Next Article in Journal
Cell Type-Specific TGF-β Mediated EMT in 3D and 2D Models and Its Reversal by TGF-β Receptor Kinase Inhibitor in Ovarian Cancer Cell Lines
Next Article in Special Issue
Epithelial–Mesenchymal Transition (EMT)
Previous Article in Journal
Highly Efficient and Stable Removal of Arsenic by Live Cell Fabricated Magnetic Nanoparticles
Previous Article in Special Issue
Proneural-Mesenchymal Transition: Phenotypic Plasticity to Acquire Multitherapy Resistance in Glioblastoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Key Role of Epithelial to Mesenchymal Transition (EMT) in Hypertensive Kidney Disease

Hypertension Unit, Department of Medicine-DIMED, University Hospital, 35128 Padova, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(14), 3567; https://doi.org/10.3390/ijms20143567
Submission received: 20 May 2019 / Revised: 12 July 2019 / Accepted: 18 July 2019 / Published: 21 July 2019
(This article belongs to the Special Issue Epithelial-Mesenchymal Transition (EMT))

Abstract

:
Accumulating evidence indicates that epithelial-to-mesenchymal transition (EMT), originally described as a key process for organ development and metastasis budding in cancer, plays a key role in the development of renal fibrosis in several diseases, including hypertensive nephroangiosclerosis. We herein reviewed the concept of EMT and its role in renal diseases, with particular focus on hypertensive kidney disease, the second leading cause of end-stage renal disease after diabetes mellitus. After discussing the pathophysiology of hypertensive nephropathy, the ‘classic’ view of hypertensive nephrosclerosis entailing hyalinization, and sclerosis of interlobular and afferent arterioles, we examined the changes occurring in the glomerulus and tubulo-interstitium and the studies that investigated the role of EMT and its molecular mechanisms in hypertensive kidney disease. Finally, we examined the reasons why some studies failed to provide solid evidence for renal EMT in hypertension.

1. Introduction

More than 20 million people have chronic kidney disease (CKD) in the US, with 3% of them developing end-stage renal disease (ESRD) (https://www.usrds.org/adr.aspx), which accounts for 7% of Medicare expenditures. In the EU, the last annual report of the European Renal Association—European Dialysis and Transplant Association (ERA-EDTA) Registry indicated that about 500 thousand people were on renal replacement therapy in 2018 [1]. Moreover, the medical spending is held to be much greater, because ESRD patients are at high risk of cardio- and cerebrovascular vascular disease, which is projected to rise even more, owing to population ageing in the next decades [2]. Arterial hypertension (HT), with a prevalence of 30–45% in the adult population, is the major cardiovascular risk factor and the second cause of ESRD after diabetes mellitus [3,4]. However, the relationship between the kidney and HT is double-threaded, in that nephropathies usually cause HT and high blood pressure damages the kidney [5].
Even though classically described as nephroangiosclerosis and hyalinosis of the glomerular tuft [6,7], hypertensive nephropathy has been more recently found to involve not only the glomerular and vascular compartments, but also the interstitium, because of the development of tubular-interstitial fibrosis (TIF) and, ultimately, ESRD [8,9].
Depending on the underlying pathophysiology and/or severity of HT, multiple mechanisms can concur to induce TIF, among which is epithelial-to-mesenchymal transition (EMT), a process originally described in organ development and metastasis budding in cancer, and later observed in several diseases [10,11]. Although potentially capable of affecting any injured epithelial tissue, whether and to what extent EMT plays role in the kidney in hypertensive nephropathy remains unclear [12,13,14]. Hence, we shall herein focus on the EMT process and its involvement in hypertensive nephropathy by exploiting a methodology to systematically review the English-written literature by means of a PICO (Patient or Problem, Intervention, Comparison, Outcome) search strategy (Table 1) with selection of the articles based on their relevance, rigorousness of study design and methodology, and appropriateness of data interpretation.

2. Tubulointerstitial Damage and Fibrosis in Hypertension

In 1993, an involvement of tubule-interstitium in the hypertensive disease was postulated for the first time by Fine [15], who put forward the ‘chronic hypoxia hypothesis’. According to this contention, glomerular injury, primarily determined by HT, would induce hypoxia and endothelial damage of peritubular capillaries [15,16], thus stimulating tubular inflammation and fibrosis. Although appealing, this hypothesis remained unproven until one study showed peritubular capillaries rarefaction and tubulointerstitial scarring in human biopsies of CKD hypertensive patients [17] and models of hypertensive disease documented an involvement of the interstitium in HT [8,9,18,19].
Fibroblasts are strategically posed in the interstitium to sense circulating toxins or factors and environmental changes and can be activated by local tissue injury to differentiate into myofibroblasts that synthesize extracellular matrix deposition and collagens. However, fibroblasts can also act as inflammatory effector cells: They express immune receptors as Toll-like receptors (TLRs) and activate NF-κB signaling with ensuing release of cytokines and chemokines that amplify kidney damage. In the hypertensive nephropathy, TIF usually associates with glomerular damage, loss of nephrons, and decline in renal function. Nephrons are then focally replaced by fibrotic tissue along a scarring process, finally leading to ESRD [20]. Thus, chronic and progressive tubular damage invariably leads to TIF.
Several factors can induce TIF. One of them is angiotensin II (Ang II), a major effector of the renin angiotensin aldosterone system (RAAS), which acts as a potent vasoconstricting and pro-inflammatory factor. Using a transgenic renin-dependent model of severe HT and cardiovascular and renal damage created with insertion of the mouse renin gene into the rat genome, the TG(mRen2)27 rat, we found that a blockade of the Ang II type 1 (AT1) receptor with irbesartan prevented TIF, thus supporting a fibrogenic role of Ang II [8]. Interestingly, in this model, TIF was also prevented by bosentan, a mixed ETA-ETB endothelin receptor antagonist, but worsened by BMS-182874, a selective ETA receptor antagonist, suggesting a key role of endothelin-1 (ET-1) acting via the ETB receptor subtype in triggering renal fibrosis [8].
It is important to note that dual inhibition of a angiotensin-converting enzyme (ACE) and a neutral endopeptidase also prevented TIF; moreover, this favorable effect was abolished by the bradykinin B2 receptor antagonist icatibant, thus implicating the B2 subtype receptor in counterbalancing the deleterious effects of Ang II in the tubule-interstitium [9].
Ang II can promote TIF not only by activating resident fibroblasts, but also by inducing proteinuria or EMT (see later “EMT in Hypertensive Nephropathy”). When hypertensive nephropathy is associated with proteinuria, filtered proteins are held to exert a further toxic effect on proximal tubular cells, which can amplify inflammation [21]. Pro-inflammatory factors, such as monocyte chemoattractant protein-1 (MCP-1/CCL2), regulated upon activation normal T-cell expressed and secreted/Chemokine (C-C motif) ligand 5 (RANTES/ CCL5), and fractalkine/CX3CL1 are chemoattractant for monocytes/macrophages and T lymphocytes, which stimulate synthesis of transforming growth factor β1 (TGFβ1), a potent fibroblast activator [14].

3. Epithelial-to-Mesenchymal Transition (EMT)

Renal epithelial cells arise during embryogenesis by mesenchymal-to-epithelial transition (MET) [22,23]. For decades the dogma was that epithelial cells maintain its phenotype until death, but, after the seminal study by Gros in 2014 [24], the concept that epithelial cells can switch back to a mesenchymal phenotype via epithelial-to-mesenchymal transition (EMT) was promptly accepted [25,26]. First described in embryogenesis [24], EMT was then observed in inflammation [27], fibrosis [28,29,30], wound healing [31], and cancer progression [32,33,34].
Epithelia are characterized by apical-basal polarity and tight cell–cell junctions between neighboring cells, accounting for cell integrity and stability, whereas mesenchymal and interstitial tissues are formed by elongated and spindle-shaped cells, known as fibroblast-like cells, which have loose cell–cell interactions. Because of their mobility, these cells can migrate and create an anchorage distantly.
The cells undergoing EMT progressively lose epithelial markers, such as E-cadherin, which is the main component of adherent junctions, and acquire markers of mesenchymal phenotype, such as α smooth muscle actin (αSMA) [35]. Loss of cell junctions allows epithelial cells to translate into loss of adhesion, change morphology, and loss apical-basal polarity. After synthesis of matrix metalloproteinases (MMPs), cells cleave and invade the basal lamina to move towards the interstitial space. Migration is driven by newly formed matrix of fibronectin and type I collagen. Upon completion of the epithelial-to-mesenchymal phenotype switch, the cells acquire a myofibroblast phenotype synthesizing αSMA and matrix proteins, the major constituents of TIF [35].
A number of epithelial and mesenchymal markers have been used to track EMT, and a lot of factors have been identified as drivers of EMT. Not unexpectedly, most factors were also found to be involved in embryogenesis, tissue repair, and cancer, suggesting parallels between normal development and malignant growth. TGFβ, a driver of EMT in cancer, is also the most potent factor triggering EMT in the renal cells. After the binding of TGF-β to receptor II TβR-II, TGF-β-bound TβR-II phosphorylates type I receptor (TβR-I) leads to the activation of Smad 2 and 3. Activated Smad2/3 can form a complex with Smad4, which translocates into the nucleus, where it interacts with transcriptional factors and transcriptional co-activators, ultimately inducing synthesis of mesenchymal proteins [36,37,38].

4. Markers of EMT

In 1995, Strutz et al. detected S100A4 in tubular cells with persistent inflammation. Since S100A4 was considered a typical protein of mesenchymal cells and only seldom found in tubular cells of healthy kidneys, S100A4 was renamed fibroblast-specific protein–1 (FSP-1) and proposed to be a marker of cells undergoing EMT [39]. However, other investigators were not able to localize FSP-1/S100A4 in activated fibroblasts, and found this protein in immune cells and occasionally in endothelial cells. Hence, the specificity of FSP-1/S100A4 as a marker of EMT was challenged [40,41].
Other markers of EMT have later been proposed. Vimentin is expressed in mesenchymal cells, but also in injured tubular cells, and/or regenerating cells [42,43,44], which led some investigators to consider it as a marker of EMT and others to propose it as a marker of tubular regenerating activity.
αSMA is characteristically expressed in mesenchymal cells, such as vascular smooth muscle cells, pericytes, and interstitial cells around injured tubules and myofibroblasts [45,46]. In contrast, proximal tubular cells contain F-actin, a different isoform that constitutes cytoskeleton microfilaments and cannot bind antibodies against αSMA. However, detection of αSMA cannot be used as a marker specific of EMT, because it cannot discriminate between mesenchymal cells physiologically present in the kidney or myofibroblasts.
The decrease or loss of markers of epithelial cells, such as E-cadherin or ZO-1, is strongly suggestive of EMT, but could also denote tubular cell damage with no transition into mesenchymal cells.
Collagens and proteins that participate to collagen synthesis, such as the collagen-specific molecular chaperone HSP 47 and prolyl-4-hydroxylase, have been also used as markers [47].
Invasion of the basal membrane and cell migration are markers of fully completed EMT. Hence, different markers of EMT have been suggested, with some markers (loss of E-cadherin and ZO-1, synthesis of αSMA and vimentin) indicating the early phenotypic cell switch and others (collagens, matrix invasion) hinting the full transformation of epithelial cells into mesenchymal cells. However, because of the lack of specific markers of transforming cells, per se, no marker can provide an unambiguous demonstration of EMT. As discussed below, EMT is a dynamic and transient process that is difficult to visualize. Only an integrated approach evaluating early and late steps can allow identification of EMT.

5. EMT as Mediator of TIF

Using a mouse model of severe renal fibrosis, i.e., unilateral ureteral obstruction, Yang and Liu first described EMT in the kidney as an orchestrated process consisting of the classical four steps, i.e., loss of epithelial cell adhesion, de novo synthesis of αSMA, disruption of tubular basement membrane, and cell migration and invasion [48]. In the same model, Iwano et al. found that up to 36% of interstitial matrix-producing cells derived from tubular cells through EMT [35]. However, using models characterized by less pronounced TIF, such as overload proteinuria [48], other investigators found that the contribution of EMT was very scant, and, Humphreys et al. [49], using the model of unilateral ureteral obstruction, were unable to reproduce data from Iwano [35], suggesting that, although renal epithelial cells can acquire mesenchymal markers in vitro, they do not contribute to interstitial myofibroblast cells in vivo. Thus, whether EMT actually contributes to TIF in vivo in all types of nephropathies remains controversial [50,51].
According to a popular view, EMT would be a process by which epithelial cells under stress conditions escape from the unfriendly microenvironment [52]. After an injury, tubular epithelial cells express genes, such as Wnt, Notch, and Hedgehog, which confer resistance to apoptosis and are involved in normal development [52]. However, during physiologic development differentiated epithelial cells proliferate and in an injured kidney epithelial cells acquire partial or full mesenchymal characteristics, thereby preventing proliferation and replacement of damaged cells. Hence, tubular epithelial cells are not locked in a differentiated state, but are ‘plastic cells’ that transform into a different phenotype, but are unable to reach terminal differentiation [52].
Two mechanisms have been proposed: (1) Differentiation of epithelial cells needs “turning off” of Wnt and Notch genes, a process that does not happen in damaged cells; (2) the over-expression of transcription factors, such as Twist, Zeb1, Zeb2, Snai1, and Snai2 favors, mesenchymal reprogramming, preventing terminal differentiation. Elegant studies by Kalluri’s and Nieto’s groups support such contentions: mice over-expressing Snai1 exhibited epithelial plasticity, myofibroblast accumulation, and inflammation, whereas an ablation of Twist and Snai1 induced expression of markers of differentiated cells, such as aquaporins and solute transporters [29,30].
In contrast to other strategies used by epithelial cells to avoid threatening conditions, such as apoptosis or necrosis, EMT is hardly identifiable because of the difficulty of catching exactly the time-point when cells transform into mesenchymal cells and pass the basement membrane [51]. Whether EMT is too rapid to be detected or visualization of EMT needs an integrated methodological approach remains a matter of debate.

6. EMT in Hypertensive Nephropathy

Several in vitro studies showed that some factors that induce HT, fibrosis, or both can also drive EMT, suggesting that EMT is a common mechanism underlying TIF [13,53,54,55]. However, only a few studies provided evidences of EMT in vivo in hypertensive nephropathy [12,56,57].
Ang II was found to induce EMT in cell culture models [13,53,54,55,58]. In the rat proximal tubular cell line (NRK52-E), EMT was identified as a morphological change of the epithelial cells from the typical cobblestone pattern to elongated, spindle-shaped mesenchymal cells, and also as a reduction in E-cadherin and synthesis of αSMA protein expression [54]. In a rat model of accelerated ESRD induced by a 14-day infusion of Ang II after preinjury by injection of Habu venom, Falkner et al. observed a progressive increase in αSMA staining, starting from 48 h. However, peritubular interstitial myofibroblasts were also detected after 7 days of Ang II alone infusion, confirming capability of Ang II of inducing EMT [50]. It is important to note that a blockade of the AT1 receptor prevented EMT in models of HT, caused by partial nephrectomy [56] or unilateral ureteral obstruction [57], supporting a role of Ang II in EMT.
As mentioned above, we recently provided evidence that ET-1 can also induce EMT in the kidney [12]: In the same kidney sections of TG(mRen2)27 rats, where we observed TIF, we found a decrease in the epithelial marker E-cadherin along with an increase in the mesenchymal markers αSMA and S100A4, i.e., phenotypic changes that indicated the occurrence of EMT. We also demonstrated co-expression of the epithelial with mesenchymal markers, thus strongly supporting EMT as the mechanism underlying fibrosis in this Ang-II-dependent model of severe hypertension [12].
Moreover, in vitro experiments confirmed that ET-1 can drive EMT in the proximal tubular cells [12]. Exposure of HK2 cells to ET-1 caused disruption of cell junctions and synthesis of αSMA, causing blunting of E-cadherin along with the increase of mesenchymal markers, synthesis of MMP-9, and cell migration [12]. All steps of EMT, such as TIF involved the ETB receptor subtype, which is the predominant receptor subtype in the kidney tubular cells (Figure 1) [12].

7. Conclusions

In summary, EMT entails a dynamic sequential process characterized by progressive loss of epithelial markers and acquirement of a mesenchymal phenotype marker that is key to several physiological and pathological processes that need cell plasticity. Although evidences obtained in a model of severe Ang II-dependent hypertension strongly support the contention that EMT contributes to TIF, which is the major determinant of ESRD, likely via the activation of the AT1 and ETB receptors, it remains debated if EMT contributes to TIF in other forms of nephropathy, which are characterized by less prominent TIF.

Author Contributions

Conceptualization, T.M.S. and G.P.R.; literature analysis, T.M.S., M.P., and B.C.; writing—original draft preparation and editing, T.M.S. and B.C.; funding acquisition, T.M.S. and G.P.R.

Funding

This work was supported by grants from the Ministry of Health (RF2011-02352318) to TMS and from the University of Padova (DOR1625891/16; DOR1670784/16; BIRD163255/16) to TMS and GPR.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

αSMAAlpha smooth muscle actin
Ang IIAngiotensin II
ACEAngiotensin-converting enzyme
AT1Angiotensin II type 1 receptor
HTArterial hypertension
CKDChronic kidney disease
ET-1Endothelin-1
ESRDEnd-stage renal disease
EMTEpithelial-to-mesenchymal transition
ERA-EDTAEuropean Renal Association—European Dialysis and Transplant Association
FSP-1Fibroblast-specific protein–1
METMesenchymal-to-epithelial transition
MMPMetalloproteinases
MCP-1Monocyte Chemoattractant Protein-1
RANTESRegulated upon Activation Normal T-cell Expressed and Secreted
RAASRenin angiotensin aldosterone system
TGFβR-ITGF-β receptors type I
TLRsToll-like receptors
TGFβ1Transforming growth factor β1
TIFTubular-interstitial fibrosis

References

  1. Kramer, A.; Pippias, M.; Noordzij, M.; Stel, V.S.; Afentakis, N.; Ambühl, P.M.; Andrusev, A.M.; Fuster, E.A.; Monzón, F.E.A.; Åsberg, A.; et al. The European Renal Association—European Dialysis and Transplant Association (ERA-EDTA) Registry Annual Report 2015: A summary. Clin. Kidney J. 2018, 11, 108–122. [Google Scholar] [CrossRef]
  2. Levey, A.S.; Coresh, J. Chronic kidney disease. Lancet 2012, 379, 165–180. [Google Scholar] [CrossRef]
  3. Williams, B.; Mancia, G.; Spiering, W.; Rosei, E.A.; Azizi, M.; Burnier, M.; Clement, D.; Coca, A.; De Simone, G.; Dominiczak, A.; et al. 2018 Practice Guidelines for the management of arterial hypertension of the European Society of Cardiology and the European Society of Hypertension. J. Hypertens. 2018, 36, 2284–2309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Udani, S.; Lazich, I.; Bakris, G.L. Epidemiology of hypertensive kidney disease. Nat. Rev. Nephrol. 2011, 7, 11–21. [Google Scholar] [CrossRef]
  5. Ritz, E. The Kidney: Both culprit and victim. Hypertension 2009, 54, 25–26. [Google Scholar] [CrossRef] [PubMed]
  6. Hill, G.S.; Heudes, D.; Bariéty, J. Morphometric study of arterioles and glomeruli in the aging kidney suggests focal loss of autoregulation. Kidney Int. 2003, 63, 1027–1036. [Google Scholar] [CrossRef] [PubMed]
  7. Hill, G.S. Hypertensive nephrosclerosis. Curr. Opin. Nephrol. Hypertens. 2008, 17, 266–270. [Google Scholar] [CrossRef] [PubMed]
  8. Seccia, T.M.; Maniero, C.; Belloni, A.S.; Guidolin, D.; Poulose, P.; Pessina, A.C.; Rossi, G.P. Role of angiotensin II, endothelin-1 and L-type calcium channel in the development of glomerular, tubulointerstitial and perivascular fibrosis. J. Hypertens. 2008, 26, 2022–2029. [Google Scholar] [CrossRef] [PubMed]
  9. Seccia, T.M.; Belloni, A.S.; Guidolin, D.; Sticchi, D.; Nussdorfer, G.G.; Pessina, A.C.; Rossi, G.P. The renal antifibrotic effects of angiotensin-converting enzyme inhibition involve bradykinin B2 receptor activation in angiotensin II-dependent hypertension. J. Hypertens. 2006, 24, 1419–1427. [Google Scholar] [CrossRef] [PubMed]
  10. Ribatti, D. Epithelial-mesenchymal transition in morphogenesis, cancer progression and angiogenesis. Exp. Cell Res. 2017, 353, 1–5. [Google Scholar] [CrossRef] [PubMed]
  11. Piera-Velazquez, S.; Li, Z.; Jimenez, S.A. Role of endothelial-mesenchymal transition (EndoMT) in the pathogenesis of fibrotic disorders. Am. J. Pathol. 2011, 179, 1074–1080. [Google Scholar] [CrossRef] [PubMed]
  12. Seccia, T.M.; Caroccia, B.; Gioco, F.; Piazza, M.; Buccella, V.; Guidolin, D.; Guerzoni, E.; Montini, B.; Petrelli, L.; Pagnin, E.; et al. Endothelin-1 Drives Epithelial-Mesenchymal Transition in Hypertensive Nephroangiosclerosis. J. Am. Heart Assoc. 2016, 5. [Google Scholar] [CrossRef] [PubMed]
  13. Carvajal, G.; Rodriguez-Vita, J.; Rodrigues-Díez, R.; Sánchez-López, E.; Rupérez, M.; Cartier, C.; Esteban, V.; Ortiz, A.; Egido, J.; Mezzano, S.A.; et al. Angiotensin II activates the Smad pathway during epithelial mesenchymal transdifferentiation. Kidney Int. 2008, 74, 585–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Macconi, D.; Remuzzi, G.; Benigni, A. Key fibrogenic mediators: Old players. Renin-angiotensin system. Kidney Int. Suppl. 2014, 4, 58–64. [Google Scholar] [CrossRef] [PubMed]
  15. Fine, L.G.; Ong, A.C.M.; Norman, J.T. Mechanisms of tubulo-interstitiaI injury in progressive renal diseases. Eur. J. Clin. Investig. 1993, 23, 259–265. [Google Scholar] [CrossRef]
  16. Fine, L.G.; Norman, J.T. Chronic hypoxia as a mechanism of progression of chronic kidney diseases: From hypothesis to novel therapeutics. Kidney Int. 2008, 74, 867–872. [Google Scholar] [CrossRef]
  17. Choi, Y.J.; Chakraborty, S.; Nguyen, V.; Nguyen, C.; Kim, B.K.; Shim, S.I.; Suki, W.N.; Truong, L.D. Peritubular capillary loss is associated with chronic tubulointerstitial injury in human kidney: Altered expression of vascular endothelial growth factor. Hum. Pathol. 2000, 31, 1491–1497. [Google Scholar] [CrossRef]
  18. Gross, O.; Schulze-Lohoff, E.; Koepke, M.-L.; Beirowski, B.; Addicks, K.; Bloch, W.; Smyth, N.; Weber, M. Antifibrotic, nephroprotective potential of ACE inhibitors vs AT1 antagonist in a murine model of renal fibrosis. Nephrol. Dial. Transpl. 2004, 19, 1716–1723. [Google Scholar] [CrossRef]
  19. Hartner, A.; Porst, M.; Klanke, B.; Cordasic, N.; Hilgers, K.F.; Veelken, R. Angiotensin II formation in the kidney and nephrosclerosis in Ren-2 hypertensive rats. Nephrol. Dial. Transpl. 2006, 21, 1778–1785. [Google Scholar] [CrossRef] [Green Version]
  20. Kriz, W.; Kaissiing, B.; Le Hir, M. Epithelial-mesenchymal transition (EMT) in kidney fibrosis: Fact or fantasy? J. Clin. Investig. 2011, 121, 468–474. [Google Scholar] [CrossRef]
  21. Liu, B.C.; Tang, T.T.; Lv, L.L.; Lan, H.Y. Renal tubule injury: A driving force toward chronic kidney disease. Kidney Int. 2018, 93, 568–579. [Google Scholar] [CrossRef] [PubMed]
  22. Galichon, P.; Finianos, S.; Hertig, A. EMT-MET in renal disease: Should we curb our enthusiasm? Cancer Lett. 2013, 341, 24–29. [Google Scholar] [CrossRef] [PubMed]
  23. Hay, E.D.; Zuk, A. Transformations between epithelium and mesenchyme: Normal, pathological, and experimentally induced1. Am. J. Kidney Dis. 1995, 26, 678–690. [Google Scholar] [CrossRef]
  24. Gros, J.; Tabin, C.J. Vertebrate limb bud formation is initiated by localized epithelial-to-mesenchymal transition. Science 2014, 343, 1253–1256. [Google Scholar] [CrossRef] [PubMed]
  25. Stark, K.; Vainio, S.; Vassileva, G.; McMahon, A.P. Epithelial transformation of metanephric mesenchyme in the developing kidney regulated by Wnt-4. Nature 1994, 372, 679–683. [Google Scholar] [CrossRef]
  26. Thiery, J.P.; Acloque, H.; Huang, R.Y.J.; Nieto, M.A. Epithelial-mesenchymal transitions in development and disease. Cell 2009, 139, 871–890. [Google Scholar] [CrossRef]
  27. Neilson, E.G. Mechanisms of Disease: Fibroblasts—A new look at an old problem. Nat. Clin. Pract. Nephrol. 2006, 2, 101–108. [Google Scholar] [CrossRef]
  28. Strutz, F.; Neilson, E.G. New insights into mechanisms of fibrosis in immune renal injury. Springer Semin. Immunopathol. 2003, 24, 459–476. [Google Scholar] [CrossRef]
  29. Lovisa, S.; LeBleu, V.S.; Tampe, B.; Sugimoto, H.; Vadnagara, K.; Carstens, J.L.; Wu, C.-C.; Hagos, Y.; Burckhardt, B.C.; Pentcheva-Hoang, T.; et al. Epithelial-to-mesenchymal transition induces cell cycle arrest and parenchymal damage in renal fibrosis. Nat. Med. 2015, 21, 998–1009. [Google Scholar] [CrossRef]
  30. Grande, M.T.; Sánchez-Laorden, B.; López-Blau, C.; A De Frutos, C.; Boutet, A.; Arévalo, M.; Rowe, R.G.; Weiss, S.J.; López-Novoa, J.M.; Nieto, M.A. Snail1-induced partial epithelial-to-mesenchymal transition drives renal fibrosis in mice and can be targeted to reverse established disease. Nat. Med. 2015, 21, 989–997. [Google Scholar] [CrossRef] [Green Version]
  31. Banerjee, P.; Venkatachalam, S.; Mamidi, M.K.; Bhonde, R.; Shankar, K.; Pal, R. Vitiligo patient-derived keratinocytes exhibit characteristics of normal wound healing via epithelial to mesenchymal transition. Exp. Dermatol. 2015, 24, 391–393. [Google Scholar] [CrossRef] [Green Version]
  32. Wei, S.C.; Fattet, L.; Tsai, J.H.; Guo, Y.; Pai, V.H.; Majeski, H.E.; Chen, A.C.; Sah, R.L.; Taylor, S.S.; Engler, A.J.; et al. Matrix stiffness drives epithelial-mesenchymal transition and tumour metastasis through a TWIST1-G3BP2 mechanotransduction pathway. Nat. Cell Biol. 2015, 17, 678–688. [Google Scholar] [CrossRef]
  33. Ruscetti, M.; Quach, B.; Dadashian, E.L.; Mulholland, D.J.; Wu, H. Tracking and Functional Characterization of Epithelial-Mesenchymal Transition and Mesenchymal Tumor Cells during Prostate Cancer Metastasis. Cancer Res. 2015, 75, 2749–2759. [Google Scholar] [CrossRef]
  34. Chen, T.; You, Y.; Jiang, H.; Wang, Z.Z. Epithelial–Mesenchymal transition (EMT): A biological process in the development, stem cell differentiation, and tumorigenesis. J. Cell. Physiol. 2017, 232, 3261–3272. [Google Scholar] [CrossRef]
  35. Iwano, M.; Plieth, D.; Danoff, T.M.; Xue, C.; Okada, H.; Neilson, E.G. Evidence that fibroblasts derive from epithelium during tissue fibrosis. J. Clin. Investig. 2002, 110, 341–350. [Google Scholar] [CrossRef]
  36. Saitoh, M.; Miyazawa, K. Transcriptional and Post-transcriptional Regulation in TGF–Mediated epithelial-mesenchymal transition. J. Biochem. 2012, 151, 563–571. [Google Scholar] [CrossRef]
  37. Griggs, L.A.; Hassan, N.T.; Malik, R.S.; Griffin, B.P.; Martinez, B.A.; Elmore, L.W.; Lemmon, C.A. Fibronectin fibrils regulate TGF-β1-induced Epithelial-Mesenchymal Transition. Matrix Biol. 2017, 60–61, 157–175. [Google Scholar] [CrossRef]
  38. Hewitson, T.D.; Holt, S.G.; Tan, S.J.; Wigg, B.; Samuel, C.S.; Smith, E.R. Epigenetic modifications to H3K9 in renal tubulointerstitial cells after unilateral ureteric obstruction and TGF-β1 stimulation. Front. Pharmacol. 2017, 8, 1–15. [Google Scholar] [CrossRef]
  39. Strutz, F.; Okada, H.; Lo, C.W.; Danoff, T.; Carone, R.L.; Tomaszewski, J.E.; Neilson, E.G. Identification and characterization of a fibroblast marker: FSP1. J. Cell Biol. 1995, 130, 393–405. [Google Scholar] [CrossRef]
  40. Inoue, T.; Plieth, D.; Venkov, C.D.; Xu, C.; Neilson, E.G. Antibodies against macrophages that overlap in specificity with fibroblasts. Kidney Int. 2005, 67, 2488–2493. [Google Scholar] [CrossRef] [Green Version]
  41. Le Hir, M.; Hegyi, I.; Cueni-Loffing, D.; Loffing, J.; Kaissling, B. Characterization of renal interstitial fibroblast-specific protein 1/S100A4-positive cells in healthy and inflamed rodent kidneys. Histochem. Cell Biol. 2005, 123, 335–346. [Google Scholar] [CrossRef] [Green Version]
  42. Zhu, M.Q.; De Broe, M.E.; Nouwen, E.J. Vimentin expression and distal tubular damage in the rat kidney. Exp. Nephrol. 1996, 4, 172–183. [Google Scholar]
  43. Witzgall, R.; Brown, D.; Schwarz, C.; Bonventre, J.V. Localization of proliferating cell nuclear antigen, vimentin, c-Fos, and clusterin in the postischemic kidney. Evidence for a heterogenous genetic response among nephron segments, and a large pool of mitotically active and dedifferentiated cells. J. Clin. Investig. 1994, 93, 2175–2788. [Google Scholar] [CrossRef]
  44. Gröne, H.J.; Weber, K.; Gröne, E.; Helmchen, U.; Osborn, M. Coexpression of keratin and vimentin in damaged and regenerating tubular epithelia of the kidney. Am. J. Pathol. 1987, 129, 1–8. [Google Scholar]
  45. Skalli, O.; Pelte, M.F.; Peclet, M.C.; Gabbiani, G.; Gugliotta, P.; Bussolati, G.; Ravazzola, M.; Orci, L. Alpha-smooth muscle actin, a differentiation marker of smooth muscle cells, is present in microfilamentous bundles of pericytes. J. Histochem. Cytochem. 1989, 37, 315–321. [Google Scholar] [CrossRef]
  46. Gabbiani, G.; Schmid, E.; Winter, S.; Chaponnier, C.; De Ckhastonay, C.; Vandekerckhove, J.; Weber, K.; Franke, W.W. Vascular smooth muscle cells differ from other smooth muscle cells: Predominance of vimentin filaments and a specific alpha-type actin. Proc. Natl. Acad. Sci. USA 1981, 78, 298–302. [Google Scholar] [CrossRef]
  47. Cheng, S.; Pollock, A.S.; Mahimkar, R.; Olson, J.L.; Lovett, D.H. Matrix metalloproteinase 2 and basement membrane integrity: A unifying mechanism for progressive renal injury. FASEB J. 2006, 20, 1898–1900. [Google Scholar] [CrossRef]
  48. Yang, J.; Liu, Y. Dissection of Key Events in Tubular Epithelial to Myofibroblast Transition and Its Implications in Renal Interstitial Fibrosis. Am. J. Pathol. 2001, 159, 1465–1475. [Google Scholar] [CrossRef] [Green Version]
  49. Humphreys, B.D.; Lin, S.-L.; Kobayashi, A.; Hudson, T.E.; Nowlin, B.T.; Bonventre, J.V.; Valerius, M.T.; McMahon, A.P.; Duffield, J.S. Fate tracing reveals the pericyte and not epithelial origin of myofibroblasts in kidney fibrosis. Am. J. Pathol. 2010, 176, 85–97. [Google Scholar] [CrossRef]
  50. Faulkner, J.L.; Szcykalski, L.M.; Springer, F.; Barnes, J.L. Origin of interstitial fibroblasts in an accelerated model of angiotensin II-induced renal fibrosis. Am. J. Pathol. 2005, 167, 1193–1205. [Google Scholar] [CrossRef]
  51. Zhu, C.; Mertens, P.R. Epithelial–Mesenchymal transition to be or not to be? Is the answer yes and no at the same time? Int. Urol. Nephrol. 2010, 42, 843–846. [Google Scholar] [CrossRef]
  52. Huang, S.; Susztak, K. Epithelial Plasticity versus EMT in Kidney Fibrosis. Trends Mol. Med. 2016, 22, 4–6. [Google Scholar] [CrossRef]
  53. Chen, L.; Liu, B.-C.; Zhang, X.-L.; Zhang, J.-D.; Liu, H.; Li, M.-X. Influence of connective tissue growth factor antisense oligonucleotide on angiotensin II-induced epithelial mesenchymal transition in HK2 cells. Acta Pharmacol. Sin. 2006, 27, 1029–1036. [Google Scholar] [CrossRef]
  54. Burns, W.C.; Thomas, M.C. Angiotensin II and its role in tubular epithelial to mesenchymal transition associated with chronic kidney disease. Cells Tissues Organs 2010, 193, 74–84. [Google Scholar] [CrossRef]
  55. Rodrigues-Díez, R.; Carvajal-González, G.; Sánchez-López, E.; Rodríguez-Vita, J.; Rodrigues Díez, R.; Selgas, R.; Ortiz, A.; Egido, J.; Mezzano, S.; Ruiz-Ortega, M. Pharmacological modulation of epithelial mesenchymal transition caused by angiotensin II. Role of ROCK and MAPK pathways. Pharm. Res. 2008, 25, 2447–2461. [Google Scholar] [CrossRef]
  56. Yang, F.; Huang, X.R.; Chung, A.C.; Hou, C.-C.; Lai, K.N.; Lan, H.Y. Essential role for Smad3 in angiotensin II-induced tubular epithelial-mesenchymal transition. J. Pathol. 2010, 221, 390–401. [Google Scholar] [CrossRef]
  57. Kellner, D.; Chen, J.; Richardson, I.; Seshan, S.V.; El Chaar, M.; Vaughan, E.; Poppas, D.; Felsen, D. Angiotensin Receptor Blockade Decreases Fibrosis and Fibroblast Expression in a Rat Model of Unilateral Ureteral Obstruction. J. Urol. 2006, 176, 806–812. [Google Scholar] [CrossRef]
  58. Hu, H.; Hu, S.; Xu, S.; Gao, Y.; Zeng, F.; Shui, H. miR-29b regulates Ang II-induced EMT of rat renal tubular epithelial cells via targeting PI3K/AKT signaling pathway. Int. J. Mol. Med. 2018, 42, 453–460. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Development of tubule-interstitial damage in the hypertensive disease. When angiotensin II and/or endothelin-1 are abnormally produced, such as under conditions characterized by excess mechanical or oxidative stress, epithelial-to-mesenchymal transition (EMT) is triggered, leading to transformation of tubular cells into myofibroblasts that produce collagens. The final event is tubulo-interstitial fibrosis. Angiotensin II (Ang II), by binding Ang II type 1 (AT1) receptors in the vascular smooth muscle cells, also favors vasoconstriction and vascular remodeling, worsening kidney damage. Vasoconstriction is also mediated by endothelin-1 via ETA and ETB receptors, located at the vascular smooth muscle cells.
Figure 1. Development of tubule-interstitial damage in the hypertensive disease. When angiotensin II and/or endothelin-1 are abnormally produced, such as under conditions characterized by excess mechanical or oxidative stress, epithelial-to-mesenchymal transition (EMT) is triggered, leading to transformation of tubular cells into myofibroblasts that produce collagens. The final event is tubulo-interstitial fibrosis. Angiotensin II (Ang II), by binding Ang II type 1 (AT1) receptors in the vascular smooth muscle cells, also favors vasoconstriction and vascular remodeling, worsening kidney damage. Vasoconstriction is also mediated by endothelin-1 via ETA and ETB receptors, located at the vascular smooth muscle cells.
Ijms 20 03567 g001
Table 1. PICO strategy used for the literature search.
Table 1. PICO strategy used for the literature search.
Kew WordOperatorKew WordOperatorKew Word
HypertensionORHigh blood pressure
AND
NephropathyORTubular damageORTubulointerstitial fibrosis
AND
Epithelial-to-mesenchymal transitionOREMTOREpithelial-to-mesenchymal transdifferentiation

Share and Cite

MDPI and ACS Style

Seccia, T.; Caroccia, B.; Piazza, M.; Rossi, G.P. The Key Role of Epithelial to Mesenchymal Transition (EMT) in Hypertensive Kidney Disease. Int. J. Mol. Sci. 2019, 20, 3567. https://doi.org/10.3390/ijms20143567

AMA Style

Seccia T, Caroccia B, Piazza M, Rossi GP. The Key Role of Epithelial to Mesenchymal Transition (EMT) in Hypertensive Kidney Disease. International Journal of Molecular Sciences. 2019; 20(14):3567. https://doi.org/10.3390/ijms20143567

Chicago/Turabian Style

Seccia, Teresa, Brasilina Caroccia, Maria Piazza, and Gian Paolo Rossi. 2019. "The Key Role of Epithelial to Mesenchymal Transition (EMT) in Hypertensive Kidney Disease" International Journal of Molecular Sciences 20, no. 14: 3567. https://doi.org/10.3390/ijms20143567

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop