YJMCC-07763; No. of pages: 9; 4C: Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Journal of Molecular and Cellular Cardiology journal homepage: www.elsevier.com/locate/yjmcc

1

Review article

4Q1

Stephanie E. Wohlgemuth a,⁎, Riccardo Calvani b, Emanuele Marzetti c a

8

a r t i c l e

9 10 11 12 13

Article history: Received 17 September 2013 Received in revised form 8 March 2014 Accepted 10 March 2014 Available online xxxx

14 15 16 17 18 19 20

Keywords: Autophagy Mitophagy Mitochondrial dynamics Quality control Hydrogen sulfide Cardioprotection

i n f o

a b s t r a c t

E

D

P

Aging is accompanied by a progressive increase in the incidence and prevalence of cardiovascular disease (CVD). Prolonged exposure to cardiovascular risk factors, together with intrinsic age-dependent declines in cardiac functionality, increases the vulnerability of the heart to both endogenous and exogenous stressors, ultimately enhancing the susceptibility to developing CVD in late life. Both increased levels of oxidative damage and the accumulation of dysfunctional mitochondria have been observed in a wide range of cardiac diseases, which may therefore represent a common ground upon which many aspects of CVD develop. In this review, we summarize the current knowledge on the mechanisms whereby oxidative stress arising from mitochondrial dysfunction is involved in the process of cardiac aging and in the pathogenesis of CVD highly prevalent in late life (e.g., heart failure and ischemic heart disease). Special emphasis is placed on recent evidence about the role played by alterations in cellular quality control systems, in particular autophagy/mitophagy and mitochondrial dynamics (fusion and fission), and their interconnections in the context of age-related CVD. Cardioprotective interventions acting through the modulation of mitochondrial autophagy (calorie restriction, calorie restriction mimetics, and the gasotransmitter hydrogen sulfide) are also presented. This article is part of a Special Issue entitled ‘Cardiac Protein Quality Control’. © 2014 Published by Elsevier Ltd.

E

35 39 37 36

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mitochondrial ROS production and targets . . . . . . . . . . . . . . . . . . . . . . Contribution of mitochondrial dysfunction and oxidative stress to cardiac aging . . . . . Mitochondrial dysfunction and cardiac pathology . . . . . . . . . . . . . . . . . . . Cellular and mitochondrial quality control . . . . . . . . . . . . . . . . . . . . . . 5.1. Mitochondrial autophagy: the degradative effector in mitochondrial QC . . . . . 5.2. Redox regulation of macroautophagy . . . . . . . . . . . . . . . . . . . . . 5.3. Regulation of selective autophagic degradation of mitochondria . . . . . . . . . 5.4. Mitochondrial dynamics as an essential component of mitochondrial quality control 6. Autophagy in the heart and its role in cardiac aging and pathology . . . . . . . . . . . 6.1. Role of autophagy in cardiac aging . . . . . . . . . . . . . . . . . . . . . . 6.2. Role of autophagy in the pathogenesis of heart diseases . . . . . . . . . . . . . 7. Conclusions and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . Disclosures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

N C O

1. 2. 3. 4. 5.

R

Contents

U

42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57

R

38 41 40

R O

Department of Animal Sciences, University of Florida, Gainesville, FL 32611, USA Institute of Crystallography, National Research Council (CNR), Bari 70126, Italy Department of Geriatrics, Neurosciences and Orthopedics, Teaching Hospital “Agostino Gemelli”, Catholic University of the Sacred Heart School of Medicine, Rome 00168, Italy

C

c

T

b

O

5 6 7

F

3

The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology

2

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

58

⁎ Corresponding author at: Department of Animal Sciences, IFAS/College of Agriculture and Life Science, University of Florida, Gainesville, FL 32611, USA. Tel.: +1 352 392 7563. E-mail address: steffiw@ufl.edu (S.E. Wohlgemuth).

http://dx.doi.org/10.1016/j.yjmcc.2014.03.007 0022-2828/© 2014 Published by Elsevier Ltd.

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

21 22 23 24 25 26 27 28 29 30 31 32 33 34

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

2. Mitochondrial ROS production and targets

88 89

Mitochondria are central to cellular redox-dependent processes, as they are both the main source of ROS and critical responders to ROSmediated changes in the cellular redox state. Most cellular ROS are partially reduced forms of molecular oxygen (O2) and their derivatives, and originate from the one-electron reduction yielding superoxide anion (O•− 2 ). This incomplete reduction occurs as a result of electron leakage during normal respiration at ETC. complexes or by the enzymatic reduction of O2 [6]. Several sites have been identified within mitochondria that can produce O•− 2 [7]. Estimates indicate that 70–80% of mitochondrial O•− 2 arise from the Q cycle (ubiquinol, QH2, to ubiquinone, Q) as part of the electron transfer to cytochrome c, catalyzed by complex III [8]. Others suggest that the majority of mitochondrial O•− 2 occurs at complex I during NADH oxidation to NAD. Interestingly, only site IIIQo (at complex III) and glycerol 3-phosphate dehydrogenase can release O•− 2 into the intermembrane space (IMS). ROS released into the IMS theoretically have easier access to the cytosol, as they only need to cross the outer mitochondrial membrane (OMM), while matrix ROS have to pass across both the inner mitochondrial membrane (IMM) and the OMM [9]. Thus, site IIIQo-derived O•− 2 may possess an advantage with regard to cytosolic signaling capacity. The quantity of mROS produced is tightly regulated at several levels. Matrix O•− 2 is converted to hydrogen peroxide (H2O2) by superoxide dismutase 2 (SOD2). H2O2 can diffuse through both IMM and OMM to access the cytosol or be converted to H2O by glutathione peroxidases (glutathione peroxidase 1, Gpx1) and peroxiredoxins 3 and 5 (Prx3 and Prx5). O•− 2 released in the IMS can exit the mitochondria through voltage-dependent anion channels and be converted into H2O2 by cytosolic SOD1 [10]. The cytosol also contains glutathione peroxidases, catalase and peroxiredoxins that can reduce H2O2 to H2O [11,12]. Very recently, Gauthier et al. [13] developed a computational model to better understand the mechanisms underlying the complex balance between mROS production and scavenging in cardiac mitochondria. The model was carefully validated against a range of independent experimental data to explore how ROS production changes as a function

81 82 83 84

90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121

T

79 80

C

77 78

E

75 76

R

73 74

R

71 72

O

69 70

C

67 68 Q3

N

65 66

U

64

F

87

62 63

O

85 86

The well-known free radical theory of aging, first proposed by Harman in 1956, posits that the intracellular production of reactive oxygen species (ROS) is the major determinant of lifespan [1]. The recognition of mitochondria as the main cellular sources of oxidants has prompted a refinement of the theory, usually referred to as the mitochondrial free radical theory of aging (MFRTA) [2]. According to this view, ROS generated by mitochondria would trigger a “vicious cycle” by primarily damaging mitochondrial DNA (mtDNA), which results in the synthesis of defective electron transport chain (ETC.) components and further oxidant emission. A large body of evidence exists both in support to and against the MFRTA (reviewed in [3–5]). Despite these controversies, the MFRTA has the undoubted merit of providing a theoretical framework for an enormous amount of work that has led to significant advances in the understanding of cardiovascular aging. Indeed, the role of mitochondrial oxidative stress and mitochondrial dysfunction in age-related cardiovascular pathologies is undisputed. Accumulating evidence also suggests that, in post-mitotic tissues such as the heart, mitochondrial quality control (QC) processes become inefficient in advanced age. As a result, the removal of dysfunctional mitochondria is impaired, which leads to a self-enhancing accumulation of mitochondrial damage with important implications for cardiac aging and pathology. In this review, the effects of aging on cardiac mitochondrial function are discussed, with special emphasis on recent evidence about the role played by alterations in mitochondrial ROS (mROS) production, mitochondrial QC systems and their interconnections in the context of agerelated CVD.

R O

60 61

of respiratory substrate, NAD/NADH redox potential, mitochondrial matrix pH, respiratory state, and inhibition of complex I or III [13]. The authors showed that ROS production was minimal at intermediate redox states, while it increased in both highly reduced/high mitochondrial membrane potential (ΔΨm) environments (e.g., high workload) and highly oxidized conditions (e.g., hypoxia) [13]. Specifically, in highly reduced environments, the increase in mROS levels was due to enhanced production, whereas in highly oxidative environments, it was largely linked to lower scavenging. Which of the two situations is more relevant to aging and CVD has yet to be established. ROS can cause posttranslational protein modifications to regulate signaling pathways. In this context, thiol groups on cysteine residues are proposed to be a major ROS target [14]. Redox modifications may influence mitochondrial function directly or indirectly by acting on mediators involved in most mitochondrial activities (reviewed in [6]). Interestingly, STAT3 appears to preserve ETC. activity by preventing ROS leakage at complex I [15]. Cardiac-specific deletion of STAT3 in mice results in the development of cardiac inflammatory fibrosis, dilated cardiomyopathy, and heart failure with advancing age [16]. Lack of STAT3 also eliminates the protective effects of ischemic preconditioning [17], while its overexpression in cardiomyocytes protects mice against doxorubicin toxicity, which is known to involve mitochondrial dysfunction [18]. The expression of a recombinant form of STAT3 that targets mitochondria and lacks the DNA-binding domain was found to protect the heart from ischemic damage by decreasing mROS production and attenuating cytochrome c release in a model of ischemia [19]. Accumulating evidence suggests that mROS are critically involved in cell homeostasis and exhibit a dual nature [12]. While very high quantities of mROS directly damage proteins, lipids and nucleic acids, lower levels function as signaling molecules to adapt to stress, and even lower amounts of mROS are required for normal cell function [12]. For instance, mROS are crucially involved in cardioprotective preconditioning pathways [20] and, notably, antioxidants render ischemic preconditioning ineffective [21]. The beneficial effects of low (physiological) mROS levels could also help explain the disappointing results of attempts to achieve cardioprotection in humans through antioxidant supplementation [22].

P

1. Introduction

E

59 Q2

D

2

122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158

3. Contribution of mitochondrial dysfunction and oxidative stress to 159 cardiac aging 160 Due to the high reliance of cardiomyocytes on mitochondrial energy metabolism, the heart is exposed to a high burden of mROS throughout the lifetime. Ultrastructural and biochemical analyses of heart tissues of old experimental animals and humans have shown evidence of elevated levels of oxidative damage to proteins, lipids and nucleic acids (reviewed in [23]). Noticeably, aberrant mitochondria, characterized by matrix derangement, loss of cristae and increased ROS generation, are encountered frequently in the myocardium of old rodents [24]. The incidence of the common 4977-bp mtDNA deletion, a typical consequence of oxidative stress, increases with age in the human heart and is estimated to be 5- to 15-fold higher in people over 40 years of age relative to younger individuals [25,26]. Moreover, levels of 8-oxo-7,8dihydro-2′deoxyguanosine in heart mtDNA, but not in nuclear DNA, are negatively correlated with longevity in different mammalian species [27]. Although correlative, these results suggest that mitochondrionmediated oxidative damage could be implicated in the process of heart aging. The mechanistic link among mtDNA mutations, mitochondrial dysfunction and cardiac aging has been provided by the characterization of mice with homozygous mutation in the exonucleaseencoding domain of mtDNA polymerase γ (PolG) [28]. PolG mice exhibit a premature aging phenotype, including cardiac enlargement, fibrosis and impaired systolic and diastolic functions [29]. Ultrastructurally, the heart of PolG mice is characterized by the accumulation of swollen and irregularly shaped mitochondria showing reduced

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

207 208 209 210 211 212

216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249

C

205 206

E

203 204

R

5. Cellular and mitochondrial quality control

282

In recent years, a great deal of research has been devoted to defining the role of QC systems in the context of heart senescence and CVD. Mitochondria possess their own QC systems that ensure the preservation of structure and function of mitochondrial proteins. Chaperones and ATP-dependent proteases are located at the OMM and IMM, as well as within the IMS and the mitochondrial matrix (reviewed in [50,51]). Recently, evidence has emerged for a mitochondrial-lysosomal degradative pathway, in which mitochondrion-derived vesicles (MDVs) transport mitochondrial cargo to the lysosome for degradation [52,53]. MDVs selectively enriched with oxidized proteins were found budding off from mitochondria in response to oxidative stress in different experimental models [52,53]. However, whether this process may function as an early response system to mitochondrial stress and damage in vivo has yet to be determined. The selective degradation of entire mitochondria through macroautophagy, or mitophagy, a term coined by Lemasters [54], may play an essential role in a number of cardiac pathologies associated with mitochondrial dysfunction. Notably, mitophagy is the only mechanism so far identified that sequesters damaged and dysfunctional mitochondria and delivers them to the lysosome for degradation [54]. Numerous studies have investigated mitochondrial QC and removal through mitophagy in a variety of disease states associated with oxidative stress. At the same time, it has become apparent that oxidative stress plays an important role in cell signaling and stress response, including the induction of selective degradation of dysfunctional mitochondria. In the following sections, we discuss about the mechanisms and regulation of macroautophagy, with specific emphasis on the interrelationship between oxidative stress and macroautophagy, the importance of mitochondrial dynamics for the process of mitophagy, the selective autophagic degradation of mitochondria, and the role that macroautophagy may play in CVD and cardiac aging.

283 284

F

201 202

R

199 200

N C O

197 198

U

195 196

O

Increased levels of oxidative damage have been observed in a variety of cardiac diseases, including coronary artery disease, heart failure, (pathological) left ventricular hypertrophy, diabetic cardiomyopathy, and hyperkinetic arrhythmias (reviewed in [20]). As such, oxidative stress resulting from the imbalance between ROS production and scavenging is now considered to be a common denominator marking many aspects of CVD [37]. In cardiac pathological conditions, ROS (and reactive nitrogen species, RNS) are derived from several sources, such as NAD(P)H oxidase, xanthine oxidase, uncoupled nitric oxide (NO) synthase (NOS), lipoxygenase/cyclooxygenase, myeloperoxidase, and autooxidation of various substances, especially catecholamines [37]. However, dysfunctional mitochondria are thought to be the major cellular oxidant producers in the context of CVD. The relevance of mitochondrial dyshomeostasis to cardiac pathophysiology is epitomized by the role played by mitochondria in myocardial damage during ischemia/reperfusion (I/R). Mitochondrial ultrastructural and functional abnormalities are typically observed in ischemic cardiomyocytes. Under oxygen deprivation, mitochondrial oxidative phosphorylation (OXPHOS) is inhibited leading to ATP shortage. Reduced ATP levels, in turn, impair transmembrane ionic homeostasis, most importantly Ca2+ homeostasis, resulting in Ca2 + overload. The latter eventually leads to mPTP opening which causes loss of cardiomyocyte function and viability through ΔΨm dissipation and release of apoptogenic factors [38]. Simultaneously, an excessive mROS production ensues, mainly at complexes I and III of the ETC., which results in extensive oxidative damage, the severity of which is commensurate to the duration of ischemia [39]. The mechanisms by which Ca2+ promotes ROS generation are not well understood. Excess Ca2+ may enhance mROS production, in part, through the activation of the tricarboxyl acid (TCA) cycle, which results in increased NADH formation, stimulation of NOS, and activation of ROS-generating enzymes such as α-ketoglutarate dehydrogenase (reviewed in [40]). A concomitant impairment of endogenous antioxidant defenses increases the extent of oxidative stress [41]. The recovery of mitochondrial function during the reperfusion phase is largely dependent on the duration of the ischemic insult [39]. After prolonged

193 194

R O

215

192

250 251

P

4. Mitochondrial dysfunction and cardiac pathology

190 191

O2 deprivation, extensive damage to ETC. complexes occurs, which, in conjunction with the concomitant disruption of antioxidant defenses, leads to further mROS generation and additional mitochondrial and cardiomyocyte injury during reperfusion. Notably, mPTP opening occurs mostly at the onset of reperfusion, and as a result of ROS accumulation, pH normalization and intracellular Ca2+ concentration rise [42]. Mitochondrial dysfunction and enhanced mROS generation are also involved in the pathogenesis of heart failure regardless of the etiology [43]. Indeed, mitochondria from the failing myocardium generate larger amounts of O•− 2 than those isolated from healthy hearts [44]. The increase in mROS generation was only observed when mitochondria were energized with NADH (substrate for complex I), but not with succinate (substrate for complex II). Excess O•− 2 emission, in turn, was associated with enhanced lipid peroxidation to mitochondrial membranes, decreased mtDNA copy number, reduced abundance of mitochondrial RNA transcripts, and impaired OXPHOS capacity [44]. In addition, oxidative stress impacts cardiomyocyte structure and function by activating signaling pathways involved in myocardial remodeling and failure [44–46]. Impaired mitochondrial function and increased ROS generation have been observed in the hearts of mice with type 2 diabetes mellitus [47, 48]. Interestingly, overexpression of ROS-detoxifying systems (metallothionein, catalase, and MnSOD) reverses mitochondrial dysfunction and diabetic cardiomyopathy, which suggests an important role for mROS in this condition (reviewed in [49]). In conclusion, available data indicate that mitochondrial dysfunction and mROS are likely involved in the pathogenesis of a wide spectrum of CVD. Given the important functions of ROS in physiologic conditions on the one hand, and their implication in CVD on the other, a critical research task will be to determine the “threshold” separating beneficial from detrimental ROS effects and the specific molecular targets that dictate the final outcomes of ROS actions.

D

214

188 189

T

213

activities of ETC. complexes and higher rates of protein carbonylation [28,29]. Remarkably, the PolG heart phenotype, the cardiac mtDNA mutation load and the extent of mitochondrial protein oxidation are partially rescued by the overexpression of human catalase targeted to the mitochondrial matrix [29]. It is also noteworthy that mice with heterozygous disruption of SOD2, albeit showing a normal lifespan, exhibit elevated levels of mitochondrial oxidative damage, impaired respiration by complex I with reduced respiratory control ratio, increased susceptibility to mitochondrial permeability transition pore (mPTP) opening, and enhanced apoptotic DNA fragmentation [30,31]. An important aspect to clarify is discerning the specific alterations experienced by subsarcolemmal mitochondria (SSM) and interfibrillar mitochondria (IFM) with aging. Studies have shown that cardiac IFM isolated from old rats exhibit reduced bioenergetic efficiency and calcium (Ca2 +) retention capacity relative to organelles obtained from younger rodents [32]. The effects of aging on the redox physiology of the two cardiac mitochondrial populations are still disputed. While data by Suh et al. [33] indicate that old IFM, but not SSM produce higher rates of oxidants, others have found either the opposite [34] or no differences [35]. These contrasting findings could originate from the different methods used to quantify mROS production as well as from differential procedural manipulations during isolation [36]. In conclusion, studies in rodent models and observations in humans have made a strong case in support to mitochondrial dysfunction as a contributing factor to cardiac aging. In this context, the buildup of oxidative damage to cardiomyocyte constituents originating from the lifelong exposure to mROS seems to explain many of the alterations observed in mitochondrial function and heart structure/performance during aging.

E

186 187

3

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

252 253 254 Q4 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281

285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 Q5

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

332

The process of macroautophagy can mechanistically be broken down into five steps: induction and nucleation, elongation/expansion (phagophore formation), closure and maturation of the isolation membrane (autophagosome formation), autophagosome-lysosome fusion (autolysosome formation), and lysosomal degradation. The origin of the sequestering double-membrane is debated, with the ER, Golgi apparatus, and mitochondria having been proposed as possible sources [57]. Each of the steps from sequestration to degradation is regulated by autophagy-related (Atg) proteins, at least 35 of which have been identified. As an essential QC system, macroautophagy is under tight regulation and responsive to a variety of stress signals. For a detailed description of the autophagic machinery and its regulation, the reader is referred to comprehensive, specialized reviews (for instance, [58]). It is generally accepted that ROS trigger the induction of macroautophagy [59–63]. Cysteine residues, which are susceptible to oxidative modification [14,64], are especially abundant in a number of autophagy-regulating proteins, which may be important in the redox regulation of this pathway [65,66]. In this regard, Scherz-Shouval et al. [67] proposed that ROS formation (likely H2O2) is essential for the induction of macroautophagy during starvation through oxidation of the cysteine residue 81 near the Atg4 catalytic site. Chen et al. [59] investigated the relative roles of O•− 2 and H2O2 in autophagy regulation in HeLa cells. Both glucose/glutamine/pyruvate/serum and amino acid deprivation induced •− O•− 2 and O2 plus H2O2, respectively, accompanied by increased autophagy and cytoprotection. The authors further determined that O•− 2 was the major signaling molecule in starvation-induced macroautophagy. In contrast, Dutta et al. [68] found that antimycin A-induced O•− 2 production by cardiac mitochondria, although leading to ΔΨm dissipation and oxidative damage to mtDNA, failed to induce macroautophagy or mitophagy. These contrasting findings might reflect cell type-specific redox regulation of autophagy. Surprisingly, despite the important functions of NO in cardiovascular physiology and its reactive nature, the effect of this signaling molecule on autophagy in the heart has not been explored in great detail. Rabkin and Klassen [69] analyzed gene expression in neonatal mouse cardiomyocytes exposed to a NO-donor and found that, while expression of apoptosis-related genes was induced, that of autophagyregulatory genes (Atg5l, Beclin-1 and Apgl2l) was not. On the other hand, NO has been implicated in autophagy induction in a model of cardiac oxidative stress stimulated by LPS (bacterial endotoxin lipopolysaccharide) [70]. Specifically, autophagy was induced in HL-1 cardiomyocytes (neonatal rat cardiomyocytes) and in the heart of mice challenged with LPS, and this effect was attenuated by NOS inhibition. Furthermore, autophagy was stimulated by direct exposure to NO and H2O2, and suppressed by the simultaneous administration of the

339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 Q6 367 368 369 370 371 372 373 374 375 376

C

337 338

E

335 336

R

333 334

R

328 329

O

326 327

C

324 325

N

322 323

U

320 321

5.3. Regulation of selective autophagic degradation of mitochondria

383

The recognition of mitochondria as a major source of and target for ROS and inducers of cell death [23] has stimulated intensive research into the dissection of selective degradation of mitochondria and its triggers [71]. Mild and short-term oxidative stress, induced by rotenone or H2O2, in HeLa cells and mouse embryonic fibroblasts (MEFs) stimulated selective mitophagy rather than non-selective autophagy [60]. In addition, acute bursts of mROS production induced by a mitochondriontargeted photosensitizer followed by mitochondrial depolarization activate Parkin-dependent mitophagy [63]. Increased mROS formation can be associated with mPTP opening [38,72] and loss of ΔΨm [13], both of which have been implicated in the autophagic removal of mitochondria [73]. In isolated rat hepatocytes, nutrient deprivation caused depolarization of mitochondria and their subsequent colocalization with autophagic vacuoles [75]. Inhibition or ablation of cyclophilin D, an integral component of the mPTP [38], preserved ΔΨm and prevented starvation-induced autophagy in rat hepatocytes [75], heart-derived HL-1 cells, neonatal rat cardiomyocytes and adult cardiac cells [74]. Loss of ΔΨm in HEK293 cells treated with the mitochondrial uncoupler carbonyl cyanide mchlorophenylhydrazone (CCCP) was followed by selective degradation of mitochondria [76]. However, it is worth noting that the causative involvement of mPTP opening and ΔΨm dissipation as triggers for mitophagy has been challenged by the demonstration that polarized mitochondria were sequestered by autophagosomes in rat hepatocytes following anoxia/reoxygenation [77]. These conflicting results may suggest the presence of different upstream sensors that trigger mitophagy depending on the type and severity of the applied stress. Recently, several downstream regulatory proteins associated with mitophagy have been identified, shedding light on the selective nature of the autophagic removal of mitochondria. The phosphatase and tensin homolog-induced putative kinase 1 (PINK1)/Parkin pathway plays an integral role in this context (reviewed in [78]). In healthy mitochondria, PINK1, a serine/threonine kinase with a mitochondrial targeting sequence, is imported into the mitochondria and directed to the IMM, where it is cleaved by several proteases and ultimately degraded [79]. ΔΨm dissipation, on the other hand, leads to stabilization of PINK1 at the OMM, where it promotes the recruitment of the cytosolic E3 ubiquitin ligase cytosolic Parkin and its activation [80]. In turn, activated Parkin mediates the engulfment of dysfunctional mitochondria by autophagosomes and their selective elimination [76]. Chen and Dorn [81] recently proposed a mechanism for Parkinrecruitment in cardiomyocytes by which the OMM guanosine triphosphatase mitofusin 2 (Mfn2) was activated by PINK1 and subsequently mediated Parkin recruitment from the cytosol to depolarized mitochondria. Interestingly, a conflicting report demonstrated Parkin recruitment to depolarized mitochondria even in the absence of Mfn2 in cultured fibroblasts [76], which could indicate the presence of alternative mechanisms for Parkin translocation. Once recruited, Parkin activity is induced, a process that is likely also PINK1-dependent [82]. Parkin ubiquitinates mitochondrial proteins, including its receptor Mfn2, and thereby targets depolarized mitochondria for degradation. It is still unclear how ubiquitination of OMM proteins promotes the translocation of the autophagic isolation membrane to the tagged mitochondrion, and different models have been proposed. Adaptor proteins such as sequestosome/p62 (SQSTM1/p62) may recognize the ubiquitin-tag and facilitate the subsequent autophagic sequestration through its binding domain for MAP-LC3 [71,83].

384

F

5.2. Redox regulation of macroautophagy

319

O

331

317 318

R O

330

Macroautophagy is a broad term for processes by which intracellular constituents, including proteins, protein aggregates and organelles, are sequestered within a double-membrane structure (autophagosome), delivered to the lysosome, and degraded upon fusion of the autophagosome with the lysosome (auto(phago)lysosome) [55]. The degradation products, building blocks of macromolecules, are subsequently recycled, used for energy production, or secreted. Besides degrading intracellular material in a random, non-selective manner, macroautophagy has also been shown to selectively target microorganisms (xenophagy), and cellular organelles such as peroxisomes (pexophagy), endoplasmic reticulum (ER, reticulophagy or ERphagy), ribosomes (ribophagy), lipid droplets (lipophagy), protein aggregates (aggrephagy), and mitochondria (mitophagy) [56]. In this capacity, mitophagy becomes an essential cellular QC mechanism through which damaged mitochondria are removed before they cause further harm, for instance by emitting ROS.

377 378

P

315 316

antioxidant N-acetyl-cysteine and the NO-donor or H2O2, respectively [70]. Given the incomplete understanding of redox regulation of macroautophagy, further studies are necessary to explore the role of ROS- and RNS-induced autophagy in the context of cardiac aging and CVD, to possibly identify additional targets for intervention.

D

5.1. Mitochondrial autophagy: the degradative effector in mitochondrial QC

T

314

E

4

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

379 380 381 382

385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 Q7 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477

482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504

C

460 461

E

458 459

R

6.1. Role of autophagy in cardiac aging

517

The post-mitotic nature of cardiomyocytes makes these cells especially prone to the accumulation of macromolecular and organellar oxidative damage over the lifetime [24,108]. In their “Mitochondrial– Lysosomal Axis Theory of Aging”, Brunk and Terman [109] postulated that enlarged mitochondria are a hallmark of aging and are inaccessible to autophagic removal. An age-related decline in autophagic activity may indeed exacerbate the accumulation of harmful cellular “garbage” [110,111]. Lipofuscin, an undegradable polymeric substance, also called the “age-pigment”, is primarily composed of cross-linked protein residues and lipid degradation products [112]. Terman and Brunk [112] portrayed lipofuscin-loaded lysosomes as a sink for newly synthesized lysosomal enzymes, thereby interfering with efficient lysosomal autophagy. An age-related impairment of autophagy has been reported in a variety of mammalian tissues such as the heart [113,114], liver [115], and nervous system [116,117]. The reduced expression of autophagyregulating proteins (Atg9, LC3-II and LAMP-1) observed in the heart of old rats suggests that both autophagosome maturation and processing are impaired in old age [114]. The relevance of autophagy to cardiac aging is supported by the observation that the expression of autophagyregulating proteins declines in parallel with cardiac function in mice [118]. Conversely, the induction of autophagy via rapamycin improves cardiomyocyte function in aged mice [118]. Cardiac-specific inhibition of autophagy through Atg5 knockout in mice results in a shorter lifespan and early development of left ventricular hypertrophy, reduced fractional shortening, and ultrastructural disorganization of sarcomeres [113]. In addition, Atg5-deficient mice are characterized by impairment of mitochondrial respiratory function, appearance of mitochondrial ultrastructure aberrations and accumulation of oxidatively modified mitochondrial proteins early in life [113]. Calorie restriction (CR) is one of the most robust anti-aging and cardioprotective interventions [119], which may, at least partially, act through inducing autophagy [114,120–123]. Indeed, in aged rats that underwent lifelong 40% CR, the increase in cardiac autophagic activity was associated with improved left ventricular diastolic function [124]. A similar CR regimen increased the cardiac autophagic activity in mature mice which was accompanied by reduced left ventricular mass and wall thickness, and improved cardiomyocyte contractile function [125]. Given the questionable feasibility of long-term CR in humans and the uncertainty of its efficacy, especially in the elderly population, the field of CR mimetics has become a topic of increasing scientific interest. As a general definition, CR mimetics are agents or interventions that are capable of reproducing the effects of CR without requiring significant food intake reduction. Since the identification of the first compound (2-deoxy-D-glucose) by Lane and colleagues in 1998 [126], the list of (putative) CR mimetics has grown dramatically [23], although evidence of efficacy and safety is insufficient for many of them. Resveratrol is a CR mimetic that has attracted considerable attention, in part due to its cardioprotective abilities [127]. One salient feature of this compound

518

F

456 457

R

454 455

N C O

452 453

U

450 451

508 509

O

Mitochondria are dynamic organelles that move within the cell and undergo dynamic morphological changes, which are important for the regulation of intracellular energy availability, Ca++ homeostasis, metabolite and ROS/RNS production, and apoptotic signaling [94,95]. For a detailed review of mitochondrial dynamics, the reader is referred to detailed reviews (for example, [96]). In cardiomyocytes, mitochondria have a half-life of 10–25 days [97] and undergo dynamic events both in normal conditions and during disturbance of mitochondrial homeostasis [95]. The molecular machinery that enables these changes in the healthy and diseased heart have recently been reviewed [95,98,99]. Mitochondrial dynamics is also centrally involved in the regulation mitochondrial autophagy [100]. In a seminal study, Twig et al. [101] showed that, in MEFs, mitochondrial segregation through fission and selective fusion is essential for mitochondrial QC and is required for the autophagic removal of dysfunctional and damaged mitochondria. By labeling mitochondria with a photoactivatable dye, individual mitochondria and their fate through fission and fusion events were tracked. In most cases, one of the daughter mitochondria after a fission event maintained normal ΔΨ m . The other daughter mitochondrion was depolarized and exited the fusion–fission cycle, concomitant with decreased fusion capabilities and increased probability of being sequestered for subsequent autophagy [101]. These findings have been corroborated by another study in which overexpression of the fission protein Fis1 and the use of different Fis1 mutants in MEFs

448 449

The role of autophagy in health and disease is multifaceted. Basal levels of autophagy are required for the physiological turnover of proteins and organelles [103,104]. Under conditions of cellular stress, such as starvation, hypoxia or oxidative stress, autophagy is activated and promotes cell survival [105]. Yet, if the insult is too severe, excessive autophagy may lead to impairment of function and ultimately cell death [105]. Given these vital functions, intensive research has been devoted to define the role of autophagy in cardiac aging and diseases [106,107].

R O

481

447

6. Autophagy in the heart and its role in cardiac aging and pathology 507

P

5.4. Mitochondrial dynamics as an essential component of mitochondrial quality control

445 446

D

479 480

443 444

revealed that mitochondrial fission, in conjunction with mitochon- 505 drial dysfunction, triggered mitophagy [102]. 506

T

478

Alternatively, the proteasomal degradation of ubiquitinated OMM proteins or the ubiquitinated proteins themselves may facilitate mitophagy (reviewed in [71]). How do mitochondria tagged for degradation induce the formation of the isolation (autophagosome) membrane for their ultimate sequestration and degradation? Fimia et al. [84] identified a novel player in autophagy regulation, Ambra (activating molecule in Beclin-1-regulated autophagy), and demonstrated in vitro and in vivo in mouse nervous system that Ambra is a binding partner of Beclin-1 and is required for autophagy. In a recent study, Van Humbeeck et al. [85] provided evidence that Ambra is also essential for mitochondrial clearance. The authors found that Ambra is recruited to depolarized mitochondria, where it interacts with Parkin and activates the class III phosphatidylinositol 3-kinase complex, thereby inducing phagophore formation [85]. Dorn et al. [86] found that the mitochondrial protein NIX (also known as BNIP3-like or BNIP3L), a BH3-only Bcl-2 family protein, is essential for mitophagy in erythrocytes [87]. CCCP-induced mitochondrial depolarization in MEFs and HeLa cells stimulated autophagy, Parkin translocation to mitochondria, OMM protein ubiquitination, and p62 recruitment. In contrast, Parkin, ubiquitin and p62 were not required for the stimulation of the autophagy machinery [86]. Interestingly, NIX not only controls Parkin translocation to mitochondria [86], but also possesses a LC3/GABARAP-interacting region that recruits autophagosomal membranes to mitochondria via LC3-binding [88]. The pro-apoptotic BH3-only protein BNIP3 (Bcl-2/adenovirus E1B 19 kDa interacting protein 3) is a mediator and independently implicated in cell death and regulation of mitophagy (reviewed in [89]). Myocardial oxidative stress following I/R led to increased BNIP3 activity and stimulation of cell death, suggesting a redox-sensing role of BNIP3 [90]. However, BNIP3 may play a dual role in the myocardium. Indeed, BNIP3 is proposed to be an important regulator of cardiomyocyte mitochondrial turnover via mitophagy, which may act protectively by inducing the removal of damaged organelles [89]. The tagging of mitochondria for degradation through mitophagy by BNIP3 occurs independent of mPTP opening, ROS production and increase in cytosolic Ca2+, but requires the recruitment of Parkin, the interaction between BNIP3 and LC3, and the involvement of mitochondrial fission [91–93].

E

441 442

5

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

510 511 512 513 514 515 516

519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 Q8 563 564 565 566

574 575

578

The dual nature of autophagy is especially evident in the context of various CVD conditions. For instance, in a model of pressure overload induced by aortic banding, the induction of autophagy accompanied the 581 onset of pathological left ventricular remodeling, which eventually 582 progressed into heart failure [131]. Down-regulation of the autophagic 583 flux by 50% through heterozygous disruption of Beclin-1 reduced 584 heart remodeling, while cardiac Beclin-1 overexpression exacerbated 585 hypertrophy and disease progression [131]. Similarly, moderate pres586 sure overload through transverse aortic constriction (TAC) provoked 587 cardiac hypertrophy in mice through autophagy-dependent mecha588 nisms [132]. Both TAC-induced autophagy and cardiac hypertrophy 589 were attenuated by the inhibition of histone deacetylases. Taken togeth590 er, these findings seem to point to a maladaptive function of autophagy 591 in the setting of pressure overload. In contrast, Oyabu et al. [133] iden592 tified a protective role of autophagy against angiotensin II-induced car593 diac hypertrophy, which was abolished in autophagy-deficient mice. A 594 cardioprotective role of autophagy was also shown in mice harboring 595 a cardiomyocyte-specific deletion of the autophagy-regulator Atg5 596 [134]. In adult mice, temporally controlled cardiac-specific deficiency 597 Q10 of Atg5 caused cardiac hypertrophy, left ventricular dilation and con598 tractile dysfunction, associated with sarcomere disorganization and mi599 tochondrial misalignment and aggregation. In contrast, mice with Atg5 600 deficiency early in cardiogenesis did not show such cardiac phenotype 601 under baseline conditions, but developed heart dysfunction and left 602 ventricular dilation within a week of pressure overload [134]. 603 It is worth noting that the level of autophagy was modulated to 604 different degrees in these studies. The difference between complete 605 autophagic depletion (Atg5 f/f ; MLC2v-Cre + ), reduced autophagy 606 (Beclin-1+/−), and extra-autophagic capacity (Beclin-1 overexpres607 sion) could help explain the varying effects on pathology progression 608 [135]. The fact that Atg5-deficient mice did not exhibit a cardiac pheno609 type under undisturbed conditions might point towards the existence 610 of compensatory mechanisms, which, however, may become inade611 quate under stress conditions. A partially reduced autophagic capacity, 612 on the other hand, might be barely sufficient to maintain contractile 613 function, without allowing for hypertrophic growth. Finally, an exces614 sive induction of autophagy through Beclin-1 overexpression could en615 hance hypertrophic growth and systolic dysfunction. 616 Myocardial I/R injury offers another prominent example of the two617 faced character of autophagy [136]. Early studies showed that upregula618 tion of autophagy during I/R correlated with functional recovery in 619 isolated rabbit hearts, while protracted ischemia was associated with 620 impaired autophagy and subsequent myocardial damage (reviewed in 621 [137]). In chronic ischemia, autophagy was stimulated in porcine myo622 cardium concomitant with a decline in apoptosis [138]. Similarly, au623 tophagy was upregulated in heart-derived H9c2 myoblasts subjected 624 to mild ischemia, which was associated with increased ATP concentra625 tion, preservation of ΔΨm, and significant delay of cell death [139]. Mod626 erate and severe ischemia, on the other hand, resulted in apoptosis and 627 necrosis and absence of autophagy, suggesting that the degree of au628 tophagy induction depends on the severity of the insult [139]. In 629 cardiomyocyte-derived HL-1 cells, the autophagic flux was drastically 630 reduced in response to simulated I/R (sI/R) [140]. Pharmacological or

7. Conclusions and future perspectives

663

The optimal function and turnover of mitochondria appear to be critically involved in cardiac aging and CVD. The cellular processes of mitochondrial QC offer new targets for the development of therapeutics aimed at preventing and treating cardiac pathologies as well as attenuating the deterioration of heart function with aging [148]. The cardioprotective effects of H2S and the interaction of this gasotransmitter with autophagy highlight the intricate network of cellular stress response pathways, which can act both synergistically and independently from each other. Therefore, the development of cardioprotective strategies targeting autophagy and H2S signaling requires a thorough understanding of multiple factors, including the pathological context of insult, the extent of autophagy activation in a given pathophysiological setting, and the cytoprotective pathways elicited by H2S. This knowledge is necessary to exploit the cardioprotective properties of autophagy, while avoiding the detrimental effects of its defective or excessive activation.

664

Disclosures

680

None.

681

U

N

C

O

R

R

E

C

T

579 580

F

573

O

6.2. Role of autophagy in the pathogenesis of heart diseases

571 572

631 632

R O

577

569 570 Q9

genetic suppression of autophagy sensitized cardiac cells to apoptotic cell death after sI/R, while activation of autophagy using rapamycin or Beclin-1 overexpression was cardioprotective [140]. These findings are in keeping with the observation that rapamycin protects Langendorffperfused rat hearts against I/R injury [141]. Matsui et al. [142] described different roles for autophagy induction during ischemia and subsequent reperfusion in vivo. Mild ischemia induced autophagy through AMPK-dependent inhibition of mTOR, featuring an adaptive role for autophagy similar to that occurring under nutrient deprivation. During subsequent reperfusion, autophagy induction was mediated by Beclin-1 upregulation, independent of AMPK. Myocardial injury occurring during the reperfusion phase was attenuated in Beclin-1+/− mice, which therefore suggests a maladaptive role for autophagy during reperfusion [142]. Studies exploring a novel therapeutic approach for the management of myocardial I/R injury have provided additional information on the complex role of autophagy in this setting (reviewed in [143]). The administration of the gasotransmitter hydrogen sulfide (H2S) at the time of reperfusion limits infarct size and preserves left ventricular function in an in vivo mouse model of I/R [144]. Interestingly, in pigs, infarct size and left ventricular function were both improved when H2S infusion started before the ischemic insult and continued throughout the reperfusion phase [145]. H2S infusion reduced Beclin-1 expression, suggesting a suppressive effect of the gasotransmitter on autophagy [145]. The protective effect of H2S preconditioning during I/R injury has recently been demonstrated in hepatocytes both in vitro and in vivo, in which cytoprotection elicited by H2S was concomitant with inhibition of autophagy through AKT activation [146]. Taken together, H2S may exert its cytoprotective actions through a multitude of cellular pathways, including, but likely not limited to attenuation of mitochondrial respiration and mROS production, antioxidant actions and autophagy modulation [147].

P

576

resides in its ability to activate sirtuins [127], which have been associated with the lifespan extending effects of CR. Sirtuins stimulate autophagy through deacetylation of autophagic proteins such as Atg5 and Atg7 [128]. Resveratrol has shown to extend the lifespan of C. elegans [129], and to protect the rat myocardium from I/R injury in an autophagy-dependent manner [130]. A thorough characterization of the cardiac autophagic response to CR mimetics such as resveratrol may open a new venue for the development of interventions to attenuate the detrimental effects of aging on heart structure and function.

D

567 568

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

E

6

633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662

665 666 667 668 669 670 671 672 673 674 675 676 677 678 679

Acknowledgments

682

This work was partly supported by the Centro Study Achille e Linda Lorenzon (E.M.). We apologize to the authors whose excellent work could not be cited because of the vast literature on the subject and space limitations. We sincerely thank the anonymous reviewers for valuable input and criticism.

683

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

684 Q11 685 686 687

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

R

R

E

C

D

P

R O

O

F

[32] Fannin SW, Lesnefsky EJ, Slabe TJ, Hassan MO, Hoppel CL. Aging selectively decreases oxidative capacity in rat heart interfibrillar mitochondria. Arch Biochem Biophys 1999;372:399–407. [33] Suh JH, Heath SH, Hagen TM. Two subpopulations of mitochondria in the aging rat heart display heterogenous levels of oxidative stress. Free Radic Biol Med 2003;35:1064–72. [34] Judge S, Jang YM, Smith A, Hagen T, Leeuwenburgh C. Age-associated increases in oxidative stress and antioxidant enzyme activities in cardiac interfibrillar mitochondria: implications for the mitochondrial theory of aging. FASEB J 2005;19:419–21. [35] Hofer T, Servais S, Seo AY, Marzetti E, Hiona A, Upadhyay SJ, et al. Bioenergetics and permeability transition pore opening in heart subsarcolemmal and interfibrillar mitochondria: effects of aging and lifelong calorie restriction. Mech Ageing Dev 2009;130:297–307. [36] Palmer JW, Tandler B, Hoppel CL. Biochemical differences between subsarcolemmal and interfibrillar mitochondria from rat cardiac muscle: effects of procedural manipulations. Arch Biochem Biophys 1985;236:691–702. [37] Misra MK, Sarwat M, Bhakuni P, Tuteja R, Tuteja N. Oxidative stress and ischemic myocardial syndromes. Med Sci Monit 2009;15:RA209–19. [38] Penna C, Perrelli MG, Pagliaro P. Mitochondrial pathways, permeability transition pore, and redox signaling in cardioprotection: therapeutic implications. Antioxid Redox Signal 2013;18:556–99. [39] Becker LB. New concepts in reactive oxygen species and cardiovascular reperfusion physiology. Cardiovasc Res 2004;61:461–70. [40] Peng TI, Jou MJ. Oxidative stress caused by mitochondrial calcium overload. Ann N Y Acad Sci 2010;1201:183–8. [41] Venardos KM, Perkins A, Headrick J, Kaye DM. Myocardial ischemia–reperfusion injury, antioxidant enzyme systems, and selenium: a review. Curr Med Chem 2007;14:1539–49. [42] Di Lisa F, Bernardi P. Mitochondria and ischemia–reperfusion injury of the heart: fixing a hole. Cardiovasc Res 2006;70:191–9. [43] Tsutsui H, Kinugawa S, Matsushima S. Mitochondrial oxidative stress and dysfunction in myocardial remodelling. Cardiovasc Res 2009;81:449–56. [44] Ide T, Tsutsui H, Kinugawa S, Utsumi H, Kang D, Hattori N, et al. Mitochondrial electron transport complex I is a potential source of oxygen free radicals in the failing myocardium. Circ Res 1999;85:357–63. [45] Spinale FG, Coker ML, Thomas CV, Walker JD, Mukherjee R, Hebbar L. Timedependent changes in matrix metalloproteinase activity and expression during the progression of congestive heart failure: relation to ventricular and myocyte function. Circ Res 1998;82:482–95. [46] Ago T, Kuroda J, Pain J, Fu C, Li H, Sadoshima J. Upregulation of Nox4 by hypertrophic stimuli promotes apoptosis and mitochondrial dysfunction in cardiac myocytes. Circ Res 2010;106:1253–64. [47] Boudina S, Sena S, O'Neill BT, Tathireddy P, Young ME, Abel ED. Reduced mitochondrial oxidative capacity and increased mitochondrial uncoupling impair myocardial energetics in obesity. Circulation 2005;112:2686–95. [48] Shen X, Zheng S, Metreveli NS, Epstein PN. Protection of cardiac mitochondria by overexpression of MnSOD reduces diabetic cardiomyopathy. Diabetes 2006;55:798–805. [49] Boudina S, Abel ED. Diabetic cardiomyopathy revisited. Circulation 2007;115:3213–23. [50] Baker BM, Haynes CM. Mitochondrial protein quality control during biogenesis and aging. Trends Biochem Sci 2011;36:254–61. [51] Voos W. Chaperone-protease networks in mitochondrial protein homeostasis. Biochim Biophys Acta 1833;2013:388–99. [52] Soubannier V, McLelland GL, Zunino R, Braschi E, Rippstein P, Fon EA, et al. A vesicular transport pathway shuttles cargo from mitochondria to lysosomes. Curr Biol 2012;22:135–41. [53] Soubannier V, Rippstein P, Kaufman BA, Shoubridge EA, McBride HM. Reconstitution of mitochondria derived vesicle formation demonstrates selective enrichment of oxidized cargo. PLoS One 2012;7:e52830. [54] Lemasters JJ. Selective mitochondrial autophagy, or mitophagy, as a targeted defense against oxidative stress, mitochondrial dysfunction, and aging. Rejuvenation Res 2005;8:3–5. [55] Levine B, Klionsky DJ. Development by self-digestion: molecular mechanisms and biological functions of autophagy. Dev Cell 2004;6:463–77. [56] Klionsky DJ, Cuervo AM, Dunn Jr WA, Levine B, van der Klei I, Seglen PO. How shall I eat thee? Autophagy 2007;3:413–6. [57] Parzych KR, Klionsky DJ. An overview of autophagy: morphology, mechanism, and regulation. Antioxid Redox Signal 2013. [58] Kroemer G, Marino G, Levine B. Autophagy and the integrated stress response. Mol Cell 2010;40:280–93. [59] Chen Y, Azad MB, Gibson SB. Superoxide is the major reactive oxygen species regulating autophagy. Cell Death Differ 2009;16:1040–52. [60] Frank M, Duvezin-Caubet S, Koob S, Occhipinti A, Jagasia R, Petcherski A, et al. Mitophagy is triggered by mild oxidative stress in a mitochondrial fission dependent manner. Biochim Biophys Acta 1823;2012:2297–310. [61] Li L, Chen Y, Gibson SB. Starvation-induced autophagy is regulated by mitochondrial reactive oxygen species leading to AMPK activation. Cell Signal 2013;25:50–65. [62] Scherz-Shouval R, Elazar Z. Regulation of autophagy by ROS: physiology and pathology. Trends Biochem Sci 2011;36:30–8. [63] Wang Y, Nartiss Y, Steipe B, McQuibban GA, Kim PK. ROS-induced mitochondrial depolarization initiates PARK2/PARKIN-dependent mitochondrial degradation by autophagy. Autophagy 2012;8:1462–76. [64] Cooper CE, Patel RP, Brookes PS, Darley-Usmar VM. Nanotransducers in cellular redox signaling: modification of thiols by reactive oxygen and nitrogen species. Trends Biochem Sci 2002;27:489–92.

E

T

[1] Harman D. Aging: a theory based on free radical and radiation chemistry. J Gerontol 1956;11:298–300. [2] Miquel J, Economos AC, Fleming J, Johnson Jr JE. Mitochondrial role in cell aging. Exp Gerontol 1980;15:575–91. [3] Hekimi S, Lapointe J, Wen Y. Taking a “good” look at free radicals in the aging process. Trends Cell Biol 2011;21:569–76. [4] Vina J, Borras C, Abdelaziz KM, Garcia-Valles R, Gomez-Cabrera MC. The free radical theory of aging revisited: the cell signaling disruption theory of aging. Antioxid Redox Signal 2013;19:779–87. [5] Barja G. Updating the mitochondrial free radical theory of aging: an integrated view, key aspects, and confounding concepts. Antioxid Redox Signal 2013;19:1420–45. [6] Handy DE, Loscalzo J. Redox regulation of mitochondrial function. Antioxid Redox Signal 2012;16:1323–67. [7] Brand MD. The sites and topology of mitochondrial superoxide production. Exp Gerontol 2010;45:466–72. [8] Chance B, Hollunger G. The interaction of energy and electron transfer reactions in mitochondria. I. General properties and nature of the products of succinate-linked reduction of pyridine nucleotide. J Biol Chem 1961;236:1534–43. [9] Muller FL, Liu Y, Van Remmen H. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J Biol Chem 2004;279:49064–73. [10] Han D, Antunes F, Canali R, Rettori D, Cadenas E. Voltage-dependent anion channels control the release of the superoxide anion from mitochondria to cytosol. J Biol Chem 2003;278:5557–63. [11] Collins Y, Chouchani ET, James AM, Menger KE, Cocheme HM, Murphy MP. Mitochondrial redox signalling at a glance. J Cell Sci 2012;125:801–6. [12] Sena LA, Chandel NS. Physiological roles of mitochondrial reactive oxygen species. Mol Cell 2012;48:158–67. [13] Gauthier LD, Greenstein JL, Cortassa S, O'Rourke B, Winslow RL. A computational model of reactive oxygen species and redox balance in cardiac mitochondria. Biophys J 2013;105:1045–56. [14] Finkel T. From sulfenylation to sulfhydration: what a thiolate needs to tolerate. Sci Signal 2012;5:pe10. [15] Szczepanek K, Chen Q, Larner AC, Lesnefsky EJ. Cytoprotection by the modulation of mitochondrial electron transport chain: the emerging role of mitochondrial STAT3. Mitochondrion 2012;12:180–9. [16] Jacoby JJ, Kalinowski A, Liu MG, Zhang SS, Gao Q, Chai GX, et al. Cardiomyocyterestricted knockout of STAT3 results in higher sensitivity to inflammation, cardiac fibrosis, and heart failure with advanced age. Proc Natl Acad Sci U S A 2003;100:12929–34. [17] Hilfiker-Kleiner D, Hilfiker A, Fuchs M, Kaminski K, Schaefer A, Schieffer B, et al. Signal transducer and activator of transcription 3 is required for myocardial capillary growth, control of interstitial matrix deposition, and heart protection from ischemic injury. Circ Res 2004;95:187–95. [18] Kunisada K, Negoro S, Tone E, Funamoto M, Osugi T, Yamada S, et al. Signal transducer and activator of transcription 3 in the heart transduces not only a hypertrophic signal but a protective signal against doxorubicin-induced cardiomyopathy. Proc Natl Acad Sci U S A 2000;97:315–9. [19] Szczepanek K, Chen Q, Derecka M, Salloum FN, Zhang Q, Szelag M, et al. Mitochondrial-targeted signal transducer and activator of transcription 3 (STAT3) protects against ischemia-induced changes in the electron transport chain and the generation of reactive oxygen species. J Biol Chem 2011;286:29610–20. [20] Wojtovich AP, Nadtochiy SM, Brookes PS, Nehrke K. Ischemic preconditioning: the role of mitochondria and aging. Exp Gerontol 2012;47:1–7. [21] Dost T, Cohen MV, Downey JM. Redox signaling triggers protection during the reperfusion rather than the ischemic phase of preconditioning. Basic Res Cardiol 2008;103:378–84. [22] Bjelakovic G, Nikolova D, Gluud C. Antioxidant supplements to prevent mortality. JAMA 2013;310:1178–9. [23] Marzetti E, Wohlgemuth SE, Anton SD, Bernabei R, Carter CS, Leeuwenburgh C. Cellular mechanisms of cardioprotection by calorie restriction: state of the science and future perspectives. Clin Geriatr Med 2009;25:715–32. [24] Sachs HG, Colgan JA, Lazarus ML. Ultrastructure of the aging myocardium: a morphometric approach. Am J Anat 1977;150:63–71. [25] Liu VW, Zhang C, Nagley P. Mutations in mitochondrial DNA accumulate differentially in three different human tissues during ageing. Nucleic Acids Res 1998;26:1268–75. [26] Mohamed SA, Hanke T, Erasmi AW, Bechtel MJ, Scharfschwerdt M, Meissner C, et al. Mitochondrial DNA deletions and the aging heart. Exp Gerontol 2006;41:508–17. [27] Barja G, Herrero A. Oxidative damage to mitochondrial DNA is inversely related to maximum life span in the heart and brain of mammals. FASEB J 2000;14:312–8. [28] Trifunovic A, Wredenberg A, Falkenberg M, Spelbrink JN, Rovio AT, Bruder CE, et al. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature 2004;429:417–23. [29] Dai DF, Chen T, Wanagat J, Laflamme M, Marcinek DJ, Emond MJ, et al. Agedependent cardiomyopathy in mitochondrial mutator mice is attenuated by overexpression of catalase targeted to mitochondria. Aging Cell 2010;9:536–44. [30] Williams MD, Van Remmen H, Conrad CC, Huang TT, Epstein CJ, Richardson A. Increased oxidative damage is correlated to altered mitochondrial function in heterozygous manganese superoxide dismutase knockout mice. J Biol Chem 1998;273:28510–5. [31] Van Remmen H, Williams MD, Guo Z, Estlack L, Yang H, Carlson EJ, et al. Knockout mice heterozygous for Sod2 show alterations in cardiac mitochondrial function and apoptosis. Am J Physiol Heart Circ Physiol 2001;281:H1422–32.

N C O

689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772

References

U

688

7

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 Q12 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858

D

P

R O

O

F

[98] Kane LA, Youle RJ. Mitochondrial fission and fusion and their roles in the heart. J Mol Med (Berl) 2010;88:971–9. [99] Dorn II GW. Mitochondrial dynamics in heart disease. Biochim Biophys Acta 1833;2013:233–41. [100] Gomes LC, Scorrano L. Mitochondrial morphology in mitophagy and macroautophagy. Biochim Biophys Acta 1833;2013:205–12. [101] Twig G, Elorza A, Molina AJ, Mohamed H, Wikstrom JD, Walzer G, et al. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J 2008;27:433–46. [102] Gomes LC, Scorrano L. High levels of Fis1, a pro-fission mitochondrial protein, trigger autophagy. Biochim Biophys Acta 2008;1777:860–6. [103] Komatsu M, Waguri S, Ueno T, Iwata J, Murata S, Tanida I, et al. Impairment of starvation-induced and constitutive autophagy in Atg7-deficient mice. J Cell Biol 2005;169:425–34. [104] Mizushima N. Autophagy in protein and organelle turnover. Cold Spring Harb Symp Quant Biol 2011;76:397–402. [105] Shintani T, Klionsky DJ. Autophagy in health and disease: a double-edged sword. Science 2004;306:990–5. [106] Lavandero S, Troncoso R, Rothermel BA, Martinet W, Sadoshima J, Hill JA. Cardiovascular autophagy: concepts, controversies and perspectives. Autophagy 2013;9. [107] Yamaguchi O, Otsu K. Role of autophagy in aging. J Cardiovasc Pharmacol 2012;60:242–7. [108] Feldman ML, Navaratnam V. Ultrastructural changes in atrial myocardium of the ageing rat. J Anat 1981;133:7–17. [109] Brunk UT, Terman A. The mitochondrial-lysosomal axis theory of aging: accumulation of damaged mitochondria as a result of imperfect autophagocytosis. Eur J Biochem 2002;269:1996–2002. [110] Cuervo AM, Bergamini E, Brunk UT, Droge W, Ffrench M, Terman A. Autophagy and aging: the importance of maintaining “clean” cells. Autophagy 2005;1:131–40. [111] Cuervo AM, Dice JF. Age-related decline in chaperone-mediated autophagy. J Biol Chem 2000;275:31505–13. [112] Terman A, Brunk UT. Lipofuscin. Int J Biochem Cell Biol 2004;36:1400–4. [113] Taneike M, Yamaguchi O, Nakai A, Hikoso S, Takeda T, Mizote I, et al. Inhibition of autophagy in the heart induces age-related cardiomyopathy. Autophagy 2010;6:600–6. [114] Wohlgemuth SE, Julian D, Akin DE, Fried J, Toscano K, Leeuwenburgh C, et al. Autophagy in the heart and liver during normal aging and calorie restriction. Rejuvenation Res 2007;10:281–92. [115] Donati A, Cavallini G, Paradiso C, Vittorini S, Pollera M, Gori Z, et al. Age-related changes in the regulation of autophagic proteolysis in rat isolated hepatocytes. J Gerontol A Biol Sci Med Sci 2001;56:B288–93. [116] Keller JN, Dimayuga E, Chen Q, Thorpe J, Gee J, Ding Q. Autophagy, proteasomes, lipofuscin, and oxidative stress in the aging brain. Int J Biochem Cell Biol 2004;36:2376–91. [117] Rangaraju S, Hankins D, Madorsky I, Madorsky E, Lee WH, Carter CS, et al. Molecular architecture of myelinated peripheral nerves is supported by calorie restriction with aging. Aging Cell 2009;8:178–91. [118] Hua Y, Zhang Y, Ceylan-Isik AF, Wold LE, Nunn JM, Ren J. Chronic Akt activation accentuates aging-induced cardiac hypertrophy and myocardial contractile dysfunction: role of autophagy. Basic Res Cardiol 2011;106:1173–91. [119] Speakman JR, Mitchell SE. Caloric restriction. Mol Aspects Med 2011;32:159–221. [120] Mizushima N, Yamamoto A, Matsui M, Yoshimori T, Ohsumi Y. In vivo analysis of autophagy in response to nutrient starvation using transgenic mice expressing a fluorescent autophagosome marker. Mol Biol Cell 2004;15:1101–11. [121] Hansen M, Chandra A, Mitic LL, Onken B, Driscoll M, Kenyon C. A role for autophagy in the extension of lifespan by dietary restriction in C. elegans. PLoS Genet 2008;4: e24. [122] Rubinsztein DC, Marino G, Kroemer G. Autophagy and aging. Cell 2011;146:682–95. [123] Cruzen C, Colman RJ. Effects of caloric restriction on cardiovascular aging in nonhuman primates and humans. Clin Geriatr Med 2009;25:733–43. [124] Shinmura K, Tamaki K, Sano M, Murata M, Yamakawa H, Ishida H, et al. Impact of long-term caloric restriction on cardiac senescence: caloric restriction ameliorates cardiac diastolic dysfunction associated with aging. J Mol Cell Cardiol 2011;50:117–27. [125] Han X, Turdi S, Hu N, Guo R, Zhang Y, Ren J. Influence of long-term caloric restriction on myocardial and cardiomyocyte contractile function and autophagy in mice. J Nutr Biochem 2012;23:1592–9. [126] Ingram DK, Roth GS. Glycolytic inhibition as a strategy for developing calorie restriction mimetics. Exp Gerontol 2011;46:148–54. [127] Lam YY, Peterson CM, Ravussin E. Resveratrol vs. calorie restriction: data from rodents to humans. Exp Gerontol 2013. [128] Lee IH, Cao L, Mostoslavsky R, Lombard DB, Liu J, Bruns NE, et al. A role for the NADdependent deacetylase Sirt1 in the regulation of autophagy. Proc Natl Acad Sci U S A 2008;105:3374–9. [129] Morselli E, Maiuri MC, Markaki M, Megalou E, Pasparaki A, Palikaras K, et al. Caloric restriction and resveratrol promote longevity through the Sirtuin-1-dependent induction of autophagy. Cell Death Dis 2010;1:e10. [130] Lekli I, Ray D, Mukherjee S, Gurusamy N, Ahsan MK, Juhasz B, et al. Co-ordinated autophagy with resveratrol and gamma-tocotrienol confers synergetic cardioprotection. J Cell Mol Med 2010;14:2506–18. [131] Zhu H, Tannous P, Johnstone JL, Kong Y, Shelton JM, Richardson JA, et al. Cardiac autophagy is a maladaptive response to hemodynamic stress. J Clin Invest 2007;117:1782–93. [132] Cao DJ, Wang ZV, Battiprolu PK, Jiang N, Morales CR, Kong Y, et al. Histone deacetylase (HDAC) inhibitors attenuate cardiac hypertrophy by suppressing autophagy. Proc Natl Acad Sci U S A 2011;108:4123–8.

N

C

O

R

R

E

C

T

[65] Lee J, Giordano S, Zhang J. Autophagy, mitochondria and oxidative stress: cross-talk and redox signalling. Biochem J 2012;441:523–40. [66] Zhong Y, Wang QJ, Li X, Yan Y, Backer JM, Chait BT, et al. Distinct regulation of autophagic activity by Atg14L and Rubicon associated with Beclin 1-phosphatidylinositol-3kinase complex. Nat Cell Biol 2009;11:468–76. [67] Scherz-Shouval R, Shvets E, Fass E, Shorer H, Gil L, Elazar Z. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. EMBO J 2007;26:1749–60. [68] Dutta D, Xu J, Kim JS, Dunn Jr WA, Leeuwenburgh C. Upregulated autophagy protects cardiomyocytes from oxidative stress-induced toxicity. Autophagy 2013;9:328–44. [69] Rabkin SW, Klassen SS. Nitric oxide differentially regulates the gene expression of caspase genes but not some autophagic genes. Nitric Oxide 2007;16:339–47. [70] Yuan H, Perry CN, Huang C, Iwai-Kanai E, Carreira RS, Glembotski CC, et al. LPSinduced autophagy is mediated by oxidative signaling in cardiomyocytes and is associated with cytoprotection. Am J Physiol Heart Circ Physiol 2009;296:H470–9. [71] Ashrafi G, Schwarz TL. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ 2013;20:31–42. [72] Kim JS, Jin Y, Lemasters JJ. Reactive oxygen species, but not Ca2+ overloading, trigger pH- and mitochondrial permeability transition-dependent death of adult rat myocytes after ischemia–reperfusion. Am J Physiol Heart Circ Physiol 2006;290: H2024–34. [73] Rodriguez-Enriquez S, He L, Lemasters JJ. Role of mitochondrial permeability transition pores in mitochondrial autophagy. Int J Biochem Cell Biol 2004;36:2463–72. [74] Carreira RS, Lee Y, Ghochani M, Gustafsson AB, Gottlieb RA. Cyclophilin D is required for mitochondrial removal by autophagy in cardiac cells. Autophagy 2010;6:462–72. [75] Elmore SP, Qian T, Grissom SF, Lemasters JJ. The mitochondrial permeability transition initiates autophagy in rat hepatocytes. FASEB J 2001;15:2286–7. [76] Narendra D, Tanaka A, Suen DF, Youle RJ. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol 2008;183:795–803. [77] Kim JS, Nitta T, Mohuczy D, O'Malley KA, Moldawer LL, Dunn Jr WA, et al. Impaired autophagy: a mechanism of mitochondrial dysfunction in anoxic rat hepatocytes. Hepatology 2008;47:1725–36. [78] Springer W, Kahle PJ. Regulation of PINK1-Parkin-mediated mitophagy. Autophagy 2011;7:266–78. [79] Greene AW, Grenier K, Aguileta MA, Muise S, Farazifard R, Haque ME, et al. Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep 2012;13:378–85. [80] Narendra DP, Jin SM, Tanaka A, Suen DF, Gautier CA, Shen J, et al. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol 2010;8: e1000298. [81] Chen Y, Dorn II GW. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science 2013;340:471–5. [82] Kondapalli C, Kazlauskaite A, Zhang N, Woodroof HI, Campbell DG, Gourlay R, et al. PINK1 is activated by mitochondrial membrane potential depolarization and stimulates Parkin E3 ligase activity by phosphorylating Serine 65. Open Biol 2012;2:120080. [83] Geisler S, Holmstrom KM, Skujat D, Fiesel FC, Rothfuss OC, Kahle PJ, et al. PINK1/ Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat Cell Biol 2010;12:119–31. [84] Fimia GM, Stoykova A, Romagnoli A, Giunta L, Di Bartolomeo S, Nardacci R, et al. Ambra1 regulates autophagy and development of the nervous system. Nature 2007;447:1121–5. [85] Van Humbeeck C, Cornelissen T, Hofkens H, Mandemakers W, Gevaert K, De Strooper B, et al. Parkin interacts with Ambra1 to induce mitophagy. J Neurosci 2011;31:10249–61. [86] Ding WX, Ni HM, Li M, Liao Y, Chen X, Stolz DB, et al. Nix is critical to two distinct phases of mitophagy, reactive oxygen species-mediated autophagy induction and Parkin-ubiquitin-p62-mediated mitochondrial priming. J Biol Chem 2010;285:27879–90. [87] Sandoval H, Thiagarajan P, Dasgupta SK, Schumacher A, Prchal JT, Chen M, et al. Essential role for Nix in autophagic maturation of erythroid cells. Nature 2008;454:232–5. [88] Novak I, Kirkin V, McEwan DG, Zhang J, Wild P, Rozenknop A, et al. Nix is a selective autophagy receptor for mitochondrial clearance. EMBO Rep 2010;11:45–51. [89] Gustafsson AB. Bnip3 as a dual regulator of mitochondrial turnover and cell death in the myocardium. Pediatr Cardiol 2011;32:267–74. [90] Kubli DA, Quinsay MN, Huang C, Lee Y, Gustafsson AB. Bnip3 functions as a mitochondrial sensor of oxidative stress during myocardial ischemia and reperfusion. Am J Physiol Heart Circ Physiol 2008;295:H2025–31. [91] Hamacher-Brady A, Brady NR, Logue SE, Sayen MR, Jinno M, Kirshenbaum LA, et al. Response to myocardial ischemia/reperfusion injury involves Bnip3 and autophagy. Cell Death Differ 2007;14:146–57. [92] Lee Y, Lee HY, Hanna RA, Gustafsson AB. Mitochondrial autophagy by Bnip3 involves Drp1-mediated mitochondrial fission and recruitment of Parkin in cardiac myocytes. Am J Physiol Heart Circ Physiol 2011;301:H1924–31. [93] Quinsay MN, Thomas RL, Lee Y, Gustafsson AB. Bnip3-mediated mitochondrial autophagy is independent of the mitochondrial permeability transition pore. Autophagy 2010;6:855–62. [94] Boldogh IR, Pon LA. Mitochondria on the move. Trends Cell Biol 2007;17:502–10. [95] Ong SB, Hall AR, Hausenloy DJ. Mitochondrial dynamics in cardiovascular health and disease. Antioxid Redox Signal 2013;19:400–14. [96] Scorrano L. Keeping mitochondria in shape: a matter of life and death. Eur J Clin Invest 2013;43:886–93. [97] Menzies RA, Gold PH. The turnover of mitochondria in a variety of tissues of young adult and aged rats. J Biol Chem 1971;246:2425–9.

U

859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

E

8

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 Q13 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030

S.E. Wohlgemuth et al. / Journal of Molecular and Cellular Cardiology xxx (2014) xxx–xxx

[141] Khan S, Salloum F, Das A, Xi L, Vetrovec GW, Kukreja RC. Rapamycin confers preconditioning-like protection against ischemia–reperfusion injury in isolated mouse heart and cardiomyocytes. J Mol Cell Cardiol 2006;41:256–64. [142] Matsui Y, Takagi H, Qu X, Abdellatif M, Sakoda H, Asano T, et al. Distinct roles of autophagy in the heart during ischemia and reperfusion: roles of AMP-activated protein kinase and Beclin 1 in mediating autophagy. Circ Res 2007;100:914–22. [143] King AL, Lefer DJ. Cytoprotective actions of hydrogen sulfide in ischaemia–reperfusion injury. Exp Physiol 2011;96:840–6. [144] Elrod JW, Calvert JW, Morrison J, Doeller JE, Kraus DW, Tao L, et al. Hydrogen sulfide attenuates myocardial ischemia–reperfusion injury by preservation of mitochondrial function. Proc Natl Acad Sci U S A 2007;104:15560–5. [145] Osipov RM, Robich MP, Feng J, Liu Y, Clements RT, Glazer HP, et al. Effect of hydrogen sulfide in a porcine model of myocardial ischemia–reperfusion: comparison of different administration regimens and characterization of the cellular mechanisms of protection. J Cardiovasc Pharmacol 2009;54:287–97. [146] Wang D, Ma Y, Li Z, Kang K, Sun X, Pan S, et al. The role of AKT1 and autophagy in the protective effect of hydrogen sulphide against hepatic ischemia/reperfusion injury in mice. Autophagy 2012;8:954–62. [147] Calvert JW, Coetzee WA, Lefer DJ. Novel insights into hydrogen sulfide-mediated cytoprotection. Antioxid Redox Signal 2010;12:1203–17. [148] Sciarretta S, Zhai P, Volpe M, Sadoshima J. Pharmacological modulation of autophagy during cardiac stress. J Cardiovasc Pharmacol 2012;60:235–41.

F

[133] Oyabu J, Yamaguchi O, Hikoso S, Takeda T, Oka T, Murakawa T, et al. Autophagymediated degradation is necessary for regression of cardiac hypertrophy during ventricular unloading. Biochem Biophys Res Commun 2013;441:787–92. [134] Nakai A, Yamaguchi O, Takeda T, Higuchi Y, Hikoso S, Taniike M, et al. The role of autophagy in cardiomyocytes in the basal state and in response to hemodynamic stress. Nat Med 2007;13:619–24. [135] Rothermel BA, Hill JA. Autophagy in load-induced heart disease. Circ Res 2008;103:1363–9. [136] Przyklenk K, Dong Y, Undyala VV, Whittaker P. Autophagy as a therapeutic target for ischaemia/reperfusion injury? Concepts, controversies, and challenges. Cardiovasc Res 2012;94:197–205. [137] Gustafsson AB, Gottlieb RA. Recycle or die: the role of autophagy in cardioprotection. J Mol Cell Cardiol 2008;44:654–61. [138] Yan L, Vatner DE, Kim SJ, Ge H, Masurekar M, Massover WH, et al. Autophagy in chronically ischemic myocardium. Proc Natl Acad Sci U S A 2005;102:13807–12. [139] Loos B, Genade S, Ellis B, Lochner A, Engelbrecht AM. At the core of survival: autophagy delays the onset of both apoptotic and necrotic cell death in a model of ischemic cell injury. Exp Cell Res 2011;317:1437–53. [140] Hamacher-Brady A, Brady NR, Gottlieb RA. Enhancing macroautophagy protects against ischemia/reperfusion injury in cardiac myocytes. J Biol Chem 2006;281:29776–87.

O

1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052

9

U

N C O

R

R

E

C

T

E

D

P

R O

1075

Please cite this article as: Wohlgemuth SE, et al, The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology, J Mol Cell Cardiol (2014), http://dx.doi.org/10.1016/j.yjmcc.2014.03.007

1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074

The interplay between autophagy and mitochondrial dysfunction in oxidative stress-induced cardiac aging and pathology.

Aging is accompanied by a progressive increase in the incidence and prevalence of cardiovascular disease (CVD). Prolonged exposure to cardiovascular r...
417KB Sizes 0 Downloads 4 Views