banner



Can A Kidney Repair Itself

  • Journal List
  • Fibrogenesis Tissue Repair
  • v.2; 2009
  • PMC2711960

Fibrogenesis Tissue Repair. 2009; ii: 3.

Possible mechanisms of kidney repair

Paola Romagnani

1Excellence Center for Research, Transfer and High Educational activity DENOthe, University of Florence, Florence, Italy

3Department of Medicine, Harvard Medical School, Partition of Matrix Biology, Department of Medicine Beth Israel Deaconess Medical Center, Boston, Massachusetts, U.s.a.

Raghu Kalluri

iiBeth State of israel Deaconess Medical Eye, Boston, Massachusetts, USA

Received 2009 Jan two; Accepted 2009 Jun 26.

Abstract

In most adult epithelia the process of replacing damaged or dead cells is maintained through the presence of stem/progenitor cells, which allow epithelial tissues to be repaired post-obit injury. Existing prove strongly supports the presence of stem cells in the adult kidney. Indeed, recent findings provide evidence in favour of a office for intrinsic renal cells and against a physiological role for bone marrow-derived stem cells in the regeneration of renal epithelial cells. In improver, recent studies have identified a subset of CD24+CD133+ renal progenitors within the Bowman's capsule of adult human kidney, which provides regenerative potential for injured renal epithelial cells. Intriguingly, CD24+CD133+ renal progenitors also represent common progenitors of tubular cells and podocytes during renal development. Chronic injury causes dysfunction of the tubular epithelial cells, which triggers the release of fibrogenic cytokines and recruitment of inflammatory cells to injured kidneys. The rapid interposition of scar tissue probably confers a survival advantage by preventing infectious microorganisms from invading the wound, but prevents subsequent tissue regeneration. Nevertheless, the existence of renal epithelial progenitors in the kidney suggests a possible explanation for the regression of renal lesions which has been observed in experimental animals and even in humans. Thus, manipulation of the wound repair process in gild to shift information technology towards regeneration will probably require the power to slow the rapid fibrotic response so that renal progenitor cells can let tissue regeneration rather than scar formation.

Background

Most epithelia need to constantly supplant damaged or dead cells throughout life. The procedure of continual cell replacement is critical for the maintenance of adult tissues and is typically maintained through the presence of stem cells. Stalk cells are functionally defined by their ability to self-renew and to differentiate into the jail cell lineages of their tissue of origin [1]. In one case activated, epithelial stalk cells tin can generate proliferating progeny, which are oftentimes referred to as transiently amplifying cells. In their normal environment, transiently amplifying cells will divide actively for a restricted period of time, expanding the cellular pool that will then differentiate along a particular cell lineage to brand the tissue. The physiological replacement of cells varies substantially among different epithelia. The epithelium of the intestine completely self-renews within effectually 5 days. Past contrast, interfollicular epidermis takes approximately 4 weeks to renew, whereas the lung epithelium can take as long every bit six months to be replaced. In add-on some epithelia, such as pilus follicles, present a cyclic manner of cell replacement [ane]. Similarly, the mammary gland proceeds through cycles of growth and degeneration during and following pregnancy [one]. In improver, stalk cells are critically involved in regeneration upon wounding. Unless the epithelial stalk/progenitor cells are permanently damaged, most epithelia are able to repair their tissues following injury [1]; when epithelial stalk cells are depleted, fibrotic responses occur [1].

Practise renal stalk/progenitor cells exist in the adult kidney?

The understanding of kidney repair is withal in its infancy despite the rapid advances fabricated in recent years. The kidney is one of the few organs that undergo mesenchymal-epithelial transition during development [ii]. Moreover, structures present in the adult kidney arise from reciprocal interactions between ii detached embryonic appendages, namely the ureteric bud (UB) and metanephric mesenchyme (MM) [2]. The developed kidney contains more than 24 mature cell types arranged in distinct vascular, interstitial, glomerular and tubular compartments [two]. This unique organogenesis and structural complexity of the adult kidney has presented many challenges to the identification and characterization of kidney stem cells [2,3]. Attempts to identify adult kidney stem cells were made on the basis of the broad principles of stalk cell biological science, such equally prolonged jail cell-cycling fourth dimension (label-retaining cells), ability to extrude Hoechst dye (side population cells), by restrictive jail cell culture conditions, or past using markers expressed by other stalk cells or developing kidney [3]. Existing prove strongly supports the presence of stem cells in the developed kidney. Indeed, remission of affliction and regression of renal lesions have been observed in experimental animals and even in humans [4]. Identification and knowledge of renal stem jail cell biology might help to unlock latent regenerative pathways in human kidney, which would have the potential to change medical practice as much equally the introduction of dialysis did in the twentieth century.

The origin of the cells that supercede injured tubular epithelia is non known [3,5], although several lines of evidence suggest an intrarenal source [6,seven]. Recently, putative adult kidney stem cells accept been isolated, with some show indicating that they may enable epithelial repair after injury [eight-xviii]. Several studies have suggested the existence of an interstitial renal stem cell. Ane fashion of looking for stalk cells in solid organs was a pulse of bromodeoxyuridine (BrdU) followed by a long chase menses. The quiescent stem cells, which do not divide, maintain the high levels of BrdU deposited in their genomes, whereas the dividing, more differentiated stem cells steadily dilute the BrdU incorporated into their genomes as they proliferate. Maeshima et al. [12,xiii] identified BrdU-labelled cells, which they termed renal progenitor-similar tubular cells, in the renal tubules. Oliver et al. [fourteen] identified a population of BrdU-label-retaining cells inside the interstitium of renal papilla in the rat kidney. However, the use of BrdU labelling does not seem to be a specific method for identification of stem cells [sixteen].

Other studies have identified a rare population of developed interstitial cells in rat or human being kidney [8,9], and such cells have been proposed to engraft into tubules of either developing or injured kidney tissue [8,9], suggesting that extratubular cells can traverse the basement membrane and contribute to epithelium [14]. Although intriguing, this hypothesis has recently been questioned by a study past Humphreys et al. who have developed a method for distinguishing the source of kidney tubular regeneration based on studies in transgenic rodents for the homeodomain transcriptional regulator Six2 [xviii]. In this model, the Six2 promoter drives a fusion protein of green fluorescent protein (GFP) and Cre recombinase, which is expressed transiently in renal epithelial precursors during the developmental period of active nephrogenesis [18]. GFPCre expression is not present in the adult, and this expression is non observed after injury. When Six2-GFPCre mice are crossed with a floxed STOP reporter strain, Cre-dependent removal of the stop sequence in progeny leads to constitutive and heritable expression of a marker cistron such that all mesenchyme-derived renal epithelial cells, from the Bowman'due south capsule to the junction of the connecting segment and collecting duct, are heritably labelled [xviii]. In contrast, the unabridged interstitial compartment is unlabelled [eighteen]. Thus, the maintenance of labelled tubules mail service-injury would support a model of epithelial tubule repair due to surviving tubular epithelial cells or stalk/progenitor cells localized inside the labelled nephron, while label dilution would implicate an unlabelled, interstitial stem cell in the repair process. These findings indicate that papillary interstitial cells [14] or other types of interstitial stem/progenitor cells [viii,9] do not direct contribute to renal epithelial cells regeneration. These observations likewise betoken that repair of injured nephrons is predominantly accomplished past intrinsic, surviving tubular epithelial cells or a subset of stem/progenitor cells localized inside the nephron [18].

Recent studies identified a subset of renal stem/progenitor cells inside the Bowman's capsule of adult homo kidney [15]. These renal progenitors were identified through the cess of the presence of both CD24, a surface molecule that has been used to place different types of human stalk cells [19,xx] and CD133, a marker of several types of adult tissue stem cells [21,22]. The results showed that both markers were co-expressed past a subset of parietal epithelial cells selectively localized at the urinary pole of the Bowman'southward capsule [xv] (Figure 1). Once isolated, CD24+CD133+ renal progenitors were found to lack lineage-specific markers; to express transcription factors that are characteristic of multipotent stem cells, and to exhibit self-renewal, high clonogenic efficiency, and multidifferentiation potential [15]. When injected intravenously in SCID mice that had acute renal failure (ARF), CD24+CD133+ renal progenitors regenerated tubular structures in different portions of the nephron and as well reduced the morphological and functional kidney harm [fifteen]. The identification of CD24+CD133+ renal progenitors is in agreement with results obtained in transgenic mice, which suggests that endogenous cells of the nephron are responsible for repair of injured tubular epithelium [18], and allows the hypothesis that, from the urinary pole of the Bowman's capsule, CD24+CD133+ renal progenitors might initiate the replacement and regeneration of tubular epithelial cells in adult human being kidney [15] (Figure 2).

An external file that holds a picture, illustration, etc.  Object name is 1755-1536-2-3-1.jpg

CD24+CD133+ renal progenitors localize at the urinary pole of the Bowman'south capsule in adult human kidneys. (A) Triple-label immunofluorescence for CD133, (green), CD24 (red) and β1 integrin (blue) showing that in a mature glomerulus, co-expression of CD133 and CD24 characterizes a subset of cells in the Bowman's capsule (white) localized at the urinary pole (Upwards). AA = afferent arteriola. Objective 20×. (B) High power magnification of a triple-label immunofluorescence for CD133, (green), CD24 (red) and CD106 (blue) in a subset of cells in the Bowman's sheathing (white). Sections were stained as previously reported [15].

An external file that holds a picture, illustration, etc.  Object name is 1755-1536-2-3-2.jpg

Hypothetical diagram for kidney regeneration by dissimilar types of renal and extrarenal progenitors. CD24+CD133+ renal progenitors (ruby) are localized at the urinary pole and are in close contiguity with podocytes (green) at one extremity (the vascular stalk) and with tubular renal cells (xanthous) at the other extremity. A transitional cell population (red/dark-green) displays features of either renal progenitors (ruddy) or podocytes (green) and localizes between the urinary pole and the vascular pole. At the vascular stalk of the glomerulus, the transitional cells are localized in shut continuity with cells that lack progenitor markers, only exhibit the podocyte markers and the phenotypic features of differentiated podocytes (green). On the opposite side, at the urinary pole, transitional cells (scarlet/xanthous) with a mixed phenotype between tubular cells (yellow) and progenitor cells (red). The directions of differentiation is indicated by the arrows (modified from [24]).

Is there a unique renal progenitor common to renal evolution and the repair of adult kidney?

Regenerative biological science draws on the understanding of normal developmental processes. It is generally believed that developed stem/progenitor cells stand for a rest population straight derived from the organ-specific embryonic progenitor that is involved in organogenesis during fetal life [23-25]. This prompted united states to evaluate whether co-expression of CD133 and CD24 might be useful to track down multipotent kidney stem/progenitor cells during human being embryonic life.

Nephrons, the basic functional units of the kidney, are generated repetitively during kidney organogenesis from a mesenchymal progenitor population. Development of the mature mammalian kidney results from reciprocal signalling betwixt the branching UB tips and the undifferentiated MM. This leads to the assemblage and condensation of renal epithelia progenitor cells to grade the renal vesicle, which so undergoes transformation in the S-shaped body [2]. At this phase, the proximal end of the S-shaped body becomes invaded past claret vessels, differentiates into podocytes and parietal epithelial cells, and then generates glomeruli. Simultaneously, the middle and the distal segments of the S-shaped body brainstorm to express proteins that are characteristic of tubular epithelia [2]. The existence of renal embryonic progenitors in the MM is supported by the ascertainment that some MM-derived cells display multidifferentiation potential [two,26-28]. Accordingly, in embryonic human kidneys, co-expression of CD133 and CD24 characterizes a subset of cells in the cap mesenchyme, renal vesicles and S-shaped bodies that brandish self-renewal and multidifferentiation potential (Figure 1).

Interestingly, during nephron development, co-expression of CD133 and CD24 remained selectively localized to cells of the urinary pole of the Bowman's capsule [28]. Accordingly, CD24+CD133+ renal embryonic progenitors progressively decreased during gestation and represented < ii% of whole cells in developed kidneys [15,28]. When injected into mice with ARF, CD24+CD133+ renal embryonic progenitors regenerated cells of different portions of the nephron, reduced tissue necrosis and fibrosis, and significantly improved renal part [28]. In agreement with the putative nature of stem/progenitor cells, CD24+CD133+ renal embryonic progenitors expressed high levels of the stem cell-specific transcription factor BmI-one [29,30] and could generate tubular cells of different portions of the nephron in a model of acute tubular necrosis [28]. The existence of a putative MM cell with stem jail cell properties was already suggested past studies performed by the group of Reisner [31], who demonstrated that functioning renal tissue tin be reconstituted by MM derived from kidneys of 8 weeks of gestation [31]. Indeed, fetal kidney tissue obtained from 10 to 14 weeks of gestation maintains the property to generate de novo functional nephrons, only generates a smaller number of mature glomeruli and tubuli than kidneys of 8 weeks of gestation [31]. Accordingly, CD24+CD133+ renal embryonic progenitors are enriched in kidneys of 8 to 9 weeks of gestation, essentially subtract during 10 to 14 weeks of gestation, and represent < 2% of whole renal cells in adults.

Information technology is interesting that in both fetal and adult kidney, CD24+CD133+ progenitors persist as parietal epithelial cells localized at the urinary pole of the Bowman'southward capsule, supporting the concept that CD24+CD133+ progenitors might correspond a subpopulation of renal embryonic progenitors preserved from the early on stages of nephrogenesis, and the urinary pole of the Bowman's sheathing may stand for stem cell niche, which is a specific site in adult tissues where stem cells reside (Figure 1) [32]. In agreement with this hypothesis, embryonic stem cells, in one case differentiated toward renal tubular cells, selectively migrated to the tubuloglomerular junction after injection into developing kidneys [17]. More recently, studies performed in transgenic rodents for the homeodomain transcriptional regulator Six2 have confirmed the being in the cap mesenchyme of a multipotent nephron progenitor population [33]. Indeed, Six2-expressing cells give rising to all cell types of the main body of the nephron during all stages of nephrogenesis [34]. Pulse labeling of Six2-expressing nephron progenitors at the onset of kidney development suggests that the Six2-expressing population is maintained past self-renewal and is multipotent, generating the multiple domains of the whole cortical nephron [34]. Notably, descendants of a Six2+ cell tin exist establish inside molecularly singled-out compartments of a single nephron – podocytes, proximal and distal tubule structures – farther confirming that a single multipotent progenitor is the source of both the glomerular and tubular epithelial cells that plant the adult nephron [34].

Do bone marrow-derived cells contribute to tissue repair?

Some studies accept suggested that cells from os marrow might possess a surprising caste of plasticity and could differentiate into cell types of multiple organs of the trunk [33,35-38]. Accordingly, information technology was claimed that os marrow-derived stem cells (BMSC) could contribute to the generation of new epithelial cells in functionally of import numbers after kidney injury [39,40]. In light of their ease of accessibility, BMSC seem to exist a very strong candidate for the handling of renal diseases. Several subsequent studies have examined this possibility, with contrasting results [40-60].

Sugimoto and colleagues demonstrated that in a mouse model for Alport syndrome, os marrow cells contribute to the emergence of viable podocytes which are associated with the product of new basement membrane [47]. In addition, unfractionated BMSC tin can differentiate into endothelial and mesangial cells in a model of progressive glomerulosclerosis [42,44] and, more surprisingly, they can form new tubular epithelial cells in functionally important numbers after kidney injury [43]. It has recently become articulate that BMSC might fuse with differentiated cells in various developed organs, further complicating the estimation of marrow transplantation studies [61]. Held and co-workers accept shown that cell fusion could exist induced between os marrow-derived cells and renal tubular cells under weather condition of chronic renal damage [60], apparently without impairment of cell partitioning or conferment of genetic instability [44,62]. Additional works by several groups have shown that tubular cell replacement with BMSC is much lower than originally reported, calling into question the concept that BMSC physiologically participate in the repair of kidney injury [47-57]. Importantly, the low charge per unit of functional improvement observed when using unfractionated BMSC suggests that in acute tubular injury, regenerating cells originated from intrarenal cells [6,7]. Accordingly, the absenteeism of label dilution in Six2-GFPCre mice after injury and repair confirms that bone marrow-derived cells exercise non directly contribute to repair of tubular epithelial cells [18]. Taken together, these findings provide strong evidence against a physiological function for BMSC-derived cells in regeneration of mail service-ischaemic tubules by direct replacement of epithelial cells.

However, several studies indicate that mesangial cells might originate from a component of the hematopoietic lineages [62-65], and that BMSC might largely contribute to the regeneration of mesangial cells. Imasawa et al. [44] demonstrated the involvement of os marrow-derived cells in normal mesangial cell turnover. Lethally irradiated mice given transplants of T-cell-depleted bone marrow cells from syngeneic donor transgenic for GFP manifested a fourth dimension-dependent increase in GFP-positive cells in their glomeruli. When isolated and cultured, these cells stained positive for the mesangial jail cell marker desmin and the cells contracted in response to angiotensin Ii (Ang II), confirming that bone marrow-derived cells have the potential to differentiate into glomerular mesangial cells. Similar experiments with mice transplanted with purified clonally expanded hematopoietic progenitor cells were carried out past Masuya et al. [62] to ostend the hematopoietic origin of bone marrow-derived mesangial cells.

Finally, several studies have provided evidence that circulating endothelial progenitor cells (EPC) may contribute to glomerular capillary repair. In rat hematopoietic chimeras, depression levels of bone marrow-derived cells staining for the rat endothelial cell antigen RECA-i [66] were observed and the number of these cells gradually increased over time, suggesting that EPC contribute to normal physiological glomerular endothelial jail cell turnover. Following anti-Thy-1.ane-induced glomerular injury the authors observed a fourfold increase in os marrow-derived endothelial cells in the glomeruli [66]. These data indicate that glomerular repair cannot just be attributed to migration and proliferation of resident endothelial cells just it also involves os marrow-derived cells.

Participation of circulating EPC in renal regeneration has as well been demonstrated in human being adults. Williams and Alvarez [67] were the first to depict the presence of acceptor endothelial cells in kidney allografts. Lagaaij et al. [68] reported that in human being renal transplants the extent of replacement of donor endothelial cells lining the peritubular capillaries past those of the acceptor was related to the severity of vascular injury. They suggested that this endothelial replacement could be explained past the interest of acceptor-derived EPC. Recently, male, donor-derived endothelial cells were observed in the renal macrovasculature of a female person patient who adult thrombotic microangiopathy after gender-mismatched bone marrow transplantation [69]. Taken together, these observations confirm a role for BMSC in maintenance and repair of renal mesangium and endothelium, just non of the epithelial components of renal tissue.

Towards the understanding of renal tissue regeneration

Tissue stem cells can class various lineages in response to physiological stimuli or injuries, a holding that has great potential for regenerative medicine approaches. All the same, in many cases repair of epithelial cells does non depend on cells generated from multipotent stem cells, just directly derives from the migration of epithelial cells from the neighbouring epithelia, as previously reported likewise for the skin [70,71]. Indeed, genetic analyses suggest that the tubular epithelium can be self-renewing later astute kidney injury. Interestingly however, several previous studies take demonstrated that the proximal tubule arises at a diverseness of angles from Bowman's capsule and that at to the lowest degree one office of the tubuloglomerular junction has an area of intermediate appearance, with prominent microvilli on parietal cells in humans, mammals and fish. The finding of intermediate cells, especially in growing animals, suggests that parietal epithelium may exist able to change to tubular and that this might peculiarly occur as the kidney grows, during severe renal disorders [72-75], following unilateral nephrectomy [76] or during ageing [77]. Thus, renal stalk/progenitor cells might contribute to tubular epithelium repair, but this probably occurs merely when a wound cannot spontaneously repair itself through the migration of neighbouring undamaged tubular cells (Figure 2).

Indeed, mice affected past rhabdomyolysis-induced acute tubular necrosis spontaneously recover from acute kidney injury, but mice undergoing early on handling with man renal progenitors prove a complete recovery of renal function and kidney tissue integrity that was non observed in mice treated with saline [15] and, more importantly, a meaning reduction of the severity of ARF, as revealed by the consistently lower blood urea nitrogen levels and extended areas of tubular tissue regenerated by homo renal progenitors that co-expressed markers of proximal and distal tubules. This suggests that CD24+CD133+ progenitors tin regenerate tubular cells of different portions of the nephron, in vitro and in vivo [78,79]. However, the most important goal of regenerative medicine in the kidney is regeneration of glomerular injury, since glomerular diseases together business relationship for 90% of terminate-stage kidney illness (ESKD).

Recent insights accept defined a unified concept of glomerular diseases in which podocyte injury or loss is a common determining factor, which suggests the need for rational clinical efforts to allow podocyte regeneration [80-82]. Mature podocytes are postal service-mitotic cells, which tin undergo Dna synthesis to a express degree but do not proliferate, because they arrest in the G2/Chiliad phase of the prison cell bicycle [80-82]. Notwithstanding, in nigh adult epithelia, replacement of damaged or dead cells is maintained through the presence of stalk/progenitor cells [1]. Unless the epithelial stem/progenitor cells are permanently damaged, near epithelia are able to repair their tissues following injuries [ane]. Although glomerular disorders represent the most prominent cause of ESKD, remission of the affliction and regression of renal lesions have been observed in experimental animals and even in humans [iii]. This shows that remodelling of glomerular compages is possible, which would imply regeneration of the injured podocytes and reconstitution of the glomerular tuft. The inability of the podocyte to proliferate and supervene upon injured cells suggests the existence of potential stem/progenitor cells within the adult glomerulus. Interestingly, CD24+CD133+ renal progenitors are physically located within the Bowman's capsule, the merely place in the kidney which appears to be contiguous with both tubular cells and glomerular podocytes [fifteen]. Previous studies accept suggested the beingness of transitional cells exhibiting a mixed phenotype between the parietal epithelial cells and the podocyte at the vascular pole of the glomerulus [83]. In improver, CD24+CD133+ renal progenitors represent mutual progenitors of tubular cells and podocytes during renal development [29]. Accordingly, recent studies performed in our laboratory propose that CD24+CD133+ renal progenitors tin can also regenerate glomerular podocytes in mice affected by adriamycin nephropathy, and can reduce the severity of proteinuria and of glomerular injury [84]. These results suggest that CD24+CD133+ renal progenitors can too replace and regenerate podocytes through their division and migration forth the Bowman's sheathing towards the glomerular tuft during developed life or in response to podocyte injury (Figure 2).

The response to renal injury: from regeneration to fibrosis

In humans, problems with wound healing can manifest as either delayed wound healing (which occurs with diabetes or radiation exposure) or excessive healing (every bit occurs with hypertrophic and keloid scars). Excessive healing is characterized past the deposition of large amounts of extracellular matrix and by alterations in local vascularization and prison cell proliferation. These excessive fibrotic reactions manifest in humans every bit a 'bad scar'. These commonly occur later on major injuries such every bit burns, in which case they are referred to as hypertrophic scars. They can also appear for unknown reasons after a relatively pocket-size trauma, as is the case for keloid scars, which might accept a genetic ground [85-87]. Chronic infections, toxic and metabolic injuries, and idiopathic inflammatory diseases can promote the development of a scar, leading to tissue fibrosis [85]. In many cases, patients with progressive fibrosis have a poor prognosis and oftentimes require organ transplantation [85-87].

Although fibrosis is a role of the normal pathophysiological response to injury in many tissues, extended exposure to chronic injury results in tissue fibrosis, massive deposition of extracellular matrix, scar formation, and organ failure [87]. Chronic injury causes dysfunction of the tubular epithelial cells, which triggers release of fibrogenic cytokines and recruitment of inflammatory cells to injured kidneys [88-91]. Over the years, the primary focus of tubulointerstitial fibrosis studies has been on interstitial fibroblasts and infiltrated mononuclear cells for obvious reasons [92-94]. However, fibroblast activation after injury is, in essence, a wound-healing response by which the injured kidney attempts to repair and recover from the injury. Therefore, fibroblast activation at most may be necessary, but certainly not sufficient, for development of a full-calibration of renal interstitial fibrosis.

Fibroblasts contribute to 50% of all collagen-expressing cells in the course of renal fibrosis. Renal cortical fibroblasts maintain a quiescent country in normal kidneys, but in response to injury they proliferate and activate into myofibroblasts. Fibroblasts are not especially abundant in normal kidneys as they are in lungs, lymph nodes, and spleen. When renal fibrogenesis sets in, about 36% of new fibroblasts come up from the local epithelial mesenchymal transition (EMT), about 14–xv% from the bone marrow, and the rest from local proliferation [94]. Endothelial cells also contribute to the emergence of fibroblasts during kidney fibrosis via the procedure of EMT, every bit recently demonstrated in mouse models of unilateral ureteral obstructive nephropathy, diabetic nephropathy, and Alport renal illness [95]. Although local activation of the renin-angiotensin system (RAS) and specifically Ang Ii affects all parenchymal organs, its effect is more pronounced in renal fibrosis. RAS stimulates inflammation, including the expression of cytokines, chemokines, growth factors, and reactive oxygen species [94,96]. Ang II induces vascular inflammation, endothelial dysfunction, up-regulation of adhesion molecules, and recruitment of infiltrating cells into the kidney (Figure three) [94,96].

An external file that holds a picture, illustration, etc.  Object name is 1755-1536-2-3-3.jpg

Hypothetical diagram for kidney fibrosis. Chronic injury causes dysfunction of the tubular epithelial cells, which triggers release of fibrogenic cytokines and recruitment of inflammatory cells to injured kidneys. Myelo-monocytic cells recruited from the bone marrow produce TGF-β1 in injured kidneys. In turn, TGF-β1 induces activation of collagen-producing cells, which more often than not arise from kidney resident cells through epithelial-mesenchymal transition (modified [24]).

In improver, myelo-monocytic cells recruited from the bone marrow produce TGF-β1 in injured kidneys. In plough, TGF-β1 induces activation of collagen-producing cells, which mostly arise from kidney resident cells. The potential part of tubular epithelial cells in renal fibrosis is often concealed, partly because no straight connection seems to be between tubular cells and the production and degradation of extracellular matrix, a hallmark of interstitial fibrosis. Nonetheless, molecular analyses of factor expression have constantly underlined the potential importance of tubular epithelia in the fibrotic process. For instance, while it is well known that TGF-β1 expression is increased in about all of the chronic kidney disease models studied, the expression of TGF-β receptors, which determine the specificity of TGF-β action, is frequently up-regulated predominantly in renal tubular epithelium [97], indicating that tubular epithelial cells are the in vivo natural targets of this pro-fibrotic cytokine. Hence, EMT helps to reconcile the disparity between molecular analysis and pathological findings in fibrotic kidney. Recently, pericytes were too identified as a major source of interstitial myofibroblasts in the fibrotic kidney, suggesting that either vascular injury or vascular factors are the most likely triggers for pericyte migration and differentiation into myofibroblasts (Figure 3) [98].

The aggregating of fibroblasts and an excess of collagen and other matrix components at sites of chronic inflammation pb to scar tissue germination and progressive tissue injury. These fibroblasts derive from the bone marrow, just also ascend from an EMT of cells at injury sites [88,89]. EMT is probable to be involved in the progressive fibrotic diseases of the heart, lung, liver, and kidney, and genetic models provide indisputable evidence for a crucial role for EMT in renal fibrogenesis.

Conclusion

Reversal of renal fibrosis is possible, equally observed in experimental animals and fifty-fifty in humans [4]. However, whether fibrotic kidneys tin can reverse to normal renal compages remains unresolved, and the point of no render in the development of irreversible renal fibrosis still remains to be determined. The rapid interposition of scar tissue probably confers a survival advantage by preventing infectious microorganisms from invading the wound and past inhibiting the continued mechanical deformation of larger tissues (a process that could compound the initial insult), but prevents subsequent tissue regeneration. All the same, the existence of renal epithelial progenitors in the kidney suggests a possible explanation for the regression of renal lesions, and indicates that a manipulation of the wound repair process in society to shift information technology towards regeneration [99] volition probably require the power to slow the rapid fibrotic response so that renal progenitor cells tin can regenerate functional tissue and avoid scar formation.

Abbreviations

Ang II: angiotensin II; ARF: acute renal failure; BMSC: os marrow-derived stem cells; BrdU: bromodeoxyuridine; EMT: epithelial mesenchymal transition; EPC: endothelial progenitor cells; ESKD: end phase kidney disease; GFP: green fluorescent protein; MM: metanephric mesenchyme; RAS: renin-angiotensin system; UB: ureteric bud

Competing interests

The authors declare that they have no competing interests.

Acknowledgements

The inquiry leading to these results has received funding from the European Enquiry Council Starting Grant under the European Customs's Seventh Framework Program (FP7/2007-2013), ERC grant number 205027. This study was besides supported by the Tuscany Ministry of Wellness and the Associazione Italiana per la Ricerca sul Cancro.

References

  • Blanpain C, Horsley V, Fuchs E. Epithelial stem cells: turning over new leaves. Cell. 2007;128:445–458. doi: 10.1016/j.cell.2007.01.014. [PMC free commodity] [PubMed] [CrossRef] [Google Scholar]
  • Dressler GR. The cellular basis of kidney development. Annu Rev Cell Dev Biol. 2006;22:509–529. doi: ten.1146/annurev.cellbio.22.010305.104340. [PubMed] [CrossRef] [Google Scholar]
  • Niggling MH. Regrow or repair: potential regenerative therapies for the kidney. J Am Soc Nephrol. 2006;17:2390–2401. doi: ten.1681/ASN.2006030218. [PubMed] [CrossRef] [Google Scholar]
  • Remuzzi G, Benigni A, Remuzzi A. Mechanisms of progression and regression of renal lesions of chronic nephropathies and diabetes. J Clin Invest. 2006;116:288–296. doi: x.1172/JCI27699. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Imai E, Iwatani H. The continuing story of renal repair with stem cells. J Am Soc Nephrol. 2007;18:2423–2424. doi: 10.1681/ASN.2007070769. [PubMed] [CrossRef] [Google Scholar]
  • Lin F, Moran A, Igarashi P. Intrarenal cells, non bone marrow-derived cells, are the major source for regeneration in post ischemic kidney. J Clin Invest. 2005;115:1756–1764. doi: 10.1172/JCI23015. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Duffield JS, Park KM, Hsiao LL, Kelley VR, Scadden DT, Ichimura T, Bonventre JV. Restoration of tubular epithelial cells during repair of the postischemic kidney occurs independently of os marrow-derived stalk cells. J Clin Invest. 2005;115:1743–1755. doi: 10.1172/JCI22593. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Bussolati B, Bruno S, Grange C, Buttiglieri Due south, Deregibus MC, Cantino D, Camussi G. Isolation of renal progenitor cells from developed human kidney. Am J Pathol. 2005;166:545–555. [PMC free article] [PubMed] [Google Scholar]
  • Dekel B, Zangi 50, Shezen E, Reich-Zeliger S, Eventov-Friedman S, Katchman H, Jacob-Hirsch J, Amariglio N, Rechavi G, Margalit R, Reisner Y. Isolation and characterization of nontubular sca-one+lin_multipotent stem/progenitor cells from adult mouse kidney. J Am Soc Nephrol. 2006;17:3300–3314. doi: x.1681/ASN.2005020195. [PubMed] [CrossRef] [Google Scholar]
  • Gupta S, Verfaillie C, Chmielewski D, Kren S, Eidman K, Connaire J, Heremans Y, Lund T, Blackstad M, Jiang Y, Luttun A, Rosenberg ME. Isolation and characterization of kidney-derived stem cells. J Am Soc Nephrol. 2006;17:3028–3040. doi: 10.1681/ASN.2006030275. [PubMed] [CrossRef] [Google Scholar]
  • Kitamura S, Yamasaki Y, Kinomura M, Sugaya T, Sugiyama H, Maeshima Y, Makino H. Establishment and characterization of renal progenitor like cells from S3 segment of nephron in rat adult kidney. FASEB J. 2005;nineteen:1789–1797. doi: ten.1096/fj.05-3942com. [PubMed] [CrossRef] [Google Scholar]
  • Maeshima A, Yamashita Due south, Nojima Y. Identification of renal progenitor-like tubular cells that participate in the regeneration processes of the kidney. J Am Soc Nephrol. 2003;14:3138–3146. doi: 10.1097/01.ASN.0000098685.43700.28. [PubMed] [CrossRef] [Google Scholar]
  • Maeshima A, Sakurai H, Nigam SK. Developed kidney tubular cell population showing phenotypic plasticity, tubulogenic capacity, and integration capability into developing kidney. J Am Soc Nephrol. 2006;17:188–198. doi: 10.1681/ASN.2005040370. [PubMed] [CrossRef] [Google Scholar]
  • 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. [PMC free article] [PubMed] [Google Scholar]
  • Sagrinati C, Netti GS, Mazzinghi B, Lazzeri E, Liotta F, Frosali F, Ronconi Eastward, Meini C, Gacci G, Squecco R, Carini M, Gesualdo L, Francini F, Maggi E, Annunziato F, Lasagni L, Serio M, Romagnani S, Romagnani P. Isolation and label of multipotent progenitor cells from the Bowman's capsule of developed man kidneys. J Am Soc Nephrol. 2006;17:2443–2456. doi: 10.1681/ASN.2006010089. [PubMed] [CrossRef] [Google Scholar]
  • Kiel MJ, He S, Ashkenazi R, Gentry SN, Teta Thousand, Kushner JA, Jackson TL, Morrison SJ. Haematopoietic stem cells do not asymmetrically segregate chromosomes or retain BrdU. Nature. 2007;449:238–242. doi: 10.1038/nature06115. [PMC costless commodity] [PubMed] [CrossRef] [Google Scholar]
  • Kim D, Dressler GR. Nephrogenic factors promote differentiation of mouse embryonic stem cells into renal epithelia. J Am Soc Nephrol. 2005;xvi:3527–3534. doi: 10.1681/ASN.2005050544. [PubMed] [CrossRef] [Google Scholar]
  • Humphreys BD, Valerius MT, Kobayashi A, Mugford JW, Soeung S, Duffield JS, McMahon AP, Bonventre JV. Intrinsic epithelial cells repair the kidney after injury. Prison cell Stem Jail cell. 2008;2:284–291. doi: x.1016/j.stalk.2008.01.014. [PubMed] [CrossRef] [Google Scholar]
  • Shackleton M, Vaillant F, Simpson KJ, Stingl J, Smyth GK, Asselin-Labat ML, Wu Fifty, Lindeman GJ, Visvader JE. Generation of a functional mammary gland from a unmarried stalk jail cell. Nature. 2006;439:84–88. doi: 10.1038/nature04372. [PubMed] [CrossRef] [Google Scholar]
  • Kubota H, Avarbock MR, Brinster RL. Spermatogonial stalk cells share some, but non all, phenotypic and functional characteristics with other stalk cells. Proc Natl Acad Sci USA. 2003;100:6487–6492. doi: ten.1073/pnas.0631767100. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Coskun V, Wu H, Blanchi B, Tsao S, Kim Grand, Zhao J, Biancotti JC, Hutnick L, Krueger RC, Jr, Fan 1000, de Vellis J, Sunday YE. CD133+ neural stem cells in the ependyma of mammalian postnatal forebrain. Proc Natl Acad Sci USA. 2008;105:1026–1031. doi: 10.1073/pnas.0710000105. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Shmelkov SV, St Clair R, Lyden D, Rafii S. AC133/CD133/Prominin-one. Int J Biochem Prison cell Biol. 2005;37:715–719. doi: 10.1016/j.biocel.2004.08.010. [PubMed] [CrossRef] [Google Scholar]
  • Temple S. The evolution of neural stem cells. Nature. 2001;414:112–117. doi: 10.1038/35102174. [PubMed] [CrossRef] [Google Scholar]
  • Morrison SJ, Kimble J. Asymmetric and symmetric stem-prison cell divisions in development and cancer. Nature. 2006;441:1068–1074. doi: x.1038/nature04956. [PubMed] [CrossRef] [Google Scholar]
  • Self Grand, Lagutin OV, Bowling B, Hendrix J, Cai Y, Dressler GR, Oliver Grand. Six2 is required for suppression of nephrogenesis and progenitor renewal in the developing kidney. EMBO J. 2006;25:5214–5228. doi: 10.1038/sj.emboj.7601381. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Oliver JA, Barasch J, Yang J, Herzlinger D, Al-Awqati Q. Metanephric mesenchyme contains embryonic renal stem cells. Am J Physiol Renal Physiol. 2002;283:F799–F809. [PubMed] [Google Scholar]
  • Herzlinger D, Koseki C, Mikawa T, Al-Awqati Q. Metanephric mesenchyme contains multipotent stem cells whose fate is restricted later consecration. Evolution. 1992;114:565–572. [PubMed] [Google Scholar]
  • Lazzeri E, Crescioli C, Ronconi Eastward, Mazzinghi B, Sagrinati C, Netti GS, Angelotti ML, Parente E, Ballerini L, Cosmi Fifty, Maggi L, Gesualdo L, Rotondi M, Annunziato F, Maggi East, Lasagni L, Serio G, Romagnani Due south, Vannelli GB, Romagnani P. Regenerative potential of embryonic renal multipotent progenitors in astute renal failure. J Am Soc Nephrol. 2007;18:3128–3138. doi: 10.1681/ASN.2007020210. [PubMed] [CrossRef] [Google Scholar]
  • Molofsky AV, Pardal R, Iwashita T, Park IK, Clarke MF, Morrison SJ. BmI-1 dependence distinguishes neural stem cell self-renewal from progenitor proliferation. Nature. 2003;425:962–967. doi: x.1038/nature02060. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kozakowski N, Soleiman A, Pammer J. BMI-1 expression is inversely correlated with the grading of renal clear cell carcinoma. Pathol Oncol Res. 2008;fourteen:nine–xiii. doi: x.1007/s12253-008-9006-3. [PubMed] [CrossRef] [Google Scholar]
  • Dekel B, Burakova T, Arditti FD, Reich-Zeliger S, Milstein O, Aviel-Ronen S, Rechavi Yard, Friedman N, Kaminski N, Passwell JH, Reisner Y. Human and porcine early kidney precursors every bit a new source for transplantation. Nat Med. 2003;ix:53–60. doi: x.1038/nm812. [PubMed] [CrossRef] [Google Scholar]
  • Scadden DT. The stem-jail cell niche as an entity of activeness. Nature. 2006;441:1075–1079. doi: x.1038/nature04957. [PubMed] [CrossRef] [Google Scholar]
  • Krause DS, Theise ND, Collector MI, Henegariu O, Hwang South, Gardner R, Neutzel Southward, Sharkis SJ. Multi-organ, multi-lineage engraftment by a unmarried os marrow-derived stem prison cell. Cell. 2001;105:369–377. doi: 10.1016/S0092-8674(01)00328-2. [PubMed] [CrossRef] [Google Scholar]
  • Kobayashi A, Valerius MT, Mugford JW, Carroll TJ, Self One thousand, Oliver G, McMahon AP. Six2 defines and regulates a multipotent self-renewing nephron progenitor population throughout mammalian kidney development. Cell Stem Cell. 2008;3:169–81. doi: ten.1016/j.stem.2008.05.020. [PMC gratuitous article] [PubMed] [CrossRef] [Google Scholar]
  • Phinney DG, Prockop DJ. Concise review: mesenchymal stem/multipotent stromal cells: the land of transdifferentiation and modes of tissue repair – current views. Stem Cells. 2007;25:2896–2902. doi: 10.1634/stemcells.2007-0637. [PubMed] [CrossRef] [Google Scholar]
  • Spyridonidis A, Zeiser R, Follo M, Metaxas Y, Finke J. Stem prison cell plasticity: the debate begins to clarify. Stalk Prison cell Rev. 2005;1:37–43. doi: 10.1385/SCR:1:1:037. [PubMed] [CrossRef] [Google Scholar]
  • Lagasse Due east, Connors H, Al-Dhalimy One thousand, Reitsma M, Dohse M, Osborne L, Wang X, Finegold M, Weissman IL, Grompe M. Purified hematopoietic stem cells tin differentiate into hepatocytes in vivo. Nat Med. 2000;6:1229–1234. doi: 10.1038/81326. [PubMed] [CrossRef] [Google Scholar]
  • Herzog EL, Chai L, Krause DS. Plasticity of marrow-derived stem cells. Claret. 2003;102:3483–3493. doi: 10.1182/claret-2003-05-1664. [PubMed] [CrossRef] [Google Scholar]
  • Kale S, Karihaloo A, Clark PR, Kashgarian Thousand, Krause DS, Cantley LG. Os marrow stalk cells contribute to repair of the ischemically injured renal tubule. J Clin Invest. 2003;112:42–49. [PMC free commodity] [PubMed] [Google Scholar]
  • Poulsom R, Forbes SJ, Hodivala-Dilke M, Ryan E, Wyles S, Navaratnarasah S, Jeffery R, Hunt T, Alison Yard, Cook T, Pusey C, Wright NA. Bone marrow contributes to renal parenchymal turnover and regeneration. J Pathol. 2001;195:229–235. doi: 10.1002/path.976. [PubMed] [CrossRef] [Google Scholar]
  • Morigi M, Benigni A, Remuzzi Yard, Imberti B. The regenerative potential of stem cells in acute renal failure. Jail cell Transplant. 2006;fifteen:S111–S117. doi: 10.3727/000000006783982449. [PubMed] [CrossRef] [Google Scholar]
  • Ikarashi Grand, Li B, Suwa Thou, Kawamura K, Morioka T, Yao J, Khan F, Uchiyama M, Oite T. Bone marrow cells contribute to regeneration of damaged glomerular endothelial cells. Kidney Int. 2005;67:1925–1933. doi: x.1111/j.1523-1755.2005.00291.x. [PubMed] [CrossRef] [Google Scholar]
  • Lin F, Cordes One thousand, Li L, Hood L, Couser WG, Shankland SJ, Igarashi P. Hematopoietic stem cells contribute to the regeneration of renal tubules afterwards renal ischemia-reperfusion injury in mice. J Am Soc Nephrol. 2003;14:1188–1199. doi: ten.1097/01.ASN.0000061595.28546.A0. [PubMed] [CrossRef] [Google Scholar]
  • Imasawa T, Utsunomiya Y, Kawamura T, Zhong Y, Nagasawa R, Okabe Chiliad, Maruyama N, Hosoya T, Ohno T. The potential of os marrow-derived cells to differentiate to glomerular mesangial cells. J Am Soc Nephrol. 2001;12:1401–1409. [PubMed] [Google Scholar]
  • Kunter U, Rong S, Boor P, Eitner F, Müller-Newen G, Djuric Z, van Roeyen CR, Konieczny A, Ostendorf T, Villa Fifty, Milovanceva-Popovska K, Kerjaschki D, Floege J. Transplanted mesenchymal stem cells accelerate glomerular healing in experimental glomerulonephritis. J Am Soc Nephrol. 2006;17:2202–2212. doi: 10.1681/ASN.2005080815. [PubMed] [CrossRef] [Google Scholar]
  • Bi B, Schmitt R, Israilova 1000, Nishio H, Cantley LG. Stromal cells protect confronting astute tubular injury via an endocrine consequence. J Am Soc Nephrol. 2007;eighteen:2486–2496. doi: 10.1681/ASN.2007020140. [PubMed] [CrossRef] [Google Scholar]
  • Sugimoto H, Mundel TM, Sund M, Xie L, Cosgrove D, Kalluri R. Bone-marrow-derived stem cells repair basement membrane collagen defects and contrary genetic kidney affliction. Proc Natl Acad Sci U.s.. 2006;103:7321–7326. doi: x.1073/pnas.0601436103. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Morigi M, Imberti B, Zoja C, Corna D, Tomasoni Due south, Abbate One thousand, Rottoli D, Angioletti S, Benigni A, Perico N, Alison M, Remuzzi G. Mesenchymal stalk cells are renotropic, helping to repair the kidney and improve role in acute renal failure. J Am Soc Nephrol. 2004;xv:1794–1804. doi: x.1097/01.ASN.0000128974.07460.34. [PubMed] [CrossRef] [Google Scholar]
  • Tögel F, Hu Z, Weiss K, Isaac J, Lange C, Westenfelder C. Administered mesenchymal stem cells protect against ischemic astute renal failure through differentiation-independent mechanisms. Am J Physiol Renal Physiol. 2005;289:F31–F42. doi: 10.1152/ajprenal.00007.2005. [PubMed] [CrossRef] [Google Scholar]
  • Lange C, Tögel F, Ittrich H, Clayton F, Nolte-Ernsting C, Zander AR, Westenfelder C. Administered mesenchymal stalk cells enhance recovery from ischemia/reperfusion-induced acute renal failure in rats. Kidney Int. 2005;68:1613–1617. doi: 10.1111/j.1523-1755.2005.00573.x. [PubMed] [CrossRef] [Google Scholar]
  • Tögel F, Weiss 1000, Yang Y, Hu Z, Zhang P, Westenfelder C. Vasculotropic, paracrine actions of infused mesenchymal stem cells are of import to the recovery from acute kidney injury. Am J Physiol Renal Physiol. 2007;292:F1626–F1635. doi: ten.1152/ajprenal.00339.2006. [PubMed] [CrossRef] [Google Scholar]
  • Herrera MB, Bussolati B, Bruno S, Morando L, Mauriello-Romanazzi Thou, Sanavio F, Stamenkovic I, Biancone L, Camussi Thousand. Exogenous mesenchymal stem cells localize to the kidney by means of CD44 following acute tubular injury. Kidney Int. 2007;72:430–441. doi: 10.1038/sj.ki.5002334. [PubMed] [CrossRef] [Google Scholar]
  • Li B, Morioka T, Uchiyama M, Oite T. Bone marrow cell infusion ameliorates progressive glomerulosclerosis in an experimental rat model. Kidney Int . 2006;69:323–330. doi: 10.1038/sj.ki.5000083. [PubMed] [CrossRef] [Google Scholar]
  • Kunter U, Rong S, Boor P, Eitner F, Müller-Newen G, Djuric Z, van Roeyen CR, Konieczny A, Ostendorf T, Villa L, Milovanceva-Popovska K, Kerjaschki D, Floege J. Mesenchymal stalk cells foreclose progressive experimental renal failure but maldifferentiate into glomerular adipocytes. J Am Soc Nephrol. 2007;18:1754–1764. doi: 10.1681/ASN.2007010044. [PubMed] [CrossRef] [Google Scholar]
  • Imberti B, Morigi Chiliad, Tomasoni S, Rota C, Corna D, Longaretti 50, Rottoli D, Valsecchi F, Benigni A, Wang J, Abbate M, Zoja C, Remuzzi G. Insulin-similar growth gene-1 sustains stem cell mediated renal repair. J Am Soc Nephrol. 2007;18:2921–2928. doi: 10.1681/ASN.2006121318. [PubMed] [CrossRef] [Google Scholar]
  • Herrera MB, Bussolati B, Bruno S, Fonsato Five, Romanazzi GM, Camussi Yard. Mesenchymal stem cells contribute to the renal repair of acute tubular epithelial injury. Int J Mol Med. 2004;14:1035–1041. [PubMed] [Google Scholar]
  • Prodromidi EI, Poulsom R, Jeffery R, Roufosse CA, Pollard PJ, Pusey CD, Cook HT. Os marrow-derived cells contribute to podocyte regeneration and amelioration of renal disease in a mouse model of Alport syndrome. Stalk Cells. 2006;24:2448–2455. doi: x.1634/stemcells.2006-0201. [PubMed] [CrossRef] [Google Scholar]
  • Ninichuk V, Gross O, Segerer S, Hoffmann R, Radomska E, Buchstaller A, Huss R, Akis N, Schlöndorff D, Anders HJ. Multipotent mesenchymal stem cells reduce interstitial fibrosis but do not delay progression of chronic kidney illness in collagen4A3-deficient mice. Kidney Int. 2006;70:121–129. doi: x.1038/sj.ki.5001521. [PubMed] [CrossRef] [Google Scholar]
  • Guo JK, Schedl A, Krause DS. Os marrow transplantation can benumb the progression of mesangial sclerosis. Stem Cells. 2006;24:406–415. doi: ten.1634/stemcells.2005-0139. [PubMed] [CrossRef] [Google Scholar]
  • Held PK, Al-Dhalimy K, Willenbring H, Akkari Y, Jiang Southward, Torimaru Y, Olson S, Fleming WH, Finegold M, Grompe 1000. In vivo genetic selection of renal proximal tubules. Mol Ther. 2006;xiii:49–58. doi: 10.1016/j.ymthe.2005.09.004. [PubMed] [CrossRef] [Google Scholar]
  • Rizvi AZ, Beau JR, Davies PS, Bailey As, Decker Advertizement, Willenbring H, Grompe M, Fleming WH, Wong MH. Bone marrow-derived cells fuse with normal and transformed abdominal stalk cells. Proc Natl Acad Sci The states. 2006;103:6321–6325. doi: 10.1073/pnas.0508593103. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Masuya M, Drake CJ, Fleming PA, Reilly CM, Zeng H, Hill WD, Martin-Studdard A, Hess DC, Ogawa M. Hematopoietic origin of glomerular mesangial cells. Blood. 2003;101:2215–2218. doi: 10.1182/blood-2002-04-1076. [PubMed] [CrossRef] [Google Scholar]
  • Abe T, Fleming PA, Masuya M, Minamiguchi H, Ebihara Y, Drake CJ, Ogawaa M. Granulocyte/macrophage origin of glomerular mesangial cells. Int J Hematol. 2005;82:115–118. doi: 10.1532/IJH97.05018. [PubMed] [CrossRef] [Google Scholar]
  • Takeda South, Rogers SA, Hammerman MR. Differential origin for endothelial and mesangial cells subsequently transplantation of pig fetal renal primordia into rats. Transpl Immunol. 2006;15:211–215. doi: ten.1016/j.trim.2005.10.003. [PubMed] [CrossRef] [Google Scholar]
  • Rookmaaker MB, Verhaar MC, Van Zonneveld AJ, Rabelink TJ. Progenitor cells in the kidney: Biology and therapeutic perspectives. Kidney Int. 2004;66:518–522. doi: ten.1111/j.1523-1755.2004.761_10.x. [PubMed] [CrossRef] [Google Scholar]
  • Rookmaaker MB, Tolboom H, Goldschmeding R, Zwaginga JJ, Rabelink TJ, Verhaar MC. Bone marrow-derived cells contribute to glomerular endothelial repair in experimental glomerulonephritis. Am J Pathol. 2003;163:553–562. [PMC complimentary commodity] [PubMed] [Google Scholar]
  • Williams GM, Alvarez CA. Host repopulation of the endothelium in allografts of kidneys and aorta. Surg Forum. 1969;twenty:293–294. [PubMed] [Google Scholar]
  • Lagaaij EL, Cramer-Knijnenburg GF, van Kemenade FJ, van Es LA, Bruijn JA, van Krieken JH. Endothelial cell chimerism after renal transplantation and vascular rejection. Lancet. 2001;357:33–37. doi: 10.1016/S0140-6736(00)03569-viii. [PubMed] [CrossRef] [Google Scholar]
  • Rookmaaker MB, Tolboom H, Goldschmeding R, Zwaginga JJ, Rabelink TJ, Verhaar MC. Bone-marrow-derived cells contribute to endothelial repair after thrombotic microangiopathy. Claret. 2002;99:1095. doi: 10.1182/blood.V99.3.1095. [PubMed] [CrossRef] [Google Scholar]
  • Ito M, Liu Y, Yang Z, Nguyen J, Liang F, Morris RJ, Cotsarelis G. Stem cells in the hair follicle bulge contribute to wound repair but not to homeostasis of the epidermis. Nature Med. 2005;eleven:1351–1354. doi: 10.1038/nm1328. [PubMed] [CrossRef] [Google Scholar]
  • Levy 5, Lindon C, Harfe BD, Morgan BA. Distinct stem cell populations regenerate the follicle and interfollicular epidermis. Dev Prison cell. 2005;9:855–861. doi: 10.1016/j.devcel.2005.xi.003. [PubMed] [CrossRef] [Google Scholar]
  • Finckh ES, Joske RA. The occurrence of columnar epithelium in Bowman'south capsule. J Path Bact. 1954;68:646–648. doi: 10.1002/path.1700680243. [PubMed] [CrossRef] [Google Scholar]
  • Nachman RL. Metaplasia of parietal capsular epithelium of renal glomerulus. Arch Path. 1962;73:48–52. [PubMed] [Google Scholar]
  • Kanel GC, Peters RL. Glomerular tubular reflux – a morphologic renal lesion associated with the hepatorenal syndrome. Hepatology. 1984;iv:242–246. doi: x.1002/hep.1840040212. [PubMed] [CrossRef] [Google Scholar]
  • Valdes AJ, Zhang JM. Intraglomerular tubular epithelial cells. A mark of glomerular hematuria. Arch Pathol Lab Med. 1987;111:189–191. [PubMed] [Google Scholar]
  • Andrews PM. The presence of proximal tubule-like cells in the kidney parietal epithelium in response to unilateral nephrectomy. Anat Rec. 1981;200:61–65. doi: 10.1002/ar.1092000106. [PubMed] [CrossRef] [Google Scholar]
  • Castelletto L, Goya RG. Sexual activity-related incidence of tubular metaplasia in Bowman's capsule of aging rats. Virchows Curvation B Jail cell Pathol Incl Mol Pathol. 1990;59:79–82. doi: 10.1007/BF02899390. [PubMed] [CrossRef] [Google Scholar]
  • Mazzinghi B, Ronconi E, Lazzeri E, Sagrinati C, Ballerini L, Angelotti ML, Parente E, Mancina R, Netti GS, Becherucci F, Gacci M, Carini M, Gesualdo Fifty, Rotondi M, Maggi E, Lasagni L, Serio M, Romagnani Southward, Romagnani P. Essential but differential part for CXCR4 and CXCR7 in the therapeutic homing of human being renal progenitor cells. J Exp Med. 2008;205:479–490. doi: 10.1084/jem.20071903. [PMC gratis article] [PubMed] [CrossRef] [Google Scholar]
  • Sagrinati C, Ronconi E, Lazzeri E, Lasagni L, Romagnani P. Stalk-cell approaches for kidney repair: choosing the right cells. Trends Mol Med. 2008;14:277–285. doi: ten.1016/j.molmed.2008.05.005. [PubMed] [CrossRef] [Google Scholar]
  • Wiggins RC. The spectrum of podocytopathies: a unifying view of glomerular diseases. Kidney Int. 2007;71:1205–1214. doi: 10.1038/sj.ki.5002222. [PubMed] [CrossRef] [Google Scholar]
  • D'Agati VD. Podocyte injury in focal segmental glomerulosclerosis: Lessons from animal models (a play in 5 acts) Kidney Int. 2008;73:399–406. doi: 10.1038/sj.ki.5002655. [PubMed] [CrossRef] [Google Scholar]
  • Wharram BL, Goyal One thousand, Wiggins JE, Sanden SK, Hussain South, Filipiak WE, Saunders TL, Dysko RC, Kohno K, Hozman LB, Wiggins RC. Podocyte depletion causes glomerulosclerosis: diphtheria toxin-induced podocyte depletion in rats expressing human diphtheria toxin receptor transgene. J Am Soc Nephrol. 2005;16:2941–2952. doi: 10.1681/ASN.2005010055. [PubMed] [CrossRef] [Google Scholar]
  • Bariety J, Mandet C, Colina GS, Bruneval P. Parietal podocytes in normal human glomeruli. J Am Soc Nephrol. 2006;17:2770–2780. doi: x.1681/ASN.2006040325. [PubMed] [CrossRef] [Google Scholar]
  • Ronconi E, Sagrinati C, Angelotti ML, Lazzeri E, Mazzinghi B, Ballerini Fifty, Parente Eastward, Becherucci F, Gacci M, Carini One thousand, Maggi East, Serio M, Vannelli GB, Lasagni L, Romagnani Due south, Romagnani P. Regeneration of glomerular podocytes by human renal progenitors. J Am Soc Nephrol. 2009;xx:322–32. doi: ten.1681/ASN.2008070709. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Gurtner GC, Werner South, Barrandon Y, Longaker MT. Wound repair and regeneration. Nature. 2008;453:314–321. doi: 10.1038/nature07039. [PubMed] [CrossRef] [Google Scholar]
  • Vocaliser AJ, Clark RA. Cutaneous wound healing. Due north Engl J Med. 1999;341:738–746. doi: 10.1056/NEJM199909023411006. [PubMed] [CrossRef] [Google Scholar]
  • Aarabi S, Longaker MT, Gurtner GC. Hypertrophic scar germination following burns and trauma: new approaches to treatment. PLoS Med. 2007;4:e234. doi: 10.1371/journal.pmed.0040234. [PMC costless commodity] [PubMed] [CrossRef] [Google Scholar]
  • Neilson EG. Mechanisms of affliction: Fibroblasts – a new await at an old trouble. Nat Clin Pract Nephrol. 2006;ii:101–108. doi: x.1038/ncpneph0093. [PubMed] [CrossRef] [Google Scholar]
  • Kalluri R, Neilson EG. Epithelial-mesenchymal transition and its implications for fibrosis. J Clin Invest. 2003;112:1776–1784. [PMC complimentary article] [PubMed] [Google Scholar]
  • Ruiz-Ortega Yard, Rodriguez-Vita J, Sanchez-Lopez E, Carvajal G, Egido J. TGF-beta signaling in vascular fibrosis. Cardiovasc Res. 2007;74:196–206. doi: 10.1016/j.cardiores.2007.02.008. [PubMed] [CrossRef] [Google Scholar]
  • Ruiz-Ortega M, Ruperez M, Esteban Five, Rodriguez-Vita J, Sanchez-Lopez Due east, Carvajal One thousand, Egido J. Angiotensin II: a key factor in the inflammatory and fibrotic response in kidney diseases. Nephrol Punch Transplant. 2006;21:16–20. doi: 10.1093/ndt/gfi265. [PubMed] [CrossRef] [Google Scholar]
  • Kuwano One thousand, Hagimoto Due north, Kawasaki One thousand, Yatomi T, Nakamura Due north, Nagata S, Suda T, Kunitake R, Maeyama T, Miyazaki H, Hara North. Essential roles of the Fas-Fas ligand pathway in the development of pulmonary fibrosis. J Clin Invest. 1999;104:xiii–19. doi: 10.1172/JCI5628. [PMC free commodity] [PubMed] [CrossRef] [Google Scholar]
  • Hinz B, Phan SH, Thannickal VJ, Galli A, Bochaton-Piallat ML, Gabbiani G. The myofibroblast: one part, multiple origins. Am J Pathol. 2007;170:1807–1816. doi: 10.2353/ajpath.2007.070112. [PMC gratuitous article] [PubMed] [CrossRef] [Google Scholar]
  • Kisseleva T, Uchinami H, Feirt N, Quintana-Bustamante O, Segovia JC, Schwabe RF, Brenner DA. Bone marrow-derived fibrocytes participate in pathogenesis of liver fibrosis. J Hepatol. 2006;45:429–438. doi: 10.1016/j.jhep.2006.04.014. [PubMed] [CrossRef] [Google Scholar]
  • Zeisberg EM, Potenta SE, Sugimoto H, Zeisberg M, Kalluri R. Fibroblasts in kidney fibrosis emerge via endothelial-to-mesenchymal transition. J Am Soc Nephrol. 2008;xix:2282–2287. doi: x.1681/ASN.2008050513. [PMC costless article] [PubMed] [CrossRef] [Google Scholar]
  • Sironi Fifty, Nobili E, Gianella A, Gelosa P, Tremoli E. Anti-inflammatory properties of drugs acting on the renin-angiotensin system. Drugs Today (Barc) 2005;41:609–622. doi: 10.1358/dot.2005.41.9.899613. [PubMed] [CrossRef] [Google Scholar]
  • Herzlinger D. Renal interstitial fibrosis: remembrance of things past? J Clin Invest. 2002;110:305–306. [PMC costless article] [PubMed] [Google Scholar]
  • Lin SL, Kisseleva T, Brenner DA, Duffield JS. Pericytes and perivascular fibroblasts are the main source of collagen-producing cells in obstructive fibrosis of the kidney. Am J Pathol. 2008;173:1617–1627. doi: x.2353/ajpath.2008.080433. [PMC gratis article] [PubMed] [CrossRef] [Google Scholar]
  • Romagnani P, Lasagni 50, Mazzinghi B, Lazzeri E, Romagnani South. Pharmacological modulation of stem prison cell function. Curr Med Chem. 2007;14:1129–1139. doi: x.2174/092986707780362880. [PubMed] [CrossRef] [Google Scholar]

Manufactures from Fibrogenesis & Tissue Repair are provided here courtesy of BioMed Key


Can A Kidney Repair Itself,

Source: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2711960/

Posted by: brownpospits1945.blogspot.com

0 Response to "Can A Kidney Repair Itself"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel