Minireviews Open Access
Copyright ©The Author(s) 2016. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Stem Cells. Sep 26, 2016; 8(9): 288-296
Published online Sep 26, 2016. doi: 10.4252/wjsc.v8.i9.288
Roles and regulation of bone morphogenetic protein-7 in kidney development and diseases
Taro Tsujimura, Mana Idei, Masahiro Yoshikawa, Osamu Takase, Keiichi Hishikawa, Department of Advanced Nephrology and Regenerative Medicine, Division of Tissue Engineering, University of Tokyo Hospital, Tokyo 113-8655, Japan
Author contributions: Tsujimura T and Hishikawa K designed and wrote the review; Idei M, Yoshikawa M and Takase O provided critical intellectual input to the study.
Supported by Grants-in-Aid for Young Scientists (B) (No. 15K18454 to Tsujimura T), for Scientific Research (B) (No. 15H03001 to Hishikawa K) and for Scientific Research (C) (Nos. 25461208 to Takase O, 15K09244 to Yoshikawa M and 26462400 to Idei M) from the Japan Society for the Promotion of Science.
Conflict-of-interest statement: All authors declare no conflict-of-interest for this study.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Keiichi Hishikawa, MD, PhD, Department of Advanced Nephrology and Regenerative Medicine, Division of Tissue Engineering, University of Tokyo Hospital, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-8655, Japan. hishikawa-tky@umin.ac.jp
Telephone: +81-3-38155411 Fax: +81-3-58009891
Received: April 27, 2016
Peer-review started: April 28, 2016
First decision: June 16, 2016
Revised: July 12, 2016
Accepted: July 20, 2016
Article in press: July 22, 2016
Published online: September 26, 2016

Abstract

The gene encoding bone morphogenetic protein-7 (BMP7) is expressed in the developing kidney in embryos and also in the mature organ in adults. During kidney development, expression of BMP7 is essential to determine the final number of nephrons in and proper size of the organ. The secreted BMP7 acts on the nephron progenitor cells to exert its dual functions: To maintain and expand the progenitor population and to provide them with competence to respond to differentiation cues, each relying on distinct signaling pathways. Intriguingly, in the adult organ, BMP7 has been implicated in protection against and regeneration from injury. Exogenous administration of recombinant BMP7 to animal models of kidney diseases has shown promising effects in counteracting inflammation, apoptosis and fibrosis evoked upon injury. Although the expression pattern of BMP7 has been well described, the mechanisms by which it is regulated have remained elusive and the processes by which the secretion sites of BMP7 impinge upon its functions in kidney development and diseases have not yet been assessed. Understanding the regulatory mechanisms will pave the way towards gaining better insight into the roles of BMP7, and to achieving desired control of the gene expression as a therapeutic strategy for kidney diseases.

Key Words: Bone morphogenetic protein-7, Therapeutics, Kidney, Development, Nephron progenitor cells, Disease, Regeneration, Chromatin conformation, Gene expression, Gene regulation

Core tip: Bone morphogenetic protein-7 (BMP7) plays crucial roles in both the development and regeneration of the kidney. The functions and mechanisms of this protein have been clarified extensively for these processes in the fetus and adult kidney. However, the functional differences of BMP7 secreted from different sites in the kidney remain undefined. We propose that uncovering the regulatory mechanism underlying BMP7 expression will help to solve that issue. Moreover, those data should pave the way towards development of a novel therapeutic strategy for kidney diseases via hyperactivation of the endogenous action of BMP7.



INTRODUCTION

Bone morphogenetic protein-7 (BMP7) belongs to the transforming growth factor-β (TGFβ) superfamily. It was first identified and cloned as a human homolog of the bovine osteogenic proteins, and designated as the osteogenic protein-1 (OP-1)[1,2]. Knockout mouse models of the BMP7 gene were reported subsequently. Most strikingly, these models exhibited severe retardation of kidney development and died soon after birth due to renal dysplasia. Additionally, these mice exhibited anophthalmia and polydactyly in the hind limbs. Other phenotypic changes were also observed in ribs and craniofacial bones, but the effects were not fully penetrant[3-5]. Since then, the roles of BMP7 in the various stages of kidney development have been extensively studied. Interestingly, BMP7 was found to be expressed not only during embryogenesis but also in the adult organ[6-8]. Series of studies have now shed light on the protective and regenerative functions of its expression in the mature kidney[9]. Furthermore, exogenous administration of BMP7 and its mimetic has been considered as a promising therapeutic strategy for treatment of severe kidney diseases[10].

Despite the original implication of an osteogenic property for BMP7, the BMP7 knockout mice showed a severe phenotype in the kidney. This finding clearly illustrated that the function of the gene is critically determined by its expression pattern. In support of this, Bmp4 under the control of the BMP7 locus rescued the loss of the developmental function of BMP7 in the kidney in mice[11]. Thus, uncovering the regulatory mechanism for BMP7 will be pivotal for gaining a better understanding of its functional roles and to developing therapeutic applications based upon it. In accordance with this perspective, in this review we first summarize the current knowledge regarding the function of BMP7 in kidney development and diseases, after which we provide an overview of the recent findings in the regulation of its expression, finally discussing the future directions that will most likely advance the knowledge and clinical applications of this field.

BMP7 IN EMBRYONIC KIDNEY DEVELOPMENT

During embryonic development, BMP7 is expressed in multiple tissues including the kidney, eyes, heart, limbs, forebrain, branchial arches, bones and cartilage[6,12,13]. In the mouse, expression in the developing kidney appears first in the Wolffian duct, at embryonic day (E)9.5, and persists in the ureteric bud evaginated from the duct[6]. At E11.5, BMP7 expression appears in the condensing mesenchyme that is induced by the ureteric bud. Slight expression is found in the uninduced metanephric mesenchyme as well[12]. At E13.5, the expression area extends to the pretubular aggregates and others derived from the condensed mesenchyme, including the comma-shaped and S-shaped bodies, the distal tubules and the podocytes of the developing glomeruli[6,12,14], although expression in the comma-shaped and S-shaped bodies and the distal tubules was found by some of the studies to be very weak or absent[4,14]. By E16.5, when the ureter has developed substantial branching, the expression in the ureteric epithelium becomes weaker in the medullar region, while its expression in the condensed mesenchyme in the nephrogenic zone of the developing kidney remains robust. Podocytes continue to strongly express BMP7 after their folding in glomeruli[3,4,12,13,15].

In mice, development of the kidney is severely retarded in the absence of Bmp7. In addition to the mutant kidney being smaller in size at birth, the number of nephrons is greatly reduced. These effects are accompanied by abnormal expansion of collecting ducts, which are interspersed by stromal cells and extracellular matrix. Mesenchymal stem cells and glomerulogenesis are also absent in the mutant kidney[3,4].

These developmental defects appear as early as E12[4]. As stated, the size of the mutant kidney is smaller than that of the control kidney, with the condensation of the mesenchyme appearing reduced at E12.5, although formation of pretubular aggregates was also observed[4]. At E14.5, cessation of nephrogenesis becomes apparent with loss of mesenchymal populations in the cortical region[3,4]. However, the comma-shaped and S-shaped bodies are present at this stage[3,4] and the ureteric buds are branched[3]. Moreover, the expression of marker genes, such as Pax2, Pax8 and Wnt4, seems more or less normal in the two lineages, as long as the corresponding structures are present[3,4].

These results suggested that the initial reciprocal inductive interaction between the ureteric epithelium and the metanephric mesenchymal cells takes place in the absence of BMP7[3]. Studies using terminal deoxynucleotidyl transferase (TdT) dUTP nick-end labeling (commonly known as the TUNEL assay) showed massive apoptosis in the uninduced metanephric mesenchymal cells occurring from E13.5 to E14.5, explaining the loss of the cell population that was observed in the mutant kidneys[4,12]. Thus, BMP7 is a survival and/or proliferative factor that acts to maintain and expand nephron progenitor cells, the loss of which leads to severe retardation of kidney development[3,4,12].

The molecular mechanisms that underlie the roles of BMP7 in kidney development have been uncovered by recent studies[16]. The collective BMPs are known to activate SMAD1, SMAD5 and SMAD8 transcription factors, which are associated with receptors of BMPs and are phosphorylated upon their binding. Following phosphorylation, these SMADs form a complex with SMAD4, which then binds to cis-regulatory regions to activate target genes. On the other hand, BMPs can also activate the mitogen-activated protein kinase (MAPK) pathway, mediating their downstream effects[17].

An in vitro study using a primary culture system of nephrogenic progenitor cells revealed that the proliferative role, but not the survival role, of BMP7 in this cell population is dependent on activation of the jun N-terminal kinase (JNK)-MAPK pathway[18]. Mice with knockout of the Trps1 gene, which encodes the tricho-rhino-phalangeal syndrome-1 zinc-finger transcription factor, show reduced epithelialization of mesenchymal progenitor cells[19]. Interestingly, expression of Trps1 in the kidney is dependent on BMP7, via p38 MAPK activation[19]. Therefore, a part of the kidney-related phenotype of the Bmp7-null mice should be due to the deficiency in Trps1 activation[19].

An additional role of BMP7 involving the SMAD pathway was recently reported[20]. In the nephrogenic zone, the progenitor populations are partitioned into two distinct compartments: One expressing detectable levels of both CITED1 and Six2 expression, and the other of Six2 only. Importantly, the Six2-only compartment is responsive to the canonical Wnt9/β-catenin signaling, showing appropriate differentiation and epithelialization; meanwhile, the CITED1+/Six2+ compartment is refractory to it[20]. It has been shown that BMP7 promotes the transition of nephron progenitor cells from the CITED1+/Six2+ compartment to the Six2-only compartment via the activation of the SMAD pathway. Thus, BMP7 has multiple roles both in the proliferation and differentiation processes of the metanephric mesenchymal progenitor cells, most likely involving distinct signaling pathways.

Use of the two distinct pathways, MAPK and SMAD, in kidney development has also been suggested by the findings from in vitro studies. In an ex vivo experiment, BMP7 was shown to control branching morphogenesis of the ureteric buds in a dose-dependent manner; specifically, low dosage of BMP7 was shown to induce morphogenesis, while high dosage was shown to exert an inhibitory effect[21]. A subsequent study showed that while the low-dose BMP7 induced p38 MAPK signaling, the high-dose triggered the SMAD pathway, which in turn caused negative regulation of the MAPK activity[22]. Such bimodal regulation might also take place in the developmental context. In this sense, it would be intriguing to understand how secretion of BMP7 from different sites, particularly the ureteric bud or the metanephric mesenchyme, impinges on the differential functional roles in the developmental process. To date, however, this process has remained largely unstudied.

Bmp7 expression continues in the developing kidney, even after the stage when the knockout mice present the severe abnormality. Several studies have aimed to uncover its roles in these later stages. A mouse strain that expresses the Cre recombinase under the promoter of Nphs2 was used to create a conditional knockout mouse in which BMP7 has near-complete specific deletion in the podocytes of mature glomeruli[23]. These mice presented with hypoplastic kidneys, and proximal tubules of markedly reduced size. Concomitantly, phosphorylation of p38 MAPK was significantly reduced in the proximal tubules[23].

Interesting phenotypes were observed upon deletion of the BMP7 alleles at E12.5, which was accomplished using a mouse line expressing an inducible Cre[24]. Deletion after the early stage of nephrogenesis resulted in precocious maturation of glomeruli, as well as increased apoptosis of the progenitor cells. In vitro assays further showed that BMP7 inhibits epithelialization of the mesenchymal progenitor cells, which might explain the precocious maturation that was observed[24]. These findings might appear to be contradictory to the above-mentioned model in which BMP7 is required for shifting the competence of the mesenchymal progenitor cells for the differentiation cue[20]; however, at an early stage, the metanephric mesenchymal cells do not need the BMP7/SMAD pathway for differentiation[20]. In fact, the BMP7 knockout mice can develop nephron structures adequately up to E13.5. Therefore, the reduction of BMP7 at E12.5 might guarantee or even accelerate the early phase of nephron formation. The late stage formation observed in the deletion mice might be attributed to the remaining expression of BMP7 (around 10% as compared to the controls) after the induction by tamoxifen, particularly as BMP7 exhibits dose-dependency in induction of the downstream cascades[22]. However, further studies are necessary to clarify this issue.

Overall, BMP7 mainly acts to determine and balance the fates of the mesenchymal progenitor cells, between proliferation and differentiation, to determine the final size of the mature kidney. Distinct pathways are utilized for these different roles. Although the mechanism to switch between the different pathways is largely unknown, dose-dependent regulation might contribute, at least partially. Therefore, regulation of the expression of BMP7 is expected to play a critical role in the developmental process, and this topic will be discussed later in this review.

BMP7 IN THE ADULT KIDNEY

Kidney-specific expression of BMP7 persists in the mature organ of the adult[7]. Its functional significance has been suggested by a series of studies. All the more, exogenous administration of BMP7 to injured kidneys was also shown to have therapeutic effects, including prevention of fibrosis, inflammation and apoptosis. We first summarize, here, the findings regarding the latter, and then we discuss the endogenous role of the protein at the end of this section.

Systemic administration of recombinant BMP7 to a rat model of ischemia/reperfusion injury was first shown to enhance recovery after acute injury by suppressing inflammatory responses, apoptosis and fibrosis[25]. In subsequent studies, the administration of BMP7 to a rat model of unilateral ureteral obstruction (UUO) using prevention protocols resulted in blunting of the development of injury[26,27].

Renal fibrosis is considered as a hallmark of chronic kidney diseases, although functional benefits of fibrosis have been recognized recently[28]. TGFβ1 is a key mediator of fibrosis in many tissues, including in the kidney in response to injury (reviewed in[29,30]). Binding of TGFβ1 to its type II receptor, TBRII, triggers the receptor to activate the TGFβ receptor type I (TBRI)-kinase, which in turn induces downstream cascades via phosphorylation of SMAD2 and SMAD3[31]. On the other hand, ligand binding to the type I activin-like kinase (Alk) receptors and type II serine/threonine kinase receptors (BMPRII) for BMP7 activates SMAD1/5/8 for SMAD signaling[31].

Roles of these signaling pathways in renal fibrosis were investigated in an in vitro model[32]. Incubation of mouse distal tubular epithelial cells (NP1) with TGFβ1 led to induction of epithelial-to-mesenchymal transition (EMT) that was associated with nuclear localization of phosphorylated SMAD2/3[32]. However, addition of BMP7 to this culture system reversed the EMT through phosphorylation and activation of SMAD1, which in turn transcriptionally up-regulated the expression of E-cadherin, an important adhesion molecule in epithelial cells[32].

Based on these findings, the counteraction of BMP7 against the TGFβ1-induced EMT was further tested in vivo. Intraperitoneal administration of BMP7 to a mouse model of progressive chronic kidney injury with nephrotoxic serum nephritis (NTN) led to reversal of the renal pathology and to a decline in the mortality rate[32]. The same group also showed that the BMP7 treatment could attenuate progression of chronic kidney fibrosis in two genetic mouse models, namely those of Alport’s syndrome and lupus nephritis[33].

Recent studies have revealed involvement of epigenetic regulation in the renoprotective effect of BMP7. Rasal1, the gene encoding rasGAP-activating-like protein 1, was shown to be aberrantly hypermethylated in the fibrotic condition that is induced by TGFβ1[34]. Reversal of fibrosis by the administered BMP7 was also found to be associated with active removal of methylation at Rasal1 via the 10-11 translocation enzyme-3 (Tet3)[35].

In contrast to the above findings showing the therapeutic effects of BMP7, the function of the endogenously expressed molecule in the adult kidney has not been thoroughly assessed to date. This might be partly due to the technical difficulty of eliminating BMP7 specifically in the adult kidney and not in the developing kidney, so as to avoid the developmental arrest that otherwise leads to death. However, several studies have demonstrated the pivotal role of endogenously expressed BMP7 in protection of the kidney from injuries.

Uterine sensitization-associated gene-1 (USAG1) is a BMP antagonist[36-38], and is abundantly expressed in the adult kidney[38]. The Usag1 knockout mice show resistance to apoptosis and fibrosis, and a down-regulation in the expression of inflammatory genes, all of which were reinduced upon administration of neutralizing antibodies against BMP7[39]. Thus, BMP7 appears to play a renoprotective role endogenously in the kidney, which is negatively regulated by USAG1. As mentioned above, fibrosis and inflammatory responses upon kidney injury seem to have beneficial effects as well for the renal function, serving to sustain the overall structure of the kidney[28]. In this sense, USAG1 and BMP7 might cooperatively serve to balance the progression of fibrosis, and titration of these two proteins in the progression of kidney diseases might be an exciting approach for therapeutics.

Kielin/chordin-like protein (KCP) is, on the other hand, an enhancer of BMP signaling. Interestingly, the knockout mouse of the encoding gene develops susceptibility to kidney injury, further demonstrating the protective role of BMP7 in the adult kidney[40].

Activin-like kinase 3 (Alk3) is one of the three type I receptors for BMP7[31,41]. During the progression of kidney injury, Alk3 is up-regulated, while the other receptor genes, Alk2 and Alk6, are not. Loss of Alk3 leads to more severe fibrosis and inflammatory response upon NTN-induced chronic kidney fibrosis, further supporting the theory that BMP7 exerts renoprotective functions through binding to Alk3 endogenously[41]. Furthermore, the small peptide agonist THR-123 exhibits therapeutic effects when applied to different models of kidney injuries, through its interaction with Alk3[41].

REGULATION OF BMP7 EXPRESSION IN THE EMBRYONIC AND ADULT KIDNEY

In the mouse embryo, BMP7 is expressed not only in the kidney but also in various other tissues. A recent study showed that expression in extra-nephrotic domains is mostly regulated by long-range enhancers that activate gene expression in a tissue-specific manner around the locus[42]. The previous in vivo studies had identified some of the enhancers capable of inducing reporter gene expression in the developing kidney. One such element is located in intron 1 of Bmp7, which is strongly conserved in tetrapods[43]. When heterologously linked to the lacZ reporter under control of a minimal promoter sequence (Hsp68lacZ)[44], the element induced lacZ expression in the Wolffian duct, mesonephric tubules, ureteric bud and collecting duct, from E9.5 until E12.5, but not in the metanephric mesenchymal lineage. Of note, this expression could not be recapitulated by the orthologous sequence in Xenopus tropicalis. A lacZ reporter construct under control of the endogenous promoter of BMP7 was also injected into mouse embryos together with the 4-kb upstream and the 3-kb downstream regions of the transcription start site (TSS) at each side, retaining the endogenous context[43]. This entire construct was able to induce reporter expression in the nephrogenic mesenchymal regions. Interestingly, however, none of the separated individual elements covering the upstream and downstream regions was able to drive the expression when linked to Hsp68lacZ[43].

Chromatin immunoprecipitation coupled with high-throughput sequencing (ChIP-seq) was performed in Six2+ nephron progenitor cells[45]. A region located at 98-kb downstream of the TSS of BMP7 was found to be co-bound by Six2 and β-catenin. When tested in vivo, this element induced gene expression in a compartment of Six2- renal vesicles, a part of the BMP7 expression domain[45]. The ChIP-seq also identified a Six2 binding site in intron 1, an evolutionarily conserved region adjacent to the intron 1 enhancer, though a reporter construct including this region did not show enhancer activity in the developing kidney[43]. These results suggest that BMP7 expression in the different compartments of the developing kidney, notably the ureteric bud and the metanephric mesenchyme, might be regulated by a distinct set of enhancers that are active in the respective domains, possibly interacting in a cooperative manner with each other. However, to determine whether or not these elements actually contribute to the BMP7 expression in the kidney, testing by deletion of the respective regions should be performed in future studies.

Bmp7 is highly and specifically expressed in the adult kidney under normal physiologic conditions[7]. The main expression domains are the ureter, collecting duct, thick ascending limb, distal convoluted tubules and podocytes in the glomerulus[46,47]. BMP7 expression in these cell types might also be regulated by cis-regulatory enhancers that are embedded around the locus. However, to date, no such enhancer elements have been described for the expression in adult kidney. Acetylation of K27 of histone H3 (H3K27ac) is associated with enhancer activity of the marked regions. We compared released data from the ENCODE project of ChIP-seq for H3K27ac in kidney tissues at different time points, ranging from E14.5 to the adult stage[48]. In Figure 1, the regions with peaks are more or less common between different stages, but there are some striking differences. Notably, the region of the intron 1 enhancer is highly acetylated at the embryonic stages, but the mark is almost diminished for the adult kidney. These data might suggest that different sets of cis-regulatory regions contribute to the expression at different stages.

Figure 1
Figure 1 Landscape of the enhances in the kidney and topological chromatin domains around the BMP7 locus. Exons and introns of the genes within the locus (chr2: 172,250,000-172,850,000 mm9) are represented by blue boxes and arrowed dashes, respectively. Names of the genes are indicated above or below the respective boxes. The ChIP-seq signals of H3K27ac in the kidney tissue at different stages [E14.5, E15.5, E16.5, postnatal day (P)0 and 8-wk-old; indicated to the right] are shown with pink plots. H3K4me1 and H3K27me3 signals in the adult kidney are shown with green and red plots, respectively. The ChIP-seq data were obtained from the ENCODE project[48]; the Data Coordination Consortium accession numbers are ENCSR703ZPF, ENCSR000CAF and ENCSR000CFP for H3K27ac, H3K4me1 and H3K27me3, respectively. Enhancer candidates are represented by pink ovals at the bottom: The RV enhancer is bound by Six2 and β-catenin, and induces reporter expression in the Six2- compartment of the renal vesicle[45]; the intron 1 enhancer drives reporter expression in the developing ureteric buds[43]; the Six2 binding site next to the intron 1 enhancer was identified by ChIP-seq[45] but is not sufficient to induce gene expression in the developing kidney[43]. Note that the H3K27ac mark around the intron 1 enhancer during embryogenesis diminishes in the adult stage. The chromatin domains identified by 4C-seq are shown at the top (indicated by whiskered lines), together with the actual results of the 4C-seq that are shown below[42]. The viewpoints of the 4C-seq are indicated by arrows on the plots: Tfap2c promoter (blue), transition zone (TZ; gray), next to TZ in the BMP7 domain (yellow), and BMP7 promoter (green). The European Nucleotide Archive accession number of the 4C-seq data is ERP005557[42]. The TZ between the two domains is indicated by the purple rectangle that spans the entire diagram.

Once the kidney is damaged, the levels and sites of BMP7 expression are dynamically altered[15,47]. The injuries cause dramatic response to the cellular states in the kidney via the inflammatory response and other signaling cascades, such as that involving TGFβ1. Such responses are expected to lead to alteration of epigenetic states around the BMP7 locus; as a result, BMP7 expression would be dynamically regulated. In the following passages of this review, we review the expression dynamics of BMP7 in several kidney disease models, as reported in the literature to date, and discuss how they are regulated.

In the kidneys of the ischemia/reperfusion rat model, BMP7 expression dramatically decreases soon after reperfusion, particularly in the outer medulla and glomeruli[47,49]. This reduction might be related to the up-regulation of TGFβ1. The immunostain signal of BMP7 increases in proximal tubular epithelial cells, which are devoid of its expression in the normal situation[50]. Similar up-regulation of BMP7 in the proximal tubular epithelial cells following ischemia was confirmed in a mouse model[51]. Proximal tubular epithelial cells from human patients with proteinuric nephropathies also showed up-regulation of BMP7 as compared to that in healthy controls[52]. In an experimental model of diabetic nephropathy, BMP7 expression was decreased[53], which might be due at least partly to the concomitant increase in Tgfb1 expression. A tubular injury induced by folic acid resulted in reduced BMP7 expression at first, but was followed by a gradual recovery of expression in the regenerative phase[15]. In the cisplatin nephrotoxicity model, however, no or only a subtle increase in BMP7 expression was scored[15,54].

It has long been postulated that TGFβ1, which is an inducer of the fibrotic response upon injury, down-regulates BMP7 expression (as discussed above). On the other hand, MyoR has been implicated in the activation of Bmp7[54,55]. However it has not been assessed adequately to conclude whether these effects are direct or not.

Epigenetic regulation has been studied to understand the direct linkages between various cellular states and BMP7 expression. In TGFβ1-induced EMT in human renal proximal tubular epithelial cells, BMP7 is slightly down-regulated[56]. Treatment with trichostatin A, a histone deacetylase (HDAC) inhibitor, however, led to deposition of acetylated histones around the promoter of BMP7 and to induction of its expression, thereby counteracting the fibrosis[56]. Consistently, in an ischemia/reperfusion mouse model, down-regulation of HDAC5 was found to be involved in the activation of BMP7 during the regenerative phase following injury, probably via acetylation of histones[51].

Another layer of epigenetic regulation might also impinge on the expression of BMP7 in kidney. The topological conformation of chromatin has recently been recognized as an important determinant of transcriptional regulation of genes. Particularly, a topologically-associating domain (TAD; a compartmentalized block of the genome, in which the genomic regions preferentially associate with each other) was characterized as a ground-state structure that facilitates and constrains the interaction between enhancers and promoters of genes within it[57-59].

Bmp7 is flanked by a large intergenic region, at the other side of which a developmental gene Tfap2c is located (Figure 1). A recent study showed that the locus is conformationally partitioned into two adjacent domains, one for BMP7 and the other for Tfap2c, by function of a region at the boundary, termed TZ[42]. At this locus, the action of enhancers is limited to genes located within the same domain that they belong to. In Figure 1, the TZ and 4C (circular chromatin conformation capture) plots that describe the physical domain structure were overlaid on the ChIP-seq map. It is apparent that the acetylation only extends within the BMP7 domain and not to the neighboring one, underlining the importance of the topological structure for the regulation of BMP7 in the kidney as well (Figure 1).

TADs seem to be more or less stable in different cell types. This might be due to the fact that CTCF, a ubiquitous DNA binding protein, greatly contributes to the formation of the domain structures[57,60]. However, the topological structure is also a function of other epigenetic modifications, such as transcription of constitutive genes and polycomb group proteins[57,61]. Indeed, the extent that enhancers can activate genes is sometimes different among different enhancers at the same locus[42,59]. Therefore, it might be possible that the topological structure is subject to regulation for the dynamic expression of BMP7 in response to kidney injuries.

CONCLUSION

BMP7 plays an important role in development and diseases of the kidney. In development, BMP7 is critical both in proliferation and maintenance of the kidney’s mesenchymal stem cells and in shifting their competence to respond to differentiation cues. Consequently, BMP7 is a critical determinant of nephron numbers and the size of the organ. At the adult stage, BMP7 is implicated in protection and regeneration of the kidney upon injury. Moreover, administration of BMP7 and its mimetic exerts therapeutic effects in conditions of both acute kidney injuries and chronic kidney diseases.

Precise regulation of the BMP7 gene is critical to its function in the kidney. At the embryonic stage, the major expression sites are metanephric mesenchyme and ureteric buds, two different lineages that interact with each other. Studies to date have indicated that the active locale of BMP7 is the mesenchymal cells, rather than the ureteric epithelia, but the functional difference of the expression sites remains elusive. In this sense, it will be insightful to identify cis-regulatory elements that induce BMP7 expression in the different cell populations and to test impacts of mutations in the elements on kidney development. The regulatory mechanisms of the adult kidney also remain elusive. Identification of the enhancers will provide insight into the still opaque role of BMP7 in the adult kidney.

Understanding the epigenetic mechanism may not only clarify the dynamic regulation of the gene, but also open up a new avenue for therapeutics for kidney diseases through control of the expression of Bmp7. Different layers of epigenetic regulation, such as DNA methylation, histone modifications and higher-order chromatin conformation, almost certainly will bear a role in achieving delicate control of the gene’s expression. Each of the layers represents a possible target for therapeutics. Indeed, inhibition of HDAC has already shown a promising effect in augmenting BMP7 expression in a TGFβ1-induced fibrosis model[56]. Furthermore, recent advances in genome editing tools, such as the CRISPR/Cas9 system, might allow us to control epigenetic modifications, including higher-order chromatin conformation, in a locus-specific manner to optimize the gene expression[62]. To this end, it will be beneficial to further deepen our understanding both of the role and regulation of BMP7 in kidney development and diseases.

ACKNOWLEDGMENTS

We thank to the ENCODE consortium and the laboratory of Prof. Bing Ren (University of California San Diego, CA, United States) for the use of the released data of ChIP-seq.

Footnotes

Manuscript source: Invited manuscript

Specialty type: Cell and tissue engineering

Country of origin: Japan

Peer-review report classification

Grade A (Excellent): 0

Grade B (Very good): B, B, B

Grade C (Good): 0

Grade D (Fair): 0

Grade E (Poor): 0

P- Reviewer: Bhimma R, Cheng TH, Gharaee-Kermani M S- Editor: Ji FF L- Editor: A E- Editor: Lu YJ

References
1.  Ozkaynak E, Rueger DC, Drier EA, Corbett C, Ridge RJ, Sampath TK, Oppermann H. OP-1 cDNA encodes an osteogenic protein in the TGF-beta family. EMBO J. 1990;9:2085-2093.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Salazar VS, Gamer LW, Rosen V. BMP signalling in skeletal development, disease and repair. Nat Rev Endocrinol. 2016;12:203-221.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 491]  [Cited by in F6Publishing: 514]  [Article Influence: 64.3]  [Reference Citation Analysis (0)]
3.  Dudley AT, Lyons KM, Robertson EJ. A requirement for bone morphogenetic protein-7 during development of the mammalian kidney and eye. Genes Dev. 1995;9:2795-2807.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 861]  [Cited by in F6Publishing: 809]  [Article Influence: 27.9]  [Reference Citation Analysis (0)]
4.  Luo G, Hofmann C, Bronckers AL, Sohocki M, Bradley A, Karsenty G. BMP-7 is an inducer of nephrogenesis, and is also required for eye development and skeletal patterning. Genes Dev. 1995;9:2808-2820.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 777]  [Cited by in F6Publishing: 733]  [Article Influence: 25.3]  [Reference Citation Analysis (0)]
5.  Zouvelou V, Luder HU, Mitsiadis TA, Graf D. Deletion of BMP7 affects the development of bones, teeth, and other ectodermal appendages of the orofacial complex. J Exp Zool B Mol Dev Evol. 2009;312B:361-374.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 61]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
6.  Lyons KM, Hogan BL, Robertson EJ. Colocalization of BMP 7 and BMP 2 RNAs suggests that these factors cooperatively mediate tissue interactions during murine development. Mech Dev. 1995;50:71-83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 345]  [Cited by in F6Publishing: 362]  [Article Influence: 12.5]  [Reference Citation Analysis (0)]
7.  Ozkaynak E, Schnegelsberg PN, Oppermann H. Murine osteogenic protein (OP-1): high levels of mRNA in kidney. Biochem Biophys Res Commun. 1991;179:116-123.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Archdeacon P, Detwiler RK. Bone morphogenetic protein 7 (BMP7): a critical role in kidney development and a putative modulator of kidney injury. Adv Chronic Kidney Dis. 2008;15:314-320.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 12]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
9.  Yanagita M. Inhibitors/antagonists of TGF-β system in kidney fibrosis. Nephrol Dial Transplant. 2012;27:3686-3691.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 75]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
10.  Tampe D, Zeisberg M. Potential approaches to reverse or repair renal fibrosis. Nat Rev Nephrol. 2014;10:226-237.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 132]  [Cited by in F6Publishing: 136]  [Article Influence: 13.6]  [Reference Citation Analysis (0)]
11.  Oxburgh L, Dudley AT, Godin RE, Koonce CH, Islam A, Anderson DC, Bikoff EK, Robertson EJ. BMP4 substitutes for loss of BMP7 during kidney development. Dev Biol. 2005;286:637-646.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 47]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
12.  Dudley AT, Robertson EJ. Overlapping expression domains of bone morphogenetic protein family members potentially account for limited tissue defects in BMP7 deficient embryos. Dev Dyn. 1997;208:349-362.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 4]  [Reference Citation Analysis (0)]
13.  Godin RE, Takaesu NT, Robertson EJ, Dudley AT. Regulation of BMP7 expression during kidney development. Development. 1998;125:3473-3482.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Vukicevic S, Kopp JB, Luyten FP, Sampath TK. Induction of nephrogenic mesenchyme by osteogenic protein 1 (bone morphogenetic protein 7). Proc Natl Acad Sci USA. 1996;93:9021-9026.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Tanaka M, Endo S, Okuda T, Economides AN, Valenzuela DM, Murphy AJ, Robertson E, Sakurai T, Fukatsu A, Yancopoulos GD. Expression of BMP-7 and USAG-1 (a BMP antagonist) in kidney development and injury. Kidney Int. 2008;73:181-191.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 61]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
16.  Nishinakamura R, Sakaguchi M. BMP signaling and its modifiers in kidney development. Pediatr Nephrol. 2014;29:681-686.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 35]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
17.  Oxburgh L, Brown AC, Fetting J, Hill B. BMP signaling in the nephron progenitor niche. Pediatr Nephrol. 2011;26:1491-1497.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
18.  Blank U, Brown A, Adams DC, Karolak MJ, Oxburgh L. BMP7 promotes proliferation of nephron progenitor cells via a JNK-dependent mechanism. Development. 2009;136:3557-3566.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 98]  [Cited by in F6Publishing: 95]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
19.  Gai Z, Zhou G, Itoh S, Morimoto Y, Tanishima H, Hatamura I, Uetani K, Ito M, Muragaki Y. Trps1 functions downstream of BMP7 in kidney development. J Am Soc Nephrol. 2009;20:2403-2411.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 35]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
20.  Brown AC, Muthukrishnan SD, Guay JA, Adams DC, Schafer DA, Fetting JL, Oxburgh L. Role for compartmentalization in nephron progenitor differentiation. Proc Natl Acad Sci USA. 2013;110:4640-4645.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 110]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
21.  Piscione TD, Phan T, Rosenblum ND. BMP7 controls collecting tubule cell proliferation and apoptosis via Smad1-dependent and -independent pathways. Am J Physiol Renal Physiol. 2001;280:F19-F33.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Hu MC, Wasserman D, Hartwig S, Rosenblum ND. p38MAPK acts in the BMP7-dependent stimulatory pathway during epithelial cell morphogenesis and is regulated by Smad1. J Biol Chem. 2004;279:12051-12059.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 51]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
23.  Kazama I, Mahoney Z, Miner JH, Graf D, Economides AN, Kreidberg JA. Podocyte-derived BMP7 is critical for nephron development. J Am Soc Nephrol. 2008;19:2181-2191.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 53]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
24.  Tomita M, Asada M, Asada N, Nakamura J, Oguchi A, Higashi AY, Endo S, Robertson E, Kimura T, Kita T. BMP7 maintains undifferentiated kidney progenitor population and determines nephron numbers at birth. PLoS One. 2013;8:e73554.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 31]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
25.  Vukicevic S, Basic V, Rogic D, Basic N, Shih MS, Shepard A, Jin D, Dattatreyamurty B, Jones W, Dorai H. Osteogenic protein-1 (bone morphogenetic protein-7) reduces severity of injury after ischemic acute renal failure in rat. J Clin Invest. 1998;102:202-214.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 257]  [Cited by in F6Publishing: 237]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
26.  Hruska KA, Guo G, Wozniak M, Martin D, Miller S, Liapis H, Loveday K, Klahr S, Sampath TK, Morrissey J. Osteogenic protein-1 prevents renal fibrogenesis associated with ureteral obstruction. Am J Physiol Renal Physiol. 2000;279:F130-F143.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Morrissey J, Hruska K, Guo G, Wang S, Chen Q, Klahr S. Bone morphogenetic protein-7 improves renal fibrosis and accelerates the return of renal function. J Am Soc Nephrol. 2002;13 Suppl 1:S14-S21.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Kaissling B, Lehir M, Kriz W. Renal epithelial injury and fibrosis. Biochim Biophys Acta. 2013;1832:931-939.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 93]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
29.  Branton MH, Kopp JB. TGF-beta and fibrosis. Microbes Infect. 1999;1:1349-1365.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 447]  [Cited by in F6Publishing: 460]  [Article Influence: 18.4]  [Reference Citation Analysis (0)]
30.  Lan HY. Diverse roles of TGF-β/Smads in renal fibrosis and inflammation. Int J Biol Sci. 2011;7:1056-1067.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 405]  [Cited by in F6Publishing: 478]  [Article Influence: 36.8]  [Reference Citation Analysis (0)]
31.  Moustakas A, Heldin CH. The regulation of TGFbeta signal transduction. Development. 2009;136:3699-3714.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 626]  [Cited by in F6Publishing: 628]  [Article Influence: 41.9]  [Reference Citation Analysis (0)]
32.  Zeisberg M, Hanai J, Sugimoto H, Mammoto T, Charytan D, Strutz F, Kalluri R. BMP-7 counteracts TGF-beta1-induced epithelial-to-mesenchymal transition and reverses chronic renal injury. Nat Med. 2003;9:964-968.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1051]  [Cited by in F6Publishing: 1049]  [Article Influence: 50.0]  [Reference Citation Analysis (0)]
33.  Zeisberg M, Bottiglio C, Kumar N, Maeshima Y, Strutz F, Müller GA, Kalluri R. Bone morphogenic protein-7 inhibits progression of chronic renal fibrosis associated with two genetic mouse models. Am J Physiol Renal Physiol. 2003;285:F1060-F1067.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 224]  [Cited by in F6Publishing: 241]  [Article Influence: 11.5]  [Reference Citation Analysis (0)]
34.  Bechtel W, McGoohan S, Zeisberg EM, Müller GA, Kalbacher H, Salant DJ, Müller CA, Kalluri R, Zeisberg M. Methylation determines fibroblast activation and fibrogenesis in the kidney. Nat Med. 2010;16:544-550.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 455]  [Cited by in F6Publishing: 485]  [Article Influence: 34.6]  [Reference Citation Analysis (0)]
35.  Tampe B, Tampe D, Müller CA, Sugimoto H, LeBleu V, Xu X, Müller GA, Zeisberg EM, Kalluri R, Zeisberg M. Tet3-mediated hydroxymethylation of epigenetically silenced genes contributes to bone morphogenic protein 7-induced reversal of kidney fibrosis. J Am Soc Nephrol. 2014;25:905-912.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 91]  [Cited by in F6Publishing: 99]  [Article Influence: 9.9]  [Reference Citation Analysis (0)]
36.  Avsian-Kretchmer O, Hsueh AJ. Comparative genomic analysis of the eight-membered ring cystine knot-containing bone morphogenetic protein antagonists. Mol Endocrinol. 2004;18:1-12.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 181]  [Cited by in F6Publishing: 190]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
37.  Laurikkala J, Kassai Y, Pakkasjärvi L, Thesleff I, Itoh N. Identification of a secreted BMP antagonist, ectodin, integrating BMP, FGF, and SHH signals from the tooth enamel knot. Dev Biol. 2003;264:91-105.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 205]  [Cited by in F6Publishing: 209]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
38.  Yanagita M, Oka M, Watabe T, Iguchi H, Niida A, Takahashi S, Akiyama T, Miyazono K, Yanagisawa M, Sakurai T. USAG-1: a bone morphogenetic protein antagonist abundantly expressed in the kidney. Biochem Biophys Res Commun. 2004;316:490-500.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 113]  [Cited by in F6Publishing: 108]  [Article Influence: 5.4]  [Reference Citation Analysis (0)]
39.  Yanagita M, Okuda T, Endo S, Tanaka M, Takahashi K, Sugiyama F, Kunita S, Takahashi S, Fukatsu A, Yanagisawa M. Uterine sensitization-associated gene-1 (USAG-1), a novel BMP antagonist expressed in the kidney, accelerates tubular injury. J Clin Invest. 2006;116:70-79.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 120]  [Cited by in F6Publishing: 116]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
40.  Lin J, Patel SR, Cheng X, Cho EA, Levitan I, Ullenbruch M, Phan SH, Park JM, Dressler GR. Kielin/chordin-like protein, a novel enhancer of BMP signaling, attenuates renal fibrotic disease. Nat Med. 2005;11:387-393.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 149]  [Cited by in F6Publishing: 154]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
41.  Sugimoto H, LeBleu VS, Bosukonda D, Keck P, Taduri G, Bechtel W, Okada H, Carlson W, Bey P, Rusckowski M. Activin-like kinase 3 is important for kidney regeneration and reversal of fibrosis. Nat Med. 2012;18:396-404.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 181]  [Cited by in F6Publishing: 181]  [Article Influence: 15.1]  [Reference Citation Analysis (0)]
42.  Tsujimura T, Klein FA, Langenfeld K, Glaser J, Huber W, Spitz F. A discrete transition zone organizes the topological and regulatory autonomy of the adjacent tfap2c and bmp7 genes. PLoS Genet. 2015;11:e1004897.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 50]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
43.  Adams D, Karolak M, Robertson E, Oxburgh L. Control of kidney, eye and limb expression of BMP7 by an enhancer element highly conserved between species. Dev Biol. 2007;311:679-690.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 17]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
44.  Sasaki H, Hogan BL. Enhancer analysis of the mouse HNF-3 beta gene: regulatory elements for node/notochord and floor plate are independent and consist of multiple sub-elements. Genes Cells. 1996;1:59-72.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 80]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
45.  Park JS, Ma W, O’Brien LL, Chung E, Guo JJ, Cheng JG, Valerius MT, McMahon JA, Wong WH, McMahon AP. Six2 and Wnt regulate self-renewal and commitment of nephron progenitors through shared gene regulatory networks. Dev Cell. 2012;23:637-651.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 192]  [Cited by in F6Publishing: 183]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
46.  Gould SE, Day M, Jones SS, Dorai H. BMP-7 regulates chemokine, cytokine, and hemodynamic gene expression in proximal tubule cells. Kidney Int. 2002;61:51-60.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 184]  [Cited by in F6Publishing: 178]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
47.  Simon M, Maresh JG, Harris SE, Hernandez JD, Arar M, Olson MS, Abboud HE. Expression of bone morphogenetic protein-7 mRNA in normal and ischemic adult rat kidney. Am J Physiol. 1999;276:F382-F389.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  ENCODE Project. An integrated encyclopedia of DNA elements in the human genome. Nature. 2012;489:57-74.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14285]  [Cited by in F6Publishing: 11834]  [Article Influence: 986.2]  [Reference Citation Analysis (0)]
49.  Almanzar MM, Frazier KS, Dube PH, Piqueras AI, Jones WK, Charette MF, Paredes AL. Osteogenic protein-1 mRNA expression is selectively modulated after acute ischemic renal injury. J Am Soc Nephrol. 1998;9:1456-1463.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Villanueva S, Céspedes C, Vio CP. Ischemic acute renal failure induces the expression of a wide range of nephrogenic proteins. Am J Physiol Regul Integr Comp Physiol. 2006;290:R861-R870.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 121]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
51.  Marumo T, Hishikawa K, Yoshikawa M, Fujita T. Epigenetic regulation of BMP7 in the regenerative response to ischemia. J Am Soc Nephrol. 2008;19:1311-1320.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 77]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
52.  Rudnicki M, Eder S, Perco P, Enrich J, Scheiber K, Koppelstätter C, Schratzberger G, Mayer B, Oberbauer R, Meyer TW. Gene expression profiles of human proximal tubular epithelial cells in proteinuric nephropathies. Kidney Int. 2007;71:325-335.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 60]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
53.  Wang SN, Lapage J, Hirschberg R. Loss of tubular bone morphogenetic protein-7 in diabetic nephropathy. J Am Soc Nephrol. 2001;12:2392-2399.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Kamiura N, Hirahashi J, Matsuzaki Y, Idei M, Takase O, Fujita T, Takato T, Hishikawa K. Basic helix-loop-helix transcriptional factor MyoR regulates BMP-7 in acute kidney injury. Am J Physiol Renal Physiol. 2013;304:F1159-F1166.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 6]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
55.  Hishikawa K, Marumo T, Miura S, Nakanishi A, Matsuzaki Y, Shibata K, Ichiyanagi T, Kohike H, Komori T, Takahashi I. Musculin/MyoR is expressed in kidney side population cells and can regulate their function. J Cell Biol. 2005;169:921-928.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 95]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
56.  Yoshikawa M, Hishikawa K, Marumo T, Fujita T. Inhibition of histone deacetylase activity suppresses epithelial-to-mesenchymal transition induced by TGF-beta1 in human renal epithelial cells. J Am Soc Nephrol. 2007;18:58-65.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 159]  [Cited by in F6Publishing: 164]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
57.  Dixon JR, Selvaraj S, Yue F, Kim A, Li Y, Shen Y, Hu M, Liu JS, Ren B. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature. 2012;485:376-380.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4690]  [Cited by in F6Publishing: 4325]  [Article Influence: 360.4]  [Reference Citation Analysis (0)]
58.  Nora EP, Lajoie BR, Schulz EG, Giorgetti L, Okamoto I, Servant N, Piolot T, van Berkum NL, Meisig J, Sedat J. Spatial partitioning of the regulatory landscape of the X-inactivation centre. Nature. 2012;485:381-385.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2145]  [Cited by in F6Publishing: 1963]  [Article Influence: 163.6]  [Reference Citation Analysis (0)]
59.  Symmons O, Uslu VV, Tsujimura T, Ruf S, Nassari S, Schwarzer W, Ettwiller L, Spitz F. Functional and topological characteristics of mammalian regulatory domains. Genome Res. 2014;24:390-400.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 367]  [Cited by in F6Publishing: 298]  [Article Influence: 29.8]  [Reference Citation Analysis (0)]
60.  Rao SS, Huntley MH, Durand NC, Stamenova EK, Bochkov ID, Robinson JT, Sanborn AL, Machol I, Omer AD, Lander ES. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell. 2014;159:1665-1680.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6253]  [Cited by in F6Publishing: 4551]  [Article Influence: 455.1]  [Reference Citation Analysis (0)]
61.  Schoenfelder S, Sugar R, Dimond A, Javierre BM, Armstrong H, Mifsud B, Dimitrova E, Matheson L, Tavares-Cadete F, Furlan-Magaril M. Polycomb repressive complex PRC1 spatially constrains the mouse embryonic stem cell genome. Nat Genet. 2015;47:1179-1186.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 320]  [Cited by in F6Publishing: 255]  [Article Influence: 28.3]  [Reference Citation Analysis (0)]
62.  Dominguez AA, Lim WA, Qi LS. Beyond editing: repurposing CRISPR-Cas9 for precision genome regulation and interrogation. Nat Rev Mol Cell Biol. 2016;17:5-15.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 566]  [Cited by in F6Publishing: 558]  [Article Influence: 62.0]  [Reference Citation Analysis (0)]