devil’s triangle in kidney diseases: oxidative stress...

60
International Journal of Nephrology Devil’s Triangle in Kidney Diseases: Oxidative Stress, Mediators, and Inflammation Guest Editors: Ays ˛ e Balat, Halima Resic, Guido Bellinghieri, and Ali Anarat

Upload: others

Post on 18-Aug-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology

Devil’s Triangle in Kidney Diseases: Oxidative Stress, Mediators, and Inflammation

Guest Editors: Ays e Balat, Halima Resic, Guido Bellinghieri, and Ali Anarat

Page 2: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Devil’s Triangle in Kidney Diseases: OxidativeStress, Mediators, and Inflammation

Page 3: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology

Devil’s Triangle in Kidney Diseases: OxidativeStress, Mediators, and Inflammation

Guest Editors: Ayse Balat, Halima Resic, Guido Bellinghieri,and Ali Anarat

Page 4: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Copyright © 2012 Hindawi Publishing Corporation. All rights reserved.

This is a special issue published in “International Journal of Nephrology.” All articles are open access articles distributed under theCreative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided theoriginal work is properly cited.

Page 5: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Editorial Board

Anil K. Agarwal, USAZiyad Al-Aly, USAAlessandro Amore, ItalyFranca Anglani, ItalyNuket Bavbek, TurkeyRichard Fatica, USASimin Goral, USAW. H. Horl, Austria

Kamel S. Kamel, CanadaKazunari I. Kaneko, JapanDavid B. Kershaw, USAAlejandro Martın-Malo, SpainTej Mattoo, USATibor Nadasdy, USAFrank Park, USAJochen Reiser, USA

Laszlo Rosivall, HungaryMichael J. Ross, USAJames E. Springate, USAVladimır Tesar, Czech RepublicGreg Tesch, AustraliaSuresh C. Tiwari, IndiaJaime Uribarri, USA

Page 6: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Contents

Devil’s Triangle in Kidney Diseases: Oxidative Stress, Mediators, and Inflammation, Ayse Balat,Halima Resic, Guido Bellinghieri, and Ali AnaratVolume 2012, Article ID 156286, 2 pages

Urotensin-II: More Than a Mediator for Kidney, Ayse Balat and Mithat BuyukcelikVolume 2012, Article ID 249790, 7 pages

Antioxidants in Kidney Diseases: The Impact of Bardoxolone Methyl, Jorge Rojas-Rivera, Alberto Ortiz,and Jesus EgidoVolume 2012, Article ID 321714, 11 pages

Induction of Oxidative Stress in Kidney, Emin OzbekVolume 2012, Article ID 465897, 9 pages

Inflammation and Oxidative Stress in Obesity-Related Glomerulopathy, Jinhua Tang, Haidong Yan,and Shougang ZhuangVolume 2012, Article ID 608397, 11 pages

Reactive Oxygen Species Modulation of Na/K-ATPase Regulates Fibrosis and Renal Proximal TubularSodium Handling, Jiang Liu, David J. Kennedy, Yanling Yan, and Joseph I. ShapiroVolume 2012, Article ID 381320, 14 pages

Page 7: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Hindawi Publishing CorporationInternational Journal of NephrologyVolume 2012, Article ID 156286, 2 pagesdoi:10.1155/2012/156286

Editorial

Devil’s Triangle in Kidney Diseases: Oxidative Stress,Mediators, and Inflammation

Ayse Balat,1 Halima Resic,2 Guido Bellinghieri,3 and Ali Anarat4

1 Division of Pediatric Nephrology, Faculty of Medicine, University of Gaziantep, PK 34, 27310 Gaziantep, Turkey2 Division of Nephrology, Clinical Center University of Sarajevo, Sarajevo, Bosnia and Herzegovina3 Nephrology and Dialysis Unit and Postgraduate School of Nephrology, Messina University, Via Consolare Pompea No. 1867 Ganzirri,98165 Messina, Italy

4 Department of Pediatric Nephrology, Faculty of Medicine, University of Cukurova, 01330 Adana, Turkey

Correspondence should be addressed to Ayse Balat, [email protected]

Received 13 November 2012; Accepted 13 November 2012

Copyright © 2012 Ayse Balat et al. This is an open access article distributed under the Creative Commons Attribution License,which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

This issue of the International Journal of Nephrology focusedon kidney diseases within a devil’s triangle, oxidative stress(OS), mediators, inflammation, specifically relating to theclinical significance of identification, and prevention.

Every creature in need of oxygen faces OS. It has a criticalrole in the molecular mechanisms of renal injury in severalkidney diseases, and many complications of these diseasesare mediated by OS, mediators, and inflammation. Thereis a complex relationship between these three; mostly theyinduce each other. While some of the diseases themselves cancontribute to OS, reactive oxygen species (ROS) producedby activated leukocytes and endothelial cells in sites ofinflammation cause tissue damage. Although inflammationlooks dangerous for the organism, it is a normal reactionof organs and tissues to protect themselves against severalinvasion(s). It enables the immune system to remove theinjurious stimuli and initiate the healing process of tissues.However, the interactions between OS, mediators, andinflammation may result in glomerular damage, proteinuria,electrolyte, and volume instabilities which cause nephronloss, on the long view. Detailed studies on this topic areincluded in this issue.

The kidney can easily be damaged by ROS, due to therich structure of long-chain polyunsaturated fatty acids.The article by E. Ozbek summarizes the induction of OSwithin kidney in several conditions, including diabetes,hypertension, hypercholesterolemia, obesity, aging, urinaryobstruction, environmental toxins, and molecular mecha-nisms of these inductions in the light of existing literaturedata.

Diabetic nephropathy is one of the most commonmicrovascular complications of type 1 and type 2 diabetesmellitus and the leading cause of end-stage renal diseaseworldwide [1].

Rojas-Rivera et al. reviewed the biological bases ofoxidative stress and its role especially on diabetic nephropa-thy, as well as the role of the Keap1-Nrf2 pathway, andrecent clinical trials targeting this pathway with bardoxolonemethyl, a novel synthetic triterpenoid with antioxidant andanti-inflammatory properties.

Obesity continues to be a public health problem through-out the world. Epidemiologic studies have shown that 66% ofadults and 16% of children and adolescents are overweight orobese [2]. Obesity-related glomerulopathy is an increasingcause of end-stage renal diseases. J. Tang et al. stressed thechronic low-grade systemic inflammation in obesity anddiscussed the roles of inflammation and oxidative stressin the progression of obesity-related glomerulopathy andpossible treatment modalities to prevent kidney injury inobesity, such as the usage of anti-IL-6 receptor antibody,TNF-α antagonist, adiponectin, nutritional and surgicalinterventions to reduce OS.

Hypertension is an another important global healthissue both in adults and children. It is one of the majorrisk factors for the progression of kidney diseases. Therelationship between blood pressure and dietary sodiumand salt sensitivity has been well known, and renal sodiumhandling is a key determinant of long-term blood pressureregulation [3]. There is a limited knowledge in the literatureregarding the role of ROS-mediated fibrosis and renal

Page 8: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

2 International Journal of Nephrology

proximal tubule sodium reabsorption through the Na/K-ATPase. S. Liu et al. reviewed the possible role of ROSin the regulation of Na/K-ATPase activity. The authorsemphasized the importance of further researches whetherROS signaling is a link between the Na/K-ATPase/c-Srccascade and NHE3 regulation and how OS, stimulated byhigh salt and cardiotonic steroids, regulates Na/K-ATPase/c-Src signaling in renal sodium handling and fibrosis.

Urotensin-II is the most potent mammalian vasocon-strictor identified to date, almost tenfold more potent thanendothelin-I [4]. A. Balat and M. Buyukcelik discussedthe role of urotensin-II on renal hemodynamics and itspossible role on several kidney diseases, such as the minimalchange nephrotic syndrome. The article includes a detaileddiscussion of urotensin-II immunoreactivity in renal biopsyspecimens of children with membranoproliferative glomeru-lonephritis, membranous nephropathy, IgA nephropathy,Henoch-Schonlein nephritis, and focal segmental glomeru-losclerosis. Because of its complex relation with OS andother mediators, authors describe it as “more than amediator” in glomerular diseases. They briefly mention fromthe effectiveness of U-II antagonism, as a new promisingpharmacological treatment target in some kidney diseases.

Given the potential impact of OS, mediators, andinflammation trio, the importance of prevention has comeinto question. Strong evidence indicates the importance ofnew molecules that are able to diminish them which inturn may help to decrease the prevalence and/or progressionof several kidney diseases. Therefore, further researches areneeded to the better understanding of the molecular andclinical mechanisms of this triad. They may help to providenew therapeutical strategies to control several complicationsin patients with kidney diseases.

Ayse BalatHalima Resic

Guido BellinghieriAli Anarat

References

[1] A. A. Lopes, “End-stage renal disease due to diabetes inracial/ethnic minorities and disadvantaged populations.,” Eth-nicity & Disease, vol. 19, supplement 1, no. 1, pp. S47–S51, 2009.

[2] Y. Wang and M. A. Beydoun, “The obesity epidemic inthe United States–gender, age, socioeconomic, racial/ethnic,and geographic characteristics: a systematic review and meta-regression analysis,” Epidemiologic Reviews, vol. 29, no. 1, pp.6–28, 2007.

[3] A. C. Guyton, “Blood pressure control–special role of thekidneys and body fluids,” Science, vol. 252, no. 5014, pp. 1813–1816, 1991.

[4] J. J. Maguire and A. P. Davenport, “Is urotensin-II the newendothelin?” British Journal of Pharmacology, vol. 137, no. 5,pp. 579–588, 2002.

Page 9: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Hindawi Publishing CorporationInternational Journal of NephrologyVolume 2012, Article ID 249790, 7 pagesdoi:10.1155/2012/249790

Review Article

Urotensin-II: More Than a Mediator for Kidney

Ayse Balat1, 2 and Mithat Buyukcelik1

1 Division of Pediatric Nephrology, Faculty of Medicine, University of Gaziantep, 27310 Gaziantep, Turkey2 Division of Pediatric Nephrology and Rheumatology, Faculty of Medicine, University of Gaziantep, 27310 Gaziantep, Turkey

Correspondence should be addressed to Ayse Balat, [email protected]

Received 30 November 2011; Accepted 6 September 2012

Academic Editor: Ali Anarat

Copyright © 2012 A. Balat and M. Buyukcelik. This is an open access article distributed under the Creative Commons AttributionLicense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properlycited.

Human urotensin-II (hU-II) is one of the most potent vasoconstrictors in mammals. Although both hU-II and its receptor, GPR14,are detected in several tissues, kidney is a major source of U-II in humans. Recent studies suggest that U-II may have a possibleautocrine/paracrine functions in kidney and may be an important target molecule in studying renal pathophysiology. It has severaleffects on tubular transport and probably has active role in renal hemodynamics. Although it is an important peptide in renalphysiology, certain diseases, such as hypertension and glomerulonephritis, may alter the expression of U-II. As might be expected,oxidative stress, mediators, and inflammation are like a devil’s triangle in kidney diseases, mostly they induce each other. Sincethere is a complex relationship between U-II and oxidative stress, and other mediators, such as transforming growth factor β1 andangiotensin II, U-II is more than a mediator in glomerular diseases. Although it is an ancient peptide, known for 31 years, it lookslike that U-II will continue to give new messages as well as raising more questions as research on it increases. In this paper, wemainly discuss the possible role of U-II on renal physiology and its effect on kidney diseases.

1. Introduction

Although urotensin-II was firstly identified in a neurohemalorgan of fish in 1980s [1, 2], only recently it became a majorfocus of clinical and experimental researches [3].

Human urotensin-II (hU-II) is a cyclic peptide of 11amino acids cleaved from a larger prepro-U-II precursorpeptide of about 130 amino acids [1, 3, 4]. The gene encodingthis peptide is located at 1p36 and contains 5 exons [5]. It isa ligand for the orphan G-protein-coupled receptor, knownas GPR14 [3, 6, 7].

Although human prepro-U-II mRNA is expressed mainlyin the brain and spinal cord [4, 8], both hU-II and itsreceptor are also detected in other organs and tissues, such askidney, spleen, smooth muscle, endothelium, small intestine,thymus, prostate, pituitary, and adrenal gland [1, 9–11].

Being almost tenfold more potent than endothelin-I(ET), it is the most potent mammalian vasoconstrictoridentified to date [3, 12]. It circulates in the plasma of healthyindividuals and acts as a circulating vasoactive hormoneand as a locally acting paracrine or autocrine factor incardiovascular regulation [4, 13].

Although it mainly has a vasoconstrictor effect, regionaldifferences may be seen in its effects in various vascularbeds and blood vessels of some species. For example, it hasa vasodilatory effect on the small arteries of rats [14, 15],and on the resistance arteries of humans, through release ofendothelium-derived hyperpolarizing factor (EDHF), nitricoxide (NO) [14–16].

It has been shown that the potent vasoconstrictoractions of U-II is mediated by Ca+2 mobilization throughactivation of a number of signaling pathways including Ca+2

channels, tyrosine kinase, p38 mitogen-activated proteinkinase (p38MAPK), and extracellular signal-regulated kinase1 and 2 (ERK1/2) [17, 18].

Since U-II and its receptor have been demonstrated inmouse, monkey [19], and human kidneys [20, 21], it isacceptable to consider that U-II is synthesized, secreted, andcleared by the kidneys [6, 22–24].

Interestingly, Mosenkis et al. [25] showed that hU-II wasalso present in 2 surgically anephric subjects. Although thisfinding is inconsistent with the conclusion that the kidneysare the primary source for production of U-II, as the authorsstressed, the high density of U-II and its receptor in renal

Page 10: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

2 International Journal of Nephrology

tissues suggest that U-II is metabolically active in the kidneyeven though it is produced in outside of the kidneys.

In this paper, we mainly discuss the possible role of U-IIon renal physiology and its effect on kidney diseases.

2. Effect of U-II on Renal Hemodynamics

Because of its potent vasoconstrictor effect, U-II attractedthe interest of researchers in general hemodynamy. In fact,hemodynamic responses to U-II show regional heterogeneityin relation to its receptor localization, even in the differencesof functional state of the endothelium [26].

However, in contrast to animal studies, Wilkinson et al.found no vasoactive responses to hU-II in vivo in man[27]. They injected hU-II intra-arterially to healthy malevolunteers, and despite the high-circulating hU-II levels,no change was seen in systemic hemodynamics, ECGs ofsubjects, and hU-II had no effect on hand vein diameter.

However, another study published in the same year [28]demonstrated that U-II produces potent vasoconstrictionin humans in vivo. They showed that U-II induced dose-dependent reduction in forearm blood flow (FBF) of healthyvolunteers, and FBF returned to baseline values within30 min.

Known data in the literature show that the kidney is amajor source of U-II in humans [29], primates, mice [19],and rats [11]. It has been found in the urine of humans[22, 24] and rats [11] at a concentration far exceeding thatof plasma. In humans the renal clearance of U-II has foundto be greater than that of creatinine, suggesting that urinaryU-II is derived primarily from the kidney [22]. An animalstudy also revealed that there was an arteriovenous con-centration gradient for U-II across the renal circulation inanaesthetised sheep [30]. As well as U-II, its receptor has alsobeen localized to the mammalian kidneys, such as in human[22], monkey, mouse [19], and rat kidneys [11]. It has beenshown that the medulla, especially tubular component of thekidney, is the principal site of U-II receptor expression in therat kidney [11, 31].

Shenouda et al. [20] demonstrated that U-II was mostlypresent in the epithelial cells of tubules and ducts, withgreater intensity in the distal convoluted tubules in normalhuman kidneys. Moderate U-II immunoreactivity was seenin the endothelial cells of renal capillaries, but only focalimmunoreactivity was found in the endothelial cells of theglomeruli. We also observed similar results in kidneys ofchildren [21], and these findings suggest that hU-II maycontribute to the pathophysiology of human kidneys (Figure1(a)) [21, 32].

Expression of U-II mostly in tubules may suggest itsprobable active role in renal hemodynamics. Loretz andBern [33] demonstrated that U-II stimulated Na transport inthe teleost urinary bladder, while Ovcharenko et al.’s study[34] indicated that short-term adminisration of U-II didnot influence sodium (Na) handling by the kidney in rats.However, Zhang et al. [35] observed that infusion of U-IIdirectly into the rat renal artery increased renal blood flow(RBF), associated with a diuresis and natriuresis. In contrast,

Song et al. [11] reported that U-II caused an antinatriuresisand antidiuresis when administered as an i.v. bolus doseand stressed that this was associated with and driven byrenal hemodynamic effects leading to a marked reduction inglomerular filtration rate (GFR).

Two years later from the above study, Abdel-Razik et al.[36] searched the potential direct tubular action of urotensinin rats. They observed dose-dependent changes in GFR andurinary electrolytes. The hemodynamic effects were predom-inated at higher doses and caused a profound reduction inGFR which was accompanied by an antidiuresis and anti-natriuresis. When a lower infusion rate of rat U-II wasemployed, a tubular action to reduce electrolyte reabsorptionbecame apparent through an increase in fractional excretionof Na and potassium (K) [36].

However, in children with minimal change nephroticsyndrome (MCNS), and their healthy controls, we couldnot find any relationship between the U-II level and Na/Kexcretion [24]. Although this differences may be partiallyrelated to different biological effects of U-II in differentspecies, contradictory observations in same species underlinethe complex influence of U-II on renal hemodynamics.

It is not clear enough whether the effect of U-IIon tubular transport is direct or mediated by secondarymechanisms. However, considering the expression of U-IIreceptor in the thin ascending limb of Henle’s loop andthe inner medullary collecting duct [11] and the greaterU-II receptor mRNA and protein expression in the medullacompared with the cortex [36], together with the abundantexpression of hU-II in the proximal and distal tubules ofchildren (patients and controls) [21], it may be suggestedthat U-II may have a direct action on tubular electrolytetransport.

3. The Effect of Urotensin-II on Kidney Diseases

Since U-II and its receptor, GPR14, are expressed abundantlyin cardiorenal system [10, 11], most of the researches on it arerelated to cardiovascular and renal diseases.

Although some studies have been investigated the circu-lating levels of U-II in several diseases, such as hypertension[37], congestive heart failure [38], renal failure [3], MCNS[24], and preeclampsia-eclampsia [39], little is known aboutthe actions of this important peptide within the kidney.Some studies suggest that renal dysfunction affects theU-II levels, since the plasma U-II level has been foundelevated in renal failure [3], congestive heart failure [38], andsystemic hypertension [37], and it was found to be an inversepredictor of overall and cardiovascular mortality in patientswith moderate-to-severe chronic kidney disease (CKD)[40].

Certain diseases, such as hypertension and glomeru-lonephritis, may alter the expression of U-II. It has beenshown that both U-II and its receptor mRNA expression lev-els were up to threefold higher in spontaneously hypertensiverat (SHR) tissue compared to control Wistar-Kyoto (WKY)rats, taking into consideration that SHR is more sensitivethan WKY to the effect of U-II [41].

Page 11: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 3

(a) (b)

(c) (d)

Figure 1: (a) Localization of urotensin-II (U-II) immunoreactivity (brown color) in the normal human kidney. Weak immunostaining inglomerulus, abundant expression of U-II in tubules. Thick arrows: weak immunostaining and black arrow: strong immunostaining [21].(b) Urotensin-II immunoreactivity in membranoproliferative glomerulonephritis. Immunostaining in glomerular basement membrane,mesangium, Bowman capsule, and tubules. GBM: glomerular basement membrane, MPGN: membranoproliferative glomerulonephritis,BC: Bowman capsule, M: mesangium, PT: proximal tubule, and DT: distal tubule [21]. (c) Localization of urotensin-II immunoreactivity ina crescent [21]. (d) Urotensin-II immunoreactivity in focal segmental glomerulosclerosis. Notice the abundant expression of U-II in scleroticarea [32].

In the literature, there are no enough data on the levelof this vasoactive peptide in glomerular diseases. Recently,we firstly demonstrated that U-II was present in plasma andurine samples of 26 children with MCNS [24]. It showedimportant changes in relapse and remission periods. PlasmaU-II concentrations during relapse were significantly lowerthan in remission and in controls, whereas urinary U-II levelswere higher in relapse than in remission [24]. The plasmaU-II level showed a significant positive correlation with theplasma albumin concentration during remission. However,there was no correlation between the amount of proteinuriaand plasma/urinary U-II levels, and we could not detectany relationship between U-II levels and other clinical andlaboratory parameters (such as the age at onset of disease,number of relapses, time to remission, blood pressure, serumcreatinine, and hematological parameters). We suggestedthat the important changes in plasma and urinary U-II levelsduring relapse may be the result of heavy proteinuria ratherthan playing a role in mediating the clinical and laboratorymanifestations of MCNS. After this, it would be interestingto search the possible role(s) of this peptide in childrenwith glomerular diseases other than MCNS. Therefore, weexamined the urotensin-II immunoreactivity in renal biopsyspecimens of children with several renal diseases, includingmembranoproliferative glomerulonephritis (MPGN), mem-branous nephropathy (MGN), IgA nephropathy (IgAN),Henoch-Schonlein nephritis (HSN), and focal segmental

glomerulosclerosis (FSGS) [21, 32]. In normal human kid-ney, there was weak expression of human U-II in glomerulus,while abundant expressions were seen in proximal, distaltubules, and collecting ducts (Figure 1(a)) [21], similar to aprevious study [20].

We observed different expression pattern of U-II indifferent glomerular diseases. In MPGN and FSGS, differentfrom the normal kidneys, more dense U-II immunoreactivitywas seen in the glomerular basement membrane (GBM),glomerular mesangium, Bowman capsule (BC), and tubules(Figures 1(b), and 1(d)) [21, 32]. Interestingly, we alsoobserved U-II immunoreactivity in crescents (Figure 1(c)),and sclerotic areas in FSGS (Figure 1(d)) [21, 32].

Systolic blood pressure (BP) was positively correlatedwith mesangial expression of U-II (r = 0.418, P = 0.042),while diastolic BP was correlated with endothelial U-IIexpression in MPGN (r = 0.469, P = 0.021) [21].

In children with MGN, U-II was mostly seen in GBMand BC. We observed more dense U-II immunoreactivity indistal tubules (P = 0.030), endothelium (P = 0.009), andmesangium (P = 0.002) in children with MPGN than inMGN. Diastolic BP was positively correlated with the expres-sion of U-II in BC in children with MGN (r = 1, P = 0.000)[21].

There is no enough data about the precise role of hU-II inrenal diseases, and that was the first report demonstrating thepresence of U-II by immunohistochemically in children with

Page 12: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

4 International Journal of Nephrology

several renal diseases, suggesting that hU-II may contributeto the pathophysiology of human kidneys.

The positive correlation between BP and intensity ofU-II expression in mesangium and endothelium in MPGN,and BC in MGN was noteworthy. Considering the literaturedata about U-II, as an important physiological mediator ofvascular tone and blood pressure in humans [16], and alsoan extremely potent constrictor of renal blood vessels fromprimates [6], it is reasonable to suggest that U-II may play animportant role in the regulation of BP in MPGN and MGN.

As it has been known, mainly, two basic mechanismsare feasible in glomerulonephritis: antibody interaction withantigens in situ within the glomerulus and antibody bindingto soluble antigens in the circulation, followed by immune-complex deposition within the glomeruli [42]. The sec-ondary immune mechanisms of glomerular injury are thecascade of inflammatory mediators that are recruited topropagate renal damage following the primary glomerularattack. Some of these mediators play essential roles, whereasothers may aggravate the glomerular lesion [42]. Most of thesecondary mediators include cytokines, growth factors, reac-tive oxygen metabolites, bioactive lipids (platelet-activatingfactor and eicosanoids), proteases, and vasoactive substances,such as ET and NO [42].

Since U-II is abundantly expressed in the glomeruli inMPGN and MGN, it is reasonable to suggest that U-II mayplay a role in this mechanism, probably in the secondaryimmune mechanisms of glomerular injury, by a paracrine oran endocrine action [21]. Djordjevic et al. [43] demonstratedthat hU-II increases the levels of NADPH oxidase-derivedreactive oxygen species, leading to the activation of mitogen-activated protein kinases and protein kinase B (akt), followedby enhanced plasminogen activator inhibitor-1 expressionand increased proliferation of pulmonary arterial smoothmuscle cells. It has been also shown that exposure of the ratproximal tubular epithelial cells (NRK-52E) to transforminggrowth factor β1 (TGF-β1) or angiotensin II (Ang II)increased U-II and GPR14 mRNA expressions [44], andU-II acts synergistically with Ang II [45, 46]. As might beexpected, oxidative stress, mediators, and inflammation arelike a devil’s triangle in kidney diseases, mostly they induceeach other. Since there is a complex relationship betweenU-II and oxidative stress [43], and other mediators, suchas TGF-β1 and Ang II [44–46], U-II is probably more thana mediator in glomerular diseases and takes place in animportant part of this devil’s triangle.

Interestingly, we observed abundant U-II immunoreac-tivity in crescents and sclerotic areas in FSGS [21, 32]. Cres-cents are composed of large swollen cells arising from bothmacrophages of hematogenous origin and native parietalepithelial cells [47]. As time elapses, the cellular crescents areprogressively replaced by fibroblasts, and in more advancedstages, the fibroblastic component is entirely replaced bycollagenous lamellar materials with a few remnant cells [48].Recent reports have shown a mitogenic role for U-II throughinduction of smooth muscle cell proliferation [49, 50], andadditionally, it has been shown to induce collagen depositionby fibroblasts [51]. Zhang et al. [52] showed that U-IIcould stimulate the phenotypic conversion, migration, and

collagen synthesis in adventitial fibroblasts. Additionally, itmay act as autocrine/paracrine growth stimulators in tumorcells [53].

The pathogenesis of glomerulosclerosis is still unknown.Several factors, cytokines and growth factors, hyperlipidemiaand platelet activation, lead to an increase of mesangialmatrix production by resident cells. Several data demon-strate that abnormal glomerular growth is associated withglomerular sclerosis [54].

Since hU-II stimulates cell proliferation in adrenaltumors [55], renal epithelial cells [23], and vascular smoothmuscle cells [49] and it has been found elevated in carotidand aortic atherosclerotic plaques [56], the abundant expres-sion of U-II in crescents and sclerotic areas suggests thatU-II may also play a role in the progression of crescents andglomerular sclerosis, probably as a growth factor or as aninflammatory peptide. This hypothesis must be searched andtested in future.

MGN is an antibody-mediated disease of uncertain andimprecise pathogenesis. However, the hypotheses that it is anautoimmune disease of the kidney and that the subepithelialimmune deposits are formed in situ with an endogenousglomerular antigen are attractive [57]. The electron-densedeposits are generally located at the site of the slit diaphragm,and subepithelial space, while no electron-dense depositsare seen in the subendothelial space or in the mesangium,and hypertension at onset is associated with a less favorableoutcome in MGN [57]. In our study, hU-II expressionwas mostly seen in GBM and BC, and there was a strongpositive correlation between diastolic BP and intensity ofU-II expression in BC. These findings may increase twointeresting questions: may U-II play a role in the formationof these deposits as a mitogenic factor, as we mentionedpreviously, and may it have any role in the clinical course ofMGN by regulating the BP? However, it is difficult to answerthese questions with that study, and these hypothesis must beclarified by further detailed studies.

In kidneys of children with HSN and IgAN, similar toeach other, more dense U-II immunoreactivity was seen inGBM, glomerular mesangium, BC, proximal/distal tubules,and also in crescents [21, 32].

Although the pathogenesis of IgAN and HSN is not wellknown [58], animal studies have shown the key role ofcytokines and growth factors (particularly platelet-derivedgrowth factor and TGF-β) in the induction and resolutionof mesangial injury, and there is some evidence that these arealso involved in IgAN [58]. The similar expression patternof U-II in HSN and IgAN has been considered that U-IImay have a role in mesangial inflammation and crescentformation in these disorders [32].

Different expression pattern of U-II in several renaldiseases may give rise to thought whether the effect ofU-II gene polymorphism. Recently, we performed a prelim-inary study, and firstly investigated the possible associationbetween a coding single nucleotide polymorphism of UT-IIgene, T21M (T/C), in 87 children with childhood nephroticsyndrome (NS), 16 children with acute poststreptococcalglomerulonephritis (APSGN), and 10 children with HSN[59]. We found higher TC genotype of U-II gene in NS

Page 13: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 5

(56.3% versus 38.9%, P = 0.025), higher TT genotype inAPSGN (50.0% versus 25.9%, P < 0.001), and a positivecorrelation between TT polymorphism and the presence ofmacroscopic hematuria in APSGN (r = 0.51, P = 0.04).This study considered that urotensin-II may be an importantmediator in pathophysiology of the childhood glomeru-lonephritis, and Turkish children with TC genotype may havea higher genetic susceptibility to NS, while TT genotype ofU-II may increase the risk of APSGN.

4. Urotensin-II Antagonists as a New PromisingPharmacological Treatment Target

Several influences of U-II in cardiovascular/renal system,and the presence of its receptor in the heart, lungs, bloodvessels, kidneys, and brain, led the researchers to investigatethe role of U-II antagonists in various diseases. The mostknown U-II receptor antagonists are palosuran and urantide.Sidharta et al. [60] investigated whether palosuran, a potent,selective, and competitive antagonist of the U-II receptor,had effects in macroalbuminuric, diabetic patients who areprone to the development of renal disease. They observedan overall clinically significant reduction of 24.3% in the 24-hour urinary albumin excretion rate.

In an experimental study, it has been shown that long-term treatment of streptozotocin-induced diabetic rats withpalosuran improved survival, increased insulin, and slowedthe increase in glycemia, glycosylated hemoglobin, andserum lipids. Furthermore, palosuran increased renal bloodflow and delayed the development of proteinuria and renaldamage [61].

These two researches suggest that U-II receptor antago-nism might be a new therapeutic approach for the treatmentand/or prevention of diabetic nephropathy.

The tolerability and safety, pharmacokinetics, and phar-macodynamics of palosuran were evaluated in also healthyyoung men with a double-blinded placebo-controlled singleascending dose designed study [62]. It has been shown thatpalosuran was well tolerated, and no serious adverse eventsor dose-related adverse events were reported. However, as theauthors stressed, the results of this entry-into-humans studywarrant further investigation of the therapeutic potential ofpalosuran.

Recently, Nitescu et al. [63] examined the effects ofanother selective U-II receptor antagonist, urantide, on renalhemodynamics, oxygenation, and function in endotoxemicrats. However, they found that urantide had no statisticallysignificant effects on any of the investigated variables (kidneyfunction, renal blood flow, cortical and outer medullaryperfusion, and oxygen tension) in these rats.

In spite of different results about the effectiveness of U-IIantagonism, it appears that the therapeutic potential of U-IIantagonists may be the focus of research interest in the nearfuture.

In summary, U-II may be an important mediator, in factprobably more than a mediator, in kidney diseases. Whetherthe observed findings which are primary or secondaryto these pathological conditions still remain unclear, they

suggest a possible role of U-II in the pathophysiology ofseveral kidney diseases. It looks like that U-II will continueto give new messages as well as raising more questions asresearch on it increases. Further, detailed studies are neededto address the exact role(s) of this peptide in renal diseases.

References

[1] D. Pearson, J. E. Shively, and B. R. Clark, “Urotensin II: asomatostatin-like peptide in the caudal neurosecretory systemof fishes,” Proceedings of the National Academy of Sciences of theUnited States of America, vol. 77, no. 8 I, pp. 5021–5024, 1980.

[2] H. A. Bern, D. Pearson, B. A. Larson, and R. S. Nishioka,“Neurohormones from fish tails: the caudal neurosecretorysystem—I. “Urophysiology” and the caudal neurosecretorysystem of fishes,” Recent Progress in Hormone Research, vol. 41,pp. 533–552, 1985.

[3] K. Totsune, K. Takahashi, Z. Arihara et al., “Role of urotensinII in patients on dialysis,” Lancet, vol. 358, no. 9284, pp. 810–811, 2001.

[4] Y. Coulouarn, I. Lihrmann, S. Jegou et al., “Cloning of thecDNA encoding the urotensin II precursor in frog andhuman reveals intense expression of the urotensin II gene inmotoneurons of the spinal cord,” Proceedings of the NationalAcademy of Sciences of the United States of America, vol. 95, no.26, pp. 15803–15808, 1998.

[5] K. L. Ong, L. Y. F. Wong, Y. B. Man et al., “Haplotypes in theurotensin II gene and urotensin II receptor gene are associatedwith insulin resistance and impaired glucose tolerance,”Peptides, vol. 27, no. 7, pp. 1659–1667, 2006.

[6] R. S. Ames, H. M. Sarau, J. K. Chambers et al., “Humanurotensin-II is a potent vasoconstrictor and agonist for theorphan receptor GPR14,” Nature, vol. 401, no. 6750, pp. 282–286, 1999.

[7] Q. Liu, S. S. Pong, Z. Zeng et al., “Identification of urotensinII as the endogenous ligand for the orphan G-protein-coupledreceptor GPR14,” Biochemical and Biophysical Research Com-munications, vol. 266, no. 1, pp. 174–178, 1999.

[8] J. Gartlon, F. Parker, D. C. Harrison et al., “Central effectsof urotensin-II following ICV administration in rats,” Psy-chopharmacology, vol. 155, no. 4, pp. 426–433, 2001.

[9] D. Onan, R. D. Hannan, and W. G. Thomas, “Urotensin II: theold kid in town,” Trends in Endocrinology and Metabolism, vol.15, no. 4, pp. 175–182, 2004.

[10] J. J. Maguire, R. E. Kuc, and A. P. Davenport, “Orphan-receptor ligand human urotensin II: receptor localizationhuman tissues and comparison of vasoconstrictor responseswith endothelin-1,” British Journal of Pharmacology, vol. 131,no. 3, pp. 441–446, 2000.

[11] W. Song, A. E. S. Abdel-Razik, W. Lu et al., “Urotensin II andrenal function in the rat,” Kidney International, vol. 69, no. 8,pp. 1360–1368, 2006.

[12] J. J. Maguire and A. P. Davenport, “Is urotensin-II the newendothelin?” British Journal of Pharmacology, vol. 137, no. 5,pp. 579–588, 2002.

[13] J. J. Maguire, R. E. Kuc, K. E. Wiley, M. J. Kleinz, and A. P. Dav-enport, “Cellular distribution of immunoreactive urotensin-II in human tissues with evidence of increased expressionin atherosclerosis and a greater constrictor response of smallcompared to large coronary arteries,” Peptides, vol. 25, no. 10,pp. 1767–1774, 2004.

[14] F. E. Bottrill, S. A. Douglas, C. R. Hiley, and R. White, “Humanurotensin-II is an endothelium-dependent vasodilator in rat

Page 14: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

6 International Journal of Nephrology

small arteries,” British Journal of Pharmacology, vol. 130, no. 8,pp. 1865–1870, 2000.

[15] J. Qi, J. Du, X. Tang, J. Li, B. Wei, and C. Tang, “The upregu-lation of endothelial nitric oxide synthase and urotensin-II isassociated with pulmonary hypertension and vascular diseasesin rats produced by aortocaval shunting,” Heart and Vessels,vol. 19, no. 2, pp. 81–88, 2004.

[16] J. Affolter and D. J. Webb, “Urotensin II: a new mediator incardiopulmonary regulation?” Lancet, vol. 358, no. 9284, pp.774–775, 2001.

[17] U. Lehner, A. Velic, R. Schroter, E. Schlatter, and A. Sindic,“Ligands and signaling of the G-protein-coupled receptorGPR14, expressed in human kidney cells,” Cellular Physiologyand Biochemistry, vol. 20, no. 1–4, pp. 181–192, 2007.

[18] C. D. Proulx, B. J. Holleran, P. Lavigne, E. Escher, G.Guillemette, and R. Leduc, “Biological properties and func-tional determinants of the urotensin II receptor,” Peptides, vol.29, no. 5, pp. 691–699, 2008.

[19] N. A. Elshourbagy, S. A. Douglas, U. Shabon et al., “Molec-ular and pharmacological characterization of genes encodingurotensin-II peptides and their cognate G-protein-coupledreceptors from the mouse and monkey,” British Journal ofPharmacology, vol. 136, no. 1, pp. 9–22, 2002.

[20] A. Shenouda, S. A. Douglas, E. H. Ohlstein, and A. Giaid,“Localization of urotensin-II immunoreactivity in normalhuman kidneys and renal carcinoma,” Journal of Histochem-istry and Cytochemistry, vol. 50, no. 7, pp. 885–889, 2002.

[21] A. Balat, M. Karakok, K. Yilmaz, and Y. Kibar, “Urotensin-IIimmunoreactivity in children with chronic glomerulonephri-tis,” Renal Failure, vol. 29, no. 5, pp. 573–578, 2007.

[22] M. Matsushita, M. Shichiri, T. Imai et al., “Co-expression ofurotensin II and its receptor (GPR14) in human cardiovascu-lar and renal tissues,” Journal of Hypertension, vol. 19, no. 12,pp. 2185–2190, 2001.

[23] M. Matsushita, M. Shichiri, N. Fukai et al., “Urotensin II isan autocrine/paracrine growth factor for the porcine renalepithelial cell line, LLCPK1,” Endocrinology, vol. 144, no. 5, pp.1825–1831, 2003.

[24] A. Balat, I. H. Pakir, F. Gok, R. Anarat, and S. Sahi-noz, “Urotensin-II levels in children with minimal changenephrotic syndrome,” Pediatric Nephrology, vol. 20, no. 1, pp.42–45, 2005.

[25] A. Mosenkis, T. M. Danoff, N. Aiyar, J. Bazeley, and R. R.Townsend, “Human Urotensin II in the plasma of anephricsubjects,” Nephrology Dialysis Transplantation, vol. 22, no. 4,pp. 1269–1270, 2007.

[26] S. M. Gardiner, J. E. March, P. A. Kemp et al., “Regional het-erogeneity in the haemodynamic responses to urotensin IIinfusion in relation to UT receptor localisation,” BritishJournal of Pharmacology, vol. 147, no. 6, pp. 612–621, 2006.

[27] I. B. Wilkinson, J. T. Affolter, S. L. De Haas et al., “High plasmaconcentrations of human urotensin II do not alter local orsystemic hemodynamics in man,” Cardiovascular Research, vol.53, no. 2, pp. 341–347, 2002.

[28] F. Bohm and J. Pernow, “Urotensin II evokes potent vasocon-striction in humans in vivo,” British Journal of Pharmacology,vol. 135, no. 1, pp. 25–27, 2002.

[29] H. P. Nothacker, Z. Wang, A. M. McNeill et al., “Identificationof the natural ligand of an orphan G-protein-coupled receptorinvolved in the regulation of vasoconstriction,” Nature CellBiology, vol. 1, no. 6, pp. 383–385, 1999.

[30] C. J. Charles, M. T. Rademaker, A. M. Richards, and T. G.Yandle, “Urotensin II: evidence for cardiac, hepatic and renalproduction,” Peptides, vol. 26, no. 11, pp. 2211–2214, 2005.

[31] J. Disa, L. E. Floyd, R. M. Edwards, S. A. Douglas, and N.V. Aiyar, “Identification and characterization of binding sitesfor human urotensin-II in Sprague-Dawley rat renal medullausing quantitative receptor autoradiography,” Peptides, vol. 27,no. 6, pp. 1532–1537, 2006.

[32] A. Balat, “Urotensin-II in children with renal diseases: newmessages from an ancient peptide,” Salud i Ciencia, vol. 15,no. 6, pp. 961–965, 2007.

[33] C. A. Loretz and H. A. Bern, “Stimulation of sodium transportacross the teleost urinary bladder by urotensin II,” General andComparative Endocrinology, vol. 43, no. 3, pp. 325–330, 1981.

[34] E. Ovcharenko, Z. Abassi, I. Rubinstein, A. Kaballa, A.Hoffman, and J. Winaver, “Renal effects of human urotensin-IIin rats with experimental congestive heart failure,” NephrologyDialysis Transplantation, vol. 21, no. 5, pp. 1205–1211, 2006.

[35] A. Y. Zhang, Y. F. Chen, D. X. Zhang et al., “Urotensin II is anitric oxide-dependent vasodilator and natriuretic peptide inthe rat kidney,” American Journal of Physiology, vol. 285, no. 4,pp. F792–F798, 2003.

[36] A. E. S. Abdel-Razik, E. J. Forty, R. J. Balment, and N. Ashton,“Renal haemodynamic and tubular actions of urotensin II inthe rat,” Journal of Endocrinology, vol. 198, no. 3, pp. 617–624,2008.

[37] B. M. Y. Cheung, R. Leung, Y. B. Man, and L. Y. F. Wong,“Plasma concentration of urotensin II is raised in hyperten-sion,” Journal of Hypertension, vol. 22, no. 7, pp. 1341–1344,2004.

[38] F. D. Russell, D. Meyers, A. J. Galbraith et al., “Elevated plasmalevels of human urotensin-II immunoreactivity in congestiveheart failure,” American Journal of Physiology, vol. 285, no. 4,pp. H1576–H1581, 2003.

[39] O. Balat, F. Aksoy, I. Kutlar et al., “Increased plasma levelsof Urotensin-II in preeclampsia-eclampsia: a new mediator inpathogenesis?” European Journal of Obstetrics Gynecology andReproductive Biology, vol. 120, no. 1, pp. 33–38, 2005.

[40] P. Ravani, G. Tripepi, P. Pecchini, F. Mallamaci, F. Malberti,and C. Zoccali, “Urotensin II is an inverse predictor of deathand fatal cardiovascular events in chronic kidney disease,”Kidney International, vol. 73, no. 1, pp. 95–101, 2008.

[41] A. E. S. Abdel-Razik, R. J. Balment, and N. Ashton, “Enhancedrenal sensitivity of the spontaneously hypertensive rat tourotensin II,” American Journal of Physiology, vol. 295, no. 4,pp. F1239–F1247, 2008.

[42] A. A. Eddy, “Immune mechanisms of glomerular injury,” inPediatric Nephrology, T. M. Barratt, E. D. Avner, and W. E.Harmon, Eds., pp. 641–668, Lippincott Williams & Wilkins,Baltimore, Md, USA, 4th edition, 1999.

[43] T. Djordjevic, R. S. BelAiba, S. Bonello, J. Pfeilschifter, J. Hess,and A. Gorlach, “Human urotensin II is a novel activator ofNADPH oxidase in human pulmonary artery smooth musclecells,” Arteriosclerosis, Thrombosis, and Vascular Biology, vol.25, no. 3, pp. 519–525, 2005.

[44] L. Tian, C. Li, J. Qi et al., “Diabetes-induced upregulation ofurotensin II and its receptor plays an important role in TGF-β1-mediated renal fibrosis and dysfunction,” American Journalof Physiology, vol. 295, no. 5, pp. E1234–E1242, 2008.

[45] Y. X. Wang, Y. J. Ding, Y. Z. Zhu, Y. Shi, T. Yao, and Y. C. Zhu,“Role of PKC in the novel synergistic action of urotensin II andangiotensin II and in urotensin II-induced vasoconstriction,”American Journal of Physiology, vol. 292, no. 1, pp. H348–H359, 2007.

[46] N. S. Lamarre and R. J. Tallarida, “A quantitative studyto assess synergistic interactions between urotensin II and

Page 15: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 7

angiotensin II,” European Journal of Pharmacology, vol. 586,no. 1–3, pp. 350–351, 2008.

[47] S. Rosen, “Crescentic glomerulonephritis. Occurrence, mech-anisms and prognosis,” Pathology Annual, vol. 10, pp. 37–61,1975.

[48] B. R. Cole and L. Salinas-Madrigal, “Acute proliferativeglo-merulonephritis and crescentic glomerulonephritis,” in Pedi-atric Nephrology, T. M. Barratt, E. D. Avner, and W. E. Har-mon, Eds., pp. 669–689, Lippincott Williams & Wilkins, Balti-more, Md, USA, 4th edition, 1999.

[49] V. Sauzeau, E. Le Mellionnec, J. Bertoglio, E. Scalbert, P.Pacaud, and G. Loirand, “Human urotensin II-induced con-traction and arterial smooth muscle cell proliferation aremediated by RhoA and Rho-kinase,” Circulation Research, vol.88, no. 11, pp. 1102–1104, 2001.

[50] T. Watanabe, R. Pakala, T. Katagiri, and C. R. Benedict,“Synergistic effect of urotensin II with mildly oxidized LDL onDNA synthesis in vascular smooth muscle cells,” Circulation,vol. 104, no. 1, pp. 16–18, 2001.

[51] A. Tzanidis, R. D. Hannan, W. G. Thomas et al., “Directactions of Urotensin II on the heart: implications for cardiacfibrosis and hypertrophy,” Circulation Research, vol. 93, no. 3,pp. 246–253, 2003.

[52] Y. G. Zhang, J. Li, Y. G. Li, and R. H. Wei, “Urotensin IIinduces phenotypic differentiation, migration, and collagensynthesis of adventitial fibroblasts from rat aorta,” Journal ofHypertension, vol. 26, no. 6, pp. 1119–1126, 2008.

[53] K. Takahashi, K. Totsune, T. Kitamuro, M. Sone, O. Murakami,and S. Shibahara, “Three vasoactive peptides, endothelin-1,adrenomedullin and urotensin-II, in human tumour cell linesof different origin: expression and effects on proliferation,”Clinical Science, vol. 103, no. 48, supplement, pp. 35S–38S,2002.

[54] P. Niaudet, “Steroid-resistant idiopathic nephrotic syndrome,”in Pediatric Nephrology, T. M. Barratt, E. D. Avner, and W. E.Harmon, Eds., pp. 749–763, Lippincott Williams & Wilkins,Baltimore, Md, USA, 4th edition, 1999.

[55] K. Takahashi, K. Totsune, O. Murakami et al., “Expression ofurotensin II and its receptor in adrenal tumors and stimula-tion of proliferation of cultured tumor cells by urotensin II,”Peptides, vol. 24, no. 2, pp. 301–306, 2003.

[56] N. Bousette, L. Patel, S. A. Douglas, E. H. Ohlstein, and A.Giaid, “Increased expression of urotensin II and its cognatereceptor GPR14 in atherosclerotic lesions of the human aorta,”Atherosclerosis, vol. 176, no. 1, pp. 117–123, 2004.

[57] S. P. Makker, “Membranous glomerulonephropathy,” in Pedi-atric Nephrology, T. M. Barratt, E. D. Avner, and W. E.Harmon, Eds., Lippincott Williams & Wilkins, Baltimore, Md,USA, 4th edition, 1999.

[58] R. H. R. White, N. Yoshikawa, and J. Feehally J, “IgAnephropathy and Henoch-Schonlein nephritis,” in PediatricNephrology, T. M. Barratt, E. D. Avner, and W. E. Harmon,Eds., pp. 691–706, Lippincott Williams & Wilkins, Baltimore,Md, USA, 4th edition, 1999.

[59] A. Balat, S. O. Balci, and M. Buyukcelik, “Urotensin-II genepolymorphism in children with glomerular diseases: newmessages from an ancient peptide?” in Proceedings of the WorldCongress of Nephrology Meeting (ERA ’09), Milan, Italy, 2009.

[60] P. N. Sidharta, F. D. Wagner, H. Bohnemeier et al., “Pharma-codynamics and pharmacokinetics of the urotensin II receptorantagonist palosuran in macroalbuminuric, diabetic patients,”Clinical Pharmacology and Therapeutics, vol. 80, no. 3, pp.246–256, 2006.

[61] M. Clozel, P. Hess, C. Qiu, S. S. Ding, and M. Rey, “Theurotensin-II receptor antagonist palosuran improves pancre-atic and renal function in diabetic rats,” Journal of Pharmacol-ogy and Experimental Therapeutics, vol. 316, no. 3, pp. 1115–1121, 2006.

[62] P. N. Sidharta, P. L. M. Van Giersbergen, and J. Dingemanse,“Pharmacokinetics and pharmacodynamics of the urotensin-II receptor antagonist palosuran in healthy male subjects,”Journal of Clinical Pharmacology, vol. 49, no. 10, pp. 1168–1175, 2009.

[63] N. Nitescu, E. Grimberg, and G. Guron, “Urotensin-II recep-tor antagonism does not improve renal haemodynamics orfunction in rats with endotoxin-induced acute kidney injury,”Clinical and Experimental Pharmacology and Physiology, vol.37, no. 12, pp. 1170–1175, 2010.

Page 16: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Hindawi Publishing CorporationInternational Journal of NephrologyVolume 2012, Article ID 321714, 11 pagesdoi:10.1155/2012/321714

Review Article

Antioxidants in Kidney Diseases:The Impact of Bardoxolone Methyl

Jorge Rojas-Rivera, Alberto Ortiz, and Jesus Egido

Laboratory of Renal and Vascular Pathology, Division of Nephrology and Hypertension, IIS Fundacion Jimenez Dıaz,Autonoma University of Madrid, 28040 Madrid, Spain

Correspondence should be addressed to Jorge Rojas-Rivera, [email protected]

Received 15 September 2011; Revised 2 April 2012; Accepted 10 April 2012

Academic Editor: Ali Anarat

Copyright © 2012 Jorge Rojas-Rivera et al. This is an open access article distributed under the Creative Commons AttributionLicense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properlycited.

Drugs targeting the renin-angiotensin-aldosterone system (RAAS) are the mainstay of therapy to retard the progression ofproteinuric chronic kidney disease (CKD) such as diabetic nephropathy. However, diabetic nephropathy is still the first causeof end-stage renal disease. New drugs targeted to the pathogenesis and mechanisms of progression of these diseases beyond RAASinhibition are needed. There is solid experimental evidence of a key role of oxidative stress and its interrelation with inflammationon renal damage. However, randomized and well-powered trials on these agents in CKD are scarce. We now review the biologicalbases of oxidative stress and its role in kidney diseases, with focus on diabetic nephropathy, as well as the role of the Keap1-Nrf2pathway and recent clinical trials targeting this pathway with bardoxolone methyl.

1. Background

Chronic kidney disease (CKD) is a serious public healthproblem, which carries a high morbidity and mortality[1]. CKD is characterized by a progressive loss of renalfunction, chronic inflammation, oxidative stress, vascular re-modeling, and glomerular and tubulointerstitial scarring.CKD treatment still represents a clinical challenge. Diabeticnephropathy (DN) is the leading cause of CKD and end-stage renal disease (ESRD) [2]. The renin-angiotensin-aldosterone system (RAAS) is a major pathway involvedin the pathogenesis and progression of DN [3, 4], andRAAS blockade is an effective therapeutic strategy to reduceproteinuria and slow progression of diabetic and nondiabeticCKD. However targeting the system sets off compensatorymechanisms that may increase angiotensin II, aldosterone,or renin, and partial RAAS blockade does not preventprogression in all CKD patients. Angiotensin II (AT II) is thekey mediator of the RAAS [5–7]. Animal models of experi-mental diabetes, clinical trials, and metanalysis have clearlydemonstrated the effectiveness of angiotensin-converting

enzyme inhibitors (ACEIs) or angiotensin receptor blockers(ARBs) therapy to improve glomerular/tubulointerstitialdamage, reduce proteinuria, and decrease CKD progression,independently of blood pressure (BP) control [8–13]. DualRAAS blockade with ACEI plus ARB inhibits compensatoryAT II activity resulting from ACE-independent pathwaysand limits compensatory AT production induced by AT1receptor blockade. This combination reduced proteinuriaby 25–45% in DN [14–16]. Results are worse for DN withdiminished kidney function or nonproteinuric CKD withischemic renal injury, probably due to advanced structuralrenal changes [13, 17, 18] and adverse effects; such acutedeterioration of renal function or hyperkalemia is morefrequent. The aldosterone antagonists spironolactone andeplerenone reduce albuminuria by 30–60% and slow CKDprogression in experimental models [19–21] and clinicalstudies [22–25] in DN. These agents abrogated the “aldos-terone breakthrough” phenomenon and its proinflammatoryand profibrotic effects. ACEI/ARB therapy increases renin.Aliskiren, a direct renin inhibitor, was beneficial in animalmodels of diabetic/hypertensive nephropathy [26, 27] and

Page 17: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

2 International Journal of Nephrology

reduced albuminuria in clinical DN [28]. In a multicenterand double-blind, randomized clinical trial in hyperten-sive type 2 DM patients with nephropathy, aliskiren pluslosartan at maximal dose was 20% more effective thanlosartan/placebo to reduce albuminuria without adverseeffects, independent of BP control [29].

A number of other strategies have been tried. AdequateBP and glucose control are part of standard care of DNpatients. Intensive glucose control has more impact on GFRif early instituted in patients with type 1 DM but this maynot necessarily apply to patients with type 2 DM or withadvanced CKD [30]. A trial of the vitamin D activatorparicalcitol missed the primary endpoint of albuminuriareduction in DN and caused a transient decrease in eGFR[31]. The nephroprotective effect of statins on CKD foundin experimental models has not been conclusively provenin clinical studies [32]. A 1-year dose-ranging study ofpirfenidone suggested better preservation of eGFR by pir-fenidone in a small number of diabetic nephropathy patients[33]. The selective endothelin antagonist atrasentan reducedalbuminuria in a short-term (8 weeks) study in a smallnumber of diabetic patients while receiving RAS inhibitorsbut did not assess long-term renal function [34]. Heartfailure patients or with peripheral edema were excluded.

In spite of all this experimental and clinical evidence,there are 35–40% of patients with DN that progress toadvanced renal disease or ESRD. The risk of progressionto ESRD is still clinically relevant in other proteinuricnephropathies [35, 36]. Novel therapeutic targets are neededin CKD that are based on a clear understanding of thepathogenesis of CKD progression beyond the RAAS.

2. Oxidative Stress and Kidney Disease

Oxidative stress and inflammation promote kidney andvascular injury [37–42]. Several factors induce ROS in renalcells, such as inflammatory cytokines, Toll-like receptors,Angiotensin II, bradykinin, arachidonic acid, thrombin,growth factors, and mechanical pressure. NADPH oxidases,now renamed Nox enzymes, are key ROS generators inresponse to these stimuli [43, 44].

In acute kidney injury (AKI) induced by ischemic-reperfusion injury, sepsis or acute rejections ROS contributeof to endothelial and tubular injury [45, 46]. In murinemodels of AKI, bardoxolone methyl decreased functionaland structural renal injury and increased the expressionof protective genes (Nrf2, PPARγ, HO-1) on glomerularendothelium, cortical peritubular capillaries, tubules, andinterstitial leukocytes [47].

ROS contribute to hypertension-induced kidney andvascular injury [41, 43, 48, 49]. The chronic complications ofdiabetes are characterized by a defect in Nrf2 signaling andits adaptive response to oxidative stress. This is a potentialmechanism for cellular stress hypersensitivity and tissuedamage [50]. Diabetic nephropathy is characterized by initialhyperfiltration, albuminuria and subsequent loss of renalfunction, thickening of basement membranes, expansionof mesangial matrix and interstitial fibrosis, and podocytes

and renal cell damage [51]. ROS production in response tohyperglycemia, protein kinase C (PKC), advanced glycosy-lation end products (AGEs), free fatty acids, inflammatorycytokines, and TGF-beta1 contributes to these changes [43,44, 52–54]. These stimuli activate the NADPH/NADPHoxidase system in renal cells. Oxidative stress induced byhyperglycemia or glucose degradation products may causeleukocyte or renal cell apoptosis and release of extracellularmatrix [52, 55–59]. PKC activates NF-kappaB, extending theinflammatory response [60]. TGF-beta signaling is key tothe excessive matrix formation [61, 62]. The activation ofNrf2 is increased in diabetic nephropathy and can amelioratemesangial damage via partial inhibition of TGF-beta1 andreduction of extracellular matrix deposition [63]. ACEinhibitors lower TGF-beta in urine from DN patients. Inrat DN glomerular HO-1 is increased, evidencing oxidativestress [52, 64]. ROS can activate several transcription factorssuch as NF-kappaB, AP-1, Sp1, which in turn affect theexpression of mediators of inflammation, fibrosis, and celldeath [52, 65] (Figure 1(a)).

ROS also contribute to renal injury in experimentalglomerulonephritis. In experimental anti-Thy 1 glomeru-lonephritis, ROS enhance cell proliferation and matrixaccumulation and fibrosis and this is improved by theantioxidant alpha-lipoic acid [66, 67]. ROS also regu-late the immune response [68]. In nephrotoxic nephritis,neutrophils promote glomerular TNF-alpha expression viaH2O2 production [69]. TNF-alpha is a key mediator ofglomerular injury [70]. Interstitial inflammatory leukocytesin proliferative glomerulonephritis locally generate ROS andcontribute to sodium retention [71]. Angiotensin II pro-motes ROS-mediated F-actin cytoskeleton rearrangement,resulting in podocyte injury [72]. In cultured podocytesAT1R signaling activates Rac-1 and NADPH oxidase toproduce additional ROS and downregulates the antioxidantprotein peroxiredoxin (Prdx2) [73]. In experimental passiveHeymann nephritis, a model of membranous nephropa-thy, C5b-9 activation promotes ROS-mediated injury inglomerular cells [74, 75]. In this regard, evidence foroxidative stress, the glomerular neoexpression of aldosereductase (AR) and SOD2, and the appearance of anti-AR and anti-SOD2 autoantibodies was recently reported inhuman membranous nephropathy suggesting that oxidativestress may generate new autoimmune targets [76].

In lupus nephritis, multiple abnormalities in T and B cellslead to autoimmune renal inflammation and ROS produc-tion [77]. Nrf2-knockout mice showed impaired antioxidantactivity, increased oxidative stress, and a lupus-like autoim-mune nephritis with glomerular injury, impaired kidneyfunction, and a shortened lifespan. Thus, Nrf2 deficiencycould lead to systemic autoimmune inflammation withenhanced lymphoproliferation [78]. In antineutrophil cyto-plasmic antibodies (ANCAs-) associated vasculitis, ANCA-activated neutrophils and monocytes release MPO andgenerate ROS, producing endothelial and tissue damage [79,80]. ANCAs also promote ROS-dependent dysregulation ofneutrophil apoptosis [81] (Figure 1(a)).

In proteinuric nephropathies and independently of eti-ology, the presence of albumin in urine activates proximal

Page 18: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 3

AGE-RAGE

Hyperglycemia

RAAS activation

Increased vasoconstrictors

JAK/STAT

signaling

Albuminuria

Loss of kidney function

NADPH

oxidase

Tubulointerstitial

fibrosis

PKCGrowth factors

Cytokines

Inflammatory

cells

Podocyte, mesangial, endothelial,

and tubular cell damage

Chronic immunological disorder

Antibody production

(auto-anti-Abs, ANCAs)

Complement system

activation

Cytokines

Tubulointerstitial

inflammationGlomerulosclerosisTubular atrophy

Salt retention

Salt-sensitive

hypertension

Vasoconstriction

Ischemia

Progression to

ESRD

↑ ROS

↑ TGF-β1

↑ ECM/EMT

(a)

Figure 1: Continued.

Page 19: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

4 International Journal of Nephrology

Albuminuria

Activation of PKC

ROS

Tubulointerstitialinflammation and fibrosis

Release of chemokines (MCP-1)and attraction of macrophages

Loss of

kidney function

Progression to

ESRD

Oxidation of albumin

RCG

Autooxidation offatty acids bounds

to albumin

RCG

Activation of

NADPH oxidase

Activation ofNF-κB in PTC

(b)

Figure 1: Overview of interrelation of ROS with other key pathogenic factors in kidney disease. (a) Role of ROS in diabetic nephropathyand immune-mediated glomerulonephritis. ROS are induced in renal cells in response to high glucose, AGE, and cytokines. PKC, NADPHoxidase, and mitochondrial metabolism are key to ROS generation. ROS activate signal transduction cascade and transcription factors,leading to upregulation of genes and proteins involved in renal cell injury, glomerular and interstitial extracellular matrix deposition, andrecruitment of inflammatory cells, promoting albuminuria and progression of chronic kidney disease. (b) Role of albuminuria and ROS intubular damage and progression of CKD. Albuminuria injures PTC and activates them to release chemokines that attract macrophages andpromote tubulointerstitial fibrosis. Membrane NADPH oxidase is the main source of the ROS. It is possible the generation of other reactivespecies, as carbonyl groups derived from abnormal oxidation of albumin and fatty acids bound to albumin. Abs: antibodies, AGE: advancedglycation end products, ANCA: antineutrophil cytoplasmic antibodies, ECM: extracellular matrix, EMT: epithelial-mesenchymal transition,ESRD: end-stage renal disease, MCP-1: monocyte chemoattractant protein-1, NADPH: nicotinamide adenine dinucleotide phosphate, NF-kappa B: nuclear factor kappa B, PKC: protein kinase C, PTC: proximal tubular cell, RAAS: renin-angiotensin-aldosterone system, ROS:reactive oxygen species, RCG: reactive carbonyl groups, TGF-β1: transforming growth factor beta 1.

Page 20: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 5

tubules to release chemokines that promote interstitialinflammation [82]. Albumin uptake by tubular cells activatesPKC and ROS generation via NADPH oxidase which acti-vates NF-kappaB and production of inflammatory mediators[83, 84] (Figure 1(b)).

Accelerated atherogenesis is common in early and ad-vanced stages of CKD [40, 85]. Oxidative stress contributesto the Immune inflammation-Renal injury-Atherosclerosiscomplex (IRA paradigm), a condition present in AKI and inCKD, and accelerated atherogenesis [86, 87].

A simplified overview of several components of oxidativestress and its interrelations with other key elements ofpathogenesis and progression in kidney disease is shown inFigure 1(a).

3. Oxidative Stress and the Keap1-Nrf2 Pathway

Reactive oxygen species (ROS) include superoxide anion(SOA), hydrogen peroxide (H2O2), and hydroxyl radical. ROSare formed continuously as by-product of aerobic metab-olism. Sources of ROS include the mitochondrial electrontransport chain, metabolism of arachidonate by cyclooxyge-nases or lipoxygenases, cytochrome P450 enzymes, NADPHoxidases, or nitric oxide synthetases [88]. ROS contributeto killing bacteria, and genetic defects of NADPH oxidasecause chronic granulomatosis [89]. However, ROS may causechemical damage to DNA, proteins, and unsaturated lipidsand lead to cell death. ROS contribute to multiple pathologicprocesses [48, 90]. In this regard, homeostasis is maintainedthrough a complex set of antioxidants mechanisms that pre-vent oxidative stress-induced injury. The main mechanismsare enzymes that catalyze antioxidant reactions: glutathioneperoxidase, superoxide dismutase, catalase, Hem-oxygenase(HO-1), NADPH-quinone oxidoreductase and glutamate-cysteine ligase. These enzymes are encoded by stress-responsegenes or phase 2 genes that contain antioxidant responseelements (AREs) in their regulatory regions [88, 91]. Nrf2is the principal transcription factor that binds to the AREpromoting transcription. Actin-tethered Keap1 is a cytosolicrepressor that binds to and retains Nrf2 in the cytoplasm,promoting its proteasomal degradation. Inducers of phase2 genes modify specific cysteine residues of Keap1 resultingin conformational changes that render Keap1 unable torepress Nrf2. Consequently, Nrf2 activates the transcriptionof phase 2 genes. Oleanolic acid activates the ARE-Keap1-Nrf2 pathway, resulting in reduced proinflammatory activityof the IKK-beta/NF-kappaB pathway, increases productionof antioxidant/reductive molecules, and decreases oxidativestress, thereby restoring redox homeostasis in areas ofinflammation. In various cell lines, this results in inhibitionof proliferation, promotion of differentiation and apoptosisinduction [91–93]. Synthetic analogues of oleanolic acid,named triterpenoids, are potent anti-inflammatory agentsthat activate the ARE-Keap1-Nrf2 pathway [91]. Bardox-olone methyl, also known as CDDOMe, is a triterpenoidwhose nephroprotective action has been recently explored inhumans.

4. Antioxidant Agents in Kidney Disease

Epidemiological studies have demonstrated associationbetween inflammatory and oxidative stress markers withcardiovascular and renal outcomes in CKD and ESRD [40,94–96]. Experimental data in animal models of renal diseasesuggest beneficial effects of antioxidants agents, but results inhuman studies are limited and controversial.

In early experimental diabetes mellitus in hypertensiverats, the administration of tempol, an antioxidant SODmimetic, corrected the oxidative imbalance and improvedoxidative stress-induced renal injury, decreasing albuminuriaand fibrosis [97]. Similar protection was afforded by theantioxidants N-acetyl-L-cysteine (thiol) and kallistatin inDahl salt-sensitive rats [98, 99]. In spontaneously hyper-tensive rats, a lifelong antioxidant-rich diet diminished theseverity of hypertension, improved oxidative stress and ame-liorated abnormalities of antioxidant enzyme expressionsand activities in contrast to regular diet [100]. In sum-mary, in models of hypertensive rats, synthetic and naturalantioxidants induced renal and endothelial protection withreduction of oxidative stress. In a model of ischaemia reper-fusion and cyclosporin toxicity after unilateral nephrectomy,the blockage of the mitochondrial enzymes monoamineoxidases with pargyline 28 days following surgery preventedH2O2 production and improved renal function and renalinflammation (lower IL-1β and TNF-α gene expression)[101]. Pargyline administrated before ischemia reperfusionsignificantly reduced apoptosis, necrosis, and fibrosis. Thiseffect was associated to decreased expression of TGF-β1, col-lagen types I, III, and IV and to the normalization of SOD1,catalase, and inflammatory gene expression. In models ofrenal chronic failure (5/6 nephrectomy rats) [102], AST-120,an oral carbonic adsorbent, improved the oxidative stress inendothelial cells, measured as oxidized/unoxidized albuminratio. This effect was reached reducing the blood levels ofindoxyl sulfate, an uremic toxin that induces ROS. In anothermodel of remnant kidney, the administration of omega-3fatty acids, an effective compound in mitigating atheroscle-rosis, significantly lowered several components of oxidativestress and markers of inflammatory and fibrotic response.Furthermore, it attenuated tubulointerstitial fibrosis andinflammation in the remnant kidney [103]. In anti-Thy1glomerulonephritis, the treatment with parthenolide, ananti-inflammatory agent related to the triterpenoid family,diminished renal inflammation via NF-kappaB inhibition,decreased MCP-1 and iNOS, and improved proteinuria,tubular, and glomerular damage [104]. The beneficial effectof exogenous antioxidants shown in animal models withhypertension or chronic renal failure has not been demon-strated in people with clinical hypertension [96, 105] orCKD.

5. Nephroprotection by Bardoxolone Methyl

Bardoxolone was initially described as an agent that pro-tected cells from radiation-induced damage (radiation mit-igator) through Nrf2-dependent and -independent pathways[106]. In humans, its potential antineoplasic activity was

Page 21: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

6 International Journal of Nephrology

RAASoveractivation

BARD

ACEI/ARB or

MRA/DRI therapy

CKDprogression

Increased antioxidant response

Decreased ROS generation

Decreased inflammation

Nrf2 activation

Other mechanisms

Aldosterone leakage phenomenon

Prorenin effects

Transcription of genes with

ARE sequences in promoters

Keap1

+

↑ Cell injury

↑ TGF-β1

↑ NFκB

↑ ROS

Figure 2: Overview of RAAS blockers and bardoxolone in pathogenic pathways in kidney diseases. RAAS blockers target several pathogenicpathways in kidney injury, including those generating ROS. However, there are several escape mechanisms (aldosterone breakthrough,increased prorenin effects) as well as less sensitive lesions (significant loss of kidney function, ischemic disease, persistent immune activity).BARD promotes activation of the Nrf2 transcription factor, that is released of the inhibitory Keap1 protein and migrates to the nucleuswhere it regulates transcription of genes containing ARE sequences in their promoters. These phase 2 response genes are collectivelyinvolved in the reduction of ROS and inhibition of NF-kappaB. Thus, BARD could promote renal protection through antioxidants and anti-inflammatory effects be promoting the activity of the Nrf2 transcription factor and inhibiting the activity of the NF-kappaB transcriptionfactor. ACE/ACEIs: angiotensin converting enzyme/angiotensin converting enzyme inhibitors, ARBs: angiotensin receptor blockers, AREs:antioxidant response elements, BARD: bardoxolone methyl, CKD: chronic kidney disease, DRI: direct renin inhibitor, mineralocorticoidreceptor antagonists, Keap1: Kelch-like ECH-associated protein 1, MRA: mineralocorticoid receptor antagonists, NF-kappaB: nuclear factorkappa B, Nrf2: nuclear factor (erythroid derive 2)-like 2, RAAS: renin-angiotensin-aldosterone system, ROS: reactive oxygen species, TGFβ-1:transforming growth factor beta 1.

evaluated. In phase 1 trials in oncologic patients, bar-doxolone unexpectedly improved kidney function, assessedas serum creatinine and creatinine clearance, especially inpatients with previous CKD. These findings lead to evaluatepotential nephroprotective actions in patients with CKD and

type 2 DM, first in an exploratory phase II open-label trialand then in a larger randomized clinical trial. In the first trial[107], 20 patients older than 18 years, with moderate-severediabetic CKD, were evaluated after 8 weeks of bardoxoloneat increasing oral doses of 25 to 75 mg/day. Notably, there

Page 22: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 7

was a significant increase in estimated GFR at 4 weeks(+2.8 mL/min/1.73 m2) with 25 mg/day and at 8 weeks(+7.2 mL/min/1.73 m2) with 75 mg/day. Serum creatinineand BUN decreased and creatinine clearance increased,without changes in total excretion or tubular secretion ofcreatinine. Unfortunately, GFR was not measured. Bloodpressure did not change, and albuminuria had a small albeitnot statistically significant increase. Markers of vascularinjury and inflammation were improved by treatment withbardoxolone, suggesting a potential beneficial effect onendothelial injury. There were not changes in urine NGALor NAG adjusted for creatinine concentration, suggestinglack of significant renal toxicity associated to bardoxolone.There were few adverse effects, mainly muscle spasms anda self-limited increase of hepatic enzymes without a truehepatic toxicity. A short followup and an open-label designwithout control group are major limitations of this study,and they do not allow drawing solid conclusions about theefficacy and long-term safety of this drug on relevant renaloutcomes.

The beneficial effect on eGFR was confirmed in a larger,multicenter, double-blind, randomized trial [108]. This trialrandomized 227 patients with moderate-severe CKD andtype 2 DM, with stable treatment with ACEI/ARB, to bardox-olone 25, 75 or 150 mg/day or placebo for 52 weeks. Patientswere categorized by GFR, urinary albumin-creatinine ratio(UACR), and HbA1c. Patients with hepatic dysfunction orrecent cardiovascular events were excluded. A significantimprovement in the primary endpoint (change of GFR at 24weeks) was observed in all bardoxolone groups (+8.2, +11.4and +10.4 mL/min/1.73 m2 resp.) versus 0 mL/min/1.73m2

in the placebo group. The secondary endpoint (changeof GFR at 52 weeks) also was significantly improved inbardoxolone groups (+5.8, +10.5 and +9.3 mL/min/1.73 m2

resp. versus 0). More patients in the placebo group had a GFRdecrease ≥25% with respect to baseline value at 24 and 52weeks. Additionally, serum BUN, phosphorus, and uric acidwere significantly lower at 24 and 52 weeks in all bardoxolonegroups when compared to placebo.

Potential unwanted effects included a mild but significantincrease of UACR and decreased serum magnesium. UACRincreased in patients receiving 75 or 150 mg/day bardoxoloneversus placebo. This was observed at 24 and 52 weeks oftreatment, but UACR decreased when patients stopped thetherapy, suggesting that this effect is reversible. There wasan inverse correlation between changes in serum BUN,phosphorus, uric acid, magnesium, and changes in eGFR,as well as a direct correlation between changes in eGFR andchanges in UACR, suggesting that changes in eGFR may bethe basis for the other observed changes. Interestingly, therewas a trend toward higher systolic BP values in the 75 mgbardoxolone group, which was observed despite weight lossand that will merit close attention in further trials. The mainadverse effects were muscle spasms (63% of patients in the75 mg group) and nausea (25%).

Another significant effect was loss of body weight. Thisappears to be related to decreased appetite and/or nauseaand may be a welcome addition to the therapeutic arma-mentarium for patients with increased body mass index

(BMI). Indeed, weight loss was more evident in patients withhigher (>35 kg/m2) BMI (mean change −10 kg). However,and perhaps worryingly, it was also observed in patients withnormal BMI (−3 kg over 52 weeks).

The increased eGFR and effects on systolic BP and albu-minuria are interesting results on surrogates renal variableswhich requires more long-term studies. These parametersand, more importantly, cardiac and renal hard end-points(cardiovascular death and progression to ESRD) will bestudied at 2 years of followup in an ongoing randomizedclinical trial in 1600 patients older than 18 years withadvanced CKD (stage 4) and type 2 DM [109]. This study willcompare bardoxolone versus placebo in patients receivingstandard of care.

6. Conclusions and Recommendations

RAAS blockade is the mainstay of current therapy to slowprogression of diabetic and nondiabetic CKD, but thisstrategy is frequently not enough. Consequently, an impor-tant number of patients progress to ESRD. There is solidexperimental evidence for a key role of ROS and oxidativestress and their interplay with RAAS and inflammation, inthe pathogenesis of CKD. Bardoxolone methyl, a novel syn-thetic triterpenoid with antioxidant and anti-inflammatoryproperties, has shown to improve kidney function in patientswith advanced DN already receiving RAAS blockers, withfew adverse events. This may be a welcomed addition to thetherapeutic armamentarium if data are confirmed in larger,longer trials (Figure 2). However, the relative importanceand eventual management of the observed influence ofbardoxolone on UACR, magnesium, and body weight mustbe further studied.

Acknowledgments

This paper is supported by the following Grants: ISCIIIand FEDER funds CP04/00060, FIS PS09/00447, SociedadEspanola de Nefrologia, ISCIII-RETIC REDinREN/RD06/0016, Comunidad de Madrid/FRACM/S-BIO0283/2006,S2010/BMD-2378, Programa Intensificacion Actividad In-vestigadora (ISCIII/Agencia Laın-Entralgo/CM) to AO, andISCIII-Redes RECAVA (RD06/0014/0035), ISCIII fundsPI10/00072, EUS2008/03565 to JE and Fundacion Lilly,cvREMOD.

References

[1] Center for Disease Control and Prevention, “Prevalence ofchronic kidney disease and associated risk factors: UnitedStates, 1999–2004,” Morbidity and Mortality Weekly Report,vol. 56, pp. 161–165, 2007.

[2] S. Wild, G. Roglic, A. Green, R. Sicree, and H. King, “Globalprevalence of diabetes: estimates for the year 2000 andprojections for 2030,” Diabetes Care, vol. 27, no. 5, pp. 1047–1053, 2004.

[3] P. Ruggenenti, P. Cravedi, and G. Remuzzi, “The RAAS in thepathogenesis and treatment of diabetic nephropathy,” NatureReviews Nephrology, vol. 6, no. 6, pp. 319–330, 2010.

Page 23: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

8 International Journal of Nephrology

[4] T. Morgan, “Renin, angiotensin, sodium and organ damage,”Hypertension Research, vol. 26, no. 5, pp. 349–354, 2003.

[5] C. M. Ferrario, “Role of angiotensin II in cardiovasculardisease—therapeutic implications of more than a centuryof research,” Journal of the Renin-Angiotensin-AldosteroneSystem, vol. 7, no. 1, pp. 3–14, 2006.

[6] M. E. Cooper, “The role of the renin-angiotensin-aldosteronesystem in diabetes and its vascular complications,” AmericanJournal of Hypertension, vol. 17, no. 11, pp. 16S–20S, 2004.

[7] U. C. Brewster and M. A. Perazella, “The renin-angiotensin-aldosterone system and the kidney: effects on kidney disease,”American Journal of Medicine, vol. 116, no. 4, pp. 263–272,2004.

[8] H. M. Siragy and R. M. Carey, “Protective role of the angi-otensin AT2 receptor in a renal wrap hypertension model,”Hypertension, vol. 33, no. 5, pp. 1237–1242, 1999.

[9] M. A. Ondetti, B. Rubin, and D. W. Cushman, “Design of spe-cific inhibitors of angiotensin converting enzyme: new classof orally active antihypertensive agents,” Science, vol. 196, no.4288, pp. 441–444, 1977.

[10] G. Remuzzi, “Randomised placebo-controlled trial of effectof ramipril on decline in glomerular filtration rate andrisk of terminal renal failure in proteinuric, non-diabeticnephropathy,” The Lancet, vol. 349, no. 9069, pp. 1857–1863,1996.

[11] Y. Taguma, Y. Kitamoto, and G. Futaki, “Effect of captopril onheavy proteinuria in azotemic diabetics,” The New EnglandJournal of Medicine, vol. 313, no. 26, pp. 1617–1620, 1985.

[12] B. M. Brenner, M. E. Cooper, D. de Zeeuw et al., “Effects oflosartan on renal and cardiovascular outcomes in patientswith type 2 diabetes and nephropathy,” The New EnglandJournal of Medicine, vol. 345, no. 12, pp. 861–869, 2001.

[13] E. J. Lewis, L. G. Hunsicker, W. R. Clarke et al., “Renoprotec-tive effect of the angiotensin-receptor antagonist irbesartanin patients with nephropathy due to type 2 diabetes,” TheNew England Journal of Medicine, vol. 345, no. 12, pp. 851–860, 2001.

[14] P. Jacobsen, S. Andersen, K. Rossing, B. V. Hansen, and H.H. Parving, “Dual blockade of the renin-angiotensin systemin type 1 patients with diabetic nephropathy,” NephrologyDialysis Transplantation, vol. 17, no. 6, pp. 1019–1024, 2002.

[15] C. E. Mogensen, S. Neldam, I. Tikkanen et al., “Randomisedcontrolled trial of dual blockade of renin-angiotensin systemin patients with hypertension, microalbuminuria, and non-insulin dependent diabetes: the candesartan and lisinoprilmicroalbuminuria (CALM) study,” British Medical Journal,vol. 321, no. 7274, pp. 1440–1444, 2000.

[16] P. Jacobsen, S. Andersen, B. R. Jensen, and H. H. Parving,“Additive effect of ACE inhibition and angiotensin II receptorblockade in type I diabetic patients with diabetic nephropa-thy,” Journal of the American Society of Nephrology, vol. 14,no. 4, pp. 992–999, 2003.

[17] P. Ruggenenti, L. Mosconi, F. Sangalli et al., “Glomerular size-selective dysfunction in NIDDM is not ameliorated by ACEinhibition or by calcium channel blockade,” Kidney Interna-tional, vol. 55, no. 3, pp. 984–994, 1999.

[18] Z. Zhang, S. Shahinfar, W. F. Keane et al., “Importance ofbaseline distribution of proteinuria in renal outcomes trials:lessons from the reduction of endpoints in NIDDM with theangiotensin II antagonist losartan (RENAAL) study,” Journalof the American Society of Nephrology, vol. 16, no. 6, pp. 1775–1780, 2005.

[19] W. Huang, C. Xu, K. W. Kahng, N. A. Noble, W. A. Border,and Y. Huang, “Aldosterone and TGF-β1 synergistically

increase PAI-1 and decrease matrix degradation in rat renalmesangial and fibroblast cells,” American Journal of Physiolo-gy, vol. 294, no. 6, pp. F1287–F1295, 2008.

[20] M. Nagase, S. Shibata, S. Yoshida, T. Nagase, T. Gotoda, andT. Fujita, “Podocyte injury underlies the glomerulopathy ofDahl salt-hypertensive rats and is reversed by aldosteroneblocker,” Hypertension, vol. 47, no. 6, pp. 1084–1093, 2006.

[21] M. Briet and E. L. Schiffrin, “Aldosterone: effects on the kid-ney and cardiovascular system,” Nature Reviews Nephrology,vol. 6, no. 5, pp. 261–273, 2010.

[22] A. Struthers, H. Krum, and G. H. Williams, “A comparison ofthe aldosterone-blocking agents eplerenone and spironolac-tone,” Clinical Cardiology, vol. 31, no. 4, pp. 153–158, 2008.

[23] S. Bianchi, R. Bigazzi, and V. M. Campese, “Long-term effectsof spironolactone on proteinuria and kidney function inpatients with chronic kidney disease,” Kidney International,vol. 70, no. 12, pp. 2116–2123, 2006.

[24] U. F. Mehdi, B. Adams-Huet, P. Raskin, G. L. Vega, and R. D.Toto, “Addition of angiotensin receptor blockade or miner-alocorticoid antagonism to maximal angiotensin-convertingenzyme inhibition in diabetic nephropathy,” Journal of theAmerican Society of Nephrology, vol. 20, no. 12, pp. 2641–2650, 2009.

[25] S. D. Navaneethan, S. U. Nigwekar, A. R. Sehgal, and G.F. M. Strippoli, “Aldosterone antagonists for preventing theprogression of chronic kidney disease: a systematic reviewand meta-analysis,” Clinical Journal of the American Societyof Nephrology, vol. 4, no. 3, pp. 542–551, 2009.

[26] Y. Huang, S. Wongamorntham, J. Kasting et al., “Renin in-creases mesangial cell transforming growth factor-β1 andmatrix proteins through receptor-mediated, angiotensin II-independent mechanisms,” Kidney International, vol. 69, no.1, pp. 105–113, 2006.

[27] A. Ichihara, F. Suzuki, T. Nakagawa et al., “Prorenin receptorblockade inhibits development of glomerulosclerosis in dia-betic angiotensin II type 1a receptor-deficient mice,” Journalof the American Society of Nephrology, vol. 17, no. 7, pp. 1950–1961, 2006.

[28] F. Persson, P. Rossing, H. Reinhard et al., “Renal effects ofaliskiren compared with and in combination with irbesartanin patients with type 2 diabetes, hypertension, and albumin-uria,” Diabetes Care, vol. 32, no. 10, pp. 1873–1879, 2009.

[29] H. H. Parving, F. Persson, J. B. Lewis, E. J. Lewis, and N.K. Hollenberg, “Aliskiren combined with losartan in type2 diabetes and nephropathy,” The New England Journal ofMedicine, vol. 358, no. 23, pp. 2433–2446, 2008.

[30] I. H. De Boer, W. Sun, P. A. Cleary et al., “Intensive diabetestherapy and glomerular filtration rate in type 1 diabetes,” TheNew England Journal of Medicine, vol. 265, pp. 2366–2376,2011.

[31] D. de Zeeuw, R. Agarwal, M. Amdahl et al., “Selectivevitamin D receptor activation with paricalcitol for reductionof albuminuria in patients with type 2 diabetes (VITALstudy): a randomised controlled trial,” The Lancet, vol. 376,no. 9752, pp. 1543–1551, 2010.

[32] M. K. Rutter, H. R. Prais, V. Charlton-Menys et al., “Pro-tection against nephropathy in diabetes with atorvastatin(PANDA): a randomized double-blind placebo-controlledtrial of high- vs. low-dose atorvastatin (1),” DiabeticMedicine, vol. 28, no. 1, pp. 100–108, 2011.

[33] K. Sharma, J. H. Ix, A. V. Mathew et al., “Pirfenidone fordiabetic nephropathy,” Journal of the American Society ofNephrology, vol. 22, no. 6, pp. 1144–1151, 2011.

Page 24: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 9

[34] D. E. Kohan, Y. Pritchett, M. Molitch et al., “Addition ofatrasentan to renin-angiotensin system blockade reducesalbuminuria in diabetic nephropathy,” Journal of the Amer-ican Society of Nephrology, vol. 22, no. 4, pp. 763–772, 2011.

[35] H. Trachtman, F. C. Fervenza, D. S. Gipson et al., “Aphase 1, single-dose study of fresolimumab, an anti-TGF-β antibody, in treatment-resistant primary focal segmentalglomerulosclerosis,” Kidney International, vol. 79, no. 11, pp.1236–1243, 2011.

[36] M. Waldman and H. A. Austin III, “Controversies in thetreatment of idiopathic membranous nephropathy,” NatureReviews Nephrology, vol. 5, no. 8, pp. 469–479, 2009.

[37] M. Ruiz-Ortega and A. Ortiz, “Angiotensin II and reactiveoxygen species,” Antioxidants and Redox Signaling, vol. 7, no.9-10, pp. 1258–1260, 2005.

[38] A. Lazaro, J. Gallego-Delgado, P. Justo et al., “Long-termblood pressure control prevents oxidative renal injury,”Antioxidants and Redox Signaling, vol. 7, no. 9-10, pp. 1285–1293, 2005.

[39] F. Neria, M. A. Castilla, R. F. Sanchez et al., “Inhibitionof JAK2 protects renal endothelial and epithelial cells fromoxidative stress and cyclosporin A toxicity,” Kidney Interna-tional, vol. 75, no. 2, pp. 227–234, 2009.

[40] V. Cachofeiro, M. Goicochea, S. G. de Vinuesa, P. Oubıa, V.Lahera, and J. Luo, “Oxidative stress and inflammation, a linkbetween chronic kidney disease and cardiovascular disease,”Kidney International, vol. 74, supplement 111, pp. S4–S9,2008.

[41] N. D. Vaziri and B. Rodriguez-Iturbe, “Mechanisms of dis-ease: oxidative stress and inflammation in the pathogenesisof hypertension,” Nature Clinical Practice Nephrology, vol. 2,no. 10, pp. 582–593, 2006.

[42] H. J. Kim, T. Sato, N. D. Vaziri, and B. Rodriguez-Iturbe,“Role of intrarenal angiotensin system activation, oxidativestress, inflammation, and impaired nuclear factor-erythroid-2-related factor 2 activity in the progression of focal glomeru-losclerosis,” Journal of Pharmacology and Experimental Ther-apeutics, vol. 337, no. 3, pp. 583–590, 2011.

[43] E. N. Wardle, “Cellular oxidative processes in relation to renaldisease,” American Journal of Nephrology, vol. 25, no. 1, pp.13–22, 2005.

[44] M. Pleskova, K. F. Beck, M. H. Behrens et al., “Nitricoxide down-regulates the expression of the catalytic NADPHoxidase subunit Nox1 in rat renal mesangial cells,” The FASEBJournal, vol. 20, no. 1, pp. 139–141, 2006.

[45] B. A. Molitoris and J. Marrs, “The role of cell adhesion mol-ecules in ischemic acute renal failure,” American Journal ofMedicine, vol. 106, no. 5, pp. 583–592, 1999.

[46] M. A. R. C. Daemen, C. Van’t Veer, G. Denecker et al., “Inhi-bition of apoptosis induced by ischemia-reperfusion preventsinflammation,” Journal of Clinical Investigation, vol. 104, no.5, pp. 541–549, 1999.

[47] J. M. Alonso de Vega, J. Diaz, E. Serrano, and L. F. Carbonell,“Oxidative stress in critically ill patients with systemicinflammatory response syndrome,” Critical Care Medicine,vol. 30, no. 8, pp. 1782–1786, 2002.

[48] Y. Taniyama and K. K. Griendling, “Reactive oxygen speciesin the vasculature: molecular and cellular mechanisms,” Hy-pertension, vol. 42, no. 6, pp. 1075–1081, 2003.

[49] M. Ruiz-Ortega, J. Egido, O. Lorenzo, M. Ruperez, S. Konig,and B. Wittig, “Angiotensin II activates nuclear transcriptionfactor κB through AT1 and AT2 in vascular smooth musclecells molecular mechanisms,” Circulation Research, vol. 86,no. 12, pp. 1266–1272, 2000.

[50] M. S. Bittar and F. Al-Mulla, “A defect in Nrf2 signaling con-stitutes a mechanism for cellular stress hypersensitivity intype 2 diabetes,” American Journal of Physiology, vol. 301, no.6, pp. E1119–E1129, 2011.

[51] R. M. Mason and N. A. Wahab, “Extracellular matrix me-tabolism in diabetic nephropathy,” Journal of the AmericanSociety of Nephrology, vol. 14, no. 5, pp. 1358–1373, 2003.

[52] H. B. Lee, M. R. Yu, Y. Yang, Z. Jiang, and H. Ha, “React-ive oxygen species-regulated signaling pathways in diabeticnephropathy,” Journal of the American Society of Nephrology,vol. 14, no. 8, supplement 3, pp. S241–S245, 2003.

[53] M. Brownlee, “Biochemistry and molecular cell biology ofdiabetic complications,” Nature, vol. 414, no. 6865, pp. 813–820, 2001.

[54] M. A. Lal, H. Brismar, A. C. Eklof, and A. Aperia, “Role ofoxidative stress in advanced glycation end product-inducedmesangial cell activation,” Kidney International, vol. 61, no.6, pp. 2006–2014, 2002.

[55] P. Justo, A. B. Sanz, J. Egido, and A. Ortiz, “3,4-Dideoxy-glucosone-3-ene induces apoptosis in renal tubular epithelialcells,” Diabetes, vol. 54, no. 8, pp. 2424–2429, 2005.

[56] B. P. S. Kang, S. Frencher, V. Reddy, A. Kessler, A. Mal-horta, and L. G. Meggs, “High glucose promotes mesangialcell apoptosis by oxidant-dependent mechanism,” AmericanJournal of Physiology, vol. 284, no. 3, pp. F455–F466, 2003.

[57] M. C. Iglesias-de La Cruz, P. Ruiz-Torres, J. AlcamI et al.,“Hydrogen peroxide increases extracellular matrix mRNAthrough TGF-β in human mesangial cells,” Kidney Interna-tional, vol. 59, no. 1, pp. 87–95, 2001.

[58] M. P. Catalan, B. Santamarıa, A. Reyero, A. Ortiz, J.Egido, and A. Ortiz, “3,4-Di-deoxyglucosone-3-ene pro-motes leukocyte apoptosis,” Kidney International, vol. 68, no.3, pp. 1303–1311, 2005.

[59] A. Ortiz, C. Lorz, P. Justo, M. P. Catalan, and J. Egido, “Con-tribution of apoptotic cell death to renal injury,” Journal ofCellular and Molecular Medicine, vol. 5, no. 1, pp. 18–32,2001.

[60] T. Inoguchi, T. Sonta, H. Tsubouchi et al., “Protein kinase C-dependent increase in reactive oxygen species (ROS) produc-tion in vascular tissues of diabetes: role of vascular NAD(P)Hoxidase,” Journal of the American Society of Nephrology, vol.14, no. 8, supplement 3, pp. S227–S232, 2003.

[61] F. N. Zyyadeh, “Mediators of diabetic renal disease: the casefor TGF-β as a major mediator,” Journal of the AmericanSociety of Nephrology, vol. 15, supplement 1, pp. S55–S57,2004.

[62] C. Weigert, U. Sauer, K. Brodbeck, A. Pfeiffer, H. U. Haring,and E. D. Schleicher, “AP-1 proteins mediate hyperglycemia-induced activation of the human TGF-β1 promoter inmesangial cells,” Journal of the American Society of Nephrol-ogy, vol. 11, no. 11, pp. 2007–2016, 2000.

[63] T. Jiang, Z. Huang, Y. Lin, Z. Zhang, D. Fang, and D. D.Zhang, “The protective role of Nrf2 in streptozotocin-induced diabetic nephropathy,” Diabetes, vol. 59, no. 4, pp.850–860, 2010.

[64] J. M. Li and A. M. Shah, “ROS generation by nonphagocyticNADPH oxidase: potential relevance in diabetic nephropa-thy,” Journal of the American Society of Nephrology, vol. 14,no. 3, pp. S221–S226, 2003.

[65] A. B. Sanz, M. D. Sanchez-Nino, A. M. Ramos et al., “NF-κB in renal inflammation,” Journal of the American Society ofNephrology, vol. 21, no. 8, pp. 1254–1262, 2010.

Page 25: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

10 International Journal of Nephrology

[66] S. V. Shah, “Oxidants and iron in chronic kidney disease,”Kidney International, Supplement, vol. 66, no. 91, pp. S50–S55, 2004.

[67] M. N. Budisavljevic, L. Hodge, K. Barber et al., “Oxidativestress in the pathogenesis of experimental mesangial prolif-erative glomerulonephritis,” American Journal of Physiology,vol. 285, no. 6, pp. F1138–F1148, 2003.

[68] S. Kantengwa, L. Jornot, C. Devenoges, and L. P. Nicod, “Su-peroxide anions induce the maturation of human dendriticcells,” American Journal of Respiratory and Critical CareMedicine, vol. 167, no. 3, pp. 431–437, 2003.

[69] Y. Suzuki, C. Gomez-Guerrero, I. Shirato et al., “Pre-existingglomerular immune complexes induce polymorphonuclearcell recruitment through an Fc receptor-dependent respira-tory burst: potential role in the perpetuation of immunenephritis,” Journal of Immunology, vol. 170, no. 6, pp. 3243–3253, 2003.

[70] A. Ortiz, S. Gonzalez-Cuadrado, C. Bustos et al., “Tumornecrosis factor as a mediator of glomerular damage,” Journalof Nephrology, vol. 8, no. 1, pp. 27–34, 1995.

[71] B. Rodriguez-Iturbe, J. Herrera-Acosta, and R. J. Johnson,“Interstitial inflammation, sodium retention, and the patho-genesis of nephrotic edema: a unifying hypothesis,” KidneyInternational, vol. 62, no. 4, pp. 1379–1384, 2002.

[72] B. Rodriguez-Iturbe, N. D. Vaziri, J. Herrera-Acosta, and R.J. Johnson, “Oxidative stress, renal infiltration of immunecells, and salt-sensitive hypertension: all for one and one forall,” American Journal of Physiology, vol. 286, no. 4, pp. F606–F616, 2004.

[73] H. H. Hsu, S. Hoffmann, G. S. Di Marco et al., “Downregula-tion of the antioxidant protein peroxiredoxin 2 contributesto angiotensin II-mediated podocyte apoptosis,” KidneyInternational, vol. 80, pp. 959–969, 2011.

[74] T. J. Neale, P. P. Ojha, M. Exner et al., “Proteinuria in passiveHeymann nephritis is associated with lipid peroxidation andformation of adducts on type IV collagen,” Journal of ClinicalInvestigation, vol. 94, no. 4, pp. 1577–1584, 1994.

[75] W. G. Couser and M. Nangaku, “Cellular and molecularbiology of membranous nephropathy,” Journal of Nephrology,vol. 19, no. 6, pp. 699–705, 2006.

[76] M. Prunotto, M. L. Carnevali, G. Candiano et al., “Autoim-munity in membranous nephropathy targets aldose reduc-tase and SOD2,” Journal of the American Society of Nephrol-ogy, vol. 21, no. 3, pp. 507–519, 2010.

[77] M. H. Foster, “T cells and B cells in lupus nephritis,” Seminarsin Nephrology, vol. 27, no. 1, pp. 47–58, 2007.

[78] K. Yoh, K. Itoh, A. Enomoto et al., “Nrf2-deficient femalemice develop lupus-like autoimmune nephritis,” KidneyInternational, vol. 60, no. 4, pp. 1343–1353, 2001.

[79] R. J. Falk and J. C. Jennette, “ANCA are pathogenicoh—ohyes they are!,” Journal of the American Society of Nephrology,vol. 13, no. 7, pp. 1977–1979, 2002.

[80] J. P. Gaut, J. Byun, H. D. Tran et al., “Myeloperoxidaseproduces nitrating oxidants in vivo,” Journal of Clinical In-vestigation, vol. 109, no. 10, pp. 1311–1319, 2002.

[81] L. Harper, Y. Ren, J. Savill, D. Adu, and C. O. S. Sav-age, “Antineutrophil cytoplasmic antibodies induce reactiveoxygen-dependent dysregulation of primed neutrophil apop-tosis and clearance by macrophages,” American Journal ofPathology, vol. 157, no. 1, pp. 211–220, 2000.

[82] M. Gomez-Chiarri, A. Ortiz, S. Gonzalez-Cuadrado et al.,“Interferon-inducible protein-10 is highly expressed in ratswith experimental nephrosis,” American Journal of Pathology,vol. 148, no. 1, pp. 301–311, 1996.

[83] S. Tang, J. C. K. Leung, K. Abe et al., “Albumin stimulatesinterleukin-8 expression in proximal tubular epithelial cellsin vitro and in vivo,” Journal of Clinical Investigation, vol. 111,no. 4, pp. 515–527, 2003.

[84] M. Morigi, D. Macconi, C. Zoja et al., “Protein overload-induced NF-κB activation in proximal tubular cells requiresH2O2 through a PKC-dependent pathway,” Journal of theAmerican Society of Nephrology, vol. 13, no. 5, pp. 1179–1189,2002.

[85] R. Stocker and J. F. Keaney, “Role of oxidative modificationsin atherosclerosis,” Physiological Reviews, vol. 84, no. 4, pp.1381–1478, 2004.

[86] J. Himmelfarb, P. Stenvinkel, T. A. Ikizler, and R. M. Hakim,“Perspectives in renal medicine: the elephant in uremia:oxidant stress as a unifying concept of cardiovascular diseasein uremia,” Kidney International, vol. 62, no. 5, pp. 1524–1538, 2002.

[87] S. Swaminathan and S. V. Shah, “Novel inflammatory mech-anisms of accelerated atherosclerosis in kidney disease,”Kidney International, vol. 80, pp. 453–463, 2011.

[88] P. Champe and R. A. Harvey, Pentose Phosphate Pathwayand NADPH in Biochemistry, Lippincott Williams & Wilkins,Philadelphia, Pa, USA, 4th edition, 2008.

[89] C. K. Sen, “Antioxidant and redox regulation of cellular sig-naling: Introduction,” Medicine and Science in Sports andExercise, vol. 33, no. 3, pp. 368–370, 2001.

[90] K. K. Griendling, D. Sorescu, and M. Ushio-Fukai,“NAD(P)H oxidase: role in cardiovascular biology and dis-ease,” Circulation Research, vol. 86, no. 5, pp. 494–501, 2000.

[91] A. T. Dinkova-Kostova, K. T. Liby, K. K. Stephenson etal., “Extremely potent triterpenoid inducers of the phase2 response: correlations of protection against oxidant andinflammatory stress,” Proceedings of the National Academy ofSciences of the United States of America, vol. 102, no. 12, pp.4584–4589, 2005.

[92] W. Li, T. O. Khor, C. Xu et al., “Activation of Nrf2-antioxidantsignaling attenuates NFκB-inflammatory response and elicitsapoptosis,” Biochemical Pharmacology, vol. 76, no. 11, pp.1485–1489, 2008.

[93] R. K. Thimmulappa, H. Lee, T. Rangasamy et al., “Nrf2 is acritical regulator of the innate immune response and survivalduring experimental sepsis,” Journal of Clinical Investigation,vol. 116, no. 4, pp. 984–995, 2006.

[94] F. Lacy, M. T. Kailasam, D. T. O’Connor, G. W. Schmid-Schonbein, and R. J. Parmer, “Plasma hydrogen peroxideproduction in human essential hypertension: role of heredity,gender, and ethnicity,” Hypertension, vol. 36, no. 5, pp. 878–884, 2000.

[95] L. J. Dixon, S. M. Hughes, K. Rooney et al., “Increased su-peroxide production in hypertensive patients with diabetesmellitus: role of nitric oxide synthase,” American Journal ofHypertension, vol. 18, no. 6, pp. 839–843, 2005.

[96] A. M. Roggensack, Y. Zhang, and S. T. Davidge, “Evidencefor peroxynitrite formation in the vasculature of women withpreeclampsia,” Hypertension, vol. 33, no. 1 I, pp. 83–89, 1999.

[97] Y. Quiroz, A. Ferrebuz, N. D. Vaziri, and B. Rodriguez-Iturbe,“Effect of chronic antioxidant therapy with superoxidedismutase-mimetic drug, tempol, on progression of renaldisease in rats with renal mass reduction,” Nephron, vol. 112,no. 1, pp. e31–e42, 2009.

[98] L. Zhang, S. Fujii, J. Igarashi, and H. Kosaka, “Effects of thiolantioxidant on reduced nicotinamide adenine dinucleotidephosphate oxidase in hypertensive Dahl salt-sensitive rats,”

Page 26: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 11

Free Radical Biology and Medicine, vol. 37, no. 11, pp. 1813–1820, 2004.

[99] B. Shen, M. Hagiwara, Y. Y. Yao, L. Chao, and J. Chao, “Salu-tary effect of kallistatin in salt-induced renal injury, inflam-mation, and fibrosis via antioxidative stress,” Hypertension,vol. 51, no. 5, pp. 1358–1365, 2008.

[100] C. D. Zhan, R. K. Sindhu, J. Pang, A. Ehdaie, and N. D. Vaziri,“Superoxide dismutase, catalase and glutathione peroxidasein the spontaneously hypertensive rat kidney: effect ofantioxidant-rich diet,” Journal of Hypertension, vol. 22, no.10, pp. 2025–2033, 2004.

[101] R. Chaaya, C. Alfarano, C. Guilbeau-Frugier et al., “Pargylinereduces renal damage associated with ischaemia-reperfusionand cyclosporin,” Nephrology Dialysis Transplantation, vol.26, no. 2, pp. 489–498, 2011.

[102] K. Shimoishi, M. Anraku, K. Kitamura et al., “An oral adsor-bent, AST-120 protects against the progression of oxidativestress by reducing the accumulation of indoxyl sulfate inthe systemic circulation in renal failure,” PharmaceuticalResearch, vol. 24, no. 7, pp. 1283–1289, 2007.

[103] W. S. An, H. J. Kim, K. H. Cho, and N. D. Vaziri, “Omega-3 fatty acid supplementation attenuates oxidative stress,inflammation, and tubulointerstitial fibrosis in the remnantkidney,” American Journal of Physiology, vol. 297, no. 4, pp.F895–F903, 2009.

[104] O. Lopez-Franco, Y. Suzuki, G. Sanjuan et al., “Nuclearfactor-κB inhibitors as potential novel anti-inflammatoryagents for the treatment of immune glomerulonephritis,”American Journal of Pathology, vol. 161, no. 4, pp. 1497–1505,2002.

[105] C. K. Roberts, D. Won, S. Pruthi et al., “Effect of a short-term diet and exercise intervention on oxidative stress, in-flammation, MMP-9, and monocyte chemotactic activity inmen with metabolic syndrome factors,” Journal of AppliedPhysiology, vol. 100, no. 5, pp. 1657–1665, 2006.

[106] S. B. Kim, U. Eskiocak, P. Ly et al., “Bardoxolone-methyl(CDDO-Me): an antioxidant, antiinflammatory modulatoris a novel radiation countermeasure and mitigator,” inProceedings of the 22nd Annual NASA Space Radiation Inves-tigators’ Workshop, 2011.

[107] P. E. Pergola, M. Krauth, J. W. Huff et al., “Effect of bardox-olone methyl on kidney function in patients with T2D andstage 3b-4 CKD,” American Journal of Nephrology, vol. 33, no.5, pp. 469–476, 2011.

[108] P. E. Pergola, P. Raskin, R. D. Toto et al., “Bardoxolone methyland kidney function in CKD with type 2 diabetes,” The NewEngland Journal of Medicine, vol. 365, no. 4, pp. 327–336,2011.

[109] “Bardoxolone methyl evaluation in patients with chronickidney disease and type 2 diabetes: the ocurrence of renalevents (BEACON),” Identifier: NCT01351675, http://www.clinicaltrials.gov/.

Page 27: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Hindawi Publishing CorporationInternational Journal of NephrologyVolume 2012, Article ID 465897, 9 pagesdoi:10.1155/2012/465897

Review Article

Induction of Oxidative Stress in Kidney

Emin Ozbek

Department of Urology, Okmeydani Research & Education Hospital, Sisli, 34384 Istanbul, Turkey

Correspondence should be addressed to Emin Ozbek, [email protected]

Received 1 December 2011; Revised 27 January 2012; Accepted 6 February 2012

Academic Editor: Ayse Balat

Copyright © 2012 Emin Ozbek. This is an open access article distributed under the Creative Commons Attribution License, whichpermits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Oxidative stress has a critical role in the pathophysiology of several kidney diseases, and many complications of these diseasesare mediated by oxidative stress, oxidative stress-related mediators, and inflammation. Several systemic diseases such ashypertension, diabetes mellitus, and hypercholesterolemia; infection; antibiotics, chemotherapeutics, and radiocontrast agents;and environmental toxins, occupational chemicals, radiation, smoking, as well as alcohol consumption induce oxidative stress inkidney. We searched the literature using PubMed, MEDLINE, and Google scholar with “oxidative stress, reactive oxygen species,oxygen free radicals, kidney, renal injury, nephropathy, nephrotoxicity, and induction”. The literature search included only articleswritten in English language. Letters or case reports were excluded. Scientific relevance, for clinical studies target populations, andstudy design, for basic science studies full coverage of main topics, are eligibility criteria for articles used in this paper.

1. Introduction

Free radicals are chemical species with a single unpairedelectron, which is highly reactive as it seeks to pair with anew free electron, and as a result of these reactions, otherfree radicals or paired electrons occur and radical featuremay be lost. If newly formed free radical occurs, it is alsounstable and it can react with another molecule to produceanother free radical or a nonradical molecule occurs becauseof the paired electrons of the newly formed molecule. Thus,a chain reaction of free radicals occurs, leading to damagingbiological systems and tissues. In aerobic conditions, allbiological systems are exposed to oxidative stress (OS), eithergenerated internally or as by-products. The great majorityof these free radicals are mainly oxygen radicals and otherreactive oxygen species (ROS) [1].

The well-known ROS are superoxide ion (O2 • −), hy-drogen peroxide (H2O2), and peroxyl radical (OH•), and thereactive nitrogen species (RNS) are nitric oxide (NO) andperoxynitrite (ONOO−). Peroxynitrite generates from rapidchemical interaction between NO and O2•−. Main sites ofROS produced in living organisms are mitochondrial elec-tron transport system, peroxisomal fatty acid, cytochromeP-450, and phagocytic cells [2–6]. Several extracellular andintracellular factors such as hormones, growth factors,

proinflammatory cytokines, physical environmental factors(like ultraviolet irradiation), nutrient metabolism, and thedetoxification of various xenobiotics affect production of OS[7–13].

In physiological conditions, ROS produced in the courseof normal conditions are completely inactivated by cellularand extracellular defence mechanisms. This means thatnormally there is a balance between prooxidant (or oxidant)and antioxidant defence systems. In certain pathologicalconditions, increased generation of ROS and/or depletionof antioxidant defence system leads to enhanced ROSactivity and OS, resulting tissue damage. OS causes tissuedamage by different mechanisms including promoting lipidperoxidation, DNA damage, and protein modification. Theseprocesses have been implicated in the pathogenesis of severalsystemic diseases including kidney.

Several systemic diseases such as hypertension, diabetesmellitus, metabolic syndrome, and hypercholesterolemia; in-fection; antibiotics and chemotherapeutic agents, and radio-contrast agents mainly excreted from kidney; and envi-ronmental toxins especially heavy metals such as lead andmercury, occupational chemicals such as urban fine particles,radiation, smoking, as well as alcohol consumption inducerenal OS. The kidney is an organ highly vulnerable to

Page 28: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

2 International Journal of Nephrology

damage caused by ROS, likely due to the abundance of long-chain polyunsaturated fatty acids in the composition of renallipids. In recent years, OSs have become one of the mostpopular topics in research of molecular mechanism of renaldiseases. The aim of this paper is to summarize the condi-tions inducing OS in kidney and molecular mechanisms ofthis induction and kidney damage.

2. Diabetes Mellitus and Induction ofOxidative Stress in Kidney

Recent years, diabetes and diabetic kidney disease continueto increase worldwide. In the USA, diabetes-associated kid-ney disease is a major cause of all new cases of end-stagekidney disease. All diabetic patients are considered to be atrisk for nephropathy. Today we have not specific markersto expect development of end-stage renal disease. Clinicallycontrol of blood sugar level and blood pressure regulationsare important two parameters to the prevention of diabeticnephropathy [14, 15].

There are huge amount of in vitro and in vivo studiesregarding explanation of mechanism of diabetes-mellitus-induced nephropathy. All of these mechanisms are a conse-quence of uncontrolled elevation of blood glucose level. Cur-rently the proposed mechanism is the glomerular hyperfiltra-tion/hypertension hypothesis. According to this hypothesis,diabetes leads to increased glomerular hyperfiltration anda resultant increased glomerular pressure. This increasedglomerular pressure leads to damage to glomerular cellsand to development of focal and segmental glomeruloscle-rosis [16, 17]. Angiotensin II inhibitors reduce glomerularpressure and prevent albuminuria. Increased angiotensin IIlevel induces OS through activation of NADPH oxidase,stimulating inflammatory cytokines, and so forth . . . [18, 19].

The mechanism by which hyperglycemia causes freeradical generation thus causes OS to be complex. Increasedblood glucose promotes glycosylation of circulator andcellular protein and may initiate a series of autooxidationreactions that culminate in the formation and accumulationof advanced glycosylation end-products (AGEs) in tissues.The AGEs have oxidizing potential and promote tissuedamage by oxygen-free radicals [20].

In experimental studies, formation of OS increases be-cause of high level of blood glucose. Sadi et al. showed thatin diabetic rat kidney antioxidant enzyme, namely, catalase(CAT) and glutathion peroxidase (GSHPx), activities werefound to be reduced; however, ∝-lipoic acid and vitamin Cadministration increased these antioxidant enzyme activities[21]. Increased OS is the common finding in tissues effectingfrom diabetes, including kidney. Reddi et al. showed thattransforming growth factor β1 (TGF-β1) is prooxidant andSe (selenium) deficiency increases OS via this growth factor.In addition Se deficiency may simulate hyperglycemic con-ditions. Se supplementation to diabetic rats prevents forformation of OS and renal structural injury [22]. Chen etal. showed that nitrosative stress increases in diabetic ratmodel [23]. These results show the induction of oxidativeand nitrosative stress in rat kidney. These may have a role

in pathophysiology of diabetes-induced morphological andfunctional changes of kidney.

3. Hypertension, Hypercholesterolemia,Obesity, and Aging Induce OxidativeStress in Kidney

Hypertension is one of the major causes of development ofrenal failure. Key regulator of this pathology is OS. Renalartery stenosis is the most common cause of secondaryhypertension and may lead to deterioration of renal functionand ischemic nephropathy. Chade et al. showed that across-talk between hypoperfusion and atherosclerosis tointeractively increased OS, inflammation, and tubular injuryin the stenotic kidney [24]. In experimental atheroscleroticrenovascular disease (simulated by concurrent hypercholes-terolemia and renal artery stenosis), it was reported thatthe activity of both CuZn and MnSOD isoforms wassignificantly decreased; however, protein expression of boththe NAD (P)H-oxidase subunits p67phox and p47phox,nitrotyrosine, inducible nitric oxide synthase (iNOS), andnuclear factor kappa-B (NFκB) increased. Furthermore,tubular and glomerular protein expression of nitrotyrosineis significantly elevated [25]. Chronic blockade of OS withantioxidant improves OS in kidney. All of these molecularabnormalities suggest increased OS in rat kidney.

Noeman et al. showed that high-fat diet-induced obesityis accompanied by increased hepatic, cardiac, and renal tissueOS, which is characterized by reduction in the antioxidantenzymes activities and glutathione levels, that correlate withthe increase in MDA and protein carbonyl (PCO) levels [26].Increased cytokine release (inflammation-related cytokinessuch as tumor necrosis factor-∝ and adiponectin) and renalmacrophage infiltration have been shown to contribute torenal injury in models of obesity. Chow et al. reportedthat monocyte chemoattractant protein-1 (MCP-1) is apotent stimulator of macrophage recruitment. It is increasedin adipose tissue during obesity and in diabetic kidneys,suggesting that inflammation of these tissues may be MCP-1 dependent [27]. Knight et al. also showed an increase inrenal macrophage-specific CD68-positive staining in a modelof obesity and hypertension [28]. From these results, wecan say that macrophages are the source of increased OSand renal injury in diabetes and obesity-induced renalinjury. Aging is associated with increased OS. Most of age-dependent changes in the kidney such as excessive fibrosis, ageneral lack of regenerative ability, and an increase in apop-tosis in cells that determine healthy renal functions are oftenrelated to excess OS [29]. At a molecular level, with agingincreased mutations in nuclear and mitochondrial DNA(mtDNA), increased lipofuscin and AGEs, increased OS, andincreased apoptosis have been observed. Proximal tubularcells contain large numbers of mitochondria and are the mostreliant upon oxidative phosphorylation and most susceptibleto oxidant-induced apoptosis and mutations [30]. Recentstudies showed that antiaging gen, klotho, is important inrenal aging and OS-induced renal damage. The klotho geneencodes the klotho protein, a single transmembrane protein

Page 29: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 3

of the beta-glycosidase family [31]. Mice overexpressingklotho exhibit an extension in lifespan and resistance to oxi-dative injury. Klotho is predominantly expressed in thekidney, with its highest expression in cells of the distal con-voluted tubule [32, 33]. Overexpression of klotho has beenfound to enhance resistance to OS through the upregulationof manganese superoxide dismutase (MnSOD). Yamamotoet al. found that klotho modulates MnSOD in a FoxO-dependent process [34]. MnSOD is found within the mito-chondria where it acts as primary scavenger of oxidants.

4. Urinary Obstruction, Urolithiasis,Infection, Ischemia Reperfusion Injury,Transplantation of Kidney, and Inductionof Oxidative Stress in Kidney

Most clinical and experimental studies have shown that OSis increased in kidney and systemic circulation. Huang et al.reported that the activities of catalase and manganesesuperoxide dismutase were elevated in early stage of ethyleneglycol-induced urolithiasis model in rats; however, on day42 almost all antioxidant enzyme activities were attenuatedexcept those of CAT. In this experiment, the possible mecha-nism that causes free radical elevation in the kidney may bedifferent in the course of ethylene glycol-induced urolithiasis.Initially systemic circulation may bring the toxic substancesto the kidney, and eventually these substances cause to pro-duce free radicals. In the late stage progressive accumulationof leucocytes and defective antioxidant enzyme activitiesmay cause kidney to remain under huge amount of OS[35, 36]. In our experimental urolithiasis studies, we showeddecreased antioxidant enzyme activities and involvement ofNFκB and p38-MAPK (mitogen-activated protein kinase)signaling pathways, related to OS in rat kidney [37–40]. Inin vitro cell culture studies using proximal tubular origin linederived from pig proximal tubules (LLC-PK1), and collectingduct origin Madin-Darby canine kidney (MDCK) cell lines,it has been reported that calcium phosphate crystals causecellular injury by increasing ROS [41].

Today, extra corporeal shock wave lithotripsy (ESWL)is widely used for the treatment of renal stones in selectedrenal cases. In our work, we showed increased expression ofinducible NO synthase (iNOS) and NFκB, indirect evidenceof increased OS [42]. Recently, Gecit et al. showed thatESWL treatment produces OS and causes impairment in theantioxidant and trace element levels in the kidneys of rats[43]. ESWL is associated with greater prevalence of hyper-tension [44]. Ischemia insult and increased renal OS andconsequent endothelial dysfunction may be possible mech-anism of hypertension after ESWL.

Urinary obstruction, especially ureteral obstruction dueto urolithiasis, is a common urological problem seen inurology practice. Unilateral ureteral obstruction (UUO)leads to decreased renal MnSOD and CAT protein expressionin a time-dependent manner. Increased 4-hydroxynoneal (4-HNE) stain for ROS products in the renal tubulointerstitialcompartment occurred after 4 hr of UUO in the kidney. Theauthors explain renal tubular apoptosis in UUO rat model

explained by increasing ROS in this study [45]. Variousmarkers of OS increase in UUO rat kidneys such as the oxida-tively damaged protein product Ne-carboxymethyl-lysine(CML); the marker of DNA oxidant damage, 8-hydroxy-2′′-deoxyguanosine (8-OHdG)); and lipid peroxidation mark-ers such as malondialdehyde (MDA), 8-iso prostaglandinF2∝ (8-iPGF2∝), and 4-HNE or 4-hydroxy-hexenal (4-HHE). OS response molecules like heat shock protein-70(HSP-70), heat-shock protein-27, and heme oxygenase-1(HO-1) [46–51] are strongly expressed after UUO. Micethat are genetically deficient endogenous antioxidant enzymeCAT are more susceptible to UUO-induced renal damagethan normal wild type mice. Furthermore, increased renalconcentrations of ROS have been observed in obstructedkidneys, together with decreased activities of the majorprotective antioxidant enzymes SOD, CAT, and glutathioneperoxidase [52, 53]. UUO-induced nephrotoxicity and renalfibrosis is thought to be secondary to increased OS inkidney. In the literature there is some information about theamelioration of antioxidant and reactive oxygen scavengeragents against UUO-induced renal damage [54, 55]. For anexcellent detailed review, please refer to [56].

Infection is another inducer of OS in kidney. There area lot of experiments showing the increased oxidative stressand decreased antioxidant defence mechanism, antioxidantenzyme systems in kidney due to infection [57–60].

ROS are important mediators exerting toxic effectson various organs, including kidney during ischemia-reperfusion (IR) injury. A large body of evidences indicatethe role of increased OS in the kidney and protective roleof antioxidants and ROS scavengers in IR injury-inducednephropathy in the literature [61–63]. OS also has a role asa mediator of injury in chronic allograft tubular atrophy andinterstitial fibrosis in rat kidney [64]. Renal transplantationis another OS inducer in kidney in human and animals.OS increases in transplanted kidney because of pretransplantand posttransplant conditions. If there is preexisting diseasessuch as chronic kidney failure, inflammation, and diabetesmellitus, kidneys are more sensitive to OS during reperfusioninjury. Postoperative immunosuppressive agents are amongmany risk factors inducing OS in kidney [65].

5. Antibiotics, Antineoplastic Agents,Immunosuppressants, Analgesics,Nonsteroidal Antiinflammatory Drugs,and Radiocontrast Agents InduceOxidative Stress in Kidney

Antibiotics, commonly used aminoglycosides, are nephro-toxic agents. Their nephrotoxicity is mainly attributed toinduction of OS and depletion of antioxidat enzyme ac-tivities in kidney. In our experiments, we showed thatiNOS/NFκB/p38MAPK pathway, OS taking place in thisaxis, is involved in gentamicin-induced nephrotoxicity [66,67]. Our and other studies showed the protective effect ofanti oxidants and reactive oxygen scavenger agents againstgentamicin-induced nephrotoxicity [68–70].

Page 30: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

4 International Journal of Nephrology

Antineoplastic agents are commonly used for the treat-ment of metastatic cancers. Some of these are nephrotoxic.Excess ROS production and depressed antioxidant defencemechanism are responsible from nephrotoxicity. Cisplatin isthe well-known and commonly used antineoplastic andnephrotoxic agent. Other nephrotoxic anticancer agents arecarboplatin, methotrexate, doxorubicin, cyclosporine, andadriamycin. Immunosuppressant such as sirolimus and cy-closporine leads to nephrotoxicity via OS [71–80].

Cisplatin is one of the commonly used potent antitumordrugs and cisplatin-based combination protocols are used asfront-line therapy for several human malignancies. Cisplatinis toxic to the renal proximal tubules and dose dependent[81, 82]. Several studies have reported the role of OS regard-ing cisplatin-induced nephrotoxicity. Cisplatin is known toaccumulate in mitochondria of renal tubular epithelial cellstogether with ROS; renal tubular cell mitochondrial dysfunc-tion is also important in cisplatin-induced nephrotoxicity.Mitochondria also continuously scavenge ROS via the actionof antioxidant enzymes such as SOD, GSHPx, CAT, andglutathione S-transferase. Studies have demonstrated thatcisplatin induces ROS in renal epithelial cells primarilyby decreasing the activity of antioxidant enzymes andby depleting intracellular concentrations of GSH. In vitrostudies using LLC-PK1 cells, which is characteristic of renalproximal tubular epithelium, also showed the role of ROS incisplatin nephrotoxicity [83–89].

In this era, analgesics, especially paracetamol andacetaminophen (APAP), and nonsteroidal antiinflammatorydrugs (NSAIDs) are widely used throught the world. Parac-etamol and APAP are nephrotoxic drugs. Several in vitroand in vivo studies showed that analgesics nephrotoxicity iscaused by increased ROS in kidney. Zhao et al. showed theincreased ROS, nitric oxide, and MDA levels, together withdepleted glutathione (GSH) concentration in the kidney ofrats. However, rhein, Chinese herb, can attenuate APAP-in-duced nephrotoxicity in a dose-dependent manner [90]. Weshowed in our experiment a significant increase in MDAand decreases in GSHPx, CAT, and SOD activities in APAP-treated rat kidneys. These findings support the inductionof OS in rat kidney by APAP. Significant beneficial changeswere noted in serum and tissue OS indicators in rats treatedwith strong antioxidant pineal hormone melatonin andcurcumin [91, 92]. Ghosh et al. reported increased OSand TNF-alpha production in rat tissues [93]. Efrati et al.reported that diclofenac (NSAID) leads to nephrotoxicity byincreasing intrarenal ROS in rat kidney, and antioxidant, N-acetylcysteine, prevents kidney damage [94].

Contrast-induced nephropathy (CIN) is a major clinicalconcern, particularly with imaging procedures. CIN is thethird most common cause of acute kidney injury in hospi-talized patients [95]. Experimental findings in vitro and invivo illustrate enhanced hypoxia and the formation of ROSwithin the kidney following the administration of iodinatedcontrast media, which may play a role in the developmentof CIN. Studies indeed support this possibility, suggestinga protective effect of ROS scavenging or reduced ROSformation with the administration of N-acetyl cysteine andbicarbonate infusion, respectively [96–99].

6. Alcohol, Smoking, Environmental Toxins,Radiation, and Mobile Phones InduceOxidative Stress in Kidney

Ethanol and its metabolites are excreted into urine, and itscontent in the urine is higher than that of the blood andthe liver. Chronic alcohol administration decreases the renaltubular reabsorption and reduces renal function. Functionalabnormalities of renal tubules may be associated withethanol-induced changes in membrane composition andlipid peroxidation. Because of high content of long-chain-polyunsaturated fatty acids, kidney is highly sensitive to OSdamage [100].

Recently it is reported that ethanol administration causeda significant decrease in the levels of antioxidant enzymeCAT, SOD, and GSHPx activities and increases MDA inkidney of the rats [101]. Shankar et al. showed that renalmetabolism of ethanol via Cytochrome P450 2E1 (CYP2E1)and antidiuretic hormone-1 led to production of renal OS,and activation of MAPK induces CYP24A1 resulting inreducing circulating 1,25 (OH)2 D3 concentrations [102].Pathogenesis of aldosterone/salt-induced renal injury simi-larly is attributed to increased ROS and activation of MAPKin rat kidney [103]. In other studies, authors showed thatchronic ethanol administration and cigarette smoke expo-sure may cause renal injury by increasing oxidative andnitrosative stress in rat kidney [104, 105].

Epidemiological studies have shown that smoking is anaccelerating risk factor for the development of nephropathy,in which TGFβ1 plays a role in diabetic patients. Cell culturestudies using mesangial cell showed that smoking couldincrease TGFβ1, probably due to increased oxidative stressand PKCβ (protein kinase C beta) activation. This findingsupports the concept that smoking is a risk factor fordevelopment of diabetic nephropathy by increasing OS inkidney [106]. Similarly smoking and alcohol together in-crease OS and suppress antioxidant defence mechanism inkidney [104, 105].

In modern era, especially industrialised countries envi-ronmental toxins such as air pollutions, substances in storedfoods, radiations, as well as heavy metals in waters especiallyin underdeveloped countries are major health problem.Ochratoxin A (OTA), a mycotoxin, produced by fungiof improperly stored food products, has been linked tothe genesis of several disease states in both animals andhumans. It has been reported as nephrotoxic, carcinogenic,teratogenic, immunotoxic, and hepatotoxic in laboratoryand domestic animals [107–109]. In primary rat kidney cellsand in vivo experiments, it has been shown that OTA inducesOS and depletes antioxidant systems. These events mightrepresent pivotal factors in the chain of cellular events leadinginto nephrotoxicity of OTA. Cadmium (Cd) is known tobe a widespread environmental contaminant and a potentialtoxin that may adversely affect public health. Cigarette smokeand food (from contaminated soil and water) are importantnonindustrial sources of exposure to Cd. Cd accumulatesin the kidney because of its preferential uptake by receptor-mediated endocytosis of freely filtered and metallothionein

Page 31: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 5

bound Cd (Cd-MT) in the proximal tubule. Internalised Cd-MT is degraded in endosomes and lysosomes, releasing freeCd into the cytosol, where it can generate ROS and activatecell death pathways [110].

Another environmental nephrotoxic agent is diazi-non (O,O-diethyl-O-[2-isopropyl-6-methyl-4-pyrimidinyl]phosphorothioate). It is an organophosphate insecticide andhas been used worldwide in agriculture and domestic for sev-eral years. Shah et al. showed that diazinon exposure depletesantioxidant enzymes and induces OS in rat kidney [111].

Increased air pollution as a result of industry is anotherlife threatening health problem. Boor et al. showed renal,vascular, and cardiac fibrosis in rats exposed to passive smok-ing and industrial dust fibre amosite. Authors explain thesechanges by increased OS in these tissues [112]. Nanosizedfraction of particulate air pollutions have been reported totranslocate from the airways into the bloodstream and acton different organs such as lungs, heart, liver, and kidneys.Nemmar et al. examined the distribution and the patholog-ical changes of diesel exhaust particles (DEPs) on systolicblood pressure (SBP), systemic inflammation, oxidativestress, and morphological alterations in lungs, heart, liver,and kidneys in Wistar rats. They showed that DEPs causeinflammation especially in lungs and pulmonary tissue, andthese pathological changes are attributed to increased OSand inflammatory cytokines in these tissues [113]. Lead andcadmium nephrotoxicity are also related to increased OS inkidney [114, 115].

Radiation is an important inducer of OS. For diagnosticand therapeutic purposes, radiation is commonly used.Chronic OS after total body irradiation is thought to be thecause of radiation nephropathy in rats. Authors looked forevidence of OS after total-body irradiation in a rat model;focusing on the period before that there is physiologicallysignificant renal damage. No statistically significant increasein urinary 8-isoprostane (a marker of lipid peroxidation) orcarbonylated proteins (a marker of protein oxidation) wasfound over the first 42 days after irradiation, while a small butstatistically significant increase in urinary 8-hydroxydeoxy-guanosine (a marker of DNA oxidation) was detected at 35–55 days. In renal tissues, they found no significant increasein either DNA or protein oxidation products over the first89 days after irradiation. They suggest that if chronic OSis a part of the pathogenesis of radiation nephropathy,it does not leave widespread or easily detectable evidencebehind. Emre et al. investigated the effect of extremely low-frequency electromagnetic field (ELF-EMF) with pulse trainsexposure on lipid peroxidation and hence oxidative stress inthe rat liver and kidney tissue. They found increases in thelevels of oxidative stress indicators, and the flow cytometricdata suggested a possible relationship between the exposureto magnetic field and the cell death; however, there weresignificantly lower necrotic cell percentages in experimentalanimals compared to either unexposed or sham controlgroups [116]. These results suggest the inductive effect ofradiation on OS in kidney.

For the last two decades, a large number of studieshave investigated the effects of mobile phone radiation onthe human and animal. Cellular target and tissue damage

are different. Male reproductive system is among the mostaffected system [117, 118]. Increased OS plays a centralrole in radiofrequency-electromagnetic-waves- (RF-EMW-)induced tissue damage. Devrim et al. examined the effect ofRF-EMW on oxidant and antioxidant status in erythrocytesand kidney, heart, liver, and ovary tissues from rats andpossible protective role of vitamin C. It was observed thatMDA level, xanthine oxydase (XO), and GSH-Px activitiessignificantly increased in the EMR group as comparedwith those of the control group in the erythrocytes. Inthe kidney tissues, it was found that MDA level and CATactivity significantly increased, whereas XO and adenosinedeaminase (ADA) activities decreased in the cellular phonegroup as compared with those of the control group. However,in the heart tissues, it was observed that MDA level, ADA,and XO activities significantly decreased in the cellular phonegroup as compared with those of the control group. Theyconcluded that RF-EMR at the frequency generated by a cellphone causes OS and peroxidation in the erythrocytes andkidney tissues from rats. In the erythrocytes, vitamin C seemsto make partial protection against the OS [119]. Ozguneret al. reported similar results in rat experiments and theyalso reported that preventive effect of Caffeic acid phenethylester (CAPE), a flavonoid-like compound, is one of the majorcomponents of honeybee propolis and melatonin against RF-EMW-induced nephrotoxicity [120–122].

7. Conclusion

There is huge amount of literature concerning the linkbetween the OS and renal diseases. Systemic diseases such ashypertension, diabetes mellitus, and hypercholesterolemia;infection; antibiotics, chemotherapeutics, and radiocontrastagents; and environmental toxins, occupational chemicals,radiation, smoking, as well as alcohol consumption inducerenal OS. The kidney is a highly vulnerable organ to damagecaused by ROS, due to the abundance of long-chain-poly-unsaturated fatty acids. Antioxidant and reactive oxygenscavengers have been shown to be effective in animals forprotecting kidney, but it is hard to translocate these results tohumans. This may be due to short duration of animal studies,dose differences between animals and humans, and differentpathophysiologic processes between animals and humans.For understanding these steps, drugs will be developed toalter the main process(es) responsible for increased OS. Inthis paper, the conditions inducing OS in kidney andmolecular mechanisms of this induction and kidney damagehave been summarized. I hope that this paper will aid inthe understanding of this complex system and directing newresearch effort.

References

[1] N. Bashan, J. Kovsan, I. Kachko, H. Ovadia, and A. Rudich,“Positive and negative regulation of insulin signaling byreactive oxygen and nitrogen species,” Physiological Reviews,vol. 89, no. 1, pp. 27–71, 2009.

[2] R. S. Balaban, S. Nemoto, and T. Finkel, “Mitochondria,oxidants, and aging,” Cell, vol. 120, no. 4, pp. 483–495, 2005.

Page 32: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

6 International Journal of Nephrology

[3] K. B. Beckman and B. N. Ames, “The free radical theory ofaging matures,” Physiological Reviews, vol. 78, no. 2, pp. 547–581, 1998.

[4] B. Chance, H. Sies, and A. Boveris, “Hydroperoxide meta-bolism in mammalian organs,” Physiological Reviews, vol. 59,no. 3, pp. 527–605, 1979.

[5] K. Staniek and H. Nohl, “Are mitochondria a permanentsource of reactive oxygen species?” Biochimica et BiophysicaActa, vol. 1460, no. 2-3, pp. 268–275, 2000.

[6] H. Aguilaniu, L. Gustafsson, M. Rigoulet, and T. Nystrom,“Asymmetric inheritance of oxidatively damaged proteinsduring cytokinesis,” Science, vol. 299, no. 5613, pp. 1751–1753, 2003.

[7] S. Corda, C. Laplace, E. Vicaut, and J. Duranteau, “Rapidreactive oxygen species production by mitochondria in endo-thelial cells exposed to tumor necrosis factor-alpha is medi-ated by ceramide,” American Journal of Respiratory Cell andMolecular Biology, vol. 24, no. 6, pp. 762–768, 2001.

[8] C. M. Haynes, E. A. Titus, and A. A. Cooper, “Degradation ofmisfolded proteins prevents ER-derived oxidative stress andcell death,” Molecular Cell, vol. 15, no. 5, pp. 767–776, 2004.

[9] Y. Y. He and D. P. Hader, “UV-B-induced formation ofreactive oxygen species and oxidative damage of the Cyano-bacterium anabaena sp.: protective effects of ascorbic acidand N-acetyl-L-cysteine,” Journal of Photochemistry andPhotobiology, vol. 66, no. 2, pp. 115–124, 2002.

[10] A. Gomez-Mendikute and M. P. Cajaraville, “Comparativeeffects of cadmium, copper, paraquat and benzo[a]pyreneon the actin cytoskeleton and production of reactive oxygenspecies (ROS) in mussel haemocytes,” Toxicology in Vitro, vol.17, no. 5-6, pp. 539–546, 2003.

[11] A. Stankiewicz, E. Skrzydlewska, and M. Makieła, “Effectsof amifostine on liver oxidative stress caused by cyclophos-phamide administration to rats,” Drug Metabolism and DrugInteractions, vol. 19, no. 2, pp. 67–82, 2002.

[12] M. Sundaresan, Z. X. Yu, V. J. Ferrans, K. Irani, and T.Finkel, “Requirement for generation of H2O2 for platelet-derived growth factor signal transduction,” Science, vol. 270,no. 5234, pp. 296–299, 1995.

[13] M. Kanda, Y. Ihara, H. Murata et al., “Glutaredoxin modu-lates platelet-derived growth factor-dependent cell signalingby regulating the redox status of low molecular weight pro-tein-tyrosine phosphatase,” Journal of Biological Chemistry,vol. 281, no. 39, pp. 28518–28528, 2006.

[14] A. I. Adler, R. J. Stevens, S. E. Manley, R. W. Bilous, C. A.Cull, and R. R. Holman, “Development and progression ofnephropathy in type 2 diabetes: the United KingdomProspective Diabetes Study (UKPDS 64),” Kidney Interna-tional, vol. 63, no. 1, pp. 225–232, 2003.

[15] C. Robert and T. Stanton, “Oxidative stress and diabetickidney disease,” Current Diabetes Reports, vol. 11, pp. 330–336, 2011.

[16] S. Anderson and B. M. Brenner, “Pathogenesis of dia-betic glomerulopathy: hemodynamic considerations,” Dia-betes/Metabolism Reviews, vol. 4, no. 2, pp. 163–177, 1988.

[17] R. Zatz, B. R. Dunn, T. W. Meyer et al., “Prevention ofdiabetic glomerulopathy by pharmacological amelioration ofglomerular capillary hypertension,” Journal of Clinical Inves-tigation, vol. 77, no. 6, pp. 1925–1930, 1986.

[18] A. M. Garrido and K. K. Griendling, “NADPH oxidases andangiotensin II receptor signaling,” Molecular and CellularEndocrinology, vol. 302, no. 2, pp. 148–158, 2009.

[19] P. K. Mehta and K. K. Griendling, “Angiotensin II cell sig-naling: physiological and pathological effects in the cardio-vascular system,” American Journal of Physiology, vol. 292, no.1, pp. C82–C97, 2007.

[20] E. Mansouri, M. Panahi, M. A. Ghaffari, and A. Ghorbani,“Effects of grape seed proanthocyanidin extract on oxidativestress induced by diabetes in rat kidney,” Iranian BiomedicalJournal, vol. 15, no. 3, pp. 100–106, 2011.

[21] G. Sadi, N. Eryilmaz, E. Tutuncuoglu, S. Cingir, and T. Guray,“Changes in expression profiles of antioxidant enzymesin diabetic rat kidneys,” Diabetes/Metabolism Research andReviews. In press.

[22] A. S. Reddi and J. S. Bollineni, “Selenium-deficient dietinduces renal oxidative stress and injury via TGF-beta1 innormal and diabetic rats,” Kidney International, vol. 59, no.4, pp. 1342–1353, 2001.

[23] Y. J. Chen and J. Quilley, “Fenofibrate treatment of diabeticrats reduces nitrosative stress, renal cyclooxygenase-2 expres-sion, and enhanced renal prostaglandin release,” Journal ofPharmacology and Experimental Therapeutics, vol. 324, no. 2,pp. 658–663, 2008.

[24] A. R. Chade, M. Rodriguez-Porcel, J. P. Grande et al., “Dis-tinct renal injury in early atherosclerosis and renovasculardisease,” Circulation, vol. 106, no. 9, pp. 1165–1171, 2002.

[25] A. R. Chade, M. Rodriguez-Porcel, J. Herrmann et al.,“Beneficial effects of antioxidant vitamins on the stenotickidney,” Hypertension I, vol. 42, no. 4, pp. 605–612, 2003.

[26] S. A. Noeman, H. E. Hamooda, and A. A. Baalash, “Biochem-ical study of oxidative stress markers in the liver, kidney andheart of high fat diet induced obesity in rats,” Diabetology &Metabolic Syndrome, vol. 3, no. 1, pp. 1–17, 2011.

[27] F. Y. Chow, D. J. Nikolic-Paterson, F. Y. Ma, E. Ozols, B.J. Rollins, and G. H. Tesch, “Monocyte chemoattractantprotein-1-induced tissue inflammation is critical for thedevelopment of renal injury but not type 2 diabetes in obesedb/db mice,” Diabetologia, vol. 50, no. 2, pp. 471–480, 2007.

[28] S. F. Knight, J. Yuan, S. Roy, and J. D. Imig, “Simvastatinand tempol protect against endothelial dysfunction and renalinjury in a model of obesity and hypertension,” AmericanJournal of Physiology, vol. 298, no. 1, pp. F86–F94, 2010.

[29] A. M. Hall and R. J. Unwin, “The not so “mighty chondrion”:emergence of renal diseases due to mitochondrial dysfunc-tion,” Nephron, vol. 105, no. 1, pp. p1–p10, 2007.

[30] C. J. Percy, D. Power, and G. C. Gobe, “Renal ageing: changesin the cellular mechanism of energy metabolism and oxidanthandling,” Nephrology, vol. 13, no. 2, pp. 147–152, 2008.

[31] D. Telci, A. U. Dogan, E. Ozbek et al., “KLOTHO genepolymorphism of G395A is associated with kidney stones,”American Journal of Nephrology, vol. 33, no. 4, pp. 337–343,2011.

[32] H. Masuda, H. Chikuda, T. Suga, H. Kawaguchi, and M.Kuro, “Regulation of multiple ageing-like phenotypes by in-ducible klotho gene expression in klotho mutant mice,”Mechanisms of Ageing and Development, vol. 126, no. 12, pp.1274–1283, 2005.

[33] A. L. Negri, “The klotho gene: a gene predominantly ex-pressed in the kidney is a fundamental regulator of aging andcalcium/phosphorus metabolism,” Journal of Nephrology, vol.18, no. 6, pp. 654–658, 2005.

[34] M. Yamamoto, J. D. Clark, J. V. Pastor et al., “Regulation ofoxidative stress by the anti-aging hormone klotho,” Journal ofBiological Chemistry, vol. 280, no. 45, pp. 38029–38034, 2005.

Page 33: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 7

[35] H. S. Huang, M. C. Ma, J. Chen, and C. F. Chen, “Changesin the oxidant-antioxidant balance in the kidney of ratswith nephrolithiasis induced by ethylene glycol,” Journal ofUrology, vol. 167, no. 6, pp. 2584–2593, 2002.

[36] H. S. Huang, C. F. Chen, C. T. Chien, and J. Chen, “Pos-sible biphasic changes of free radicals in ethylene glycol-induced nephrolithiasis in rats,” British Journal of UrologyInternational, vol. 85, no. 9, pp. 1143–1149, 2000.

[37] Y. O. Ilbey, E. Ozbek, A. Simsek, M. Cekmen, A. Somay, andA. I. Tasci, “Pyrrolidine dithiocarbamate treatment preventsethylene glycol-induced urolithiasis through inhibition ofNF-kB and p38-MAPK signaling pathways in rat kidney,”Archivio Italiano di Urologia e Andrologia, vol. 82, no. 2, pp.87–94, 2010.

[38] Y. O. Ilbey, E. Ozbek, A. Simsek, M. Cekmen, A. Somay, andA. I. Tasci, “Effects of pomegranate juice on hyperoxaluria-induced oxidative stress in the rat kidneys,” Renal Failure, vol.31, no. 6, pp. 522–531, 2009.

[39] V. Tugcu, E. Kemahli, E. Ozbek et al., “Protective effect of apotent antioxidant, pomegranate juice, in the kidney of ratswith nephrolithiasis induced by ethylene glycol,” Journal ofEndourology, vol. 22, no. 12, pp. 2723–2731, 2008.

[40] V. Tugcu, E. Ozbek, E. Kemahli et al., “Rapid communi-cation: protective effect of a nuclear factor kappa B in-hibitor, pyrolidium dithiocarbamate, in the kidney of ratswith nephrolithiasis induced by ethylene glycol,” Journal ofEndourology, vol. 21, no. 9, pp. 1097–1106, 2007.

[41] K. Aihara, K. J. Byer, and S. R. Khan, “Calcium phosphate-induced renal epithelial injury and stone formation: involve-ment of reactive oxygen species,” Kidney International, vol.64, no. 4, pp. 1283–1291, 2003.

[42] M. Bas, V. Tugcu, E. Kemahli et al., “Curcumin preventsshock-wave lithotripsy-induced renal injury through inhi-bition of nuclear factor kappa-B and inducible nitric oxidesynthase activity in rats,” Urological Research, vol. 37, no. 3,pp. 159–164, 2009.

[43] I. Gecit, S. Kavak, I. Meral et al., “Effects of shock waves onoxidative stress, antioxidant enzyme and element levels inkidney of rats,” Biological Trace Element Research, vol. 144,no. 1–3, pp. 1069–1076, 2011.

[44] P. V. Barbosa, A. A. Makhlouf, D. Thorner, R. Ugarte, andM. Monga, “Shock wave lithotripsy associated with greaterprevalence of hypertension,” Urology, vol. 78, no. 1, pp. 22–25, 2011.

[45] C. H. Yeh, H. S. Chiang, T. Y. Lai, and C. T. Chien, “Unilateralureteral obstruction evokes renal tubular apoptosis via theenhanced oxidative stress and endoplasmic reticulum stressin the rat,” Neurourol Urodyn, vol. 30, no. 3, pp. 472–479,2011.

[46] N. Kawada, T. Moriyama, A. Ando et al., “Increased oxidativestress in mouse kidneys with unilateral ureteral obstruction,”Kidney International, vol. 56, no. 3, pp. 1004–1013, 1999.

[47] B. Pat, T. Yang, C. Kong, D. Watters, D. W. Johnson, and G.Gobe, “Activation of ERK in renal fibrosis after unilateralureteral obstruction: modulation by antioxidants,” KidneyInternational, vol. 67, no. 3, pp. 931–943, 2005.

[48] T. Moriyama, N. Kawada, K. Nagatoya, M. Horio, E. Imai,and M. Hori, “Oxidative stress in tubulointerstitial injury:therapeutic potential of antioxidants towards interstitialfibrosis,” Nephrology Dialysis Transplantation, vol. 15, supple-ment 6, pp. 47–49, 2000.

[49] H. Sugiyama, M. Kobayashi, D. H. Wang et al., “Telmisartaninhibits both oxidative stress and renal fibrosis after unilateral

ureteral obstruction in acatalasemic mice,” Nephrology Dial-ysis Transplantation, vol. 20, no. 12, pp. 2670–2680, 2005.

[50] A. Kamijo-Ikemori, T. Sugaya, A. Obama et al., “Liver-typefatty acid-binding protein attenuates renal injury inducedby unilateral ureteral obstruction,” American Journal ofPathology, vol. 169, no. 4, pp. 1107–1117, 2006.

[51] L. Meng, V. van Putten, L. Qu et al., “Altered expressionof genes profiles modulated by a combination of AstragaliRadix and Angelicae Sinensis Radix in obstructed rat kidney,”Planta Medica, vol. 76, no. 13, pp. 1431–1438, 2010.

[52] R. Sunami, H. Sugiyama, D. H. Wang et al., “Acatalasemiasensitizes renal tubular epithelial cells to apoptosis and exac-erbates renal fibrosis after unilateral ureteral obstruction,”American Journal of Physiology, vol. 286, no. 6, pp. F1030–F1038, 2004.

[53] W. Manucha, L. Carrizo, C. Ruete, H. Molina, and P. Valles,“Angiotensin II type I antagonist on oxidative stress andheat shock protein 70 (HSP 70) expression in obstructivenephropathy,” Cellular and Molecular Biology, vol. 51, no. 6,pp. 547–555, 2005.

[54] E. Ozbek, Y. O. Ilbey, M. Ozbek, A. Simsek, M. Cekmen,and A. Somay, “Melatonin attenuates unilateral ureteral ob-struction-induced renal injury by reducing oxidative stress,iNOS, MAPK, and NF-kB expression,” Journal of Endourol-ogy, vol. 23, no. 7, pp. 1165–1173, 2009.

[55] M. Zecher, C. Guichard, M. J. Velasquez, G. Figueroa, and R.Rodrigo, “Implications of oxidative stress in the pathophys-iology of obstructive uropathy,” Urological Research, vol. 37,no. 1, pp. 19–26, 2009.

[56] A. Dendooven, D. A. Ishola Jr., T. Q. Nguyen et al., “Oxidativestress in obstructive nephropathy,” International Journal ofExperimental Pathology, vol. 92, no. 3, pp. 202–210, 2011.

[57] L. Tothova, J. Hodosy, N. Kamodyova et al., “Bactofectionwith toll-like receptor 4 in a murine model of urinary tractinfection,” Current Microbiology, vol. 62, no. 6, pp. 1739–1742, 2011.

[58] C. H. Han, S. H. Kim, S. H. Kang et al., “Protective effectsof cranberries on infection-induced oxidative renal damagein a rabbit model of vesico-ureteric reflux,” British Journal ofUrology International, vol. 100, no. 5, pp. 1172–1175, 2007.

[59] S. Celik, S. Gorur, O. Aslantas, S. Erdogan, S. Ocak, and S.Hakverdi, “Caffeic acid phenethyl ester suppresses oxidativestress in Escherichia coli pyelonephritis in rats,” Molecular andCellular Biochemistry, vol. 297, no. 1-2, pp. 131–138, 2007.

[60] G. Sener, H. Tugtepe, A. Velioglu-Ogunc, S. Cetinel, N.Gedik, and B. C. Yegen, “Melatonin prevents neutrophil-mediated oxidative injury in Escherichia coli-induced pyelo-nephritis in rats,” Journal of Pineal Research, vol. 41, no. 3, pp.220–227, 2006.

[61] A. C. Seguro, L. F. Poli de Figueiredo, and M. H. Shimizu,“N-acetylcysteine (NAC) protects against acute kidney injury(AKI) following prolonged pneumoperitoneum in the rat,”Journal of Surgical Research. In press.

[62] P. Cassis, N. Azzollini, S. Solini et al., “Both darbepoetinalfa and carbamylated erythropoietin prevent kidney graftdysfunction due to ischemia/reperfusion in rats,” Transplan-tation, vol. 92, no. 3, pp. 271–279, 2011.

[63] E. Esposito, S. Mondello, R. di Paola et al., “Glutaminecontributes to ameliorate inflammation after renal ischemia/reperfusion injury in rats,” Naunyn-Schmiedeberg’s Archivesof Pharmacology, vol. 383, no. 5, pp. 493–508, 2011.

[64] A. Djamali, “Oxidative stress as a common pathwayto chronic tubulointerstitial injury in kidney allografts,”

Page 34: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

8 International Journal of Nephrology

American Journal of Physiology, vol. 293, no. 2, pp. F445–F455, 2007.

[65] M. Nafar, Z. Sahraei, J. Salamzadeh, S. Samavat, and N. D.Vaziri, “Oxidative stress in kidney transplantation: causes,consequences, and potential treatment,” Iranian Journal ofKidney Diseases, vol. 5, no. 6, pp. 357–372, 2011.

[66] E. Ozbek, M. Cekmen, Y. O. Ilbey, A. Simsek, E. C. Polat, andA. Somay, “Atorvastatin prevents gentamicin-induced renaldamage in rats through the inhibition of p38-MAPK and NF-kappaB pathways,” Renal Failure, vol. 31, no. 5, pp. 382–392,2009.

[67] V. Tugcu, E. Ozbek, A. I. Tasci et al., “Selective nuclearfactor kappa-B inhibitors, pyrolidium dithiocarbamate andsulfasalazine, prevent the nephrotoxicity induced by gentam-icin,” British Journal of Urology International, vol. 98, no. 3,pp. 680–686, 2006.

[68] A. Maniu, M. Perde-Schrepler, and M. Cosgarea, “Protectiveeffect of L-N-acetylcysteine against gentamycin ototoxicityin the organ cultures of the rat cochlea,” Romanian Journalof Morphology and Embryology, vol. 52, no. 1, pp. 159–164,2011.

[69] E. Ozbek, Y. Turkoz, E. Sahna, F. Ozugurlu, B. Mizrak,and M. Ozbek, “Melatonin administration prevents thenephrotoxicity induced by gentamicin,” British Journal ofUrology International, vol. 85, no. 6, pp. 742–746, 2000.

[70] E. Ozbek, Y. O. Ilbey, A. Simsek, M. Cekmen, F. Mete, andA. Somay, “Rosiglitazone, peroxisome proliferator receptor-gamma agonist, ameliorates gentamicin-induced nephrotox-icity in rats,” International Urology and Nephrology, vol. 42,no. 3, pp. 579–587, 2010.

[71] D. Giustarini, I. Dalle-Donne, E. Paccagnini, A. Milzani,and R. Rossi, “Carboplatin-induced alteration of the thiolhomeostasis in the isolated perfused rat kidney,” Archives ofBiochemistry and Biophysics, vol. 488, no. 1, pp. 83–89, 2009.

[72] V. Schmitz, J. Klawitter, J. Bendrick-Peart et al., “Metabolicprofiles in urine reflect nephrotoxicity of sirolimus andcyclosporine following rat kidney transplantation,” Nephron,vol. 111, no. 4, pp. e80–e91, 2009.

[73] V. K. Kolli, P. Abraham, B. Isaac, and D. Selvakumar, “Neu-trophil infiltration and oxidative stress may play a critical rolein methotrexate-induced renal damage,” Chemotherapy, vol.55, no. 2, pp. 83–90, 2009.

[74] T. A. Ajith, M. S. Aswathy, and U. Hema, “Protectiveeffect of Zingiber officinale roscoe against anticancer drugdoxorubicin-induced acute nephrotoxicity,” Food and Chem-ical Toxicology, vol. 46, no. 9, pp. 3178–3181, 2008.

[75] P. Galletti, C. I. di Gennaro, V. Migliardi et al., “Diverseeffects of natural antioxidants on cyclosporin cytotoxicity inrat renal tubular cells,” Nephrology Dialysis Transplantation,vol. 20, no. 8, pp. 1551–1558, 2005.

[76] K. P. Malarkodi, A. V. Balachandar, and P. Varalakshmi,“Protective effect of lipoic acid on adriamycin inducedlipid peroxidation in rat kidney,” Molecular and CellularBiochemistry, vol. 247, no. 1-2, pp. 9–13, 2003.

[77] A. Atessahin, A. O. Ceribasi, and S. Yilmaz, “Lycopene, acarotenoid, attenuates cyclosporine-induced renal dysfunc-tion and oxidative stress in rats,” Basic and Clinical Phar-macology and Toxicology, vol. 100, no. 6, pp. 372–376, 2007.

[78] N. Tirkey, G. Kaur, G. Vij, and K. Chopra, “Curcumin,a diferuloylmethane, attenuates cyclosporine-induced renaldysfunction and oxidative stress in rat kidneys,” BMCPharmacology, vol. 5, article 15, 2005.

[79] B. Naghizadeh, S. M. T. Mansouri, and N. V. Mashhadian,“Crocin attenuates cisplatin-induced renal oxidative stress in

rats,” Food and Chemical Toxicology, vol. 48, no. 10, pp. 2650–2655, 2010.

[80] P. Kalaiselvi, V. Pragasam, S. Chinnikrishnan, C. K. Veena,R. Sundarapandiyan, and P. Varalakshmi, “Counteractingadriamycin-induced oxidative stress by administration ofN-acetyl cysteine and vitamin E,” Clinical Chemistry andLaboratory Medicine, vol. 43, no. 8, pp. 834–840, 2005.

[81] A. D. Langerak and L. P. Dreisbach, Chemotherapy Regimensand Cancer Care, Landes Bioscience, Georgetown, Tex, USA,2001.

[82] M. H. Hanigan and P. Devarajan, “Cisplatin nephrotoxicity:molecular mechanisms,” Cancer Therapy, vol. 1, pp. 47–61,2003.

[83] C. Richter, V. Gogvadze, R. Laffranchi et al., “Oxidants inmitochondria: from physiology to diseases,” Biochimica etBiophysica Acta, vol. 1271, no. 1, pp. 67–74, 1995.

[84] M. Gemba and N. Fukuishi, “Amelioration by ascorbic acidof cisplatin-induced injury in cultured renal epithelial cells,”Contributions to nephrology, vol. 95, pp. 138–142, 1991.

[85] Q. Huang, R. T. Dunn, S. Jayadev et al., “Assessment ofcisplatin-induced nephrotoxicity by microarray technology,”Toxicological Sciences, vol. 63, no. 2, pp. 196–207, 2001.

[86] K. Husain, C. Morris, C. Whitworth, G. L. Trammell,L. P. Rybak, and S. M. Somani, “Protection by ebselenagainst cisplatin-induced nephrotoxicity: antioxidant sys-tem,” Molecular and Cellular Biochemistry, vol. 178, no. 1-2,pp. 127–133, 1998.

[87] M. Kruidering, B. van de Water, E. de Heer, G. J. Mulder,and J. F. Nagelkerke, “Cisplatin-induced nephrotoxicity inporcine proximal tubular cells: mitochondrial dysfunctionby inhibition of complexes I to iv of the respiratory chain,”Journal of Pharmacology and Experimental Therapeutics, vol.280, no. 2, pp. 638–649, 1997.

[88] Y. Kawai and M. Gemba, “Cisplatin-induced renal injury inLLC-PK1 cells,” in Proceedings of the 6th World Congress onAlternatives & Animal Use in the Life Sciences, pp. 453–456,Tokyo, Japan, August 2007.

[89] Y. Kawai, T. Nakao, N. Kunimura, Y. Kohda, and M. Gemba,“Relationship of intracellular calcium and oxygen radicals tocisplatin-related renal cell injury,” Journal of PharmacologicalSciences, vol. 100, no. 1, pp. 65–72, 2006.

[90] Y. L. Zhao, G. D. Zhou, H. B. Yang et al., “Rhein protectsagainst acetaminophen-induced hepatic and renal toxicity,”Food and Chemical Toxicology, vol. 49, no. 8, pp. 1705–1710,2011.

[91] Y. O. Ilbey, E. Ozbek, M. Cekmen et al., “Melatoninprevents acetaminophen-induced nephrotoxicity in rats,”International Urology and Nephrology, vol. 41, no. 3, pp. 695–702, 2009.

[92] M. Cekmen, Y. O. Ilbey, E. Ozbek, A. Simsek, A. Somay,and C. Ersoz, “Curcumin prevents oxidative renal damageinduced by acetaminophen in rats,” Food and ChemicalToxicology, vol. 47, no. 7, pp. 1480–1484, 2009.

[93] J. Ghosh, J. Das, P. Manna, and P. C. Sil, “Acetaminopheninduced renal injury via oxidative stress and TNF-alpha pro-duction: therapeutic potential of arjunolic acid,” Toxicology,vol. 268, no. 1-2, pp. 8–18, 2010.

[94] S. Efrati, S. Berman, Y. Siman-Tov et al., “N-acetylcysteineattenuates NSAID-induced rat renal failure by restor-ing intrarenal prostaglandin synthesis,” Nephrology DialysisTransplantation, vol. 22, no. 7, pp. 1873–1881, 2007.

[95] R. E. Cronin, “Contrast-induced nephropathy: pathogenesisand prevention,” Pediatric Nephrology, vol. 25, no. 2, pp. 191–204, 2010.

Page 35: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 9

[96] G. L. Bakris, N. Lass, A. O. Gaber, J. D. Jones, and J.C. Burnett Jr., “Radiocontrast medium-induced declines inrenal function: a role for oxygen free radicals,” AmericanJournal of Physiology, vol. 258, no. 1, pp. F115–F120, 1990.

[97] M. Araujo and W. J. Welch, “Oxidative stress and nitricoxide in kidney function,” Current Opinion in Nephrology andHypertension, vol. 15, no. 1, pp. 72–77, 2006.

[98] S. N. Heyman, S. Rosen, M. Khamaisi, J. M. Idee, and C.Rosenberger, “Reactive oxygen species and the pathogenesisof radiocontrast-induced nephropathy,” Investigative Radiol-ogy, vol. 45, no. 4, pp. 188–195, 2010.

[99] C. Haller and I. Hizoh, “The cytotoxicity of iodinatedradiocontrast agents on renal cells in vitro,” InvestigativeRadiology, vol. 39, no. 3, pp. 149–154, 2004.

[100] R. Rodrigo and G. Rivera, “Renal damage mediated byoxidative stress: a hypothesis of protective effects of redwine,” Free Radical Biology and Medicine, vol. 33, no. 3, pp.409–422, 2002.

[101] O. A. Adaramoye and A. Aluko, “Methanolic extractof Cnidoscolus aconitifolius attenuates renal dysfunctioninduced by chronic ethanol administration in Wistar rats,”Alcohol and Alcoholism, vol. 46, no. 1, Article ID agq082, pp.4–9, 2011.

[102] K. Shankar, X. Liu, R. Singhal et al., “Chronic ethanolconsumption leads to disruption of vitamin D3 homeostasisassociated with induction of renal 1,25 dihydroxyvitaminD3-24-hydroxylase (CYP24A1),” Endocrinology, vol. 149, no.4, pp. 1748–1756, 2008.

[103] A. Nishiyama, L. Yao, Y. Nagai et al., “Possible contributionsof reactive oxygen species and mitogen-activated proteinkinase to renal injury in aldosterone/salt-induced hyperten-sive rats,” Hypertension, vol. 43, no. 4, pp. 841–848, 2004.

[104] Y. Cigremis, Y. Turkoz, M. Tuzcu et al., “The effects ofchronic exposure to ethanol and cigarette smoke on theformation of peroxynitrite, level of nitric oxide, xanthineoxidase and yeloperoxidase activities in rat kidney,” Molecularand Cellular Biochemistry, vol. 291, no. 1-2, pp. 127–138,2006.

[105] Y. Cigremis, Y. Turkoz, M. Akgoz, and M. Sozmen, “Theeffects of chronic exposure to ethanol and cigarette smokeon the level of reduced glutathione and malondialdehyde inrat kidney,” Urological Research, vol. 32, no. 3, pp. 213–218,2004.

[106] C. Mur, J. Claria, S. Rodela et al., “Cigarette smoke concen-trate increases 8-epi-PGF2alpha and TGFbeta1 secretion inrat mesangial cells,” Life Sciences, vol. 75, no. 5, pp. 611–621,2004.

[107] E. O’Brien and D. R. Dietrich, “Ochratoxin A: the continuingenigma,” Critical Reviews in Toxicology, vol. 35, no. 1, pp. 33–60, 2005.

[108] H. G. Kamp, G. Eisenbrand, J. Schlatter, K. Wurth, andC. Janzowski, “Ochratoxin A: induction of (oxidative) DNAdamage, cytotoxicity and apoptosis in mammalian cell linesand primary cells,” Toxicology, vol. 206, no. 3, pp. 413–425,2005.

[109] A. M. Domijan, M. Peraica, A. L. Vrdoljak, B. Radic, V.Zlender, and R. Fuchs, “The involvement of oxidative stressin ochratoxin A and fumonisin B 1 toxicity in rats,” MolecularNutrition and Food Research, vol. 51, no. 9, pp. 1147–1151,2007.

[110] N. Johri, G. Jacquillet, and R. Unwin, “Heavy metal poison-ing: the effects of cadmium on the kidney,” BioMetals, vol. 23,no. 5, pp. 783–792, 2010.

[111] M. D. Shah and M. Iqbal, “Diazinon-induced oxidative stressand renal dysfunction in rats,” Food and Chemical Toxicology,vol. 48, no. 12, pp. 3345–3353, 2010.

[112] P. Boor, S. Casper, P. Celec et al., “Renal, vascular andcardiac fibrosis in rats exposed to passive smoking and indus-trial dust fibre amosite,” Journal of Cellular and MolecularMedicine, vol. 13, no. 11-12, pp. 4484–4491, 2009.

[113] A. Nemmar, S. Al-Salam, S. Zia, S. Dhanasekaran, M.Shudadevi, and B. H. Ali, “Time-course effects of systemicallyadministered diesel exhaust particles in rats,” ToxicologyLetters, vol. 194, no. 3, pp. 58–65, 2010.

[114] L. G. Navarro-Moreno, M. A. Quintanar-Escorza, S. Gon-zalez et al., “Effects of lead intoxication on intercellularjunctions and biochemical alterations of the renal proximaltubule cells,” Toxicology in Vitro, vol. 23, no. 7, pp. 1298–1304, 2009.

[115] L. Wang, J. Cao, D. Chen, X. Liu, H. Lu, and Z. Liu, “Roleof oxidative stress, apoptosis, and intracellular homeostasisin primary cultures of rat proximal tubular cells exposed tocadmium,” Biological Trace Element Research, vol. 127, no. 1,pp. 53–68, 2009.

[116] M. Lenarczyk, E. P. Cohen, B. L. Fish et al., “Chronicoxidative stress as a mechanism for radiation nephropathy,”Radiation Research, vol. 171, no. 2, pp. 164–172, 2009.

[117] N. R. Desai, K. K. Kesari, and A. Agarwal, “Pathophysiologyof cell phone radiation: oxidative stress and carcinogenesiswith focus on male reproductive system,” Reproductive Biol-ogy and Endocrinology, vol. 22, no. 7, p. 114, 2009.

[118] A. Agarwal, N. R. Desai, K. Makker et al., “Effects ofradiofrequency electromagnetic waves (RF-EMW) from cel-lular phones on human ejaculated semen: an in vitro pilotstudy,” Fertility and Sterility, vol. 92, no. 4, pp. 1318–1325,2009.

[119] E. Devrim, I. B. Ergude, B. Kılıcoglu, E. Yaykaslı, R. Cetin,and I. Durak, “Effects of electromagnetic radiation use onoxidant/antioxidant status and DNA turn-over enzyme activ-ities in erythrocytes and heart, kidney, liver, and ovary tissuesfrom rats: possible protective role of vitamin C,” ToxicologyMechanisms and Methods, vol. 18, no. 9, pp. 679–683, 2008.

[120] M. F. Ozguner, F. Oktem, A. Ayata, A. Koyu, and H. R. Yilmaz,“A novel antioxidant agent caffeic acid phenethyl esterprevents long-term mobile phone exposure-induced renalimpairment in rat. Prognostic value of malondialdehyde, N-acetyl-beta-d-glucosaminidase and nitric oxide determina-tion,” Molecular and Cellular Biochemistry, vol. 277, no. 1-2,pp. 73–80, 2005.

[121] F. Ozguner, F. Oktem, A. Armagan et al., “Comparativeanalysis of the protective effects of melatonin and caffeicacid phenethyl ester (CAPE) on mobile phone-induced renalimpairment in rat,” Molecular and Cellular Biochemistry, vol.276, no. 1-2, pp. 31–37, 2005.

[122] F. Oktem, F. Ozguner, H. Mollaoglu, A. Koyu, and E. Uz,“Oxidative damage in the kidney induced by 900-MHz-emitted mobile phone: protection by melatonin,” Archives ofMedical Research, vol. 36, no. 4, pp. 350–355, 2005.

Page 36: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Hindawi Publishing CorporationInternational Journal of NephrologyVolume 2012, Article ID 608397, 11 pagesdoi:10.1155/2012/608397

Research Article

Inflammation and Oxidative Stress inObesity-Related Glomerulopathy

Jinhua Tang,1, 2 Haidong Yan,1 and Shougang Zhuang1, 2

1 Department of Nephrology, Shanghai East Hospital, Tongji University School of Medicine, Shanghai 200120, China2 Department of Medicine, Alpert Medical School of Brown University, Rhode Island Hospital-Middle House 301,593 Eddy Street, Providence, RI 02903, USA

Correspondence should be addressed to Haidong Yan, [email protected] and Shougang Zhuang, [email protected]

Received 2 September 2011; Accepted 6 February 2012

Academic Editor: Ayse Balat

Copyright © 2012 Jinhua Tang et al. This is an open access article distributed under the Creative Commons Attribution License,which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Obesity-related glomerulopathy is an increasing cause of end-stage renal disease. Obesity has been considered a state of chroniclow-grade systemic inflammation and chronic oxidative stress. Augmented inflammation in adipose and kidney tissues promotesthe progression of kidney damage in obesity. Adipose tissue, which is accumulated in obesity, is a key endocrine organ thatproduces multiple biologically active molecules, including leptin, adiponectin, resistin, that affect inflammation, and subsequentderegulation of cell function in renal glomeruli that leads to pathological changes. Oxidative stress is also associated with obesity-related renal diseases and may trigger the initiation or progression of renal damage in obesity. In this paper, we focus oninflammation and oxidative stress in the progression of obesity-related glomerulopathy and possible interventions to preventkidney injury in obesity.

1. Introduction

Obesity has become a heavy public health problem in theUnited States, with a prevalence among adults increasing to32% from 13% between the 1960s and 2004 [1]. Currently,66% of adults and 16% of children and adolescents are over-weight or obese [1]. Although obesity has long been recog-nized as an independent risk factor for cardiovascular dis-eases and diabetes mellitus, newer research points to obesityas an important risk factor for chronic kidney diseases(CKDs) [2–4]. In 1974, Weisinger et al. [5] firstly reportedthat massive obese patients developed nephrotic-range pro-teinuria. Subsequent studies confirmed that obesity couldinduce renal injury, namely, obesity-related glomerulopathy(ORG) [6–8]. A large-scale clinicopathologic study including6818 renal biopsies from 1986 to 2000 revealed a progressiveincrease in biopsy incidence of ORG from 0.2% in 1986–1990to 2.0% in 1996–2000 [8]. The tenfold increase in incidenceof ORG over 15 years suggests a newly emerging epidemic[8].

The clinical characteristics of subjects with ORG typi-cally manifest with nephrotic or subnephrotic proteinuria,

accompanied by renal insufficiency [8–10]. Histologically,ORG presents as focal segmental glomerulosclerosis (FSGS)and glomerular hypertrophy or glomerular hypertrophyalone and relatively decreased podocyte density and numberand mild foot process fusion [8, 11, 12]. Clinically, it isdistinguished from idiopathic FSGS (I-FSGS) by its lowerincidence of nephrotic syndrome, more benign course, andslower progression of proteinuria and renal failure [8, 11].

ORG is an increasing cause of end-stage renal disease(ESRD). The pathophysiology of ORG remains incomplete-ly understood. Potential mechanisms by which obesity af-fects renal physiology include altered renal hemodynam-ics, insulin resistance, hyperlipidemia, activation of renin-angiotensin-aldosterone system (RAAS), inflammation, andoxidative stress. Increases in both glomerular filtration rate(GFR) and renal plasma flow (RPF) were observed inobese subjects and animals [13, 14]. This likely occurs be-cause of afferent arteriolar dilation as a result of proximalsalt reabsorption, coupled with efferent renal arteriolar vaso-constriction as a result of elevated angiotensin II (AngII)[15]. These effects may contribute to hyperfiltration, glom-erulomegaly, and later focal glomerulosclerosis [8, 9]. Insulin

Page 37: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

2 International Journal of Nephrology

resistance can raise the transcapillary pressure gradient andcause hydrostatic pressure and hyperfiltration by reducingnorepinephrine-induced efferent arteriolar constriction [16],leading to glomerular hypertrophy and sclerosis. Hyperin-sulinemia also has been shown to stimulate the synthe-sis of growth factors such as insulin-like growth factor-(IGF-) 1 and IGF-2 and transforming growth factor-β1(TGF-β1), which accelerate production of extracellularmatrix and promote glomerular hypertrophy and sclerosis[17, 18]. Hyperlipidemia may promote glomerulosclerosisthrough mechanisms that involve engagement of low-densitylipoprotein receptors on mesangial cells, direct podocytetoxicity, oxidative cellular injury, macrophage chemotaxis,and increase renal expression of sterol regulatory element-binding proteins (SREBP-1 and SREBP-2), resulting in therenal accumulation of cholesterol and triglycerides and to-gether with significant renal increase of fibrogenic cytokines[19, 20]. Adipose tissue is the major source of the com-ponents of RAAS. Obese subjects usually have increasesin plasma renin activity, angiotensinogen, angiotensin-con-verting enzyme activity, and circulating AngII, which triggeror promote renal damage by renal hemodynamic changesand nonhemodynamic pathways such as hyperinsulinemia,oxidative stress, and inflammation [21–23]. Inflammatoryabnormalities and oxidative stress are characteristic findingsof obesity and play important roles in the renal damage as-sociated with obesity, which will be discussed in detail in thefollowing.

2. Inflammation in Obesity-RelatedGlomerulopathy

Recent studies have demonstrated that obesity causes chroniclow-grade systemic inflammation and thus contributes tothe development of systemic metabolic dysfunction that isassociated with obesity-related disorders and renal disease[24–27]. Levels of some inflammatory markers and cytokinessuch as C-reactive protein (CRP), tumor necrosis factor-α(TNF-α), interleukin-6 (IL-6), and macrophage migrationinhibitory factor (MIF) are elevated, whereas concentra-tions of adiponectin, a protein hormone that exerts anti-inflammatory activities, are reduced in obesity [28–34].

2.1. Leptin. Leptin is a 16-kDa-peptide hormone encoded byobese (ob) gene that is mainly produced by adipose tissue.Leptin serves as a regulator of energy balance by binding tothe full-length leptin receptors—obese receptor b (Ob-Rb)in the hypothalamus, leading to reduction in food intakeand elevation in temperature and energy expenditure [35].Leptin receptors can be classified as secreted-forms (Ob-Re),short-forms (ob-Ra, c, d, and f) mainly expressed in periph-eral tissue, and long-forms (ob-Rb) predominantly expressedin hypothalamus. The kidney expresses abundant concentra-tions of the truncated isoform of the leptin receptor Ob-Ra, but only a small amount of the full-length receptorOb-Rb [36]. Leptin production is associated with increasedsize of adipocytes and is positively correlated with thebody mass index (BMI) [37]. Increased circulating leptin, a

marker of leptin resistance, is common in obesity. Obesity-induced leptin resistance injures numerous peripheral tissuesincluding kidney, liver, myocardium, and vasculature [36,38]. Leptin results in the development of renal disease bybinding to its specific receptors in renal endothelial cells andmesangial cells. In glomerular endothelial cells, leptin stimu-lates cellular proliferation, TGF-β1 synthesis, and type IV col-lagen production [36, 39]. In mesangial cells, leptin upregu-lates synthesis of the TGF-β type II receptor, but not TGF-β1, and stimulates glucose transport and type I collagen pro-duction through signal transduction pathways involvingphosphatidylinositol-3-kinase [40]. Leptin also stimulateshypertrophy, but not proliferation in cultured rat mesangialcells [41]. However, both those cell types increase their ex-pression of extracellular matrix in response to leptin. Trans-genic mice with leptin overexpression demonstrated an in-crease in collagen type IV and fibronectin mRNA in thekidney [41]. Leptin is involved in the development ofglomerulosclerosis through a paracrine TGF-β pathway (be-tween glomerular endothelial and mesangial cells) that pro-motes the deposition of extracellular matrix, proteinuria,and, eventually, glomerulosclerosis [39]. Infusion of leptininto normal rats for 3 weeks fosters the development of focalglomerulosclerosis and proteinuria [36].

Leptin also has proinflammatory actions through itsinteraction with mediators of innate and adaptive immunityand CRP [38]. Leptin regulates components of innate andadaptive immunity, including T lymphocytes and mono-cytes/macrophages [42, 43]. Central leptin administrationin ob/ob mice accelerates renal macrophage infiltrationthrough the melanocortin system [44]. Leptin stimulatescentral T-cell production and a peripheral shift in favor ofT helper (Th) 1 adaptive immune responses (proinflamma-tory) as opposed to Th2 responses (anti-inflammatory) [38].Leptin has been shown to modulate adaptive immunity byenhancing T-cell survival and stimulating production of pro-inflammatory cytokines such as IFN-γ and IL-2 [45]. Leptinalso has structural and functional resemblance to proinflam-matory cytokines, such as IL-6 [42], and may modulate CRP,a leptin-interacting protein [46].

Therefore, these direct and indirect effects of leptin onthe kidney, including stimulating cellular proliferation andhypertrophy, increasing extracellular matrix expression, andexhibiting proinflammatory activities, may partially explainobesity-related kidney disease.

2.2. Adiponectin. Adiponectin is a 30 kDa adipocyte-derivedprotein hormone encoded by the adipose most abundantgene transcript 1 (APM1), [47] which plays a role in thesuppression of inflammation-associated metabolic disorders.Two distinct adiponectin receptors, AdipoR1 and R2, havebeen cloned [48]. AdipoR1 is expressed ubiquitously whileAdipoR2 is most abundant in the liver. Adiponectin is highlyabundant in human serum, but its level is decreased inmost obese animal and human subjects, particularly in thosewith visceral obesity [49–51]. Recent clinical studies show anegative association of adiponectin in obese patients, [52,53] suggesting that adiponectin may play a key role in the

Page 38: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 3

development of obesity-related albuminuria and alterationof renal function.

Studies with the adiponectin knockout mouse provideevidence that adiponectin can regulate podocyte functionand thus contribute to the initial development of albumin-uria [37, 53]. Sharma et al. showed that knockout of adipo-nectin in mice increased albuminuria and caused fusionof podocyte foot processes [53]. In cell culture studieswith podocytes, incubation with adiponectin potently de-creased permeability to albumin, induced translocation ofzona occludens-1 (ZO-1) to the plasma membrane, and re-duced the renal predominant NADPH oxidase Nox 4,largely via a 5′-AMP-activated-protein kinase- (AMPK-)dependent pathway. Treatment of the adiponectin knockoutmice with exogenous adiponectin was able to decrease albu-minuria and improve podocyte morphology [53]. Chronichyperadiponectinemia significantly alleviated the progres-sion of proteinuria in early-stage diabetic nephropathy byseveral mechanisms. It led to an increase in nephrin expres-sion, improvement of the endothelial dysfunction due todecreases in endothelin 1 (ET-1) and plasminogen activatorinhibitor 1 (PAI-1), and an increase in endothelial nitricoxide synthase (eNOS) expression in the renal cortex [54].

Recent studies suggest that adiponectin exerts anti-in-flammatory effects by suppressing TNF-α-induced activationof nuclear factor-κB (NF-κB) in human aortic endothelialcells and aortic smooth muscle cells through inhibition ofIκB phosphorylation [55, 56] and inhibition of vascular celladhesion molecule 1 (VCAM-1) and intercellular adhesionmolecule 1 (ICAM-1) expression [57], thereby reducing mo-nocyte adhesion and macrophage-induced cytokine produc-tion [58] and CRP expression in human adipose tissue [59].Human adipose tissue expressed CRP, which was negativelycorrelated with adiponectin expression in adipose tissue [59].Low levels of adiponectin are associated with higher levelsof highly-sensitive C-reactive protein (hs-CRP) and IL-6[49], two inflammatory mediators that are involved in theinitiation and progression of atherosclerosis and renal dis-ease. Therefore, hypoadiponectinemia contributes to devel-opment of a low-grade systemic chronic inflammation state,suggesting that hypoadiponectinemia may play a causativerole in the systemic and vascular inflammation commonlyfound in obesity and obesity-related disorders, including re-nal injury, through its proinflammatory effects.

2.3. Resistin. Resistin, also known as adipocyte-specificsecretory factor (ADSF) or as found in inflammatory zone(Fizz), is a cysteine-rich 12.5-kDa polypeptide that belongsto a small family called resistin-like molecules (RELMs) [60,61]. In rodents, resistin is secreted from white adipocytes [62,63]. In human, it is produced largely by macrophages andexpressed in adipose tissue predominantly by nonadipocyteresident inflammatory cells [64–66]. Current evidence sug-gests that resistin has been variably associated with obesity,insulin resistance, inflammation, and renal dysfunction.Resistin levels are elevated in both genetic and diet-inducedanimal models of obesity [62, 67]. Studies of obese subjectshave frequently noted higher serum levels of resistin as wellas direct correlations between resistin level and adiposity as

measured by BMI [68, 69]. There has been a link betweencirculating resistin and low-grade inflammation that accom-pany obesity [70]. Resistin is associated with elevated CRPand white blood cells, suggesting that the role of resistinmay be a component of obesity-related inflammation [70].It has recently been found that resistin involves in theregulation of proinflammatory cytokine expression. Resistinstrongly upregulates IL-6 and TNF-α in human peripheralblood mononuclear cells (PBMCs) via NF-κB pathway [71].Human resistin enhanced secretion of proinflammatory cy-tokines, TNF-α and IL-12 in macrophages by NF-κB-de-pendent pathway [72]. Studies also show that increased levelsof resistin in patients with CKD are associated with declinedrenal function and inflammation [73, 74]. This suggests thatresistin may play an important role in obesity and obesity-as-sociated disease by triggering the release of other proinflam-matory cytokines.

2.4. Inflammatory Markers. A growing body of evidence in-dicates that obesity-related glomerulopathy is associatedwith upregulation of inflammatory mediators [75]. Obesityleads to adipose tissue macrophage infiltration in white adi-pose tissue and increased levels in proinflammatory cyto-kines. Several inflammatory mediators released from adipo-cytes and macrophages, such as TNF-α, IL-6, IL-1β, CRP,monocyte chemoattractant protein 1 (MCP-1), PAI-1, andMIF, contribute to a low level of chronic inflammatory statein obesity and may be responsible for renal injury in obesity-associated glomerulopathy. An emerging pattern of geneexpression was observed in adipose tissue in mice fed highfat-fed diets, indicating a shift toward global upregulationof inflammatory genes, including TNF-α, IL-6, and MCP-1[76].

TNF-α, a proinflammatory cytokine, is predominantlyproduced by macrophages infiltrating adipose tissue [77, 78]and can also be produced by the kidney [79]. This cytokineis involved in the genesis of inflammation and contributesto obesity-associated insulin resistance [80–82]. Within thekidney, AngII, advanced glycation end-products (AGEs), andoxidized low-density lipoprotein (LDL) can stimulate TNF-αsynthesis from renal cells to initiate local damage [83–86]. Arecent study demonstrates that TNF-α reduces the expressionof Klotho, a protein expressed by renal cells, through an NF-κB-dependent mechanism, which contribute to renal injury[87]. TNF-α enhances the expression of PAI-1 in human adi-pose tissue and plasma PAI-1 levels in obesity subjects andis responsible for reduced fibrinolysis and also a componentof extracellular matrix, leading to renal fibrosis and terminalrenal failure [88–90]. TNF-α also has been shown to inducethe expression of MCP-1 via p38 mitogen-activated proteinkinase (MAPK) signaling pathway in renal mesangial cells[91]. MCP-1, a key regulator in recruiting monocytes to theglomeruli, may also contribute to renal damage at a laterstage of kidney disease in obesity.

IL-6 is another important proinflammatory mediatorsystemically secreted from adipose tissue and locally pro-duced in the kidney. Studies have demonstrated the positiverelationship between BMI and plasma IL-6 concentrations[92–94]. Studies also suggest that IL-6 plays a key role in

Page 39: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

4 International Journal of Nephrology

the development of renal disease. Endogenous IL-6 enhancesthe degree of renal injury, dysfunction, and inflammationcaused by ischemia/reperfusion by promoting the expressionof adhesion molecules and subsequent oxidative stress [95].Transgenic knockout of IL-6 ameliorates renal injury asmeasured by serum creatinine and histology [96]. Blockingthe IL-6 receptor prevents progression of proteinuria andrenal lipid deposit as well as the mesangial cell proliferationassociated with severe hyperlipoproteinemia [97]. IL-6 alsostimulates the synthesis of CRP [98], which is well known asboth a marker and important risk factor of atherosclerosis inthe general population and CKD patients [99].

MIF was initially described as an immunomodulatoryfactor isolated from the supernatants of T lymphocytes andwas found to inhibit the random migration of macrophages[100]. Subsequent studies have indicated that MIF acts asa proinflammatory cytokine and pituitary-derived hormonethat potentiates endotoxemia [101]. MIF also plays a path-ogenic role in kidney disease. In a rat model of immuno-logically induced crescentic antiglomerular basement mem-brane glomerulonephritis, treatment with anti-MIF anti-body reduced proteinuria, prevented the loss of renal func-tion, attenuated histological damage including glomerularcrescent formation, inhibited renal leukocytic infiltrationand activation, and reduced IL-1β expression by both intrin-sic kidney cells and macrophages [102]. Thus, MIF is a keymediator of the inflammatory and immune response andplays a pathological role in immune-mediated renal injury.

3. Oxidative Stress inObesity-Related Glomerulopathy

3.1. Obesity and Oxidative Stress. Oxidative stress is causedby an imbalance between increased production of reactiveoxygen species (ROS) and/or reduced antioxidant activity,leading to oxidative damage to cells or tissue including lip-ids, proteins and DNA. It is known that oxidative stress isinvolved in pathological processes of various diseases, suchas cancer, diabetes mellitus, hypertension, and cardiovasculardisease. Studies have suggested that obesity is associated withincreased oxidative stress [103, 104]. Analysis of oxidativemarkers in obesity subjects indicates that oxidative damageis associated with increased BMI and percentage of body fat[105, 106]. Conversely, parameters of antioxidant capacityare inversely related to the amount of body fat and centralobesity [107, 108]. The possible mechanisms of obesity-related oxidative stress include increased oxygen consump-tion and subsequent production of free radicals derivedfrom the increase in mitochondrial respiration, diminishedantioxidant capacity, fatty acid oxidation, lipid oxidizability,and cell injury causing increased rates of free radical form-ation [104, 109, 110]. It is also reported that the increasein obesity-associated oxidative stress is due to the presenceof excessive adipose tissue accumulation [111]. Accumulatedadipose tissue generates an immune response leading to thesecretion of proinflammatory cytokines, including TNF-α,IL-1, and IL-6, which lead to increased generation of ROS.Excessive fat accumulation also stimulates nicotinamide

adenine dinucleotide phosphate (NADPH) oxidase activity,which contributes to ROS production [112]. ROS, in return,augmented the expressions of NADPH oxidase (NOX) sub-units, including NOX4 and PU.1 in adipocytes, establishing avicious cycle that augments oxidative stress in white adiposetissue and blood [112].

3.2. Oxidative Stress and Inflammation. Oxidative stress inadipocyte seems to be responsible for the low-grade proin-flammatory state commonly observed in obesity [113, 114].Oxidative stress is increasingly viewed as a major upstreamcomponent in cell-signaling cascades involved in inflamma-tory responses, stimulating the expression of proinflamma-tory cytokines. ROS activate redox-sensitive transcriptionfactors, particularly NF-κB, inducing the release of proin-flammatory cytokines and the expression of adhesion mole-cules and growth factors, including TNF-α, IL-6, IL-1β, TGF-β1, connective tissue growth factor, IGF-1, platelet-derivedgrowth factor, and VCAM-1 [115, 116]. H2O2 stimulates IL-4 and IL-6 gene expression and cytokine secretion by anapurinic/apyrimidinic-endonuclease/redox-factor-1- (APE/Ref-1-) dependent pathway [117]. ROS increased the expres-sion levels of PAI-1, IL-6, and MCP-1 through NADPHoxidase pathway [112]. Oxidized high-density lipoprotein(HDL) enhances proinflammatory properties such as TNF-α and MCP-1 in renal mesangial cells partly via CD36and LDL receptor-1 and via MAPK and NF-κB pathways[118]. Oxidized LDL can stimulate TNF-α synthesis fromrenal cells and initiate local effects of renal damage [85].Increased ROS production and MCP-1 secretion fromaccumulated fat may cause infiltration of macrophages andinflammation in adipose tissue of obesity [112]. Moreover,enhanced macrophage migration induces the release ofproinflammatory cytokines, which further stimulates thegeneration of ROS [119, 120]. Therefore, oxidative stress-induced cytokine production is likely to further increaseoxidative stress levels, setting a vicious cycle [121] that maypromote the progression of kidney damage in obesity.

3.3. Oxidative Stress Leads to Renal Injury in Obesity. Oxida-tive stress has been commonly identified in obesity-relatedrenal diseases and may be the mechanism underlying theinitiation or progression of renal injury in obesity [122, 123].Previous studies suggest that oxidative stress triggers, at anearly age, the onset of kidney lesions and functional im-pairment in Zucker obese (ZO) fa/fa rats, a good model ofobesity-related renal disease, in absence of hyperglycaemia,hypertension, and inflammation [124]. ROS play a crucialrole in mediating renal injury [125–127]. ROS are highlyreactive molecules that oxidize lipids and proteins, cause cell-ular injury, and promote glomerular and renal tubule injuryand associated proteinuria [128].

ROS are produced by various cells, such as vascularcells, inflammatory cells, and renal cells, and have distinctfunction on different types of cells, such as endothelial dys-function, inflammatory gene expression, and renal tubuleion transport. A major source for vascular and renal ROS is aNox family of nonphagocytic NAD(P)H oxidases, includingthe prototypic Nox2 homolog-based NAD(P)H oxidase, as

Page 40: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 5

well as other NAD(P)H oxidases, such as Nox1 and Nox4[129]. Numerous reports indicate that within the kidney,NAD(P)H oxidase, an enzyme that produces superoxide(O2

•−) by transferring electrons from NADH/NADPH tomolecular oxygen and thereby forming O2

−, H+, andNAD+/NADP+, is capable of modulating renal epithelial iontransport [130, 131]. NAP(D)H oxidase-derived ROS canalter renal pressure natriuresis and blood pressure regulationthrough its effects on renal hemodynamics and renal tubularsodium transport. Recent data suggest that NADPH oxidase-mediated oxidative injury to the proximal tubule, like thatseen in the glomerulus, contributes to proteinuria in insulin-resistant states [128].

Oxidative stress also plays an important role in the path-ogenesis of renal damage through its effects on vascular bio-logy. ROS are generated by all types of vascular cells, in-cluding endothelial, smooth muscle, and adventitial cells.ROS influence vascular cell growth, migration, proliferation,and activation [132, 133]. Physiologically, ROS can mediatecellular function, receptor signals, and immune responseson vascular cells. In pathophysiological condition, ROS con-tribute to progressive vascular dysfunction and remodel-ing through oxidative damage caused by decreased nitricoxide (NO) bioavailability, impaired endothelium-depend-ent vasodilatation and endothelial cell growth, apoptosis oranoikis, endothelial cell migration, and activation of adhe-sion molecules and inflammatory reactions [134, 135].

4. Potential Interventions toPrevent Renal Injury

4.1. Anti-Inflammation Therapy. Obesity accelerates the pro-gression of renal injury, associated with augmented inflam-mation in adipose and kidney tissues [136]. Anti-inflam-mation therapy might be a potential treatment for renaldamage. IL-6 is a key inflammatory molecule in renaldiseases. Studies in IL-6 transgenic mice suggested that highconcentrations of IL-6 contribute to development of renal in-jury [137]. Treatment with anti-IL-6 receptor antibodyMR16-1 prevented progression of proteinuria, renal lipiddeposit, and the mesangial cell proliferation in hypercho-lesterolemia-induced renal injury [97]. TNF-α is anotherimportant proinflammatory cytokine in the development ofrenal diseases. Inhibition of TNF-α by etanercept, a TNF-α antagonist, also decreased blood pressure and protectedthe kidney through reduction of renal NF-κB, oxidativestress, and inflammation [138]. TNF-α blockade increasesrenal Cyp2c23 expression and slows the progression of renaldamage in salt-sensitive hypertension [139]. TNF-α inhibi-tion also reduces renal injury in deoxycorticosterone-acetate-(DOCA-) salt hypertensive rats via suppression of renal cor-tical NF-κB activity [140]. Adiponectin is an adipose-secret-ed hormone with anti-inflammatory properties. Treatmentof adiponectin-knockout mice with adenovirus-mediatedadiponectin results in amelioration of albuminuria, glom-erular hypertrophy, and tubulointerstitial fibrosis and redu-ces the elevated levels of VCAM-1, MCP-1, TNF-α, TGF-β1,collagen type I/III, and NADPH oxidase components [141].Adiponectin prevents glomerular and tubulointerstitial

injury through modulating inflammation and oxidativestress [141].

Excessive fat accumulation contributes to macrophageinfiltration in adipose tissue and increased production ofproinflammatory cytokines, such as TNF-α, IL-8, and IL-6 [142–145]. Consequently, it is possible that weight lossmay be a potential method to reduce inflammation. Evidenceindicates that weight loss induced by nutritional interventionor gastric surgery markedly improves the systemic and adi-pose tissue inflammatory states linked to obesity [143, 146,147]. Studies by gene profiling analysis have shown thatcaloric restriction-induced weight reduction leads to the reg-ulation of a wide variety of inflammation-related moleculesin human adipose tissue [148]. Weight loss globally improvesthe inflammatory profile of obese subjects through a de-crease of proinflammatory factors and an increase of anti-inflammatory molecules in white adipose tissue [148]. Roux-en-Y-Gastric-Bypass- (RYGB-) induced weight loss has beenshown to reduce MCP-1, IL-18, IL-6, and TNF-α concentra-tions [149, 150]. A longer-term weight reduction induced byRYGB in corpulence also prevails in regulating circulatingcytokine concentrations [151]. Weight loss also amelioratesthe low-grade inflammation state that leads to glomerulardysfunction in obesity. In morbidly obese individuals withglomerular hyperfiltration, weight loss by surgical inter-ventions normalizes glomerular filtration rate (GFR) andreduces blood pressure and microalbuminuria [152]. Weightloss improves renal function as shown by reduced levels ofserum creatinine and improved creatinine clearance [153].

4.2. Antioxidant Intervention. Since ROS play a key role inthe pathogenesis of renal injury such as glomerulosclerosisand tubulointerstitial fibrosis, approaches to reduce oxidativestress by antioxidants supplementation, nutritional and sur-gical interventions may have renoprotective effects. Garciniaprotects against obesity-induced nephropathy by attenuatingoxidative stress through reduced lipid peroxidation and levelsof oxidized LDL [154]. The obese Zucker rat is a goodmodel for studying obesity-related kidney disease because itdevelops proteinuria, glomerular hypertrophy, and focal seg-mental glomerulosclerosis [155–157]. Using these rats, it hasbeen demonstrated that nephropathy is associated with oxi-dative stress, and supplementation with an antioxidantebselen improved kidney damage by ameliorating protein-uria and renal focal and segmental sclerosis [158]. Chronicebselen therapy also improved vasculopathy with lipiddeposits, tubulointerstitial scarring, and inflammation [158].Other antioxidants also have renoprotective effects. For ex-ample, administration of grape seed proanthocyanidin ex-tract (GSPE), an efficient phytochemical antioxidant, canprotect against the nephrotoxicity effects induced by cisplatinand gentamicin [159, 160] and reverse experimental myo-globinuric acute renal failure [161]. Quercetin, a flavonoidthat exhibits antioxidant properties in many diseases, couldalso protect the rat kidney against lead-induced injury andimprove renal function [162]. Thus, antioxidants may be apotential therapeutic to prevent the renal damage in ORG.

Nutritional and surgical interventions are additional ap-proaches to reduce oxidative stress and prevent kidney injury

Page 41: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

6 International Journal of Nephrology

in obesity. Caloric restriction and protein restriction reducefree radicals and ROS formation and inhibit accumulationof oxidative biomarkers in animal models [163]. In geneti-cally obese animals, diet restriction can prevent or greatlydelay the onset of specific degenerative lesions, in particularglomerulonephritis associated with obesity [164]. Sinceadipose tissue mass in obesity contributes to oxidative stress,bariatric surgery-induced weight loss also results in decreas-ing systemic oxidative stress in adiposity [111]. Weight lossinduced by diet restriction or bariatric surgery not only im-proves inflammation state but also reduces oxidative stressstate in obesity, which may protect renal function in obesity-related glomerulopathy.

5. Conclusion

Obesity causes chronic low-grade inflammation and sys-temic and local oxidative stress, which may play a pivotal rolein the initiation or progression of obesity-associated glomer-ulopathy. Elevated inflammation in obesity is the result ofthe production of adipokines and increased inflammatorycytokines and decreased anti-inflammatory factors. Oxida-tive stress is triggered by an imbalance between increasedproduction of ROS and/or reduced antioxidant activity. Bothinflammation and oxidative stress induce damage to renaltubule and glomerulus and result in endothelial dysfunctionin the kidney. Therefore, anti-inflammation and antioxidantinterventions may be the potential therapies to prevent andtreat obesity-related renal diseases.

References

[1] Y. Wang and M. A. Beydoun, “The obesity epidemic inthe United States—gender, age, socioeconomic, racial/ethnic,and geographic characteristics: a systematic review and meta-regression analysis,” Epidemiologic Reviews, vol. 29, no. 1, pp.6–28, 2007.

[2] C. Y. Hsu, C. E. McCulloch, C. Iribarren, J. Darbinian, and A.S. Go, “Body mass index and risk for end-stage renal disease,”Annals of Internal Medicine, vol. 144, no. 1, pp. 21–28, 2006.

[3] H. Kramer, A. Luke, A. Bidani, G. Cao, R. Cooper, and D.McGee, “Obesity and prevalent and incident CKD: the hy-pertension detection and follow-up program,” AmericanJournal of Kidney Diseases, vol. 46, no. 4, pp. 587–594, 2005.

[4] K. Reynolds, D. Gu, P. Muntner et al., “Body mass index andrisk of ESRD in China,” American Journal of Kidney Diseases,vol. 50, no. 5, pp. 754–764, 2007.

[5] J. R. Weisinger, R. L. Kempson, F. L. Eldridge, and R. S.Swenson, “The nephrotic syndrome: a complication of mas-sive obesity,” Annals of Internal Medicine, vol. 81, no. 4, pp.440–447, 1974.

[6] J. C. Jennette, L. Charles, and W. Grubb, “Glomerulomegalyand focal segmental glomerulosclerosis associated with obe-sity and sleep-apnea syndrome,” American Journal of KidneyDiseases, vol. 10, no. 6, pp. 470–472, 1987.

[7] R. R. Verani, “Obesity-associated focal segmental glomeru-losclerosis: pathological features of the lesion and relation-ship with cardiomegaly and hyperlipidemia,” American Jour-nal of Kidney Diseases, vol. 20, no. 6, pp. 629–634, 1992.

[8] N. Kambham, G. S. Markowitz, A. M. Valeri, J. Lin, and V.D. D’Agati, “Obesity-related glomerulopathy: an emerging

epidemic,” Kidney International, vol. 59, no. 4, pp. 1498–1509, 2001.

[9] M. Praga, E. Hernandez, E. Morales et al., “Clinical featuresand long-term outcome of obesity-associated focal segmentalglomerulosclerosis,” Nephrology Dialysis Transplantation, vol.16, no. 9, pp. 1790–1798, 2001.

[10] R. D. Adelman, I. G. Restaino, U. S. Alon, and D. L.Blowey, “Proteinuria and focal segmental glomerulosclerosisin severely obese adolescents,” Journal of Pediatrics, vol. 138,no. 4, pp. 481–485, 2001.

[11] H. M. Chen, S. J. Li, H. P. Chen, Q. W. Wang, L. S. Li, andZ. H. Liu, “Obesity-related glomerulopathy in China: a caseseries of 90 patients,” American Journal of Kidney Diseases,vol. 52, no. 1, pp. 58–65, 2008.

[12] H. M. Chen, Z. H. Liu, C. H. Zeng, S. J. Li, Q. W. Wang, andL. S. Li, “Podocyte Lesions in Patients With Obesity-RelatedGlomerulopathy,” American Journal of Kidney Diseases, vol.48, no. 5, pp. 772–779, 2006.

[13] A. Chagnac, T. Weinstein, A. Korzets, E. Ramadan, J. Hirsch,and U. Gafter, “Glomerular hemodynamics in severe obesity,”American Journal of Physiology, vol. 278, no. 5, pp. F817–F822, 2000.

[14] J. R. Henegar, S. A. Bigler, L. K. Henegar, S. C. Tyagi, and J.E. Hall, “Functional and structural changes in the kidney inthe early stages of obesity,” Journal of the American Society ofNephrology, vol. 12, no. 6, pp. 1211–1217, 2001.

[15] I. M. Wahba and R. H. Mak, “Obesity and obesity-initiatedmetabolic syndrome: mechanistic links to chronic kidneydisease,” Clinical Journal of the American Society of Nephrol-ogy, vol. 2, no. 3, pp. 550–562, 2007.

[16] L. A. Juncos and S. Ito, “Disparate effects of insulin onisolated rabbit afferent and efferent arterioles,” Journal ofClinical Investigation, vol. 92, no. 4, pp. 1981–1985, 1993.

[17] J. Frystyk, C. Skjærbæk, E. Vestbo, S. Fisker, and H. Ørskov,“Circulating levels of free insulin-like growth factors in obesesubjects: the impact of type 2 diabetes,” Diabetes/MetabolismResearch and Reviews, vol. 15, no. 5, pp. 314–322, 1999.

[18] K. Aihara, Y. Ikeda, S. Yagi et al., “Transforming growthfactor-beta1 as a common target molecule for developmentof cardiovascular diseases, renal insufficiency and metabolicsyndrome,” Cardiology Reseach Practice, vol. 2011, Article ID175381, 2011.

[19] J. A. Joles, U. Kunter, U. Janssen et al., “Early mechanisms ofrenal injury in hypercholesterolemic or hypertriglyceridemicrats,” Journal of the American Society of Nephrology, vol. 11,no. 4, pp. 669–683, 2000.

[20] T. Jiang, Z. Wang, G. Proctor et al., “Diet-induced obesity inC57BL/6J mice causes increased renal lipid accumulation andglomerulosclerosis via a sterol regulatory element-bindingprotein-1c-dependent pathway,” The Journal of BiologicalChemistry, vol. 280, no. 37, pp. 32317–32325, 2005.

[21] J. E. Hall, “The kidney, hypertension, and obesity,” Hyperten-sion, vol. 41, no. 3, pp. 625–633, 2003.

[22] S. P. Bagby, “Obesity-initiated metabolic syndrome and thekidney: a recipe for chronic kidney disease?” Journal of theAmerican Society of Nephrology, vol. 15, no. 11, pp. 2775–2791, 2004.

[23] F. A. El-Atat, S. N. Stas, S. I. Mcfarlane, and J. R. Sowers,“The relationship between hyperinsulinemia, hypertensionand progressive renal disease,” Journal of the American Societyof Nephrology, vol. 15, no. 11, pp. 2816–2827, 2004.

[24] J. Kurokawa, H. Nagano, O. Ohara et al., “Apoptosis inhibitorof macrophage (AIM) is required for obesity-associatedrecruitment of inflammatory macrophages into adipose

Page 42: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 7

tissue,” Proceedings of the National Academy of Sciences of theUnited States of America, vol. 108, no. 29, pp. 12072–12077,2011.

[25] N. Ouchi, J. L. Parker, J. J. Lugus et al., “Adipokinesin inflammation and metabolic disease,” Nature ReviewsImmunology, vol. 11, no. 2, pp. 85–97, 2011.

[26] F. Y. Chow, D. J. Nikolic-Paterson, F. Y. Ma, E. Ozols, B. J.Rollins, and G. H. Tesch, “Monocyte chemoattractant pro-tein-1-induced tissue inflammation is critical for the devel-opment of renal injury but not type 2 diabetes in obese db/dbmice,” Diabetologia, vol. 50, no. 2, pp. 471–480, 2007.

[27] G. Egger and J. Dixon, “Non-nutrient causes of low-grade,systemic inflammation: support for a “canary in the mine-shaft” view of obesity in chronic disease,” Obesity Reviews,vol. 12, no. 5, pp. 339–345, 2011.

[28] M. Visser, L. M. Bouter, G. M. McQuillan, M. H. Wener, andT. B. Harris, “Elevated C-reactive protein levels in overweightand obese adults,” Journal of the American Medical Associa-tion, vol. 282, no. 22, pp. 2131–2135, 1999.

[29] M. G. Farb, S. Bigornia, M. Mott et al., “Reduced adiposetissue inflammation represents an intermediate cardiometa-bolic phenotype in obesity,” Jounal of the American College ofCardiology, vol. 58, no. 3, pp. 232–237, 2011.

[30] S. Khanna and A. M. Mali, “Evaluation of tumor necrosisfactor-alpha (TNF-alpha) levels in plasma and their cor-relation with periodontal status in obese and non-obesesubjects,” Journal of Indian Society Periodontology, vol. 14, no.4, pp. 217–221, 2010.

[31] A. Virdis, F. Santini, R. Colucci et al., “Vascular generation oftumor necrosis factor-alpha reduces nitric oxide availabilityin small arteries from visceral fat of obese patients,” Jounalof the American College of Cardiology, vol. 58, no. 3, pp. 238–247, 2011.

[32] S. La Vignera, R. Condorelli, S. Bellanca et al., “Obesity isassociated with a higher level of pro-inflammatory cytokinesin follicular fluid of women undergoing medically assistedprocreation (PMA) programs,” European Review of MedicalPharmacology Sciences, vol. 15, no. 3, pp. 267–273, 2011.

[33] T. S. Church, M. S. Willis, E. L. Priest et al., “Obesity, ma-crophage migration inhibitory factor, and weight loss,” In-ternational Journal of Obesity, vol. 29, no. 6, pp. 675–681,2005.

[34] C. Buechler, J. Wanninger, and M. Neumeier, “Adiponectin, akey adipokine in obesity related liver diseases,” World Journalof Gastroenterology, vol. 17, no. 23, pp. 2801–2811, 2011.

[35] S. Kshatriya, K. Liu, A. Salah et al., “Obesity hypertension:the regulatory role of leptin,” International Journal of Hyper-tension, vol. 2011, Article ID 270624, 8 pages, 2011.

[36] G. Wolf, S. Chen, D. C. Han, and F. N. Ziyadeh, “Leptin andrenal disease,” American Journal of Kidney Diseases, vol. 39,no. 1, pp. 1–11, 2002.

[37] K. Sharma, “The link between obesity and albuminuria: adi-ponectin and podocyte dysfunction,” Kidney International,vol. 76, no. 2, pp. 145–148, 2009.

[38] S. S. Martin, A. Qasim, and M. P. Reilly, “Leptin resistance:a possible interface of inflammation and metabolism inobesity-related cardiovascular disease,” Journal of the Amer-ican College of Cardiology, vol. 52, no. 15, pp. 1201–1210,2008.

[39] G. Wolf, A. Hamann, D. C. Han et al., “Leptin stimulatesproliferation and TGF-β expression in renal glomerular end-othelial cells: potential role in glomerulosclerosis,” Kidney In-ternational, vol. 56, no. 3, pp. 860–872, 1999.

[40] D. C. Han, M. Isono, S. Chen et al., “Leptin stimulates type Icollagen production in db/db mesangial cells: glucose uptake

and TGF-β type II receptor expression,” Kidney International,vol. 59, no. 4, pp. 1315–1323, 2001.

[41] G. Wolf and F. N. Ziyadeh, “Leptin and renal fibrosis,” Con-tributions to Nephrology, vol. 151, pp. 175–183, 2006.

[42] Q. L. Lam and L. Lu, “Role of leptin in immunity,” Cellular &Molecular Immunology, vol. 4, no. 1, pp. 1–13, 2007.

[43] I. S. Farooqi, G. Matarese, G. M. Lord et al., “Beneficialeffects of leptin on obesity, T cell hyporesponsiveness, andneuroendocrine/metabolic dysfunction of human congenitalleptin deficiency,” Journal of Clinical Investigation, vol. 110,no. 8, pp. 1093–1103, 2002.

[44] M. Tanaka, T. Suganami, S. Sugita et al., “Role of centralleptin signaling in renal macrophage infiltration,” EndocrineJournal, vol. 57, no. 1, pp. 61–72, 2010.

[45] S. Loffreda, S. Q. Yang, and H. Z. Lin, “Leptin regulates pro-inflammatory immune responses,” The Federation of Ameri-can Societies for Experimental Biology Journal, vol. 12, no. 1,pp. 57–65, 1998.

[46] K. Chen, F. Li, J. Li et al., “Induction of leptin resistancethrough direct interaction of C-reactive protein with leptin,”Nature Medicine, vol. 12, no. 4, pp. 425–432, 2006.

[47] F. Vasseur, F. Lepretre, C. Lacquemant, and P. Froguel, “Thegenetics of adiponectin,” Current Diabetes Reports, vol. 3, no.2, pp. 151–158, 2003.

[48] T. Yamauchi, J. Kamon, Y. Ito et al., “Cloning of adiponectinreceptors that mediate antidiabetic metabolic effects,” Na-ture, vol. 423, no. 6941, pp. 762–769, 2003.

[49] S. Engeli, M. Feldpausch, K. Gorzelniak et al., “Associationbetween adiponectin and mediators of inflammation inobese women,” Diabetes, vol. 52, no. 4, pp. 942–947, 2003.

[50] E. Hu, P. Liang, and B. M. Spiegelman, “AdipoQ is a noveladipose-specific gene dysregulated in obesity,” The Journal ofBiological Chemistry, vol. 271, no. 18, pp. 10697–10703, 1996.

[51] Y. Arita, S. Kihara, N. Ouchi et al., “Paradoxical decreaseof an adipose-specific protein, adiponectin, in obesity,”Biochemical and Biophysical Research Communications, vol.257, no. 1, pp. 79–83, 1999.

[52] Y. Yano, S. Hoshide, J. Ishikawa et al., “Differential impactsof adiponectin on low-grade albuminuria between obese andnonobese persons without diabetes,” Journal of Clinical Hy-pertension (Greenwich), vol. 9, no. 10, pp. 775–782, 2007.

[53] K. Sharma, S. RamachandraRao, G. Qiu et al., “Adiponectinregulates albuminuria and podocyte function in mice,”Journal of Clinical Investigation, vol. 118, no. 5, pp. 1645–1656, 2008.

[54] S. Nakamaki, H. Satoh, A. Kudoh et al., “Adiponectin reducesproteinuria in streptozotocin-induced diabetic Wistar rats,”Experimental Biology and Medicine (Maywood), vol. 236, no.5, pp. 614–620, 2011.

[55] N. Ouchi, S. Kihara, Y. Arita et al., “Novel modulator forendothelial adhesion molecules: adipocyte-derived plasmaprotein adiponectin,” Circulation, vol. 100, no. 25, pp. 2473–2476, 1999.

[56] N. Ouchi, S. Kihara, Y. Arita et al., “Adiponectin, an adi-pocyte-derived plasma protein, inhibits endothelial NF-κBsignaling through a cAMP-dependent pathway,” Circulation,vol. 102, no. 11, pp. 1296–1301, 2000.

[57] R. S. Rosenson, “Effect of fenofibrate on adiponectin and in-flammatory biomarkers in metabolic syndrome patients,”Obesity, vol. 17, no. 3, pp. 504–509, 2009.

[58] T. Yokota, K. Oritani, I. Takahashi et al., “Adiponectin, a newmember of the family of soluble defense collagens, negativelyregulates the growth of myelomonocytic progenitors and thefunctions of macrophages,” Blood, vol. 96, no. 5, pp. 1723–1732, 2000.

Page 43: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

8 International Journal of Nephrology

[59] N. Ouchi, S. Kihara, T. Funahashi et al., “Reciprocal associa-tion of C-reactive protein with adiponectin in blood streamand adipose tissue,” Circulation, vol. 107, no. 5, pp. 671–674,2003.

[60] K. H. Kim, K. Lee, Y. S. Moon, and H. S. Sul, “A cysteine-rich adipose tissue-specific secretory factor inhibits adipocytedifferentiation,” The Journal of Biological Chemistry, vol. 276,no. 14, pp. 11252–11256, 2001.

[61] P. G. McTernan, C. M. Kusminski, and S. Kumar, “Resistin,”Current Opinion in Lipidology, vol. 17, no. 2, pp. 170–175,2006.

[62] C. M. Steppan, S. T. Bailey, S. Bhat et al., “The hormoneresistin links obesity to diabetes,” Nature, vol. 409, no. 6818,pp. 307–312, 2001.

[63] M. W. Rajala, Y. Lin, M. Ranalletta et al., “Cell type-specificexpression and coregulation of murine resistin and resistin-like molecule-α in adipose tissue,” Molecular Endocrinology,vol. 16, no. 8, pp. 1920–1930, 2002.

[64] L. Patel, A. C. Buckels, I. J. Kinghorn et al., “Resistin is ex-pressed in human macrophages and directly regulated byPPARγ activators,” Biochemical and Biophysical ResearchCommunications, vol. 300, no. 2, pp. 472–476, 2003.

[65] H. S. Jung, K. H. Park, Y. M. Cho et al., “Resistin is sec-reted from macrophages in atheromas and promotes athe-rosclerosis,” Cardiovascular Research, vol. 69, no. 1, pp. 76–85, 2006.

[66] J. N. Fain, P. S. Cheema, S. W. Bahouth, and M. L. Hiler,“Resistin release by human adipose tissue explants in pri-mary culture,” Biochemical and Biophysical Research Commu-nications, vol. 300, no. 3, pp. 674–678, 2003.

[67] M. W. Rajala, Y. Qi, H. R. Patel et al., “Regulation of resistinexpression and circulating levels in obesity, diabetes, andfasting,” Diabetes, vol. 53, no. 7, pp. 1671–1679, 2004.

[68] M. A. Lazar, “Resistin- and obesity-associated metabolic dis-eases,” Hormone and Metabolic Research, vol. 39, no. 10, pp.710–716, 2007.

[69] M. Owecki, A. Miczke, and E. Nikisch, “Serum resistin con-centrations are higher in human obesity but independentfrom insulin resistance,” Experimental and Clinical Endocri-nology and Diabetes, vol. 119, no. 2, pp. 117–121, 2011.

[70] O. A. Mojiminiyi and N. A. Abdella, “Associations of resistinwith inflammation and insulin resistance in patients withtype 2 diabetes mellitus,” Scandinavian Journal of Clinical andLaboratory Investigation, vol. 67, no. 2, pp. 215–225, 2007.

[71] M. Bokarewa, I. Nagaev, L. Dahlberg, U. Smith, and A.Tarkowski, “Resistin, an adipokine with potent proinflam-matory properties,” Journal of Immunology, vol. 174, no. 9,pp. 5789–5795, 2005.

[72] N. Silswal, A. K. Singh, B. Aruna, S. Mukhopadhyay, S.Ghosh, and N. Z. Ehtesham, “Human resistin stimulatesthe pro-inflammatory cytokines TNF-α and IL-12 in macro-phages by NF-κB-dependent pathway,” Biochemical and Bi-ophysical Research Communications, vol. 334, no. 4, pp. 1092–1101, 2005.

[73] J. Axelsson, A. Bergsten, A. R. Qureshi et al., “Elevatedresistin levels in chronic kidney disease are associated withdecreased glomerular filtration rate and inflammation, butnot with insulin resistance,” Kidney International, vol. 69, no.3, pp. 596–604, 2006.

[74] R. Kawamura, Y. Doi, H. Osawa et al., “Circulating resistinis increased with decreasing renal function in a generalJapanese population: the Hisayama Study,” Nephrology Dial-ysis Transplantation, vol. 25, no. 10, pp. 3236–3240, 2010.

[75] Z. G. Xu, L. Lanting, N. D. Vaziri et al., “Upregulation ofangiotensin II type 1 receptor, inflammatory mediators, andenzymes of arachidonate metabolism in obese zucker ratkidney: reversal by angiotensin II type 1 receptor blockade,”Circulation, vol. 111, no. 15, pp. 1962–1969, 2005.

[76] J. DeFuria, G. Bennett, K. J. Strissel et al., “Dietary blueberryattenuates whole-body insulin resistance in high fat-fed miceby reducing adipocyte death and its inflammatory sequelae,”Journal of Nutrition, vol. 139, no. 8, pp. 1510–1516, 2009.

[77] S. P. Weisberg, D. McCann, M. Desai, M. Rosenbaum, R. L.Leibel, and A. W. Ferrante, “Obesity is associated with ma-crophage accumulation in adipose tissue,” Journal of ClinicalInvestigation, vol. 112, no. 12, pp. 1796–1808, 2003.

[78] J. N. Fain, S. W. Bahouth, and A. K. Madan, “TNFalpharelease by the nonfat cells of human adipose tissue,” Interna-tional Jounal of Obessity Related Metabolic Disorders, vol. 28,no. 4, pp. 616–622, 2004.

[79] J. Egido, M. Gomez-Chiarri, A. Ortiz et al., “Role of tumornecrosis factor-α in the pathogenesis of glomerular diseases,”Kidney International, Supplement, no. 39, pp. S-59–S-64,1993.

[80] J. K. Sethi and G. S. Hotamisligil, “The role of TNFα inadipocyte metabolism,” Seminars in Cell and DevelopmentalBiology, vol. 10, no. 1, pp. 19–29, 1999.

[81] C. Tsigos, I. Kyrou, E. Chala et al., “Circulating tumornecrosis factor alpha concentrations are higher in abdominalversus peripheral obesity,” Metabolism, vol. 48, no. 10, pp.1332–1335, 1999.

[82] D. E. Moller, “Potential role of TNF-α in the pathogenesis ofinsulin resistance and type 2 diabetes,” Trends in Endocrinol-ogy and Metabolism, vol. 11, no. 6, pp. 212–217, 2000.

[83] G. Rashid, S. Benchetrit, D. Fishman, and J. Bernheim,“Effect of advanced glycation end-products on gene expres-sion and synthesis of TNF-α and endothelial nitric oxidesynthase by endothelial cells,” Kidney International, vol. 66,no. 3, pp. 1099–1106, 2004.

[84] M. Ruiz-Ortega, M. Ruperez, O. Lorenzo et al., “AngiotensinII regulates the synthesis of proinflammatory cytokines andchemokines in the kidney,” Kidney International, Supplement,vol. 62, no. 82, pp. S12–S22, 2002.

[85] S. Jovinge, M. P. S. Ares, B. Kallin, and J. Nilsson, “Humanmonocytes/macrophages release TNF-α in response to Ox-LDL,” Arteriosclerosis, Thrombosis, and Vascular Biology, vol.16, no. 12, pp. 1573–1579, 1996.

[86] M. H. S. Al-Dahr and E. H. Jiffri, “Increased adipose tissueexpression of tumor necrosis factor-alpha and insulin resis-tance in obese subjects with type II diabetes,” World Journalof Medical Sciences, vol. 5, no. 2, pp. 30–35, 2010.

[87] J. A. Moreno, M. C. Izquierdo, M. D. Sanchez-Nino et al.,“The inflammatory cytokines TWEAK and TNFα reducerenal klotho expression through NFκB.,” Journal of theAmerican Society Nephrology, vol. 22, no. 7, pp. 1315–1325,2011.

[88] Y. L. Cao, Y. X. Wang, D. F. Wang, X. Meng, and J. Zhang,“Correlation between omental TNF-α protein and plasmaPAI-1 in obesity subjects,” International Journal of Cardiology,vol. 128, no. 3, pp. 399–405, 2008.

[89] M. Cigolini, M. Tonoli, L. Borgato et al., “Expression of plas-minogen activator inhibitor-1 in human adipose tissue: a rolefor TNF-α?” Atherosclerosis, vol. 143, no. 1, pp. 81–90, 1999.

[90] J. P. Rerolle, A. Hertig, G. Nguyen, J. D. Sraer, and E.P. Rondeau, “Plasminogen activator inhibitor type 1 is apotential target in renal fibrogenesis,” Kidney International,vol. 58, no. 5, pp. 1841–1850, 2000.

Page 44: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 9

[91] K. Matoba, D. Kawanami, S. Ishizawa, Y. Kanazawa, T.Yokota, and K. Utsunomiya, “Rho-kinase mediates TNF-α-induced MCP-1 expression via p38 MAPK signaling pathwayin mesangial cells,” Biochemical and Biophysical ResearchCommunications, vol. 402, no. 4, pp. 725–730, 2010.

[92] L. Roytblat, M. Rachinsky, A. Fisher et al., “Raised interleu-kin-6 levels in obese patients,” Obesity Research, vol. 8, no. 9,pp. 673–675, 2000.

[93] H. S. Park, J. Y. Park, and R. Yu, “Relationship of obesity andvisceral adiposity with serum concentrations of CRP, TNF-αand IL-6,” Diabetes Research and Clinical Practice, vol. 69, no.1, pp. 29–35, 2005.

[94] M. Maachi, L. Pieroni, and E. Bruckert, “Systemic low-gradeinflammation is related to both circulating and adipose tissueTNFalpha, leptin and IL-6 levels in obese women,” Interna-tional Journal of Obesity Related Metabolic Disorders, vol. 28,no. 8, pp. 993–997, 2004.

[95] N. S. A. Patel, P. K. Chatterjee, R. Di Paola et al., “Endogenousinterleukin-6 enhances the renal injury, dysfunction, andinflammation caused by ischemia/reperfusion,” Journal ofPharmacology and Experimental Therapeutics, vol. 312, no. 3,pp. 1170–1178, 2005.

[96] M. L. Kielar, R. John, M. Bennett et al., “Maladaptive role ofIL-6 in ischemic acute renal failure,” Journal of the AmericanSociety of Nephrology, vol. 16, no. 11, pp. 3315–3325, 2005.

[97] M. Tomiyama-Hanayama, H. Rakugi, M. Kohara et al.,“Effect of interleukin-6 receptor blockage on renal injuryin apolipoprotein E-deficient mice,” American Journal ofPhysiology, vol. 297, no. 3, pp. F679–F684, 2009.

[98] C. Zoccali, F. Mallamaci, and G. Tripepi, “Adipose tissue asa source of inflammatory cytokines in health and disease:focus on end-stage renal disease,” Kidney International, Sup-plement, vol. 63, no. 84, pp. S65–S68, 2003.

[99] A. P. Burke, R. P. Tracy, F. Kolodgie et al., “Elevated C-reactiveprotein values and atherosclerosis in sudden coronary death:association with different pathologies,” Circulation, vol. 105,no. 17, pp. 2019–2023, 2002.

[100] H. Lue, R. Kleemann, T. Calandra, T. Roger, and J. Bern-hagen, “Macrophage migration inhibitory factor (MIF):mechanisms of action and role in disease,” Microbes andInfection, vol. 4, no. 4, pp. 449–460, 2002.

[101] J. Bernhagen, T. Calandra, R. A. Mitchell et al., “MIF is apituitary-derived cytokine that potentiates lethal endotox-aemia,” Nature, vol. 365, no. 6448, pp. 756–759, 1993.

[102] H. Y. Lan, M. Bacher, N. Yang et al., “The pathogenic role ofmacrophage migration inhibitory factor in immunologicallyinduced kidney disease in the rat,” Journal of ExperimentalMedicine, vol. 185, no. 8, pp. 1455–1465, 1997.

[103] J. F. Keaney Jr., M. G. Larson, R. S. Vasan et al., “Obesity andsystemic oxidative stress: clinical correlates of oxidative stressin the Framingham study,” Arteriosclerosis, Thrombosis, andVascular Biology, vol. 23, no. 3, pp. 434–439, 2003.

[104] H. K. Vincent and A. G. Taylor, “Biomarkers and potentialmechanisms of obesity-induced oxidant stress in humans,”International Journal of Obesity, vol. 30, no. 3, pp. 400–418,2006.

[105] E. Pihl, K. Zilmer, T. Kullisaar, C. Kairane, A. Magi, andM. Zilmer, “Atherogenic inflammatory and oxidative stressmarkers in relation to overweight values in male formerathletes,” International Journal of Obesity, vol. 30, no. 1, pp.141–146, 2006.

[106] S. J. Piva, M. M. M. F. Duarte, I. B. M. Da Cruz et al.,“Ischemia-modified albumin as an oxidative stress biomarkerin obesity,” Clinical Biochemistry, vol. 44, no. 4, pp. 345–347,2011.

[107] C. Chrysohoou, D. B. Panagiotakos, C. Pitsavos et al., “Theimplication of obesity on total antioxidant capacity in ap-parently healthy men and women: the ATTICA study,” Nu-trition, Metabolism and Cardiovascular Diseases, vol. 17, no.8, pp. 590–597, 2007.

[108] J. Hartwich, J. Goralska, D. Siedlecka, A. Gruca, M. Trzos, andA. Dembinska-Kiec, “Effect of supplementation with vitaminE and C on plasma hsCRP level and cobalt-albumin bindingscore as markers of plasma oxidative stress in obesity,” Genesand Nutrition, vol. 2, no. 1, pp. 151–154, 2007.

[109] L. A. Brown, C. J. Kerr, and P. Whiting, “Oxidant stress inhealthy normal-weight, overweight, and obese individuals,”Obesity (Silver Spring), vol. 17, no. 3, pp. 460–466, 2009.

[110] H. K. Vincent, S. K. Powers, A. J. Dirkset et al., “Mechanismfor obesity-induced increase in myocardial lipid peroxi-dation,” International Journal of Obesity Related MetabolicDisorders, vol. 25, no. 3, pp. 378–388, 2001.

[111] N. Gletsu-Miller, J. M. Hansen, D. P. Jones et al., “Loss of totaland visceral adipose tissue mass predicts decreases in oxi-dative stress after weight-loss surgery,” Obesity (Silver Spring),vol. 17, no. 3, pp. 439–446, 2009.

[112] S. Furukawa, T. Fujita, M. Shimabukuro et al., “Increasedoxidative stress in obesity and its impact on metabolic syn-drome,” Journal of Clinical Investigation, vol. 114, no. 12, pp.1752–1761, 2004.

[113] J. P. Bastard, M. Maachi, C. Lagathu et al., “Recent advancesin the relationship between obesity, inflammation, and in-sulin resistance,” European Cytokine Network, vol. 17, no. 1,pp. 4–12, 2006.

[114] S. de Ferranti and D. Mozaffarian, “The perfect storm: obe-sity, adipocyte dysfunction, and metabolic consequences,”Clinical Chemistry, vol. 54, no. 6, pp. 945–955, 2008.

[115] Y. Lavrovsky, B. Chatterjee, R. A. Clark, and A. K. Roy, “Roleof redox-regulated transcription factors in inflammation,aging and age-related diseases,” Experimental Gerontology,vol. 35, no. 5, pp. 521–532, 2000.

[116] A. L. Y. Tan, J. M. Forbes, and M. E. Cooper, “AGE, RAGE,and ROS in diabetic nephropathy,” Seminars in Nephrology,vol. 27, no. 2, pp. 130–143, 2007.

[117] B. Frossi, M. De Carli, K. C. Daniel, J. Rivera, and C. Pucillo,“Oxidative stress stimulates IL-4 and IL-6 production in mastcells by an APE/Ref-1-dependent pathway,” European Journalof Immunology, vol. 33, no. 8, pp. 2168–2177, 2003.

[118] M. Zhang, X. Gao, J. Wu et al., “Oxidized high-density lipo-protein enhances inflammatory activity in rat mesangialcells,” Diabetes/Metabolism Research and Reviews, vol. 26, no.6, pp. 455–463, 2010.

[119] F. N. Ziyadeh and G. Wolf, “Pathogenesis of the podocy-topathy and proteinuria in diabetic glomerulopathy,” CurrentDiabetes Reviews, vol. 4, no. 1, pp. 39–45, 2008.

[120] H. E. Hohmeier, V. V. Tran, G. Chen, R. Gasa, and C. B.Newgard, “Inflammatory mechanisms in diabetes: Lessonsfrom the β-cell,” International Journal of Obesity, vol. 27,supplement 3, pp. S12–S16, 2003.

[121] A. A. Elmarakby and J. C. Sullivan, “Relationship betweenoxidative stress and inflammatory cytokines in diabetic ne-phropathy,” Cardiovascular Therapeutics, vol. 30, no. 1, pp.49–59, 2012.

[122] S. Darouich, R. Goucha, M. H. Jaafoura, S. Zekri, H. B.Maiz, and A. Kheder, “Clinicopathological characteristicsof obesity-associated focal segmental glomerulosclerosis,”Ultrastructural Pathology, vol. 35, no. 4, pp. 176–182, 2011.

[123] J. E. Quigley, A. A. Elmarakby, S. F. Knight et al., “Obesityinduced renal oxidative stress contributes to renal injury in

Page 45: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

10 International Journal of Nephrology

salt-sensitive hypertension,” Clinical and Experimental Phar-macology and Physiology, vol. 36, no. 7, pp. 724–728, 2009.

[124] B. Poirier, M. Lannaud-Bournoville, M. Conti et al., “Oxida-tive stress occurs in absence of hyperglycaemia and inflam-mation in the onset of kidney lesions in normotensive obeserats,” Nephrology Dialysis Transplantation, vol. 15, no. 4, pp.467–476, 2000.

[125] E. A. Jaimes, P. Hua, R. X. Tian, and L. Raij, “Human glom-erular endothelium: interplay among glucose, free fatty acids,angiotensin II, and oxidative stress,” American Journal ofPhysiology, vol. 298, no. 1, pp. F125–F132, 2010.

[126] H. Fujii, K. Kono, K. Nakai et al., “Oxidative and nitrosativestress and progression of diabetic nephropathy in type 2diabetes,” American Journal of Nephrology, vol. 31, no. 4, pp.342–352, 2010.

[127] V. B. O’Donnell and B. A. Freeman, “Interactions betweennitric oxide and lipid oxidation pathways: implications forvascular disease,” Circulation Research, vol. 88, no. 1, pp. 12–21, 2001.

[128] J. Habibi, M. R. Hayden, J. R. Sowers et al., “Nebivololattenuates redox-sensitive glomerular and tubular mediatedproteinuria in obese rats,” Endocrinology, vol. 152, no. 2, pp.659–668, 2011.

[129] T. M. Paravicini and R. M. Touyz, “NADPH oxidases, reactiveoxygen species, and hypertension: clinical implications andtherapeutic possibilities,” Diabetes care, vol. 31, supplement2, pp. S170–S180, 2008.

[130] C. Schreck and P. M. O’Connor, “NAD(P)H oxidase andrenal epithelial ion transport,” American Journal of Physiol-ogy, vol. 300, no. 5, pp. R1023–R1029, 2011.

[131] S. H. M. Bengtsson, L. M. Gulluyan, G. J. Dusting, and G. R.Drummond, “Novel isoforms of NADPH oxidase in vascularphysiology and pathophysiology,” Clinical and ExperimentalPharmacology and Physiology, vol. 30, no. 11, pp. 849–854,2003.

[132] R. M. Touyz, F. Tabet, and E. L. Schiffrin, “Redox-dependentsignalling by angiotensin II and vascular remodelling in hy-pertension,” Clinical and Experimental Pharmacology andPhysiology, vol. 30, no. 11, pp. 860–866, 2003.

[133] H. Cai, “Hydrogen peroxide regulation of endothelial func-tion: origins, mechanisms, and consequences,” Cardiovascu-lar Research, vol. 68, no. 1, pp. 26–36, 2005.

[134] L. M. Yung, F. P. Leung, X. Yao, Z. Y. Chen, and Y. Huang,“Reactive oxygen species in vascular wall,” Cardiovascularand Hematological Disorders, vol. 6, no. 1, pp. 1–19, 2006.

[135] H. Cai and D. G. Harrison, “Endothelial dysfunction incardiovascular diseases: the role of oxidant stress,” CirculationResearch, vol. 87, no. 10, pp. 840–844, 2000.

[136] L. J. Ma, B. A. Corsa, J. Zhou et al., “Angiotensin type 1receptor modulates macrophage polarization and renal in-jury in obesity,” American Journal of Physiology, vol. 300, no.5, pp. 1203–1213, 2011.

[137] Y. Shima, M. Iwano, K. Yoshizaki, T. Tanaka, I. Kawase, andN. Nishimoto, “All-trans-retinoic acid inhibits the develop-ment of mesangial proliferative glomerulonephritis in inter-leukin-6 transgenic mice,” Nephron, vol. 100, no. 1, pp. e54–62, 2005.

[138] M. Venegas-Pont, M. B. Manigrasso, S. C. Grifoni et al.,“Tumor necrosis factor-α antagonist etanercept decreasesblood pressure and protects the kidney in a mouse model ofsystemic lupus erythematosus,” Hypertension, vol. 56, no. 4,pp. 643–649, 2010.

[139] A. A. Elmarakby, J. E. Quigley, D. M. Pollock, and J. D. Imig,“Tumor necrosis factor α blockade increases renal Cyp2c23

expression and slows the progression of renal damage in salt-sensitive hypertension,” Hypertension, vol. 47, no. 3, pp. 557–562, 2006.

[140] A. A. Elmarakby, J. E. Quigley, J. D. Imig, J. S. Pollock, andD. M. Pollock, “TNF-α inhibition reduces renal injury inDOCA-salt hypertensive rats,” American Journal of Physiol-ogy, vol. 294, no. 1, pp. R76–R83, 2008.

[141] K. Ohashi, H. Iwatani, S. Kihara et al., “Exacerbation of albu-minuria and renal fibrosis in subtotal renal ablation modelof adiponectin-knockout mice,” Arteriosclerosis, Thrombosis,and Vascular Biology, vol. 27, no. 9, pp. 1910–1917, 2007.

[142] J. P. Bastard, C. Jardel, J. Delattre, B. Hainque, E. Bruckert,and F. Oberlin, “Evidence for a link between adiposetissue interleukin-6 content and serum C-reactive proteinconcentrations in obese subjects,” Circulation, vol. 99, no. 16,pp. 2221–2222, 1999.

[143] J. P. Bastard, C. Jardel, E. Bruckert et al., “Elevated levels ofinterleukin 6 are reduced in serum and subcutaneous adiposetissue of obese women after weight loss,” Journal of ClinicalEndocrinology and Metabolism, vol. 85, no. 9, pp. 3338–3342,2000.

[144] V. Mohamed-Ali, L. Flower, J. Sethi et al., “β-adrenergic regu-lation of IL-6 release from adipose tissue: in vivo and in vitrostudies,” Journal of Clinical Endocrinology and Metabolism,vol. 86, no. 12, pp. 5864–5869, 2001.

[145] H. Bruunsgaard and B. K. Pedersen, “Age-related inflamma-tory cytokines and disease,” Immunology and Allergy Clinicsof North America, vol. 23, no. 1, pp. 15–39, 2003.

[146] R. Cancello, C. Henegar, N. Viguerie et al., “Reduction ofmacrophage infiltration and chemoattractant gene expres-sion changes in white adipose tissue of morbidly obesesubjects after surgery-induced weight loss,” Diabetes, vol. 54,no. 8, pp. 2277–2286, 2005.

[147] L. K. Forsythe, J. M. W. Wallace, and M. B. E. Livingstone,“Obesity and inflammation: the effects of weight loss,”Nutrition Research Reviews, vol. 21, no. 2, pp. 117–133, 2008.

[148] K. Clement, N. Viguerie, C. Poitou et al., “Weight loss regu-lates inflammation-related genes in white adipose tissue ofobese subjects,” FASEB Journal, vol. 18, no. 14, pp. 1657–1669, 2004.

[149] G. H. Schernthaner, H. P. Kopp, S. Kriwanek et al., “Effect ofmassive weight loss induced by bariatric surgery on serumlevels of interleukin-18 and monocyte-chemoattractant-protein-1 in morbid obesity,” Obesity Surgery, vol. 16, no. 6,pp. 709–715, 2006.

[150] J. M. Bruun, C. Verdich, S. Toubro, A. Astrup, and B.Richelsen, “Association between measures of insulin sensitiv-ity and circulating levels of interleukin-8, interleukin-6 andtumor necrosis factor-α. Effect of weight loss in obese men,”European Journal of Endocrinology, vol. 148, no. 5, pp. 535–542, 2003.

[151] K. Clement, N. Viguerie, C. Poitou et al., “Weight loss regu-lates inflammation-related genes in white adipose tissue ofobese subjects,” FASEB Journal, vol. 18, no. 14, pp. 1657–1669, 2004.

[152] S. D. Navaneethan, H. Yehnert, F. Moustarah, M. J. Schreiber,P. R. Schauer, and S. Beddhu, “Weight loss interventions inchronic kidney disease: a systematic review and meta-ana-lysis,” Clinical Journal of the American Society of Nephrology,vol. 4, no. 10, pp. 1565–1574, 2009.

[153] K. Masuo, H. Rakugi, T. Ogihara, M. D. Esler, and G. W.Lambert, “Effects of weight loss on renal function in over-weight Japanese men,” Hypertension Research, vol. 34, no. 8,pp. 915–921, 2011.

Page 46: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 11

[154] K. A. Amin, H. H. Kamel, and M. A. Abd Eltawab, “Protec-tive effect of Garcinia against renal oxidative stress and bio-markers induced by high fat and sucrose diet,” Lipids inHealth and Disease, vol. 10, article 6, 2011.

[155] M. D. Gades, H. Van Goor, G. A. Kaysen, P. R. Johnson, B. A.Horwitz, and J. S. Stern, “Brief periods of hyperphagia causerenal injury in the obese Zucker rat,” Kidney International,vol. 56, no. 5, pp. 1779–1787, 1999.

[156] P. R. Johnson, J. S. Stern, B. A. Horwitz, R. E. Harris, and S.F. Greene, “Longevity in obese and lean male and female ratsof the Zucker strain: prevention of hyperphagia,” AmericanJournal of Clinical Nutrition, vol. 66, no. 4, pp. 890–903, 1997.

[157] P. G. Schmitz, M. P. O’Donnell, B. L. Kasiske, S. A. Katz, andW. F. Keane, “Renal injury in obese Zucker rats: glomerularhemodynamic alterations and effects of enalapril,” AmericanJournal of Physiology, vol. 263, no. 3, pp. F496–F502, 1992.

[158] P. N. Chander, O. Gealekman, S. V. Brodsky et al., “Nephro-pathy in Zucker diabetic fat rat is associated with oxidativeand nitrosative stress: prevention by chronic therapy witha peroxynitrite scavenger ebselen,” Journal of the AmericanSociety of Nephrology, vol. 15, no. 9, pp. 2391–2403, 2004.

[159] A. A. Saad, M. I. Youssef, and L. K. El-Shennawy, “Cisplatininduced damage in kidney genomic DNA and nephrotoxicityin male rats: the protective effect of grape seed proantho-cyanidin extract,” Food and Chemical Toxicology, vol. 47, no.7, pp. 1499–1506, 2009.

[160] J. Safa, H. Argani, B. Bastani et al., “Protective effect of grapeseed extract on gentamicin-induced acute kidney injury,”Iranian Journal of Kidney Diseases, vol. 4, no. 4, pp. 285–291,2010.

[161] V. Stefanovic, V. Savic, P. Vlahovic, T. Cvetkovic, S. Najman,and M. Mitic-Zlatkovic, “Reversal of experimental myo-globinuric acute renal failure with bioflavonoids from seedsof grape,” Renal Failure, vol. 22, no. 3, pp. 255–266, 2000.

[162] C. M. Liu, J. Q. Ma, and Y. Z. Sun, “Quercetin protects the ratkidney against oxidative stress-mediated DNA damage andapoptosis induced by lead,” Environmental Toxicology andPharmacology, vol. 30, no. 3, pp. 264–271, 2010.

[163] R. Gredilla and G. Barja, “The role of oxidative stress inrelation to caloric restriction and longevity,” Endocrinology,vol. 146, no. 9, pp. 3713–3717, 2005.

[164] J. S. Stern, M. D. Gades, C. M. Wheeldon, and A. T. Borchers,“Calorie restriction in obesity: prevention of kidney diseasein rodents,” Journal of Nutrition, vol. 131, no. 3, pp. 913S–917S, 2001.

Page 47: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

Hindawi Publishing CorporationInternational Journal of NephrologyVolume 2012, Article ID 381320, 14 pagesdoi:10.1155/2012/381320

Review Article

Reactive Oxygen Species Modulation of Na/K-ATPase RegulatesFibrosis and Renal Proximal Tubular Sodium Handling

Jiang Liu,1 David J. Kennedy,2 Yanling Yan,1, 3 and Joseph I. Shapiro1

1 Department of Medicine, College of Medicine University of Toledo, 3000 Arlington Avenue, Toledo, OH 43614, USA2 Department of Cell Biology, Lerner Research Institute, Cleveland Clinic, 9500 Euclid Avenue, Cleveland, OH 44195, USA3 Institute of Biomedical Engineering, Yanshan University, Qinhuangdao 066004, China

Correspondence should be addressed to Joseph I. Shapiro, [email protected]

Received 7 October 2011; Accepted 7 November 2011

Academic Editor: Ayse Balat

Copyright © 2012 Jiang Liu et al. This is an open access article distributed under the Creative Commons Attribution License,which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

The Na/K-ATPase is the primary force regulating renal sodium handling and plays a key role in both ion homeostasis and bloodpressure regulation. Recently, cardiotonic steroids (CTS)-mediated Na/K-ATPase signaling has been shown to regulate fibrosis,renal proximal tubule (RPT) sodium reabsorption, and experimental Dahl salt-sensitive hypertension in response to a high-saltdiet. Reactive oxygen species (ROS) are an important modulator of nephron ion transport. As there is limited knowledge regardingthe role of ROS-mediated fibrosis and RPT sodium reabsorption through the Na/K-ATPase, the focus of this review is to examinethe possible role of ROS in the regulation of Na/K-ATPase activity, its signaling, fibrosis, and RPT sodium reabsorption.

1. Introduction

According to the American Heart Association (AHA), over70 million people in the United States aged 20 and olderhave high blood pressure (BP). The cause of 90–95 percentof the cases of high BP is unknown, yet in the last decadethe associated morbidity and mortality from high BP hasincreased precipitously. In the 2008 AHA Scientific State-ment [1], excessive dietary salt intake is listed as one ofthe major lifestyle factors which significantly contributesto the development of hypertension and tends to be morepronounced in typical salt-sensitive patients. Modest dietarysalt restriction and diuretic therapy, therefore, are recom-mended for treatment of resistant hypertension, especiallyin the salt-sensitive subgroup [1, 2]. Renal sodium handlingis a key determinant of long-term BP regulation [3]. Therelationship between dietary sodium, salt sensitivity, andBP has been established on an epidemiological and clinicalbasis. It is estimated that hypertension affects 25% to 35%of the world population aged 18 and older [4], and morehypertensive subjects (∼50%) are significantly salt sensitivethan normotensive subjects (∼25%) [5]. In the DASH-Sodium clinical trial, BP reduction was correlated withsodium restriction in the salt-sensitive subjects regardless

of diet [6]. Interestingly, animal renal cross-transplantationexperiments [7–10] and studies of human renal transplanta-tion [11] demonstrate that BP levels “travel with the donor’skidney,” providing compelling evidence for the role of renalfunction in the pathogenesis of hypertension. In clinical andexperimental models, renal proximal tubule (RPT) sodiumhandling accounts for over 60% reabsorption of filteredsodium and is an independent determinant of BP responseto salt intake, playing a critical role in the pathogenesisof salt-sensitive hypertension. Recently, accumulating dataindicate that cardiotonic steroids (CTS) signaling throughthe Na/K-ATPase contribute to RPT sodium handling andsalt sensitivity [12–18].

2. CTS and Na/K-ATPase Signaling

CTS (also known as endogenous digitalis-like substances)include plant-derived digitalis drugs, such as digoxin andouabain, and vertebrate-derived aglycones such as bufalinand marinobufagenin (MBG) [16, 18]. Recent studies haveidentified both ouabain and MBG as endogenous steroidswhose production, and secretion are regulated by stimuliincluding angiotensin II (Ang II) [18–21]. The Na/K-ATPase

Page 48: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

2 International Journal of Nephrology

belongs to the P-type ATPases family and consists of twononcovalently linked α and β subunits [22–24]. Several αand β isoforms, expressed in a tissue-specific manner, havebeen identified and functionally characterized [22–25]. Theα1 subunit contains multiple structural motifs that interactwith soluble, membrane and structural proteins includingSrc, caveolin-1, phospholipase C-γ, PI3 kinase, IP3 receptor,BiP, calnexin, cofilin, and ankyrin [26–36]. Binding to theseproteins not only regulates the ion-pumping function ofthe enzyme, but it also conveys signal-transducing functionsto the Na/K-ATPase [18, 32, 37–39]. In LLC-PK1 cells,the Na/K-ATPase α1 subunit and Src form a functionalreceptor in which the binding of CTS to the α1 subunitactivates Src and consequent signaling cascade [39]. Thesignaling function of the Na/K-ATPase regulates numerouscell functions in various types of organs and cells includingcell motility, cell proliferation, cancer, endothelin release,glycogen synthesis, apoptosis, hypertension, intracellularcalcium signaling, cardiac hypertrophy, cardiac remodeling,renal remodeling, epithelial cell tight junction, vascular tonehomeostasis, and sodium homeostasis [26, 40–59]. The topicof CTS-Na/K-ATPase signaling and its downstream patho-physiological implications has been extensively reviewed inthe last few years [12–16, 18, 21, 32, 39, 59–64], and we willnot discuss them in detail in this review.

3. CTS and Sodium Homeostasis

Endogenous CTS were first proposed many years ago tofunction as a natriuretic hormone. Although their patho-physiological significance has been a subject of debate formany years [65], the concept of a natriuretic hormone issupported by the experimental observations in animal mod-els of ouabain-induced natriuresis [66, 67]. Recently, severalelegant reports have confirmed this concept with differentapproaches. Gene replacement studies have unequivocallydemonstrated an important role of endogenous CTS inregulation of renal sodium excretion and BP [68, 69]. Intransgenic mice expressing ouabain-sensitive Na/K-ATPaseα1 subunit, a significant observation is an augmentednatriuretic response to both acute salt load and ouabaininfusion, indicating that the ouabain-binding site of Na/K-ATPase α1 subunit participates in the natriuretic response tosalt load by responding to endogenous ouabain. Moreover,the augmented natriuretic response in the ouabain-sensitiveα1 isoform mice can be blocked with administration ofan anti-digoxin antibody fragment [68, 69]. In normalmale Wistar rats, endogenous circulatory ouabain hasphysiological roles controlling vasculature tone and sodiumhomeostasis, showing endogenous ouabain regulates renalsodium excretion in normal animals [70]. A significantinhibition of natriuresis is observed in rats that were passivelyimmunized with anti-ouabain antibody and in rats thatwere actively immunized with ouabain-albumin antigen, inwhich endogenous ouabain levels were reduced in both cases.Like ouabain, MBG has both natriuretic and vasoconstrictoreffects [12, 71, 72]. High salt intake induced an initial tran-sient stimulation of ouabain and a subsequent progressive

increase of MBG both in Dahl salt-sensitive (S) rats and inhumans [12, 73]. In Dahl S rats, the increase in natriuresisstimulated by an acute salt loading is prevented by admin-istration of both an anti-ouabain antibody and an anti-MBG antibody [72]. Furthermore, endogenous CTS havealso been implied in age-related increases in salt sensitivityof BP in human normotensive subjects [73]. It is estimatedthat approximately 50% of humans with untreated essentialhypertension have significantly elevated levels of endogenousouabain [74] which is involved in the regulation of vasculartone homeostasis through stimulation of interaction betweenthe Na/K-ATPase and sodium/calcium exchanger [59]. Innormotensive human males, both a high salt diet and system-atic sodium depletion (by hydrochlorothiazide) significantlyincrease plasma ouabain concentration [75].

Release of endogenous ouabain from the brain hypotha-lamus stimulated by a high salt diet leads to inhibition ofthe Na/K-ATPase activity and central sympathetic activation(reviewed in [76, 77]), which plays a crucial role in the pres-sor effects of high salt intake in spontaneously hypertensiverats (SHR) and Dahl S rats. In Dahl S rats, elevation in brainouabain increased MBG secretion from the adrenal cortex,and this effect was blocked by the AT1 receptor antagonistlosartan [78, 79]. The data suggest that the observed CTS-induced natriuretic effect might be a result of the combinedeffects of ouabain and MBG [60]. Furthermore, CTS notonly induced hypertension in rats but also caused significantcardiovascular remodeling and natriuresis independent oftheir effect on BP [49, 51, 70, 80, 81].

4. CTS and Oxidative Stress in Cardiac andRenal Fibrosis

CTS binding to the Na/K-ATPase induces cellular reactiveoxygen species (ROS) production and its downstream effects,such as cardiac and renal fibrosis, and these effects can beprevented by ROS scavenging [80, 82–84]. In kidney andheart, the central role of CTS in the development of fibrosishas been demonstrated both in vivo, in the partial (5/6th)nephrectomy model, and in vitro cell culture, includingcardiac and renal fibroblasts. 5/6th nephrectomy increasescirculating levels of MBG and stimulates cardiac fibrosisin both rat and mouse [80, 84, 85]. Rats subjected to5/6th nephrectomy develop systemic oxidant stress that issimilar to that seen in rats subjected to MBG infusion asevidenced by significant elevation of plasma carbonylatedprotein. Active immunization against MBG and reductionof circulating levels of MBG by adrenalectomy substan-tially attenuate 5/6th nephrectomy and MBG infusion-mediated cardiac fibrosis and oxidant stress, an effect thatis independent of BP [80, 84]. In primary culture of ratcardiac and human dermal fibroblasts as well as a cellline derived from rat renal fibroblasts, MBG and ouabainstimulate [3H]proline incorporation as well as gene andprotein expression of collagen [86]. MBG induced a PLC-dependant translocation of PKC-δ to the nucleus, resultingin the phosphorylation and degradation of transcriptionfactor Friend leukemia integration-1 (Fli-1), a negative

Page 49: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 3

Partial (5/6th) nephrectomy

Fli-1-p

ROS

Na/K-ATPase/c-Src

Shift of cardiac and renal function curve

CTS

PLC

Fli-1

Fibrosis

EMT

Collagen production

AntioxidantsImmunizationSpironolactoneCanrenone

dermal fibroblastsRenal cortexRenal proximal tubules

PKC-δ

Snail

Cardiac, renal, and

Figure 1: Schematic illustration of the effect of CTS on fibrosis.

regulator of collagen synthesis [86]. Both CTS-inducedNa/K-ATPase signaling and oxidative stress are necessary inthe pathogenesis of cardiac and renal fibrosis as evidencedby CTS-stimulated phosphorylation of both Src and MAPKwhich is effectively blocked not only by ROS scavengingand Src inhibition but also through possible competitiveinhibition of CTS binding to Na/K-ATPase by spironolactoneand canrenone [80, 84, 87, 88]. MBG infusion causes renalfibrosis mainly in the cortex of the kidney by stimulationof the transcription factor Snail expression and its nuclearlocalization in the tubular epithelia, which is associated withepithelial-to-mesenchymal transition (EMT) during renalfibrosis [89]. This EMT phenomenon is also demonstratedin LLC-PK1 cells, indicating that MBG may cause damageof renal proximal tubules. CTS-induced fibrosis in heart andkidney might shift the cardiac and renal function curve tofavor a higher set point of pressure natriuresis (Figure 1).

5. Na/K-ATPase Signaling and RPTSodium Handling

It has been postulated for decades that endogenous CTSstimulated by increased sodium intake increases natriuresisand diuresis by directly inhibiting renal tubular Na/K-ATPase to prevent renal reabsorption of filtered sodium [90–92]. There is accumulating evidence supporting this ideaunder conditions such as high salt intake, chronic renalsodium retention, renal ischemia, uremic cardiomyopathy,and volume expansion in various animal models and humanbeings [12, 13, 15, 60, 62, 75, 93–107]. Although thedirect inhibition of the Na/K-ATPase enzymatic activityand sodium reabsorption in RPTs by CTS has not beenvalidated, recent observations indicate that ligand-mediatedRPT sodium reabsorption via Na/K-ATPase/c-Src signalingcounterbalances the sodium retention-mediated increasesin BP, such as that seen in salt-sensitive hypertension

[108–114]. Ouabain, a ligand of the Na/K-ATPase, acti-vates the Na/K-ATPase/c-Src signaling pathway and sub-sequently redistributes basolateral Na/K-ATPase and apicalsodium/hydrogen exchanger isoform 3 (NHE3) in RPTs,leading to reduced RPT sodium reabsorption and increasedurinary sodium excretion.

In LLC-PK1 cells, ouabain activates Na/K-ATPase/c-Srcsignaling pathways and reduces cell surface Na/K-ATPase andNHE3, leading to a significant inhibition of active transcel-lular 22Na+ flux from the apical to basolateral compartment[108–112]. MBG, an important CTS species, and depro-teinated extract of serum (derived from patients with chronicrenal failure) also induce Na/K-ATPase endocytosis andinhibition of active transepithelial 22Na+ flux. However, it isstill not clear how ouabain-activated Na/K-ATPase signalingregulates NHE3, since, at concentrations of ouabain used,no significant change in intracellular Na+ concentration wasobserved [108]. Interestingly, this phenomenon is either notobserved or is much less significant in MDCK cells (a caninerenal distal tubule cell line). These in vitro data suggest thatCTS-Na/K-ATPase signaling has a profound effect on RPTsodium handling, but not in distal tubules.

Different species of endogenous CTS show differences inthe kinetics and tissue actions in response to salt loading inboth animal models and in human salt-sensitive hypertensivepatients [16, 18, 60, 75, 79, 105]. The Dahl salt-resistant(R) and salt-sensitive (S) strains were developed by selectivebreeding of the outbred Sprague-Dawley rat strain forresistance or susceptibility to the hypertensive effects of highdietary sodium [115]. RPT sodium handling is a criticaldeterminant of the different BP responses in these strains[7, 116–118], and there is no Na/K-ATPase α1 gene (Atp1a1)difference between R and S rats [119]. In comparison toDahl S rats that eliminate excessive sodium mainly throughpressure natriuresis at the expense of an elevated systolicBP, a major response to salt loading in the Dahl R rats is agreater reduction in renal sodium reabsorption to eliminateexcessive sodium without raising BP. Our recent in vivoobservation [114] is in agreement with this hypothesis andour in vitro observations in LLC-PK1 cells. Specifically inisolated RPTs, both a high salt diet and ouabain are ableto activate Na/K-ATPase/c-Src signaling pathways, leading tothe redistribution of Na/K-ATPase and NHE3 in the Dahl Rbut not in the S rats. The R rats show significant increases intotal urinary sodium excretion and RPT-mediated fractionalsodium excretion, without BP elevation [114]. While theBP response to salt loading in R and S rats involves manyregulatory factors [117], our data indicate that impairmentof the RPT Na/K-ATPase/c-Src signaling contributes to saltsensitivity. Since it failed to confirm the possible differenceof Na/K-ATPase α1 gene (Atp1a1) between R and S rats[119], other factor(s) must be present to prevent activationof Na/K-ATPase/c-Src signaling in the S rats. The possibleeffect of CTS on RPT Na/K-ATPase signaling and sodiumreabsorption is summarized in Figure 2.

The SHR rat is an established model of human essentialhypertension with the characteristic of vascular resistance.SHR rat develops hypertension spontaneously at the ageof 7–15 weeks, regardless of salt loading, mainly through

Page 50: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

4 International Journal of Nephrology

ROS

RPT Na/K-ATPase

c-Src ROS

Redistribution ofNa/K-ATPase and NHE3

RPT sodium reabsorption

Desensitization ofNa/K-ATPase

CTS

Antioxidants

High-salt diet

Salt sensitivity

Figure 2: Schematic illustration of the effects of ROS and Na/K-ATPase signaling on RPT sodium reabsorption. CTS: cardiotonicsteroids; RPT: renal proximal tubules.

increase in peripheral vascular resistance [120] includingrenal vascular resistance. Interestingly, either reduction ofdietary vitamin E or caloric restriction without sodiumrestriction prevents the development of hypertension [121,122]. Like Dahl S rats, SHR rats eliminate excessive sodiumvia a pressure-natriuresis mechanism but have significantlower BP response to a high salt diet and higher systolic BPto eliminate the same amount of sodium when comparedto Dahl S rats [123–126]. Most interestingly, comparingto control Wistar-Kyoto (WKY) rats, regulation of theNa/K-ATPase and NHE3 with both aging and oxidativestress have been shown contributed to the developmentand maintenance of hypertension in the SHR [127–135].In renal cortical (proximal) tubules, activity and proteinlevels of NHE-3 are significantly higher in SHR than age-matched WKY rats at all stages during the developmentand maintenance of hypertension. When NHE3 function isdetermined by the rate of bicarbonate reabsorption by invivo stationary microperfusion in RPT, young SHR rats showhigher NHE3 activity than adult SHR and WKY rats, andthis is accompanied by changes in NHE3 phosphorylationand distribution. In young SHR rats, the RPT Na/K-ATPaseactivity is significantly higher than in age-matched WKY,which can be prevented by treatment of the diuretic drughydrochlorothiazide. However, the Na-K-ATPase activity inmedullary thick ascending limb, cortical thick ascendinglimb, and distal tubule is significantly lower in young SHRrats than in age-matched WKY rats. There were no signif-icant differences in Na/K-ATPase activity in these nephronsegments in adult SHR and WKY rats, nor in collecting ductsegments of young and adult SHR and age-matched WKYrats. Interestingly, however, the Na/K-ATPase α1 subunitgene expression is lower in both young and adult SHR ratsthan age-matched WKY rats, indicating a posttranscriptionalregulation.

Recent study indicates that SHR rats have enhancedrenal superoxide generation and NADPH oxidase (NOX)expression in both vascular and renal tissue before and

after development of hypertension [136]. Elevated basal levelof superoxide inhibits proximal tubule NHE3 activity andfluid reabsorption in SHR rats in comparison to WKY rats.This effect is prevented by the NOX inhibitor apocyninor knockdown of the critical NOX subunit p22phox withsmall interfering RNA, indicating that increased basal levelof superoxide impairs RPT function [137]. Furthermore,in Sprague-Dawley rats, oxidative stress impairs Ang II-mediated regulation of NHE3 [138].

6. Oxidative Stress and Regulationof Na/K-ATPase

Many stimuli including hypoxia and dopamine induce acell- and tissue-specific endocytosis and exocytosis of Na/K-ATPase and a change in Na/K-ATPase activity [33, 61, 139].Ouabain-stimulated ROS generation functions as a secondmessenger in ouabain-activated Na/K-ATPase signaling inisolated rat cardiac myocytes [82, 140–142]. Binding ofouabain to the Na/K-ATPase activates Src kinase, which inturn transactivates EGFR, leading to activation of the Ras-Raf-MEK-ERK pathway [32, 39, 63]. Ras activation leads tothe activation of MAPK and increase in [Ca2+]i which resultin opening of mitochondrial ATP-sensitive K+ channels[142] and generation of mitochondrial ROS [82, 141]. ROSsubsequently activate NF-κB [141, 143] and slow [Ca2+]i

oscillations at nanomolar ouabain concentrations [40].Additionally, ouabain-induced generation of ROS in neona-tal myocytes is antagonized by overexpression of a dominantnegative Ras as well as myxothiazol/diphenyleneiodonium,indicating a mitochondrial origin of the Ras-dependent ROSgeneration [82]. Ouabain also stimulates ROS generation inother cell types [144–146]. Conversely, oxidative stress canactivate the Na/K-ATPase signaling. Both a bolus of H2O2

and exogenously added glucose oxidase (which generates asustained low level of H2O2 by consuming glucose) activatesNa/K-ATPase signaling in cardiac myocytes [140]. Pretreat-ment with the antioxidant N-acetyl cysteine (NAC) preventsouabain-Na/K-ATPase signaling and its downstream effects.Moreover, infusion of CTS causes ROS generation andprotein oxidation in experimental animals [80, 147].

The redox sensitivity of the Na/K-ATPase was first dem-onstrated in electric eels with treatment of H2O2 [148]. Thisphenomenon was further observed in a wide range of animalspecies, tissues, and other species of ROS such as hypochlor-ous anion, hydroxyl radicals, superoxide, hyperchloriteanions, and singlet oxygen. It has been shown that the Na/K-ATPase in skeletal muscle is redox-sensitive and infusionof NAC attenuated ROS-mediated inhibition of maximalpump activity [149]. In dog kidney, oxidative modificationof kidney Na/K-ATPase by H2O2 was accompanied by adecrease in the amount of sulfhydryl (SH) groups [150] and,importantly, oxidative modification can result in formationof Na/K-ATPase oligomeric structure [151]. The differencesin the number, location, and accessibility of SH groups inNa/K-ATPase isozymes might predict their oxidative stability[152]. Different antioxidant enzymes, natural or syntheticantioxidants, and some inhibitors of oxidase activity can

Page 51: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 5

attenuate ROS mediated inhibition of Na/K-ATPase activity.ROS are known to inhibit Na/K-ATPase activity in differenttypes of cells as well. Interestingly, the Na/K-ATPase in ratcerebellar granule cells are redox-sensitive with an “optimalredox potential range,” where ROS levels out of this “optimalrange” are capable of inhibiting Na/K-ATPase activity [153].While the Na/K-ATPase does not contain heme groups, itdoes contain cysteine residues located in α subunit cytosolicloops which may determine the redox sensitivity of the αsubunit. The sensitivity of the Na/K-ATPase to redox andoxygen status, the regulatory factors which govern theseinteractions, and the implied molecular mechanism wererecently reviewed [154].

Increases in ROS can oxidize the Na/K-ATPase α/βsubunits and its independent regulator FXYD proteins. Thisoxidation inhibits its activity and promotes its susceptibilityto degradation by proteasomal and endosomal/lysosomalproteolytic pathways in different cell types including cardiacmyocytes, vascular smooth muscle cells, and RPTs [155–166]. It appears that the oxidized modification of the Na/K-ATPase is a reversible, redox-sensitive modification. How-ever, purified enzyme has also been shown to be irreversiblyinhibited upon exposure to hydrogen peroxide, the super-oxide anion, and the hydroxyl radical [158]. The regulationof renal Na/K-ATPase α1 by oxidants is not dependent onthe ouabain sensitivity of the α1 subunit per se. It appearsthat the α2 and α3 subunits are more sensitive to oxidantsthan the α1 subunit, and ouabain-sensitive α1 (canine)and insensitive-α1 (rat) have similar sensitivity to oxidants,suggesting the regulation of α1 by ROS is not species specific.Furthermore, ROS accelerates degradation of the oxidativelydamaged Na/K-ATPase and the α2 and α3 subunits, whichalso appear to be more susceptible to degradation than the α1subunit [159]. These studies indicate that differential oxidantsensitivities of the Na/K-ATPase subunits are dictated by theprimary sequences of different subunits and different subunitcompositions of the various tissues may contribute to theirrelative susceptibilities to oxidant stress. In the RPT cellline originated from WKY rats, cadmium, a ROS generator,stimulates ROS production and causes a toxic oxidativedamage of the Na/K-ATPase. Oxidative damage increasesNa/K-ATPase degradation through both the proteasomaland endo-/lysosomal proteolytic pathways [163]. In purifiedrenal Na/K-ATPase, peroxynitrite (ONOO−) causes tyrosinenitration and cysteine thiol group modification of the Na/K-ATPase, but only cysteine thiol group modification is impliedin the inhibition of the enzyme activity since glutathione isunable to reverse the inhibition [167]. In isolated rat renalRPTs, peroxynitrite and its signaling participates in Ang II-induced regulation of renal Na/K-ATPase activity [168].

Ang II inhibits the Na/K-ATPase via PKC-dependentNOX activation. The dependence of Ang II-induced Na/K-ATPase inhibition on NOX and superoxide, as well asreversible oxidative modification of the Na/K-ATPase,strongly suggests a role for redox signaling [164–166]. Incardiac myocytes and pig kidney, oxidative stress inducesglutathionylation of the β1 subunit of the Na/K-ATPase.In purified pig renal Na/K-ATPase, peroxynitrite inhibitsthe enzymatic activity by stabilization of the enzyme in an

E2-prone conformation. At the same time, FXYD proteinsreverse oxidative stress-induced inhibition of the Na/K-ATPase by facilitating deglutathionylation of the β1 subunit.Moreover, both tyrosine kinase c-Src and cell membranestructural component lipid rafts, which are critical in Na/K-ATPase/c-Src signaling, are also critical in redox signalingplatform formation [169–172]. These observations, alongwith the fact that c-Src is redox-sensitive [173] and itsactivation regulates NOX-derived superoxide generation[174], suggests a redox-sensitive Na/K-ATPase/c-Src signal-ing cascade and its possible role in ROS regulation, althoughthe mechanism is not clear. Our unpublished data suggestthat certain basal levels of ROS might be required for theinitiation of ouabain-Na/K-ATPase/c-Src signaling. In LLC-PK1 cells, pretreatment with higher concentrations of NAC(5 and 10 mM for 30 min), but not with lower concentrationof 1 mM, can prevent ouabain-stimulated c-Src activationand the redistribution of Na/K-ATPase and NHE3. Thissuggests that ROS may stabilize Na/K-ATPase in a certainconformational status in order to facilitate ouabain bindingto the Na/K-ATPase α1 subunit and favor ouabain-Na/K-ATPase/c-Src signaling. Some pertinent questions remainto be resolved, such as how ROS interacts with and influ-ences the Na/K-ATPase/c-Src signaling cascade and whetherouabain-induced ROS boosts the Na/K-ATPase signaling bya positive feedback mechanism and chronically desensitizesthe signaling cascade by stimulating Na/K-ATPase/c-Srcendocytosis (Figure 2).

7. Oxidative Stress and Regulationof Renal Function

ROS function as important intracellular and extracellularsecond messengers to modulate many signaling molecules.Physiological concentrations of ROS play an importantrole in normal redox signaling, while pathological levelsof ROS contribute to renal and vascular dysfunction andremodeling through oxidative damage [175–177]. Geneticfactor(s) partially contribute to high basal ROS levels and thedevelopment of hypertension [178, 179].

Oxidative stress has been shown to regulate BP andsodium handling in various animal models. High saltintake increases oxidative stress, and this has importantimplications for the regulation of cardiovascular and renalsystems. An increase in oxidative stress is both a cause andconsequence of hypertension [177, 180–183]. Renal NOX ispresent in the renal cortex, medulla, and vasculature. NOX,the major source of superoxide in the kidney, is of particularinterest because of its prominent expression and implicationin pathophysiology. In the kidney, increased oxidative stressinfluences a number of physiologic processes including renalsodium handling in the proximal tubule [137, 168, 184, 185]and thick ascending limb [186], renal medulla blood flow[183, 187–190], descending vasa recta contraction [191, 192]in addition to interactions with other regulatory systemssuch as the dopamine signaling pathway [193]. Increasedoxidative stress has been shown to contribute to saltsensitivity [194, 195]. In macula densa cells, NOX isoform

Page 52: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

6 International Journal of Nephrology

Nox2 is responsible for salt-induced superoxide generation,while Nox4 regulates basal ROS [196]. In the medullar thickascending limb of the loop of Henle, both exogenous andendogenous superoxides stimulate sodium absorption [197].Also, in this nephron segment, NOX-induced generation ofsuperoxide enhances sodium absorption by reduction of thebioavailability of nitric oxide (NO) to prevent NO-inducedreduction of NaCl absorption [198], which contributes tosalt-sensitive hypertension observed in Dahl salt-sensitiverats [190].

In PRTs, in particular, increases in ROS inhibit the Na/K-ATPase as well as the apical NHE3 and sodium/glucosecotransporter, in order to promote RPT sodium excretionunder certain circumstances [137, 168, 184, 185]. Whileelevated basal level of superoxide inhibits proximal tubuleNHE3 activity and fluid reabsorption in SHR rats in com-parison to WKY rats [137], peroxynitrite and its signalingparticipates in Ang II-induced regulation of renal Na/K-ATPase activity in isolated rat renal RPTs [168]. In thepathogenesis of diabetic nephropathy, a high level of glucose-induced ROS generation induced by stimulation of mito-chondrial metabolism and NOX activity in RPT primarycultures leads to inhibition of the expression and activityof the sodium/glucose cotransport system [185]. In maleWistar rats, a high salt diet (3% NaCl for 2 weeks) promotessodium/water excretion and urinary 8-isoprostane excretion,a marker of oxidative stress, which can be attenuated bytreatment with apocynin, an NOX inhibitor. The salt loadingleads to increased generation of ROS and a state of oxidativestress in the cortex but not to such a degree in the medulla[199].

RPT sodium and/or fluid absorption in the normal rat isreduced by inhibition of NO synthesis, while NO promotesRPT Na+ and/or fluid reabsorption [200]. In immortalizedand freshly isolated RPTs from the WKY and SHR rats,the basal level of membrane NOX activity is greater inSHRs [201]. Moreover, NOX-induced superoxide genera-tion inhibits RPT fluid reabsorption in SHRs [137]. InSprague-Dawley rats treated with the oxidant L-buthioninesulfoximine, Ang II overstimulates RPT Na/K-ATPase andNOX and leads to increased sodium reabsorption, whichis prevented by administration of the superoxide scavengerTempol [202].

High salt diet, which is well documented for its stim-ulation of systematic oxidative stress, regulates the activityand distribution of the Na/K-ATPase and NHE3 in differentanimal models [203–207]. PRT sodium reabsorption signif-icantly affects water and sodium homeostasis by regulatingredistribution of ion transporters in response to high saltintake [208]. Regulation of NHE3 activity and distribution aswell as PRT fluid reabsorption contribute to the developmentand maintenance of hypertension in young and adult SHRrats [127, 137, 209].

It has become clear that antioxidant agents such asTempol and enzymes such as heme oxygenase-1 (HO-1)exhibit a beneficial and protective effect on BP in variousanimal models of hypertension. As an example, inhibitionof HO activity reduces renal medullary blood flow [84,210, 211], total renal blood flow (RBF) [212], glomerular

filtration rate (GFR), and renal production of nitric oxide[213]. Inhibition of HO-1 increases mean arterial pressurein Sprague-Dawley rats [214, 215] and SHR rats [216].Induction of HO-1 reduces BP in SHR rats [217–220]. Theeffect of HO-1 on BP is presumably through the carbonmonoxide (CO)/HO system and depression of cytochromep-450-derived 20-HETE. In the Dahl salt-dependent modelof systemic hypertension, induction of HO-1 occurred in thevasculature and is accompanied by endothelial dysfunction[221, 222]. In Dahl salt-sensitive rats, a high salt dietincreases HO-1 expression and CO generation in aorta andcoronary arteries [221, 223], and, in Dahl salt-sensitive ratsfed low salt diet, induction of HO-1 expression attenuatesoxidative stress and reduces proteinuria and renal injury[224]. However, most studies examining the contributionof HO-1 to BP regulation have focused on the vasculature,that is, pressure-natriuresis and arterial BP, and so the roleof HO-1 in RPT-mediated sodium handling is still poorlyunderstood. Nevertheless, increased HO expression in RPTscould result in an increased ability to buffer locally producedoxidants, leading to their neutralization.

The beneficial effect of antioxidants is controversial andnot seen in most clinic trials with administration of antioxi-dants (reviewed in [177, 225]). The Dietary Approaches toStop Hypertension (DASH) study and subsequent studieshave demonstrated that lower BP associated with reduceddietary salt intake may be related to reductions in oxidativestress [6, 226–228]. Interestingly, however, while a com-bination antioxidant supplement (with an ascorbic acid,synthetic vitamin E, and β-carotene) had no improvementon BP after 5-year treatment [229], another combinationantioxidant supplement (zinc, ascorbic acid, α-tocopherol,and β-carotene) did result in a significant reduction insystolic BP [230]. Other studies have also shown that certainantioxidants, such as glutathione and vitamin C, have ablood-pressure-lowering effect [231, 232]. However, antiox-idant supplementation may be ineffective or even dangerous[233] due to the possible “over-antioxidant-buffering” effectof excessive antioxidant supplementation. In this scenario,excess antioxidants might become pro-oxidants (by pro-viding H+) if they cannot promptly be reduced by thefollowing antioxidant in the biological antioxidant chain.Thus, it appears that the balance of the ROS status, withina physiological range, may be more important to maintainbeneficial ROS signaling.

8. Perspective

Renal sodium handling is a key determinant of blood pres-sure. ROS status, among others, is an important regulatorof vasculature and sodium handling. However, the effectof ROS and Na/K-ATPase, especially of their interaction,on RPT sodium reabsorption has only been explored in alimited fashion. Coordinated regulation of two major iontransporters, the basolateral Na/K-ATPase and the apicalNHE3, has been implicated in the counterbalancing of highsalt intake (volume expansion) mediated blood pressureincrease. The Na/K-ATPase/c-Src signaling regulates this

Page 53: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 7

coordinated regulatory mechanism and impairment of thissignaling cascade contributes to experimental Dahl salt-sensitive hypertension. Both the Na/K-ATPase (α and βsubunit) and its proximal signaling partner c-Src are redoxsensitive, suggesting a redox-sensitive Na/K-ATPase/c-Srcsignaling complex. However, the mechanisms remain largelyto be elucidated since the available data is limited [184,234]. Nevertheless, some pertinent questions remain to beaddressed. In the future, it will be important to explorewhether ROS signaling is a link between the Na/K-ATPase/c-Src cascade and NHE3 regulation and how oxidative stressstimulated by high salt and CTS regulates Na/K-ATPase/c-Src signaling in renal sodium handling and fibrosis.

References

[1] D. A. Calhoun, D. Jones, S. Textor et al., “Resistant hyper-tension: siagnosis, evaluation, and treatment a scientificstatement from the american heart association professionaleducation committee of the council for high blood pressureresearch,” Hypertension, vol. 51, no. 6, pp. 1403–1419, 2008.

[2] F. J. He and G. A. MacGregor, “Effect of longer-term modestsalt reduction on blood pressure,” Cochrane Database ofSystematic Reviews, no. 3, Article ID CD004937, 2004.

[3] A. C. Guyton, “Blood pressure control—special role of thekidneys and body fluids,” Science, vol. 252, no. 5014, pp.1813–1816, 1991.

[4] J. A. Staessen, J. Wang, G. Bianchi, and W. H. Birkenhager,“Essential hypertension,” Lancet, vol. 361, no. 9369, pp.1629–1641, 2003.

[5] M. H. Weinberger, J. Z. Miller, and F. C. Luft, “Definitionsand characteristics of sodium sensitivity and blood pressureresistance,” Hypertension, vol. 8, no. 6, pp. II127–II134, 1986.

[6] F. M. Sacks, L. P. Svetkey, W. M. Vollmer et al., “Effects onblood pressure of reduced dietary sodium and the dietaryapproaches to stop hypertension (DASH) diet,” New EnglandJournal of Medicine, vol. 344, no. 1, pp. 3–10, 2001.

[7] L. K. Dahl, M. Heine, and K. Thompson, “Genetic influenceof the kidneys on blood pressure. Evidence from chronicrenal homografts in rats with opposite predispositions tohypertension,” Circulation Research, vol. 40, no. 4, pp. 94–101, 1974.

[8] G. Bianchi et al., “The hypertensive role of the kidneyin spontaneously hypertensive rats,” Clinical Science andMolecular Medicine, vol. 23, no. 45, Supplement 1, pp. S135–S139, 1973.

[9] J. Heller, G. Schubert, J. Havlickova, and K. Thurau, “Therole of the kidney in the development of hypertension:a transplantation study in the Prague hypertensive rat,”Pflugers Archiv European Journal of Physiology, vol. 425, no.3-4, pp. 208–212, 1993.

[10] D. A. Morgan, G. F. DiBona, and A. L. Mark, “Effects of inter-strain renal transplantation of NaCl-induced hypertension inDahl rats,” Hypertension, vol. 15, no. 4, pp. 436–442, 1990.

[11] J. J. Curtis, R. G. Luke, and H. P. Dustan, “Remissionof essential hypertension after renal transplantation,” NewEngland Journal of Medicine, vol. 309, no. 17, pp. 1009–1015,1983.

[12] A. Y. Bagrov and J. I. Shapiro, “Endogenous digitalis:pathophysiologic roles and therapeutic applications,” NatureClinical Practice Nephrology, vol. 4, no. 7, pp. 378–392, 2008.

[13] V. M. Buckalew, “Endogenous digitalis-like factors. Anhistorical overview,” Frontiers in Bioscience, vol. 10, no. 1, pp.2325–2334, 2005.

[14] O. V. Fedorova, J. I. Shapiro, and A. Y. Bagrov, “Endoge-nous cardiotonic steroids and salt-sensitive hypertension,”Biochimica et Biophysica Acta, vol. 1802, no. 12, pp. 1230–1236, 2010.

[15] J. Liu and Z. J. Xie, “The sodium pump and cardiotonicsteroids-induced signal transduction protein kinases andcalcium-signaling microdomain in regulation of transportertrafficking,” Biochimica et Biophysica Acta, vol. 1802, no. 12,pp. 1237–1245, 2010.

[16] W. Schoner and G. Scheiner-Bobis, “Role of endogenouscardiotonic steroids in sodium homeostasis,” NephrologyDialysis Transplantation, vol. 23, no. 9, pp. 2723–2729, 2008.

[17] P. Meneton, X. Jeunemaitre, H. E. De Wardener, and G.A. Macgregor, “Links between dietary salt intake, renalsalt handling, blood pressure, and cardiovascular diseases,”Physiological Reviews, vol. 85, no. 2, pp. 679–715, 2005.

[18] W. Schoner and G. Scheiner-Bobis, “Endogenous and exoge-nous cardiac glycosides: their roles in hypertension, saltmetabolism, and cell growth,” American Journal of Physiology,Cell Physiology, vol. 293, no. 2, pp. C509–C536, 2007.

[19] J. M. Hamlyn, M. P. Blaustein, S. Bova et al., “Identificationand characterization of a ouabain-like compound fromhuman plasma,” Proceedings of the National Academy ofSciences of the United States of America, vol. 88, no. 14, pp.6259–6263, 1991.

[20] J. Laredo, J. R. Shah, Z. R. Lu, B. P. Hamilton, and J. M.Hamlyn, “Angiotensin II stimulates secretion of endogenousouabain from bovine adrenocortical cells via angiotensintype 2 receptors,” Hypertension, vol. 29, no. 1, pp. 401–407,1997.

[21] A. Y. Bagrov, J. I. Shapiro, and O. V. Fedorova, “Endogenouscardiotonic steroids: physiology, pharmacology, and noveltherapeutic targets,” Pharmacological Reviews, vol. 61, no. 1,pp. 9–38, 2009.

[22] K. J. Sweadner, “Isozymes of the Na+/K+-ATPase,” Biochimicaet Biophysica Acta, vol. 988, no. 2, pp. 185–220, 1989.

[23] J. H. Kaplan, “Biochemistry of Na,K-ATPase,” Annual Reviewof Biochemistry, vol. 71, pp. 511–535, 2002.

[24] G. Blanco and R. W. Mercer, “Isozymes of the Na-K-ATPase:heterogeneity in structure, diversity in function,” AmericanJournal of Physiology, vol. 275, no. 5, pp. F633–F650, 1998.

[25] G. Sanchez, A. N. T. Nguyen, B. Timmerberg, J. S. Tash,and G. Blanco, “The Na,K-ATPase α4 isoform from humanshas distinct enzymatic properties and is important for spermmotility,” Molecular Human Reproduction, vol. 12, no. 9, pp.565–576, 2006.

[26] S. P. Barwe, G. Anilkumar, S. Y. Moon et al., “Novel role forNa,K-ATPase in phosphatidylinositol 3-kinase signaling andsuppression of cell motility,” Molecular Biology of the Cell, vol.16, no. 3, pp. 1082–1094, 2005.

[27] A. T. Beggah and K. Geering, “α and β subunits of Na,K-ATPase interact with BiP and calnexin,” Annals of the NewYork Academy of Sciences, vol. 834, pp. 537–539, 1997.

[28] M. S. Feschenko, R. K. Wetzel, and K. J. Sweadner,“Phosphorylation of Na,K-ATPase by protein kinases. Sites,susceptibility, and consequences,” Annals of the New YorkAcademy of Sciences, vol. 834, pp. 479–488, 1997.

[29] C. Jordan, B. Puschel, R. Koob, and D. Drenckhahn, “Iden-tification of a binding motif for ankyrin on the α-subunit ofNa+,K+-ATPase,” Journal of Biological Chemistry, vol. 270, no.50, pp. 29971–29975, 1995.

Page 54: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

8 International Journal of Nephrology

[30] K. Lee, J. Jung, M. Kim, and G. Guidotti, “Interaction of theα subunit of Na,K-ATPase with, cofilin,” Biochemical Journal,vol. 353, no. 2, pp. 377–385, 2001.

[31] J. Tian, T. Cai, Z. Yuan et al., “Binding of Src to Na+/K+-ATPase forms a functional signaling complex,” MolecularBiology of the Cell, vol. 17, no. 1, pp. 317–326, 2006.

[32] Z. Xie and T. Cai, “Na+-K+–ATPase-mediated signal trans-duction: from protein interaction to cellular function,” MolInterv, vol. 3, no. 3, pp. 157–168, 2003.

[33] G. A. Yudowski, R. Efendiev, C. H. Pedemonte, A. I. Katz,P. O. Berggren, and A. M. Bertorello, “Phosphoinositide-3kinase binds to a proline-rich motif in the Na+,K+-ATPaseα subunit and regulates its trafficking,” Proceedings of theNational Academy of Sciences of the United States of America,vol. 97, no. 12, pp. 6556–6561, 2000.

[34] Z. Zhang, P. Devarajan, A. L. Dorfman, and J. S. Mor-row, “Structure of the ankyrin-binding domain of α-Na,K-ATPase,” Journal of Biological Chemistry, vol. 273, no. 30, pp.18681–18684, 1998.

[35] H. Song, M. Y. Lee, S. P. Kinsey, D. J. Weber, and M.P. Blaustein, “An N-terminal sequence targets and tethersNa+ pump α2 subunits to specialized plasma membranemicrodomains,” Journal of Biological Chemistry, vol. 281, no.18, pp. 12929–12940, 2006.

[36] S. Zhang, S. Malmersjo, J. Li et al., “Distinct role of the N-terminal tail of the Na,K-ATPase catalytic subunit as a signaltransducer,” Journal of Biological Chemistry, vol. 281, no. 31,pp. 21954–21962, 2006.

[37] J. H. Kaplan, “A moving new role for the sodium pump inepithelial cells and carcinomas,” Science’s STKE, vol. 2005, no.289, p. pe31, 2005.

[38] J. D. Kaunitz, “Membrane transport proteins: not just fortransport anymore,” American Journal of Physiology, vol. 290,no. 5, pp. F995–F996, 2006.

[39] Z. Li and Z. Xie, “The Na/K-ATPase/Src complex and car-diotonic steroid-activated protein kinase cascades,” PflugersArchiv European Journal of Physiology, vol. 457, no. 3, pp.635–644, 2009.

[40] O. Aizman, P. Uhlen, M. Lal, H. Brismar, and A. Aperia,“Ouabain, a steroid hormone that signals with slow calciumoscillations,” Proceedings of the National Academy of Sciencesof the United States of America, vol. 98, no. 23, pp. 13420–13424, 2001.

[41] A. Aydemir-Koksoy, J. Abramowitz, and J. C. Allen,“Ouabain-induced signaling and vascular smooth muscle cellproliferation,” Journal of Biological Chemistry, vol. 276, no.49, pp. 46605–46611, 2001.

[42] J. Li, S. Zelenin, A. Aperia, and O. Aizman, “Low doses ofouabain protect from serum deprivation-triggered apoptosisand stimulate kidney cell proliferation via activation of NF-κB,” Journal of the American Society of Nephrology, vol. 17, no.7, pp. 1848–1857, 2006.

[43] R. Saunders and G. Scheiner-Bobis, “Ouabain stimulatesendothelin release and expression in human endothelial cellswithout inhibiting the sodium pump,” European Journal ofBiochemistry, vol. 271, no. 5, pp. 1054–1062, 2004.

[44] O. Kotova, L. Al-Khalili, S. Talia et al., “Cardiotonic steroidsstimulate glycogen synthesis in human skeletal muscle cellsvia a Src- and ERK1/2-dependent mechanism,” Journal ofBiological Chemistry, vol. 281, no. 29, pp. 20085–20094, 2006.

[45] S. J. Khundmiri, M. A. Metzler, M. Ameen, V. Amin,M. J. Rane, and N. A. Delamere, “Ouabain induces cellproliferation through calcium-dependent phosphorylationof Akt (protein kinase B) in opossum kidney proximal tubule

cells,” American Journal of Physiology, Cell Physiology, vol.291, no. 6, pp. C1247–C1257, 2006.

[46] L. Trevisi, B. Visentin, F. Cusinato, I. Pighin, and S. Luciani,“Antiapoptotic effect of ouabain on human umbilical veinendothelial cells,” Biochemical and Biophysical Research Com-munications, vol. 321, no. 3, pp. 716–721, 2004.

[47] P. Ferrari, M. Ferrandi, G. Valentini, P. Manunta, andG. Bianchi, “Targeting Ouabain- and Adducin-dependentmechanisms of hypertension and cardiovascular remodelingas a novel pharmacological approach,” Medical Hypotheses,vol. 68, no. 6, pp. 1307–1314, 2007.

[48] W. C. Golden and L. J. Martin, “Low-dose ouabain protectsagainst excitotoxic apoptosis and up-regulates nuclear Bcl-2in vivo,” Neuroscience, vol. 137, no. 1, pp. 133–144, 2006.

[49] X. Jiang, Y. P. Ren, and Z. R. Lu, “Ouabain induces cardiacremodeling in rats independent of blood pressure,” ActaPharmacologica Sinica, vol. 28, no. 3, pp. 344–352, 2007.

[50] M. Li, Q. Wang, and L. Guan, “Effects of ouabain onproliferation, intracellular free calcium and c-myc mRNAexpression in vascular smooth muscle cells,” Journal ofComparative Physiology B, vol. 177, no. 5, pp. 589–595, 2007.

[51] R. Skoumal, I. Szokodi, J. Aro et al., “Involvement of endoge-nous ouabain-like compound in the cardiac hypertrophicprocess in vivo,” Life Sciences, vol. 80, no. 14, pp. 1303–1310,2007.

[52] K. Taguchi, H. Kumanogoh, S. Nakamura, and S. Maekawa,“Ouabain-induced isoform-specific localization change ofthe Na+,K+-ATPase α subunit in the synaptic plasma mem-brane of rat brain,” Neuroscience Letters, vol. 413, no. 1, pp.42–45, 2007.

[53] J. C. Thundathil, M. Anzar, and M. M. Buhr, “Na+/K+ ATPaseas a signaling molecule during bovine sperm capacitation,”Biology of Reproduction, vol. 75, no. 3, pp. 308–317, 2006.

[54] I. Larre, A. Ponce, R. Fiorentino, L. Shoshani, R. G. Contr-eras, and M. Cereijido, “Contacts and cooperation betweencells depend on the hormone ouabain,” Proceedings of theNational Academy of Sciences of the United States of America,vol. 103, no. 29, pp. 10911–10916, 2006.

[55] A. N. T. Nguyen, D. P. Wallace, and G. Blanco, “Ouabainbinds with high affinity to the Na,K-ATPase in humanpolycystic kidney cells and induces extracellular signal-regulated kinase activation and cell proliferation,” Journal ofthe American Society of Nephrology, vol. 18, no. 1, pp. 46–57,2007.

[56] R. L. Dmitrieva and P. A. Doris, “Ouabain is a potentpromoter of growth and activator of ERK1/2 in ouabain-resistant rat renal epithelial cells,” Journal of BiologicalChemistry, vol. 278, no. 30, pp. 28160–28166, 2003.

[57] I. Larre, A. Lazaro, R. G. Contreras et al., “Ouabainmodulates epithelial cell tight junction,” Proceedings of theNational Academy of Sciences of the United States of America,vol. 107, no. 25, pp. 11387–11392, 2010.

[58] A. K. Rajasekaran and S. A. Rajasekaran, “Role of Na-K-ATPase in the assembly of tight junctions,” American Journalof Physiology, vol. 285, no. 3, pp. F388–F396, 2003.

[59] M. P. Blaustein, J. Zhang, L. Chen et al., “The pump, theexchanger, and endogenous ouabain: signaling mechanismsthat link salt retention to hypertension,” Hypertension, vol.53, no. 2, pp. 291–298, 2009.

[60] A. Y. Bagrov and O. V. Fedorova, “Cardenolide and bufa-dienolide ligands of the sodium pump. How they worktogether in NaCl sensitive hypertension,” Frontiers in Bio-science, vol. 10, no. 1, pp. 2250–2256, 2005.

Page 55: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 9

[61] A. M. Bertorello and J. I. Sznajder, “The dopamine paradoxin lung and kidney epithelia: sharing the same target butoperating different signaling networks,” American Journal ofRespiratory Cell and Molecular Biology, vol. 33, no. 5, pp. 432–437, 2005.

[62] J. Liu and J. I. Shapiro, “Regulation of sodium pumpendocytosis by cardiotonic steroids: molecular mechanismsand physiological implications,” Pathophysiology, vol. 14, no.3-4, pp. 171–181, 2007.

[63] Z. Xie and A. Askari, “Na+/K+-ATPase as a signal transducer,”European Journal of Biochemistry, vol. 269, no. 10, pp. 2434–2439, 2002.

[64] A. Aperia, “New roles for an old enzyme: Na,K-ATPaseemerges as an interesting drug target,” Journal of InternalMedicine, vol. 261, no. 1, pp. 44–52, 2007.

[65] R. A. Kelly and T. W. Smith, “The search for the endogenousdigitalis: an alternative hypothesis,” American Journal ofPhysiology, vol. 256, no. 5, pp. C937–C950, 1989.

[66] R. Foulkes, R. G. Ferrario, P. Salvati, and G. Bianchi,“Differences in ouabain-induced natriuresis between isolatedkidneys of Milan hypertensive and normotensive rats,”Clinical Science, vol. 82, no. 2, pp. 185–190, 1992.

[67] N. A. Yates and J. G. McDougall, “Effect of volume expansionon the natriuretic response to ouabain infusion,” RenalPhysiology and Biochemistry, vol. 18, no. 6, pp. 311–320, 1995.

[68] I. Dostanic-Larson, J. W. van Huysse, J. N. Lorenz, and J. B.Lingrel, “The highly conserved cardiac glycoside binding siteof Na,K-ATPase plays a role in blood pressure regulation,”Proceedings of the National Academy of Sciences of the UnitedStates of America, vol. 102, no. 44, pp. 15845–15850, 2005.

[69] E. L. Loreaux, B. Kaul, J. N. Lorenz, and J. B. Lingrel,“Ouabain-sensitive α1 Na,K-ATPase enhances natriureticresponse to saline load,” Journal of the American Society ofNephrology, vol. 19, no. 10, pp. 1947–1954, 2008.

[70] M. Nesher, M. Dvela, V. U. Igbokwe, H. Rosen, and D.Lichtstein, “Physiological roles of endogenous ouabain innormal rats,” American Journal of Physiology, Heart andCirculatory Physiology, vol. 297, no. 6, pp. H2026–H2034,2009.

[71] O. V. Fedorova, N. I. Kolodkin, N. I. Agalakova, E. G. Lakatta,and A. Y. Bagrov, “Marinobufagenin, an endogenous α-1 sodium pump ligand, in hypertensive Dahl salt-sensitiverats,” Hypertension, vol. 37, no. 2, pp. 462–466, 2001.

[72] O. V. Fedorova, M. I. Talan, N. I. Agalakova, E. G. Lakatta,and A. Y. Bagrov, “Endogenous ligand of α1 sodium pump,marinobufagenin, is a novel mediator of sodium chloride-dependent hypertension,” Circulation, vol. 105, no. 9, pp.1122–1127, 2002.

[73] D. E. Anderson, O. V. Fedorova, C. H. Morrell et al., “Endoge-nous sodium pump inhibitors and age-associated increases insalt sensitivity of blood pressure in normotensives,” AmericanJournal of Physiology, Regulatory Integrative and ComparativePhysiology, vol. 294, no. 4, pp. R1248–R1254, 2008.

[74] P. Manunta, P. Stella, R. Rivera et al., “Left ventricularmass, stroke volume, and ouabain-like factor in essentialhypertension,” Hypertension, vol. 34, no. 3, pp. 450–456,1999.

[75] P. Manunta, B. P. Hamilton, and J. M. Hamlyn, “Salt intakeand depletion increase circulating levels of endogenousouabain in normal men,” American Journal of Physiology,Regulatory Integrative and Comparative Physiology, vol. 290,no. 3, pp. R553–R559, 2006.

[76] B. S. Huang, M. S. Amin, and F. H. H. Leenen, “The centralrole of the brain in salt-sensitive hypertension,” CurrentOpinion in Cardiology, vol. 21, no. 4, pp. 295–304, 2006.

[77] J. W. van Huysse, “Endogenous brain Na pumps, brainouabain-like substance and the α2 isoform in salt-dependenthypertension,” Pathophysiology, vol. 14, no. 3-4, pp. 213–220,2007.

[78] O. V. Fedorova, I. A. Zhuravin, N. I. Agalakova et al.,“Intrahippocampal microinjection of an exquisitely low doseof ouabain mimics NaCl loading and stimulates a bufadieno-lide Na/K-ATPase inhibitor,” Journal of Hypertension, vol. 25,no. 9, pp. 1834–1844, 2007.

[79] O. V. Fedorova, N. I. Agalakova, M. I. Talan, E. G. Lakatta,and A. Y. Bagrov, “Brain ouabain stimulates peripheralmarinobufagenin via angiotensin II signalling in NaCl-loaded Dahl-S rats,” Journal of Hypertension, vol. 23, no. 8,pp. 1515–1523, 2005.

[80] D. J. Kennedy, S. Vetteth, S. M. Periyasamy et al., “Central rolefor the cardiotonic steroid marinobufagenin in the pathogen-esis of experimental uremic cardiomyopathy,” Hypertension,vol. 47, no. 3, pp. 488–495, 2006.

[81] M. Ferrandi, I. Molinari, P. Barassi, E. Minotti, G. Bianchi,and P. Ferrari, “Organ hypertrophic signaling within caveolaemembrane subdomains triggered by ouabain and antago-nized by PST 2238,” Journal of Biological Chemistry, vol. 279,no. 32, pp. 33306–33314, 2004.

[82] J. Liu, J. Tian, M. Haas, J. I. Shapiro, A. Askari, and Z. Xie,“Ouabain interaction with cardiac Na+/K+-ATPase initiatessignal cascades independent of changes in intracellular Na+

and Ca2+ concentrations,” Journal of Biological Chemistry,vol. 275, no. 36, pp. 27838–27844, 2000.

[83] S. Priyadarshi, B. Valentine, C. Han et al., “Effect of green teaextract on cardiac hypertrophy following 5/6 nephrectomy inthe rat,” Kidney International, vol. 63, no. 5, pp. 1785–1790,2003.

[84] J. Elkareh, D. J. Kennedy, B. Yashaswi et al., “Marinobufa-genin stimulates fibroblast collagen production and causesfibrosis in experimental uremic cardiomyopathy,” Hyperten-sion, vol. 49, no. 1, pp. 215–224, 2007.

[85] D. J. Kennedy, J. Elkareh, A. Shidyak et al., “Partial nephrec-tomy as a model for uremic cardiomyopathy in the mouse,”American Journal of Physiology, vol. 294, no. 2, pp. F450–F454, 2008.

[86] J. Elkareh, S. M. Periyasamy, A. Shidyak et al., “Marinobufa-genin induces increases in procollagen expression in a processinvolving protein kinase C and Fli-1: implications for uremiccardiomyopathy,” American Journal of Physiology, vol. 296,no. 5, pp. F1219–F1226, 2009.

[87] J. Tian, A. Shidyak, S. M. Periyasamy et al., “Spironolactoneattenuates experimental uremic cardiomyopathy by antago-nizing marinobufagenin,” Hypertension, vol. 54, no. 6, pp.1313–1320, 2009.

[88] N. El-Okdi, S. Smaili, V. Raju et al., “Effects of cardiotonicsteroids on dermal collagen synthesis and wound healing,”Journal of Applied Physiology, vol. 105, no. 1, pp. 30–36, 2008.

[89] L. V. Fedorova, V. Raju, N. El-Okdi et al., “The cardiotonicsteroid hormone marinobufagenin induces renal fibrosis:implication of epithelial-to-mesenchymal transition,” Amer-ican Journal of Physiology, vol. 296, no. 4, pp. F922–F934,2009.

[90] M. P. Blaustein, “Sodium ions, calcium ions, blood pressureregulation, and hypertension: a reassessment and a hypothe-sis,” American Journal of Physiology, Cell Physiology, vol. 232,no. 5, pp. C165–C173, 1977.

Page 56: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

10 International Journal of Nephrology

[91] F. J. Haddy, M. B. Pamnani, and D. L. Clough, “Humoral fac-tors and the sodium-potassium pump in volume expandedhypertension,” Life Sciences, vol. 24, no. 23, pp. 2105–2117,1979.

[92] H. E. de Wardener and E. M. Clarkson, “Concept ofnatriuretic hormone,” Physiological Reviews, vol. 65, no. 3,pp. 658–759, 1985.

[93] F. J. Haddy and H. W. Overbeck, “The role of humoral agentsin volume expanded hypertension,” Life Sciences, vol. 19, no.7, pp. 935–947, 1976.

[94] H. E. de Wardener and E. M. Clarkson, “Concept ofnatriuretic hormone,” Physiological Reviews, vol. 65, no. 3,pp. 658–759, 1985.

[95] J. G. McDougall and N. A. Yates, “Natriuresis and inhibitionof Na+/K+-ATPase: modulation of response by physiologicalmanipulation,” Clinical and Experimental Pharmacology andPhysiology, vol. 25, supplement, pp. S57–S60, 1998.

[96] N. A. Yates and J. G. McDougall, “Effects of direct renalarterial infusion of bufalin and ouabain in conscious sheep,”British Journal of Pharmacology, vol. 108, no. 3, pp. 627–630,1993.

[97] N. A. Yates and J. G. McDougall, “Interaction of exogenousouabain and chronic mineralocorticoid treatment in thekidney of the conscious sheep,” Clinical and ExperimentalPharmacology and Physiology, vol. 24, no. 1, pp. 57–63, 1997.

[98] H. E. de Wardener and G. A. MacGregor, “Dahl’s hypothesisthat a saluretic substance may be responsible for a sustainedrise in arterial pressure: its possible role in essential hyperten-sion,” Kidney International, vol. 18, no. 1, pp. 1–9, 1980.

[99] O. V. Fedorova, N. I. Agalakova, C. H. Morrell, E. G.Lakatta, and A. Y. Bagrov, “ANP differentially modulatesmarinobufagenin-induced sodium pump inhibition in kid-ney and aorta,” Hypertension, vol. 48, no. 6, pp. 1160–1168,2006.

[100] O. V. Fedorova, E. G. Lakatta, and A. Y. Bagrov, “EndogenousNa,K pump ligands are differentially regulated during acuteNaCl loading of dahl rats,” Circulation, vol. 102, no. 24, pp.3009–3014, 2000.

[101] W. Schoner and G. Scheiner-Bobis, “Role of endogenouscardiotonic steroids in sodium homeostasis,” NephrologyDialysis Transplantation, vol. 23, no. 9, pp. 2723–2729, 2008.

[102] M. Yoshika, Y. Komiyama, M. Konishi et al., “Novel digitalis-like factor, marinobufotoxin, isolated from cultured Y-1 cells,and its hypertensive effect in rats,” Hypertension, vol. 49, no.1, pp. 209–214, 2007.

[103] E. Ritz et al., “Uremic Cardiomyopathy-An EndogenousDigitalis Intoxication?” Journal of the American Society ofNephrology, vol. 17, no. 6, pp. 1493–1497, 2006.

[104] A. Y. Bagrov, J. I. Shapiro, and O. V. Fedorova, “Endogenouscardiotonic steroids: physiology, pharmacology, and noveltherapeutic targets,” Pharmacological Reviews, vol. 61, no. 1,pp. 9–38, 2009.

[105] F. J. Haddy and M. B. Pamnani, “Role of ouabain-like factorsand NA-K-ATPase inhibitors in hypertension—some old andrecent findings,” Clinical and Experimental Hypertension, vol.20, no. 5-6, pp. 499–508, 1998.

[106] W. Schoner and G. Scheiner-Bobis, “Endogenous and exoge-nous cardiac glycosides: their roles in hypertension, saltmetabolism, and cell growth,” American Journal of Physiology,Cell Physiology, vol. 293, no. 2, pp. C509–C536, 2007.

[107] J. Tian, S. Haller, S. Periyasamy et al., “Renal ischemiaregulates marinobufagenin release in humans,” Hypertension,vol. 56, no. 5, pp. 914–919, 2010.

[108] H. Cai, L. Wu, W. Qu et al., “Regulation of apical NHE3trafficking by ouabain-induced activation of the basolateralNa+,K+-ATPase receptor complex,” American Journal ofPhysiology, Cell Physiology, vol. 294, no. 2, pp. C555–C563,2008.

[109] J. Liu, R. Kesiry, S. M. Periyasamy, D. Malhotra, Z. Xie, and J.I. Shapiro, “Ouabain induces endocytosis of plasmalemmalNa/K-ATPase in LLC-PK1 cells by a clathrin-dependentmechanism,” Kidney International, vol. 66, no. 1, pp. 227–241, 2004.

[110] J. Liu, M. Liang, L. Liu, D. Malhotra, Z. Xie, and J. I. Shapiro,“Ouabain-induced endocytosis of the plasmalemmal Na/K-ATPase in LLC-PK1 cells requires caveolin-1,” Kidney Inter-national, vol. 67, no. 5, pp. 1844–1854, 2005.

[111] J. Liu, S. M. Periyasamy, W. Gunning et al., “Effects of cardiacglycosides on sodium pump expression and function in LLC-PK1 and MDCK cells,” Kidney International, vol. 62, no. 6,pp. 2118–2125, 2002.

[112] S. Oweis, L. Wu, P. R. Kiela et al., “Cardiac glycosidedownregulates NHE3 activity and expression in LLC-PK 1cells,” American Journal of Physiology, vol. 290, no. 5, pp.F997–F1008, 2006.

[113] S. M. Periyasamy, J. Liu, F. Tanta et al., “Salt loadinginduces redistribution of the plasmalemmal Na/K-ATPase inproximal tubule cells,” Kidney International, vol. 67, no. 5, pp.1868–1877, 2005.

[114] J. Liu, Y. Yan, L. Liu et al., “Impairment of Na/K-ATPasesignaling in renal proximal tubule contributes to Dahl salt-sensitive hypertension,” Journal of Biological Chemistry, vol.286, no. 26, pp. 22806–22813, 2011.

[115] L. K. Dahl, M. Heine, and L. Tassinari, “Role of geneticfactors in susceptibility to experimental hypertension due tochronic excess salt ingestion,” Nature, vol. 194, no. 4827, pp.480–482, 1962.

[116] L. K. Dahl, K. D. Knudsen, and J. Iwai, “Humoral transmis-sion of hypertension: evidence from parabiosis,” CirculationResearch, vol. 24, no. 5, supplement, pp. 21–33, 1969.

[117] J. P. Rapp, “Dahl salt-susceptible and salt-resistant rats. Areview,” Hypertension, vol. 4, no. 6, pp. 753–763, 1982.

[118] J. P. Rapp and H. Dene, “Development and characteristics ofinbred strains of Dahl salt-sensitive and salt-resistant rats,”Hypertension, vol. 7, no. 3, pp. 340–349, 1985.

[119] M. Mokry and E. Cuppen, “The Atp1a1 gene from inbredDahl salt sensitive rats does not contain the A1079T missensetransversion,” Hypertension, vol. 51, no. 4, pp. 922–927, 2008.

[120] Y. Yamori, “The development of Spontaneously HypertensiveRat (SHR) and of various spontaneous rat models, andtheir implications,” in Experimental and Genetic Models ofHypertension, D. Ganten and W. de Jong, Eds., vol. 4 ofHandbook of Hypertension, pp. 224–239, Elsevier, New York,NY, USA, 1984.

[121] C. R. Pace-Asciak and M. C. Carrara, “Reduction in dietaryvitamin E prevents onset of hypertension in developingspontaneously hypertensive rats,” Experientia, vol. 35, no. 12,pp. 1561–1562, 1979.

[122] J. B. Young, D. Mullen, and L. Landsberg, “Caloric restrictionlowers blood pressure in the spontaneously hypertensive rat,”Metabolism, vol. 27, no. 12, pp. 1711–1714, 1978.

[123] N. Adams and D. A. Blizard, “Genetic and maternalinfluences in rat models of spontaneous and salt-inducedhypertension,” Developmental Psychobiology, vol. 24, no. 7,pp. 507–519, 1991.

Page 57: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 11

[124] W. J. Arendshorst and W. H. Beierwaltes, “Renal tubularreabsorption in spontaneously hypertensive rats,” The Amer-ican Journal of Physiology, vol. 237, no. 1, pp. F38–47, 1979.

[125] W. H. Beierwaltes and W. J. Arendshorst, “Renal functionof conscious spontaneously hypertensive rats,” CirculationResearch, vol. 42, no. 5, pp. 721–726, 1978.

[126] W. J. Arendshorst, “Autoregulation of renal blood flow inspontaneously hypertensive rats,” Circulation Research, vol.44, no. 3, pp. 344–349, 1979.

[127] R. O. Crajoinas, L. M. A. Lessa, L. R. Carraro-Lacroix etal., “Posttranslational mechanisms associated with reducedNHE3 activity in adult vs. young prehypertensive SHR,”American Journal of Physiology, vol. 299, no. 4, pp. F872–F881, 2010.

[128] M. S. LaPointe, C. Sodhi, A. Sahai, and D. Batlle, “Na+/H+

exchange activity and NHE-3 expression in renal tubulesfrom the spontaneously hypertensive rat,” Kidney Interna-tional, vol. 62, no. 1, pp. 157–165, 2002.

[129] L. C. Garg, N. Narang, and S. McArdle, “Na-K-ATPase innephron segments of rats developing spontaneous hyperten-sion,” The American Journal of Physiology, vol. 249, no. 6, pp.F863–869, 1985.

[130] J. L. Cangiano, C. Rodriguez-Sargent, S. Opava-Stitzer, andM. Martinez-Maldonado, “Renal Na+,K+-ATPase in wean-ling and adult spontaneously hypertensive rats,” Proceedingsof the Society for Experimental Biology and Medicine, vol. 177,no. 2, pp. 240–246, 1984.

[131] L. C. Garg and N. Narang, “Differences in renal tubular Na-K-adenosine triphosphatase in spontaneously hypertensiveand normotensive rats,” Journal of Cardiovascular Pharma-cology, vol. 8, no. 1, pp. 186–189, 1986.

[132] L. C. Garg and N. Narang, “Effects of hydrochlorothiazideon Na-K-ATPase activity along the rat nephron,” KidneyInternational, vol. 31, no. 4, pp. 918–922, 1987.

[133] G. A. Morduchowicz, D. Sheikh-Hamad, O. D. Jo, E. P. Nord,D. B. N. Lee, and N. Yanagawa, “Increased Na+/H+ antiportactivity in the renal brush border membrane of SHR,” KidneyInternational, vol. 36, no. 4, pp. 576–581, 1989.

[134] A. L. Hayward, C. A. Hinojos, B. Nurowska et al., “Alteredsodium pump alpha and gamma subunit gene expressionin nephron segments from hypertensive rats,” Journal ofHypertension, vol. 17, no. 8, pp. 1081–1087, 1999.

[135] J. I. Kourie, “Interaction of reactive oxygen species with iontransport mechanisms,” American Journal of Physiology, CellPhysiology, vol. 275, no. 1, pp. C1–C24, 1998.

[136] T. Chabrashvili, A. Tojo, M. L. Onozato et al., “Expressionand cellular localization of classic NADPH oxidase subunitsin the spontaneously hypertensive rat kidney,” Hypertension,vol. 39, no. 2, pp. 269–274, 2002.

[137] C. Panico, Z. Luo, S. Damiano, F. Artigiano, P. Gill, and W.J. Welch, “Renal proximal tubular reabsorption is reducedin adult spontaneously hypertensive rats: roles of superoxideand Na+/H+ exchanger 3,” Hypertension, vol. 54, no. 6, pp.1291–1297, 2009.

[138] A. A. Banday and M. F. Lokhandwala, “Angiotensin II-mediated biphasic regulation of proximal tubular Na+/H+

exchanger 3 is impaired during oxidative stress,” AmericanJournal of Physiology, Renal Physiology, vol. 301, no. 2, pp.F364–F370, 2011.

[139] L. A. Dada, E. Novoa, E. Lecuona, H. Sun, and J. I. Sznajder,“Role of the small GTPase RhoA in the hypoxia-induceddecrease of plasma membrane Na,K-ATPase in A549 cells,”Journal of Cell Science, vol. 120, no. 13, pp. 2214–2222, 2007.

[140] L. Liu, J. Li, J. Liu et al., “Involvement of Na+/K+-ATPase in hydrogen peroxide-induced hypertrophy in car-diac myocytes,” Free Radical Biology and Medicine, vol. 41,no. 10, pp. 1548–1556, 2006.

[141] Z. Xie, P. Kometiani, J. Liu, J. Li, J. I. Shapiro, and A. Askari,“Intracellular reactive oxygen species mediate the linkage ofNa+/K+- ATPase to hypertrophy and its marker genes incardiac myocytes,” Journal of Biological Chemistry, vol. 274,no. 27, pp. 19323–19328, 1999.

[142] J. Tian, J. Liu, K. D. Garlid, J. I. Shapiro, and Z. Xie,“Involvement of mitogen-activated protein kinases and reac-tive oxygen species in the inotropic action of ouabain oncardiac myocytes. A potential role for mitochondrial KATPchannels,” Molecular and Cellular Biochemistry, vol. 242, no.1-2, pp. 181–187, 2003.

[143] A. Miyakawa-Naito, P. Uhlen, M. Lal et al., “Cell signal-ing microdomain with Na,K-ATPase and inositol 1,4,5-trisphosphate receptor generates calcium oscillations,” Jour-nal of Biological Chemistry, vol. 278, no. 50, pp. 50355–50361,2003.

[144] M. Kajikawa, S. Fujimoto, Y. Tsuura et al., “Ouabainsuppresses glucose-induced mitochondrial ATP productionand insulin release by generating reactive oxygen species inpancreatic islets,” Diabetes, vol. 51, no. 8, pp. 2522–2529,2002.

[145] Y. T. Huang, S. C. Chueh, C. M. Teng, and J. H. Guh, “Inves-tigation of ouabain-induced anticancer effect in humanandrogen-independent prostate cancer PC-3 cells,” Biochem-ical Pharmacology, vol. 67, no. 4, pp. 727–733, 2004.

[146] A. Boldyrev, E. Bulygina, M. Yuneva, and W. Schoner,“Na/K-ATPase regulates intracellular ROS level in cerebellumneurons,” Annals of the New York Academy of Sciences, vol.986, pp. 519–521, 2003.

[147] D. J. Kennedy, S. Vetteth, M. Xie et al., “Ouabain decreasessarco(endo)plasmic reticulum calcium ATPase activity in rathearts by a process involving protein oxidation,” AmericanJournal of Physiology, Heart and Circulatory Physiology, vol.291, no. 6, pp. H3003–H3011, 2006.

[148] I. M. Glynn, ““Transport adenosinetriphosphatase” in elec-tric organ. The relation between ion transport and oxidativephosphorylation,” The Journal of Physiology, vol. 169, no. 2,pp. 452–465, 1963.

[149] M. J. McKenna, I. Medved, C. A. Goodman et al., “N-acetylcysteine attenuates the decline in muscle Na+, K+-pump activity and delays fatigue during prolonged exercisein humans,” Journal of Physiology, vol. 576, no. 1, pp. 279–288, 2006.

[150] A. Boldyrev and E. Kurella, “Mechanism of oxidative damageof dog kidney Na/K-ATPase,” Biochemical and BiophysicalResearch Communications, vol. 222, no. 2, pp. 483–487, 1996.

[151] D. Dobrota, M. Matejovicova, E. G. Kurella, and A. A.Boldyrev, “Na/K-ATPase under oxidative stress: molecularmechanisms of injury,” Cellular and Molecular Neurobiology,vol. 19, no. 1, pp. 141–149, 1999.

[152] E. G. Kurella, O. V. Tyulina, and A. A. Boldyrev, “Oxidativeresistance of Na/K-ATPase,” Cellular and Molecular Neurobi-ology, vol. 19, no. 1, pp. 133–140, 1999.

[153] I. Petrushanko, N. Bogdanov, E. Bulygina et al., “Na-K-ATPase in rat cerebellar granule cells is redox sensitive,”American Journal of Physiology, Regulatory Integrative andComparative Physiology, vol. 290, no. 4, pp. R916–R925,2006.

Page 58: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

12 International Journal of Nephrology

[154] A. Bogdanova, A. Boldyrev, and M. Gassmann, “Oxygen-and redox-induced regulaton of the Na/K ATPase,” CurrentEnzyme Inhibition, vol. 2, no. 1, pp. 37–59, 2006.

[155] D. Z. Ellis, J. Rabe, and K. J. Sweadner, “Global loss ofNa,K-ATPase and its nitric oxide-mediated regulation in atransgenic mouse model of amyotrophic lateral sclerosis,”Journal of Neuroscience, vol. 23, no. 1, pp. 43–51, 2003.

[156] M. S. Kim and T. Akera, “O2 free radicals: cause of ischemia-reperfusion injury to cardiac Na+,K+-ATPase,” AmericanJournal of Physiology, Heart and Circulatory Physiology, vol.252, no. 2, pp. H252–H257, 1987.

[157] Z. Xie, Y. Wang, A. Askari, W. H. Huang, J. E. Klaunig, andA. Askari, “Studies on the specificity of the effects of oxygenmetabolites on cardiac sodium pump,” Journal of Molecularand Cellular Cardiology, vol. 22, no. 8, pp. 911–920, 1990.

[158] W. H. Huang, Y. Wang, and A. Askari, “(Na+ + K+)-ATPase:inactivation and degradation induced by oxygen radicals,”International Journal of Biochemistry, vol. 24, no. 4, pp. 621–626, 1992.

[159] W.-H. Huang, Y. Wang, A. Askari, N. Zolotarjova, andM. Ganjeizadeh, “Different sensitivities of the Na+/K+-ATPase isoforms to oxidants,” Biochimica et Biophysica Acta,Biomembranes, vol. 1190, no. 1, pp. 108–114, 1994.

[160] N. Zolotarjova, C. Ho, R. L. Mellgren, A. Askari, and W. -H.Huang, “Different sensitivities of native and oxidized formsof Na+/K+-ATPase to intracellular proteinases,” Biochimica etBiophysica Acta, Biomembranes, vol. 1192, no. 1, pp. 125–131,1994.

[161] Z. Xie, M. Jack-Hays, Y. Wang et al., “Different oxidantsensitivities of the α1 and α2 isoforms of Na+/K+-ATPaseexpressed in baculovirus-infected insect cells,” Biochemicaland Biophysical Research Communications, vol. 207, no. 1, pp.155–159, 1995.

[162] M. Mense, G. Stark, and H. J. Apell, “Effects of freeradicals on partial reactions of the Na,K-ATPase,” Journal ofMembrane Biology, vol. 156, no. 1, pp. 63–71, 1997.

[163] F. Thevenod and J. M. Friedmann, “Cadmium-mediatedoxidative stress in kidney proximal tubule cells inducesdegradation of Na+/K+-ATPase through proteasomal andendo- /lysosomal proteolytic pathways,” FASEB Journal, vol.13, no. 13, pp. 1751–1761, 1999.

[164] C. N. White, G. A. Figtree, C. C. Liu et al., “AngiotensinII inhibits the Na+-K+ pump via PKC-dependent activationof NADPH oxidase,” American Journal of Physiology, CellPhysiology, vol. 296, no. 4, pp. C693–C700, 2009.

[165] G. A. Figtree, C. C. Liu, S. Bibert et al., “Reversible oxidativemodification: a key mechanism of Na+-K+ pump regulation,”Circulation Research, vol. 105, no. 2, pp. 185–193, 2009.

[166] S. Bibert, C.-C. Liu, G. A. Figtree et al., “FXYD pro-teins reverse inhibition of the Na+-K+ pump mediated byglutathionylation of its β1 subunit,” Journal of BiologicalChemistry, vol. 286, no. 21, pp. 18562–18572, 2011.

[167] M. S. Reifenberger, K. L. Arnett, C. Gatto, and M. A.Milanick, “The reactive nitrogen species peroxynitrite is apotent inhibitor of renal Na-K-ATPase activity,” AmericanJournal of Physiology, vol. 295, no. 4, pp. F1191–F1198, 2008.

[168] C. Zhang, S. Z. Imam, S. F. Ali, and P. R. Mayeux,“Peroxynitrite and the regulation of Na+,K+-ATPase activityby angiotensin II in the rat proximal tubule,” Nitric Oxide,vol. 7, no. 1, pp. 30–35, 2002.

[169] A. Y. Zhang, F. Yi, G. Zhang, E. Gulbins, and P. L. Li, “Lipidraft clustering and redox signaling platform formation incoronary arterial endothelial cells,” Hypertension, vol. 47, no.1, pp. 74–80, 2006.

[170] L. Zuo, M. Ushio-Fukai, S. Ikeda, L. Hilenski, N. Patrushev,and R. W. Alexander, “Caveolin-1 is essential for activationof Rac1 and NAD(P)H oxidase after angiotensin II type 1receptor stimulation in vascular smooth muscle cells: rolein redox signaling and vascular hypertrophy,” Arteriosclerosis,Thrombosis, and Vascular Biology, vol. 25, no. 9, pp. 1824–1830, 2005.

[171] R. M. Touyz, “Lipid rafts take center stage in endothelial cellredox signaling by death receptors,” Hypertension, vol. 47, no.1, pp. 16–18, 2006.

[172] W. Han, H. Li, V. A. M. Villar et al., “Lipid rafts keep NADPHoxidase in the inactive state in human renal proximal tubulecells,” Hypertension, vol. 51, no. 2, pp. 481–487, 2008.

[173] P. N. Seshiah, D. S. Weber, P. Rocic, L. Valppu, Y. Taniyama,and K. K. Griendling, “Angiotensin II stimulation ofNAD(P)H oxidase activity: upstream mediators,” CirculationResearch, vol. 91, no. 5, pp. 406–413, 2002.

[174] R. M. Touyz, G. Yao, and E. L. Schiffrin, “c-Src inducesphosphorylation and translocation of p47phox: role insuperoxide generation by angiotensin II in human vascularsmooth muscle cells,” Arteriosclerosis, Thrombosis, and Vas-cular Biology, vol. 23, no. 6, pp. 981–987, 2003.

[175] T. M. Paravicini and R. M. Touyz, “Redox signaling inhypertension,” Cardiovascular Research, vol. 71, no. 2, pp.247–258, 2006.

[176] C. S. Wilcox and D. Gutterman, “Focus on oxidative stressin the cardiovascular and renal systems,” American Journal ofPhysiology, Heart and Circulatory Physiology, vol. 288, no. 1,pp. H3–H6, 2005.

[177] R. M. Touyz, “Reactive oxygen species, vascular oxidativestress, and redox signaling in hypertension: what is theclinical significance?” Hypertension, vol. 44, no. 3, pp. 248–252, 2004.

[178] F. Lacy, M. T. Kailasam, D. T. O’Connor, G. W. Schmid-Schonbein, and R. J. Parmer, “Plasma hydrogen peroxideproduction in human essential hypertension: role of heredity,gender, and ethnicity,” Hypertension, vol. 36, no. 5, pp. 878–884, 2000.

[179] F. Lacy, D. T. O’Connor, and G. W. Schmid-Schonbein,“Plasma hydrogen peroxide production in hypertensivesand normotensive subjects at genetic risk of hypertension,”Journal of Hypertension, vol. 16, no. 3, pp. 291–303, 1998.

[180] N. D. Vaziri and B. Rodrıguez-Iturbe, “Mechanisms of dis-ease: oxidative stress and inflammation in the pathogenesisof hypertension,” Nature Clinical Practice Nephrology, vol. 2,no. 10, pp. 582–593, 2006.

[181] C. S. Wilcox, “Oxidative stress and nitric oxide deficiencyin the kidney: a critical link to hypertension?” AmericanJournal of Physiology, Regulatory, Integrative and ComparativePhysiology, vol. 289, no. 4, pp. R913–R935, 2005.

[182] W. J. Welch, “Intrarenal oxygen and hypertension,” Clinicaland Experimental Pharmacology and Physiology, vol. 33, no.10, pp. 1002–1005, 2006.

[183] C. Kitiyakara, T. Chabrashvili, Y. Chen et al., “Salt intake,oxidative stress, and renal expression of NADPH oxidaseand superoxide dismutase,” Journal of the American Societyof Nephrology, vol. 14, no. 11, pp. 2775–2782, 2003.

[184] C. Schreck and P. M. O’Connor, “NAD(P)H oxidase andrenal epithelial ion transport,” American Journal of Physiol-ogy, Regulatory Integrative and Comparative Physiology, vol.300, no. 5, pp. R1023–R1029, 2011.

[185] H. J. Han, J. L. Yun, H. P. Su, H. L. Jang, and M. Taub,“High glucose-induced oxidative stress inhibits Na+/glucose

Page 59: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

International Journal of Nephrology 13

cotransporter activity in renal proximal tubule cells,” Amer-ican Journal of Physiology, vol. 288, no. 5, pp. F988–F996,2005.

[186] R. Juncos, N. J. Hong, and J. L. Garvin, “Differential effectsof superoxide on luminal and basolateral Na+/H+ exchangein the thick ascending limb,” American Journal of Physiology,Regulatory Integrative and Comparative Physiology, vol. 290,no. 1, pp. R79–R83, 2006.

[187] A. W. Cowley, T. Mori, D. Mattson, and A. P. Zou, “Role ofrenal NO production in the regulation of medullary bloodflow,” American Journal of Physiology, Regulatory Integrativeand Comparative Physiology, vol. 284, no. 6, pp. R1355–R1369, 2003.

[188] A. Makino, M. M. Skelton, A. P. Zou, and A. W. Cowley,“Increased renal medullary H2O2 leads to hypertension,”Hypertension, vol. 42, no. 1, pp. 25–30, 2003.

[189] N. E. Taylor and A. W. Cowley Jr., “Effect of renalmedullary H2O2 on salt-induced hypertension and renalinjury,” American Journal of Physiology, Regulatory Integrativeand Comparative Physiology, vol. 289, no. 6, pp. R1573–R1579, 2005.

[190] N. E. Taylor, P. Glocka, M. Liang, and A. W. Cowley,“NADPH oxidase in the renal medulla causes oxidative stressand contributes to salt-sensitive hypertension in Dahl S rats,”Hypertension, vol. 47, no. 4, pp. 692–698, 2006.

[191] C. Cao, A. Edwards, M. Sendeski et al., “Intrinsic nitric oxideand superoxide production regulates descending vasa rectacontraction,” American Journal of Physiology, vol. 299, no. 5,pp. F1056–F1064, 2010.

[192] A. Edwards, C. Cao, and T. L. Pallone, “Cellular mechanismsunderlying nitric oxide-induced vasodilation of descendingvasa recta,” American Journal of Physiology, Renal Physiology,vol. 300, no. 2, pp. F441–F456, 2011.

[193] Z. Yang, L. D. Asico, P. Yu et al., “D5 dopamine receptorregulation of reactive oxygen species production, NADPHoxidase, and blood pressure,” American Journal of Physiology,Regulatory Integrative and Comparative Physiology, vol. 290,no. 1, pp. R96–R104, 2006.

[194] L. Kopkan and D. S. Majid, “Superoxide contributes todevelopment of salt sensitivity and hypertension inducedby nitric oxide deficiency,” Hypertension., vol. 46, no. 4, pp.1026–1031, 2005.

[195] L. Kopkan, A. Hess, Z. Huskova, L. Cervenka, L. G. Navar,and D. S. A. Majid, “High-salt intake enhances superoxideactivity in enos knockout mice leading to the development ofsalt sensitivity,” American Journal of Physiology, vol. 299, no.3, pp. F656–F663, 2010.

[196] R. Zhang, P. Hording, J. L. Garvin et al., “Isoforms andfunctions of NAD(P)H oxidase at the macula densa,” Hyper-tension, vol. 53, no. 3, pp. 556–563, 2009.

[197] P. A. Ortiz and J. L. Garvin, “Superoxide stimulates NaClabsorption by the thick ascending limb,” American Journal ofPhysiology, vol. 283, no. 5, pp. F957–F962, 2002.

[198] T. Mori and A. W. Cowley, “Angiotensin II-NAD(P)Hoxidase-stimulated superoxide modifies tubulovascularnitric oxide cross-talk in renal outer medulla,” Hypertension,vol. 42, no. 4, pp. 588–593, 2003.

[199] E. J. Johns, B. O’Shaughnessy, S. O’Neill, B. Lane, andV. Healy, “Impact of elevated dietary sodium intake onNAD(P)H oxidase and SOD in the cortex and medulla ofthe rat kidney,” American Journal of Physiology, RegulatoryIntegrative and Comparative Physiology, vol. 299, no. 1, pp.R234–R240, 2010.

[200] T. Wang, “Nitric oxide regulates HCO3 and Na+ transportby a cGMP-mediated mechanism in the kidney proximaltubule,” American Journal of Physiology, Renal Physiology, vol.272, no. 2, pp. F242–F248, 1997.

[201] H. Li, W. Han, A. M. Van Villar et al., “D1-like receptorsregulate nadph oxidase activity and subunit expression inlipid raft microdomains of renal proximal tubule cells,”Hypertension, vol. 53, no. 6, pp. 1054–1061, 2009.

[202] A. A. Banday and M. F. Lokhandwala, “Loss of biphasiceffect on Na/K-ATPase activity by angiotensin II involvesdefective angiotensin type 1 receptor-nitric oxide signaling,”Hypertension, vol. 52, no. 6, pp. 1099–1105, 2008.

[203] O. W. Moe, A. Tejedor, M. Levi, D. W. Seldin, P. A.Preisig, and R. J. Alpern, “Dietary NaCl modulates Na+-H+ antiporter activity in renal cortical apical membranevesicles,” American Journal of Physiology - Renal Fluid andElectrolyte Physiology, vol. 260, no. 1, pp. F130–F137, 1991.

[204] L. E. Yang, M. B. Sandberg, A. D. Can, K. Pihakaski-Maunsbach, and A. A. McDonough, “Effects of dietary salton renal Na+ transporter subcellular distribution, abun-dance, and phosphorylation status,” American Journal ofPhysiology, vol. 295, no. 4, pp. F1003–F1016, 2008.

[205] J. Liu, Y. Yan, L. Liu et al., “Impairment of Na/K-ATPasesignaling in renal proximal tubule contributes to Dahl salt-sensitive hypertension,” Journal of Biological Chemistry, vol.286, no. 26, pp. 22806–22813, 2011.

[206] K. A. Fisher, S. H. Lee, J. Walker, C. Dileto-Fang, L.Ginsberg, and S. R. Stapleton, “Regulation of proximaltubule sodium/hydrogen antiporter with chronic volumecontraction,” American Journal of Physiology, vol. 280, no. 5,pp. F922–F926, 2001.

[207] E. Silva and P. Soares-da-Silva, “Reactive oxygen species andthe regulation of renal Na+-K+-ATPase in opossum kidneycells,” American Journal of Physiology, Regulatory Integrativeand Comparative Physiology, vol. 293, no. 4, pp. R1764–R1770, 2007.

[208] A. A. McDonough, “Mechanisms of proximal tubule sodiumtransport regulation that link extracellular fluid volume andblood pressure,” American Journal of Physiology, RegulatoryIntegrative and Comparative Physiology, vol. 298, no. 4, pp.R851–R861, 2010.

[209] L. E. Yang, P. K. K. Leong, and A. A. McDonough, “Reducingblood pressure in SHR with enalapril provokes redistributionof NHE3, NaPi2, and NCC and decreases NaPi2 and ACEabundance,” American Journal of Physiology, vol. 293, no. 4,pp. F1197–F1208, 2007.

[210] T. L. Pallone, “Control of renal Na+ excretion by hemeoxygenase,” Hypertension, vol. 49, no. 1, pp. 23–24, 2007.

[211] A. P. Zou, H. Billington, N. Su, and A. W. Cowley, “Expres-sion and actions of heme oxygenase in the renal medulla ofrats,” Hypertension, vol. 35, no. 1, pp. 342–347, 2000.

[212] F. Rodriguez, F. Zhang, S. Dinocca, and A. Nasjletti, “Nitricoxide synthesis influences the renal vascular response toheme oxygenase inhibition,” American Journal of Physiology,vol. 284, no. 6, pp. F1255–F1262, 2003.

[213] B. Arregui, B. Lopez, M. G. Salom, F. Valero, C. Navarro, andF. J. Fenoy, “Acute renal hemodynamic effects of dimanganesedecacarbonyl and cobalt protoporphyrin,” Kidney Interna-tional, vol. 65, no. 2, pp. 564–574, 2004.

[214] R. A. Johnson, M. Lavesa, B. Askari, N. G. Abraham,and A. Nasjletti, “A heme oxygenase product, presumablycarbon monoxide, mediates a vasodepressor function inrats,” Hypertension, vol. 25, no. 2, pp. 166–169, 1995.

Page 60: Devil’s Triangle in Kidney Diseases: Oxidative Stress ...downloads.hindawi.com/journals/specialissues/352592.pdfoxidative stress, mediators, and inflammation are like a devil’s

14 International Journal of Nephrology

[215] H. Hirakawa and Y. Hayashida, “Autonomic cardiovascularresponses to heme oxygenase inhibition in conscious rats,”Hypertension, vol. 48, no. 6, pp. 1124–1129, 2006.

[216] R. A. Johnson, M. Lavesa, K. Deseyn, M. J. Scholer, and A.Nasjletti, “Heme oxygenase substrates acutely lower bloodpressure in hypertensive rats,” American Journal of Physiology,vol. 271, no. 3, pp. H1132–H1138, 1996.

[217] R. D. Levere, P. Martasek, B. Escalante, M. L. Schwartzman,and N. G. Abraham, “Effect of heme arginate administrationon blood pressure in spontaneously hypertensive rats,”Journal of Clinical Investigation, vol. 86, no. 1, pp. 213–219,1990.

[218] D. Sacerdoti, B. Escalante, N. G. Abraham, J. C. McGiff,R. D. Levere, and M. L. Schwartzman, “Treatment with tinprevents the development of hypertension in spontaneouslyhypertensive rats,” Science, vol. 243, no. 4889, pp. 388–390,1989.

[219] H. E. Sabaawy, F. Zhang, X. Nguyen et al., “Humanheme oxygenase-1 gene transfer lowers blood pressureand promotes growth in spontaneously hypertensive rats,”Hypertension, vol. 38, no. 2, pp. 210–215, 2001.

[220] R. Wang, R. Shamloul, X. Wang, Q. Meng, and L. Wu,“Sustained normalization of high blood pressure in spon-taneously hypertensive rats by implanted hemin pump,”Hypertension, vol. 48, no. 4, pp. 685–692, 2006.

[221] F. K. Johnson, W. Durante, K. J. Peyton, and R. A. Johnson,“Heme oxygenase inhibitor restores arteriolar nitric oxidefunction in dahl rats,” Hypertension, vol. 41, no. 1, pp. 149–155, 2003.

[222] F. J. Teran, R. A. Johnson, B. K. Stevenson et al., “Hemeoxygenase-derived carbon monoxide promotes arteriolarendothelial dysfunction and contributes to salt-inducedhypertension in Dahl salt-sensitive rats,” American Journalof Physiology, Regulatory Integrative and Comparative Physi-ology, vol. 288, no. 3, pp. R615–R622, 2005.

[223] R. A. Johnson, F. J. Teran, W. Durante, K. J. Peyton, and F.K. Johnson, “Enhanced heme oxygenase-mediated coronaryvasodilation in Dahl salt-sensitive hypertension,” AmericanJournal of Hypertension, vol. 17, no. 1, pp. 25–30, 2004.

[224] P. Katavetin, R. Inagi, T. Miyata et al., “Erythropoietininduces heme oxygenase-1 expression and attenuates oxida-tive stress,” Biochemical and Biophysical Research Communi-cations, vol. 359, no. 4, pp. 928–934, 2007.

[225] T. Munzel, T. Gori, R. M. Bruno, and S. Taddei, “Is oxidativestress a therapeutic target in cardiovascular disease?” Euro-pean Heart Journal, vol. 31, no. 22, pp. 2741–2748, 2010.

[226] L. J. Appel, T. J. Moore, E. Obarzanek et al., “A clinical trialof the effects of dietary patterns on blood pressure,” NewEngland Journal of Medicine, vol. 336, no. 16, pp. 1117–1124,1997.

[227] E. R. Miller III, L. J. Appel, and T. H. Risby, “Effect of dietarypatterns on measures of lipid peroxidation: results from arandomized clinical trial,” Circulation, vol. 98, no. 22, pp.2390–2395, 1998.

[228] J. H. John, S. Ziebland, P. Yudkin, L. S. Roe, and H. A.W. Neil, “Effects of fruit and vegetable consumption onplasma antioxidant concentrations and blood pressure: arandomised controlled trial,” Lancet, vol. 359, no. 9322, pp.1969–1974, 2002.

[229] R. Collins, J. Armitage, S. Parish, P. Sleight, and R. Peto,“MRC/BHF Heart Protection Study of antioxidant vitaminsupplementation in 20 536 high-risk individuals: a ran-domised placebo-controlled trial,” Lancet, vol. 360, no. 9326,pp. 23–33, 2002.

[230] H. F. Galley, J. Thornton, P. D. Howdle, B. E. Walker, and N.R. Webster, “Combination oral antioxidant supplementationreduces blood pressure,” Clinical Science, vol. 92, no. 4, pp.361–365, 1997.

[231] A. Ceriello, D. Giugliano, A. Quatraro, and P. J. Lefebvre,“Anti-oxidants show an anti-hypertensive effect in diabeticand hypertensive subjects,” Clinical Science, vol. 81, no. 6, pp.739–742, 1991.

[232] A. Ceriello, E. Motz, and D. Giugliano, “Hypertension andascorbic acid,” Lancet, vol. 355, no. 9211, pp. 1272–1273,2000.

[233] H.-Y. Huang, B. Caballero, S. Chang et al., “The efficacyand safety of multivitamin and mineral supplement use toprevent cancer and chronic disease in adults: a systematicreview for a National Institutes of Health state-of-the-scienceconference,” Annals of Internal Medicine, vol. 145, no. 5, pp.372–385, 2006.

[234] J. L. Garvin and P. A. Ortiz, “The role of reactive oxygenspecies in the regulation of tubular function,” Acta Physio-logica Scandinavica, vol. 179, no. 3, pp. 225–232, 2003.