Sepsis-induced acute kidney injury

Hernando Gómez, John A Kellum, Hernando Gómez, John A Kellum

Abstract

Purpose of review: Sepsis is a common and frequently fatal condition in which mortality has been consistently linked to increasing organ dysfunction. For example, acute kidney injury (AKI) occurs in 40-50% of septic patients and increases mortality six to eight-fold. However, the mechanisms by which sepsis causes organ dysfunction are not well understood and hence current therapy remains reactive and nonspecific.

Recent findings: Recent studies have challenged the previous notion that organ dysfunction is solely secondary to hypoperfusion, by showing, for example, that AKI occurs in the setting of normal or increased renal blood flow; and that it is characterized not by acute tubular necrosis or apoptosis, but rather by heterogeneous areas of colocalized sluggish peritubular blood flow and tubular epithelial cell oxidative stress. Evidence has also shown that microvascular dysfunction, inflammation, and the metabolic response to inflammatory injury are fundamental pathophysiologic mechanisms that may explain the development of sepsis-induced AKI.

Summary: The implications of these findings are significant because in the context of decades of negative clinical trials in the field, the recognition that other mechanisms are at play opens the possibility to better understand the processes of injury and repair, and provides an invaluable opportunity to design mechanism-targeted therapeutic interventions.

Conflict of interest statement

Conflicts of interest

There are no conflicts of interest.

Figures

FIGURE 1
FIGURE 1
The figure summarizes the mechanisms that are thought to participate in the development of microvascular dysfunction. Damage-associated molecular patterns and pathogen-associated molecular patterns resulting from invading pathogens and the subsequent immune response activate leukocytes and endothelial cells. Activation of, and injury to endothelial cells directly or through oxidative stress, induces alterations in protective mechanisms (see text). In addition, damaged or activated endothelial cells undergo shedding of their glycocalyx which exposes adhesion molecules to circulating leukocytes and platelets, promoting adhesion of both, and rolling and transmigration of leukocytes; alters the barrier function of the capillary, resulting in capillary leak and formation of edema; alters sensing of shear stress forces which are necessary to regulate tone and couple blood flow to changing circumstances; and causes profound alterations in blood flow distribution. DAMPs, damage-associated molecular patterns; eNOS, endothelial-derived nitric oxide synthase; iNOS, inducible nitric oxide synthase; NO, nitric oxide. Adapted with permission from [19].
FIGURE 2
FIGURE 2
The figure represents a potential model to explain how the interactions between sepsis-induced microvascular dysfunction, inflammation and the metabolic response from TECs interact. It is still unknown if these are the only interactions, if the directions of the interactions are correct, and what is the sequence of events that spins this injury-response cycle into reverberation. Here, the response of the TEC to injury is governed by metabolic reprogramming. Through this, the TEC is able to optimize energy production and prioritize energy expenditure, while supporting the necessary supply of carbon, nitrogen and other components for the synthesis of proteins and structural components necessary to mount an innate immune response. This may be achieved early on by activating a series of master regulators of cellular metabolism that ultimately are capable of switching between aerobic glycolysis and OXPHOS, and engaging other processes necessary for the survival of the cell such as autophagy (mitophagy) and biogenesis. In addition, cell cycle arrest may be a complementary important protective strategy to avoid the overtaxing energy expenditure and the risk of replicating damaged DNA during mitosis, particularly in the context of scarce energy resources. AKI, acute kidney injury; DAMPs, damage-associated molecular patterns; OXPHOS, oxidative phosphorylation; TEC, tubular epithelial cell. Bottom panel is adapted with permission from [37], and cell cycle panel is adapted with permission from [38■].

References

    1. Hoste EA, Bagshaw SM, Bellomo R, et al. Epidemiology of acute kidney injury in critically ill patients: the multinational AKI-EPI study. Intensive Care Med. 2015;41:1411–1423. This is the first multinational cross-sectional study assessing the epidemiology of AKI in ICU patients based on the KDIGO criteria, demonstrating that AKI is very common affecting about half of the critically ill population.
    1. Uchino S, Kellum JA, Bellomo R, et al. Acute renal failure in critically ill patients: a multinational, multicenter study. JAMA. 2005;294:813–818.
    1. Thakar CV, Christianson A, Freyberg R, et al. Incidence and outcomes of acute kidney injury in intensive care units: a Veterans Administration study. Crit Care Med. 2009;37:2552–2558.
    1. Murugan R, Kellum JA. Acute kidney injury: what’s the prognosis? Nat Rev Nephrol. 2011;7:209–217.
    1. Langenberg C, Wan L, Egi M, et al. Renal blood flow in experimental septic acute renal failure. Kidney Int. 2006;69:1996–2002.
    1. Prowle JR, Ishikawa K, May CN, Bellomo R. Renal blood flow during acute renal failure in man. Blood Purif. 2009;28:216–225.
    1. Murugan R, Karajala-Subramanyam V, Lee M, et al. Genetic and Inflammatory Markers of Sepsis (GenIMS) Investigators Acute kidney injury in nonsevere pneumonia is associated with an increased immune response and lower survival. Kidney Int. 2010;77:527–535.
    1. Hotchkiss RS, Swanson PE, Freeman BD, et al. Apoptotic cell death in patients with sepsis, shock, and multiple organ dysfunction. Crit Care Med. 1999;27:1230–1251.
    1. Takasu O, Gaut JP, Watanabe E, et al. Mechanisms of cardiac and renal dysfunction in patients dying of sepsis. Am J Respir Crit Care Med. 2013;187:509–517.
    1. De Backer D, Creteur J, Preiser J-C, et al. Microvascular blood flow is altered in patients with sepsis. Am J Respir Crit Care Med. 2002;166:98–104.
    1. Wang Z, Holthoff JH, Seely KA, et al. Development of oxidative stress in the peritubular capillary microenvironment mediates sepsis-induced renal microcirculatory failure and acute kidney injury. Am J Pathol. 2012;180:505–516.
    1. Seely KA, Holthoff JH, Burns ST, et al. Hemodynamic changes in the kidney in a pediatric rat model of sepsis-induced acute kidney injury. Am J Physiol Renal Physiol. 2011;301:F209–F217.
    1. Singer M, De Santis V, Vitale D, Jeffcoate W. Multiorgan failure is an adaptive, endocrine-mediated, metabolic response to overwhelming systemic inflammation. Lancet. 2004;364:545–548.
    1. Spronk PE, Ince C, Gardien MJ, et al. Nitroglycerin in septic shock after intravascular volume resuscitation. Lancet. 2002;360:1395–1396.
    1. De Backer D, Donadello K, Taccone FS, et al. Microcirculatory alterations: potential mechanisms and implications for therapy. Ann Intensive Care. 2011;1:27.
    1. Tiwari MM, Brock RW, Megyesi JK, et al. Disruption of renal peritubular blood flow in lipopolysaccharide-induced renal failure: role of nitric oxide and caspases. Am J Physiol Renal Physiol. 2005;289:F1324–F1332.
    1. Holthoff JH, Wang Z, Seely KA, et al. Resveratrol improves renal microcirculation, protects the tubular epithelium, and prolongs survival in a mouse model of sepsis-induced acute kidney injury. Kidney Int. 2012;81:370–378.
    1. Bezemer R, Legrand M, Klijn E, et al. Real-time assessment of renal cortical microvascular perfusion heterogeneities using near-infrared laser speckle imaging. Opt Express. 2010;18:15054–15061.
    1. Zafrani L, Payen D, Azoulay E, Ince C. The microcirculation of the septic kidney. Semin Nephrol. 2015;35:75–84.
    1. Tyml K, Wang X, Lidington D, Ouellette Y. Lipopolysaccharide reduces intercellular coupling in vitro and arteriolar conducted response in vivo. Am J Physiol Heart Circ Physiol. 2001;281:H1397–H1406.
    1. Sprague AH, Khalil RA. Inflammatory cytokines in vascular dysfunction and vascular disease. Biochem Pharmacol. 2009;78:539–552.
    1. Dyson A, Bezemer R, Legrand M, et al. Microvascular and interstitial oxygen tension in the renal cortex and medulla studied in a 4-h rat model of LPS-induced endotoxemia. Shock. 2011;36:83–89.
    1. Almac E, Siegemund M, Demirci C, Ince C. Microcirculatory recruitment maneuvers correct tissue CO2 abnormalities in sepsis. Minerva Anestesiol. 2006;72:507–519.
    1. Prowle JR, Echeverri JE, Ligabo EV, et al. Fluid balance and acute kidney injury. Nat Rev Nephrol. 2010;6:107–115.
    1. Bagshaw SM, Brophy PD, Cruz D, Ronco C. Fluid balance as a biomarker: impact of fluid overload on outcome in critically ill patients with acute kidney injury. Crit Care. 2008;12:169.
    1. Hollenberg SM, Ahrens TS, Annane D, et al. Practice parameters for hemodynamic support of sepsis in adult patients: 2004 update. Crit Care Med. 2004;32:1928–1948.
    1. Prowle JR, Kirwan CJ, Bellomo R. Fluid management for the prevention and attenuation of acute kidney injury. Nat Rev Nephrol. 2013;10:37–47.
    1. Rajendram R, Prowle JR. Venous congestion: are we adding insult to kidney injury in sepsis? Crit Care. 2014;18:104.
    1. Cruces P, Salas C, Lillo P, et al. The renal compartment: a hydraulic view. Intensive Care Med Exp. 2014;2:26.
    1. Zarbock A, Gomez H, Kellum JA. Sepsis-induced acute kidney injury revisited: pathophysiology, prevention and future therapies. Curr Opin Crit Care. 2014;20:588–595. Excellent review of the pathophysiology of sepsis-induced AKI.
    1. Cunha FQ, Assreuy J, Moss DW, et al. Differential induction of nitric oxide synthase in various organs of the mouse during endotoxaemia: role of TNF-alpha and IL-1-beta. Immunology. 1994;81:211–215.
    1. Trzeciak S, Cinel I, Phillip Dellinger R, et al. Microcirculatory Alterations in Resuscitation and Shock (MARS) Investigators Resuscitating the microcirculation in sepsis: the central role of nitric oxide, emerging concepts for novel therapies, and challenges for clinical trials. Acad Emerg Med. 2008;15:399–413.
    1. Chauhan SD, Seggara G, Vo PA, et al. Protection against lipopolysaccharide-induced endothelial dysfunction in resistance and conduit vasculature of iNOS knockout mice. FASEB J. 2003;17:773–775.
    1. Heemskerk S, Masereeuw R, Russel FG, Pickkers P. Selective iNOS inhibition for the treatment of sepsis-induced acute kidney injury. Nat Rev Nephrol. 2009;5:629–640.
    1. Rabelink TJ, van Zonneveld A-J. Coupling eNOS uncoupling to the innate immune response. Arterioscler Thromb Vasc Biol. 2006;26:2585–2587.
    1. Weinbaum S, Tarbell JM, Damiano ER. The structure and function of the endothelial glycocalyx layer. Annu Rev Biomed Eng. 2007;9:121–167.
    1. Gomez H, Ince C, De Backer D, et al. A unified theory of sepsis-induced acute kidney injury: inflammation, microcirculatory dysfunction, bioenergetics, and the tubular cell adaptation to injury. Shock. 2014;41:3–11.
    1. Emlet DR, Shaw AD, Kellum JA. Sepsis-associated AKI: epithelial cell dysfunction. Semin Nephrol. 2015;35:85–95. Excellent review on the pathophysiology of TEC dysfunction in the context of sepsis.
    1. Ljungqvist A. Ultrastructural demonstration of a connection between afferent and efferent juxtamedullary glomerular arterioles. Kidney Int. 1975;8:239–244.
    1. Mårtensson J, Bellomo R. Sepsis-induced acute kidney injury. Crit Care Clin. 2015;31:649–660.
    1. Hsiao H-W, Tsai K-L, Wang L-F, et al. The decline of autophagy contributes to proximal tubular dysfunction during sepsis. Shock. 2012;37:289–296.
    1. Singh P, Okusa MD. The role of tubuloglomerular feedback in the pathogenesis of acute kidney injury. Contrib Nephrol. 2011;174:12–21.
    1. Ferraro E, Cecconi F. Autophagic and apoptotic response to stress signals in mammalian cells. Arch Biochem Biophys. 2007;462:210–219.
    1. Singer M. The role of mitochondrial dysfunction in sepsis-induced multiorgan failure. Virulence. 2014;5:66–72.
    1. Hochachka PW, Buck LT, Doll CJ, Land SC. Unifying theory of hypoxia tolerance: molecular/metabolic defense and rescue mechanisms for surviving oxygen lack. Proc Natl Acad Sci U S A. 1996;93:9493–9498.
    1. Schmidt C, Höcherl K, Schweda F, Bucher M. Proinflammatory cytokines cause down-regulation of renal chloride entry pathways during sepsis. Crit Care Med. 2007;35:2110–2119.
    1. Mandel LJ, Balaban RS. Stoichiometry and coupling of active transport to oxidative metabolism in epithelial tissues. Am J Physiol. 1981;240:F357–F371.
    1. Good DW, George T, Watts BA. Lipopolysaccharide directly alters renal tubule transport through distinct TLR4-dependent pathways in basolateral and apical membranes. Am J Physiol Renal Physiol. 2009;297:F866–F874.
    1. Frauwirth KA, Riley JL, Harris MH, et al. The CD28 signaling pathway regulates glucose metabolism. Immunity. 2002;16:769–777.
    1. Yang L, Xie M, Yang M, et al. PKM2 regulates the Warburg effect and promotes HMGB1 release in sepsis. Nat Commun. 2014;5:4436.
    1. Waltz P, Carchman E, Gomez H, Zuckerbraun B. Sepsis results in an altered renal metabolic and osmolyte profile. J Surg Res. 2016;202:8–12.
    1. Escobar DA, Botero-Quintero AM, Kautza BC, et al. Adenosine monophosphate-activated protein kinase activation protects against sepsis-induced organ injury and inflammation. J Surg Res. 2014;194:262–272. Experimental study that demonstrated that preemptive activation of the master energy regulator AMP-activated protein kinase protects rodents from sepsis-induced AKI.
    1. Opal S, Ellis JL, Suri V, et al. Pharmacological Sirt1 activation improves motrtality and markedly alters transcriptional profiles that accompany experimental sepsis. Shock. 2016;45:411–418. Experimental study demonstrating that activation of master regulators of metabolism, like Sirt1, in the renal TEC can convey protection to animals exposed to a septic insult.
    1. Han SH, Malaga-Dieguez L, Chinga F, et al. Deletion of Lkb1 in renal tubular epithelial cells leads to CKD by altering metabolism. J Am Soc Nephrol. 2015;27:439–453. Experimental study demonstrating the role of metabolic regulation in the progression of AKI to CKD and the development of renal fibrosis. Importantly, the study demonstrates that the profibrotic phenotype is reversible by manipulating downstream mediators in the lacking the liver kinase B1 pathway.
    1. Kang HM, Ahn SH, Choi P, et al. Defective fatty acid oxidation in renal tubular epithelial cells has a key role in kidney fibrosis development. Nat Med. 2015;21:37–46. Experimental study which demonstrates in vitro that inhibition of FAO in TECs causes ATP depletion, cell death, dedifferentiation, and lipid deposition, all consistent with changes that occur during renal fibrosis. They further demonstrate that restoring FAO by pharmacologic or genetic means protected mice from tubulointerstitial fibrosis induced by folic acid.
    1. Carré JE, Singer M. Cellular energetic metabolism in sepsis: the need for a systems approach. Biochim Biophys Acta. 2008;1777:763–771.
    1. Vanhorebeek I, Gunst J, Derde S, et al. Mitochondrial fusion, fission, and biogenesis in prolonged critically ill patients. J Clin Endocrinol Metab. 2012;97:E59–E64.
    1. Green DR, Galluzzi L, Kroemer G. Mitochondria and the autophagy-inflammation-cell death axis in organismal aging. Science. 2011;333:1109–1112.
    1. Waltz P, Carchman EH, Young AC, et al. Lipopolysaccaride induces autophagic signaling in macrophages via a TLR4, heme oxygenase-1 dependent pathway. Autophagy. 2011;7:315–320.
    1. Frank M, Duvezin-Caubet S, Koob S, et al. Mitophagy is triggered by mild oxidative stress in a mitochondrial fission dependent manner. Biochim Biophys Acta. 2012;1823:2297–2310.
    1. Wang Y, Nartiss Y, Steipe B, et al. ROS-induced mitochondrial depolarization initiates PARK2/PARKIN-dependent mitochondrial degradation by autophagy. Autophagy. 2012;8:1462–1476.
    1. Gunst J, Derese I, Aertgeerts A, et al. Insufficient autophagy contributes to mitochondrial dysfunction, organ failure, and adverse outcome in an animal model of critical illness. Crit Care Med. 2013;41:182–194.
    1. Howell GM, Gomez H, Collage RD, et al. Augmenting autophagy to treat acute kidney injury during endotoxemia in mice. PLoS One. 2013;8:e69520.
    1. Carchman EH, Whelan S, Loughran P, et al. Experimental sepsis-induced mitochondrial biogenesis is dependent on autophagy, TLR4, and TLR9 signaling in liver. FASEB J. 2013;27:4703–4711.
    1. Tran MT, Zsengeller ZK, Berg AH, et al. PGC1α drives NAD biosynthesis linking oxidative metabolism to renal protection. Nature. 2016;531:528–532.
    1. Bartz RR, Fu P, Suliman HB, et al. Staphylococcus aureus sepsis induces early renal mitochondrial DNA repair and mitochondrial biogenesis in mice. PLoS One. 2014;9:e100912–e100919. Experimental study that demonstrates the role of activation of mitochondrial biogenesis as a protective defense mechanism against sepsis-induced renal injury. Sepsis induces mitochondrial damage, and subsequent activation of renal mtDNA repair protein alongside the activation of biogenesis, which ultimately maintains energy homeostasis and potentially influences organ-level repair processes.
    1. MacGarvey NC, Suliman HB, Bartz RR, et al. Activation of mitochondrial biogenesis by heme oxygenase-1-mediated NF-E2-related factor-2 induction rescues mice from lethal Staphylococcus aureus sepsis. Am J Respir Crit Care Med. 2012;185:851–861.
    1. Mizushima N, Levine B, Cuervo AM, Klionsky DJ. Autophagy fights disease through cellular self-digestion. Nature. 2008;451:1069–1075.
    1. Levine B, Yuan J. Autophagy in cell death: an innocent convict? J Clin Invest. 2005;115:2679–2688.
    1. Ciechomska IA, Goemans CG, Tolkovsky AM. Molecular links between autophagy and apoptosis. Methods Mol Biol. 2008;445:175–193.
    1. Carchman EH, Rao J, Loughran PA, et al. Heme oxygenase-1-mediated autophagy protects against hepatocyte cell death and hepatic injury from infection/sepsis in mice. Hepatology. 2011;53:2053–2062.
    1. Yang Q-H, Liu D-W, Long Y, et al. Acute renal failure during sepsis: potential role of cell cycle regulation. J Infect. 2009;58:459–464.
    1. Meersch M, Schmidt C, Van Aken H, et al. Urinary TIMP-2 and IGFBP7 as early biomarkers of acute kidney injury and renal recovery following cardiac surgery. PLoS One. 2014;9:e93460. The observational trial identified two inducers of G1 cell cycle arrest, tissue inhibitor of metalloproteinases 2 and insulin-like growth factor-binding protein 7, as sensitive and specific biomarkers to predict AKI early after cardiac surgery and to predict renal recovery.
    1. Bihorac A, Chawla LS, Shaw AD, et al. Validation of cell-cycle arrest biomarkers for acute kidney injury using clinical adjudication. Am J Respir Crit Care Med. 2014;189:932–939.
    1. Kashani K, Al-Khafaji A, Ardiles T, et al. Discovery and validation of cell cycle arrest biomarkers in human acute kidney injury. Crit Care. 2013;17:R25.

Source: PubMed

3
S'abonner