis was performed using R (R-Group) and Bioconductor59 statistics packages in combination with other widely known GeneChip analysis tools. Briefly, for preprocessing of the chip data, the raw data were normalized across the chips by log scale robust multi-array analysis (GCRMA) method,60 which has been shown to give more accurate results than the standard algorithm supplied by Affymetrix, and the probe sets with absent calls will be eliminated for downstream analysis. For genes with multiple probe sets on the chip, the median of the expression values of duplicated probe sets was calculated to represent the expression value for corresponding genes. For identification of differentially expressed genes between two groups, the statistical package Significance Analysis of Microarrays61 was used.

Quantitative Real-Time PCR

For mature miRNA quantitative real-time PCR, Ncode miRNA First-Strand cDNA Synthesis kit and Ncode SYBR Green miRNA quantitative real-time PCR kit (Invitrogen) were used following the manufacturer's instructions. For precursor miRNA quantitative real-time PCR, reverse transcriptions using pre-miRNA–specific primers (premiR-30a, 5′-GCAGCTGCAAACATCCGACTGAA-3′; premiR-30c-2, 5′-AAGGCAGAGAGAGTAAACAGCCTT-3′; and miR-30d, 5′-GAGCCAGTAGCAGCAAACATCT-3′) and SuperScript reverse transcriptase III (Invitrogen) were conducted, followed by quantitative PCR using these precursor-specific primers and the primers corresponding to mature miRNA sequences. For mRNA quantitative real-time PCR, SYBR Green PCR master mix (Applied Biosystems). The primers used were as follows: RAMP1-AU067747 (forward 5′-GGGATGAGAAGAGCATCCAGAAT-3′, reverse 5′-CGGGTACTGATGCGAGCTTTGTC-3′); Ccbe1(AV264768; forward 5′-GAC CGCGAGAGACACCAAAAGC-3′, reverse 5′-CCCATCGTCCTCAAGGATGTAG-3′); Hsp20 (forward 5′-CCCCAGTGTGGCGTTACCCACAG-3′, reverse 5′-GCGGTGGAACTCTCGAGCAAT-3′); Scel (forward 5′-AACAGGGTTTTCAGGACGTG-3′, reverse 5′-TGTTTGCATTGAAGGAGCTG-3′); Plod2 (forward 5′- GCCGTCTGGTCCAGCAGTGGAAT-3′, reverse 5′-GGCCTGGAAAATTTTGCATTTG-3′); Dusp14 (forward 5′-GGCTCCAG CTCCCTGGAA ATCCT-3′, reverse 5′-GGGAGCGTGCTGTGACCTCTGGA-3′); AI790298 (forward 5′- CGGCAAGCAGCTGCTGGTGATC-3′, reverse 5′-CCGGTGCTCAGTCACATCAAAGT-3′); Gadd45b (forward 5′-GGCGAGCGACAACGCGGTTCAG-3′, reverse 5′-CATCCTCCTCTTCTTCGTCTATGG-3′); Serpinb6b (forward 5′-CAAGAAATGCCTTTCAATGTCAC-3′, reverse 5′-ATTTCTTATAAGTTATCTCCTTTTCC-3′); Sbk1 (forward 5′-CCCTTCATCATCAAGGTCTTTGAC-3′, reverse 5′-CTGCCTGCTGTGCATGAAGTCCA-3′); Ager (forward 5′-CCCTCCTCAGGTCCACTGGATAA-3′, reverse 5′-TGTGACCCTGATGCTGACAGGA-3′); Acvr2b (forward 5′-CGGCCTGGCTGTTCGGTTTGAG-3′, reverse 5′-GCCCATGGCGTACATGTCGATAC-3′); Tagln (forward 5′-CCTTCCAGTCCACAAACGACCAA-3′, reverse 5′-GGCCACACTGCACTACAATCCA-3′); Nphs2 (forward 5′-CGGTGGAAGCTGAGGCACAAAGA-3′, reverse 5′-GCGACTGAAGAGTGTGCAAGTAT-3′). ABI real-time PCR machine 7900HT was used for all kinds of quantitative PCR described.

Identification of miRNA with Enriched Target Sequence in Upregulated Transcripts

To identify the potential miRNA candidates whose suppression by dicer knockout caused upregulation of their targeting genes, we used Sanger's miRNA target database (http://www.miRNA.org) to search for miRNA targeting 3′ UTR sequences of significantly upregulated genes and the occurrence for each miRNA was obtained. For determination of the significance of miRNA occurrence, the same number of genes as miRNA targets in upregulated genes were randomly chosen from the chip gene list and the occurrence of a miRNA was counted. Such permutation steps were repeated 10,000 times. The P value was calculated as the percentage of occurrences of a miRNA for 10,000 randomly generated gene lists greater than that for upregulated genes and then adjusted by false discovery rate.

DISCLOSURES

None.

Supplementary Material

[Supplemental Data]

Acknowledgments

This work was supported by National Institutes of Health grants RO1DK060043, RO1DK056077, RO1DK073960, and UO1DK060995 to E.P.B.

Part of this work was presented at the annual meeting of the American Society of Nephrology; October 31 through November 7, 2007; San Francisco, CA.

Microscopy was performed at the Mount Sinai-Microscopy Shared Resource Facility, supported with funding from 5R24CA095823, DBI-9724504, and 1510 RRO 9145-01.

Published online ahead of print. Publication date available at www.jasn.org.

Supplemental information for this article is available online at http://www.jasn.org/.

See related Occasional Observation, “Dicer Cuts the Kidney,” on pages 2043–2046.

REFERENCES

  • 1.Pillai RS, Bhattacharyya SN, Filipowicz W: Repression of protein synthesis by miRNAs: How many mechanisms? Trends Cell Biol 17: 118–126, 2007 [DOI] [PubMed] [Google Scholar]
  • 2.Lee Y, Kim M, Han J, Yeom KH, Lee S, Baek SH, Kim VN: MicroRNA genes are transcribed by RNA polymerase II. EMBO J 23: 4051–4060, 2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Borchert GM, Lanier W, Davidson BL: RNA polymerase III transcribes human microRNAs. Nat Struct Mol Biol 13: 1097–1101, 2006 [DOI] [PubMed] [Google Scholar]
  • 4.Gregory RI, Chendrimada TP, Cooch N, Shiekhattar R: Human RISC couples microRNA biogenesis and posttranscriptional gene silencing. Cell 123: 631–640, 2005 [DOI] [PubMed] [Google Scholar]
  • 5.Bohnsack MT, Czaplinski K, Gorlich D: Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA 10: 185–191, 2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Grishok A, Pasquinelli AE, Conte D, Li N, Parrish S, Ha I, Baillie DL, Fire A, Ruvkun G, Mello CC: Genes and mechanisms related to RNA interference regulate expression of the small temporal RNAs that control C. elegans developmental timing. Cell 106: 23–34, 2001 [DOI] [PubMed] [Google Scholar]
  • 7.Bernstein E, Kim SY, Carmell MA, Murchison EP, Alcorn H, Li MZ, Mills AA, Elledge SJ, Anderson KV, Hannon GJ: Dicer is essential for mouse development. Nat Genet 35: 215–217, 2003 [DOI] [PubMed] [Google Scholar]
  • 8.Cobb BS, Nesterova TB, Thompson E, Hertweck A, O'Connor E, Godwin J, Wilson CB, Brockdorff N, Fisher AG, Smale ST, Merkenschlager M: T cell lineage choice and differentiation in the absence of the RNase III enzyme dicer. J Exp Med 201: 1367–1373, 2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Cobb BS, Hertweck A, Smith J, O'Connor E, Graf D, Cook T, Smale ST, Sakaguchi S, Livesey FJ, Fisher AG, Merkenschlager M: A role for dicer in immune regulation. J Exp Med 203: 2519–2527, 2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Murchison EP, Stein P, Xuan Z, Pan H, Zhang MQ, Schultz RM, Hannon GJ: Critical roles for dicer in the female germline. Genes Dev 21: 682–693, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Tang F, Kaneda M, O'Carroll D, Hajkova P, Barton SC, Sun YA, Lee C, Tarakhovsky A, Lao K, Surani MA: Maternal microRNAs are essential for mouse zygotic development. Genes Dev 21: 644–648, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Harfe BD, McManus MT, Mansfield JH, Hornstein E, Tabin CJ: The RNaseIII enzyme dicer is required for morphogenesis but not patterning of the vertebrate limb. Proc Natl Acad Sci U S A 102: 10898–10903, 2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Schaefer A, O'Carroll D, Tan CL, Hillman D, Sugimori M, Llinas R, Greengard P: Cerebellar neurodegeneration in the absence of microRNAs. J Exp Med 204: 1553–1558, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Kobayashi T, Lu J, Cobb BS, Rodda SJ, McMahon AP, Schipani E, Merkenschlager M, Kronenberg HM: Dicer-dependent pathways regulate chondrocyte proliferation and differentiation. Proc Natl Acad Sci U S A 105: 1949–1954, 2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Chen JF, Murchison EP, Tang R, Callis TE, Tatsuguchi M, Deng Z, Rojas M, Hammond SM, Schneider MD, Selzman CH, Meissner G, Patterson C, Hannon GJ, Wang DZ: Targeted deletion of dicer in the heart leads to dilated cardiomyopathy and heart failure. Proc Natl Acad Sci U S A 105: 2111–2116, 2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.van Rooij E, Sutherland LB, Qi X, Richardson JA, Hill J, Olson EN: Control of stress-dependent cardiac growth and gene expression by a microRNA. Science 316: 575–579, 2007 [DOI] [PubMed] [Google Scholar]
  • 17.Thai TH, Calado DP, Casola S, Ansel KM, Xiao C, Xue Y, Murphy A, Frendewey D, Valenzuela D, Kutok JL, Schmidt-Supprian M, Rajewsky N, Yancopoulos G, Rao A, Rajewsky K: Regulation of the germinal center response by microRNA-155. Science 316: 604–608, 2007 [DOI] [PubMed] [Google Scholar]
  • 18.Rodriguez A, Vigorito E, Clare S, Warren MV, Couttet P, Soond DR, van Dongen S, Grocock RJ, Das PP, Miska EA, Vetrie D, Okkenhaug K, Enright AJ, Dougan G, Turner M, Bradley A: Requirement of bic/microRNA-155 for normal immune function. Science 316: 608–611, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kloosterman WP, Lagendijk AK, Ketting RF, Moulton JD, Plasterk RH: Targeted inhibition of miRNA maturation with morpholinos reveals a role for miR-375 in pancreatic islet development. PLoS Biol 5: e203, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Poy MN, Eliasson L, Krutzfeldt J, Kuwajima S, Ma X, Macdonald PE, Pfeffer S, Tuschl T, Rajewsky N, Rorsman P, Stoffel M: A pancreatic islet-specific microRNA regulates insulin secretion. Nature 432: 226–230, 2004 [DOI] [PubMed] [Google Scholar]
  • 21.Bandres E, Agirre X, Ramirez N, Zarate R, Garcia-Foncillas J: MicroRNAs as cancer players: Potential clinical and biological effects. DNA Cell Biol 26: 273–282, 2007 [DOI] [PubMed] [Google Scholar]
  • 22.Dykxhoorn DM: MicroRNAs in viral replication and pathogenesis. DNA Cell Biol 26: 239–249, 2007 [DOI] [PubMed] [Google Scholar]
  • 23.Kato M, Zhang J, Wang M, Lanting L, Yuan H, Rossi JJ, Natarajan R: MicroRNA-192 in diabetic kidney glomeruli and its function in TGF-beta-induced collagen expression via inhibition of E-box repressors. Proc Natl Acad Sci U S A 104: 3432–3437, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Kriz W, Gretz N, Lemley KV: Progression of glomerular diseases: Is the podocyte the culprit? Kidney Int 54: 687–697, 1998 [DOI] [PubMed] [Google Scholar]
  • 25.Wiggins RC: The spectrum of podocytopathies: A unifying view of glomerular diseases. Kidney Int 71: 1205–1214, 2007 [DOI] [PubMed] [Google Scholar]
  • 26.Akhtar M, Al Mana H: Molecular basis of proteinuria. Adv Anat Pathol 11: 304–309, 2004 [DOI] [PubMed] [Google Scholar]
  • 27.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 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Miner JH: Renal basement membrane components. Kidney Int 56: 2016–2024, 1999 [DOI] [PubMed] [Google Scholar]
  • 29.Miner JH, Patton BL: Laminin-11. Int J Biochem Cell Biol 31: 811–816, 1999 [DOI] [PubMed] [Google Scholar]
  • 30.Jarad G, Cunningham J, Shaw AS, Miner JH: Proteinuria precedes podocyte abnormalities in Lamb2−/− mice, implicating the glomerular basement membrane as an albumin barrier. J Clin Invest 116: 2272–2279, 2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Schiffer M, Bitzer M, Roberts IS, Kopp JB, ten Dijke P, Mundel P, Bottinger EP: Apoptosis in podocytes induced by TGF-beta and Smad7. J Clin Invest 108: 807–816, 2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Schiffer M, Mundel P, Shaw AS, Bottinger EP: A novel role for the adaptor molecule CD2-associated protein in transforming growth factor-beta-induced apoptosis. J Biol Chem 279: 37004–37012, 2004 [DOI] [PubMed] [Google Scholar]
  • 33.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 [PubMed] [Google Scholar]
  • 34.Moeller MJ, Sanden SK, Soofi A, Wiggins RC, Holzman LB: Podocyte-specific expression of cre recombinase in transgenic mice. Genesis 35: 39–42, 2003 [DOI] [PubMed] [Google Scholar]
  • 35.Schmitter D, Filkowski J, Sewer A, Pillai RS, Oakeley EJ, Zavolan M, Svoboda P, Filipowicz W: Effects of dicer and argonaute down-regulation on mRNA levels in human HEK293 cells. Nucleic Acids Res 34: 4801–4815, 2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Wu L, Fan J, Belasco JG: MicroRNAs direct rapid deadenylation of mRNA. Proc Natl Acad Sci U S A 103: 4034–4039, 2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Bagga S, Bracht J, Hunter S, Massirer K, Holtz J, Eachus R, Pasquinelli AE: Regulation by let-7 and lin-4 miRNAs results in target mRNA degradation. Cell 122: 553–563, 2005 [DOI] [PubMed] [Google Scholar]
  • 38.Lim LP, Lau NC, Garrett-Engele P, Grimson A, Schelter JM, Castle J, Bartel DP, Linsley PS, Johnson JM: Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 433: 769–773, 2005 [DOI] [PubMed] [Google Scholar]
  • 39.Esau C, Davis S, Murray SF, Yu XX, Pandey SK, Pear M, Watts L, Booten SL, Graham M, McKay R, Subramaniam A, Propp S, Lollo BA, Freier S, Bennett CF, Bhanot S, Monia BP: miR-122 regulation of lipid metabolism revealed by in vivo antisense targeting. Cell Metab 3: 87–98, 2006 [DOI] [PubMed] [Google Scholar]
  • 40.Linsley PS, Schelter J, Burchard J, Kibukawa M, Martin MM, Bartz SR, Johnson JM, Cummins JM, Raymond CK, Dai H, Chau N, Cleary M, Jackson AL, Carleton M, Lim L: Transcripts targeted by the microRNA-16 family cooperatively regulate cell cycle progression. Mol Cell Biol 27: 2240–2252, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Hebert C, Norris K, Scheper MA, Nikitakis N, Sauk JJ: High mobility group A2 is a target for miRNA-98 in head and neck squamous cell carcinoma. Mol Cancer 6: 5, 2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Ivanovska I, Ball AS, Diaz RL, Magnus JF, Kibukawa M, Schelter JM, Kobayashi SV, Lim L, Burchard J, Jackson AL, Linsley PS, Cleary MA: MicroRNAs in the miR-106b family regulate p21/CDKN1A and promote cell cycle progression. Mol Cell Biol 28: 2167–2174, 2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Wendt TM, Tanji N, Guo J, Kislinger TR, Qu W, Lu Y, Bucciarelli LG, Rong LL, Moser B, Markowitz GS, Stein G, Bierhaus A, Liliensiek B, Arnold B, Nawroth PP, Stern DM, D'Agati VD, Schmidt AM: RAGE drives the development of glomerulosclerosis and implicates podocyte activation in the pathogenesis of diabetic nephropathy. Am J Pathol 162: 1123–1137, 2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Guo J, Ananthakrishnan R, Qu W, Lu Y, Reiniger N, Zeng S, Ma W, Rosario R, Yan SF, Ramasamy R, D'Agati V, Schmidt AM: RAGE mediates podocyte injury in adriamycin-induced glomerulosclerosis. J Am Soc Nephrol 19: 951–972, 2008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Abel M, Ritthaler U, Zhang Y, Deng Y, Schmidt AM, Greten J, Sernau T, Wahl P, Andrassy K, Ritz E: Expression of receptors for advanced glycosylated end-products in renal disease. Nephrol Dial Transplant 10: 1662–1667, 1995 [PubMed] [Google Scholar]
  • 46.Leclerc E, Fritz G, Weibel M, Heizmann CW, Galichet A: S100B and S100A6 differentially modulate cell survival by interacting with distinct RAGE (receptor for advanced glycation end products) immunoglobulin domains. J Biol Chem 282: 31317–31331, 2007 [DOI] [PubMed] [Google Scholar]
  • 47.Schilling D, Pittelkow MR, Kumar R: IEX-1, an immediate early gene, increases the rate of apoptosis in keratinocytes. Oncogene 20: 7992–7997, 2001 [DOI] [PubMed] [Google Scholar]
  • 48.Kumar R, Lutz W, Frank E, Im HJ: Immediate early gene X-1 interacts with proteins that modulate apoptosis. Biochem Biophys Res Commun 323: 1293–1298, 2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Kobayashi N, Reiser J, Schwarz K, Sakai T, Kriz W, Mundel P: Process formation of podocytes: Morphogenetic activity of microtubules and regulation by protein serine/threonine phosphatase PP2A. Histochem Cell Biol 115: 255–266, 2001 [DOI] [PubMed] [Google Scholar]
  • 50.Zou J, Yaoita E, Watanabe Y, Yoshida Y, Nameta M, Li H, Qu Z, Yamamoto T: Upregulation of nestin, vimentin, and desmin in rat podocytes in response to injury. Virchows Arch 448: 485–492, 2006 [DOI] [PubMed] [Google Scholar]
  • 51.Fan GC, Chu G, Mitton B, Song Q, Yuan Q, Kranias EG: Small heat-shock protein Hsp20 phosphorylation inhibits beta-agonist-induced cardiac apoptosis. Circ Res 94: 1474–1482, 2004 [DOI] [PubMed] [Google Scholar]
  • 52.Fan GC, Chu G, Kranias EG: Hsp20 and its cardioprotection. Trends Cardiovasc Med 15: 138–141, 2005 [DOI] [PubMed] [Google Scholar]
  • 53.Fu Y, Liu HW, Forsythe SM, Kogut P, McConville JF, Halayko AJ, Camoretti-Mercado B, Solway J: Mutagenesis analysis of human SM22: Characterization of actin binding. J Appl Physiol 89: 1985–1990, 2000 [DOI] [PubMed] [Google Scholar]
  • 54.Zhang JC, Kim S, Helmke BP, Yu WW, Du KL, Lu MM, Strobeck M, Yu Q, Parmacek MS: Analysis of SM22alpha-deficient mice reveals unanticipated insights into smooth muscle cell differentiation and function. Mol Cell Biol 21: 1336–1344, 2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Morgan KG, Gangopadhyay SS: Invited review: Cross-bridge regulation by thin filament-associated proteins. J Appl Physiol 91: 953–962, 2001 [DOI] [PubMed] [Google Scholar]
  • 56.Rautavuoma K, Takaluoma K, Sormunen R, Myllyharju J, Kivirikko KI, Soininen R: Premature aggregation of type IV collagen and early lethality in lysyl hydroxylase 3 null mice. Proc Natl Acad Sci U S A 101: 14120–14125, 2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Kreidberg JA, Donovan MJ, Goldstein SL, Rennke H, Shepherd K, Jones RC, Jaenisch R: Alpha 3 beta 1 integrin has a crucial role in kidney and lung organogenesis. Development 122: 3537–3547, 1996 [DOI] [PubMed] [Google Scholar]
  • 58.Takemoto M, Asker N, Gerhardt H, Lundkvist A, Johansson BR, Saito Y, Betsholtz C: A new method for large scale isolation of kidney glomeruli from mice. Am J Pathol 161: 799–805, 2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Gentleman RC, Carey VJ, Bates DM, Bolstad B, Dettling M, Dudoit S, Ellis B, Gautier L, Ge Y, Gentry J, Hornik K, Hothorn T, Huber W, Iacus S, Irizarry R, Leisch F, Li C, Maechler M, Rossini AJ, Sawitzki G, Smith C, Smyth G, Tierney L, Yang JY, Zhang J: Bioconductor: Open software development for computational biology and bioinformatics. Genome Biol 5: R80, 2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Irizarry RA, Hobbs B, Collin F, Beazer-Barclay YD, Antonellis KJ, Scherf U, Speed TP: Exploration, normalization, and summaries of high density oligonucleotide array probe level data. Biostatistics 4: 249–264, 2003 [DOI] [PubMed] [Google Scholar]
  • 61.Tusher VG, Tibshirani R, Chu G: Significance analysis of microarrays applied to the ionizing radiation response. Proc Natl Acad Sci U S A 98: 5116–5121, 2001 [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

[Supplemental Data]

Articles from Journal of the American Society of Nephrology : JASN are provided here courtesy of American Society of Nephrology

RESOURCES