Skip to main content
Molecular Medicine logoLink to Molecular Medicine
. 2008 Apr 23;14(7-8):502–516. doi: 10.2119/2008-00006.Legrand

Renal Hypoxia and Dysoxia After Reperfusion of the Ischemic Kidney

Matthieu Legrand 1,2,, Egbert G Mik 3, Tanja Johannes 4, Didier Payen 2, Can Ince 1
PMCID: PMC2386087  PMID: 18488066

Abstract

Ischemia is the most common cause of acute renal failure. Ischemic-induced renal tissue hypoxia is thought to be a major component in the development of acute renal failure in promoting the initial tubular damage. Renal oxygenation originates from a balance between oxygen supply and consumption. Recent investigations have provided new insights into alterations in oxygenation pathways in the ischemic kidney. These findings have identified a central role of microvascular dysfunction related to an imbalance between vasoconstrictors and vasodilators, endothelial damage and endothelium–leukocyte interactions, leading to decreased renal oxygen supply. Reduced microcirculatory oxygen supply may be associated with altered cellular oxygen consumption (dysoxia), because of mitochondrial dysfunction and activity of alternative oxygen-consuming pathways. Alterations in oxygen utilization and/or supply might therefore contribute to the occurrence of organ dysfunction. This view places oxygen pathways’ alterations as a potential central player in the pathogenesis of acute kidney injury. Both in regulation of oxygen supply and consumption, nitric oxide seems to play a pivotal role. Furthermore, recent studies suggest that, following acute ischemic renal injury, persistent tissue hypoxia contributes to the development of chronic renal dysfunction. Adaptative mechanisms to renal hypoxia may be ineffective in more severe cases and lead to the development of chronic renal failure following ischemia-reperfusion. This paper is aimed at reviewing the current insights into oxygen transport pathways, from oxygen supply to oxygen consumption in the kidney and from the adaptation mechanisms to renal hypoxia. Their role in the development of ischemia-induced renal damage and ischemic acute renal failure are discussed.

INTRODUCTION

Acute renal failure (ARF) is a common condition that develops in 5% of hospitalized patients. Of the patients who develop ARF, ~10% eventually require renal replacement therapy (1). Among critical care patients who have acute renal failure and survive, 2% to 10% develop terminal renal failure and require long-term dialysis (2,3). There is clear evidence that the occurrence of acute renal failure is associated with excess early and late mortality, and also with high financial costs (2,4,5). The kidneys are particularly susceptible to ischemic injury in many clinical conditions such as renal transplantation (6), treatment of suprarenal aneurysms (7), renal artery reconstructions, contrast agent–induced nephropathy (8), cardiac arrest, and shock. One reason for renal sensitivity to ischemia is that the kidney microvasculature is highly complex and must meet a high energy demand. Under normal, steady-state conditions, the oxygen (O2) supply to the renal tissues is well in excess of oxygen demand. Under pathological conditions, however, the delicate balance of oxygen supply compared with demand is easily disturbed owing to the unique arrangement of the renal microvasculature and its increasing numbers of diffusive shunting pathways (9,10). The renal microvasculature is serially organized, with almost all descending vasa recta emerging from the efferent arterioles of the juxtamedullary glomeruli. Adequate tissue oxygenation thus partially depends on the maintenance of medullary perfusion by adequate cortical perfusion. This, combined with the low amount of medullary blood flow [~10% of total renal blood flow (11)] in the U-shaped microvasculature of the medulla, allows O2 shunting between the descending and ascending vasa recta and contributes to the high sensitivity of the medulla and corticomedullary junction to decreased O2 supply (1215). The effects of limited O2 supply are aggravated by the high O2 demand associated with the high tubular O2 consumption necessary for solute exchange (16) and the high rate of aerobic glycolysis (17). It is these conditions, working in parallel, that make the kidney highly susceptible to hypoxic conditions (18,19).

Whereas past investigations have focused mainly on tubular injury as the main cause of ischemia-related acute renal failure, increasing evidence implicates alterations in the intrarenal microcirculation pathway (2022) and O2 handling (2325). Indeed, although acute tubular necrosis (ATN) has classically been believed to be the leading cause of ARF (6,26), data from biopsies in patients with ATN have shown few or no changes consistent with tubular necrosis (27). The role played by microvascular dysfunction, however, has generated increasing interest. The complex pathophysiology of ischemic ARF includes the inevitable reperfusion phase associated with oxidative stress (28,29), cellular dysfunction, and altered signal transduction (30). During this process, alterations in oxygen transport pathways can result in cellular hypoxia and/or dysoxia. In this context, the distinction between hypoxia and dysoxia is that cellular hypoxia refers to the condition of decreased availability of oxygen due to inadequate convective delivery from the microcirculation. Cellular dysoxia, in contrast, refers to a pathological condition where the ability of mitochondria to perform oxidative phosphorylation is limited, regardless of the amount of available oxygen. The latter condition is associated with mitochondrial failure and/or activation of alternative pathways for oxygen consumption (31). Thus, we would expect that an optimal balance between oxygen supply and demand is essential to reducing damage from renal ischemia-reperfusion (I/R) injury (Figure 1). As discussed below, many factors can lead to a reduced oxygen supply at the microvascular level (32), including endothelial damage and leukocyte plugging. The negative effects of reduced oxygen supply on tissue oxygen levels may be aggravated by altered cellular oxygen consumption (dysoxia), due possibly to mitochondrial dysfunction and activity of alternative oxygen-consuming pathways (factors are discussed below).

Figure 1.

Figure 1

Synthetic representation of mechanisms involved in renal tissue hypoxia leading to acute and chronic renal failure after ischemia-reperfusion. Renal tissue hypoxia reflects an imbalance between oxygen supply and oxygen consumption. Renal tissue hypoxia participates to further tubular damage and acute renal failure. Recently the role of renal hypoxia has been proposed in development of chronic renal failure. HIF, hypoxia-inducible factor; –, protect; +, promote.

Factors that disrupt this balance can be expected to contribute to the pathogenesis of ARF. This article provides an overview of how renal oxygenation pathways are affected by I/R in the kidney and how these processes affect organ failure. We will review the contribution of factors related to the pathogenesis of ischemic acute renal failure that involve imbalances between renal O2 supply and demand. We will also discuss the adaptive mechanisms that respond to acute renal ischemia and how imbalances in renal oxygenation imbalance contribute to the development of ischemia-induced acute renal failure. We will conclude with how our current understanding of the pathogenesis of ARF may affect future therapeutic strategies.

THE NO-REFLOW PHENOMENON JEOPARDIZES RENAL O2 SUPPLY

Reperfusion after ischemic injury results in a partial recovery of renal blood flow, with a reduction in renal blood flow of up to 50% after reperfusion (33). This persistent reduction in total and regional renal blood flow (3436) was described several decades ago as the “no-reflow” phenomenon (37). Recent work using intravital microscopy techniques has provided new insights into the central role of microvascular alterations and tubular damage in ischemic kidney injury leading to renal dysfunction (21). Many factors have been proposed to explain this phenomenon, such as an imbalance between vasodilatory and vasoconstrictive mediators, endothelial congestion, increased endothelial permeability leading to interstitial edema that compresses the peritubular capillaries, increased leukocyte adherence, and the extravascular accumulation of leukocytes. Persistent reductions in renal tissue oxygenation due to reperfusion have been directly measured by O2 electrodes and phosphorescence measurements in animal models of reperfusion injury and sepsis (38,39). The persistence of depressed regional blood flow and consequent renal tissue oxygenation is thought to play a pivotal role in the occurrence of acute renal failure after an ischemic injury, especially in the corticomedullary junction and the outer medulla (9).

Endothelial dysfunction plays a central role in jeopardizing renal O2 delivery at the microcirculatory level. In the next section, we discuss the role of the endothelium in the development of (micro)vascular dysfunction and its contribution to ischemia-induced renal injury.

MICROVASCULAR DYSFUNCTION AND THE BALANCE BETWEEN VASOCONSTRICTORS AND VASODILATORS

Most of the descending vasa recta providing blood to the medulla arise from efferent arterioles of juxtacorti-comedullary nephrons (13). Vascular smooth muscle cells (including pericytes) around these descending vasa recta have vasoconstrictive properties (40). Thus tonic modulation of afferent and efferent arterioles and the descending vasa recta contributes to the regulation of renal O2 supply and medullary blood flow, respectively. The endothelium plays a key role in flow regulation, permeability, cell trafficking, signaling, and immunologic functions (32). In this context, endothelial dysfunction is defined as impaired vasorelaxation in response to endothelium-dependent vasodilators, such as acetylcholine (ACh). Such endothelial dysfunction has been shown to participate in postischemic vasoconstriction following reperfusion (41). Although the mechanism behind this dysfunction remains unclear and various contributing factors have been suggested, the nitric oxide (NO) pathway is thought to be a main player (42,43). NO is well known for its role in vascular tone regulation by acting on vascular smooth muscle cells to induce vasodilation (40,44). NO is generated by the enzymatic transformation of L-arginine and O2 into L-citrulline and NO; this reaction is catalyzed by an enzyme called nitric oxide synthase (NOS), which exists in three different isoforms, all of which are found in the kidney (42,45). Whereas endothelial NOS (eNOS) is found in the vasa recta, inner medullary collecting duct, and glomeruli (46,47), inducible NOS (iNOS) can be expressed by vascular smooth muscle cells, renal tubular cells (39), and immune cells such as monocytes, macrophages, and neutrophils (42,4850). In the kidney, nNOS is expressed in cells of the macula densa and seems to participate in tubuloglomerular feedback, but has minimal effect on medullary perfusion (51,52), in contrast to eNOS (53). NOS inhibitors have been shown to induce hypoxic medullary damage in models of renal ischemia, highlighting the role of NO in maintaining regional renal blood flow under these conditions (54). Despite having a common end product (NO), activating different types of NOS can have very different effects on renal tissue, depending on both the temporal profile and topography of NO production by each NOS enzyme (42). iNOS-derived NO appears to participate in vascular dysfunction (55,56), and it is therefore thought that NO generated by iNOS is harmful, leading to tissue damage (43,55,57). Several in vivo and in vitro investigations have demonstrated that inhibiting the expression or activity of iNOS (58,59) or even the absence of iNOS itself (60) can prevent renal I/R injury. The two main pathways of iNOS-derived NO-induced damage are thought to be the generation of peroxynitrite, causing endothelial dysfunction (60,61), and the inhibition of eNOS-derived NO, resulting in the inability to buffer vasoconstrictive agents (45,58). Conversely, infusing nonselective NOS inhibitors that block eNOS activity has been shown to worsen both cortical and medullary blood flow and oxygenation (45,62). Indeed, eNOS-derived NO may act in a paracrine fashion, affecting surrounding cells and preventing vascular dysfunction by inhibiting platelet aggregation and preventing leukocyte activation (63), as well as having a direct vasodilatory effect (54). The inability of hypoxic endothelial cells to produce NO could thus have detrimental effects and participate in renal damage (64).

To highlight the role played by endothelial cell dysfunction after renal ischemia, Brodsky et al. (65) intravenously injected human umbilical vein endothelial cells (HUVECs) into a rat kidney transiently exposed to ischemia. HUVEC injection improved blood flow in peritubular capillaries as well as renal function. In the same study, implanting cells expressing functional eNOS also improved renal function, suggesting a pivotal role for the dysfunction of both endothelial cells and eNOS in postischemic renal injury. Incomplete restoration of endothelium-dependent vasodilation, however, suggests that other pathways could be involved in this dysfunction, such as prostaglandins or endothelium-derived hyperpolarizing factor (EDHF) (66,67). Furthermore, during I/R, a lack of tetrahydrobiopterin (BH4)—an essential cofactor of NOS—has also been implicated in endothelial dysfunction after I/R injury (56). Deficits in BH4 lead to eNOS uncoupling and the oxidation of O2, which causes the enzyme to produce superoxides and generate peroxynitrite instead of NO (68). The administration of BH4 before I/R has been shown to protect the kidney from damage following I/R (69).

O2 pressure can also interfere with NO bioavailability (55). For example, NO is generated in response to decreased arterial O2 partial pressure, thereby causing vasodilation and an increase in capillary recruitment and blood flow (7072). In addition, medullary vasoconstriction induced by indomethacin and radiological contrast media causes hypoxia, and has been shown to increase NO availability (72). Hypoxia can increase NO release from red blood cells, possibly through modifying the allosteric and electronic properties of hemoglobin, such as by decreasing the amount of the S-nitrosylated form of hemoglobin (73). Another mechanism responsible for hypoxia-induced NO release by red blood cells occurs through the nitrites (74).

In addition to decreased vasodilation proprieties, arterioles can exhibit increased reactivity to vasoconstrictive agents (75). Many vasoconstrictive agents have been suggested to accompany the postischemic increase in vascular tone, including endothelin, angiotensin II (76), thromboxane A2 (77), leukotriene C4 (78), adenosine (14), endothelium-derived prostaglandin H2, and sympathetic nerve stimulation (13,18). These abnormalities in vascular reactivity have been associated with increases in cytosolic Ca2+ (79); such increases in intracellular Ca2+ may also cause cell death by inducing the activation of proteases, phospholipases, and pro-apoptotic pathways (80). Of all the vasoconstrictive agents, endothelin seems to be particularly important, because endothelin-1 levels have been found to increase after ischemia, and blockage of the endothelin pathway by endothelin A receptor antagonists (70,76) has been shown to protect against I/R injury. NO can counteract the effects of endothelin after I/R (81). It seems, however, that further blockage of endothelin B (ETB) receptors can be detrimental to the kidney by interfering with the recovery process (82). Eicosanoids are another class of mediators that exert complex effects on microcirculatory function after I/R. They are produced from arachidonic acid through three major enzymatic pathways: cyclooxygenase (COX), lipoxygenase (LO), and cytochrome P450 monooxygenase (83). It has long been recognized that prostanoids (made by COX-1 and -2) play a critical role in modulating renal blood flow and glomerular hemodynamics (84,85). Besides regulating the vascular tone of the glomerular afferent arterioles, prostanoids with vasodilatory properties, such as prostaglandin E2 and prostacyclin (PGI2), can oppose the effects of vasoconstrictors (catecholamines, angiotensin, and vasopressin) in hypovolemic conditions or conditions of decreased cardiac output (8690). Thus, these mediators can influence medullary blood flow by controlling the vascular tone of the descending vasa recta, which likely requires the activation of the prostanoid receptors EP2, EP4, or IP (91). Leukotriene A4–derived eicosanoids (from the LO pathway) have been shown to act on the microcirculation by enhancing vasoconstriction and vascular permeability (via LTC4 and LTD4), and by mediating the local inflammatory response to increase polymorphonuclear leukocyte aggregation and adhesion to the endothelium (LTB4) (9294). Hydroxyeicosatetraenoic acids (HETEs) and epoxyeicosatrienoic acids (EETs) from the P450 monooxygenase pathway can constrict renal vessels and glomerular mesangial cells and have been implicated in the tubuloglomerular feedback and renal blood flow autoregulation of the kidney (95,96). Further details on the complex renal pathophysiology of eicosanoids can be found in a recent dedicated review (83).

ENDOTHELIAL INJURY

Structural damage to both endothelial and vascular smooth muscle cells occurs rapidly after an ischemic insult (97). Reorganization of F-actin in endothelial smooth cells occurs after renal I/R (98). Cytoskeletal modifications observed in vitro in endothelial cells undergoing ATP depletion (99,100) or exposure to oxidants (such as H2O2) may be responsible for increased endothelial permeability and interstitial edema (101103). Furthermore, fluorescence microscopy studies have demonstrated that microvascular flow can be impaired in peritubular capillaries owing to endothelial cell rupture or aggregation, causing intravascular obstruction after renal I/R (32,41,104,105).

Sutton et al. (20) suggested another possible mechanism involved in increasing endothelial permeability. They observed a change in endothelial cell–cell junctions characterized by a rapid loss of adherens junctions (V-E cadherin) following I/R in an in vivo rat model. The disruption of adherens junctions can increase paracellular permeability and leakage from the vascular bed to the surrounding tissue (106). Endothelial cell swelling could also participate in the perfusion impairment by externally compressing capillaries and shrinking the vascular lumen (107,108), in addition to the interstitial edema (97). Indeed, endothelial cell swelling is a classic histological observation following I/R (109). Although the role of ion channel impairment in ischemia-induced ARF remains a matter of debate, such impairments have been suggested to take part in cellular volume dysregulation (110). I/R can induce membrane depolarization of endothelial cells by inactivating flow-sensitive K+ channels and hyperpolarization or depolarization through the inhibition of inwardly or outwardly rectifying K+ currents, respectively (111,112). How such changes cause volume dysregulation, however, remains to be determined. Furthermore, regional endothelial injury may activate endothelial cells in remote organs, such as the lungs, leading to acute lung injury, which further contributes to renal damage via systemic hypoxia (113). In addition, the injury of endothelial cells may induce a procoagulative response and thereby contribute to vascular obstruction and deficits in renal perfusion (114,115).

Our understanding of the exact pathways involved in renal ischemia and endothelial cell dysfunction following I/R remains largely incomplete and needs further investigation. It should be emphasized that even if it is often described as an organ, the endothelium cannot be seen as a homogeneous structure. Indeed, endothelial cells differ in structure and function at different sites of the vascular tree and demonstrate diverse responses to hypoxic stimuli (116). Because of this endothelial heterogeneity and the broad range of tissue O2 pressures throughout the kidney, endothelial cells in different regions of the kidney will likely show different levels of susceptibility to ischemic or hypoxic injury (117,118).

ACTIVATION OF INFLAMMATION AND ENDOTHELIAL–LEUKOCYTE INTERACTIONS

Endothelial–leukocyte interactions have classically been associated with renal I/R injury (119). Although polymorphonuclear neutrophils (PMNs) have been implicated for many years in organ damage during I/R (120), research has also implicated monocytes, macrophages, and T cells (121,122). Previous studies assessed PMN activation/infiltration using enzymatic biomarkers, such as myeloperoxidase or chloroacetate esterase. Nevertheless, these assays are not specific for PMN activation (123). Using more direct staining methods, only a few infiltrating PMNs have been observed in the renal parenchyma after I/R, whereas the number of monocytes or macrophages seems to be much higher and occurs much earlier in the time course of reperfusion (121,124). This occurred notably in the outer medulla, which is known to be sensitive to ischemic injury (124). Additionally, some studies performed with PMN-depleted animals failed to show a protective effect on renal I/R injury compared with control animals (125,126). Similarly, ischemic acute renal failure has been reported in neutropenic patients, arguing against a key role for PMNs in postischemic renal damage (127). Therefore, the concept of a more prominent role for mononuclear leukocytes in this setting has recently emerged. For example, Takada et al. (128) showed that renal function could be improved by inhibiting the infiltration of mononuclear leukocytes via blockade of the B7-CD28 co-stimulation pathway, one of the costimulatory pathways necessary for T-cell activation (129). The B7-1 pathway was further implicated in human ischemic renal tissue by indirect immunohistochemical staining of endothelial cells from the ascending vasa recta in the outer medulla (130). Moreover, CD4/CD8 double-knockout mice exhibited decreased renal dysfunction following renal ischemia, with the CD4+ subtype appearing to be the major contributor to I/R injury (131). Nevertheless, the exact mechanism by which mononuclear leukocytes are activated in this setting remains largely unknown, especially given the absence of any defined antigen. Several detailed reviews can be found on this topic (121,123,127,132).

The activation of the inflammatory cascade can contribute to microcirculatory dysfunction by capillary plugging caused by the interaction between leukocytes, platelets, and red blood cells. Such plugging results in the so-called noreflow phenomenon occurring after I/R (133). The adhesion of leukocytes to the vascular endothelium following upregulation of adhesion molecules may occur via a cytokine-release pathway (134). Modulation of the inflammatory response has been shown to improve renal function after I/R. For instance, the anti-inflammatory agent, α-melanocyte stimulating hormone (α-MSH), protects the kidney against ischemic injury but also inhibits neutrophil chemotaxis (135). Similarly, infusion of a tumor necrosis factor (TNF)-α binding protein decreased neutrophil infiltration and preserved renal function, suggesting a deleterious role for the up-regulation of TNF-α in renal I/R (136). In addition, levels of the pro-inflammatory cytokines IL-6 and IL-8 in plasma have been found to predict mortality in patients with acute renal failure (137).

DNA repair processes such as poly(ADP-ribose) polymerase (PARP) activation may aggravate endothelial dysfunction and tubular damage (138,139). PARP-1 activation has been implicated in endothelial injury, as illustrated by finding a reduced number of adherent and rolling leukocyte as well as red blood cell trapping in PARP-1−/− mice after I/R (140). This mechanism is thought to involve inhibition of the PARP-1–induced inflammatory cascade. PARP-1 is involved in the regulation of inflammatory genes (141), such as nuclear factor (NF)-κB, which are implicated in leukocyte activation (134). Furthermore, overactivation of PARP-1 could lead to cell death by the activation of pro-apoptotic pathways or energy failure following the cellular depletion of its substrate, NAD+ (142).

Leukocyte activation may lead to further endothelial damage by sustaining the inflammatory response, releasing cytokines and proteases and inducing oxidative stress (32). Thus, targeting adhesion molecules may help prevent is-chemic damage to the kidney. Intercellular cell adhesion molecule type A (ICAM-1) deficiency or blockage has been demonstrated to provide protection from moderate renal I/R injury (120,143). Systemic moderate hypoxia alone can induce an increase in leukocyte adhesion mediated by CD11/CD18 integrins, even in the absence of complete ischemia (133,144). L-, E-, and P-selectins, expressed on the surface of leukocytes, endothelial cells, and platelets, respectively, are involved in the infiltration of leukocytes via their interaction with the endothelium by inducing rolling and sticking of leukocytes to the endothelium (145,146). Inhibition of P-selectin has been found both to improve renal blood flow in the early stage of ischemic renal injury and to confer renoprotection (145,147). Interestingly, if the infiltration of T cells occurs within the first hours after ischemia, monocyte/macrophage and T-cell infiltration occurs days following the injury (123,148).

Complex interactions exist between tubular injury, microvascular injury, and inflammation after renal I/R. On the one hand, insults to the tubule cells promote the liberation of a number of inflammatory mediators, such as TNF-α, interleukin (IL)-6, TGF-β, and chemotactic cytokines (RANTES, monocyte chemotactic protein-1, ENA-78, Gro-α, and IL-8) (132,148,149). On the other hand, chemokine production can promote leukocyte–endothelium interactions and leukocyte activation, resulting in renal blood flow impairment and the expansion of tubular damage.

In summary, renal ischemia triggers the activation of multiple mechanisms leading to (micro)vascular dysfunction. These mechanisms include functional endothelial dysfunction, probably involving NO as the central mediator. Endothelial dysfunction is correlated with direct endothelial injury and structural damage associated with increased vascular permeability, tissue congestion, vasomotor disorders, and inflammatory and hemostatic activation (see Figure 2). Together, these alterations compromise global and local renal perfusion, and decrease renal oxygen supply for extended periods of time after the initial ischemic insult has been effectively resolved.

Figure 2.

Figure 2

Interaction between microvascular injury and tubular damage before (A) and after (B) renal ischemia-reperfusion. PMN, polymorphonuclear.

CHANGES IN O2 CONSUMPTION AND UTILIZATION

Renal O2 Consumption Depends on Renal Hemodynamics and Electrolyte Reabsorption

Adequate medullary tissue oxygenation, in terms of balanced oxygen supply and demand, depends on the maintenance of medullary perfusion by adequate cortical perfusion and also on the high rate of O2 consumption required for active electrolyte transport (13). Furthermore, renal blood flow is closely associated with renal sodium transport [the primary O2-consuming function of the kidney (16)] by limiting/increasing solute delivery to the medullary nephrons, thus reducing/increasing metabolic demands (150).

As mentioned above, in addition to having a limited O2 supply due to the anatomy of the microcirculation, the sensitivity of the medulla to hypoxic conditions results from this high O2 consumption. Renal sodium transport is the main O2-consuming function of the kidney and is closely linked to renal blood flow for sodium transport, particularly in the thick ascending limbs of the loop of Henle and the S3 segments of the proximal tubules (19). Medullary renal blood flow is also highly dependent on cortical perfusion, with almost all descending vasa recta emerging from the efferent arteriole of juxtamedullary glomeruli (151). Thus, factors involved in the glomerular filtration rate strongly influence medullary oxygenation by modifying the amount of solute delivered to the tubules and local O2 consumption. Therefore, a slight decrease in renal perfusion pressure leads to decreased Na+ reabsorptive work and a subsequent increase in medullary pO2 due to a reduction in tubular O2 consumption (150). However, a profound reduction in cortical perfusion can disrupt medullary blood flow and lead to an imbalance between O2 supply and O2 consumption, if the decline in regional blood flow exceeds the decrease in tubular metabolic rate. Similarly, inhibition of tubular reabsorption by diuretics increases medullary pO2 by decreasing the activity of Na+/K+-ATPases and local O2 consumption (152). If cortical and medullary regional renal blood flow is interdependent, however, differential regulation and region-specific responses to vasoconstrictive and vasodilatory agents must exist (13). We previously showed that μpO2 in both the cortex and the outer medulla can vary independently, based on differences in O2 supply and consumption (38). This balance/imbalance in O2 renal consumption and delivery, and the components that affect this balance are illustrated in Figure 1. Dysoxia induced by I/R, in combination with impaired oxygen delivery due to microvascular dysfunction, may lead to the further deterioration of oxygen availability at the cellular level.

Mitochondrial Activity and NO-Mediated O2 Consumption

The medulla has been found to be the main site of production of NO in the kidney. In addition to the actions described above, NO appears to be a key regulator of renal tubule cell metabolism by inhibiting the activity of the Na+-K+-2Cl cotransporter and reducing Na+/H+ ex-change (153). Because superoxide (O2) is required to inhibit solute transport activity (154), it was assumed that these effects were mediated by peroxynitrite (OONO). Indeed, mitochondrial nNOS upregulation, together with an increase in NO production, has been shown to increase mitochondrial peroxynitrite generation, which in turn can induce cytochrome c release and promote apoptosis (155). NO has also been shown to directly compete with O2 at the mitochondrial level (156). These findings support the idea that NO acts as an endogenous regulator to match O2 supply to O2 consumption (157), especially in the renal medulla (158).

NO can bind cytochrome oxidase and cause reversible inhibition of mitochondrial respiration at nanomolar concentrations (159). Cytochrome c oxidase (cytochrome aa3, complex IV) is the terminal complex of the mitochondrial respiratory chain, which is responsible for about 90% of O2 consumption in mammals (23). NO reversibly binds to the O2 binding site of cytochrome oxidase and acts as a potent, rapid, and reversible inhibitor of cytochrome oxidase in competition with molecular O2 (160,161). This inhibition could be dependent on O2 level, because the IC50 (the concentration of NO that reduces the specified response by half) decreases with reduction in O2 concentration (156,162). The inhibition of electron flux at the cytochrome oxidase level switches the electron transport chain to a reduced state, and consequently leads to depolarization of the mitochondrial membrane potential and electron leakage. Together, these changes lead to the generation of superoxide anions (O2−•), especially by complexes I and III (156). Moncada and Erusalimsky (156) proposed that NO-induced cytochrome oxidase inhibition has a biphasic effect. A transient inhibition of cytochrome oxidase by NO can decrease the generation of O2−•, leading to potentially beneficial effects on hypoxic adaptation mechanisms. However, progression to a later phase presents a harmful response, generating a large amount of free O2 radicals. The reduction of superoxide dismutase (which converts O2−• into hydrogen peroxide [H2O2]) and the depletion of glutathione (the substrate of glutathione peroxidase, which converts H2O2 into water) result in the availability of even more O2−• that can react with NO to generate peroxynitrite (163). In addition, reactive O2 species are produced by infiltrating leukocytes, further disrupting the balance between oxidants and antioxidants. This leads to irreversible damage to the mitochondrial structures, causing downstream activation of apoptotic and other cell death pathways (25) (see Figure 3). Furthermore, reactive oxygen species (ROS) production can further impair arteriolar blood flow by inducing endothelial cell damage, leukocyte adhesion (164), and glycocalyx damage (165).

Figure 3.

Figure 3

Influence of NO and O2 on mitochondria activity. The mitochondrial electron transport chain includes four multiple subunit enzyme complexes of proteins (complexes I–IV), ATP synthase (complex V), and the adenine nucleotide translocator (ANT). Electrons are collected at the complexes I (NADH dehydrogenase) and II (succinate dehydrogenase) after oxidation of nicotinamide adenine dinucleotide (NADH) and of succinate, respectively, is transferred to coenzyme Q (CoQ) to give reduced CoQ. Electrons from reduced CoQ are transferred first to complex III (cytochrome bc1), then to cytochrome c (cyt c) and followed by complex IV (cytochrome c oxidase, COX), and finally to oxygen (O2) to give water (H2O). An electrochemical gradient is generated between the mitochondrial matrix and the intermembrane space by pumping protons through the inner membrane using the energy generated by the electron flow. This electrochemical gradient is then used to reintroduce protons through the complex V (ATP synthase) and to convert inorganic phosphate and ADP to ATP. The ATP, generated in mitochondria, is then exchanged for cytosolic ADP by the ANT. With increase in NO/O2 ratio, NO competitively inhibits complex IV associated with an electron leakage of the complexes chain and finally favors the generation of superoxide anions (O2−•). Hydrogen peroxide (H2O2) generated from O2−• by superoxide dismutase (SOD) contributes to hypoxia adaptation response via stabilization of HIF-α. If prolonged production of O2−• overwhelms capacity of the mitochondrial’s antioxidant system including the superoxide dismutase (MnSOD) and glutathione peroxidase (GSHpx), formation of large amount of peroxynitrite (ONOO–) and H2O2, precursor of the hydroxyl radical (OH•), via the Fenton reaction occurs. Peroxynitrite further inhibits the mitochondrial Mn-superoxide dismutase by nitration. Cytosolic (xanthine oxydase and cytochrome P450 reductases, and NOS, myeloperoxidase) and membrane (NADPH oxidase, the cytochrome P450) intramitochondrial ROS overproduction leads to cell necrosis or apoptosis.

To summarize, whereas the NO/O2 ratio can act as a regulator of cellular O2 consumption by matching decreases in O2 delivery to decreases in cellular O2 cellular (166), the inhibitory effect of NO on mitochondrial respiration under hypoxic conditions further impairs cellular aerobic metabolism. This leads to a state of “cytopathic hypoxia,” as described in the sepsis literature (167). Several lines of evidence exist demonstrating impairments in mitochondria function following renal I/R (161). For example, Coremans et al. (25) used a NADH fluorescence technique to show that I/R directly affects mitochondrial activity in an isolated kidney. One area of current investigation focuses on unraveling the role of the site of NO production on the modulation of mitochondrial respiration. Indeed, the site of NO production seems to be of major importance. Stimulation of eNOS by bradykinin increases the inhibition of mitochondrial respiration, supporting the idea of a regulatory role for eNOS (168). Furthermore, Clementi et al. (169) found that only cell-secreted NO competes with O2 and to regulate mitochondrial respiration. In addition to the three isoforms mentioned above, an α-isoform of neuronal NOS, the mitochondrial isoform (mNOS) located in the inner mitochondrial membrane, has also been shown to regulate mitochondrial respiration (61). These data support a role for NO in the balanced regulation of renal O2 supply and O2 consumption after renal I/R. However, the relationships between the determinants of O2 supply, O2 consumption, and renal function, and their relation to renal damage, remain largely unknown. Future research should aim to elucidate the mechanisms that lead from decreased O2 supply to changes in O2 consumption, and to determine how adaptive or damaging their respective contributions are in the setting of renal I/R. Our group recently developed a method for measuring mitochondrial pO2 in living cells in vitro by the excitation of endogenous mitochondrial protoporphyrin IX and measurement of the decay of delayed fluorescence (170). If this method proves to be feasible in vivo, it will allow insight into the balance between oxygen supply and oxygen utilization at the intracellular level.

MILD AND LONG-TERM RENAL RESPONSE TO HYPOXIA

Sustained Endothelial Activation and Vascular Rarefaction

Ischemic renal failure leads to persistent endothelial activation, mainly in the form of endothelium–leukocyte interactions and the activation of adhesion molecules. This persistent activation can compromise renal blood flow, prevent the recovery of adequate tissue oxygenation, and jeopardize tubular cell survival despite the initial recovery of renal tubular function (22,171). A 30% to 50% reduction in microvascular density was seen 40 weeks after renal ischemic injury in a rat model (104). Vascular rarefaction has been proposed to induce chronic hypoxia, resulting in tubulointerstitial fibrosis via the molecular activation of fibrogenic factors such as transforming growth factor (TGF)- β, collagen, and fibronectin (172), all of which may play an important role in the progression of chronic renal disease. Responses to hypoxia may represent an important element that prevents the onset of chronic renal failure, postrecovery. This possibility has been suggested by studies in which therapeutics that improve renal oxygenation or capillary density were shown to prevent delayed renal dysfunction after I/R. Likewise, Basile et al. (173) have shown that treatment with L-arginine, starting 3 d after renal I/R, increased renal blood flow, decreased pimonidizole staining (a hypoxia-sensitive marker), and attenuated renal damage. Kang et al. (174) used VEGF administration to stimulate angiogenesis that had been impaired in a remnant kidney model. VEGF helped maintain capillary density, reduced renal interstitial fibrosis, and stabilized renal function, although tissue oxygenation was not measured in that study. Impairment of the regenerative capacity of endothelial cells, possibly due to low activity of the telomere-synthesizing enzyme, telomerase, may be involved in postischemic vascular rarefaction (175). Progenitor cells have been suggested to be involved in endothelial repair because inhibiting the apoptosis of progenitor cells has been shown to reduce the loss of peritubular capillaries and restore kidney function in an adriamycin-induced nephrosis model (176178). However, whether deficiencies in progenitor cell-mediated endothelial repair plays a role in delayed vascular rarefaction remains to be determined (179).

Rarefaction of peritubular capillaries and hypoxia precede the development of tubulointerstitial fibrosis in several animal models, supporting a link between decreased peritubular capillary density, renal hypoxia, and progression to chronic renal failure after acute injury. Additional studies are necessary to support this hypothesis, however.

Adaptation to Hypoxia

Oxygen deprivation is an important determinant of cellular function. Over the last decade, the role of hypoxia-inducible factors (HIFs) in O2 supply and adaptation to hypoxic conditions has found increasing support. HIFs are O2-sensitive transcription factors involved in O2-dependent gene regulation that mediate cellular adaptation to O2 deprivation and tissue protection under hypoxic conditions in the kidney (180,181).

HIF is a heterodimer composed of an O2-regulated α-subunit (HIF-1 α or HIF-2α) and a constitutive β-subunit. In the presence of O2, two critical prolines of the HIF-α are hydroxylated by prolyl hydroxylases, which require molecular O2 as a substrate (182). The hydroxylated prolyl residues are recognized by the von Hippel–Lindau protein as a component of a ubiquitin ligase, which targets HIF α for destruction through the ubiquitin-proteasome pathway (183). Under hypoxic conditions, HIF is not degraded, but instead accumulates in the cell and induces transcription of target genes after binding to a hypoxia-response element (HRE) following translocation into the nucleus (184). ROS generated by mitochondria seem to be required for HIF- α stabilization and thus contribute an O2-sensoring function for the mitochondria (166,185).

The effect of NO on HIF stabilization is a matter of debate, with conflicting published results. NO generation can promote HIF-1α accumulation in a cGMP-independent manner (186). However, Hagen et al. (187) showed that NO may reduce the activation of HIF in hypoxia via the inhibitory effect of NO on cytochrome oxidase. Therefore, it seems that NO has pleiotropic effects on HIF expression, with various responses related to different pathways.

HIF-1α upregulates a number of factors implicated in cytoprotection, including angiogenic growth factors—such as vascular endothelial growth factor (VEGF), endothelial progenitor cell recruitment via the endothelial expression of SDF-1, hemeoxygenase 1 (HO-1), and erythropoietin (EPO)—and vasomotor regulation (184). VEGF is known to act on the microcirculation through various mechanisms. VEGF can stimulate endothelial cell proliferation and differentiation, increase vascular permeability, and mediate endothelium-dependent vasodilation (188). Furthermore, VEGF promotes monocyte chemotaxis and the expression of adhesion molecules. Although expression of VEGF has been shown to increase in renal diseases states or during hypoxia (188), Kanellis et al. (189) did not find any increase in VEGF expression in a rat model of renal I/R, but did observe upregulation of VEGF receptor 2 (VEGFR-2). Although it has been speculated that stimulation of VEGFR-2 by tubular cells could play a role in preventing the apoptosis of adjacent endothelial cells, this question remains to be addressed. VEGF administration has also been shown to accelerate the repair of injured capillaries, probably through nitric oxide generation, in an experimental model of thrombotic microangiopathy (190). HO-1 produces carbon monoxide (a potent vasodilator) while degrading heme, which may preserve tissue blood flow during reperfusion (191). Thus, it has been suggested that the induction of HO-1 can protect the kidney from ischemic damage by decreasing oxidative damage and NO generation (192,193). Finally, in addition to its anti-apoptotic properties, EPO may restore renal microcirculation, and thus protect the kidney from ischemic damage (176), by stimulating the mobilization and differentiation of progenitor cells toward an endothelial phenotype (194) and inducing NO release from eNOS (195).

HIF may also be involved in potentially deleterious responses to hypoxia, including the promotion of fibrosis. Metabolic regulation by HIF has been found to increase the expression of glucose transporter 1 and glycolytic enzymes, leading to a faster depletion of glucose resources (196). Although the precise role of HIF in hypoxia-induced cell death has yet to be determined, HIF has also been implicated in regulating apoptotic and cell death pathways. A detailed description of HIF pathway activation is beyond the scope of this review, but excellent reviews on this topic have been published (182,197).

Inter- and intracell type differences exist in the ability of cells to induce HIF activity in response to systemic hypoxia (184). In the kidney, HIF-1α is expressed in tubular cells, whereas HIF-2 α is expressed in peritubular cells, renal interstitial fibroblasts, and endothelial cells. Differences in HIF expression among tubular cell types have been proposed to underlie their wide disparity in vulnerability to hypoxic injury. Rosenberger et al. (180) showed that cells from the collecting duct and medullary thick ascending limbs (mTALs) have markedly different HIF responses, which were correlated with their viability during hypoxic stress and inversely correlated with tissue damage. The authors showed that HIF expression does not increase continuously with increasing severity of cellular hypoxia, and they proposed that above a certain hypoxia threshold, responses to HIF expression fail and injury may become irreversible. Therefore, the induction of the HIF response may be effective only under conditions of slight to mild hypoxic injury and ineffective in more severe cases. Further investigations will be necessary to assess potentially beneficial therapeutic strategies that target the HIF pathway.

THERAPEUTIC IMPLICATIONS AND PERSPECTIVES

Correcting intravascular volume depletion with fluid administration and maintaining adequate renal perfusion pressure with vasopressors in vasoplegic states are both first-line strategies to preserve renal oxygen supply. There is a definite need for fluid administration during the early stages of sepsis (198,199), in dehydration states (200), or for preventing contrast-induced nephropathy (8,201), however, the volume and choice of fluid remains a matter of debate. The main goal of fluid infusion is to reestablish adequate circulation and ultimately to restore tissue perfusion and oxygenation.

Correction of systemic hemodynamics does not necessarily result in the recovery of microcirculatory perfusion and oxygenation (202). Several lines of evidence show a difference in microcirculatory oxygenation and renal oxygen consumption between different fluid regimens (38). High-molecular-weight hydroxyethyl starch solutions (≥200 kDa) appear to be deleterious in terms of renal function (203,204); low-molecular-weight starch solutions, on the other hand, may improve renal tissue oxygenation by decreasing blood viscosity (205) and reducing vascular resistance, along with preventing starch molecule accumulation in the renal tubules and the disturbance of glomerular hemodynamics (206). Likewise, because low doses of dopamine are ineffective in terms of prevention or treatment of acute renal failure (201,207), norepinephrine (208) and vasopressin (209) are probably the drugs of choice for improving function in the renal macro-circulation and microcirculation. Because reaching supraphysiological hemodynamic values are at least doubtful (210) and may be harmful (211,212), the optimal fluid volume that should be given remains difficult to estimate. Indeed, based on the oxygen supply/consumption concept, maneuvers that systematically enhance the glomerular filtration rate in patients with acute renal failure may be harmful in increasing solute delivery to the tubules, and thereby the renal oxygen consumption (67). Thurau and Boylan (213) used the provocative term “acute renal success” to concisely describe the adaptive role of glomerular hemodynamics in matching oxygen supply to consumption with the ultimate goal of preventing tubular damage.

Use of pharmacological interventions that act at the microcirculatory level may be a successful strategy to overcome ischemia-induced vascular damage and prevent ARF. Activated protein C (APC), an endogenous vitamin K–dependent serine protease with multiple biological activities, may meet these criteria (207). Along with antithrombotic and profibrinolytic properties, APC can reduce the chemotaxis and interactions of leukocytes with activated endothelium (214). However, although APC has been shown to improve survival in septic shock and to shorten the time to resolution of cardiovascular and respiratory dysfunction (215), renal dysfunction was not improved in the largest study published so far (215). Nevertheless, the criteria used for renal dysfunction in that study (that is, oliguria) are questionable, because oliguria cannot be considered a fully accurate marker of kidney damage. Furthermore, animal studies have suggested a protective effect of APC in sepsis-induced kidney injury (216,217).

Drugs with pleiotropic effects on the vasculature, such as EPO (194) and statins (218), have the potential to prevent ischemia-induced renal failure by enhancing endothelial NO synthase expression and/or improving vascular permeability, leading to improved microcirculatory function. Based on the mechanisms mentioned above, the use of immunosuppressive drugs and approaches to inhibit leukocyte adhesion warrant further investigation, and may yield promising renoprotective effects (219,220).

Finally, despite being an attractive pathophysiological model, neither drugs with renal vasodilatory effects (that is, dopamine, fenoldopam, endothelin receptor blockers, adenosine antagonists) nor agents that decrease renal oxygen consumption (that is, loop diuretics) have been shown to protect the kidney from ischemic damage (200,201). Recent common standards for the diagnosis and classification of acute kidney injury (221), and new tools for early diagnosis of renal injury (222), may reduce the striking gap between the plethora of drugs found to be efficient in animal models and their almost constant failure to improve outcome in the clinical setting.

We have to bear in mind that a magic bullet to treat the highly complex condition that is renal I/R is not in sight yet, and further research is needed. We can expect that understanding the balance between O2 delivery and O2 consumption, as well as the function of O2-consuming pathways (that is, mitochondrial function, ROS generation), will be central to this treatment strategy.

CONCLUSION

Growing evidence indicates that microvascular dysfunction occurs after acute renal ischemic injury, mainly through an imbalance between vasoconstrictors and vasodilators, with NO being the main mediator. These deleterious effects of NO are thought to be associated with the NO generated by the induction of iNOS and its contribution to oxidative stress, both resulting in vascular dysfunction and tissue damage. Ischemic injury also leads to structural damage to the endothelium and leukocyte infiltration. Consequently, renal tissue hypoxia is proposed to promote the initial tubular damage, leading to acute organ dysfunction. Furthermore, recent studies suggest that, after acute ischemic renal injury, persistent tissue hypoxia contributes to the development of chronic renal dysfunction. In addition to a decrease in renal O2 supply, O2-consuming pathways are affected over the course of renal I/R. Responses to the inhibition of the mitochondrial electron transport chain during I/R may contribute to either hypoxia adaptation or cell death, depending on the length and magnitude of the process. Research is necessary to clarify the respective contributions of O2 supply (for example, to determine the precise pO2 threshold that leads to kidney damage) and O2 utilization changes to renal (dys)function in the context of I/R. These studies should aim to unravel the mechanisms leading from a decrease in O2 supply to changes in O2 consumption and determine the extent to which these responses are adaptive or damaging in the setting of renal I/R. The results of these studies will hopefully lead to the development of new treatment strategies to protect the kidney from I/R injury.

ACKNOWLEDGMENTS

The authors acknowledge Richard Milstein from Skylab Industry for help with the illustration in Figure 2.

Footnotes

Online address: http://www.molmed.org

REFERENCES

  • 1.Brady HR, Singer GG. Acute renal failure. Lancet. 1995;346:1533–40. doi: 10.1016/s0140-6736(95)92057-9. [DOI] [PubMed] [Google Scholar]
  • 2.Bagshaw SM. The long-term outcome after acute renal failure. Curr Opin Crit Care. 2006;12:561–6. doi: 10.1097/01.ccx.0000247445.71381.72. [DOI] [PubMed] [Google Scholar]
  • 3.Korkeila M, Ruokonen E, Takala J. Costs of care, long-term prognosis and quality of life in patients requiring renal replacement therapy during intensive care. Intensive Care Med. 2000;26:1824–31. doi: 10.1007/s001340000726. [DOI] [PubMed] [Google Scholar]
  • 4.Manns B, et al. Cost of acute renal failure requiring dialysis in the intensive care unit: clinical and resource implications of renal recovery. Crit Care Med. 2003;31:449–55. doi: 10.1097/01.CCM.0000045182.90302.B3. [DOI] [PubMed] [Google Scholar]
  • 5.Bell M, Martling CR. Long-term outcome after intensive care: can we protect the kidney? Crit Care. 2007;11:147. doi: 10.1186/cc5959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Hoste EA, Kellum JA. Acute kidney injury: epidemiology and diagnostic criteria. Curr Opin Crit Care. 2006;12:531–7. doi: 10.1097/MCC.0b013e3280102af7. [DOI] [PubMed] [Google Scholar]
  • 7.Ellenberger C, et al. Incidence, risk factors and prognosis of changes in serum creatinine early after aortic abdominal surgery. Intensive Care Med. 2006;32:1808–16. doi: 10.1007/s00134-006-0308-1. [DOI] [PubMed] [Google Scholar]
  • 8.Wong GT, Irwin MG. Contrast-induced nephropathy. Br J Anaesth. 2007;99:474–83. doi: 10.1093/bja/aem237. [DOI] [PubMed] [Google Scholar]
  • 9.Eckardt KU, et al. Role of hypoxia in the pathogenesis of renal disease. Kidney Int Suppl. 2005:S46–51. doi: 10.1111/j.1523-1755.2005.09909.x. [DOI] [PubMed] [Google Scholar]
  • 10.Leong CL, Anderson WP, O’Connor PM, Evans RG. Evidence that renal arterial-venous oxygen shunting contributes to dynamic regulation of renal oxygenation. Am J Physiol Renal Physiol. 2007;292:F1726–33. doi: 10.1152/ajprenal.00436.2006. [DOI] [PubMed] [Google Scholar]
  • 11.Janssen WM, Beekhuis H, de Bruin R, de Jong PE, de Zeeuw D. Noninvasive measurement of intrarenal blood flow distribution: kinetic model of renal 123I-hippuran handling. Am J Physiol. 1995;269:F571–80. doi: 10.1152/ajprenal.1995.269.4.F571. [DOI] [PubMed] [Google Scholar]
  • 12.Pallone TL, Robertson CR, Jamison RL. Renal medullary microcirculation. Physiol Rev. 1990;70:885–920. doi: 10.1152/physrev.1990.70.3.885. [DOI] [PubMed] [Google Scholar]
  • 13.Pallone TL, Silldorff EP, Turner MR. Intrarenal blood flow: microvascular anatomy and the regulation of medullary perfusion. Clin Exp Pharmacol Physiol. 1998;25:383–92. doi: 10.1111/j.1440-1681.1998.tb02220.x. [DOI] [PubMed] [Google Scholar]
  • 14.Joo JD, et al. Acute and delayed renal protection against renal ischemia and reperfusion injury with A1 adenosine receptors. Am J Physiol Renal Physiol. 2007;293:F1847–57. doi: 10.1152/ajprenal.00336.2007. [DOI] [PubMed] [Google Scholar]
  • 15.Whitehouse T, Stotz M, Taylor V, Stidwill R, Singer M. Tissue oxygen and hemodynamics in renal medulla, cortex, and corticomedullary junction during hemorrhage-reperfusion. Am J Physiol Renal Physiol. 2006;291:F647–53. doi: 10.1152/ajprenal.00475.2005. [DOI] [PubMed] [Google Scholar]
  • 16.Rosen S, Epstein FH, Brezis M. Determinants of intrarenal oxygenation: factors in acute renal failure. Ren Fail. 1992;14:321–5. doi: 10.3109/08860229209106636. [DOI] [PubMed] [Google Scholar]
  • 17.Cohen JJ. Is the function of the renal papilla coupled exclusively to an anaerobic pattern of metabolism? Am J Physiol. 1979;236:F423–33. doi: 10.1152/ajprenal.1979.236.5.F423. [DOI] [PubMed] [Google Scholar]
  • 18.Pallone TL, Zhang Z, Rhinehart K. Physiology of the renal medullary microcirculation. Am J Physiol Renal Physiol. 2003;284:F253–66. doi: 10.1152/ajprenal.00304.2002. [DOI] [PubMed] [Google Scholar]
  • 19.Brezis M, Rosen S. Hypoxia of the renal medulla: its implications for disease. N Engl J Med. 1995;332:647–55. doi: 10.1056/NEJM199503093321006. [DOI] [PubMed] [Google Scholar]
  • 20.Sutton TA, Mang HE, Campos SB, Sandoval RM, Yoder MC, Molitoris BA. Injury of the renal microvascular endothelium alters barrier function after ischemia. Am J Physiol Renal Physiol. 2003;285:F191–8. doi: 10.1152/ajprenal.00042.2003. [DOI] [PubMed] [Google Scholar]
  • 21.Yamamoto T, et al. Intravital videomicroscopy of peritubular capillaries in renal ischemia. Am J Physiol Renal Physiol. 2002;282:F1150–5. doi: 10.1152/ajprenal.00310.2001. [DOI] [PubMed] [Google Scholar]
  • 22.Horbelt M, et al. Acute and chronic microvascular alterations in a mouse model of ischemic acute kidney injury. Am J Physiol Renal Physiol. 2007;293:F688–95. doi: 10.1152/ajprenal.00452.2006. [DOI] [PubMed] [Google Scholar]
  • 23.Jassem W, Fuggle SV, Rela M, Koo DD, Heaton ND. The role of mitochondria in ischemia/reperfusion injury. Transplantation. 2002;73:493–9. doi: 10.1097/00007890-200202270-00001. [DOI] [PubMed] [Google Scholar]
  • 24.Pan Y, et al. Multiple factors affecting cellular redox status and energy metabolism modulate hypoxia-inducible factor prolyl hydroxylase activity in vivo and in vitro. Mol Cell Biol. 2007;27:912–25. doi: 10.1128/MCB.01223-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Coremans JM, Van Aken M, Naus DC, Van Velthuysen ML, Bruining HA, Puppels GJ. Pretransplantation assessment of renal viability with NADH fluorimetry. Kidney Int. 2000;57:671–83. doi: 10.1046/j.1523-1755.2000.00889.x. [DOI] [PubMed] [Google Scholar]
  • 26.Lameire N, Van Biesen W, Vanholder R. Acute renal failure. Lancet. 2005;365:417–30. doi: 10.1016/S0140-6736(05)17831-3. [DOI] [PubMed] [Google Scholar]
  • 27.Rosen S, Heyman SN. Difficulties in understanding human “acute tubular necrosis”: limited data and flawed animal models. Kidney Int. 2001;60:1220–4. doi: 10.1046/j.1523-1755.2001.00930.x. [DOI] [PubMed] [Google Scholar]
  • 28.Schnackenberg CG. Physiological and pathophysiological roles of oxygen radicals in the renal microvasculature. Am J Physiol Regul Integr Comp Physiol. 2002;282:R335–42. doi: 10.1152/ajpregu.00605.2001. [DOI] [PubMed] [Google Scholar]
  • 29.Li C, Jackson RM. Reactive species mechanisms of cellular hypoxia-reoxygenation injury. Am J Physiol Cell Physiol. 2002;282:C227–41. doi: 10.1152/ajpcell.00112.2001. [DOI] [PubMed] [Google Scholar]
  • 30.Schrier RW, Wang W, Poole B, Mitra A. Acute renal failure: definitions, diagnosis, pathogenesis, and therapy. J Clin Invest. 2004;114:5–14. doi: 10.1172/JCI22353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Connett RJ, Honig CR, Gayeski TE, Brooks GA. Defining hypoxia: a systems view of VO2, glycolysis, energetics, and intracellular PO2. J Appl Physiol. 1990;68:833–42. doi: 10.1152/jappl.1990.68.3.833. [DOI] [PubMed] [Google Scholar]
  • 32.Sutton TA, Fisher CJ, Molitoris BA. Microvascular endothelial injury and dysfunction during ischemic acute renal failure. Kidney Int. 2002;62:1539–49. doi: 10.1046/j.1523-1755.2002.00631.x. [DOI] [PubMed] [Google Scholar]
  • 33.Arendshorst WJ, Finn WF, Gottschalk CW. Pathogenesis of acute renal failure following temporary renal ischemia in the rat. Circ Res. 1975;37:558–68. doi: 10.1161/01.res.37.5.558. [DOI] [PubMed] [Google Scholar]
  • 34.Olof P, Hellberg A, Kallskog O, Wolgast M. Red cell trapping and postischemic renal blood flow: differences between the cortex, outer and inner medulla. Kidney Int. 1991;40:625–31. doi: 10.1038/ki.1991.254. [DOI] [PubMed] [Google Scholar]
  • 35.Mason J, Torhorst J, Welsch J. Role of the medullary perfusion defect in the pathogenesis of ischemic renal failure. Kidney Int. 1984;26:283–93. doi: 10.1038/ki.1984.171. [DOI] [PubMed] [Google Scholar]
  • 36.Thiel G, de Rougemont D, Kriz W, Mason J, Torhorst J, Wolgast M. The role of reduced medullary perfusion in the genesis of acute ischemic renal failure: summary of a roundtable discussion. Nephron. 1982;31:321–3. doi: 10.1159/000182674. [DOI] [PubMed] [Google Scholar]
  • 37.Summers WK, Jamison RL. The no reflow phenomenon in renal ischemia. Lab Invest. 1971;25:635–43. [PubMed] [Google Scholar]
  • 38.Johannes T, Mik EG, Nohe B, Raat NJ, Unertl KE, Ince C. Influence of fluid resuscitation on renal microvascular PO2 in a normotensive rat model of endotoxemia. Crit Care. 2006;10:R88. doi: 10.1186/cc4948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Johannes T, Mik EG, Ince C. Nonresuscitated endotoxemia induces microcirculatory hypoxic areas in the renal cortex in the rat. Shock. 2008 May 19;2008 doi: 10.1097/SHK.0b013e31817c02a5. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 40.Pallone TL, Silldorff EP. Pericyte regulation of renal medullary blood flow. Exp Nephrol. 2001;9:165–70. doi: 10.1159/000052608. [DOI] [PubMed] [Google Scholar]
  • 41.Molitoris BA, Sandoval R, Sutton TA. Endothelial injury and dysfunction in ischemic acute renal failure. Crit Care Med. 2002;30:S235–40. doi: 10.1097/00003246-200205001-00011. [DOI] [PubMed] [Google Scholar]
  • 42.Kone BC, Baylis C. Biosynthesis and homeostatic roles of nitric oxide in the normal kidney. Am J Physiol. 1997;272:F561–78. doi: 10.1152/ajprenal.1997.272.5.F561. [DOI] [PubMed] [Google Scholar]
  • 43.Guan Z, Gobe G, Willgoss D, Endre ZH. Renal endothelial dysfunction and impaired autoregulation after ischemia-reperfusion injury result from excess nitric oxide. Am J Physiol Renal Physiol. 2006;291:F619–28. doi: 10.1152/ajprenal.00302.2005. [DOI] [PubMed] [Google Scholar]
  • 44.Ito S, Abe K. Contractile properties of afferent and efferent arterioles. Clin Exp Pharmacol Physiol. 1997;24:532–5. doi: 10.1111/j.1440-1681.1997.tb01241.x. [DOI] [PubMed] [Google Scholar]
  • 45.Cowley AW, Jr, Mori T, Mattson D, Zou AP. Role of renal NO production in the regulation of medullary blood flow. Am J Physiol Regul Integr Comp Physiol. 2003;284:R1355–69. doi: 10.1152/ajpregu.00701.2002. [DOI] [PubMed] [Google Scholar]
  • 46.Wu F, Park F, Cowley AW, Jr, Mattson DL. Quantification of nitric oxide synthase activity in microdissected segments of the rat kidney. Am J Physiol. 1999;276:F874–81. doi: 10.1152/ajprenal.1999.276.6.F874. [DOI] [PubMed] [Google Scholar]
  • 47.Arima S, Ito S. Isolated juxtaglomerular apparatus as a tool for exploring glomerular hemodynamics: application of microperfusion techniques. Exp Nephrol. 2000;8:304–11. doi: 10.1159/000020683. [DOI] [PubMed] [Google Scholar]
  • 48.Kone BC. Localization and regulation of nitric oxide synthase isoforms in the kidney. Semin Nephrol. 1999;19:230–41. [PubMed] [Google Scholar]
  • 49.Buttery LD, Evans TJ, Springall DR, Carpenter A, Cohen J, Polak JM. Immunochemical localization of inducible nitric oxide synthase in endotoxin-treated rats. Lab Invest. 1994;71:755–64. [PubMed] [Google Scholar]
  • 50.Tojo A, et al. Immunocytochemical localization of distinct isoforms of nitric oxide synthase in the juxtaglomerular apparatus of normal rat kidney. J Am Soc Nephrol. 1994;4:1438–47. doi: 10.1681/ASN.V471438. [DOI] [PubMed] [Google Scholar]
  • 51.Brown R, Ollerstam A, Persson AE. Neuronal nitric oxide synthase inhibition sensitizes the tubuloglomerular feedback mechanism after volume expansion. Kidney Int. 2004;65:1349–56. doi: 10.1111/j.1523-1755.2004.00509.x. [DOI] [PubMed] [Google Scholar]
  • 52.Grzelec-Mojzesowicz M, Sadowski J. Renal tissue NO and intrarenal haemodynamics during experimental variations of NO content in anaesthetised rats. J Physiol Pharmacol. 2007;58:149–63. [PubMed] [Google Scholar]
  • 53.Ito S, Carretero OA, Abe K. Role of nitric oxide in the control of glomerular microcirculation. Clin Exp Pharmacol Physiol. 1997;24:578–81. doi: 10.1111/j.1440-1681.1997.tb02094.x. [DOI] [PubMed] [Google Scholar]
  • 54.Agmon Y, Peleg H, Greenfeld Z, Rosen S, Brezis M. Nitric oxide and prostanoids protect the renal outer medulla from radiocontrast toxicity in the rat. J Clin Invest. 1994;94:1069–75. doi: 10.1172/JCI117421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Goligorsky MS, Brodsky SV, Noiri E. NO bioavailability, endothelial dysfunction, and acute renal failure: new insights into pathophysiology. Semin Nephrol. 2004;24:316–23. doi: 10.1016/j.semnephrol.2004.04.003. [DOI] [PubMed] [Google Scholar]
  • 56.Gunnett CA, Lund DD, McDowell AK, Faraci FM, Heistad DD. Mechanisms of inducible nitric oxide synthase-mediated vascular dysfunction. Arterioscler Thromb Vasc Biol. 2005;25:1617–22. doi: 10.1161/01.ATV.0000172626.00296.ba. [DOI] [PubMed] [Google Scholar]
  • 57.Yu L, Gengaro PE, Niederberger M, Burke TJ, Schrier RW. Nitric oxide: a mediator in rat tubular hypoxia/reoxygenation injury. Proc Natl Acad Sci U S A. 1994;91:1691–5. doi: 10.1073/pnas.91.5.1691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Noiri E, Peresleni T, Miller F, Goligorsky MS. In vivo targeting of inducible NO synthase with oligodeoxynucleotides protects rat kidney against ischemia. J Clin Invest. 1996;97:2377–83. doi: 10.1172/JCI118681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Chatterjee PK, et al. Inhibition of inducible nitric oxide synthase reduces renal ischemia/reperfusion injury. Kidney Int. 2002;61:862–71. doi: 10.1046/j.1523-1755.2002.00234.x. [DOI] [PubMed] [Google Scholar]
  • 60.Wangsiripaisan A, Gengaro PE, Nemenoff RA, Ling H, Edelstein CL, Schrier RW. Effect of nitric oxide donors on renal tubular epithelial cell-matrix adhesion. Kidney Int. 1999;55:2281–8. doi: 10.1046/j.1523-1755.1999.00484.x. [DOI] [PubMed] [Google Scholar]
  • 61.Schild L, Reinheckel T, Reiser M, Horn TF, Wolf G, Augustin W. Nitric oxide produced in rat liver mitochondria causes oxidative stress and impairment of respiration after transient hypoxia. FASEB J. 2003;17:2194–201. doi: 10.1096/fj.02-1170com. [DOI] [PubMed] [Google Scholar]
  • 62.Brezis M, Heyman SN, Dinour D, Epstein FH, Rosen S. Role of nitric oxide in renal medullary oxygenation. Studies in isolated and intact rat kidneys. J Clin Invest. 1991;88:390–5. doi: 10.1172/JCI115316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Chung AW, et al. Platelet-leukocyte aggregation induced by PAR agonists: regulation by nitric oxide and matrix metalloproteinases. Br J Pharmacol. 2004;143:845–55. doi: 10.1038/sj.bjp.0705997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Phelan MW, Faller DV. Hypoxia decreases constitutive nitric oxide synthase transcript and protein in cultured endothelial cells. J Cell Physiol. 1996;167:469–76. doi: 10.1002/(SICI)1097-4652(199606)167:3<469::AID-JCP11>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
  • 65.Brodsky SV, et al. Endothelial dysfunction in ischemic acute renal failure: rescue by transplanted endothelial cells. Am J Physiol Renal Physiol. 2002;282:F1140–9. doi: 10.1152/ajprenal.00329.2001. [DOI] [PubMed] [Google Scholar]
  • 66.Lieberthal W, Wolf EF, Rennke HG, Valeri CR, Levinsky NG. Renal ischemia and reperfusion impair endothelium-dependent vascular relaxation. Am J Physiol. 1989;256:F894–900. doi: 10.1152/ajprenal.1989.256.5.F894. [DOI] [PubMed] [Google Scholar]
  • 67.Rosenberger C, Rosen S, Heyman SN. Renal parenchymal oxygenation and hypoxia adaptation in acute kidney injury. Clin Exp Pharmacol Physiol. 2006;33:980–8. doi: 10.1111/j.1440-1681.2006.04472.x. [DOI] [PubMed] [Google Scholar]
  • 68.Rabelink TJ, van Zonneveld AJ. Coupling eNOS uncoupling to the innate immune response. Arterioscler Thromb Vasc Biol. 2006;26:2585–7. doi: 10.1161/01.ATV.0000250932.24151.50. [DOI] [PubMed] [Google Scholar]
  • 69.Kakoki M, et al. Effects of tetrahydrobiopterin on endothelial dysfunction in rats with ischemic acute renal failure. J Am Soc Nephrol. 2000;11:301–9. doi: 10.1681/ASN.V112301. [DOI] [PubMed] [Google Scholar]
  • 70.Bertuglia S, Giusti A. Role of nitric oxide in capillary perfusion and oxygen delivery regulation during systemic hypoxia. Am J Physiol Heart Circ Physiol. 2005;288:H525–31. doi: 10.1152/ajpheart.00426.2004. [DOI] [PubMed] [Google Scholar]
  • 71.Crawford JH, et al. Hypoxia, red blood cells, and nitrite regulate NO-dependent hypoxic vasodilation. Blood. 2006;107:566–74. doi: 10.1182/blood-2005-07-2668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Heyman SN, Goldfarb M, Darmon D, Brezis M. Tissue oxygenation modifies nitric oxide bioavailability. Microcirculation. 1999;6:199–203. [PubMed] [Google Scholar]
  • 73.Jia L, Bonaventura C, Bonaventura J, Stamler JS. S-nitrosohaemoglobin: a dynamic activity of blood involved in vascular control. Nature. 1996;380:221–6. doi: 10.1038/380221a0. [DOI] [PubMed] [Google Scholar]
  • 74.Gladwin MT. Role of the red blood cell in nitric oxide homeostasis and hypoxic vasodilation. Adv Exp Med Biol. 2006;588:189–205. doi: 10.1007/978-0-387-34817-9_17. [DOI] [PubMed] [Google Scholar]
  • 75.Conger J, Robinette J, Villar A, Raij L, Shultz P. Increased nitric oxide synthase activity despite lack of response to endothelium-dependent vasodilators in postischemic acute renal failure in rats. J Clin Invest. 1995;96:631–8. doi: 10.1172/JCI118078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Jerkic M, et al. Relative roles of endothelin-1 and angiotensin II in experimental post-ischaemic acute renal failure. Nephrol Dial Transplant. 2004;19:83–94. doi: 10.1093/ndt/gfg521. [DOI] [PubMed] [Google Scholar]
  • 77.Masumura H, Kunitada S, Irie K, Ashida S, Abe Y. A thromboxane A2 synthase inhibitor, DP-1904, prevents rat renal injury. Eur J Pharmacol. 1991;193:321–7. doi: 10.1016/0014-2999(91)90146-h. [DOI] [PubMed] [Google Scholar]
  • 78.Klausner JM, et al. Postischemic renal injury is mediated by neutrophils and leukotrienes. Am J Physiol. 1989;256:F794–802. doi: 10.1152/ajprenal.1989.256.5.F794. [DOI] [PubMed] [Google Scholar]
  • 79.Conger JD, Robinette JB, Schrier RW. Smooth muscle calcium and endothelium-derived relaxing factor in the abnormal vascular responses of acute renal failure. J Clin Invest. 1988;82:532–7. doi: 10.1172/JCI113628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Harriman JF, Liu XL, Aleo MD, Machaca K, Schnellmann RG. Endoplasmic reticulum Ca(2+) signaling and calpains mediate renal cell death. Cell Death Differ. 2002;9:734–41. doi: 10.1038/sj.cdd.4401029. [DOI] [PubMed] [Google Scholar]
  • 81.Kurata H, et al. Protective effect of nitric oxide on ischemia/reperfusion-induced renal injury and endothelin-1 overproduction. Eur J Pharmacol. 2005;517:232–9. doi: 10.1016/j.ejphar.2005.05.026. [DOI] [PubMed] [Google Scholar]
  • 82.Forbes JM, Hewitson TD, Becker GJ, Jones CL. Simultaneous blockade of endothelin A and B receptors in ischemic acute renal failure is detrimental to long-term kidney function. Kidney Int. 2001;59:1333–41. doi: 10.1046/j.1523-1755.2001.0590041333.x. [DOI] [PubMed] [Google Scholar]
  • 83.Hao CM, Breyer MD. Physiologic and pathophysiologic roles of lipid mediators in the kidney. Kidney Int. 2007;71:1105–15. doi: 10.1038/sj.ki.5002192. [DOI] [PubMed] [Google Scholar]
  • 84.DiBona GF. Prostaglandins and nonsteroidal anti-inflammatory drugs: effects on renal hemodynamics. Am J Med. 1986;80:12–21. doi: 10.1016/0002-9343(86)90928-9. [DOI] [PubMed] [Google Scholar]
  • 85.Catella-Lawson F, et al. Effects of specific inhibition of cyclooxygenase-2 on sodium balance, hemodynamics, and vasoactive eicosanoids. J Pharmacol Exp Ther. 1999;289:735–41. [PubMed] [Google Scholar]
  • 86.Llinas MT, Lopez R, Rodriguez F, Roig F, Salazar FJ. Role of COX-2-derived metabolites in regulation of the renal hemodynamic response to norepinephrine. Am J Physiol Renal Physiol. 2001;281:F975–82. doi: 10.1152/ajprenal.2001.281.5.F975. [DOI] [PubMed] [Google Scholar]
  • 87.Rodriguez F, Llinas MT, Gonzalez JD, Rivera J, Salazar FJ. Renal changes induced by a cyclooxygenase-2 inhibitor during normal and low sodium intake. Hypertension. 2000;36:276–81. doi: 10.1161/01.hyp.36.2.276. [DOI] [PubMed] [Google Scholar]
  • 88.Ling H, et al. Effect of hypoxia on proximal tubules isolated from nitric oxide synthase knockout mice. Kidney Int. 1998;53:1642–6. doi: 10.1046/j.1523-1755.1998.00913.x. [DOI] [PubMed] [Google Scholar]
  • 89.Farge D, De la Coussaye JE, Beloucif S, Fratacci MD, Payen DM. Interactions between hemodynamic and hormonal modifications during PEEP-induced antidiuresis and antinatriuresis. Chest. 1995;107:1095–100. doi: 10.1378/chest.107.4.1095. [DOI] [PubMed] [Google Scholar]
  • 90.Payen DM, et al. Hemodynamic, gas exchange, and hormonal consequences of LBPP during PEEP ventilation. J Appl Physiol. 1987;62:61–70. doi: 10.1152/jappl.1987.62.1.61. [DOI] [PubMed] [Google Scholar]
  • 91.Imig JD. Eicosanoids and renal vascular function in diseases. Clin Sci (Lond) 2006;111:21–34. doi: 10.1042/CS20050251. [DOI] [PubMed] [Google Scholar]
  • 92.Lewis RA, Austen KF, Soberman RJ. Leukotrienes and other products of the 5-lipoxygenase pathway. Biochemistry and relation to pathobiology in human diseases. N Engl J Med. 1990;323:645–55. doi: 10.1056/NEJM199009063231006. [DOI] [PubMed] [Google Scholar]
  • 93.Pace-Asciak CR, Asotra S. Biosynthesis, catabolism, and biological properties of HPETEs, hydroperoxide derivatives of arachidonic acid. Free Radic Biol Med. 1989;7:409–33. doi: 10.1016/0891-5849(89)90125-1. [DOI] [PubMed] [Google Scholar]
  • 94.Samuelsson B, Dahlen SE, Lindgren JA, Rouzer CA, Serhan CN. Leukotrienes and lipoxins: structures, biosynthesis, and biological effects. Science. 1987;237:1171–6. doi: 10.1126/science.2820055. [DOI] [PubMed] [Google Scholar]
  • 95.Yiu SS, Zhao X, Inscho EW, Imig JD. 12-Hydroxyeicosatetraenoic acid participates in angiotensin II afferent arteriolar vasoconstriction by activating L-type calcium channels. J Lipid Res. 2003;44:2391–9. doi: 10.1194/jlr.M300183-JLR200. [DOI] [PubMed] [Google Scholar]
  • 96.Katoh T, Takahashi K, DeBoer DK, Serhan CN, Badr KF. Renal hemodynamic actions of lipoxins in rats: a comparative physiological study. Am J Physiol. 1992;263:F436–42. doi: 10.1152/ajprenal.1992.263.3.F436. [DOI] [PubMed] [Google Scholar]
  • 97.Kwon O, et al. Backleak, tight junctions, and cell-cell adhesion in postischemic injury to the renal allograft. J Clin Invest. 1998;101:2054–64. doi: 10.1172/JCI772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Leemreis JR, Versteilen AM, Sipkema P, Groeneveld AB, Musters RJ. Digital image analysis of cytoskeletal F-actin disintegration in renal microvascular endothelium following ischemia/reperfusion. Cytometry A. 2006;69:973–8. doi: 10.1002/cyto.a.20269. [DOI] [PubMed] [Google Scholar]
  • 99.Genesca M, Sola A, Hotter G. Actin cytoskeleton derangement induces apoptosis in renal ischemia/reperfusion. Apoptosis. 2006;11:563–71. doi: 10.1007/s10495-006-4937-1. [DOI] [PubMed] [Google Scholar]
  • 100.Noll T, Muhs A, Besselmann M, Watanabe H, Piper HM. Initiation of hyperpermeability in energy-depleted coronary endothelial monolayers. Am J Physiol. 1995;268:H1462–70. doi: 10.1152/ajpheart.1995.268.4.H1462. [DOI] [PubMed] [Google Scholar]
  • 101.Kevil CG, Oshima T, Alexander B, Coe LL, Alexander JS. H(2)O(2)-mediated permeability: role of MAPK and occludin. Am J Physiol Cell Physiol. 2000;279:C21–30. doi: 10.1152/ajpcell.2000.279.1.C21. [DOI] [PubMed] [Google Scholar]
  • 102.Kevil CG, Oshima T, Alexander JS. The role of p38 MAP kinase in hydrogen peroxide mediated endothelial solute permeability. Endothelium. 2001;8:107–16. doi: 10.3109/10623320109165320. [DOI] [PubMed] [Google Scholar]
  • 103.Ochoa L, Waypa G, Mahoney JR, Jr, Rodriguez L, Minnear FL. Contrasting effects of hypochlorous acid and hydrogen peroxide on endothelial permeability: prevention with cAMP drugs. Am J Respir Crit Care Med. 1997;156:1247–55. doi: 10.1164/ajrccm.156.4.96-10115. [DOI] [PubMed] [Google Scholar]
  • 104.Basile DP, Donohoe D, Roethe K, Osborn JL. Renal ischemic injury results in permanent damage to peritubular capillaries and influences long-term function. Am J Physiol Renal Physiol. 2001;281:F887–99. doi: 10.1152/ajprenal.2001.281.5.F887. [DOI] [PubMed] [Google Scholar]
  • 105.Sutton TA, Horbelt M, Sandoval RM. Imaging vascular pathology. Nephron Physiol. 2006;103:p82–5. doi: 10.1159/000090624. [DOI] [PubMed] [Google Scholar]
  • 106.Dejana E. Endothelial cell-cell junctions: happy together. Nat Rev Mol Cell Biol. 2004;5:261–70. doi: 10.1038/nrm1357. [DOI] [PubMed] [Google Scholar]
  • 107.Soullier S, Gayrard N, Méjean C, Swarcz I, Mourad G, Argilés A. Molecular mechanisms involved in kidney ischemia-reperfusion [in French] Nephrol Ther. 2005;1:315–21. doi: 10.1016/j.nephro.2005.05.005. [DOI] [PubMed] [Google Scholar]
  • 108.Johnston WH, Latta H. Glomerular mesangial and endothelial cell swelling following temporary renal ischemia and its role in the noreflow phenomenon. Am J Pathol. 1977;89:153–66. [PMC free article] [PubMed] [Google Scholar]
  • 109.Flores J, DiBona DR, Beck CH, Leaf A. The role of cell swelling in ischemic renal damage and the protective effect of hypertonic solute. J Clin Invest. 1972;51:118–26. doi: 10.1172/JCI106781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Nilius B, Droogmans G. Ion channels and their functional role in vascular endothelium. Physiol Rev. 2001;81:1415–59. doi: 10.1152/physrev.2001.81.4.1415. [DOI] [PubMed] [Google Scholar]
  • 111.Orie NN, Fry CH, Clapp LH. Evidence that inward rectifier K+ channels mediate relaxation by the PGI2 receptor agonist cicaprost via a cyclic AMP-independent mechanism. Cardiovasc Res. 2006;69:107–15. doi: 10.1016/j.cardiores.2005.08.004. [DOI] [PubMed] [Google Scholar]
  • 112.Bychkov R, et al. Hydrogen peroxide, potassium currents, and membrane potential in human endothelial cells. Circulation. 1999;99:1719–25. doi: 10.1161/01.cir.99.13.1719. [DOI] [PubMed] [Google Scholar]
  • 113.Kramer AA, Postler G, Salhab KF, Mendez C, Carey LC, Rabb H. Renal ischemia/reperfusion leads to macrophage-mediated increase in pulmonary vascular permeability. Kidney Int. 1999;55:2362–7. doi: 10.1046/j.1523-1755.1999.00460.x. [DOI] [PubMed] [Google Scholar]
  • 114.Okajima K. Prevention of endothelial cell injury by activated protein C: the molecular mechanism(s) and therapeutic implications. Curr Vasc Pharmacol. 2004;2:125–33. doi: 10.2174/1570161043476429. [DOI] [PubMed] [Google Scholar]
  • 115.Peek GJ, Firmin RK. The inflammatory and coagulative response to prolonged extra-corporeal membrane oxygenation. ASAIO J. 1999;45:250–63. doi: 10.1097/00002480-199907000-00003. [DOI] [PubMed] [Google Scholar]
  • 116.Xu XP, Pollock JS, Tanner MA, Myers PR. Hypoxia activates nitric oxide synthase and stimulates nitric oxide production in porcine coronary resistance arteriolar endothelial cells. Cardiovasc Res. 1995;30:841–7. [PubMed] [Google Scholar]
  • 117.Aird WC. Endothelial cell heterogeneity. Crit Care Med. 2003;31:S221–30. doi: 10.1097/01.CCM.0000057847.32590.C1. [DOI] [PubMed] [Google Scholar]
  • 118.Stevens T, et al. NHLBI workshop report: endothelial cell phenotypes in heart, lung, and blood diseases. Am J Physiol Cell Physiol. 2001;281:C1422–33. doi: 10.1152/ajpcell.2001.281.5.C1422. [DOI] [PubMed] [Google Scholar]
  • 119.Sheridan AM, Bonventre JV. Cell biology and molecular mechanisms of injury in ischemic acute renal failure. Curr Opin Nephrol Hypertens. 2000;9:427–34. doi: 10.1097/00041552-200007000-00015. [DOI] [PubMed] [Google Scholar]
  • 120.Dragun D, Haller H. Diapedesis of leukocytes: antisense oligonucleotides for rescue. Exp Nephrol. 1999;7:185–92. doi: 10.1159/000020599. [DOI] [PubMed] [Google Scholar]
  • 121.Burne-Taney MJ, Rabb H. The role of adhesion molecules and T cells in ischemic renal injury. Curr Opin Nephrol Hypertens. 2003;12:85–90. doi: 10.1097/00041552-200301000-00014. [DOI] [PubMed] [Google Scholar]
  • 122.Savransky V, Molls RR, Burne-Taney M, Chien CC, Racusen L, Rabb H. Role of the T-cell receptor in kidney ischemia-reperfusion injury. Kidney Int. 2006;69:233–8. doi: 10.1038/sj.ki.5000038. [DOI] [PubMed] [Google Scholar]
  • 123.Ysebaert DK, et al. T cells as mediators in renal ischemia/reperfusion injury. Kidney Int. 2004;66:491–6. doi: 10.1111/j.1523-1755.2004.761_4.x. [DOI] [PubMed] [Google Scholar]
  • 124.Friedewald JJ, Rabb H. Inflammatory cells in ischemic acute renal failure. Kidney Int. 2004;66:486–91. doi: 10.1111/j.1523-1755.2004.761_3.x. [DOI] [PubMed] [Google Scholar]
  • 125.Thornton MA, Winn R, Alpers CE, Zager RA. An evaluation of the neutrophil as a mediator of in vivo renal ischemic-reperfusion injury. Am J Pathol. 1989;135:509–15. [PMC free article] [PubMed] [Google Scholar]
  • 126.Paller MS. Effect of neutrophil depletion on ischemic renal injury in the rat. J Lab Clin Med. 1989;113:379–86. [PubMed] [Google Scholar]
  • 127.Rabb H, O’Meara YM, Maderna P, Coleman P, Brady HR. Leukocytes, cell adhesion molecules and ischemic acute renal failure. Kidney Int. 1997;51:1463–8. doi: 10.1038/ki.1997.200. [DOI] [PubMed] [Google Scholar]
  • 128.Takada M, Chandraker A, Nadeau KC, Sayegh MH, Tilney NL. The role of the B7 co-stimulatory pathway in experimental cold ischemia/reperfusion injury. J Clin Invest. 1997;100:1199–203. doi: 10.1172/JCI119632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.De Greef KE, et al. Anti-B7-1 blocks mononuclear cell adherence in vasa recta after ischemia. Kidney Int. 2001;60:1415–27. doi: 10.1046/j.1523-1755.2001.00944.x. [DOI] [PubMed] [Google Scholar]
  • 130.Chandraker A, Takada M, Nadeau KC, Peach R, Tilney NL, Sayegh MH. CD28-b7 blockade in organ dysfunction secondary to cold ischemia/reperfusion injury. Kidney Int. 1997;52:1678–84. doi: 10.1038/ki.1997.502. [DOI] [PubMed] [Google Scholar]
  • 131.Rabb H, et al. Pathophysiological role of T lymphocytes in renal ischemia-reperfusion injury in mice. Am J Physiol Renal Physiol. 2000;279:F525–31. doi: 10.1152/ajprenal.2000.279.3.F525. [DOI] [PubMed] [Google Scholar]
  • 132.Bonventre JV, Weinberg JM. Recent advances in the pathophysiology of ischemic acute renal failure. J Am Soc Nephrol. 2003;14:2199–210. doi: 10.1097/01.asn.0000079785.13922.f6. [DOI] [PubMed] [Google Scholar]
  • 133.Rabb H, et al. Role of CD11a and CD11b in ischemic acute renal failure in rats. Am J Physiol. 1994;267:F1052–8. doi: 10.1152/ajprenal.1994.267.6.F1052. [DOI] [PubMed] [Google Scholar]
  • 134.Rao RM, Yang L, Garcia-Cardena G, Luscinskas FW. Endothelial-dependent mechanisms of leukocyte recruitment to the vascular wall. Circ Res. 2007;101:234–7. doi: 10.1161/CIRCRESAHA.107.151860b. [DOI] [PubMed] [Google Scholar]
  • 135.Chiao H, Kohda Y, McLeroy P, Craig L, Housini I, Star RA. Alpha-melanocyte-stimulating hormone protects against renal injury after ischemia in mice and rats. J Clin Invest. 1997;99:1165–72. doi: 10.1172/JCI119272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Donnahoo KK, Meng X, Ayala A, Cain MP, Harken AH, Meldrum DR. Early kidney TNF-alpha expression mediates neutrophil infiltration and injury after renal ischemia-reperfusion. Am J Physiol. 1999;277:R922–9. doi: 10.1152/ajpregu.1999.277.3.R922. [DOI] [PubMed] [Google Scholar]
  • 137.Simmons EM, et al. Plasma cytokine levels predict mortality in patients with acute renal failure. Kidney Int. 2004;65:1357–65. doi: 10.1111/j.1523-1755.2004.00512.x. [DOI] [PubMed] [Google Scholar]
  • 138.Martin DR, Lewington AJ, Hammerman MR, Padanilam BJ. Inhibition of poly(ADP-ribose) polymerase attenuates ischemic renal injury in rats. Am J Physiol Regul Integr Comp Physiol. 2000;279:R1834–40. doi: 10.1152/ajpregu.2000.279.5.R1834. [DOI] [PubMed] [Google Scholar]
  • 139.Chatterjee PK, Zacharowski K, Cuzzocrea S, Otto M, Thiemermann C. Inhibitors of poly (ADP-ribose) synthetase reduce renal ischemia-reperfusion injury in the anesthetized rat in vivo. FASEB J. 2000;14:641–51. doi: 10.1096/fasebj.14.5.641. [DOI] [PubMed] [Google Scholar]
  • 140.Zheng J, Devalaraja-Narashimha K, Singaravelu K, Padanilam BJ. Poly(ADP-ribose) polymerase-1 gene ablation protects mice from ischemic renal injury. Am J Physiol Renal Physiol. 2005;288:F387–98. doi: 10.1152/ajprenal.00436.2003. [DOI] [PubMed] [Google Scholar]
  • 141.Kota BP, Huang TH, Roufogalis BD. An overview on biological mechanisms of PPARs. Pharmacol Res. 2005;51:85–94. doi: 10.1016/j.phrs.2004.07.012. [DOI] [PubMed] [Google Scholar]
  • 142.Devalaraja-Narashimha K, Singaravelu K, Padanilam BJ. Poly(ADP-ribose) polymerase-mediated cell injury in acute renal failure. Pharmacol Res. 2005;52:44–59. doi: 10.1016/j.phrs.2005.02.022. [DOI] [PubMed] [Google Scholar]
  • 143.Kelly KJ, Williams WW, Jr, Colvin RB, Bonventre JV. Antibody to intercellular adhesion molecule 1 protects the kidney against ischemic injury. Proc Natl Acad Sci U S A. 1994;91:812–6. doi: 10.1073/pnas.91.2.812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Molitoris BA, Marrs J. The role of cell adhesion molecules in ischemic acute renal failure. Am J Med. 1999;106:583–92. doi: 10.1016/s0002-9343(99)00061-3. [DOI] [PubMed] [Google Scholar]
  • 145.Takada M, Nadeau KC, Shaw GD, Marquette KA, Tilney NL. The cytokine-adhesion molecule cascade in ischemia/reperfusion injury of the rat kidney: inhibition by a soluble P-selectin ligand. J Clin Invest. 1997;99:2682–90. doi: 10.1172/JCI119457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Singbartl K, Forlow SB, Ley K. Platelet, but not endothelial, P-selectin is critical for neutrophil-mediated acute postischemic renal failure. FASEB J. 2001;15:2337–44. doi: 10.1096/fj.01-0199com. [DOI] [PubMed] [Google Scholar]
  • 147.Jayle C, et al. Protective role of selectin ligand inhibition in a large animal model of kidney ischemia-reperfusion injury. Kidney Int. 2006;69:1749–55. doi: 10.1038/sj.ki.5000335. [DOI] [PubMed] [Google Scholar]
  • 148.Burne-Taney MJ, Kofler J, Yokota N, Weisfeldt M, Traystman RJ, Rabb H. Acute renal failure after whole body ischemia is characterized by inflammation and T cell-mediated injury. Am J Physiol Renal Physiol. 2003;285:F87–94. doi: 10.1152/ajprenal.00026.2003. [DOI] [PubMed] [Google Scholar]
  • 149.Lieberthal W. Biology of ischemic and toxic renal tubular cell injury: role of nitric oxide and the inflammatory response. Curr Opin Nephrol Hypertens. 1998;7:289–95. doi: 10.1097/00041552-199805000-00009. [DOI] [PubMed] [Google Scholar]
  • 150.Brezis M, Heyman SN, Epstein FH. Determinants of intrarenal oxygenation. II Hemodynamic effects. Am J Physiol. 1994;267:F1063–8. doi: 10.1152/ajprenal.1994.267.6.F1063. [DOI] [PubMed] [Google Scholar]
  • 151.O’Connor PM, Kett MM, Anderson WP, Evans RG. Renal medullary tissue oxygenation is dependent on both cortical and medullary blood flow. Am J Physiol Renal Physiol. 2006;290:F688–94. doi: 10.1152/ajprenal.00275.2005. [DOI] [PubMed] [Google Scholar]
  • 152.Brezis M, Agmon Y, Epstein FH. Determinants of intrarenal oxygenation. I Effects of diuretics. Am J Physiol. 1994;267:F1059–62. doi: 10.1152/ajprenal.1994.267.6.F1059. [DOI] [PubMed] [Google Scholar]
  • 153.Koivisto A, Pittner J, Froelich M, Persson AE. Oxygen-dependent inhibition of respiration in isolated renal tubules by nitric oxide. Kidney Int. 1999;55:2368–75. doi: 10.1046/j.1523-1755.1999.00474.x. [DOI] [PubMed] [Google Scholar]
  • 154.Varela M, Herrera M, Garvin JL. Inhibition of Na-K-ATPase in thick ascending limbs and is diminished by a by NO depends on O2− high-salt diet. Am J Physiol Renal Physiol. 2004;287:F224–30. doi: 10.1152/ajprenal.00427.2003. [DOI] [PubMed] [Google Scholar]
  • 155.Vinas JL, Sola A, Hotter G. Mitochondrial NOS upregulation during renal I/R causes apoptosis in a peroxynitrite-dependent manner. Kidney Int. 2006;69:1403–9. doi: 10.1038/sj.ki.5000361. [DOI] [PubMed] [Google Scholar]
  • 156.Moncada S, Erusalimsky JD. Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat Rev Mol Cell Biol. 2002;3:214–20. doi: 10.1038/nrm762. [DOI] [PubMed] [Google Scholar]
  • 157.Laycock SK, et al. Role of nitric oxide in the control of renal oxygen consumption and the regulation of chemical work in the kidney. Circ Res. 1998;82:1263–71. doi: 10.1161/01.res.82.12.1263. [DOI] [PubMed] [Google Scholar]
  • 158.Gladwin MT, et al. Nitrite as a vascular endocrine nitric oxide reservoir that contributes to hypoxic signaling, cytoprotection, and vasodilation. Am J Physiol Heart Circ Physiol. 2006;291:H2026–35. doi: 10.1152/ajpheart.00407.2006. [DOI] [PubMed] [Google Scholar]
  • 159.Brown GC. Nitric oxide regulates mitochondrial respiration and cell functions by inhibiting cytochrome oxidase. FEBS Lett. 1995;369:136–9. doi: 10.1016/0014-5793(95)00763-y. [DOI] [PubMed] [Google Scholar]
  • 160.Davidson SM, Duchen MR. Effects of NO on mitochondrial function in cardiomyocytes: pathophysiological relevance. Cardiovasc Res. 2006;71:10–21. doi: 10.1016/j.cardiores.2006.01.019. [DOI] [PubMed] [Google Scholar]
  • 161.Boveris A, Costa LE, Poderoso JJ, Carreras MC, Cadenas E. Regulation of mitochondrial respiration by oxygen and nitric oxide. Ann N Y Acad Sci. 2000;899:121–35. doi: 10.1111/j.1749-6632.2000.tb06181.x. [DOI] [PubMed] [Google Scholar]
  • 162.Takehara Y, et al. Oxygen-dependent reversible inhibition of mitochondrial respiration by nitric oxide. Cell Struct Funct. 1996;21:251–8. doi: 10.1247/csf.21.251. [DOI] [PubMed] [Google Scholar]
  • 163.Toledo-Pereyra LH, Lopez-Neblina F, Toledo AH. Reactive oxygen species and molecular biology of ischemia/reperfusion. Ann Transplant. 2004;9:81–3. [PubMed] [Google Scholar]
  • 164.Bertuglia S, Giusti A. Microvascular oxygenation and oxidative stress during postischemic reperfusion: PO2, ROS, and NO during reperfusion. Adv Exp Med Biol. 2005;566:23–9. doi: 10.1007/0-387-26206-7_4. [DOI] [PubMed] [Google Scholar]
  • 165.Rubio-Gayosso I, Platts SH, Duling BR. Reactive oxygen species mediate modification of glycocalyx during ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol. 2006;290:H2247–56. doi: 10.1152/ajpheart.00796.2005. [DOI] [PubMed] [Google Scholar]
  • 166.Chandel NS, Budinger GR, Choe SH, Schumacker PT. Cellular respiration during hypoxia: role of cytochrome oxidase as the oxygen sensor in hepatocytes. J Biol Chem. 1997;272:18808–16. doi: 10.1074/jbc.272.30.18808. [DOI] [PubMed] [Google Scholar]
  • 167.Brealey D, et al. Mitochondrial dysfunction in a long-term rodent model of sepsis and organ failure. Am J Physiol Regul Integr Comp Physiol. 2004;286:R491–7. doi: 10.1152/ajpregu.00432.2003. [DOI] [PubMed] [Google Scholar]
  • 168.Adler S, et al. Endothelial nitric oxide synthase plays an essential role in regulation of renal oxygen consumption by NO. Am J Physiol Renal Physiol. 2001;280:F838–43. doi: 10.1152/ajprenal.2001.280.5.F838. [DOI] [PubMed] [Google Scholar]
  • 169.Clementi E, Brown GC, Foxwell N, Moncada S. On the mechanism by which vascular endothelial cells regulate their oxygen consumption. Proc Natl Acad Sci U S A. 1999;96:1559–62. doi: 10.1073/pnas.96.4.1559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Mik EG, et al. Mitochondrial PO2 measured by delayed fluorescence of endogenous protoporphyrin IX. Nat Methods. 2006;3:939–45. doi: 10.1038/nmeth940. [DOI] [PubMed] [Google Scholar]
  • 171.Rabelink TJ, Wijewickrama DC, de Koning EJ. Peritubular endothelium: the Achilles heel of the kidney? Kidney Int. 2007;72:926–30. doi: 10.1038/sj.ki.5002414. [DOI] [PubMed] [Google Scholar]
  • 172.Norman JT, Fine LG. Intrarenal oxygenation in chronic renal failure. Clin Exp Pharmacol Physiol. 2006;33:989–96. doi: 10.1111/j.1440-1681.2006.04476.x. [DOI] [PubMed] [Google Scholar]
  • 173.Basile DP, Donohoe DL, Roethe K, Mattson DL. Chronic renal hypoxia after acute ischemic injury: effects of L-arginine on hypoxia and secondary damage. Am J Physiol Renal Physiol. 2003;284:F338–48. doi: 10.1152/ajprenal.00169.2002. [DOI] [PubMed] [Google Scholar]
  • 174.Kang DH, Hughes J, Mazzali M, Schreiner GF, Johnson RJ. Impaired angiogenesis in the remnant kidney model: II. Vascular endothelial growth factor administration reduces renal fibrosis and stabilizes renal function. J Am Soc Nephrol. 2001;12:1448–57. doi: 10.1681/ASN.V1271448. [DOI] [PubMed] [Google Scholar]
  • 175.Yang J, et al. Telomerized human micro-vasculature is functional in vivo. Nat Biotechnol. 2001;19:219–24. doi: 10.1038/85655. [DOI] [PubMed] [Google Scholar]
  • 176.Sharples EJ, et al. Erythropoietin protects the kidney against the injury and dysfunction caused by ischemia-reperfusion. J Am Soc Nephrol. 2004;15:2115–24. doi: 10.1097/01.ASN.0000135059.67385.5D. [DOI] [PubMed] [Google Scholar]
  • 177.Ribatti D, Nico B, Crivellato E, Vacca A. Endothelial progenitor cells in health and disease. Histol Histopathol. 2005;20:1351–8. doi: 10.14670/HH-20.1351. [DOI] [PubMed] [Google Scholar]
  • 178.Basile DP. The endothelial cell in ischemic acute kidney injury: implications for acute and chronic function. Kidney Int. 2007;72:151–6. doi: 10.1038/sj.ki.5002312. [DOI] [PubMed] [Google Scholar]
  • 179.Li J, Deane JA, Campanale NV, Bertram JF, Ricardo SD. The contribution of bone marrow-derived cells to the development of renal interstitial fibrosis. Stem Cells. 2007;25:697–706. doi: 10.1634/stemcells.2006-0133. [DOI] [PubMed] [Google Scholar]
  • 180.Rosenberger C, et al. Hypoxia-inducible factors and tubular cell survival in isolated perfused kidneys. Kidney Int. 2006;70:60–70. doi: 10.1038/sj.ki.5000395. [DOI] [PubMed] [Google Scholar]
  • 181.Rosenberger C, et al. Up-regulation of HIF in experimental acute renal failure: evidence for a protective transcriptional response to hypoxia. Kidney Int. 2005;67:531–42. doi: 10.1111/j.1523-1755.2005.67110.x. [DOI] [PubMed] [Google Scholar]
  • 182.Safran M, Kaelin WG., Jr HIF hydroxylation and the mammalian oxygen-sensing pathway. J Clin Invest. 2003;111:779–83. doi: 10.1172/JCI18181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Haase VH. The VHL/HIF oxygen-sensing pathway and its relevance to kidney disease. Kidney Int. 2006;69:1302–7. doi: 10.1038/sj.ki.5000221. [DOI] [PubMed] [Google Scholar]
  • 184.Haase VH. Hypoxia-inducible factors in the kidney. Am J Physiol Renal Physiol. 2006;291:F271–81. doi: 10.1152/ajprenal.00071.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Mansfield KD, et al. Mitochondrial dysfunction resulting from loss of cytochrome c impairs cellular oxygen sensing and hypoxic HIF-alpha activation. Cell Metab. 2005;1:393–9. doi: 10.1016/j.cmet.2005.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Sandau KB, Fandrey J, Brune B. Accumulation of HIF-1alpha under the influence of nitric oxide. Blood. 2001;97:1009–15. doi: 10.1182/blood.v97.4.1009. [DOI] [PubMed] [Google Scholar]
  • 187.Hagen T, Taylor CT, Lam F, Moncada S. Redistribution of intracellular oxygen in hypoxia by nitric oxide: effect on HIF1alpha. Science. 2003;302:1975–8. doi: 10.1126/science.1088805. [DOI] [PubMed] [Google Scholar]
  • 188.Schrijvers BF, Flyvbjerg A, De Vriese AS. The role of vascular endothelial growth factor (VEGF) in renal pathophysiology. Kidney Int. 2004;65:2003–17. doi: 10.1111/j.1523-1755.2004.00621.x. [DOI] [PubMed] [Google Scholar]
  • 189.Kanellis J, et al. Renal ischemia-reperfusion increases endothelial VEGFR-2 without increasing VEGF or VEGFR-1 expression. Kidney Int. 2002;61:1696–706. doi: 10.1046/j.1523-1755.2002.00329.x. [DOI] [PubMed] [Google Scholar]
  • 190.Kim YG, et al. Vascular endothelial growth factor accelerates renal recovery in experimental thrombotic microangiopathy. Kidney Int. 2000;58:2390–9. doi: 10.1046/j.1523-1755.2000.00422.x. [DOI] [PubMed] [Google Scholar]
  • 191.Takahashi T, Morita K, Akagi R, Sassa S. Protective role of heme oxygenase-1 in renal ischemia. Antioxid Redox Signal. 2004;6:867–77. doi: 10.1089/ars.2004.6.867. [DOI] [PubMed] [Google Scholar]
  • 192.Shimizu H, et al. Protective effect of heme oxygenase induction in ischemic acute renal failure. Crit Care Med. 2000;28:809–17. doi: 10.1097/00003246-200003000-00033. [DOI] [PubMed] [Google Scholar]
  • 193.Salom MG, et al. Heme oxygenase-1 induction improves ischemic renal failure: role of nitric oxide and peroxynitrite. Am J Physiol Heart Circ Physiol. 2007;293:H3542–9. doi: 10.1152/ajpheart.00977.2007. [DOI] [PubMed] [Google Scholar]
  • 194.Aydin Z, Duijs J, Bajema IM, van Zonneveld AJ, Rabelink TJ. Erythropoietin, progenitors, and repair. Kidney Int. 2007;72(Suppl):S16–20. doi: 10.1038/sj.ki.5002483. [DOI] [PubMed] [Google Scholar]
  • 195.Beleslin-Cokic BB, Cokic VP, Yu X, Weksler BB, Schechter AN, Noguchi CT. Erythropoietin and hypoxia stimulate erythropoietin receptor and nitric oxide production by endothelial cells. Blood. 2004;104:2073–80. doi: 10.1182/blood-2004-02-0744. [DOI] [PubMed] [Google Scholar]
  • 196.Biju MP, Akai Y, Shrimanker N, Haase VH. Protection of HIF-1-deficient primary renal tubular epithelial cells from hypoxia-induced cell death is glucose dependent. Am J Physiol Renal Physiol. 2005;289:F1217–26. doi: 10.1152/ajprenal.00233.2005. [DOI] [PubMed] [Google Scholar]
  • 197.Schofield CJ, Ratcliffe PJ. Oxygen sensing by HIF hydroxylases. Nat Rev Mol Cell Biol. 2004;5:343–54. doi: 10.1038/nrm1366. [DOI] [PubMed] [Google Scholar]
  • 198.Rivers E, et al. Early goal-directed therapy in the treatment of severe sepsis and septic shock. N Engl J Med. 2001;345:1368–77. doi: 10.1056/NEJMoa010307. [DOI] [PubMed] [Google Scholar]
  • 199.Rivers EP, et al. The influence of early hemodynamic optimization on biomarker patterns of severe sepsis and septic shock. Crit Care Med. 2007;35:2016–24. doi: 10.1097/01.ccm.0000281637.08984.6e. [DOI] [PubMed] [Google Scholar]
  • 200.Lameire NH, De Vriese AS, Vanholder R. Prevention and nondialytic treatment of acute renal failure. Curr Opin Crit Care. 2003;9:481–90. doi: 10.1097/00075198-200312000-00004. [DOI] [PubMed] [Google Scholar]
  • 201.Venkataraman R, Kellum JA. Prevention of acute renal failure. Chest. 2007;131:300–8. doi: 10.1378/chest.06-1246. [DOI] [PubMed] [Google Scholar]
  • 202.De Backer D, et al. The effects of dobutamine on microcirculatory alterations in patients with septic shock are independent of its systemic effects. Crit Care Med. 2006;34:403–8. doi: 10.1097/01.ccm.0000198107.61493.5a. [DOI] [PubMed] [Google Scholar]
  • 203.Schortgen F, et al. Effects of hydroxyethylstarch and gelatin on renal function in severe sepsis: a multicentre randomised study. Lancet. 2001;357:911–6. doi: 10.1016/S0140-6736(00)04211-2. [DOI] [PubMed] [Google Scholar]
  • 204.Brunkhorst FM, et al. Intensive insulin therapy and pentastarch resuscitation in severe sepsis. N Engl J Med. 2008;358:125–39. doi: 10.1056/NEJMoa070716. [DOI] [PubMed] [Google Scholar]
  • 205.Yuruk K, Almac E, Ince C. Hydroxyethyl starch solutions and their effect on the microcirculation and tissue oxygenation. Transfus Altern Transfus Med. 2007;9:164–72. [Google Scholar]
  • 206.Blasco V, Leone M, Antonini F, Geissler A, Albanese J, Martin C. Comparison of the novel hydroxyethylstarch 130/0.4 and hydroxyethylstarch 200/0.6 in brain-dead donor resuscitation on renal function after transplantation. Br J Anaesth. 2008;100:504–8. doi: 10.1093/bja/aen001. [DOI] [PubMed] [Google Scholar]
  • 207.Sakr Y, et al. Does dopamine administration in shock influence outcome? Results of the Sepsis Occurrence in Acutely Ill Patients (SOAP) Study. Crit Care Med. 2006;34:589–97. doi: 10.1097/01.CCM.0000201896.45809.E3. [DOI] [PubMed] [Google Scholar]
  • 208.Faivre V, et al. Cardiac and renal effects of levosimendan, arginine vasopressin, and norepinephrine in lipopolysaccharide-treated rabbits. Anesthesiology. 2005;103:514–21. doi: 10.1097/00000542-200509000-00014. [DOI] [PubMed] [Google Scholar]
  • 209.Albert M, Losser MR, Hayon D, Faivre V, Payen D. Systemic and renal macro- and micro-circulatory responses to arginine vasopressin in endotoxic rabbits. Crit Care Med. 2004;32:1891–8. doi: 10.1097/01.ccm.0000139708.10718.e3. [DOI] [PubMed] [Google Scholar]
  • 210.Gattinoni L, et al. A trial of goal-oriented hemodynamic therapy in critically ill patients. SvO2 Collaborative Group. N Engl J Med. 1995;333:1025–32. doi: 10.1056/NEJM199510193331601. [DOI] [PubMed] [Google Scholar]
  • 211.Hayes MA, Timmins AC, Yau EH, Palazzo M, Hinds CJ, Watson D. Elevation of systemic oxygen delivery in the treatment of critically ill patients. N Engl J Med. 1994;330:1717–22. doi: 10.1056/NEJM199406163302404. [DOI] [PubMed] [Google Scholar]
  • 212.Durairaj L, Schmidt GA. Fluid therapy in resuscitated sepsis: less is more. Chest. 2008;133:252–63. doi: 10.1378/chest.07-1496. [DOI] [PubMed] [Google Scholar]
  • 213.Thurau K, Boylan JW. Acute renal success: the unexpected logic of oliguria in acute renal failure. Am J Med. 1976;61:308–15. doi: 10.1016/0002-9343(76)90365-x. [DOI] [PubMed] [Google Scholar]
  • 214.Macias WL, et al. New insights into the protein C pathway: potential implications for the biological activities of drotrecogin alfa (activated) Crit Care. 2005;9(Suppl 4):S38–45. doi: 10.1186/cc3747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Bernard GR, et al. Efficacy and safety of recombinant human activated protein C for severe sepsis. N Engl J Med. 2001;344:699–709. doi: 10.1056/NEJM200103083441001. [DOI] [PubMed] [Google Scholar]
  • 216.Gupta A, Rhodes GJ, Berg DT, Gerlitz B, Molitoris BA, Grinnell BW. Activated protein C ameliorates LPS-induced acute kidney injury and downregulates renal INOS and angiotensin 2. Am J Physiol Renal Physiol. 2007;293:F245–54. doi: 10.1152/ajprenal.00477.2006. [DOI] [PubMed] [Google Scholar]
  • 217.Gupta A, et al. Role of protein C in renal dysfunction after polymicrobial sepsis. J Am Soc Nephrol. 2007;18:860–7. doi: 10.1681/ASN.2006101167. [DOI] [PubMed] [Google Scholar]
  • 218.Dear JW, et al. Sepsis-induced organ failure is mediated by different pathways in the kidney and liver: acute renal failure is dependent on MyD88 but not renal cell apoptosis. Kidney Int. 2006;69:832–6. doi: 10.1038/sj.ki.5000165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Ventura CG, Coimbra TM, de Campos SB, de Castro I, Yu L, Seguro AC. Mycophenolate mofetil attenuates renal ischemia/reperfusion injury. J Am Soc Nephrol. 2002;13:2524–33. doi: 10.1097/01.asn.0000030143.73830.3c. [DOI] [PubMed] [Google Scholar]
  • 220.Ysebaert DK, et al. Effect of immunosuppression on damage, leukocyte infiltration, and regeneration after severe warm ischemia/reperfusion renal injury. Kidney Int. 2003;64:864–73. doi: 10.1046/j.1523-1755.2003.00150.x. [DOI] [PubMed] [Google Scholar]
  • 221.Ricci Z, Cruz D, Ronco C. The RIFLE criteria and mortality in acute kidney injury: A systematic review. Kidney Int. 2008;73:538–46. doi: 10.1038/sj.ki.5002743. [DOI] [PubMed] [Google Scholar]
  • 222.Bonventre JV. Diagnosis of acute kidney injury: from classic parameters to new biomarkers. Contrib Nephrol. 2007;156:213–9. doi: 10.1159/000102086. [DOI] [PubMed] [Google Scholar]

Articles from Molecular Medicine are provided here courtesy of The Feinstein Institute for Medical Research at North Shore LIJ

RESOURCES