Autophagy in kidney disease and aging - CyberLeninka

0 downloads 0 Views 810KB Size Report
region).33 Furthermore, activation of autophagy at the whole- body level extends the life span ..... No kidney abnormalities in Atg5-/- pups. 113-117. Beclin1À/À.
review

www.kidney-international.org

Autophagy in kidney disease and aging: lessons from rodent models 1,2

Olivia Lenoir , Pierre-Louis Tharaux

1,2,3,4*

and Tobias B. Huber

OPEN

4,5,6,7*

1

Paris Cardiovascular Research Centre—PARCC, Institut National de la Santé et de la Recherche Médicale (INSERM), Paris, France; Université Paris Descartes, Sorbonne Paris Cité, Paris, France; 3Nephrology Division, Georges Pompidou European Hospital, Assistance Publique-Hôpitaux de Paris, Paris, France; 4FRIAS, Freiburg Institute for Advanced Studies and Center for Biological System Analysis— ZBSA, Freiburg, Germany; 5Department of Medicine IV, Faculty of Medicine, University of Freiburg, Germany; 6BIOSS Center for Biological Signalling Studies, Albert-Ludwigs-University Freiburg, Freiburg, Germany; and 7Center for Systems Biology (ZBSA), Albert-LudwigsUniversity, Freiburg, Germany 2

Autophagy is a highly regulated lysosomal protein degradation pathway that removes protein aggregates and damaged or excess organelles to maintain intracellular homeostasis and cell integrity. Dysregulation of autophagy is involved in the pathogenesis of a variety of metabolic and age-related diseases. Growing evidence suggests that autophagy is implicated in cell injury during renal diseases, both in the tubulointerstitial compartment and in glomeruli. Nevertheless, the impact of autophagy on renal disease progression and aging is still not fully understood. This review summarizes the recent advances in understanding the role of autophagy for kidney disease and aging. Kidney International (2016) j.kint.2016.04.014

-, -–-;

http://dx.doi.org/10.1016/

KEYWORDS: acute kidney injury; aging; autophagy; endothelium; glomerulus; kidney; kidney transplantation; mTOR; podocyte; polycystic kidney disease Copyright ª 2016, International Society of Nephrology. Published by Elsevier Inc. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

A

utophagy (“self-eating” in Greek) is a highly regulated lysosomal protein degradation pathway that removes protein aggregates and damaged organelles to maintain intracellular homeostasis and cell integrity.1–3 This process was first described in 1957 by Sam Clark Jr.,4 but the term autophagy was coined by Christian de Duve only in the late 1950s.5 The autophagy process is well conserved in the evolution from yeast to mammals.6,7 Characterization of the molecular regulators of autophagy was first described in the 1990s with the identification of autophagy-related genes (ATGs) in autophagy-defective yeast cells,8,9 Then mammal orthologs and the molecular machinery were identified.10 Three types of autophagy have been described to date as follows: macroautophagy, the most studied form and the focus of this review; microautophagy; and chaperonemediated autophagy. All differ in their mechanisms and functions11: microautophagy involves the engulfment of small cytoplasmic cargos within lysosomal membrane invaginations12 and chaperone-mediated autophagy involves the heat shock cognate protein 70–mediated recruitment of KFERQ motif–bearing proteins to the lysosome,13,14 whereas macroautophagy is the most prevalent and probably less selective type and is referred to in this review as autophagy. In this review, we describe the pathway of autophagy and highlight its role in renal physiology, renal aging, and kidney diseases. We will also discuss the potential implication of manipulating autophagy as a potential novel renoprotective therapeutic strategy. AUTOPHAGY Functions of autophagy

Correspondence: Pierre-Louis Tharaux, INSERM Paris Cardiovascular Research Centre, 56, rue Leblanc, 75015 Paris, France. E-mail: pierre-louis. [email protected] or Tobias B. Huber, University Hospital Freiburg, Nephrology, Breisacherstrasse 66, Freiburg, BW 79106, Germany. E-mail: [email protected] *These authors contributed equally to this work. Received 10 October 2015; revised 17 April 2016; accepted 20 April 2016 Kidney International (2016) -, -–-

Autophagy has been demonstrated to be essential for a number of fundamental biological activities,15 such as the maintenance of cellular homeostasis and cellular stress response, particularly in postmitotic cells. Autophagy participates in recycling of organelles such as mitochondria (mitophagy),16,17 endoplasmic reticulum (ER), and peroxisomes7,18–22; the clearance of polyubiquitinated proteins aggregates23; and lipid degradation (lipophagy).24 Dysregulation of autophagy is involved in the pathogenesis of a variety of metabolic and age-related diseases.25–31 The major role of 1

review

autophagy is to provide metabolic precursors for survival in stress conditions and to serve as a quality control by clearing misfolded proteins and others cellular debris. Interestingly, genetic evidence (i.e., polymorphisms in ATGs) indicates that autophagy is implicated in several immune diseases associated with kidney dysfunction, such as systemic lupus erythematosus (autophagy related 5 gene [ATG5])32 and rheumatoid arthritis (the PR domain 1 gene [PRDM1]-ATG5 intergenic region).33 Furthermore, activation of autophagy at the wholebody level extends the life span of various model organisms, including mice.28 There is a growing amount of evidence that dysregulation of the autophagic pathway is implicated in the pathogenesis of kidney aging and in several renal diseases such as acute kidney injury (AKI), polycystic kidney disease (PKD), diabetic nephropathy (DN), obstructive nephropathy, focal and segmental glomerulosclerosis (FSGS), and potentially other kidney diseases.34–37 Molecular mechanisms of autophagy

Autophagy proceeds through at least five steps. Autophagosomes are initiated by expansion and sealing of a small vesicle made of a double-membrane structure called phagophore or isolation membrane. The origin of the autophagosome

O Lenoir et al.: Autophagy in kidney disease and aging

membrane may involve different sources, such as lipid droplets, mitochondria, Golgi, ER, plasma membrane, and recycling endosomes.38,39 The phagophore is generated at a specialized site known as the phagophore assembly site or preautophagosomal structure. Sixteen ATG proteins comprise the conserved core ATG machinery that catalyzes formation of the phagophore and its expansion into autophagosomes in all eukaryotes40 (Figure 1). The metabolites generated in the lysosomes/vacuoles are subsequently transported in the cytoplasm and recycled. Macroautophagy has recently been shown to achieve some selectivity for lipid droplets, ribosomes, pathogens, surplus reticulum, peroxisomes, enzymes, and proteins aggregates in processes that are called lipophagy, ribophagy, xenophagy, reticulophagy, pexophagy, zymophagy, and aggrephagy, respectively (reviewed and discussed in Birgisdottir et al.,41 Johansen and Lamark,42 Johansen and Lamark,43 Mochida et al.,44 Khaminets et al.,45 and Khaminets et al.46). Selectivity of autophagy is controlled by autophagy receptors that physically associate with the autophagy compartment by interacting simultaneously with cargo and Atg8- or microtubule-associated protein 1A/1B–light chain 3 (LC3)/g-aminobutyric acid receptor–associated protein–like

Figure 1 | Autophagy pathway. The autophagic pathway responds to signals from the environment. Nutrients status controls autophagy through the mTOR signaling pathway. The class I PI3K-AKT can also activate the mTORC1 complex in response to insulin and other growth factors, acting as a negative regulator of autophagy. Activation of AMPK in response to low energy status (i) inhibits the mTORC1 complex and (ii) activates the ULK1 complex through ULK1 and ATG13 phosphorylation, thereby acting as a positive regulator of autophagy in response to energy depletion. The BECN1/class III PI3K complex, which is inactivated by Bcl2, and the class I PI3K-AKT complex also regulate autophagy. Upon activation, the Beclin1-VPS34-ATG14L-P150 complex will generate PI3P, which promotes autophagosomal membrane nucleation in coordination with the ULK1-ATG13-ATG101-FIP200 complex. The ATG5-ATG12-ATG16L complex and the LC3-phosphatidylethanolamine system will play roles in cargo recruitment, membrane elongation, and autophagosome maturation. Lysosome fusion to autophagosome to form autolysosome is mediated through SNARE proteins, and lysosomal hydrolases degrade proteins, lipids, and nucleic acids. AKT, protein kinase B; AMPK, adenosine monophosphate–activated protein kinase; ATG, autophagy-related gene; BCL2, B-cell lymphoma 2; BCLXL, B-cell lymphoma–extra large; DEPTOR, DEP domain-containing mTOR-interacting protein; FIP200, FAK family kinase-interacting protein of 200 kDa; LC3, microtubule-associated protein 1 light chain 3; MLST8, target of rapamycin complex subunit LST8; mTOR, mammalian target of rapamycin; mTORC1, mTOR complex 1; PI3K, phosphatidylinositol-3-kinase; PI3P, phosphatidylinositol-3-phosphate; PRAS40, proline-rich v-akt murine thymoma viral oncogene homolog 1 substrate 40 kDa; Raptor, regulatory-associated protein of mTOR; SNARE, soluble NSF attachment protein receptor; ULK1, Unc51-like kinase 1. 2

Kidney International (2016) -, -–-

O Lenoir et al.: Autophagy in kidney disease and aging

proteins on autophagosomes.47 The best-characterized cargo receptors are probably sequestosome 1 (SQSTM1)/p62 and neighbor of breast cancer 1 gene (BRCA1) 1, which can sequester aggregated proteins and form the SQSTM1/p62 bodies. One of the prevailing mechanisms to regulate and often stimulate selective autophagy pathways includes phosphorylation of both cargo signals and autophagy receptors. For example, the phosphate and tensin homolog–induced putative kinase protein 1 and ubiquitin (UB) ligase Parkin act cooperatively in sensing mitochondrial functional state and directing damaged mitochondria for disposal via the autophagy pathway (mitophagy). Putative kinase protein 1 phosphorylates UB to activate the UB ligase Parkin, which builds UB chains on mitochondrial outer membrane proteins, where they act to recruit autophagy receptors such as NDP52 and optineurin.48,49 The fast-evolving field of UB-dependent and independent signals in selective autophagy has been reviewed recently.46 Monitoring autophagy

The accumulation of autophagosomes can be used to monitor autophagy. It is important to keep in mind that accumulation of autophagosomes can be caused by an increased autophagic flux or can be due to the inhibition of autophagosomelysosome fusion or inhibition of lysosomal enzyme function.50,51 The formation and maturation of autophagosomes depends on several genes, including microtubule-associated protein 1 light chain 3 (MAP1LC3A/B), beclin 1 gene (BECN1), and ATGs.1,26,52,53 Microtubule-associated protein 1 light chain 3A/B is a soluble protein that belongs to the UBrelated system. The cytoplasmic LC3 is conjugated to phosphatidylethanolamine to produce LC3–PE (or LC3-II), which is subsequently incorporated into to the double-membrane structure of the autophagosome. Therefore, detecting the intracellular level of LC3-II and conversion of cytoplasmic LC3 into LC3-II by immunofluorescence and Western blot analysis can be used as marker for autophagy. However, accumulation of LC3-positive vacuoles or LC3II protein could be due to impaired autophagosome maturation or an acceleration of autophagic flux, which can be further differentiated by inhibition of autolysosomal degradation with bafilomycin A1 or chloroquine. Additionally, the lysosomal turnover of p62, which interacts with LC3-II, can also be used to monitor the autophagic flux.54 The use of an monomeric red fluorescent protein–green fluorescent protein tandem fluorescent-tagged LC3 has also been proposed for visualizing the autophagic flux.55 We refer readers to landmark articles and methodological guidelines.50,51,56,57 Regulation of autophagy

Autophagy is regulated by the major nutrient-sensing pathways, including the mammalian target of rapamycin complex 1 (mTORC1), adenosine monophosphate–activated protein kinase (AMPK), and oxidized nicotinamide adenine dinucleotide–dependent histone deacetylase (sirtuin 1 [SIRT1]).1,58–62 The most powerful autophagy inducer is Kidney International (2016) -, -–-

review

starvation by induction of low levels of amino acids, glucose, adenosine triphosphate, and growth factors. The activated unc-51 like autophagy activating kinase 1 (ULK1) complex (ULK1-ATG13-ATG101–FAK family kinase-interacting protein of 200 kDa) is recruited to the ER,52 where it activates the class III phosphoinositide 3–kinase complex (beclin 1 [BECN1]-ATG14L-Vps34-p150) by phosphorylating BECN1.63 The complex generates phosphatidylinositol 3– phosphate, which recruits effectors to initiate autophagosome formation.64,65 The lipid kinase activity of this phosphoinositide 3–kinase complex I is enhanced upon dimerization provoked by nuclear receptor binding factor 2 binding to Atg14L.66 Membrane elongation, autophagosome maturation, and cargo recruitment depend on two conjugation complexes: the ATG5-ATG12-ATG16L1 complex and the LC3phosphatidylethanolamine complex.40 LC3 and SQSTM1 have primordial roles in cargo selection. LC3 is also essential for autophagosomal membrane closure.67 Nutrition-related regulation of autophagy. Most of the studies on autophagy regulation focus on the role of mTORC1 complex. Mammalian target of rapamycin (mTOR) is a serine/threonine protein kinase whose activity is associated with nutrient levels and redox status. The function of TORC1 as a repressor of autophagy is conserved in mammals and mTORC1 inhibits autophagy in many mammalian cell types.68 It has been shown that mTORC1 interacts with the ULK1 complex and inhibits autophagy induction through phosphorylation of ULK1 and ATG13. Upon starvation, inhibitory phosphorylation of the serine/threonine-protein kinase ULK1 at Ser757 is removed, thereby allowing autophagy induction.58,62,69 The functions of mTORC1 in glomerular physiology have been well documented.70–72 In situations of low energy, cellular adenosine monophosphate levels rise, leading to the activation of AMPK.73 AMPK is ubiquitously expressed and plays a major role in cellular energy homeostasis by sensing levels of intracellular calcium and adenosine triphosphate/adenosine monophosphate ratio. AMPK is strongly expressed in several renal cells including podocytes, mesangial cells, glomerular endothelial cells and tubular cells. AMPK is activated through phosphorylation on Ser72 by the calcium calmodulindependent protein kinase kinase and the tumor suppressor protein serine/threonine kinase 11.74,75 AMPK is an autophagy inducer by phosphorylating ULK1 on Ser317 and Ser777.62 Once activated, AMPK phosphorylates tuberous sclerosis complex 2,76 which forms a complex with tuberous sclerosis complex 1 protein, which acts as a guanosine triphosphatase–activating protein toward Ras homolog enriched in brain.77–79 Additionally, tuberous sclerosis complex 2 phosphorylation by AMPK is a primer for further phosphorylation by glycogen synthase kinase 3 beta, which leads to Ras homolog enriched in brain–dependent inhibition of mTORC1.80 AMPK was also shown to directly phosphorylate raptor.81 However, the relative contribution of raptor phosphorylation to overall mTORC1 regulation by AMPK and the context in which raptor is implicated in mTORC1 3

review

regulation are not well understood. The link between autophagy and AMPK regulation in kidney cells is only incompletely understood.34 The oxidized nicotinamide adenine dinucleotide–dependent histone deacetylase (SIRT1) also regulates autophagy through different targets and effectors.82,83 Sirtuins are a family of histone and nonhistone deacetylases that are regulated by nicotinamide adenine dinucleotide as a cofactor linking their activity to the metabolic status of the cell.84 SIRT1 has been demonstrated to be an inducer of autophagy by directly and indirectly acting on components of the autophagy machinery. SIRT1 coimmunoprecipitates with ATG5, ATG7, and ATG8/LC3 and can deacetylate these proteins in a nicotinamide adenine dinucleotide–dependent manner in vitro.61 Through deacetylation of transcription factors that in turn activate autophagy genes, SIRT1 can also indirectly modulate autophagy. This is the case for members of the forkhead box O family.85,86 Interestingly, SIRT1mediated forkhead box O1 activation was shown to increase expression of Rab7, a small guanosine triphosphatase that mediates late autophagosome-lysosome fusion. Furthermore, forkhead box O3 deacetylation by SIRT1 can upregulate multiple autophagy-related genes such as unc-51 like autophagy activating kinase 2 gene (ULK2), BECN1, phosphatidylinositol 3–kinase catalytic subunit type 3 gene (VPS34), B-cell lymphoma 2 (BCL2)/adenovirus E1B 19kDa interacting protein 3 gene (Bnip3) and BCL2/adenovirus E1B 19kDa interacting protein 3 gene-like (Bnip3L), autophagy related 12 gene (Atg12), autophagy related 4B gene (Atg4B), LC3, and gaminobutyric acid receptor gene (GABAR)–amino acid permase like gene (APL1).87 Recent evidence demonstrated SIRT1 involvement in renal pathophysiology88–93 and the aging kidney (see later). Stress-induced regulation of autophagy (oxidative stress, ER stress, hypoxia). In addition to nutrient starvation, intracel-

lular stress effectors can also modulate autophagy levels. Most of the following evidence has been collected from nonrenal cellular models. For example, reactive oxygen species can influence LC3 production through activation of c-Jun Nterminal protein kinase 194 and activation of protein kinase RNA-like endoplasmic reticulum kinase.95 In the absence of clearly deciphered interaction between ER stress and autophagy in the kidney, we will refer to a recent review of the literature demonstrating that ER stress serves dual roles by favoring both induction and inhibition of autophagy.96 Modulation by hypoxia: interplay with oxidative stress and proapoptotic and metabolic pathways. Evidence has accu-

mulated that hypoxia plays a significant role in the pathogenesis and progression of chronic renal disease.97 Depending on the severity and duration of the variations in oxygen tension, hypoxia has been demonstrated to regulate autophagy pathways differently. For instance, chronic and moderate hypoxia induces autophagy through hypoxiainducible factor-1 alpha and PKCd–c-Jun N-terminal protein kinase 1–implicated pathways.98 Conversely, rapid and 4

O Lenoir et al.: Autophagy in kidney disease and aging

severe oxygen fluctuation could induce autophagy via hypoxia-inducible factor–independent pathways. Hypoxia can also activate autophagy through induction of BCL2/adenovirus E15 19kDa interacting protein 3 (BNIP3),99 which can disrupt the inhibitory interaction between BECN1 and BCL2.100,98,101–103 Under conditions of hypoxia, BNIP3 transcription triggers mitochondrial dysfunction, formation of autophagosomes, and promotion of the apoptosis process in neonatal cardiac myocytes.104 By causing mitochondrial dysfunction, BNIP3 or BCL2/ adenovirus E1B 19kDa interacting protein 3-like may increase production of reactive oxygen species, which can activate autophagy.105 Competition by BNIP3 or BCL2/adenovirus E1B 19kDa interacting protein 3-like for binding to BCL2 (or a related protein) could also liberate BECN1 from BCL2 complexes and activate autophagy.106 In fact, this interplay between proapoptotic pathways and proautophagy pathways may lead to misinterpretation of the role of autophagy. For example, overexpression of Bcl2 homolgy domain 3–only proteins at some stages of kidney diseases may increase autophagy by freeing BECN1 from BCL2, but the net result may still be proapoptotic.107,108 Finally, BNIP3 binds to and inhibits Ras homolog enriched in brain,109 which is an upstream activator of mTOR. Therefore, BNIP3 may activate autophagy by repressing mTOR. Oxygen is also an essential regulator of cellular metabolism affecting protein translation, autophagy, and cell growth through mTOR regulation. In response to hypoxia, regulation of mTORC1 occurs through a conserved pathway that is independent of energy signaling pathways.110,111 Under conditions of oxygen deprivation, mTORC1 dissociates from the ULK1 to phosphorylate ATG13/FAK family kinase-interacting protein of 200 kDa complex and, thus, initiates autophagy in cells.79 Regulation of mTORC1 and autophagy by hypoxia, however, seems to be highly cell type–dependent and context dependent. Another level of chronic hypoxia-dependent regulation is activation of tuberous sclerosis complex 2 protein by the reduction of adenosine triphosphate depletion and upregulation of AMPK phosphorylation.112 Overall, both laboratory evidence and clinical evidence highlight hypoxia as an important inducer of autophagy. Future studies will have to elucidate whether the steep oxygen gradient in the kidney is a physiological modulator of autophagy. AUTOPHAGY IN AGING AND DISEASED KIDNEYS

To date, the roles of autophagy in the diseased kidney have been studied mainly in proximal tubular cells and podocytes. The availability of mouse models genetically targeting the autophagy machinery has facilitated our understanding of the role of autophagy in kidney physiology and pathology (Tables 1, 2, and 3). Autophagy in glomerular aging and diseases

Podocytes have been identified as the most fragile cell types of the glomerulus, and a decrease in the number of podocytes Kidney International (2016) -, -–-

review

O Lenoir et al.: Autophagy in kidney disease and aging

Table 1 | Effects on kidneys of global genetic targeting of autophagy genes

Name

Renal disease Tissue model

Atg3/, Atg5/, Atg7/, Atg9/, Atg16L1/ Beclin1/

Global

N/A

Global

N/A

FIP200/

Global

N/A

Ambra1gt/gt

Global

N/A

Ulk1/

Global None

Atg4C/ Map1lc3/ Gabarap/

Global None Global None Global None

Sirt1þ/-

Global

Map1lc3-GFP

Global None

UUO

IRI CAG-RFP-EGFP-Map1lc3 Global

IRI

Phenotype/renal phenotype

Comments

References

Neonatal lethal

No kidney abnormalities in Atg5-/- pups

113–117

Early embryonic lethal (E7.5) with defects in proamniotic canal closure Embryonic lethal (E13.5–E16.5) because of defective heart and liver development Embryonic lethal (wE14) with defects in neural tube development Viable. Increased reticulocyte number with Kidney not studied delayed mitochondrial clearance Viable. Increased tumorigenesis Kidney not studied None Normal renal histology in young adults None Gabarap(-/-) mice have reduced urinary excretion of P(i), higher Na(þ)-dependent (32)P(i) uptake in BBM vesicles, and increased expression of NaPi-IIa in renal BBM compared with Gabarap(þ/þ) mice. The expression of Na(þ)/H(þ) exchanger regulatory factor 1, an important scaffold for the apical expression of NaPi-IIa, is also increased in Gabarap(-/-) mice Increased oxidative stress, apoptosis and Lower autophagy activity and decreased BNIP3 fibrosis after UUO expression None Demonstrate high level of autophagy in podocytes at basal state None Demonstrate that autophagic activity in the tubule system is low at basal state and increased after IRI None Demonstrate that autophagic activity in the proximal tubule system is increased after IRI

118, 119

120

121

122

123 124 125, 126

88

56, 141

192

127

Atg3, autophagy related 3 gene; Atg4C, autophagy related 4C gene; Atg5, autophagy related 5 gene; Atg7, autophagy related 7 gene; Atg9, autophagy related 9 gene; Atg16L1, autophagy related 16-like 1 gene; BBM, brush border membrane; EGFP, enhanced green fluorescent protein; FIP200, FAK family kinase-interacting protein of 200 kDa gene; GABARAP, g-aminobutyric acid receptor–associated pro; GFP, green fluorescent protein; IRI, ischemia-reperfusion injury; LC3, microtubule-associated protein 1A/1B–light chain 3 gene; LC3B, microtubule-associated protein 1A/1B–light chain 3B gene; N/A, not applicable; RFP, red fluorescent protein; SIRT1, sirtuin 1 gene; ULK1, unc-51 like autophagy activating kinase 1; UUO, unilateral ureteral obstruction.

and/or in their function is directly correlated with disease progression in several renal diseases, such as DN, IgA nephropathy, HIV-associated nephropathies, or FSGS.132–134 Glomerular endothelial cells are also altered in several renal diseases (i.e., primary hypertension, DN, and other chronic kidney diseases). Mechanisms of podocyte and glomerular endothelial cell injuries are incompletely understood, but ER stress, mitochondrial damage, and oxidative stress are implicated in most types of glomerulosclerosis.135–138 Autophagy and glomerular maintenance during aging. Aging induces the accumulation of damaged organelles and mitochondria as well as protein aggregates that are physiological causes of organ dysfunction. The kidney is particularly susceptible to age-related renal damage such as tubular atrophy, interstitial fibrosis, and glomerulosclerosis. Elderly individuals are particularly sensitive to ischemia and toxic stress and show high rates of end-stage renal disease and chronic kidney disease.139,140 In glomeruli, podocytes are terminally differentiated postmitotic cells. Therefore, their capacities for regeneration are limited and they require efficient cellular mechanisms to “clean” themselves from protein aggregates and altered organelles that will accumulate throughout a lifetime. Recent evidence suggests that autophagy might be this cleaning mechanism allowing podocyte maintenance during aging. Kidney International (2016) -, -–-

Unlike other renal and nonrenal cell types, podocytes have a high level of basal autophagy, suggesting that autophagy has important functions in this cell type.56,141 Furthermore, podocyte-specific deletion of Atg5 results in proteinuria, loss of podocytes, and aging-related glomerulosclerosis, thus demonstrating the primordial function of autophagy for podocyte maintenance.141 Importantly, Atg5 knockout podocytes showed the following typical age-related alteration: lipofuscin accumulation, an increase in oxidized proteins, and UB-positive and SQSTM1-positive protein aggregates.141 Concomitant with such results, podocyte-specific Vps34 deletion leads to spontaneous development of glomerulosclerosis, although defective autophagy was not primarily responsible for the severe phenotype observed in Vps34-deficient podocytes.142–144 Finally, pharmacological inhibition of autophagy induces podocyte apoptosis by activating the apoptotic pathway of ER stress.145 Autophagy in FSGS. FSGS is a glomerular disease associated with podocytopathy that is associated with a poor prognosis and evolves easily into end-stage renal disease.146 Several pieces of evidence suggest that autophagy is implicated in podocytopathy during FSGS.141,147–150 In an initial study it was demonstrated that podocyte-specific deficiency of Atg5 (Nphs2.Cre Atg5lox/lox mice) sensitizes mice for the 5

review

O Lenoir et al.: Autophagy in kidney disease and aging

Table 2 | Effects of kidney cell–specific deletion of autophagy genes Name fl/fl

Nphs2-cre Atg5

Nphs2-cre Vps34fl/fl

Tissue

Cre-driven expression

Renal disease model

Podocyte

Nphs2

None

Phenotype/renal phenotype

Age-dependent late-onset glomerulosclerosis. Podocytes loss associated with ubiquinated protein aggregates. LPS-, PAN-, doxorubicin- Increased sensitivity to the development (Adriamycin-), and of proteinuria bortezomib-induced proteinuria STZ-induced DN Increased sensitivity to the development of podocyte injury and apoptosis HFD-induced proteinuria HFD-fed podo-Atg5/ mice develop albuminuria with severe podocyte injury None Massive proteinuria and early lethality. Podocyte degeneration and early-onset glomerulosclerosis

Podocyte

Nphs2

Cdh5-cre Atg5fl/fl

Endothelial

Cdh5

STZ-induced DN

Six2-cre Atg5fl/fl

Podocytes, PEC, tubules

Six2

None

Six2-cre Atg7fl/fl

Podocytes, PEC, tubules

Six2

None

KAP-cre Atg5fl/fl

Proximal tubule

KAP

Cisplatin and IRI FFA-albumin overload

PEPCK-cre Atg7fl/fl Pax8.rtTA;tetO.Cre Atg5fl/fl

Proximal tubule Proximal, distal tubules l and collecting duct

PEPCK Pax8

Cisplatin and IRI None

IRI Ksp-cre Atg5fl/fl

Distal tubule

Ksp

Comments

None

Increased sensitivity to the development of DN Tubulointerstitial damage and FSGS by 4 mo. Lethality by 5 mo FSGS by 4 mo with moderate albuminuria. Less tubular injury than for the Six2-cre Atg5fl/fl mice. Survive until 1 yr Increased sensitivity to AKI Increased sensitivity to the development of tubular lesions Increased sensitivity to AKI Few aberrant ultrastructure in tubular cells 5 mo after autophagy blockade Severe tubular injury. P62 and oxidative injury markers accumulation Normal renal function. P62 accumulation and increased oxidative stress markers

References 141

141

167

178

Abrogated autophagic flux is not causative of the podocyte phenotype

142, 144

167

150

150

202, 230 162

201 192

192

192

AKI, acute kidney injury; Atg5, autophagy related 5 gene; Cdh5, cadherin 5 gene; DN, diabetic nephropathy; FFA, free fatty acid; FSGS, focal and segmental glomerulosclerosis; HFD, high-fat diet; IRI, ischemia-reperfusion injury; KAP, kidney androgen regulated protein gene; Ksp, cadherin 16, KSP-cadherin gene; LPS, lipopolysaccharide; Nphs2, nephrosis 2, idiopathic, steroid-resistant (Podocin) gene; PAN, puromycin aminonucleoside nephrosis; Pax8, paired box 8 gene; PEPCK, phosphoenolpyruvate carboxykinase 2, mitochondrial gene; Six2, SIX2 homeobox 2 gene; STZ, streptozotocin; Vps34, phosphatidylinositol 3– kinase gene.

development of several renal diseases, including FSGS as a general theme of podocyte biology and glomerular disease.141 Podocyte injury was characterized by ER stress, accumulation of oxidized proteins, and altered mitochondria.141 These results could also be confirmed using Six2.Cre Atg5lox/lox mice or Six2.Cre Atg7lox/lox mice.150 The more dramatic phenotype caused by deletion of autophagy throughout the nephron than that reported in models of autophagy deletion in tubular or podocyte compartments solely unravels a strong effect of autophagy on nephron development or integration of physiological cross talk between the distinct parts of the nephron. Autophagy was also found to have protective properties in puromycin-aminonucleoside–treated human podocytes.151 A decreased number of autophagic vacuoles was described in podocytes from patients with FSGS.149 Finally, pharmacological induction of autophagy delays the progression of 6

FSGS. Indeed, Wu et al. found that rapamycin could reduce podocyte injury by inducing autophagy.152 Concomitantly, inhibition of autophagy by 3-methyladenine treatment sensitizes rats to the development of puromycin-aminonucleoside–induced FSGS, whereas rapamycin partially protects podocytes from puromycin-aminonucleoside–induced nephropathy.149 Autophagy in other nondiabetic glomerular diseases. A limited amount of studies brought direct evidence for autophagy in other models of glomerulopathies, principally antibody-mediated glomerular diseases. A major role for B-cell autophagy was shown in a mouse model of lupus nephritis, the spontaneous mutant mouse Y-linked autoimmune accelerator with duplication of the Toll-like receptor 7 (Tlr7) gene. Autophagy in CD19þ cells mediates Toll-like receptor 7–dependent autoimmunity and Kidney International (2016) -, -–-

review

O Lenoir et al.: Autophagy in kidney disease and aging

Table 3 | Effects of kidney cell–specific deletion of potential modulators of autophagy Name

Tissue fl/fl

Cre-driven expression Renal disease model

Nphs2-cre Tsc1

Podocytes

Nphs2

Nphs2-cre Raptorfl/fl

Podocytes

Nphs2

Nphs2.rtTA;tetO Raptorfl/fl

Podocytes

Nphs2

Nphs2-cre mTorfl/fl

Podocytes

Nphs2

Ksp-cre Raptorfl/fl

Distal tubule

Ksp

Pax8.rtTA;tetO.Cre Raptorfl/fl

Pax8

Nphs2-cre Sirt1fl/fl

Proximal, distal, and collecting duct tubules Podocytes

Nphs2

Nphs2-cre Sirt1fl/fl

Podocytes

Nphs2

Proximal tubule

Npt2

Sirt1Tg (Npt2)

g-GT-cre Sirt1fl/fl

Proximal tubule

g-GT

None

Phenotype/renal phenotype

Comments

DN-like phenotype: glomerular No link with autophagy membrane thickening, established mesangial expansion, podocyte loss, and proteinuria None Increased proteinuria by 4 wk of No link with autophagy age. established Podocyte degeneration and progressive glomerulosclerosis by 12 mo None Glomerular synechiae No link with autophagy established None Massive proteinuria by 4 wk mTOR inhibition of age. disrupts the Podocyte degeneration and autophagic pathway in early-onset glomerulosclerosis podocytes in vitro None Defective concentrating No link with autophagy mechanisms established and polyuria IRI More severe cell sloughing, cell No link with autophagy flattening, established and distal tubular cast formation db/db DN Increased urinary proteins No link with autophagy established NTS-induced RPGN, Increased sensitivity to podocyte No link with autophagy PS-induced podocyte injury established injury cisplatin-induced AKI Less oxidative stress and No link with autophagy decreased established cisplatin-induced renal injuries db/db DN and 5/6 Proximal tubule-specific Sirt1 No link with autophagy Nphx overexpression alleviates established diabetic albuminuria in db/db mice but not in 5/6th nephrectomized STZ-induced DN Increased sensitivity to DN No link with autophagy established

References 72

71

71

128

129

130

131

89

89

90

AKI, acute kidney injury; DN, diabetic nephropathy; IRI, ischemia-reperfusion injury; Ksp, cadherin 16, KSP-cadherin gene; mTOR, mammalian target of rapamycin; Nphs2, nephrosis 2, idiopathic, steroid-resistant (Podocin) gene; Nphx, nephrectomy; NTS, nephrotoxic serum; Pax8, paired box 8 gene; PS, protamine sulfate; RPGN, rapid and progressive glomerulonephritis; Sirt1, sirtuin 1; STZ, streptozotocin; Tsc1, tuberous sclerosis 1 gene.

inflammation because lupus nephritis development in Tlr7 transgenic (Tg) mice with B-cell–specific ablation of autophagy (Cd19-Cre Atg5lox/lox) was markedly alleviated compared with mice without B-cell–specific autophagy deficiency.153 By contrast, the role of autophagy in other cell types is protective in models of antibody-mediated nephritis, such as lupus nephritis and nephrotoxic serum nephritis. Amino acid metabolism triggers interferon-g–mediated induction of indoleamine 2,3-dioxygenase 1 enzyme activity with subsequent activation of a stress response dependent on the eIF2alpha kinase general control nonderepressible 2. Consistent with a role in prevention of apoptotic cell driven autoreactivity, myeloid deletion of general control nonderepressible 2 in lupus-prone mice resulted in increased immune cell activation, humoral autoimmunity, and renal disease.154 Interestingly, the indoleamine 2,3-dioxygenase– general control nonderepressible 2 pathway in podocytes and other cells was suggested to be a critical negative feedback Kidney International (2016) -, -–-

cascade that limits inflammatory renal injuries by inducing autophagy in a model of nephrotoxic serum nephritis.155 Increasing kidney indoleamine 2,3-dioxygenase 1 activity or treating mice with a general control nonderepressible 2 agonist induced autophagy and protected mice from nephritic kidney damage. Interestingly, kidneys from patients with antibody-driven nephropathy showed increased indoleamine 2,3-dioxygenase 1 abundance and stress gene expression, suggesting interplay between autophagy and interferong–mediated protective feedback cascade.155 Autophagy in DN. DN is the leading cause of end-stage renal disease in industrialized countries. Progressive loss of podocytes and microvascular alterations appear to most closely correlate with the functional renal decline in DN.156–158 Progressive podocyte injury characterized by cell detachment, hypertrophy, and foot process effacement plays a central role in the development of DN in both type 1 and type 2 diabetes mellitus.132,158–160 Over the past decade, several studies have provided indirect and direct evidence that 7

review

autophagy is implicated in the pathogenesis of DN.34,161,162 First, autophagy can be induced by high glucose levels in various cell types, partly through hyperglycemia-mediated production of reactive oxygen species, and it has protective effects in vitro.163–165 In podocytes, high glucose levels lead to podocyte apoptosis in vitro that is mediated through caspase-3 activation.166 High glucose concentration also induces autophagy.165,167,168 However, prolonged exposure to hyperglycemia results in depressed podocyte autophagy.169 This secondary downregulation of autophagy may play a role in disease progression and may be linked to podocyte hypertrophy through the upregulation of the cyclindependent kinase inhibitor p27/Kip1170 and activation of mTOR.71,72 Furthermore, activation of mTOR was associated with an accelerated glomerular injury in DN.71,152 Although rapamycin, resveratrol, and caloric restriction treatments affect many pathways besides the autophagy pathway, several studies have shown that they have positive effects on inflammation, tubular injury, glomerulosclerosis, and podocyte injury in rodent models of DN.93,171–177 Taurineconjugated ursodeoxycholic acid also restores high-glucose– suppressed autophagy and podocin expression in podocyte culture, and it attenuates albuminuria in diabetic mice.168 Direct evidence for the protective role of autophagy during DN is coming from recent work demonstrating that podocyte Atg5 deficiency increased high-glucose–induced apoptosis. Moreover, specific podocyte autophagy deficiency (in Nphs2.Cre Atg5lox/lox mice) resulted in accelerated diabetesinduced podocytopathy as shown in models of type I167 and type II DN.178 Thus, chronic alteration of podocyte autophagy may contribute to DN progression, and fluctuations of autophagy in the diabetic podocyte over time could be an interesting paradigm for stage-specific interventions. Histone deacetylase 4 may participate to suppression of podocyte autophagy.179 Expression level of histone deacetylase 4 was shown to negatively correlate with estimated glomerular filtration rate in individual patients with DN and was upregulated in kidney tissues of rats with streptozotocininduced diabetes as well as in diabetic db/db mice. A pattern of histone deacetylase 4 expression was prominent in cultured podocytes upon incubation with a high concentration of glucose and in podocytes from individuals with DN or FSGS. Histone deacetylase 4 inhibition sustained the high constitutive autophagy in podocytes. Alteration of proteostasis and autophagy in the tubular compartment also takes place in diabetic kidneys and in cells exposed to a high concentration of glucose or advanced glycation end products. An early work investigating the activity of chaperone-mediated autophagy in rat kidney cortical lysates upon induction of diabetes revealed that reduced chaperone-mediated autophagy contributes to accumulation of specific proteins in diabetic-induced renal hypertrophy.180 The implication of such changes in DN progression is unknown. A recent work demonstrated that ubiquitinated protein bodies were delivered to the autophagy machinery but 8

O Lenoir et al.: Autophagy in kidney disease and aging

not degraded by the impaired lysosomes of HK-2 cells exposed to advanced glycation end product–bovine serum albumin.181 Again, this study suggests that the autophagylysosome pathway is disrupted by advanced glycation end products in tubular epithelial cells during the development of DN, which results in the accumulation of abnormal proteins. Progressive and irreversible microvascular damage and loss are also observed in the diabetic kidney, and a decrease in the function and density of intrarenal microvessels has been reported in several studies.182–185 Interestingly, the treatment of diabetic mice with resveratrol, a well-known inducer of autophagy, protects from intraglomerular capillary rarefaction, suggesting that resveratrol attenuates DN by modulating angiogenesis.186 Moreover, autophagy protects from high glucose–induced senescence in cord human venous endothelial cells.187 Finally, endothelial-specific deletion of Atg5 (Cdh5.Cre Atg5lox/lox mice) leads to increased diabetes-induced renal endothelial injury, as shown by interstitial and glomerular capillary rarefactions and glomerular endothelial cells’ loss of fenestrations.167 Interestingly, epidermal growth factor receptor activation detected in diabetic glomeruli was associated with ER stress, decreased autophagy, and accelerated diabetic glomerulopathy.169 Therefore, pharmacologic epidermal growth factor receptor kinase inhibition may have therapeutic implications in DN.169 In summary, autophagy acts as a protective mechanism on both cellular layers of the glomerular filtration barrier in DN. Autophagy in the tubulointerstitial compartment Autophagy in AKI. Most cases of AKI are caused by renal

ischemia-reperfusion, sepsis, xenobiotic agents, drugs, and metals, which are risk factors for development of chronic kidney diseases.188,189 Autophagy is now generally accepted as a renoprotective cellular response in AKI of various causes.36,190–203 An in-depth review of the relationship between autophagy and AKI has been published in this journal.204 Autophagy in renal fibrosis. Unilateral ureteral obstruction (UUO) in rodents has been established as a classic model of progressive renal fibrosis.205 Autophagy markers were increased in obstructed renal tubules in a UUO model. Indeed, BECN1 and LC3-II are overexpressed during the time course of renal injury and accompany tubular atrophy and tubular cell death.77,206,207 Nevertheless, the positive or negative impact of autophagy induction in UUO is not clear. The autophagy inhibitor 3-methyladenine enhances tubular cell apoptosis and tubulointerstitial fibrosis in the obstructed kidney in rats after ureter obstruction, suggesting that autophagy induced by UUO has a renoprotective role in the obstructed kidney.77 The protective function of autophagy in tubules during ureteral obstruction was also demonstrated recently by the use of genetically modified animals deficient for the autophagy pathway.207 In renal tubular epithelial cells, autophagy regulates expression of transforming growth factor beta (TGF-b) by promoting autophagic degradation of Kidney International (2016) -, -–-

O Lenoir et al.: Autophagy in kidney disease and aging

mature TGF-b. Deletion of LC3B (in LC3-/- mice) or Beclin1 haploinsufficiency (in Bcn1þ/- mice) resulted in increased levels of mature TGFbeta and was associated with more severe fibrosis and renal tubular epithelial cell apoptosis in the obstructed kidney after UUO. Interestingly, TGF-b itself stimulates autophagy.207 On the other hand, it was found that under the stimulus of persistent UUO, the induction of autophagy in the proximal tubules contributes to cell death. Indeed, the authors found that mitochondrial structure changes during UUO and lipid peroxidation product, reduced nicotinamide adenine dinucleotide phosphate oxidase 4, and reduced nicotinamide adenine dinucleotide phosphate oxidase activity are increased in the obstructed renal cortex, whereas mitochondrial injury leads to autophagy and apoptosis through the BECN1 pathway and Bcl2 interaction. Therefore, during UUO, oxidative stress that leads to mitochondrial damage could drive autophagy-dependent cell death and apoptosis and could be a mechanism of tubular atrophy.208 The sequestration of damaged lysosomes by autophagy is another emerging role for autophagy in epithelial maintenance. This involves complex process of selective lysosome recognition, engulfment, and degradation or so-called lysophagy.209 This proved to be the case in tubular kidney cells, in which under conditions of lysosomal damage, loss of autophagy caused inhibition of lysosomal biogenesis in vitro and deterioration of AKI in vivo.210 LC3-associated phagocytosis in tubular cells. Although several studies suggested that cells from proximal tubules play a significant role in removing apoptotic cells and debris via phagocytosis or heterophagy, recent evidence suggests that proximal tubular cells also exert noncanonical autophagy in cell corpse clearance, termed LC3-associated phagocytosis. In this clearance pathway, LC3 is targeted to the phagosome in an ATG5-, ATG7-, and BECN1-dependent manner; however, this process does not classically involve the engagement of upstream autophagy proteins such as ULK1.211–213 A recent study found that kidney injury molecule-1 (KIM-1)/T-cell Ig and mucin domain 1, a phosphatidylserine phagocytosis and scavenger receptor expressed by kidney proximal tubule epithelial cells, is the dominant apoptotic cell phagocytosis receptor in these cells.214 Originally, upon ligand extracellular recognition, binding and phosphorylation of KIM-1 cytosolic domain interacted with the phosphoinositide 3–kinase pathway and stimulated LC3 lipidation and autophagy induction with progressive appearance of KIM-1-and LC3-positive organelles characteristic of autophagy. Unlike standard LC3-associated phagocytosis described in professional phagocytic cells, KIM1 expression induced phosphorylation and activation of ULK1/ATG1, which is essential for autophagosome formation. These results also suggest that KIM-1 phagosomes are degraded by autophagy. In summary, this cascade described in kidney proximal epithelial cells represents a novel mechanism by which autophagosomes generated in an ULK1-dependent manner target phagosomes for degradation, which has potential pathophysiological implications as kidney injury in mice Kidney International (2016) -, -–-

review

carrying a nonfunctional mutant KIM-1 results in a worsening of AKI with the accumulation of unphagocytosed apoptotic cells and increased inflammation.214,215 Beyond the case of AKI, we hypothesize that interplay between KIM-1–mediated phagocytosis, tubular autophagy, and inflammation may play a role in resolution of inflammation, scarring processes, and progression of chronic nephropathies. Autophagy and metabolic acidosis. Metabolic acidosis induces autophagy in proximal tubular cells in vitro and in vivo. Kidneys of acid-loaded proximal tubule–specific autophagydeficient mice (Atg5lox/lox:KAP-Cre mice) exhibited an increased number of SQSTM1/p62 and UB-positive aggregates in the kidney proximal tubular cells compared with in the vehicle-treated Atg5lox/lox:KAP-Cre mice. Furthermore, acid-loaded Atg5lox/lox:KAP-Cre mice showed significantly more severe acidosis than did the acid-loaded controls, with impaired ammoniagenesis indicating a protective role for ATG5 on ammoniagenesis. This was linked to deficiency in proper mitochondrial functions, including ammoniagenesis in Atg5lox/lox:KAP-Cre animals, as basal mitochondrial oxygen consumption rate, the adenosine triphosphate–linked oxygen consumption rate, and maximum respiratory function were significantly more reduced by acid loading in ATG5-deficient proximal tubular cells than in ATG5-competent proximal tubular cells.216 Autophagy PKDs. Several pieces of evidence suggest that autophagy could play roles in PKDs in cyst formation and growth. Indeed, electron microscopy, immunofluorescence, and immunoblot demonstrated LC3-II and BECN1 overexpression and increased number of autophagosomes in polycystic kidneys, thus suggesting autophagic flux dysregulation in PKD.217 Moreover, PKD is a cilia-related disease and autophagy was recently shown to be implicated in ciliogenesis,218 whereas primary cilia are required for the activation of autophagy.219 Therefore, the close relationship between ciliogenesis and autophagy constitutes an exciting method of investigation in the field of cilia-related disease such as PKD. Finally, autophagy modulators have been implicated in PKD progression. The PtdIns3K–v-akt murine thymoma viral oncogene homolog 1–tuberous sclerosis complex–mTOR pathway is activated in PKD, and rapamycin has shown proof of therapeutic effects on animal models of PKD220–226 but not in patients.227,228 Interestingly, a low dose of rapamycin results in protection from PKD without effects on apoptosis,225 whereas a higher dose of rapamycin increases apoptosis of cyst-lining epithelial cells in PKD.226 Moreover, it was shown recently that metformin, a known AMPK inducer, slows cyst formation.229 Even if the authors did not establish any link between AMPK activation and autophagy induction, this study suggests that autophagy induction, mediated by metformin, could be renoprotective in PKD. Further studies on the link between autophagy and cyst formation in PKD are required to clarify the functions of autophagy. Autophagy and aging. The renal tubulointerstitial compartment is not spared from aging-induced damage. Tubular atrophy, apoptosis, and fibrosis are associated with a decline 9

review

in the filtration function in the aged kidney. Experimental evidence suggests that autophagy is protective in proximal tubules cells during aging. Indeed, mice with a specific autophagy deficiency in renal proximal tubular cells (Atg5lox/lox;KSP.Cre and Atg5lox/lox; Pax8.rtTA;TetO.Cre mice) have increased accumulation of damaged mitochondria, ubiquitinated proteins, and SQSTM1-positive aggregates, which are associated with tubular apoptosis and tubulointerstitial fibrosis.192,230 These studies showed that autophagy deficiency in renal tubular cells leads to premature aging. Consequently, autophagy activation in the aging kidney would prevent tubular lesions development and protect from age-related renal dysfunction. Calorie restriction is a major inducer of autophagy and has been correlated with increased life span in eukaryotes.231,232 Interestingly, calorie restriction protects from age-related kidney damage233,234 and autophagy underlies calorie restriction–mediated renal protection.233 It was recently demonstrated that SIRT1 mediates forkhead box O3 deacetylation in response to calorie restriction and leads to restoration of autophagy in aged kidneys.233,235 Resveratrol is an antioxidant that occurs naturally in many plant parts and products, such as grapes, berries, red wine, and peanut skins. Resveratrol activates SIRT1, and previous reports have shown that resveratrol can ameliorate several types of renal injury, such as DN, ischemia-reperfusion injury, and UUO among others, in animal models through its antioxidant effect or SIRT1 activation. Therefore, resveratrol may be a promising treatment for preventing age-related renal injury.236 PUTATIVE THERAPEUTIC TARGETS AND CLINICAL IMPLICATIONS

Inducing autophagy could be a promising therapeutic strategy for the treatment of renal diseases. Nevertheless, all the cytoprotective effects of autophagy were found in rodent models of renal diseases, and so far, proof is lacking that these findings can be transferred to human diseases. To date, the only autophagy inducers commonly used in human medicine are mTOR inhibitors (sirolimus and everolimus), which are widely used as immunosuppressive agents. However, it is now well established that in humans these antiproliferative mTOR inhibitors have adverse effects on compensatory renal epithelial cell hypertrophy and recovery from ischemia following renal transplantation.237–239 Moreover, sirolimus is associated with numerous side effects, including diabetes mellitus, systemic inflammation, or edema among others.237 Thus, the use of mTOR inhibitors as autophagy inducers appears to not be a feasible strategy in many cases. Resveratrol supplementation is currently being tested in several clinical trials, essentially for safety of use, for weight loss therapy, and in elderly individuals, and it seems to be safe at the doses tested.240–243 In summary, to date there is a lack of reliable biomarkers and clear interventional strategies targeting autophagy in renal diseases. Thus, a specific approach that enables pathway-specific and kidney-selective regulation of autophagy would be required. Recently, a Tat-Beclin fusion protein has 10

O Lenoir et al.: Autophagy in kidney disease and aging

been found to selectively induce autophagy without side effects on other signaling pathways, offering novel perspectives for potential treatment strategies.244 CONCLUSION

Because autophagy is a dynamic process, instant pictures from cells or tissues cannot bring robust information regarding induction or inhibition of this homeostatic phenomenon. Therefore, information gained from animal models has been precious. The development of a conditional knockout mice enabling targeting of autophagy in rodent models of renal diseases has significantly deepened our knowledge of the role of autophagy in these conditions. Most of the direct studies of autophagy systems in vivo have used loss-of-function approaches rather than specific gain-offunction approaches, which may be seen as a limitation. Meanwhile, most of these studies are in accord regarding a protective function of autophagy in the kidney, highlighting autophagy as a novel and very promising target for renal diseases therapy. However, further studies are required to elucidate the precise and cell-specific programs regulating autophagy in renal diseases. There is also a pressing need for robust biomarkers that accurately assess the clinical utility of modulating autophagy in various disease contexts. Investigations of kidney disease–dependent triggers, sensors, and adaptors that tailor the autophagy machinery might drive biomarker discovery to achieve target specificity. Moreover, selective autophagy modulators and (cell-)specific delivery routes remain to be elucidated to ultimately hopefully offer novel treatment options for our kidney patients. DISCLOSURE

All the authors declared no competing interests. ACKNOWLEDGMENTS

This study was supported by INSERM, Joint Transnational Call 2011 for “Integrated Research on Genomics and Pathophysiology of the Metabolic Syndrome and the Diseases arising from It” from l’Agence Nationale de la Recherche (ANR) of France (PLT); the LefoulonDelalande Foundation; and the Francophone Diabetes Society (to OL). This study was further supported by the German Research Foundation (DFG): CRC 1140 (to TBH) and CRC 992 (to TBH), Heisenberg program (to TBH) and HU 1016/8-1; by the European Research Council-ERC grant 616891 (to TBH) and grant 311255 (to PLT); by the BMBF-Joint transnational grant 01KU1215 (to TBH) and STOP-FSGS 01GM1518C (to TBH); by the Else-Kröner Fresenius Stiftung-NAKSYS; and by the Excellence Initiative of the German Federal and State Governments (GSC-4, Spemann Graduate School and EXC294, BIOSS II to TBH). REFERENCES 1. Boya P, Reggiori F, Codogno P. Emerging regulation and functions of autophagy. Nat Cell Biol. 2013;15:713–720. 2. Choi AM, Ryter SW, Levine B. Autophagy in human health and disease. N Engl J Med. 2013;368:1845–1846. 3. Glick D, Barth S, Macleod KF. Autophagy: cellular and molecular mechanisms. J Pathol. 2010;221:3–12. 4. Clark SL Jr. Cellular differentiation in the kidneys of newborn mice studies with the electron microscope. J Biophys Biochem Cytol. 1957;3: 349–362. 5. De Duve C. The lysosome. Sci Am. 1963;208:64–72. Kidney International (2016) -, -–-

O Lenoir et al.: Autophagy in kidney disease and aging

6. Mikles-Robertson F, Dave C, Porter CW. Apparent autophagocytosis of mitochondria in L1210 leukemia cells treated in vitro with 4,4’-diacetyldiphenylurea-bis(guanylhydrazone). Cancer Res. 1980;40:1054–1061. 7. Novikoff AB, Shin WY. Endoplasmic reticulum and autophagy in rat hepatocytes. Proc Natl Acad Sci U S A. 1978;75:5039–5042. 8. Takeshige K, Baba M, Tsuboi S, et al. Autophagy in yeast demonstrated with proteinase-deficient mutants and conditions for its induction. J Cell Biol. 1992;119:301–311. 9. Tsukada M, Ohsumi Y. Isolation and characterization of autophagydefective mutants of Saccharomyces cerevisiae. FEBS Lett. 1993;333: 169–174. 10. Mizushima N, Noda T, Yoshimori T, et al. A protein conjugation system essential for autophagy. Nature. 1998;395:395–398. 11. Klionsky DJ. Autophagy. Curr Biol. 2005;15:R282–R283. 12. Mijaljica D, Prescott M, Devenish RJ. Microautophagy in mammalian cells: revisiting a 40-year-old conundrum. Autophagy. 2011;7:673–682. 13. Massey AC, Kaushik S, Sovak G, et al. Consequences of the selective blockage of chaperone-mediated autophagy. Proc Natl Acad Sci U S A. 2006;103:5805–5810. 14. Kaushik S, Cuervo AM. Chaperone-mediated autophagy: a unique way to enter the lysosome world. Trends Cell Biol. 2012;22:407–417. 15. Mizushima N, Levine B, Cuervo AM, et al. Autophagy fights disease through cellular self-digestion. Nature. 2008;451:1069–1075. 16. Wang K, Klionsky DJ. Mitochondria removal by autophagy. Autophagy. 2011;7:297–300. 17. Green DR, Galluzzi L, Kroemer G. Mitochondria and the autophagyinflammation-cell death axis in organismal aging. Science. 2011;33: 1109–1112. 18. Schuck S, Gallagher CM, Walter P. ER-phagy mediates selective degradation of endoplasmic reticulum independently of the core autophagy machinery. J Cell Sci. 2014;127:4078–4088. 19. Cianfanelli V, De Zio D, Di Bartolomeo S, et al. Ambra1 at a glance. J Cell Sci. 2015;128:2003–2008. 20. Deosaran E, Larsen KB, Hua R, et al. NBR1 acts as an autophagy receptor for peroxisomes. J Cell Sci. 2013;126:939–952. 21. Dunn WA Jr, Cregg JM, Kiel JA, et al. Pexophagy: the selective autophagy of peroxisomes. Autophagy. 2005;1:75–83. 22. Kim PK, Hailey DW, Mullen RT, et al. Ubiquitin signals autophagic degradation of cytosolic proteins and peroxisomes. Proc Natl Acad Sci U S A. 2008;105:20567–20574. 23. Lamark T, Johansen T. Aggrephagy: selective disposal of protein aggregates by macroautophagy. Int J Cell Biol. 2012;2012:736905. 24. Singh R, Kaushik S, Wang Y, et al. Autophagy regulates lipid metabolism. Nature. 2009;458:1131–1135. 25. Shibata M, Lu T, Furuya T, et al. Regulation of intracellular accumulation of mutant Huntingtin by Beclin 1. J Biol Chem. 2006;281:14474–14485. 26. Yang Z, Klionsky DJ. Eaten alive: a history of macroautophagy. Nat Cell Biol. 2010;12:814–822. 27. Singh R, Xiang Y, Wang Y, et al. Autophagy regulates adipose mass and differentiation in mice. J Clin Invest. 2009;119:3329–3339. 28. Rubinsztein DC, Marino G, Kroemer G. Autophagy and aging. Cell. 2011;146:682–695. 29. Melendez A, Talloczy Z, Seaman M, et al. Autophagy genes are essential for dauer development and life-span extension in C. elegans. Science. 2003;301:1387–1391. 30. Lipinski MM, Zheng B, Lu T, et al. Genome-wide analysis reveals mechanisms modulating autophagy in normal brain aging and in Alzheimer’s disease. Proc Natl Acad Sci U S A. 2010;107: 14164–14169. 31. Jia G, Cheng G, Agrawal DK. Autophagy of vascular smooth muscle cells in atherosclerotic lesions. Autophagy. 2007;3:63–64. 32. Harley JB, Alarcon-Riquelme ME, Criswell LA, et al. Genome-wide association scan in women with systemic lupus erythematosus identifies susceptibility variants in ITGAM, PXK, KIAA1542 and other loci. Nat Genet. 2008;40:204–210. 33. Raychaudhuri S, Thomson BP, Remmers EF, et al. Genetic variants at CD28, PRDM1 and CD2/CD58 are associated with rheumatoid arthritis risk. Nat Genet. 2009;41:1313–1318. 34. Kume S, Thomas MC, Koya D. Nutrient sensing, autophagy, and diabetic nephropathy. Diabetes. 2012;61:23–29. 35. Kume S, Uzu T, Maegawa H, et al. Autophagy: a novel therapeutic target for kidney diseases. Clin Exp Nephrol. 2012;16:827–832. 36. Huber TB, Edelstein CL, Hartleben B, et al. Emerging role of autophagy in kidney function, diseases and aging. Autophagy. 2012;8:1009–1031. Kidney International (2016) -, -–-

review

37. Weide T, Huber TB. Implications of autophagy for glomerular aging and disease. Cell Tissue Res. 2011;343:467–473. 38. Shpilka T, Welter E, Borovsky N, et al. Lipid droplets and their component triglycerides and steryl esters regulate autophagosome biogenesis. EMBO J. 2015;34:2117–2131. 39. Shpilka T, Elazar Z. Lipid droplets regulate autophagosome biogenesis. Autophagy. 2015;11:2130–2131. 40. Yang Z, Klionsky DJ. Mammalian autophagy: core molecular machinery and signaling regulation. Curr Opin Cell Biol. 2010;22:124–131. 41. Birgisdottir AB, Lamark T, Johansen T. The LIR motif—crucial for selective autophagy. J Cell Sci. 2013;126:3237–3247. 42. Johansen T, Lamark T. Selective autophagy mediated by autophagic adapter proteins. Autophagy. 2011;7:279–296. 43. Johansen T, Lamark T. Selective autophagy goes exclusive. Nat Cell Biol. 2014;16:395–397. 44. Mochida K, Oikawa Y, Kimura Y, et al. Receptor-mediated selective autophagy degrades the endoplasmic reticulum and the nucleus. Nature. 2015;522:359–362. 45. Khaminets A, Heinrich T, Mari M, et al. Regulation of endoplasmic reticulum turnover by selective autophagy. Nature. 2015;522:354–358. 46. Khaminets A, Behl C, Dikic I. Ubiquitin-dependent and independent signals in selective autophagy. Trends Cell Biol. 2016;26:6–16. 47. Stolz A, Ernst A, Dikic I. Cargo recognition and trafficking in selective autophagy. Nat Cell Biol. 2014;16:495–501. 48. Eiyama A, Okamoto K. PINK1/Parkin-mediated mitophagy in mammalian cells. Curr Opin Cell Biol. 2015;33:95–101. 49. Lazarou M, Sliter DA, Kane LA, et al. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature. 2015;524: 309–314. 50. Klionsky DJ, Abdalla FC, Abeliovich H, et al. Guidelines for the use and interpretation of assays for monitoring autophagy. Autophagy. 2012;8: 445–544. 51. Mizushima N. Methods for monitoring autophagy. Int J Biochem Cell Biol. 2004;36:2491–2502. 52. Mizushima N. The role of the Atg1/ULK1 complex in autophagy regulation. Curr Opin Cell Biol. 2010;22:132–139. 53. Mizushima N, Levine B. Autophagy in mammalian development and differentiation. Nat Cell Biol. 2010;12:823–830. 54. Mizushima N, Yoshimori T, Levine B. Methods in mammalian autophagy research. Cell. 2010;140:313–326. 55. Kimura S, Fujita N, Noda T, et al. Monitoring autophagy in mammalian cultured cells through the dynamics of LC3. Methods Enzymol. 2009;452: 1–12. 56. Mizushima N, Yamamoto A, Matsui M, et al. In vivo analysis of autophagy in response to nutrient starvation using transgenic mice expressing a fluorescent autophagosome marker. Mol Biol Cell. 2004;15: 1101–1111. 57. Klionsky DJ, Abdelmohsen K, Abe A, et al. Guidelines for the use and interpretation of assays for monitoring autophagy (3rd edition). Autophagy. 2016;12:1–222. 58. Hosokawa N, Hara T, Kaizuka T, et al. Nutrient-dependent mTORC1 association with the ULK1-Atg13-FIP200 complex required for autophagy. Mol Biol Cell. 2009;20:1981–1991. 59. Hoyer-Hansen M, Jaattela M. AMP-activated protein kinase: a universal regulator of autophagy? Autophagy. 2007;3:381–383. 60. Mack HI, Zheng B, Asara JM, et al. AMPK-dependent phosphorylation of ULK1 regulates ATG9 localization. Autophagy. 2012;8:1197–1214. 61. Lee IH, Cao L, Mostoslavsky R, et al. A role for the NAD-dependent deacetylase Sirt1 in the regulation of autophagy. Proc Natl Acad Sci U S A. 2008;105:3374–3379. 62. Kim J, Kundu M, Viollet B, et al. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat Cell Biol. 2011;13:132–141. 63. He C, Levine B. The Beclin 1 interactome. Curr Opin Cell Biol. 2010;22: 140–149. 64. Axe EL, Walker SA, Manifava M, et al. Autophagosome formation from membrane compartments enriched in phosphatidylinositol 3-phosphate and dynamically connected to the endoplasmic reticulum. J Cell Biol. 2008;182:685–701. 65. Polson HE, de Lartigue J, Rigden DJ, et al. Mammalian Atg18 (WIPI2) localizes to omegasome-anchored phagophores and positively regulates LC3 lipidation. Autophagy. 2010;6:506–522. 66. Lu J, He L, Behrends C, et al. NRBF2 regulates autophagy and prevents liver injury by modulating Atg14L-linked phosphatidylinositol-3 kinase III activity. Nat Commun. 2014;5:3920.

11

review

67. Fujita N, Hayashi-Nishino M, Fukumoto H, et al. An Atg4B mutant hampers the lipidation of LC3 paralogues and causes defects in autophagosome closure. Mol Biol Cell. 2008;19:4651–4659. 68. Jung CH, Ro SH, Cao J, et al. mTOR regulation of autophagy. FEBS Lett. 2010;584:1287–1295. 69. Ganley IG, Lam du H, Wang J, et al. ULK1.ATG13.FIP200 complex mediates mTOR signaling and is essential for autophagy. J Biol Chem. 2009;284:12297–12305. 70. Huber TB, Walz G, Kuehn EW. mTOR and rapamycin in the kidney: signaling and therapeutic implications beyond immunosuppression. Kidney Int. 2011;79:502–511. 71. Godel M, Hartleben B, Herbach N, et al. Role of mTOR in podocyte function and diabetic nephropathy in humans and mice. J Clin Invest. 2011;121:2197–2209. 72. Inoki K, Mori H, Wang J, et al. mTORC1 activation in podocytes is a critical step in the development of diabetic nephropathy in mice. J Clin Invest. 2011;121:2181–2196. 73. Hardie DG. Cell biology. Why starving cells eat themselves. Science. 2011;331:410–411. 74. Long YC, Zierath JR. AMP-activated protein kinase signaling in metabolic regulation. J Clin Invest. 2006;116:1776–1783. 75. Shackelford DB, Shaw RJ. The LKB1-AMPK pathway: metabolism and growth control in tumour suppression. Nat Rev Cancer. 2009;9:563–575. 76. Inoki K, Li Y, Zhu T, et al. TSC2 is phosphorylated and inhibited by Akt and suppresses mTOR signalling. Nat Cell Biol. 2002;4:648–657. 77. Kim WY, Nam SA, Song HC, et al. The role of autophagy in unilateral ureteral obstruction rat model. Nephrology. 2012;17:148–159. 78. Inoki K, Li Y, Xu T, et al. Rheb GTPase is a direct target of TSC2 GAP activity and regulates mTOR signaling. Genes Dev. 2003;17:1829–1834. 79. Alers S, Loffler AS, Wesselborg S, et al. Role of AMPK-mTOR-Ulk1/2 in the regulation of autophagy: cross talk, shortcuts, and feedbacks. Mol Cell Biol. 2012;32:2–11. 80. Inoki K, Ouyang H, Zhu T, et al. TSC2 integrates Wnt and energy signals via a coordinated phosphorylation by AMPK and GSK3 to regulate cell growth. Cell. 2006;126:955–968. 81. Gwinn DM, Shackelford DB, Egan DF, et al. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol Cell. 2008;30:214–226. 82. Ng F, Tang BL. Sirtuins’ modulation of autophagy. J Cell Physiol. 2013;228:2262–2270. 83. Lee IH, Yun J, Finkel T. The emerging links between sirtuins and autophagy. Methods Mol Biol. 2013;1077:259–271. 84. Michan S, Sinclair D. Sirtuins in mammals: insights into their biological function. Biochem J. 2007;404:1–13. 85. Huang H, Tindall DJ. Dynamic FoxO transcription factors. J Cell Science. 2007;120:2479–2487. 86. Daitoku H, Sakamaki J, Fukamizu A. Regulation of FoxO transcription factors by acetylation and protein–protein interactions. Biochim Biophys Acta. 2011;1813:1954–1960. 87. Kroemer G, Marino G, Levine B. Autophagy and the integrated stress response. Mol Cell. 2010;40:280–293. 88. He W, Wang Y, Zhang MZ, et al. Sirt1 activation protects the mouse renal medulla from oxidative injury. J Clin Invest. 2010;120:1056–1068. 89. Hasegawa K, Wakino S, Yoshioka K, et al. Kidney-specific overexpression of Sirt1 protects against acute kidney injury by retaining peroxisome function. J Biol Chem. 2010;285:13045–13056. 90. Hasegawa K, Wakino S, Simic P, et al. Renal tubular Sirt1 attenuates diabetic albuminuria by epigenetically suppressing Claudin-1 overexpression in podocytes. Nat Med. 2013;19:1496–1504. 91. Tikoo K, Singh K, Kabra D, et al. Change in histone H3 phosphorylation, MAP kinase p38, SIR 2 and p53 expression by resveratrol in preventing streptozotocin induced type I diabetic nephropathy. Free Radic Res. 2008;42:397–404. 92. Kitada M, Kume S, Imaizumi N, et al. Resveratrol improves oxidative stress and protects against diabetic nephropathy through normalization of Mn-SOD dysfunction in AMPK/SIRT1-independent pathway. Diabetes. 2011;60:634–643. 93. Kitada M, Takeda A, Nagai T, et al. Dietary restriction ameliorates diabetic nephropathy through anti-inflammatory effects and regulation of the autophagy via restoration of Sirt1 in diabetic Wistar fatty (fa/fa) rats: a model of type 2 diabetes. Exp Diabetes Res. 2011;2011:908185. 94. Webber JL. Regulation of autophagy by p38alpha MAPK. Autophagy. 2010;6:292–293.

12

O Lenoir et al.: Autophagy in kidney disease and aging

95. Liu L, Wise DR, Diehl JA, et al. Hypoxic reactive oxygen species regulate the integrated stress response and cell survival. J Biol Chem. 2008;283: 31153–31162. 96. Rashid HO, Yadav RK, Kim HR, et al. ER stress: autophagy induction, inhibition and selection. Autophagy. 2015;11:1956–1977. 97. Eckardt KU, Bernhardt WM, Weidemann A, et al. Role of hypoxia in the pathogenesis of renal disease. Kidney Int Suppl. 2005:S46–S51. 98. Mazure NM, Pouyssegur J. Hypoxia-induced autophagy: cell death or cell survival? Curr Opin Cell Biol. 2010;22:177–180. 99. Bruick RK. Expression of the gene encoding the proapoptotic Nip3 protein is induced by hypoxia. Proc Natl Acad Sci U S A. 2000;97: 9082–9087. 100. Zhang H, Bosch-Marce M, Shimoda LA, et al. Mitochondrial autophagy is an HIF-1-dependent adaptive metabolic response to hypoxia. J Biol Chem. 2008;283:10892–10903. 101. Bellot G, Garcia-Medina R, Gounon P, et al. Hypoxia-induced autophagy is mediated through hypoxia-inducible factor induction of BNIP3 and BNIP3L via their BH3 domains. Mol Cell Biol. 2009;29:2570–2581. 102. Burton TR, Gibson SB. The role of Bcl-2 family member BNIP3 in cell death and disease: NIPping at the heels of cell death. Cell Death Differ. 2009;16:515–523. 103. Qi Y, Tian X, Liu J, et al. Bnip3 and AIF cooperate to induce apoptosis and cavitation during epithelial morphogenesis. J Cell Biol. 2012;198: 103–114. 104. Ma X, Liu H, Foyil SR, et al. Autophagy is impaired in cardiac ischemiareperfusion injury. Autophagy. 2012;8:1394–1396. 105. Scherz-Shouval R, Elazar Z. Regulation of autophagy by ROS: physiology and pathology. Trends Biochem Sci. 2011;36:30–38. 106. Pattingre S, Tassa A, Qu X, et al. Bcl-2 antiapoptotic proteins inhibit Beclin 1-dependent autophagy. Cell. 2005;122:927–939. 107. Maiuri MC, Criollo A, Kroemer G. Crosstalk between apoptosis and autophagy within the Beclin 1 interactome. EMBO J. 2010;29:515–516. 108. Malik SA, Orhon I, Morselli E, et al. BH3 mimetics activate multiple proautophagic pathways. Oncogene. 2011;30:3918–3929. 109. Li Y, Wang Y, Kim E, et al. Bnip3 mediates the hypoxia-induced inhibition on mammalian target of rapamycin by interacting with Rheb. J Biol Chem. 2007;282:35803–35813. 110. DeYoung MP, Horak P, Sofer A, et al. Hypoxia regulates TSC1/2-mTOR signaling and tumor suppression through REDD1-mediated 14-3-3 shuttling. Genes Dev. 2008;22:239–251. 111. Gomez-Pinillos A, Ferrari AC. mTOR signaling pathway and mTOR inhibitors in cancer therapy. Hematol Oncol Clin North Am. 2012;26: 483–505, vii. 112. Wolff NC, Vega-Rubin-de-Celis S, Xie XJ, et al. Cell-type-dependent regulation of mTORC1 by REDD1 and the tumor suppressors TSC1/TSC2 and LKB1 in response to hypoxia. Mol Cell Biol. 2011;31:1870–1884. 113. Kuma A, Hatano M, Matsui M, et al. The role of autophagy during the early neonatal starvation period. Nature. 2004;432:1032–1036. 114. Sou YS, Waguri S, Iwata J, et al. The Atg8 conjugation system is indispensable for proper development of autophagic isolation membranes in mice. Mol Biol Cell. 2008;19:4762–4775. 115. Sou YS, Tanida I, Komatsu M, et al. Phosphatidylserine in addition to phosphatidylethanolamine is an in vitro target of the mammalian Atg8 modifiers, LC3, GABARAP, and GATE-16. J Biol Chem. 2006;281: 3017–3024. 116. Saitoh T, Fujita N, Hayashi T, et al. Atg9a controls dsDNA-driven dynamic translocation of STING and the innate immune response. Proc Natl Acad Sci U S A. 2009;106:20842–20846. 117. Saitoh T, Fujita N, Jang MH, et al. Loss of the autophagy protein Atg16L1 enhances endotoxin-induced IL-1beta production. Nature. 2008;456:264–268. 118. Qu X, Yu J, Bhagat G, et al. Promotion of tumorigenesis by heterozygous disruption of the beclin 1 autophagy gene. J Clin Invest. 2003;112:1809–1820. 119. Yue Z, Jin S, Yang C, et al. Beclin 1, an autophagy gene essential for early embryonic development, is a haploinsufficient tumor suppressor. Proc Natl Acad Sci U S A. 2003;100:15077–15082. 120. Gan B, Peng X, Nagy T, et al. Role of FIP200 in cardiac and liver development and its regulation of TNFalpha and TSC-mTOR signaling pathways. J Cell Biol. 2006;175:121–133. 121. Fimia GM, Stoykova A, Romagnoli A, et al. Ambra1 regulates autophagy and development of the nervous system. Nature. 2007;447:1121–1125.

Kidney International (2016) -, -–-

O Lenoir et al.: Autophagy in kidney disease and aging

122. Kundu M, Lindsten T, Yang CY, et al. Ulk1 plays a critical role in the autophagic clearance of mitochondria and ribosomes during reticulocyte maturation. Blood. 2008;112:1493–1502. 123. Marino G, Salvador-Montoliu N, Fueyo A, et al. Tissue-specific autophagy alterations and increased tumorigenesis in mice deficient in Atg4C/autophagin-3. J Biol Chem. 2007;282:18573–18583. 124. Cann GM, Guignabert C, Ying L, et al. Developmental expression of LC3alpha and beta: absence of fibronectin or autophagy phenotype in LC3beta knockout mice. Dev Dynam. 2008;237:187–195. 125. Reining SC, Gisler SM, Fuster D, et al. GABARAP deficiency modulates expression of NaPi-IIa in renal brush-border membranes. Am J Physiol Renal Physiol. 2009;296:F1118–F1128. 126. O’Sullivan GA, Kneussel M, Elazar Z, et al. GABARAP is not essential for GABA receptor targeting to the synapse. Eur J Neurosci. 2005;22: 2644–2648. 127. Li L, Wang ZV, Hill JA, et al. New autophagy reporter mice reveal dynamics of proximal tubular autophagy. J Am Soc Nephrol. 2014;25: 305–315. 128. Cina DP, Onay T, Paltoo A, et al. Inhibition of MTOR disrupts autophagic flux in podocytes. J Am Soc Nephrol. 2012;23:412–420. 129. Grahammer F, Haenisch N, Steinhardt F, et al. mTORC1 maintains renal tubular homeostasis and is essential in response to ischemic stress. Proc Natl Acad Sci U S A. 2014;111:E2817–E2826. 130. Liu R, Zhong Y, Li X, et al. Role of transcription factor acetylation in diabetic kidney disease. Diabetes. 2014;63:2440–2453. 131. Motonishi S, Nangaku M, Wada T, et al. Sirtuin1 maintains actin cytoskeleton by deacetylation of cortactin in injured podocytes. J Am Soc Nephrol. 2015;26:1939–1959. 132. Ziyadeh FN, Wolf G. Pathogenesis of the podocytopathy and proteinuria in diabetic glomerulopathy. Curr Diabetes Rev. 2008;4:39–45. 133. Lemley KV, Lafayette RA, Safai M, et al. Podocytopenia and disease severity in IgA nephropathy. Kidney Int. 2002;61:1475–1485. 134. Wiggins RC. The spectrum of podocytopathies: a unifying view of glomerular diseases. Kidney Int. 2007;71:1205–1214. 135. Imasawa T, Rossignol R. Podocyte energy metabolism and glomerular diseases. Int J Biochem Cell Biol. 2013;45:2109–2118. 136. Haraldsson B, Nystrom J. The glomerular endothelium: new insights on function and structure. Curr Opin Nephrol Hypertens. 2012;21:258–263. 137. Cybulsky AV. The intersecting roles of endoplasmic reticulum stress, ubiquitin-proteasome system, and autophagy in the pathogenesis of proteinuric kidney disease. Kidney Int. 2013;84:25–33. 138. Bijian K, Cybulsky AV. Stress proteins in glomerular epithelial cell injury. Contrib Nephrol. 2005;148:8–20. 139. Nitta K, Okada K, Yanai M, et al. Aging and chronic kidney disease. Kidney Blood Press Res. 2013;38:109–120. 140. Bolignano D, Mattace-Raso F, Sijbrands EJ, et al. The aging kidney revisited: a systematic review. Ageing Res Rev. 2014;14:65–80. 141. Hartleben B, Godel M, Meyer-Schwesinger C, et al. Autophagy influences glomerular disease susceptibility and maintains podocyte homeostasis in aging mice. J Clin Invest. 2010;120:1084–1096. 142. Bechtel W, Helmstadter M, Balica J, et al. Vps34 deficiency reveals the importance of endocytosis for podocyte homeostasis. J Am Soc Nephrol. 2013;24:727–743. 143. Bechtel W, Helmstadter M, Balica J, et al. The class III phosphatidylinositol 3-kinase PIK3C3/VPS34 regulates endocytosis and autophagosomeautolysosome formation in podocytes. Autophagy. 2013;9:1097–1099. 144. Chen J, Chen MX, Fogo AB, et al. mVps34 deletion in podocytes causes glomerulosclerosis by disrupting intracellular vesicle trafficking. J Am Soc Nephrol. 2013;24:198–207. 145. Fang L, Li X, Luo Y, et al. Autophagy inhibition induces podocyte apoptosis by activating the pro-apoptotic pathway of endoplasmic reticulum stress. Exp Cell Res. 2014;322:290–301. 146. D’Agati VD. The spectrum of focal segmental glomerulosclerosis: new insights. Curr Opin Nephrol Hypertens. 2008;17:271–281. 147. Kang YL, Saleem MA, Chan KW, et al. The cytoprotective role of autophagy in puromycin aminonucleoside treated human podocytes. Biochem Biophys Res Commun. 2014;443:628–634. 148. Asanuma K, Tanida I, Shirato I, et al. MAP-LC3, a promising autophagosomal marker, is processed during the differentiation and recovery of podocytes from PAN nephrosis. FASEB J. 2003;17: 1165–1167. 149. Zeng C, Fan Y, Wu J, et al. Podocyte autophagic activity plays a protective role in renal injury and delays the progression of podocytopathies. J Pathol. 2014;234:203–213. Kidney International (2016) -, -–-

review

150. Kawakami T, Gomez IG, Ren S, et al. Deficient autophagy results in mitochondrial dysfunction and FSGS. J Am Soc Nephrol. 2015;26: 1040–1052. 151. Kang YL, Saleem MA, Chan KW, et al. Trehalose, an mTOR independent autophagy inducer, alleviates human podocyte injury after puromycin aminonucleoside treatment. PloS One. 2014;9:e113520. 152. Wu L, Feng Z, Cui S, et al. Rapamycin upregulates autophagy by inhibiting the mTOR-ULK1 pathway, resulting in reduced podocyte injury. PloS One. 2013;8:e63799. 153. Weindel CG, Richey LJ, Bolland S, et al. B cell autophagy mediates TLR7-dependent autoimmunity and inflammation. Autophagy. 2015;11: 1010–1024. 154. Ravishankar B, Liu H, Shinde R, et al. The amino acid sensor GCN2 inhibits inflammatory responses to apoptotic cells promoting tolerance and suppressing systemic autoimmunity. Proc Natl Acad Sci U S A. 2015;112:10774–10779. 155. Chaudhary K, Shinde R, Liu H, et al. Amino acid metabolism inhibits antibody-driven kidney injury by inducing autophagy. J Immunol. 2015;194:5713–5724. 156. Ziyadeh FN, Hoffman BB, Han DC, et al. Long-term prevention of renal insufficiency, excess matrix gene expression, and glomerular mesangial matrix expansion by treatment with monoclonal antitransforming growth factor-beta antibody in db/db diabetic mice. Proc Natl Acad Sci U S A. 2000;97:8015–8020. 157. Ziyadeh FN. Mediators of diabetic renal disease: the case for tgf-Beta as the major mediator. J Am Soc Nephrol. 2004;15(suppl 1):S55–S57. 158. Najafian B, Alpers CE, Fogo AB. Pathology of human diabetic nephropathy. Contrib Nephrol. 2011;170:36–47. 159. Stitt-Cavanagh E, MacLeod L, Kennedy C. The podocyte in diabetic kidney disease. ScientificWorldJournal. 2009;9:1127–1139. 160. Reddy GR, Kotlyarevska K, Ransom RF, et al. The podocyte and diabetes mellitus: is the podocyte the key to the origins of diabetic nephropathy? Curr Opin Nephrol Hypertens. 2008;17:32–36. 161. Kume S, Yamahara K, Yasuda M, et al. Autophagy: emerging therapeutic target for diabetic nephropathy. Semin Nephrol. 2014;34:9–16. 162. Yamahara K, Kume S, Koya D, et al. Obesity-mediated autophagy insufficiency exacerbates proteinuria-induced tubulointerstitial lesions. J Am Soc Nephrol. 2013;24:1769–1781. 163. Park EY, Park JB. High glucose-induced oxidative stress promotes autophagy through mitochondrial damage in rat notochordal cells. Int Orthop. 2013;37:2507–2514. 164. Chen H, Luo Z, Sun W, et al. Low glucose promotes CD133mAb-elicited cell death via inhibition of autophagy in hepatocarcinoma cells. Cancer Lett. 2013;336:204–212. 165. Ma T, Zhu J, Chen X, et al. High glucose induces autophagy in podocytes. Exp Cell Res. 2013;319:779–789. 166. Susztak K, Raff AC, Schiffer M, et al. Glucose-induced reactive oxygen species cause apoptosis of podocytes and podocyte depletion at the onset of diabetic nephropathy. Diabetes. 2006;55:225–233. 167. Lenoir O, Jasiek M, Henique C, et al. Endothelial cell and podocyte autophagy synergistically protect from diabetes-induced glomerulosclerosis. Autophagy. 2015;11:1130–1145. 168. Fang L, Zhou Y, Cao H, et al. Autophagy attenuates diabetic glomerular damage through protection of hyperglycemia-induced podocyte injury. PloS One. 2013;8:e60546. 169. Zhang MZ, Wang Y, Paueksakon P, et al. Epidermal growth factor receptor inhibition slows progression of diabetic nephropathy in association with a decrease in endoplasmic reticulum stress and an increase in autophagy. Diabetes. 2014;63:2063–2072. 170. Wolf G, Schanze A, Stahl RA, et al. p27(Kip1) Knockout mice are protected from diabetic nephropathy: evidence for p27(Kip1) haplotype insufficiency. Kidney Int. 2005;68:1583–1589. 171. Mori H, Inoki K, Masutani K, et al. The mTOR pathway is highly activated in diabetic nephropathy and rapamycin has a strong therapeutic potential. Biochem Biophys Res Commun. 2009;384:471–475. 172. Sakaguchi M, Isono M, Isshiki K, et al. Inhibition of mTOR signaling with rapamycin attenuates renal hypertrophy in the early diabetic mice. Biochem Biophys Res Commun. 2006;340:296–301. 173. Yang Y, Wang J, Qin L, et al. Rapamycin prevents early steps of the development of diabetic nephropathy in rats. Am J Nephrol. 2007;27: 495–502. 174. Wen X, Wu J, Wang F, et al. Deconvoluting the role of reactive oxygen species and autophagy in human diseases. Free Radic Biol Med. 2013;65: 40–410.

13

review

175. Xu F, Wang Y, Cui W, et al. Resveratrol prevention of diabetic nephropathy is associated with the suppression of renal inflammation and mesangial cell proliferation: possible roles of Akt/NF-kappaB pathway. Int J Endocrinol. 2014;2014:289327. 176. Palsamy P, Subramanian S. Resveratrol protects diabetic kidney by attenuating hyperglycemia-mediated oxidative stress and renal inflammatory cytokines via Nrf2-Keap1 signaling. Biochim Biophys Acta. 2011;1812:719–731. 177. Chang CC, Chang CY, Wu YT, et al. Resveratrol retards progression of diabetic nephropathy through modulations of oxidative stress, proinflammatory cytokines, and AMP-activated protein kinase. J Biomed Sci. 2011;18:47. 178. Tagawa A, Yasuda M, Kume S, et al. Impaired podocyte autophagy exacerbates proteinuria in diabetic nephropathy. Diabetes. 6;65:755–767. 179. Wang X, Liu J, Zhen J, et al. Histone deacetylase 4 selectively contributes to podocyte injury in diabetic nephropathy. Kidney Int. 2014;86:712–725. 180. Sooparb S, Price SR, Shaoguang J, et al. Suppression of chaperonemediated autophagy in the renal cortex during acute diabetes mellitus. Kidney Int. 2004;65:2135–2144. 181. Liu WJ, Shen TT, Chen RH, et al. Autophagy-lysosome pathway in renal tubular epithelial cells is disrupted by advanced glycation end products in diabetic nephropathy. J Biol Chem. 2015;290:20499–20510. 182. Kawagishi T, Matsuyoshi M, Emoto M, et al. Impaired endotheliumdependent vascular responses of retinal and intrarenal arteries in patients with type 2 diabetes. Arterioscler Thromb Vasc Biol. 1999;19: 2509–2516. 183. Ishimura E, Nishizawa Y, Kawagishi T, et al. Intrarenal hemodynamic abnormalities in diabetic nephropathy measured by duplex Doppler sonography. Kidney Int. 1997;51:1920–1927. 184. Eleftheriadis T, Antoniadi G, Pissas G, et al. The renal endothelium in diabetic nephropathy. Ren Fail. 2013;35:592–599. 185. Lindenmeyer MT, Kretzler M, Boucherot A, et al. Interstitial vascular rarefaction and reduced VEGF-A expression in human diabetic nephropathy. J Am Soc Nephrol. 2007;18:1765–1776. 186. Wen D, Huang X, Zhang M, et al. Resveratrol attenuates diabetic nephropathy via modulating angiogenesis. PloS One. 2013;8:e82336. 187. Chen F, Chen B, Xiao FQ, et al. Autophagy protects against senescence and apoptosis via the RAS-mitochondria in high-glucose-induced endothelial cells. Cell Physiol Biochem. 2014;33:1058–1074. 188. Vaidya VS, Ferguson MA, Bonventre JV. Biomarkers of acute kidney injury. Ann Rev Pharmacol Toxicol. 2008;48:463–493. 189. Schrier RW, Wang W, Poole B, et al. Acute renal failure: definitions, diagnosis, pathogenesis, and therapy. J Clin Invest. 2004;114:5–14. 190. Fougeray S, Pallet N. Mechanisms and biological functions of autophagy in diseased and ageing kidneys. Nat Rev Nephrol. 2015;11: 34–45. 191. He L, Livingston MJ, Dong Z. Autophagy in acute kidney injury and repair. Nephron Clin Pract. 2014;127:56–60. 192. Liu S, Hartleben B, Kretz O, et al. Autophagy plays a critical role in kidney tubule maintenance, aging and ischemia-reperfusion injury. Autophagy. 2012;8:826–837. 193. Guan X, Qian Y, Shen Y, et al. Autophagy protects renal tubular cells against ischemia / reperfusion injury in a time-dependent manner. Cell Physiol Biochem. 2015;36:285–298. 194. Periyasamy-Thandavan S, Jiang M, Wei Q, et al. Autophagy is cytoprotective during cisplatin injury of renal proximal tubular cells. Kidney Int. 2008;74:631–640. 195. Yang C, Kaushal V, Shah SV, et al. Autophagy is associated with apoptosis in cisplatin injury to renal tubular epithelial cells. Am J Physiol Renal Physiol. 2008;294:F777–F787. 196. Inoue K, Kuwana H, Shimamura Y, et al. Cisplatin-induced macroautophagy occurs prior to apoptosis in proximal tubules in vivo. Clin Exp Nephrol. 2010;14:112–122. 197. Pallet N, Bouvier N, Legendre C, et al. Autophagy protects renal tubular cells against cyclosporine toxicity. Autophagy. 2008;4:783–791. 198. Kimura T, Takahashi A, Takabatake Y, et al. Autophagy protects kidney proximal tubule epithelial cells from mitochondrial metabolic stress. Autophagy. 2013;9:1876–1886. 199. Chargui A, Zekri S, Jacquillet G, et al. Cadmium-induced autophagy in rat kidney: an early biomarker of subtoxic exposure. Toxicol Sci. 2011;121:31–42. 200. Kimura A, Ishida Y, Wada T, et al. The absence of interleukin-6 enhanced arsenite-induced renal injury by promoting autophagy of tubular

14

O Lenoir et al.: Autophagy in kidney disease and aging

201. 202. 203.

204. 205.

206.

207.

208.

209.

210.

211.

212.

213. 214.

215. 216.

217.

218.

219. 220.

221.

222. 223.

224.

225.

226.

epithelial cells with aberrant extracellular signal-regulated kinase activation. Am J Pathol. 2010;176:40–50. Jiang M, Wei Q, Dong G, et al. Autophagy in proximal tubules protects against acute kidney injury. Kidney Int. 2012;82:1271–1283. Takahashi A, Kimura T, Takabatake Y, et al. Autophagy guards against cisplatin-induced acute kidney injury. Am J Pathol. 2012;180:517–525. Matsumoto T, Urushido M, Ide H, et al. Small heat shock protein beta-1 (HSPB1) is upregulated and regulates autophagy and apoptosis of renal tubular cells in acute kidney injury. PloS One. 2015;10:e0126229. Kaushal G, Shah S. Autophagy in acute kidney injury. Kidney Int. 2016;89:779–791. Chevalier RL, Forbes MS, Thornhill BA. Ureteral obstruction as a model of renal interstitial fibrosis and obstructive nephropathy. Kidney Int. 2009;75:1145–1152. Li L, Zepeda-Orozco D, Black R, et al. Autophagy is a component of epithelial cell fate in obstructive uropathy. Am J Pathol. 2010;176: 1767–1778. Ding Y, Kim SL, Lee SY, et al. Autophagy regulates TGF-beta expression and suppresses kidney fibrosis induced by unilateral ureteral obstruction. J Am Soc Nephrol. 2014;25:2835–2846. Xu Y, Ruan S, Wu X, et al. Autophagy and apoptosis in tubular cells following unilateral ureteral obstruction are associated with mitochondrial oxidative stress. Int J Mol Med. 2013;31:628–636. Kawabata T, Yoshimori T. Beyond starvation: an update on the autophagic machinery and its functions [e-pub ahead of print]. J Mol Cell Cardiol. 2015;15:30143–30147. Maejima I, Takahashi A, Omori H, et al. Autophagy sequesters damaged lysosomes to control lysosomal biogenesis and kidney injury. EMBO J. 2013;32:2336–2347. Sanjuan MA, Dillon CP, Tait SW, et al. Toll-like receptor signalling in macrophages links the autophagy pathway to phagocytosis. Nature. 2007;450:1253–1257. Florey O, Kim SE, Sandoval CP, et al. Autophagy machinery mediates macroendocytic processing and entotic cell death by targeting single membranes. Nat Cell Biol. 2011;13:1335–1343. Kim JY, Zhao H, Martinez J, et al. Noncanonical autophagy promotes the visual cycle. Cell. 2013;154:365–376. Brooks CR, Yeung MY, Brooks YS, et al. KIM-1-/TIM-1-mediated phagocytosis links ATG5-/ULK1-dependent clearance of apoptotic cells to antigen presentation. EMBO J. 2015;34:2441–2464. Yang L, Brooks CR, Xiao S, et al. KIM-1-mediated phagocytosis reduces acute injury to the kidney. J Clin Invest. 2015;125:1620–1636. Namba T, Takabatake Y, Kimura T, et al. Autophagic clearance of mitochondria in the kidney copes with metabolic acidosis. J Am Soc Nephrol. 2014;25:2254–2266. Belibi F, Zafar I, Ravichandran K, et al. Hypoxia-inducible factor-1alpha (HIF-1alpha) and autophagy in polycystic kidney disease (PKD). Am J Physiol Renal Physiol. 2011;300:F1235–F1243. Tang Z, Lin MG, Stowe TR, et al. Autophagy promotes primary ciliogenesis by removing OFD1 from centriolar satellites. Nature. 2013;502:254–257. Pampliega O, Orhon I, Patel B, et al. Functional interaction between autophagy and ciliogenesis. Nature. 2013;502:194–200. Belibi F, Ravichandran K, Zafar I, et al. mTORC1/2 and rapamycin in female Han:SPRD rats with polycystic kidney disease. Am J Physiol Renal Physiol. 2011;300:F236–F244. Shillingford JM, Murcia NS, Larson CH, et al. The mTOR pathway is regulated by polycystin-1, and its inhibition reverses renal cystogenesis in polycystic kidney disease. Proc Natl Acad Sci U S A. 2006;103: 5466–5471. Weimbs T. Regulation of mTOR by polycystin-1: is polycystic kidney disease a case of futile repair? Cell Cycle. 2006;5:2425–2429. Wu M, Wahl PR, Le Hir M, et al. Everolimus retards cyst growth and preserves kidney function in a rodent model for polycystic kidney disease. Kidney Blood Press Res. 2007;30:253–259. Tao Y, Kim J, Schrier RW, et al. Rapamycin markedly slows disease progression in a rat model of polycystic kidney disease. J Am Soc Nephrol. 2005;16:46–51. Zafar I, Ravichandran K, Belibi FA, et al. Sirolimus attenuates disease progression in an orthologous mouse model of human autosomal dominant polycystic kidney disease. Kidney Int. 2010;78:754–761. Shillingford JM, Piontek KB, Germino GG, et al. Rapamycin ameliorates PKD resulting from conditional inactivation of Pkd1. J Am Soc Nephrol. 2010;21:489–497. Kidney International (2016) -, -–-

O Lenoir et al.: Autophagy in kidney disease and aging

227. Walz G, Budde K, Mannaa M, et al. Everolimus in patients with autosomal dominant polycystic kidney disease. N Engl J Med. 2010;363: 830–840. 228. Serra AL, Poster D, Kistler AD, et al. Sirolimus and kidney growth in autosomal dominant polycystic kidney disease. N Engl J Med. 2010;363: 820–829. 229. Takiar V, Nishio S, Seo-Mayer P, et al. Activating AMP-activated protein kinase (AMPK) slows renal cystogenesis. Proc Natl Acad Sci U S A. 2011;108:2462–2467. 230. Kimura T, Takabatake Y, Takahashi A, et al. Autophagy protects the proximal tubule from degeneration and acute ischemic injury. J Am Soc Nephrol. 2011;22:902–913. 231. Colman RJ, Anderson RM, Johnson SC, et al. Caloric restriction delays disease onset and mortality in rhesus monkeys. Science. 2009;325: 201–204. 232. Testa G, Biasi F, Poli G, et al. Calorie restriction and dietary restriction mimetics: a strategy for improving healthy aging and longevity. Curr Pharm Des. 2014;20:2950–2977. 233. Kume S, Uzu T, Horiike K, et al. Calorie restriction enhances cell adaptation to hypoxia through Sirt1-dependent mitochondrial autophagy in mouse aged kidney. J Clin Invest. 2010;120:1043–1055. 234. McKiernan SH, Tuen VC, Baldwin K, et al. Adult-onset calorie restriction delays the accumulation of mitochondrial enzyme abnormalities in aging rat kidney tubular epithelial cells. Am J Physiol Renal Physiol. 2007;292:F1751–F1760. 235. Kitada M, Kume S, Koya D. Role of sirtuins in kidney disease. Curr Opin Nephrol Hypertens. 2014;23:75–79.

Kidney International (2016) -, -–-

review

236. Kitada M, Koya D. Renal protective effects of resveratrol. Oxid Med Cell Longev. 2013;2013:568093. 237. Pallet N, Legendre C. Adverse events associated with mTOR inhibitors. Expert Opin Drug Saf. 2013;12:177–186. 238. McTaggart RA, Gottlieb D, Brooks J, et al. Sirolimus prolongs recovery from delayed graft function after cadaveric renal transplantation. Am J Transplant. 2003;3:416–423. 239. Letavernier E, Bruneval P, Mandet C, et al. High sirolimus levels may induce focal segmental glomerulosclerosis de novo. Clin J Am Soc Nephrol. 2007;2:326–333. 240. Macedo RC, Vieira A, Marin DP, et al. Effects of chronic resveratrol supplementation in military firefighters undergo a physical fitness test—a placebo-controlled, double blind study. Chem Biol Interact. 2015;227:89–95. 241. Hobbs T, Caso R, McMahon D, et al. A novel, multi-ingredient supplement to manage elevated blood lipids in patients with no evidence of cardiovascular disease: a pilot study. Altern Ther Health Med. 2014;20:18–23. 242. Chow HH, Garland LL, Heckman-Stoddard BM, et al. A pilot clinical study of resveratrol in postmenopausal women with high body mass index: effects on systemic sex steroid hormones. J Transl Med. 2014;12:223. 243. Anton SD, Embry C, Marsiske M, et al. Safety and metabolic outcomes of resveratrol supplementation in older adults: results of a twelve-week, placebo-controlled pilot study. Exp Gerontol. 2014;57:181–187. 244. Shoji-Kawata S, Sumpter R, Leveno M, et al. Identification of a candidate therapeutic autophagy-inducing peptide. Nature. 2013;494:201–206.

15