Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Oct 9.
Published in final edited form as: Pediatr Nephrol. 2010 Jan 5;25(6):1005–1016. doi: 10.1007/s00467-009-1392-6

Molecular anatomy of the kidney: what have we learned from gene expression and functional genomics?

Bree Rumballe 1, Kylie Georgas 1, Lorine Wilkinson 1, Melissa Little 1,
PMCID: PMC3189493  NIHMSID: NIHMS189765  PMID: 20049614

Abstract

The discipline of paediatric nephrology encompasses the congenital nephritic syndromes, renal dysplasias, neonatal renal tumours, early onset cystic disease, tubulopathies and vesicoureteric reflux, all of which arise due to defects in normal kidney development. Indeed, congenital anomalies of the kidney and urinary tract (CAKUT) represent 20–30% of prenatal anomalies, occurring in 1 in 500 births. Developmental biologists have studied the anatomical and morphogenetic processes involved in kidney development for the last five decades. However, with the advent of transgenic mice, the sequencing of the genome, improvements in mutation detection and the advent of functional genomics, our understanding of the molecular basis of kidney development has grown significantly. Here we discuss how the advent of new genetic and genomics approaches has added to our understanding of kidney development and paediatric renal disease, as well as identifying areas in which we are still lacking knowledge.

Keywords: Functional genomics, Gene expression, Kidney development, Metanephros, Mesonephros, Microarray, Pronephros

An introduction to the embryology of the urinary tract

The vertebrate kidney is derived from the intermediate mesoderm (IM) located lateral to the somites. The metanephros is the final of three embryonic excretory organs that develop in a temporally and spatially distinct order, these being the pronephros, mesonephros and metanephros. The pronephros is the earliest to arise and the most primitive. Functional during development only in lower vertebrates, it consists of simple tubules that empty into the primary nephric duct. The mesonephros develops more caudally along the nephric duct as the pronephros degenerates. Mesonephric tubules are more complex, consisting of a glomerulus and proximal tubule-like structures. The mesonephros is functional in adult higher fishes and amphibians, but only during embryogenesis in mammals. The metanephros is the final of the three embryonic organs; in mammals, this forms the permanent kidney. The metanephros is derived from two key structures: the ureteric bud (UB), which branches from the nephric duct, and a region of spatially defined intermediate mesoderm termed the metanephric mesenchyme (MM). The nephric duct (ND), also referred to as the Wolffian duct, empties into the cloaca. Paired UB arise from the ND at embryonic day (E) 10.5 in the mouse (at approximately 5 weeks gestation in the human). The epithelial UB grows towards the MM in response to inductive signals from the MM. Reciprocal interactions between the UB and MM result in growth and branching of the UB to ultimately form the collecting system of the kidney. A mesenchyme-to-epithelial transition (MET) of the MM at each of the newly formed UB tips results in the development of the nephrons, the filtration units of the kidney (Fig. 1).

Fig. 1.

Fig. 1

Overview of mammalian metanephric development. a The kidney is initiated via reciprocal induction events between the epithelial ureteric bud (UB), which arises from the mesonephric duct, and the metanephric mesenchyme (MM), which condenses to form a cap mesenchyme (CM). b Around the advancing ureteric tips (UT), the CM is induced to form a pre-tubular aggregate (PA) which then undergoes a mesenchyme-to-epithelial transition to form a renal vesicle (RV) surrounded by a basement membrane. The formation of each RV represents the initiation of a single nephron. c The RV proliferates and elongates to form first a comma-shaped body (CB) and then an S-shaped body (SB). Vascular endothelial cells are drawn into the cleft at the proximal end of this structure to form the glomerular capillaries. The other end of the forming nephron fuses with the UT at the late RV stage [73]. d A single nephron in the mammalian kidney. e Continued patterning of the elongating uriniferous tubule results in segmentation into the various functional regions of the nephron, including the proximal tubule and the distal tubule with an intervening loop of Henle (LH). The recruitment of vasculature and mesangial cells into the proximal end of this tubule forms the renal corpuscle (RC). PCT Proximal convoluted tubule, PST proximal straight tubule, S1–3 segment 1–3, DCT distal convoluted tubule, DST distal straight tubule, CT connecting tubule, CCD cortical collecting duct, MCD medullary collecting duct

This level of understanding regarding the origin of the metanephros is the result of careful anatomical observations and organ culture experiments. Pioneers in this field, Clifford Grobstein and Lauri Saxén, performed classical experiments in which the UB was dissected from the MM to demonstrate that a primary inductive event from the UB is essential for initiating MET in the MM. Without the UB, the MM fails to form epithelial tubules and rapidly degenerates [1]. However, a number of other tissue types, including embryonic brain and spinal cord, can induce MET and tubulogenesis when placed in contact with the MM, indicating the factors responsible are not specific to UB [2, 3]. MET induction requires cell-cell contact. Hence, even when the inducer is separated from the MM by microporous filters, induction only occurs if cells from the inducing tissue are able to extend processes through the pores into the MM [4].

Comprehensive profiling of gene expression during kidney development

Our molecular understanding of how these early morphogenetic events occur has grown slowly, based initially upon anatomical observations of renal defects in specific mutant mouse strains. However, the development over the last 20 years of a diverse array of functional genomics technologies has led to a far more comprehensive view of kidney development. This has included the development of high-throughput expression analysis techniques. Table 1 comprehensively lists all studies performed over the last decade in which genomics approaches have been applied to the analysis of kidney development. In the late 1990s, differential display PCR (ddPCR) was introduced as a means of defining mRNA expression differences in two different samples. This technique was applied to kidney morphogenesis by Leimeister et al. [5] and Plisov et al. [6]. However, ddPCR is difficult to automate. Consequently, techniques allowing a higher throughput analysis were established, such as comparative expressed sequence tag analysis (EST), serial analysis of gene expression (SAGE), high-throughput real time PCR (RT-PCR) and microarrays. cDNA or oligonucleotide microarrays have proven to be the most popular approach for global gene expression analysis as they are progressively more comprehensive [Affymetrix (Santa Clara, CA) ver. U133 Plus 2.0 Genechips now cover >47,000 human transcripts], sensitive, quantitative, rapid, and increasingly cost effective. Research on differential kidney gene expression in vivo initially provided a global and temporal overview of organ gene expression [7], identifying genes and hence pathways showing specific dynamic patterns of expression across kidney development. However, given the diversity of cell types within the developing kidney, such data were not particularly illuminating. To address spatially restricted gene expression, Valerius et al. [8] first analysed gene expression in immortalised cell lines representing either early metanephric mesenchyme (mK3), as determined by the expression of Hoxa11/d11, collagen 1 and vimentin, or MM undergoing epithelialisation (mK4), as determined by the coexpression of Pax2/8, Wnt4, collagen IV and LFB3 (mK4). While this study defined over 100 genes showing a greater than tenfold higher gene expression in mK4 versus mK3, these data were hard to relate back to the organ itself. Indeed, in 2002, in an editorial in Genome Biology, the relevance of microarray profiling to developmental biology was questioned [9]. However, spatial analysis of kidney development dramatically improved as research moved away from cell lines to direct isolation of specific kidney subcompartments, initially using manual microdissection and then via laser capture microdissection (LCM) and fluorescence-activated cell sorting (FACS)-based isolation from fluorescently tagged transgenic animals. Spatial experiments comparing UB and MM in vivo at various embryonic time points [10, 11] and in vitro [12] highlighted large numbers of differentially regulated genes in each murine cell population. This work was advanced with the microdissection of the ureteric tip from the ureteric bud at E11.5 [13] and E12.5 [14]. The microarray analysis of cells isolated from Sall1-GFP knock-in [15] and transgenic Hoxb7-GFP mice [11] went a step further in defining the compartments of MM-derived populations and the ureteric tree/tip, respectively. Challen et al. [16] also compared the expression of the E10.5 MM to that of the adjacent intermediate mesoderm and identified a number of early MM-specific markers, including CD24, which have subsequently formed the basis for studies looking for renal progenitors. Yano et al. [17] and Higgins et al. [18] profiled macroscopic compartments of the human kidney but concluded that greater spatial specificity was required to elucidate pathways involved in specific patterning events. Brunskill et al. [19] were the first to present a global atlas of gene expression of subcompartments of the developing kidney. They achieved this by combining the techniques of laser capture microdissection and FACS of compartment-specific transgenic mice to isolate 14 major subcompartments for microarray profiling. From this analysis, they were able to identify the top 223 genes showing the most compartment-enriched gene expression patterns. Validation by in situ hybridisation further defined the distribution of mRNA expression within a restricted component of the compartment, allowing the further subdissection of segmentation during the formation of the uriniferous tubule as it unwinds [19]. This was the first step in identifying anchor genes absolutely restricted to each anatomical structure and time point in the process of kidney development. This work was performed within the Genitourinary Development Molecular Atlas Project (GUDMAP), with the objective of identifying molecular markers that would enable the development of reporter mouse strains for further subdivision of kidney development [20]. Such mice will allow further subfractionation across time and space for subsequent expression analyses as well as provide tools for genetic fate mapping through the generation of GFP:Cre:ERT2 lines driven by kidney anchor gene promoters.

Table 1.

Gene expression profiling of human and murine kidney development

Article title Technologya Experiment Source/species Outcome Reference
Screen for genes regulated
  during early kidney
  morphogenesis
Transfilter organ
  culture and ddPCR
Spatial E11 cultured mesenchymal cells
  from wildtype outbred CD1
  mice
Compared uninduced state to that
  of explants cultured with spinal
  cord; 5/13 known genes and
  4/8 novel gene tags verified by
  Northern blot and ISH
[5]
Mesenchymal-epithelial
  transition in the
  developing metanephric
  kidney: gene expression
  study by differential
  display
ddPCR, ISH validation Spatial E13 cultured MM, F344 rat Induced metanephric explants vs.
  tubular induced explants; 72
  differentially regulated mRNAs
  detected in a MET event
[6]
Changes in global gene
  expression patterns
  during development and
  maturation of the rat
  kidney
Affymetrix Rat
  Genome U34A
  GeneChip Array;
  custom data analysis
  with “The Equaliser”
  & “eBlot”
Temporal Isolated RNA from E13, 15, 17,
  19, P0, P7 and NP adult rats
873/8740 genes were found to
  vary significantly (P = 0.0025)
  during rat kidney development.
  These genes clustered into five
  clear patterns of gene
  expression, each defined by a
  unique cluster-member gene
  function, tissue distribution,
  and embryonic expression
[7]
Microarray analysis of
  novel cell lines
  representing two stages of
  metanephric mesenchyme
  differentiation
Affymetrix Mu6500,
  Mu11K and U74A
  chip sets, ISH
  validation
Spatial Clonal kidney cell lines of
  representative stages of
  uninduced and induced MM
  isolated from Hoxa11/SV40 4–
  12 week old transgenic mice
Microarray experiments were
  performed on representative
  stages of MM (mK3 early =
  prior to induction) and late mK4
  (post-induction on way to MET);
  121 differential genes showed
  >10fold difference in
  expression; identification of
  genes not previously implicated
  in MM differentiation. (Hoxa7,
  Notch1, Fst, Igf2, sFRP1,
  Cdh16, shrom, Prx1, EST
  AA123934, NfiB)
[8]
A catalogue of gene
  expression in the
  developing kidney
Affymetrix U74Av2,
  ISH validation
Temporal
  & Spatial
Microdissected wildtype murine
  UB & MM (E11.5); total
  kidneys E11.5, E12.5 (pooled),
  E13.5, E16.5 & adult
  
Identification of 428 genes with
  significantly elevated
  expression levels at E12.5 vs.
  adult kidney; 78 genes
  differentially expressed
  between E11.5UB vs. MM; 53
  in UB, 25 in MM
[10]
Changes in gene
  expression patterns in the
  ureteric bud and
  metanephric mesenchyme
  in models of kidney
  development
High-density DNA
  arrays together with
  dIVT; Affymetrix Rat
  Genome U34A Gene
  chips
Temporal
  & Spatial
In vitro models of isolated E13
  Sprague-Dawley rat UB and
  MM cultures
in vitro models for
  morphogenesis of the rat UB
  and MM examined and data
  compared to whole embryonic
  kidney at different stages of
  development; 740/8471 genes
  were highly expressed in UB;
  624/8471 genes highly
  expressed in MM
[12]
Gene expression in the
  normal adult human
  kidney assessed by
  complementary DNA
  microarray
41,859 unique cDNA
  array (Stanford
  Functional Genomics
  Facility), IHC
  validation
Spatial Human nephrectomy specimens
  removed for renal neoplasia,
  lobes separated and dissected
  into 5 fractions
RNA isolated from inner and
  outer cortex, inner and outer
  medulla, papillary tips, renal
  pelvis and glomeruli. cDNA
  microarray & IHC validation;
  1548/41859 genes with
  variable expression; distinct
  expression of >250 genes in
  cortex, 139 in glomerulus, 130
  genes in renal tubules, 58
  genes in medullary vs.
  papillary tips, 102 genes in
  papillary tips/pelvis
[18]
Identification of kidney
  mesenchymal genes by a
  combination of
  microarray analysis and
  Sall1-GFP knock-in mice
Affymetrix Murine
  Genome U74v2
  GeneChips, ISH
  validation
Spatial Isolated murine Sall1-GFP-
  positive population of E17.5
  kidneys
Compared mesenchymal vs. UB
  derivatives; reported 78% of
  top 50 mRNAs were in the
  mesenchyme; Identification of
  17 novel mesenchymal genes
[15]
Temporal and spatial
  transcriptional programs
  in murine kidney
  development
Compugen long-
  oligonucleotide
  microarrays, ISH
  validation
Temporal
  & Spatial
Wildtype CD1 murine kidneys
  from 24-h intervals from E10.5
  to P0; E10.5 MM vs. IM;
  E15.5 Hoxb7-GFP positive
  ureteric tree vs. Hoxb7–GFP-
  negative MM
Temporal: Profiling through
  developmental time course &
  boxed comparison of
  development at E10.5 vs. 11.5
  vs. 13.5 kidneys; functional
  annotation clustered into 5
  classes and pathway/network
  analysis to identify critical
  signaling pathways; Spatial:
  MM E10.5 compared to
  adjacent IM; E15.5 UB vs.
  GFP negative mesenchymal
  derivatives from the Hoxb7
  mouse; 3600 dynamically
  expressed genes identified
[11]
Novel regulators of kidney
  development from the
  tips of the ureteric bud
Affymetrix Mouse
  Genome 430 2.0
  Array, RT-PCR and
  ISH validation
Spatial 13.5 rats, E12.5 mice isolated
  UB & MM
UB tips vs. tree have distinct
  gene expression profile,
  identified 20 predicted secreted
  protein transcripts upregulated
  in UB vs MM for both rat and
  mouse, e.g. Clf-1
[14]
Spatial gene expression in
  the T-stage mouse
  metanephros
Compugen mouse long
  microarrays, ISH
  validation
Spatial Dissected murine E11.5 ureteric
  tip, ureteric tree and MM
Identified genes spatially
  expressed in regions of the
  developing kidney e.g. Ureteric
  tip epithelium: sim-1, Arg2,
  Tacstd1, Crlf-1, BMP7; ureteric
  tree epithelium: Innp1, Itm2b,
  Mkrn1, SPARC, Emu2 and
  Gsta3; MM genes: CSPG2 and
  CV-2
[13]
Microarray interrogation of
  human metanephric
  mesenchymal cells
  highlights potentially
  important molecules in
  vivo
Affymetrix human
  U133 plus 2.0 Gene
  Chips, RT-PCR and
  IHC validation
Temporal 10-week gestation human fetus
  used to generate a human MM
  cell line
Generated mesenchymal cell
  lines from first trimester
  human metanephros and
  stimulated with growth factors;
  36 genes differentially
  regulated; changes in
  components of LIF signaling
  pathway reported
[58]
Atlas of gene expression in
  the developing kidney at
  microanatomic resolution
Affymetrix Mouse
  Genome 430 2.0
  Array; ISH validation
Spatial LCM-or FACS-sorted
  subpopulations from E12.5 and E15.5
  kidneys of CD1, Six2-GFP
  and Meis1-GFP mice.
Microarray analysis of 14
  subcompartments of the
  developing kidney; 223 genes
  identified as component-
  specific; identification of gene
  expression states of intermediate
  states of nephrogenesis;
  regulatory networks explored
[19]

ddPCR, Differential display PCR; MET, mesenchyme to epithelial transition; MM, metanephric mesenchyme; IM, intermediate mesoderm; UB, ureteric bud; dIVT, double in vitro transcription; ISH, in situ hybridisation; IHC, immunohistochemistry; LCM, laser capture microdissection; FACS, fluorescence-activated cell sorting; E, embryonic day; GFP, green fluorescent protein

a

Affymetrix, Santa Clara, CA; Compugen, Toronto, ONT, Canada

Expression profiling of mutant models

As well as being used for expression profiling of normal kidney development, such genomic tools have also been applied to a growing number of mouse mutant strains to increase our understanding of the pathways interrupted to cause specific phenotypes and, concommitently, to infer normal function for many genes. Table 2 lists such studies, describing the mutant model examined, the tissue isolated for analysis and the overall conclusions of the studies. In some cases, these mouse models were of known human conditions, such as the Gata3−/− mice, where defects in this gene are implicated in the autosomal dominant HDR syndrome (hyperparathyroidism, sensorineural deafness and renal anomaly) [21], and the Hnf-1 beta knockout mice, which mimic human MODY5 (maturity-onset diabetes mellitus and congenital malformations of the kidney) [22]. The latter highlighted a role for Hnf-1 beta in the suppression of SOCS3, thereby promoting tubulogenesis. As for normal development, the application of microdissection, FACS and LCM to mutant tissue has focused such analyses to the affected subcompartment. Cui et al. [23] profiled isolated glomeruli from Pod1 knockout mice, showing that the loss of this gene resulted in a reduction of alpha 8 integrin expression. LCM of individual renal vesicles was used to analyse changes in RV-gene expression in Lim1 mutant mice, showing significant down-regulation of Dkk1 and placing this gene downstream of Lim1 [24]. One of the most interesting developments of late has been the investigation of the role of specific micro-RNAs (miRNAs) in kidney development. These short non-coding RNAs bind to the 3′ untranslated region of mRNAs, thereby regulating levels of gene expression at the posttranscriptional level. An association had been noted between the expression of specific miRNAs and a number of renal disease states, including miR-17–92 in Wilms’ tumour and miR-15a in cystic kidney disease [25]. Three independent groups have now shown that a knockout of the miRNA processing enzyme, Dicer, in the podocytes results in dramatic cytoskeletal rearrangement and dedifferentiation [2628]. This resulted in proteinuria, podocyte effacement and apoptosis. Profiling of these mice compared to littermates enabled the in silico identification of miRNA target sequences in podocyte-specific genes.

Table 2.

Gene expression profiling of murine models of kidney dysgenesis

Mutant strain
(gene)
Source/species Technologya Outcome Reference
Megalin −/− 12- to 16-week-old mice kidneys Affymetrix MullK chip B Megalin regulates vitamin D homeostasis [59]
Meprin beta−/− Adult male mouse kidneys Affymetrix MG-U74Av2 oligonucleotide
array
Changes in genes associated with immune response and cell cycle,
increase in TRAF6, Tlr6
[60]
Aqp1−/− 10-week-old mouse inner
  medulla
NIA mouse 15 K clone set GRL array Downregulation of many medullary genes involved in heatshock/stress,
  mitochondrial pathways, vasopressin receptors
[61]
Gata3−/− E9.5 mouse embryos NIA mouse 15 K cDNA array Downregulated Dad2 [21]
Pod1−/− E18.5 mouse isolated glomeruli Affymetrix GeneChip Mouse Expression
  Array 430 2.0
Almost 4000 differentially expressed genes. Fourfold reduction in
  alpha 8 integrin
[23]
KininB1 and
kininB2−/−
12- to 16-week-old male
  mouse kidneys
Agilent oligonucleotide arrays Large number of downregulated genes that are already implicated in
  renal physiology or pathology
[62]
Pax 2lacZ/+
  Pax 2GFP
GFP Pro/mesonephric cells cDNA microarrays Gata3 in the nephric duct is regulated by Pax2 and Pax8 [63]
Sprague
  Dawley Cy−/−
3-week-old rat total kidney Agilent 60-mer Rat Array G4130A Upregulation of extracellular matrix metabolism genes in PKD model [64]
Hoxd 11−/−
  Hoxd 11−/−
E11.5, E12.5, E13.5 and
  E16.5 mouse kidneys
Affymetrix GeneChip Mouse Expression
  Array 430A&B
Upregulation of Irx3, downstream target of Hox; changes in, for example,
  Hoxa11s, GATA6, Slc40a1, Cxcl12; multiple downstream pathways
[65]
c-maf−/− Mouse kidneys Affymetrix GeneChip U74A array Down-regulation of GPx3 in knockout mice, possibly regulated by c-maf [66]
Pygo1−/−/
  Pygo2−/−
El8.5 mouse kidneys Illumina Sentrix Mouse-6 expression
  Beadchip
Expression changes in candidate Wnt signaling targets but Pygol/Pygo2
  not essential for canonical Wnt signaling.
[67]
Rarb2Crtg/+;
  Lim1lz/flox
LCM isolated renal vesicles Affymetrix GeneChip Mouse Expression
  Array 430 2.0
Differential expression of Chrdl2, Bmf, myob5, pdgfrl and 9×
  downregulation of Dkk1
[24]
Aqp 1−/− 8-week-old mouse renal medulla ±
  infusion of vasopressin
cDNA microarrays Increased cyclin Dl and Egr accompanied by increased cell proliferation [68]
Hnf1 beta−/− mIMCD3 cells CHIP-chip & Affymetrix GeneChip Mouse
  Expression Array 430A&B
Hnf-1 beta binds to and regulates Socs3 to control tubulogenesis [22]
Nephrin−/− E18.5 glomeruli mouse Affymetrix GeneChip Mouse Expression
  Array 430 2.0
Nephrin not essential for cell survival but plays a role in glomerular
  differentiation and development.
[69]
Glis2mut PND25 & PND60 total kidney Agilent whole-mouse genome microarrays Upregulation of genes involved in ECM homeostasis and
  immune/inflammatory response
[70]
Wnt4 n/+ E14.5 mouse kidney Affymetrix GeneChip MG-U74Av2 arrays >50% genes expressed in renal vesicle or later stage [71]
Aqp 11−/− and
  Pkd1−/−
PND1 mouse kidneys Affymetrix GeneChip Mouse Expression
  Array 430 2.0
Comparison suggests strong similarities between the two strains
  (same changes in Myc, Egfr, Egf, Mmp12, Timp1, Tgfb1)
[72]
NPHS2-cre;
  DicerFlx/Flx
3-week-old mouse isolated
  glomeruli
Exiqon's miRCURY LNA array mmu-miR-23b, mmu-miR-24, and mmu-miR-26a identified in glomeruli [27]
NPHS2-cre;
  DicerFlx/Flx
4- to 5-week old mouse
  glomeruli
Affymetrix GeneChip Mouse Expression
  Array 430 2.0
Upregulation of podocyte specific members of mir-30 miRNA family [28]

PND, Post-natal day; GFP, green fluorescent protein; PKD model, polycystic kidney disease model; ECM, extracellular matrix

a

Exiqon, Vedbaek, Denmark; Agilent Technologies, Santa Clara, CA; Illumina, San Diego, CA

Development databases and in silico modeling

The databasing of information on the molecular basis of kidney development began in the mid-1990s with the creation of the Kidney Development Database (KDD) [29, 30]. This began as a spreadsheet of gene expression in different cellular compartments or structure in the developing kidney as well as a list of mutant mice displaying defects in renal development. The database still exists (http://golgi.ana.ed.ac.uk/kidhome.html), but for the most part it is no longer comprehensive as the flow of data from expression profiling has become too large. The National Institutes for Digestion, Diabetes and Kidney Disease have more recently supported the creation of a large public database of gene expression in the developing urogenital tract as a whole which supercedes the KDD. This Genitourinary Development Molecular Anatomy Project database (http://www.GUDMAP.org) [20] houses >255 microarrays and >7200 wholemount and section in situ hybridisation entries covering >2900 genes analysed for expression in the developing genitourinary tract (UGT) from TS17 (E10.5) to adult tissue (as of August 2009). Access to such a rich source of expression data that have been annotated to the level of detail as this will prove to be a valuable resource in the prioritisation or even identification of genes responsible for specific phenotypes. An example of this was reported recently by Berkovic et al. [31], who used single-nucleotide polymorphism chips to localise a region responsible for the action of myoclonus-renal failure syndrome (AMRF) to chromosome 4q13–21 and then used microarray-expression analysis to prioritise candidate genes. As a result, SCARB2/Limp2, a lysosomal-membrane protein, was shown to contain mutations in all three affected pedigrees. Re-analysis of a Limp2 knockout mouse verified the presence of cerebral and renal changes. There is also an extensive gene expression database documenting the expression of genes in the developing pronephros of Xenopus laevis established by the Euregene consortium (XGEbase http://www.euregene.org/xgebase).

Flies, fish, frogs and worms

While the mouse has been a key tool in our understanding of kidney development, the most powerful model organisms for functional genomics are the fly Drosophila melanogaster, the zebrafish Danio rerio, the African clawed toad Xenopus laevis and the worm Caenorhabditis elegans. Whilst only zebrafish and Xenopus are vertebrates, all of these model systems have made valuable contributions to our understanding of kidney development and disease.

Arguably, the animal that has had the most influence on the kidney field is C. elegans, where the analysis of the role of large numbers of genes involved in cilia function have significantly advanced the field of cystic kidney disease. Identification of the C. elegans genes lov-1 and pkd-2 revealed the association between cilia and autosomal dominant polycystic kidney disease (ADPKD). Other mammalian renal diseases, such as autosomal recessive PKD (ARPKD), nephronophthisis (NPHP) and Bardet–Biedl syndrome (BBS), which also have orthologues in C. elegans, share the common elements of cystic kidneys and ciliary localised gene products [32]. This link came with the realisation that the human PKD1 and PKD2 proteins, mutations in which cause PKD, localise to the primary cilia of renal epithelial cells as do the worm homologues. Cilia in the worm are critical for appropriate male mating behaviour, and a large number of genes identified as being involved in cilia morphogenesis or function have subsequently been shown to result in renal cysts when mutated in mice [33].

Teleosts, such as zebrafish, do not form a metanephros. Instead, they have a functional pronephros during embryonic and larval stages and a mesonephros acting as their excretory organ in the adults. Nevertheless, there is considerable similarity in gene expression and cell type between the paired pronephric tubules and ducts and the single midline glomerulus of the developing zebrafish pronephros and the nephron of the mammalian metanephros (Fig. 2b). Zebrafish mutant embryo models and gene knockdown using morpholino antisense oligonucleotides have proven useful for studying defects of glomerular formation and slit diaphragm morphology [34] and for identifying several genes with roles in cyst formation and tubular defects [35]. Indeed, such studies have identified more than a dozen gene disruptions, leading to cystic phenotypes affecting either the glomerulus or the pronephric tubules [36]. Some of the defects have been due to mutations in intraflagellar transport genes associated with ciliogenesis, again highlighting the link between cilia defects and PKD.

Fig. 2.

Fig. 2

Comparison of the segmental patterning of the developing excretory systems in the fly, zebrafish and Xenopus. a In the fly, there are two pairs of excretory tubules, the Malpighian tubules, which empty into the hindgut. These consist of blind-ended tubules containing nephrocytes. Hir Hindgut imaginal ring. b The zebrafish has a pair of nephrons that fuse to form a single central glomerulus. The pronephric tubules lead to pronephric ducts which empty into the cloaca. G Glomerulus, coS corpuscle of Stannius. c Xenopus contain a bilateral pair of nephrons where blood filtration occurs through a single glomus into the coelom, and filtrate enters the segmented nephron tubules via nephrostomes. Connecting tubules empty into the cloaca. G Glomus, c coelom, Ns nephrostome, PT1 proximal tubule 1, PT2 proximal tubule 2, PT3 proximal tubule 3, IT1 intermediate tubule 1, IT2 intermediate tubule 2, DT1 distal tubule 1, DT2 distal tubule 2, CT connecting tubule. Figure adapted from [74, 75] and [37] (used with permission)

In developing Xenopus, there is one pronephric glomus, a pronephric tubule, and a connecting tubule leading to the cloaca where the nephron joins its bilateral counterpart (Fig. 2c). Large-scale gene expression studies of the pronephros in Xenopus have also been performed, and after an analysis of the expression of over 240 genes, eight distinct cellular regions have been identified within the pronephric tubule and pronephric duct [37]. The pronephric nephrons can be segmented into proximal tubule, intermediate tubule, distal tubule and connecting tubule. These data are available via XGEbase (http://www.euregene.org/xgebase), which houses over 1200 in situ hybridisation (ISH) images. A large number of the expression patterns delineating different regions are represented by distinct solute transporters, and there is a direct correlation between the expression of these genes in Xenopus and that in mouse/human, suggesting that the more tractable Xenopus model system is pertinent to the understanding of development in humans. Nephron segmentation has also been investigated in Xenopus via microinjection of mRNA into Xenopus blastocysts [38]. This has allowed the investigation of the role of key transcription factors, including Lhx1, Wt1, Pax8 and Hnf1b, in nephron patterning and tubulogenesis. Gain- and loss-of-function experiments have also shown that the Irx genes are required early for maintenance of the kidney anlage, and later for segmentation of the nephron [39, 40]. Irx3 was found to be both necessary and sufficient for the formation of the Xenopus intermediate segments that are equivalent to the loop of Henle segments of the mammalian nephron [39]. Patterns of Irx1, Irx2 and Irx3 expression in both the embryonic and adult mouse nephron suggest that these genes also have roles in loop of Henle patterning in mammals [39].

Finally, even the fly has played a role in our understanding of kidney development. The secretory and absorptive Malpighian tubules in Drosophila melanogaster have provided insights into tubular morphogenesis [41]. The filtration barrier of the fly is called the nephrocyte. Like podocytes, these cells are mesodermal in origin and contain specialised slit diaphragms analogous to those in the mammalian renal podocyte [42] (Fig. 2a). Of note, the fly orthologues of Nphs1 (sticks and stones; Sns), Kirrel (kin of irre; kirre), Cd2ap (CG31012), Tjp1 (polychaetoid; pyd) and Nphs2 (Mec2) are genes critically important for podocyte structure and function in humans and mice, and all are expressed in nephrocytes in analogous intracellular locations. These cells form a size-selective barrier and remove products from the haemolymph by filtration and endocytosis [42]. However, the similarities in gene expression and the ready ability to perform mutation screens in flies may prove a rich source of understanding of renal development and disease. In addition, Liu et al. [43] have created a large-scale network diagram based on known gene function in the fly and identified roles for the JNK pathway in renal cancer based on these network predictions. This same pathway appears to be critical for the self-renewal of a population of cells within the Malpighian tubules in the fly [44]. This observation of self-renewing cells in the adult Malpighian tubules may translate to a greater understanding of the tubular regenerative capacity of the human kidney.

Lineage and selective gene deletion

While such comparative models have shown parallels to the mammalian situation, the primary animal model of kidney development has remained the mouse, and advancements in gene targeting technologies to allow for cell-specific gene deletion has revolutionised this field. Previous reviews have summarized available Cre transgenic lines targeting specific cell types/regions of the developing kidney [45, 46]. To these lists can be added three distinct alleles of Six2-cre transgenic mice [47], which allow gene inactivation in the cap mesenchyme, and the recently reported Wnt4EGFPCre, which inactivates genes from the renal vesicle stage of nephron development [48]. Another line that has been widely used to examine the role of genes in the metanephric mesenchyme of the kidney is the Pax3-Cre transgenic line originally described by Li et al. [49] and shown to result in recombination in the metanephric mesenchyme by Chang et al. [50].

These tools have contributed significantly to our understanding of kidney development. The definition of podocyte-specific expression of Nphs1, for example, has facilitated the inactivation of a large number of genes in this cell type alone. This involved the expression of the enzyme Cre-recombinase under the control of the Nphs1 promoter crossed to a strain in which sequences for the gene of interest are flanked by loxP sites recognised by Cre-recombinase. Such cell-specific gene deletion versus cell-specific promoter-driven overexpression has facilitated the development of an allelic series in which the level of vascular endothelial growth factor-A (VEGF-A) in the developing glomerulus was varied from too high to too low [51]. This revealed a requirement for exquisite control of VEGF-A expression from the podocytes in order to establish and maintain normal glomerular function. The same technology has been adapted to investigate the lineage fate of cells during development. More recently, the same podocin-Cre (Nphs2-Cre) mouse, which also expresses Cre recombinase in the podocytes, has revealed a critical role for microRNAs in the maintenance of glomerular function when crossed with a Dicer (flx/flx) strain, Dicer being the enzyme required for miRNA processing [27]. By incorporating the requirement for an inducible agent, such as tamoxifen or doxycycline, to initiate Cre-recombinase expression, specific cells can be permanently marked to visualise what they give rise to. In this way, it has been possible to prove that all segments of the nephron, other than the collecting duct, do indeed arise from the metanephric mesenchyme [47], but that only a subset of the MM, the Six2-expressing cap mesenchyme, condenses around the ureteric tips. Lineage tracing has also been applied to the intermediate mesodermal Osr1-expressing population to show that these cells can give rise to all parts of the kidney, other than the resident macrophages, but only at specific time points [52]. This has not eliminated the possibility that at least some of the Foxd1-positive interstitial cells do migrate into the kidney from another location. As well as defining fate during development, lineage tracing approaches are now being used to establish what cells give rise to what in response to damage. Humphreys et al. [53], using Six2 lineage tracing, showed that tubular repair after induced damage does not involve any cell population that was not derived from the Six2-positive cap mesenchyme. None of this would have been possible without the identification of cell types of compartment-specific gene expression patterns, many of which have and will come from gene expression analyses.

What questions remain?

As a result of these diverse functional genomics approaches, we now have a much more global and comprehensive view of which genes, gene networks and pathways are expressed across time and space in kidney development, and we have identified many more critical players in podocyte maturation and function, nephron patterning and segmentation, ureteric bud outgrowth and branching as well as in the commitment of the MM and the various fates of these cells. Studies in worms, frogs, fish and mice have revealed a key role for cilia in the onset and progression of cystic disease and the advent of lineage tracing, and Cre-mediated selective gene ablation has particularly impacted on our understanding of glomerular disease. However, there are many key questions in kidney molecular development that remain unanswered. These include the regulation of cessation of nephrogenesis, the process governing the perinatal remodeling of the medulla and a clearer understanding of the normal repair response of the kidney to different types of damage. In mammals, nephron endowment is complete before or around birth. Hartman et al. [54] examined this process in mouse and proposed that it results from the exhaustion of the nephron progenitor cells, the cap mesenchyme. In humans, cessation of nephrogenesis occurs by 36 weeks of life. Consequently, premature babies or those suffering from intrauterine growth retardation (IUGR) are known to have a lower number of nephrons. Unlike mammals, teleosts and cartilaginous fish, including the skate, retain a progenitor blastema such that if the kidney is damaged, this blastema undergoes nephron induction to replace lost nephrons. While it is accepted that the remaining cap mesenchyme is exhausted by a burst of nephron differentiation at the end of mammalian development, what triggers this loss of self renewal and stimulates terminal differentiation is not understood. Intervention to prolong nephron endowment may play an important role in the treatment of IUGR or premature babies.

At birth, the medulla of the mammalian kidney still contains a large amount of interstitial tissue between the descending loops of Henle. Hence, an efficient counter-current mechanism is not yet in place, and considerable remodeling is required in the perinatal period for the newborn to be able to concentrate urine. Nothing is known about the regulation of this maturation, although a small number of genes showing medullary defects evident in newborn mice have been identified, including a mutation of Adamts-1 [55]. Of interest, a number of genes expressed during nephron formation in the cortex are also expressed in the developing and perinatal papilla, but their roles there are unknown. This includes Wt1 and Wnt4, which are important for mesenchymal-to-epithelial transitions in nephron formation. Given the proposal of a papillary-located postnatal progenitor population in the adult kidney [56], it is possible that such cells are also involved in the normal remodeling of the papilla during medullary maturation and are used again after obstructive damage in the kidney.

Many of these key developmental questions are likely to involve epigenetic as well as genetic regulation. There is a growing appreciation of the role of non-coding RNAs, chromatin modification and DNA methylation and acetylation. A link between a key transcription factor in kidney development, Pax2, and the site-specific methylation of chromatin by histone methyltransferases has already been described [57]. Here, Pax2 can modify chromatin at different gene loci depending upon the proteins it is interacting with. Investigating these processes will represent the next phase of functional genomics and will be substantially enhanced by the advent of Next Generation (NextGen) sequencing. While this technology will complement current expression profiling by more comprehensively cataloging the gene expression of all alternate isoforms of coding as well as all non-coding transcripts, it will allow the assessment of DNA methylation and acetylation to globally determine why a given set of genes are active. The advent of genome wide lentiviral gene and shRNA libraries will also enable large-scale analysis of the overexpression or silencing of individual or combinations of genes either in cell lines or in embryo morphology screens.

Conclusion

In conclusion, based on results from two decades of functional genetics and genomics, our understanding of the molecular basis of kidney development and disease has expanded considerably. We now have a detailed and comprehensive atlas of the developing kidney which has subdivided territories into smaller domains based upon gene expression and activity. Continued complimentary use of functional genomics in all model organisms will continue to reveal the molecular markers and regulatory pathways that underpin the development of the mammalian kidney. While we have not yet reached the level of the complete temporo-spatial wiring diagram, there is no doubt that the continued application of these technologies, and new additions to the functional genomics arsenal as they come on line, will bring us to the point where we can link cause and effect in each of the morphogenetic processes required to form a normal functional kidney. By extension, our understanding of what can and does go wrong will also improve. The final link is turning that understanding into diagnosis and treatment.

References

  • 1.Grobstein C. Trans-filter induction of tubules in mouse metanephrogenic mesenchyme. Exp Cell Res. 1956;10:424–440. doi: 10.1016/0014-4827(56)90016-7. [DOI] [PubMed] [Google Scholar]
  • 2.Grobstein C. Morphogenetic interaction between embryonic mouse tissues separated by a membrane filter. Nature. 1953;172:869–870. doi: 10.1038/172869a0. [DOI] [PubMed] [Google Scholar]
  • 3.Unsworth B, Grobstein C. Induction of kidney tubules in mouse metanephrogenic mesenchyme by various embryonic mesenchymal tissues. Dev Biol. 1970;21:547–556. doi: 10.1016/0012-1606(70)90077-1. [DOI] [PubMed] [Google Scholar]
  • 4.Sariola H, Ekblom P, Henke-Fahle S. Embryonic neurons as in vitro inducers of differentiation of nephrogenic mesenchyme. Dev Biol. 1989;132:271–281. doi: 10.1016/0012-1606(89)90224-8. [DOI] [PubMed] [Google Scholar]
  • 5.Leimeister C, Bach A, Woolf AS, Gessler M. Screen for genes regulated during early kidney morphogenesis. Dev Genet. 1999;24:273–283. doi: 10.1002/(SICI)1520-6408(1999)24:3/4<273::AID-DVG10>3.0.CO;2-3. [DOI] [PubMed] [Google Scholar]
  • 6.Plisov SY, Ivanov SV, Yoshino K, Dove LF, Plisova TM, Higinbotham KG, Karavanova I, Lerman M, Perantoni AO. Mesenchymal-epithelial transition in the developing metanephric kidney: gene expression study by differential display. Genesis. 2000;27:22–31. doi: 10.1002/1526-968x(200005)27:1<22::aid-gene40>3.0.co;2-v. [DOI] [PubMed] [Google Scholar]
  • 7.Stuart RO, Bush KT, Nigam SK. Changes in global gene expression patterns during development and maturation of the rat kidney. Proc Natl Acad Sci USA. 2001;98:5649–5654. doi: 10.1073/pnas.091110798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Valerius MT, Patterson LT, Witte DP, Potter SS. Microarray analysis of novel cell lines representing two stages of metanephric mesenchyme differentiation. Mech Dev. 2002;112:219–232. doi: 10.1016/s0925-4773(02)00008-4. [DOI] [PubMed] [Google Scholar]
  • 9.Livesey R. Have microarrays failed to deliver for developmental biology? Genome Biol. 2002;3 doi: 10.1186/gb-2002-3-9-comment2009. comment 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Schwab K, Patterson LT, Aronow BJ, Luckas R, Liang HC, Potter SS. A catalogue of gene expression in the developing kidney. Kidney Int. 2003;64:1588–1604. doi: 10.1046/j.1523-1755.2003.00276.x. [DOI] [PubMed] [Google Scholar]
  • 11.Challen G, Gardiner B, Caruana G, Kostoulias X, Martinez G, Crowe M, Taylor DF, Bertram J, Little M, Grimmond SM. Temporal and spatial transcriptional programs in murine kidney development. Physiol Genomics. 2005;23:159–171. doi: 10.1152/physiolgenomics.00043.2005. [DOI] [PubMed] [Google Scholar]
  • 12.Stuart RO, Bush KT, Nigam SK. Changes in gene expression patterns in the ureteric bud and metanephric mesenchyme in models of kidney development. Kidney Int. 2003;64:1997–2008. doi: 10.1046/j.1523-1755.2003.00383.x. [DOI] [PubMed] [Google Scholar]
  • 13.Caruana G, Cullen-McEwen L, Nelson AL, Kostoulias X, Woods K, Gardiner B, Davis MJ, Taylor DF, Teasdale RD, Grimmond SM, Little MH, Bertram JF. Spatial gene expression in the T-stage mouse metanephros. Gene Expr Patterns. 2006;6:807–825. doi: 10.1016/j.modgep.2006.02.001. [DOI] [PubMed] [Google Scholar]
  • 14.Schmidt-Ott KM, Yang J, Chen X, Wang H, Paragas N, Mori K, Li JY, Lu B, Costantini F, Schiffer M, Bottinger E, Barasch J. Novel regulators of kidney development from the tips of the ureteric bud. J Am Soc Nephrol. 2005;16:1993–2002. doi: 10.1681/ASN.2004121127. [DOI] [PubMed] [Google Scholar]
  • 15.Takasato M, Osafune K, Matsumoto Y, Kataoka Y, Yoshida N, Meguro H, Aburatani H, Asashima M, Nishinakamura R. Identification of kidney mesenchymal genes by a combination of microarray analysis and Sall1-GFP knockin mice. Mech Dev. 2004;121:547–557. doi: 10.1016/j.mod.2004.04.007. [DOI] [PubMed] [Google Scholar]
  • 16.Challen GA, Martinez G, Davis MJ, Taylor DF, Crowe M, Teasdale RD, Grimmond SM, Little MH. Identifying the molecular phenotype of renal progenitor cells. J Am Soc Nephrol. 2004;15:2344–2357. doi: 10.1097/01.ASN.0000136779.17837.8F. [DOI] [PubMed] [Google Scholar]
  • 17.Yano N, Endoh M, Fadden K, Yamashita H, Kane A, Sakai H, Rifai A. Comprehensive gene expression profile of the adult human renal cortex: analysis by cDNA array hybridization. Kidney Int. 2000;57:1452–1459. doi: 10.1046/j.1523-1755.2000.00990.x. [DOI] [PubMed] [Google Scholar]
  • 18.Higgins JP, Wang L, Kambham N, Montgomery K, Mason V, Vogelmann SU, Lemley KV, Brown PO, Brooks JD, van de Rijn M. Gene expression in the normal adult human kidney assessed by complementary DNA microarray. Mol Biol Cell. 2004;15:649–656. doi: 10.1091/mbc.E03-06-0432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Brunskill EW, Aronow BJ, Georgas K, Rumballe B, Valerius MT, Aronow J, Kaimal V, Jegga AG, Grimmond S, McMahon AP, Patterson LT, Little MH, Potter SS. Atlas of gene expression in the developing kidney at microanatomic resolution. Dev Cell. 2008;15:781–791. doi: 10.1016/j.devcel.2008.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.McMahon AP, Aronow BJ, Davidson DR, Davies JA, Gaido KW, Grimmond S, Lessard JL, Little MH, Potter SS, Wilder EL, Zhang P. GUDMAP: the genitourinary developmental molecular anatomy project. J Am Soc Nephrol. 2008;19:667–671. doi: 10.1681/ASN.2007101078. [DOI] [PubMed] [Google Scholar]
  • 21.Airik R, Karner M, Karis A, Karner J. Gene expression analysis of Gata3−/− mice by using cDNA microarray technology. Life Sci. 2005;76:2559–2568. doi: 10.1016/j.lfs.2004.10.054. [DOI] [PubMed] [Google Scholar]
  • 22.Ma Z, Gong Y, Patel V, Karner CM, Fischer E, Hiesberger T, Carroll TJ, Pontoglio M, Igarashi P. Mutations of HNF-1beta inhibit epithelial morphogenesis through dysregulation of SOCS-3. Proc Natl Acad Sci USA. 2007;104:20386–20391. doi: 10.1073/pnas.0705957104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Cui S, Li C, Ema M, Weinstein J, Quaggin SE. Rapid isolation of glomeruli coupled with gene expression profiling identifies downstream targets in Pod1 knockout mice. J Am Soc Nephrol. 2005;16:3247–3255. doi: 10.1681/ASN.2005030278. [DOI] [PubMed] [Google Scholar]
  • 24.Potter SS, Hartman HA, Kwan KM, Behringer RR, Patterson LT. Laser capture-microarray analysis of Lim1 mutant kidney development. Genesis. 2007;45 doi: 10.1002/dvg.20309. 432–329. [DOI] [PubMed] [Google Scholar]
  • 25.Saal S, Harvey SJ. MicroRNAs and the kidney: coming of age. Curr Opin Nephrol Hypertens. 2009;18:317–323. doi: 10.1097/MNH.0b013e32832c9da2. [DOI] [PubMed] [Google Scholar]
  • 26.Harvey SJ, Jarad G, Cunningham J, Goldberg S, Schermer B, Harfe BD, McManus MT, Benzing T, Miner JH. Podocyte-specific deletion of dicer alters cytoskeletal dynamics and causes glomerular disease. J Am Soc Nephrol. 2008;19:2150–2158. doi: 10.1681/ASN.2008020233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Ho J, Ng KH, Rosen S, Dostal A, Gregory RI, Kreidberg JA. Podocyte-specific loss of functional microRNAs leads to rapid glomerular and tubular injury. J Am Soc Nephrol. 2008;19:2069–2075. doi: 10.1681/ASN.2008020162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Shi S, Yu L, Chiu C, Sun Y, Chen J, Khitrov G, Merkenschlager M, Holzman LB, Zhang W, Mundel P, Bottinger EP. Podocyte-selective deletion of dicer induces proteinuria and glomerulosclerosis. J Am Soc Nephrol. 2008;19:2159–2169. doi: 10.1681/ASN.2008030312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Davies JA, Brandli AW. A computer database for kidney development. Trends Genet. 1996;12:322. [Google Scholar]
  • 30.Davies JA. The kidney development database. Dev Genet. 1999;24:194–198. doi: 10.1002/(SICI)1520-6408(1999)24:3/4<194::AID-DVG2>3.0.CO;2-V. [DOI] [PubMed] [Google Scholar]
  • 31.Berkovic SF, Dibbens LM, Oshlack A, Silver JD, Katerelos M, Vears DF, Lullmann-Rauch R, Blanz J, Zhang KW, Stankovich J, Kalnins RM, Dowling JP, Andermann E, Andermann F, Faldini E, D'Hooge R, Vadlamudi L, Macdonell RA, Hodgson BL, Bayly MA, Savige J, Mulley JC, Smyth GK, Power DA, Saftig P, Bahlo M. Array-based gene discovery with three unrelated subjects shows SCARB2/LIMP-2 deficiency causes myoclonus epilepsy and glomerulosclerosis. Am J Hum Genet. 2008;82:673–684. doi: 10.1016/j.ajhg.2007.12.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Barr MM. Caenorhabditis elegans as a model to study renal development and disease: sexy cilia. J Am Soc Nephrol. 2005;16:305–312. doi: 10.1681/ASN.2004080645. [DOI] [PubMed] [Google Scholar]
  • 33.Lipton J. Mating worms and the cystic kidney: Caenorhabditis elegans as a model for renal disease. Pediatr Nephrol. 2005;20:1531–1536. doi: 10.1007/s00467-005-1958-x. [DOI] [PubMed] [Google Scholar]
  • 34.Majumdar A, Drummond IA. The zebrafish floating head mutant demonstrates podocytes play an important role in directing glomerular differentiation. Dev Biol. 2000;222:147–157. doi: 10.1006/dbio.2000.9642. [DOI] [PubMed] [Google Scholar]
  • 35.Sun Z, Amsterdam A, Pazour GJ, Cole DG, Miller MS, Hopkins N. A genetic screen in zebrafish identifies cilia genes as a principal cause of cystic kidney. Development. 2004;131:4085–4093. doi: 10.1242/dev.01240. [DOI] [PubMed] [Google Scholar]
  • 36.Drummond IA, Majumdar A, Hentschel H, Elger M, Solnica-Krezel L, Schier AF, Neuhauss SC, Stemple DL, Zwartkruis F, Rangini Z, Driever W, Fishman MC. Early development of the zebrafish pronephros and analysis of mutations affecting pronephric function. Development. 1998;125:4655–4667. doi: 10.1242/dev.125.23.4655. [DOI] [PubMed] [Google Scholar]
  • 37.Raciti D, Reggiani L, Geffers L, Jiang Q, Bacchion F, Subrizi AE, Clements D, Tindal C, Davidson DR, Kaissling B, Brandli AW. Organization of the pronephric kidney revealed by large-scale gene expression mapping. Genome Biol. 2008;9:R84. doi: 10.1186/gb-2008-9-5-r84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Jones EA. Xenopus: a prince among models for pronephric kidney development. J Am Soc Nephrol. 2005;16:313–321. doi: 10.1681/ASN.2004070617. [DOI] [PubMed] [Google Scholar]
  • 39.Reggiani L, Raciti D, Airik R, Kispert A, Brandli AW. The prepattern transcription factor Irx3 directs nephron segment identity. Genes Dev. 2007;21:2358–23570. doi: 10.1101/gad.450707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Alarcon P, Rodriguez-Seguel E, Fernandez-Gonzalez A, Rubio R, Gomez-Skarmeta JL. A dual requirement for Iroquois genes during Xenopus kidney development. Development. 2008;135:3197–3207. doi: 10.1242/dev.023697. [DOI] [PubMed] [Google Scholar]
  • 41.Jung AC, Denholm B, Skaer H, Affolter M. Renal tubule development in Drosophila: a closer look at the cellular level. J Am Soc Nephrol. 2005;16:322–328. doi: 10.1681/ASN.2004090729. [DOI] [PubMed] [Google Scholar]
  • 42.Weavers H, Prieto-Sanchez S, Grawe F, Garcia-Lopez A, Artero R, Wilsch-Brauninger M, Ruiz-Gomez M, Skaer H, Denholm B. The insect nephrocyte is a podocyte-like cell with a filtration slit diaphragm. Nature. 2009;457:322–326. doi: 10.1038/nature07526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Liu J, Ghanim M, Xue L, Brown CD, Iossifov I, Angeletti C, Hua S, Negre N, Ludwig M, Stricker T, Al-Ahmadie HA, Tretiakova M, Camp RL, Perera-Alberto M, Rimm DL, Xu T, Rzhetsky A, White KP. Analysis of Drosophila segmentation network identifies a JNK pathway factor overexpressed in kidney cancer. Science. 2009;323:1218–1222. doi: 10.1126/science.1157669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Singh SR, Hou SX. Multipotent stem cells in the Malpighian tubules of adult Drosophila melanogaster. J Exp Biol. 2009;212:413–423. doi: 10.1242/jeb.024216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Igarashi P. Kidney-specific gene targeting. J Am Soc Nephrol. 2004;15:2237–2239. doi: 10.1097/01.ASN.0000136298.09488.D7. [DOI] [PubMed] [Google Scholar]
  • 46.Rubera I, Hummler E, Beermann F. Transgenic mice and their impact on kidney research. Pflugers Arch. 2009;458:211–222. doi: 10.1007/s00424-008-0624-0. [DOI] [PubMed] [Google Scholar]
  • 47.Kobayashi A, Valerius MT, Mugford JW, Carroll TJ, Self M, Oliver G, McMahon AP. Six2 defines and regulates a multipotent self-renewing nephron progenitor population throughout mammalian kidney development. Cell Stem Cell. 2008;3:169–181. doi: 10.1016/j.stem.2008.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Shan J, Jokela T, Skovorodkin I, Vainio S. Mapping of the fate of cell lineages generated from cells that express the Wnt4 gene by time-lapse during kidney development. Differentiation. 2010;79:57–64. doi: 10.1016/j.diff.2009.08.006. [DOI] [PubMed] [Google Scholar]
  • 49.Li J, Chen F, Epstein JA. Neural crest expression of Cre recombinase directed by the proximal Pax3 promoter in transgenic mice. Genesis. 2000;26:162–164. doi: 10.1002/(sici)1526-968x(200002)26:2<162::aid-gene21>3.0.co;2-r. [DOI] [PubMed] [Google Scholar]
  • 50.Chang CP, McDill BW, Neilson JR, Joist HE, Epstein JA, Crabtree GR, Chen F. Calcineurin is required in urinary tract mesenchyme for the development of the pyeloureteral peristaltic machinery. J Clin Invest. 2004;113:1051–1058. doi: 10.1172/JCI20049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Eremina V, Baelde HJ, Quaggin SE. Role of the VEGF–a signaling pathway in the glomerulus: evidence for crosstalk between components of the glomerular filtration barrier. Nephron Physiol. 2007;106:32–37. doi: 10.1159/000101798. [DOI] [PubMed] [Google Scholar]
  • 52.Mugford JW, Sipila P, McMahon JA, McMahon AP. Osr1 expression demarcates a multi-potent population of intermediate mesoderm that undergoes progressive restriction to an Osr1-dependent nephron progenitor compartment within the mammalian kidney. Dev Biol. 2008;324:88–98. doi: 10.1016/j.ydbio.2008.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Humphreys BD, Valerius MT, Kobayashi A, Mugford JW, Soeung S, Duffield JS, McMahon AP, Bonventre JV. Intrinsic epithelial cells repair the kidney after injury. Cell Stem Cell. 2008;2:284–291. doi: 10.1016/j.stem.2008.01.014. [DOI] [PubMed] [Google Scholar]
  • 54.Hartman HA, Lai HL, Patterson LT. Cessation of renal morphogenesis in mice. Dev Biol. 2007;310:379–387. doi: 10.1016/j.ydbio.2007.08.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Mittaz L, Ricardo S, Martinez G, Kola I, Kelly DJ, Little MH, Hertzog PJ, Pritchard MA. Neonatal calyceal dilation and renal fibrosis resulting from loss of Adamts-1 in mouse kidney is due to a developmental dysgenesis. Nephrol Dial Transplant. 2005;20:419–423. doi: 10.1093/ndt/gfh603. [DOI] [PubMed] [Google Scholar]
  • 56.Oliver JA, Maarouf O, Cheema FH, Martens TP, Al-Awqati Q. The renal papilla is a niche for adult kidney stem cells. J Clin Invest. 2004;114:795–804. doi: 10.1172/JCI20921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Patel SR, Kim D, Levitan I, Dressler GR. The BRCT-domain containing protein PTIP links PAX2 to a histone H3, lysine 4 methyltransferase complex. Dev Cell. 2007;13:580–592. doi: 10.1016/j.devcel.2007.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Price KL, Long DA, Jina N, Liapis H, Hubank M, Woolf AS, Winyard PJ. Microarray interrogation of human metanephric mesenchymal cells highlights potentially important molecules in vivo. Physiol Genomics. 2007;28:193–202. doi: 10.1152/physiolgenomics.00147.2006. [DOI] [PubMed] [Google Scholar]
  • 59.Hilpert J, Wogensen L, Thykjaer T, Wellner M, Schlichting U, Orntoft TF, Bachmann S, Nykjaer A, Willnow TE. Expression profiling confirms the role of endocytic receptor megalin in renal vitamin D3 metabolism. Kidney Int. 2002;62:1672–1681. doi: 10.1046/j.1523-1755.2002.00634.x. [DOI] [PubMed] [Google Scholar]
  • 60.Norman LP, Jiang W, Han X, Saunders TL, Bond JS. Targeted disruption of the meprin beta gene in mice leads to underrepresentation of knockout mice and changes in renal gene expression profiles. Mol Cell Biol. 2003;23:1221–1230. doi: 10.1128/MCB.23.4.1221-1230.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.McReynolds MR, Taylor-Garcia KM, Greer KA, Hoying JB, Brooks HL. Renal medullary gene expression in aquaporin-1 null mice. Am J Physiol Renal Physiol. 2005;288:F315–F321. doi: 10.1152/ajprenal.00207.2004. [DOI] [PubMed] [Google Scholar]
  • 62.Bachvarov D, Bachvarova M, Koumangaye R, Klein J, Pesquero JB, Neau E, Bader M, Schanstra JP, Bascands JL. Renal gene expression profiling using kinin B1 and B2 receptor knockout mice reveals comparable modulation of functionally related genes. Biol Chem. 2006;387:15–22. doi: 10.1515/BC.2006.004. [DOI] [PubMed] [Google Scholar]
  • 63.Grote D, Souabni A, Busslinger M, Bouchard M. Pax 2/8-regulated Gata 3 expression is necessary for morphogenesis and guidance of the nephric duct in the developing kidney. Development. 2006;133:53–61. doi: 10.1242/dev.02184. [DOI] [PubMed] [Google Scholar]
  • 64.Riera M, Burtey S, Fontes M. Transcriptome analysis of a rat PKD model: Importance of genes involved in extracellular matrix metabolism. Kidney Int. 2006;69:1558–1563. doi: 10.1038/sj.ki.5000309. [DOI] [PubMed] [Google Scholar]
  • 65.Schwab K, Hartman HA, Liang HC, Aronow BJ, Patterson LT, Potter SS. Comprehensive microarray analysis of Hoxa11/ Hoxd11 mutant kidney development. Dev Biol. 2006;293:540–554. doi: 10.1016/j.ydbio.2006.02.023. [DOI] [PubMed] [Google Scholar]
  • 66.Shirota S, Yoshida T, Sakai M, Kim JI, Sugiura H, Oishi T, Nitta K, Tsuchiya K. Correlation between the expression level of c-maf and glutathione peroxidase-3 in c-maf −/− mice kidney and c-maf overexpressed renal tubular cells. Biochem Biophys Res Commun. 2006;348:501–506. doi: 10.1016/j.bbrc.2006.07.111. [DOI] [PubMed] [Google Scholar]
  • 67.Schwab KR, Patterson LT, Hartman HA, Song N, Lang RA, Lin X, Potter SS. Pygo1 and Pygo2 roles in Wnt signaling in mammalian kidney development. BMC Biol. 2007;5:15. doi: 10.1186/1741-7007-5-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Cai Q, McReynolds MR, Keck M, Greer KA, Hoying JB, Brooks HL. Vasopressin receptor subtype 2 activation increases cell proliferation in the renal medulla of AQP1 null mice. Am J Physiol Renal Physiol. 2007;293:F1858–F1864. doi: 10.1152/ajprenal.00068.2007. [DOI] [PubMed] [Google Scholar]
  • 69.Done SC, Takemoto M, He L, Sun Y, Hultenby K, Betsholtz C, Tryggvason K. Nephrin is involved in podocyte maturation but not survival during glomerular development. Kidney Int. 2008;73:697–704. doi: 10.1038/sj.ki.5002707. [DOI] [PubMed] [Google Scholar]
  • 70.Kim YS, Kang HS, Herbert R, Beak JY, Collins JB, Grissom SF, Jetten AM. Kruppel-like zinc finger protein Glis2 is essential for the maintenance of normal renal functions. Mol Cell Biol. 2008;28:2358–2367. doi: 10.1128/MCB.01722-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Valerius MT, McMahon AP. Transcriptional profiling of Wnt4 mutant mouse kidneys identifies genes expressed during nephron formation. Gene Expr Patterns. 2008;8:297–306. doi: 10.1016/j.gep.2008.02.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Okada S, Misaka T, Tanaka Y, Matsumoto I, Ishibashi K, Sasaki S, Abe K. Aquaporin-11 knockout mice and polycystic kidney disease animals share a common mechanism of cyst formation. FASEB J. 2008;22:3672–3684. doi: 10.1096/fj.08-111872. [DOI] [PubMed] [Google Scholar]
  • 73.Georgas K, Rumballe B, Valerius MT, Chiu HS, Thiagarajan RD, Lesieur E, Aronow BJ, Brunskill EW, Combes AN, Tang D, Taylor D, Grimmond SM, Potter SS, McMahon AP, Little MH. Analysis of early nephron patterning reveals a role for distal RV proliferation in fusion to the ureteric tip via a cap mesenchyme-derived connecting segment. Dev Biol. 2009;332:273–286. doi: 10.1016/j.ydbio.2009.05.578. [DOI] [PubMed] [Google Scholar]
  • 74.Drummond I. Making a zebrafish kidney: a tale of two tubes. Trends Cell Biol. 2003;13:357–365. doi: 10.1016/s0962-8924(03)00124-7. [DOI] [PubMed] [Google Scholar]
  • 75.Wingert RA, Davidson AJ. The zebrafish pronephros: a model to study nephron segmentation. Kidney Int. 2008;73:1120–1127. doi: 10.1038/ki.2008.37. [DOI] [PubMed] [Google Scholar]

RESOURCES