YJMCC-08071; No. of pages: 7; 4C: Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

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

Review article

2Q1

Mitochondrial dynamics: Orchestrating the journey to advanced age

3Q2

Agnieszka K. Biala, Rimpy Dhingra, Lorrie A. Kirshenbaum ⁎

4Q3 5 6

The Institute of Cardiovascular Sciences, St. Boniface Hospital Research Centre, Winnipeg, Manitoba R2H 2A6, Canada Department of Physiology, Faculty of Health Sciences, College of Medicine, University of Manitoba, Winnipeg, Manitoba R2H 2A6, Canada Department of Pharmacology& Therapeutics, Faculty of Health Sciences, College of Medicine, University of Manitoba, Winnipeg, Manitoba R2H 2A6, Canada

7

a r t i c l e

8 9 10 11 12

Article history: Received 19 January 2015 Received in revised form 30 March 2015 Accepted 19 April 2015 Available online xxxx

13 14 15 16 17 18

Keywords: ROS Oxidative stress Aging Mitochondrial dynamics Cell death

O

R O

i n f o

a b s t r a c t

E

D

P

Aging is a degenerative process that unfortunately is an inevitable part of life and risk factor for cardiovascular disease including heart failure. Among the several theories purported to explain the effects of age on cardiac dysfunction, the mitochondrion has emerged a central regulator of this process. Hence, it is not surprising that abnormalities in mitochondrial quality control including biogenesis and turnover have such detrimental effects on cardiac function. In fact mitochondria serve as a conduit for biological signals for apoptosis, necrosis and autophagy respectively. The removal of damaged mitochondria by autophagy/mitophagy is essential for mitochondrial quality control and cardiac homeostasis. Defects in mitochondrial dynamism fission/fusion events have been linked to cardiac senescence and heart failure. In this review we discuss the impact of aging on mitochondrial dynamics and senescence on cardiovascular health. This article is part of a Special Issue entitled: CV Aging. © 2015 Published by Elsevier Ltd.

Contents . . . . . . . . . .

. . . . . . . . . .

R

N C O

45

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

E

. . . . . . . . . .

R

1. Introduction . . . . . . . . . . . . . . . . 2. Theories of aging and the mitochondrion . . . 3. Mitochondrial damage and aging . . . . . . 4. Molecular regulation of the aging heart . . . . 5. Aging and caloric restriction . . . . . . . . . 6. Mitochondrial dynamics in the senescent heart 7. Mitophagy in the senescent heart . . . . . . Conflict of interest . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . .

C

31

35 36 37 38 39 40 41 42 43 44

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

1. Introduction

47 48

The search for the “Fountain of Youth” dates as far back as the 5th century BC in which the Greek historian Herodotus describes waters with magical rejuvenating powers that would reverse the effects of time on age. Indeed, throughout historical record there are countless anecdotes of expeditions commissioned by kings and aristocrats alike seeking the ambrosia that would promise youth and immortality. The concept of reversing or even halting the effects of age on the body and

51 52 53

U

46

49 50

19 20 21 22 23 24 25 26 27 Q4 28

T

32 30 29 34 33

F

1

⁎ Corresponding author at: Institute of Cardiovascular Sciences, St. Boniface General Hospital Research Centre Rm. 3016, 351 TachéAvenue, Winnipeg, Manitoba, Canada, R2H 2A6. Tel.: +1 204 235 3661; fax: +1 204 233 6723. E-mail address: [email protected] (L.A. Kirshenbaum).

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

0 0 0 0 0 0 0 0 0 0

restoring one's youth has withstood the tests of time to even modern day society. In fact one can easily find advertisements and infomercials for a plethora of “health products” that claim to magically reverse the effects of time and age on the body. Hence, our modern day obsession with age has become a multi-billion dollar industry world-wide driven by the desire to be “forever young”. Aging leads to a progressive decline in the normal physiological function of the organism. These age-related defects include DNA strand breaks, defective cell proliferation and repair, mitochondrial dysfunction, accumulation of oxidized proteins, and lipids, organ failure that ultimately lead to senescence and death. In fact, the inability of the body to remove cells that have become genetically unstable or irreversibly injured by apoptosis or necrosis respectively is considered a central underlying mechanism for the increased prevalence of human cancers in the aging population. This perhaps is best exemplified by the

http://dx.doi.org/10.1016/j.yjmcc.2015.04.015 0022-2828/© 2015 Published by Elsevier Ltd.

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

54 55 56 57 58 59 60 61 62 63 64 65 66 67

93 94 95 96 97 98

C

91 92

E

89 90

R

87 88

R

85 86

O

83 84

C

81 82

N

79 80

U

77 78

F

Though several theories of cellular aging have been suggested, oxidative stress from increased production of reactive oxygen species

75 76

O

102 103

74

R O

2. Theories of aging and the mitochondrion

72 73

P

101

70 71

(ROS) has been identified as a major contributing factor [5]. Adaptive mechanisms to overcome the potentially harmful effects of molecular oxygen occurred over millions of years ago of the evolutionary pressure and the appearance of the first land creatures that utilized atmospheric oxygen for aerobic respiration. Indeed, the evolution of multi-cellular organisms that relied on mitochondrial oxidative processes and respiration for ATP synthesis also required adaptive mechanisms to detoxify the lethal effects of mono- and divalent oxygen species [6–8]. Early studies in aerobic metabolism in eukaryotic cells revealed that molecular divalent oxygen served as a terminal electron acceptor in the mitochondrial electron transport chain during aerobic respiration reviewed in [6,9]. Here the sequential tetravalent reduction of molecular oxygen via the mitochondrial electron transport chain complexes (I–IV) to yield water and carbon dioxide is a critical mechanism for efficient ATP synthesis and prevention of reactive oxygen species (ROS). Univalent or monovalent reduction of molecular oxygen through electron leakage within the electron transport chain or other cellular sources leads to the generation of highly reactive ROS intermediates that include the superoxide anion, hydrogen peroxide, hydroxyl radical, peroxy radicals and other minor species [6, 10]. In addition to the mitochondrial electron transport chain, other cellular sources of ROS include membrane bound NADPH oxidases [11], H2O2-generating monoamine oxidases [12] and p66shc, and ROSinduced/ROS-released [13]. If left unchecked ROS intermediates can induce cellular injury including DNA strand breaks, oxidation of lipids, proteins and carbohydrates resulting in organelle dysfunction. Increased ROS production has been shown to disrupt normal cardiac function and contractile abnormalities resulting in heart failure reviewed in [14–16]. Both enzymatic (superoxide dismutase, catalase and glutathione peroxidase) and non-enzymatic antioxidants including water (vitamin C) and lipid soluble antioxidants (vitamins A, E) neutralize ROS and protect against oxidative stress injury [16,17]. Hence, an imbalance between excess ROS production and anti-oxidant capacity would trigger deleterious cellular defects resulting in cell death and is the fundamental tenant of the ROS theory of aging [18]. For example, increased ROS production has been reported in the aged myocardium and is believed to underlie in part the cellular defects associated with heart failure [19–21]. The mitochondrion has emerged as a convergence point for not only regulating cellular respiration but also signaling pathways linked to cell

T

99 100

fact that defects in the mechanisms that govern autophagy, which is predominantly required for catabolic recycling of damaged organelles and prevention of proteo- and lipo-toxic stress, are considered a major driving force for cellular senescence with aging. In the context of the heart, cardiac senescence is a major clinical problem since age-associated defects in contractility, calcium handling, cell metabolism and mitochondrial function have been identified as a primary underlying defects associated with senescence and heart failure reviewed in [1]. In this regard, heart failure represents a major financial and socioeconomic burden to the health care system, since individuals diagnosed with heart failure require costly long-term care. In fact, the co-morbid effects of aging and frailty coupled with cardiovascular disease are expected to escalate over the next several years in the baby boomer population. Next to cancer, age-associated complications with cardiovascular disease account for more than half the reported deaths in North America. For example, cardiac senescence resulting from the accumulation of oxidized proteins, lipids and increased ROS production from dysfunctional organelles such as mitochondria contribute to vascular remodeling, and impaired contractility (Fig. 1). Hence, a better understanding of the underlying molecular mechanisms that contribute to cardiac senescence and heart failure would be of tremendous scientific and clinical importance [2]. As a physiological/pathological process aging induces distinct biochemical changes to proteins and lipids that alter the structural integrity and function of the vasculature and myocardium. In this regard, recent evidence suggests that cellular senescence is a contributing factor to atherosclerosis. Indeed, endothelial dysfunction activates pro-inflammatory cytokines resulting in inflammation, pro-coagulatory responses and thrombosis. Endothelium-dependent vasodilation is impaired during senescence largely due to increased oxidative injury and decreased nitric oxide production. Macroscopically, arteries lose their normal elastic properties and become calcified and rigid which contributes to vascular remodeling and dysfunction [3,4].

D

68 69

A.K. Biala et al. / Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

E

2

Fig. 1. Signaling pathways involved in cardiac senescence. Figure depicts cellular senescence signaling pathways in the heart. Increased age predisposes to development of senescencerelated events resulting in endothelial cell dysfunction, fibrosis, hypertrophy, atherosclerosis, contractile dysfunction, myocardial infarction (MI) and heart failure. The hallmarks of mitochondrial senescence included mtDNA mutations, respiratory chain defects, ROS production, loss of ΔΨm, Ca2+ overload, nuclear DNA damage and telomere attrition. Perturbations in mitochondrial dynamism (fission/fusion) impair autophagy/mitophagy resulting in cellular senescence. Nicotinamide adenine dinucleotide phosphate (NADPH); p66, mono amino oxidase (MAO); reactive oxygen species (ROS); Mitofusin 1&2 (Mfn1&2), optic atrophy type 1 (Opa1); Dynamin-related Protein1 (Drp1); Mitochondrial fission 1 (Fis1), Bcl2 19 kDa interacting protein3 (Bnip3)/Bnip3-like protein (Nix/Bnip3FL); light chain protein3 (LC3).

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141

A.K. Biala et al. / Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

163 164 165 166 167 168 169 170 171 172

177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205

C

161 162

E

159 160

R

4. Molecular regulation of the aging heart

218

Shortly, after birth cardiac myocytes exit the cell cycle and lose their ability to actively replicate. This inherent property of the post-mitotic heart has been suggested to be a key underlying factor of cellular senescence and heart failure, since damaged cardiac myocytes are not replaced by new ones. The underlying mechanism for cellular senescence in the adult heart has been linked to developmental down-regulation of key cell cycle proteins as well as to telomere attrition. Telomeres, are located at the end of chromosomal DNA, and protect DNA during replication. However, since telomere length shortens with each round of DNA replication, telomeres and telomerase activity can be used as an index of cellular senescence [43–45]. In this regard, telomerase deficient mice (Terc −/−) were reported to have shorter telomere lengths than wild type mice that were accompanied with increased senescence. In addition, elevated levels of cyclin-dependent kinase inhibitor p21CIP1WAF1 were observed in the aortae of Terc −/− mice with DNA aging marker—γH2AX—in endothelial cells, as well as cell cycle inhibitors p16INKa and p19ARF [46–50]. High levels of cell cycle inhibitors p16INK4a/Rb and p14ARF/p53 have been reported in the aged myocardium [51] as well as senescent endothelial and vascular smooth muscle cells [52–54].

219 220

5. Aging and caloric restriction

239

As stated above the mitochondrion is a major source of ROS production during cellular metabolism and has been intimately linked to senescence. Pioneering studies in worms, flies and rodents conclusively revealed that life span of the model organism was dramatically increased by simply restricting the caloric intake [55]. This was attributed to a reduction in mitochondrial ROS production and presumably oxidative injury [56]. Collectively, these studies revealed for the first time an unprecedented relationship between caloric restriction, and longevity [57]. Several lines of investigation have linked the Sirtuins (Sirt 1–7) as molecular caloric sensors [58]; [59–61]. Sirt-1 is the mammalian homolog to the yeast Sir-2 longevity factor which was demonstrated to be critical for regulating life span of yeast. Sirt-1 is an NAD+ regulated histone deacetylase activated in cells during caloric restriction or nutrient stress [58,62,63]. The discovery of Sirt-1 and its activation by caloric restriction further links cellular metabolism to vital cellular processes involved in aging including apoptosis and autophagy [58]. Several proteins known to regulate metabolism and proliferation are intimately coupled to Sirt-1 highlighting the inner connectivity of these pathways. These include the tumor suppressor p53, Forkhead Box Protein O1/3 (FOXO1/3), mitochondrial biogenesis regulator, peroxisome proliferator-activated receptor gamma coactivator 1 α (PGC-1 α), nuclear factor κB (NF-κB) and mechanistic target of rapamycin (mTOR) as reviewed in [64]. By targeting these different signaling pathways, Sirt1 suppresses aging by impinging simultaneously on multiple cellular pathways crucial for cell longevity such as DNA repair, mitochondrial biogenesis, cell metabolism, and cell death thereby delaying aging and prolonging lifespan [62,65,66]. Pharmacological approaches to activate

240

F

157 158

R

155 156

N C O

153 154

U

151 152

O

In addition to its many complex biochemical functions, the mitochondrion is uniquely distinguished from other cellular organelles because it contains its own mitochondrial DNA (mtDNA). The mtDNA is circular and located within the mitochondrial matrix. It is continually turned over through replicative and degradative processes. Notably, certain proteins of the respiratory electron transport chain are encoded by mtDNA, whereas other mitochondrial including cytochrome c and other heme constituent proteins are encoded by nuclear DNA. Disorders associated with mtDNA mutations generally fall into two classes: those linked to specific, maternally-inherited mtDNA mutations and those linked to mutations in nuclear-encoded genes that maintain normal fidelity and stability of mtDNA [33]. The incidence and frequency of mtDNA mutations increase exponentially with age and contribute to cellular senescence [34,35]. The mitochondrial theory of aging was substantiated in studies in mice expressing error-prone mitochondrial DNA polymerase γ (PolgD257A) defective for proof-reading activity. The increased mutation frequency from loss of function of DNA proof-reading activity resulted in the expression of defective respiratory chain proteins and premature aging [33,36,37]. Moreover, these mitochondrial defects resulted in pre-mature aging, dilated cardiac hypertrophy and mice dying of congestive heart failure by six months of age [38]. Through the process of mitochondrial fusion, the mtDNA of one mitochondrion can be transmitted or shared with other mitochondria within the cell. Therefore, the fusion of healthy mitochondria with mitochondria harboring mutations within their mtDNA during replication could be detrimental and profoundly influence the quality of mitochondrial pool within the cell [39–41]. Tam et al. [42] used a mathematical in silico model to predict the frequency of mtDNA mutations and accumulation over time. In this study, the authors demonstrated that lowered rates of mitochondrial fission and fusion events with age resulted in a

149 150

R O

176

148

206 207

P

3. Mitochondrial damage and aging

146 147

predicted increase in the disproportionate distribution of mutant mtDNA to mitochondria. Presumably based on this model, the higher prevalence of mtDNA mutations with age would result in impaired mitochondrial respiration and organ failure. Based on this study, a critical role for age dependent regulation of mitochondrial fission–fusion events and mitochondrial turnover rates via mitophagy as discussed above is important for averting the accumulation of mitochondria with damaged mtDNA and contamination of the mitochondrial pool. Collectively, these findings highlight the importance of mitochondrial dynamism for ensuring normal mitochondrial homogeneity with increased age; this will be discussed in more detail below.

D

175

144 145

T

173 174

death by apoptosis, necrosis and autophagy during aging [22]. The link between mitochondrial perturbations and aging has been previously associated with activation of the intrinsic mitochondrial death pathway and increased apoptotic cell death [23]. In fact, mitochondrial perturbations that have been suggested to underlie cellular aging have also been intimately linked to autophagy/mitophagy. In the most classical sense autophagy is defined as a catabolic process by which macromolecular structures, including proteins, and organelles such as mitochondria are re-cycled by an elaborate autophagosome–lysosomal system into their native biochemical constituents, carbohydrates, lipids and proteins for generating ATP during caloric restriction or nutrient stress. In this context several lines of investigation have suggested that the removal of damaged mitochondria by mitophagy is critical for maintaining mitochondrial quality control and cellular homeostasis for two important and salient reasons [24]. Selective removal of dysfunctional mitochondria by mitophagy during caloric restriction would presumably confer a survival advantage by eliminating damaged mitochondria that would otherwise trigger mitochondrial-driven cell death. Moreover, mitophagy of damaged mitochondria would ensure that only healthy mitochondria are maintained within the cell for mitochondrial biogenesis [24–28]. It must be stated however that beyond a certain threshold excessive mitophagy/autophagy is maladaptive and promotes cell death [29,30]. Previous work by our laboratory demonstrated, that the inducible protein Bnip3 triggered maladaptive autophagy in cardiac myocytes and necrotic cell death following p53 activation [30]. In the same study we demonstrated that interventions that suppressed autophagy, were equivalently sufficient for suppressing cell death—indicating that at least in this context excess autophagy induced by Bnip3 was maladaptive and promoted death [30]. Hence, a growing body of experimental evidence suggests that age associated apoptosis and autophagy may not be mutually exclusive or independently regulated events but commonly linked through several Bcl-2 family proteins including Beclin-1, Bnip3, Bim, Bnip3L, Bax, Bak and others reviewed in [31,32].

E

142 143

3

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

208 209 210 211 212 213 214 215 216 217

221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238

241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267

293 294 295 296 297 298 299 300 301 302 303 304 305 306

311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331

C

291 292

E

289 290

R

287 288

R

285 286

O

283 284

C

281 282

N

279 280

U

277 278

F

There is a growing body of evidence that suggest that mitochondria are highly dynamic organelles constantly remodeling through fission and fusion to produce fragmented or highly organized mitochondrial networks. Mitochondrial fission and fusion are highly regulated events and critical for the normal growth, development and functioning of the cell reviewed in [78–80]. Recently, defects in mitochondrial dynamism (fission/fusion) have been linked to a variety of human pathologies and age-associated diseases [81,82]. Given the importance of the mitochondrion in the regulation of cellular metabolism it is not surprising that genetic abnormalities in mitochondria inherited or induced have such a dramatic and profound impact on human disease, especially in organs such as the heart which rely heavily on mitochondria for normal physiological function. The molecular signaling pathways by which mitochondria direct cell growth, autophagy and cell death remain to be elucidated. Mitochondrial dynamics is regulated by a number of cellular proteins related to the dynamin motor GTPases that regulate fission and fusion of mitochondrial outer and inner membranes. These include key fission proteins GTPase Drp1 (Dynamin related protein-1) [83] and Fis1 [84, 85] and mitochondrial fusion proteins Mitofusins 1&2 (Mfn1/2) [86,87] and Optic Atrophy 1 (Opa1) [88]. The physiological importance of mitochondrial fission and fusion process in growth, development and aging is supported by the reports that mice germ-line deleted for

275 276

O

309 310

274

R O

6. Mitochondrial dynamics in the senescent heart

272 273

P

308

270 271

either of the mitochondrial fission or fusion proteins (Drp1, Mfn2 or Opa-1) were found to be embryonic lethal [86,89]. Mice with disrupted mitochondrial fusion ability developed several mutations in mtDNA, and displayed impaired respiratory function [87]. However, mice heterozygous for the inner mitochondrial membrane fusion protein Opa-1 were viable but developed age dependent retinal and nerve degeneration [90]. Recently, an essential and critical role for mitochondrial fusion has been suggested for normal function of the heart. In this report, knockdown of OPA-1 and Mitochondrial Assembly Regulatory Factor (MARF) the functional homologue of human Mitofusins) in the heart tubes of Drosophila melanogaster increased heterogeneity in mitochondria, displaying heart tube dilation and contractile impairment [91]. In another interesting report, lipid accumulation during aging was linked to decreased expression of Mfn2 and impaired autophagy. In this study, which was conducted in both young and old mice, Zhao et al. demonstrated an elevated levels of triglycerides in the liver and heart muscles of old mice, associated with a reduction in PGC-1α, mitochondrial fusion protein, Mfn2 and declined mitochondrial function [92]. Although mitochondrial fusion is essential for normal cardiac myocyte homeostasis, in cells defective for mitophagy, mitochondrial fusion can reportedly trigger dilated cardiomyopathy from the promiscuous fusion of senescent damaged mitochondria with normal healthy ones thereby contaminating the cellular pool of mitochondria [93]. Together these studies highlight the importance of mitochondrial fission and fusion for normal development and aging. This is best illustrated by a recent report by Park et al. [94] which showed the importance of mitochondrial fission in preventing onset of senescence. Here, MARCH5, a mitochondrial-associated ubiquitin ligase was shown to regulate mitochondrial fission and fusion, MARCH5 depletion in HeLa cells resulted in highly interconnected elongated mitochondrial network, increased expression of Mfn1, coupled with increased senescence associated marker β-galactosidase [94]. Interestingly, defects in mitochondrial network and senescence in MARCH5 depleted cells was reversed by either inactivating Mfn1 or inducing fission by Drp1 overexpression, suggesting the importance of mitochondrial fission events in preventing or delaying senescence [94]. It remains to be seen whether these events are equivalently operational in the vasculature or heart muscle itself with age. Another study in young and old mice, demonstrated increased mitochondrial fusion in aged muscle concordant with increased Mfn1 and Mfn2. Interestingly, Opa-1 levels which have been implicated in mitochondrial inner membrane fusion, were down-regulated in aged muscle. The increased mitochondrial fusion in aged muscle was accompanied by lower fission rates related to a decline in Fis-1 since Drp-1 levels remained unchanged [95]. Aging related decline in cardiovascular performance is also associated with thickening and increased stiffness of arteries and impaired endothelial cell function. Endothelium produces nitric oxide and other vasoactive materials that are important for maintaining cardiovascular health. Recently endothelium dysfunction during ischemia–reperfusion (I–R) has been linked to excessive mitochondrial fission from increased oxidative stress injury [96]. In this study, it was demonstrated that endothelial cells subjected to I–R underwent changes in mitochondrial morphology and exhibited phenotypic changes including mitochondrial fission that were partially suppressed by either anti-oxidants and inhibition of NO synthase or DRP1, suggesting a tentative link between endothelial cell dysfunction and mitochondrial fragmentation induced by oxidative stress [96]. There is a growing body of literature that supports the notion that active mitochondrial dynamics is critical for delaying senescence, through either mitochondrial repair process or mitophagy, however, an alternative view by Figge et al. proposed reduced mitochondrial dynamics and fission–fusion cycles would instead delay the detrimental effects of aging and senescence [97]. Although, this study was based on an in silico modeling of mitochondrial turnover and which will require further in vivo testing for validation.

T

307

Sirt1 with resveratrol or SRT1720 were shown to protect endothelial cells from senescence through inhibition of p53–p21 pathway [67]. SRT1720 was also reported to normalize NF-κB and TNF-α levels in old mice, ameliorating endothelial dysfunction, oxidative stress injury and inflammation [68]. The metabolic link between Sirt-1 and senescence is profound. Senescent cells are typified by reduced metabolic activity and increased AMP (or ADP) to ATP ratios. This activates AMP-activated protein kinase (AMPK), a central regulator of metabolic signaling which induces growth arrest and autophagy by inhibiting mechanistic target of rapamycin (mTOR) pathway [69,70]. A recent report by the Finkel laboratory demonstrated that attenuating mTOR activity in mice expressing hypomorphic mTOR(Δ/Δ) alleles that express only 25% of normal mTORC1 and mTORC2 activity increased lifespan and decelerated aging in mice—supporting a link between metabolism, aging and life-span [71]. Moreover, a recent study by our laboratory demonstrated that mTOR inhibition with Rapamycin or hypoxia activated the Bcl-2 protein Bnip3 triggering autophagy in cardiac myocytes. Our studies demonstrated a direct functional link between nutrient sensing properties of mTOR and autophagy induced by Bnip3 [72]. Concordantly, Sirt1-mediated deacetylation of Foxo3 was shown to promote survival in a kidney model of caloric restriction, further highlighting the link between cellular senescence and metabolic signaling. Moreover, deacetylation of p53 by Sirt1 blunted p21 activation and cellular aging [73,74]. Taken together these findings collectively reveal the highly interconnected relationship between cell metabolism, growth and senescence [74]. Interestingly in contrast to Sirt-1 which is predominantly localized to the nucleus [75], Sirt-3 is reportedly localized to mitochondria and was recently demonstrated to deacetylate cyclophilin D (CypD), putative to be the regulator of the mitochondrial permeability transition pore (mPTP) [76]. In the study, deacetylation of CypD suppressed the age-dependent increase in mitochondrial swelling and cardiac dysfunction [76]. Moreover, Sirt3-deficiency was linked with increased vulnerability to ischemia–reperfusion injury, suggesting that the activation of Sirt3 may play an important role in preserving mitochondrial function through its effects on CypD during cellular stress [77]. However, it remains to be tested whether CypD inhibition by Sirt3 will protect mitochondria from mPTP in the aged myocardium, given that the study by Basso et al. demonstrated that mPTP can also occur in the absence of CypD in the liver in vitro and in vivo [2].

D

268 269

A.K. Biala et al. / Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

E

4

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

332 333 334 335 336 337 338 339 340 341 342 Q5 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397

A.K. Biala et al. / Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462

470

465 466 467 468 469

471

None.

F

Acknowledgment

472

This work was supported by grants to LAK from the Canadian Insti- 473 tutes of Health Research. AKB holds a post-doctoral fellowship from 474 Q6 the Manitoba Health Research Council and IMPACT-CIHR Fellowship. 475

O

414 415

Conflict of interest

References

R O

412 413

C

410 411

E

408 409

R

406 407

R

404 405

N C O

403

U

401 402

[1] Kovacic JC, Moreno P, Hachinski V, Nabel EG, Fuster V. Cellular senescence, vascular disease, and aging: part 1 of a 2-part review. Circulation 2011;123:1650–60. [2] Basso E, Fante L, Fowlkes J, Petronilli V, Forte MA, Bernardi P. Properties of the permeability transition pore in mitochondria devoid of Cyclophilin D. J Biol Chem 2005;280:18558–61. [3] Minamino T, Miyauchi H, Yoshida T, Ishida Y, Yoshida H, Komuro I. Endothelial cell senescence in human atherosclerosis: role of telomere in endothelial dysfunction. Circulation 2002;105:1541–4. [4] Voghel G, Thorin-Trescases N, Farhat N, Nguyen A, Villeneuve L, Mamarbachi AM, et al. Cellular senescence in endothelial cells from atherosclerotic patients is accelerated by oxidative stress associated with cardiovascular risk factors. Mech Ageing Dev 2007;128:662–71. [5] Kirshenbaum LA, Singal PK. Antioxidant changes in heart hypertrophy: significance during hypoxia-reoxygenation injury. Can J Physiol Pharmacol 1992;70:1330–5. [6] McCord JM. Superoxide radical: controversies, contradictions, and paradoxes [see comments]. Proc Soc Exp Biol Med 1995;209:112–7. [7] Fridovich I, Freeman B. Antioxidant defenses in the lung. Annu Rev Physiol 1986; 48:693–702. [8] Richter C, Gogvadze V, Laffranchi R, Schlapbach R, Schweizer M, Suter M, et al. Oxidants in mitochondria: from physiology to diseases. Biochim Biophys Acta 1995; 1271:67–74. [9] Singal PK, Kirshenbaum LA. A relative deficit in antioxidant reserve may contribute in cardiac failure. Can J Cardiol 1990;6:47–9. [10] Min L, Jian-xing X. Detoxifying function of cytochrome c against oxygen toxicity. Mitochondrion 2007;7:13–6. [11] Cortial S, Chaignon P, Iorga BI, Aymerich S, Truan G, Gueguen-Chaignon V, et al. NADH oxidase activity of Bacillus subtilis nitroreductase NfrA1: insight into its biological role. FEBS Lett 2010;584:3916–22. [12] Naoi M, Maruyama W, Inaba-Hasegawa K. Type A and B monoamine oxidase in age-related neurodegenerative disorders: their distinct roles in neuronal death and survival. Curr Top Med Chem 2012;12:2177–88. [13] Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev 2014;94:909–50. [14] Feuerstein G, Yue TL, Ma X, Ruffolo RR. Novel mechanisms in the treatment of heart failure: inhibition of oxygen radicals and apoptosis by carvedilol. Prog Cardiovasc Dis 1998;41:17–24. [15] Singal PK, Li T, Kumar D, Danelisen I, Iliskovic N. Adriamycin-induced heart failure: mechanism and modulation. Mol Cell Biochem 2000;207:77–86. [16] Singal PK, Kirshenbaum LA. A relative deficit in antioxidant reserve may contribute in cardiac failure. Can J Cardiol 1990;6:47–9. [17] Giordano FJ. Oxygen, oxidative stress, hypoxia, and heart failure. J Clin Invest 2005; 115:500–8. [18] Harman D. Free radical theory of aging: an update: increasing the functional life span. Ann N Y Acad Sci 2006;1067:10–21. [19] Das KC, Muniyappa H. Age-dependent mitochondrial energy dynamics in the mice heart: role of superoxide dismutase-2. Exp Gerontol 2013;48:947–59. [20] Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science 2005;308:1909–11. [21] Zhao K, Zhao GM, Wu D, Soong Y, Birk AV, Schiller PW, et al. Cell-permeable peptide antioxidants targeted to inner mitochondrial membrane inhibit mitochondrial swelling, oxidative cell death, and reperfusion injury. J Biol Chem 2004;279: 34682–90. [22] Dai DF, Rabinovitch PS, Ungvari Z. Mitochondria and cardiovascular aging. Circ Res 2012;110:1109–24. [23] Chiara F, Castellaro D, Marin O, Petronilli V, Brusilow WS, Juhaszova M, et al. Hexokinase II detachment from mitochondria triggers apoptosis through the permeability transition pore independent of voltage-dependent anion channels. PLoS One 2008;3 e1852.

P

As stated above mitochondria are a major source of ROS which is a contributing factor to oxidative stress during aging. Proteotoxic stress, associated with senescence induced mitochondrial ROS production and oxidative stress injury results in the accumulation of oxidized proteins, contributing to cardiac dysfunction. To this date there is a growing body of evidence suggesting that autophagy and ubiquitin proteasome pathways become defective with age [98,99], resulting in accumulation of damaged mitochondria and misfolded proteins that negatively impact the overall function of the cell. It can be seen that in this context the removal of severely oxidized proteins or organelles within myocardium, by autophagy and ubiquitin proteasome pathways would be beneficial [100]. Aberrant mitochondria are removed in the process of mitophagy in order to sustain cardiovascular dynamics [26]. In fact in a recent study young versus old mice demonstrated the importance of autophagy for preventing the deleterious effects of senescence—induced cardiomyopathy [101]. In this study, a marked decline in Beclin-1 and ULK-1, along with increased levels of p62 levels indicative of impaired autophagy and cardiac dysfunction was observed in old versus younger animals. Bnip3 and related protein Nix (for Nip-like protein X) are known to induce mitophagy through the recruitment of LC3II/GABARAP/Atg8 proteins to the damaged mitochondria for removal by mitophagy [102,103]. Recently, an essential role for Nix and Bnip3 genes in mitochondrial quality control by eliminating damaged mitochondria by autophagy has been reported to be critical for preventing age induced cardiomyopathy [103,104]. In this study expression of Bnip3 and Nix was linked to adverse cardiac remodeling leading to cardiomyopathy, while mice deficient for both Nix and Bnip3 genes (Bnip3/Nix double knock-out) displayed age-dependent cardiac dysfunction and accumulation of aberrant mitochondria presumably from the lack of sufficient mitochondrial clearance—supporting the notion that autophagy/ mitophagy is essential for removing damaged organelles [105]. However, as stated earlier the exact role of Bnip3 and Nix in mitochondrial quality control and age-induced cardiomyopathy remains poorly understood. Previous work by our laboratory demonstrated that Bnip3 triggered mitochondrial perturbations resulting in mitochondrial depolarization, mPTP opening and cell death during hypoxia. Notably, the cell death induced by Bnip3 was abrogated by inhibiting autophagy with 3-MA or Atg7 knock-down [30,106]. These findings suggest that autophagy-induced by Bnip3 at least in this context was detrimental. Based on these findings as well as those of others we speculate that Bnip3 activation likely triggers the depolarization of mitochondria which then serves as a signal receptor for recruitment of GABARAP/ Atg8 and mitochondrial clearance [103,104]. We refer the reader to the article and works by Dr. Roberta Gottlieb in this issue for more in depth and detailed analysis of autophagy/mitophagy in the heart. Several reports have suggested interplay between mitochondrial dynamics and age induced mitophagy. Lee et al. described a mechanism by which Drp1 and E3-ubiquitin ligase Parkin are recruited to mitochondria in a manner dependent upon Bnip3. Moreover inhibition of Drp1 prevented Bnip3-induced autophagy and expression of dominant negative mutant Drp1K38E or Mfn1 also prevented Bnip3 induced mitochondrial fission and autophagy [104], suggesting that fission of mitochondria is a critical stage for clearance by mitophagy. Mfn2 is a key protein of mitochondrial fusion that is regulated by PTEN-Induced Putative Kinase protein 1 (PINK1)/Parkin mitophagy system [107,108]. Accordingly, this study demonstrated that Mfn2 was critical for removal of defective mitochondria by serving as a mitochondrial docking site for ubiquitin E3-ligase Parkin. Interestingly, these elegant studies demonstrated that mitochondrial associated PINK1 acts on both Mfn2 and Parkin, it phosphorylates Mfn2 thereby recruiting Parkin to mitochondria whereby Parkin then ubiquitinates Mfn2 [107,109]. Notably the absence of Mfn2 prevented Parkin translocation to mitochondria resulting in accumulation of dysfunctional mitochondria and impaired respiration. The association of age related defects in mitochondrial dynamics

463 464

D

399 400

appears to be a critical nodal point in the maintenance of a viable normal regulation of myocardial function. Over the next several years it will be interesting to see whether interventions that selectively target metabolic signaling and mitochondrial dynamics will be sufficient to rejuvenate the senescent heart improving cardiovascular health and longevity and bring us a little closer to the elusive “Fountain of Youth”.

E

7. Mitophagy in the senescent heart

T

398

5

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

476 Q7 Q8 Q9 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534

D

P

R O

O

F

[55] Heilbronn LK, Ravussin E. Calorie restriction and aging: review of the literature and implications for studies in humans. Am J Clin Nutr 2003;78:361–9. [56] Haigis MC, Guarente LP. Mammalian sirtuins—emerging roles in physiology, aging, and calorie restriction. Genes Dev 2006;20:2913–21. [57] Rohrbach S, Aslam M, Niemann B, Schulz R. Impact of caloric restriction on myocardial ischaemia/reperfusion injury and new therapeutic options to mimic its effects. Br J Pharmacol 2014;171:2964–92. [58] Ghosh HS. The anti-aging, metabolism potential of SIRT1. Curr Opin Investig Drugs 2008;9:1095–102. [59] Chen D, Steele AD, Hutter G, Bruno J, Govindarajan A, Easlon E, et al. The role of calorie restriction and SIRT1 in prion-mediated neurodegeneration. Exp Gerontol 2008;43:1086–93. [60] Chen D, Bruno J, Easlon E, Lin SJ, Cheng HL, Alt FW, et al. Tissue-specific regulation of SIRT1 by calorie restriction. Genes Dev 2008;22:1753–7. [61] Jadiya P, Chatterjee M, Sammi SR, Kaur S, Palit G, Nazir A. Sir-2.1 modulates ‘calorie-restriction-mediated’ prevention of neurodegeneration in Caenorhabditis elegans: implications for Parkinson's disease. Biochem Biophys Res Commun 2011;413:306–10. [62] Luo XY, Qu SL, Tang ZH, Zhang Y, Liu MH, Peng J, et al. SIRT1 in cardiovascular aging. Clin Chim Acta 2014;437:106–14. [63] Morris BJ. Seven sirtuins for seven deadly diseases of aging. Free Radic Biol Med 2013;56:133–71. [64] Nogueiras R, Habegger KM, Chaudhary N, Finan B, Banks AS, Dietrich MO, et al. Sirtuin 1 and sirtuin 3: physiological modulators of metabolism. Physiol Rev 2012;92:1479–514. [65] Fabrizio P, Gattazzo C, Battistella L, Wei M, Cheng C, McGrew K, et al. Sir2 blocks extreme life-span extension. Cell 2005;123:655–67. [66] Franco SS, De FL, Ghaffari S, Brugnara C, Sinclair DA, Matte' A. Resveratrol accelerates erythroid maturation by activation of FoxO3 and ameliorates anemia in betathalassemic mice. Haematologica 2014;99:267–75. [67] Warboys CM, de LA Amini N, Luong L, Duckles H, Hsiao S, et al. Disturbed flow promotes endothelial senescence via a p53-dependent pathway. Arterioscler Thromb Vasc Biol 2014;34:985–95. [68] Gano LB, Donato AJ, Pasha HM, Hearon Jr CM, Sindler AL, Seals DR. The SIRT1 activator SRT1720 reverses vascular endothelial dysfunction, excessive superoxide production, and inflammation with aging in mice. Am J Physiol Heart Circ Physiol 2014;307:H1754–63. [69] Zwerschke W, Mazurek S, Stockl P, Hutter E, Eigenbrodt E, Jansen-Durr P. Metabolic analysis of senescent human fibroblasts reveals a role for AMP in cellular senescence. Biochem J 2003;376:403–11. [70] Hahn-Windgassen A, Nogueira V, Chen CC, Skeen JE, Sonenberg N, Hay N. Akt activates the mammalian target of rapamycin by regulating cellular ATP level and AMPK activity. J Biol Chem 2005;280:32081–9. [71] Wu JJ, Liu J, Chen EB, Wang JJ, Cao L, Narayan N, et al. Increased mammalian lifespan and a segmental and tissue-specific slowing of aging after genetic reduction of mTOR expression. Cell Rep 2013;4:913–20. [72] Dhingra R, Gang H, Wang Y, Biala AK, Aviv Y, Margulets V, et al. Bidirectional regulation of nuclear factor-kappaB and mammalian target of rapamycin signaling functionally links Bnip3 gene repression and cell survival of ventricular myocytes. Circ Heart Fail 2013;6:335–43. [73] Atkins KM, Thomas LL, Barroso-Gonzalez J, Thomas L, Auclair S, Yin J, et al. The multifunctional sorting protein PACS-2 regulates SIRT1-mediated deacetylation of p53 to modulate p21-dependent cell-cycle arrest. Cell Rep 2014;8:1545–57. [74] Warboys CM, de LA Amini N, Luong L, Duckles H, Hsiao S, et al. Disturbed flow promotes endothelial senescence via a p53-dependent pathway. Arterioscler Thromb Vasc Biol 2014;34:985–95. [75] Tanno M, Kuno A, Yano T, Miura T, Hisahara S, Ishikawa S, et al. Induction of manganese superoxide dismutase by nuclear translocation and activation of SIRT1 promotes cell survival in chronic heart failure. J Biol Chem 2010;285:8375–82. [76] Hafner AV, Dai J, Gomes AP, Xiao CY, Palmeira CM, Rosenzweig A, et al. Regulation of the mPTP by SIRT3-mediated deacetylation of CypD at lysine 166 suppresses age-related cardiac hypertrophy. Aging (Albany NY) 2010;2:914–23. [77] Porter GA, Urciuoli WR, Brookes PS, Nadtochiy SM. SIRT3 deficiency exacerbates ischemia-reperfusion injury: implication for aged hearts. Am J Physiol Heart Circ Physiol 2014;306:H1602–9. [78] Dorn GW, Scorrano L. Two close, too close: sarcoplasmic reticulum-mitochondrial crosstalk and cardiomyocyte fate. Circ Res 2010;107:689–99. [79] Eschenbacher WH, Song M, Chen Y, Bhandari P, Zhao P, Jowdy CC, et al. Two rare human mitofusin 2 mutations alter mitochondrial dynamics and induce retinal and cardiac pathology in Drosophila. PLoS One 2012;7:e44296. [80] Okamoto K, Shaw JM. Mitochondrial morphology and dynamics in yeast and multicellular eukaryotes. Annu Rev Genet 2005;39:503–36. [81] Bossy-Wetzel E, Barsoum MJ, Godzik A, Schwarzenbacher R, Lipton SA. Mitochondrial fission in apoptosis, neurodegeneration and aging. Curr Opin Cell Biol 2003; 15:706–16. [82] Iglewski M, Hill JA, Lavandero S, Rothermel BA. Mitochondrial fission and autophagy in the normal and diseased heart. Curr Hypertens Rep 2010;12:418–25. [83] Mopert K, Hajek P, Frank S, Chen C, Kaufmann J, Santel A. Loss of Drp1 function alters OPA1 processing and changes mitochondrial membrane organization. Exp Cell Res 2009;315:2165–80. [84] James DI, Parone PA, Mattenberger Y, Martinou JC. hFis1, a novel component of the mammalian mitochondrial fission machinery. J Biol Chem 2003;278: 36373–9. [85] Yoon Y, Krueger EW, Oswald BJ, McNiven MA. The mitochondrial protein hFis1 regulates mitochondrial fission in mammalian cells through an interaction with the dynamin-like protein DLP1. Mol Cell Biol 2003;23:5409–20.

N

C

O

R

R

E

C

T

[24] Dorn GW, Kitsis RN. The mitochondrial dynamism-mitophagy-cell death interactome: multiple roles performed by members of a mitochondrial molecular ensemble. Circ Res 2015;116:167–82. [25] Chen H, Chomyn A, Chan DC. Disruption of fusion results in mitochondrial heterogeneity and dysfunction. J Biol Chem 2005;280:26185–92. [26] Jimenez RE, Kubli DA, Gustafsson AB. Autophagy and mitophagy in the myocardium: therapeutic potential and concerns. Br J Pharmacol 2013. [27] Pohjoismaki JL, Boettger T, Liu Z, Goffart S, Szibor M, Braun T. Oxidative stress during mitochondrial biogenesis compromises mtDNA integrity in growing hearts and induces a global DNA repair response. Nucleic Acids Res 2012;40: 6595–607. [28] Das S, Mitrovsky G, Vasanthi HR, Das DK. Antiaging properties of a grape-derived antioxidant are regulated by mitochondrial balance of fusion and fission leading to mitophagy triggered by a signaling network of Sirt1–Sirt3–Foxo3–PINK1– PARKIN. Oxid Med Cell Longev 2014;2014:345105. [29] Mughal W, Dhingra R, Kirshenbaum LA. Striking a balance: autophagy, apoptosis, and necrosis in a normal and failing heart. Curr Hypertens Rep 2012;14:540–7. [30] Wang EY, Gang H, Aviv Y, Dhingra R, Margulets V, Kirshenbaum LA. p53 mediates autophagy and cell death by a mechanism contingent on Bnip3. Hypertension 2013;62:70–7. [31] Dhingra R, Kirshenbaum LA. Mst-1 switches between cardiac cell life and death. Nat Med 2013;19:1367–8. [32] Weidman D, Shaw J, Bednarczyk J, Regula KM, Yurkova N, Zhang T, et al. Dissecting apoptosis and intrinsic death pathways in the heart. Methods Enzymol 2008;446: 277–85. [33] 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. [34] Haendeler J, Hoffmann J, Diehl JF, Vasa M, Spyridopoulos I, Zeiher AM, et al. Antioxidants inhibit nuclear export of telomerase reverse transcriptase and delay replicative senescence of endothelial cells. Circ Res 2004;94:768–75. [35] Wei YH, Wu SB, Ma YS, Lee HC. Respiratory function decline and DNA mutation in mitochondria, oxidative stress and altered gene expression during aging. Chang Gung Med J 2009;32:113–32. [36] Kujoth GC, Hiona A, Pugh TD, Someya S, Panzer K, Wohlgemuth SE, et al. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science 2005;309:481–4. [37] Trifunovic A, Hansson A, Wredenberg A, Rovio AT, Dufour E, Khvorostov I, et al. Somatic mtDNA mutations cause aging phenotypes without affecting reactive oxygen species production. Proc Natl Acad Sci U S A 2005;102:17993–8. [38] Zhang D, Mott JL, Chang SW, Denniger G, Feng Z, Zassenhaus HP. Construction of transgenic mice with tissue-specific acceleration of mitochondrial DNA mutagenesis. Genomics 2000;69:151–61. [39] Kraytsberg Y, Nekhaeva E, Bodyak NB, Khrapko K. Mutation and intracellular clonal expansion of mitochondrial genomes: two synergistic components of the aging process? Mech Ageing Dev 2003;124:49–53. [40] Ishihara T, Ban-Ishihara R, Maeda M, Matsunaga Y, Ichimura A, Kyogoku S, et al. Dynamics of mitochondrial DNA nucleoids regulated by mitochondrial fission is essential for maintenance of homogeneously active mitochondria during neonatal heart development. Mol Cell Biol 2015;35:211–23. [41] Katada S, Mito T, Ogasawara E, Hayashi J, Nakada K. Mitochondrial DNA with a large-scale deletion causes two distinct mitochondrial disease phenotypes in mice. G3 (Bethesda), 3; 2013 1545–52. [42] Tam ZY, Gruber J, Halliwell B, Gunawan R. Mathematical modeling of the role of mitochondrial fusion and fission in mitochondrial DNA maintenance. PLoS One 2013;8:e76230. [43] Leri A, Malhotra A, Liew CC, Kajstura J, Anversa P. Telomerase activity in rat cardiac myocytes is age and gender dependent. J Mol Cell Cardiol 2000;32:385–90. [44] Rhyu MS. Telomeres, telomerase, and immortality [see comments]. J Natl Cancer Inst 1995;87:884–94. [45] Sussman MA, Anversa P. Myocardial aging and senescence: where have the stem cells gone? Annu Rev Physiol 2004;66:29–48. [46] Bhayadia R, Schmidt BM, Melk A, Homme M. Senescence-induced oxidative stress causes endothelial dysfunction. J Gerontol A Biol Sci Med Sci 2015. [47] Chen J, Huang X, Halicka D, Brodsky S, Avram A, Eskander J, et al. Contribution of p16INK4a and p21CIP1 pathways to induction of premature senescence of human endothelial cells: permissive role of p53. Am J Physiol Heart Circ Physiol 2006;290:H1575–86. [48] Holdt LM, Sass K, Gabel G, Bergert H, Thiery J, Teupser D. Expression of Chr9p21 genes CDKN2B (p15(INK4b)), CDKN2A (p16(INK4a), p14(ARF)) and MTAP in human atherosclerotic plaque. Atherosclerosis 2011;214:264–70. [49] Serrano M, Lee H, Chin L, Cordon-Cardo C, Beach D, DePinho RA. Role of the INK4a locus in tumor suppression and cell mortality. Cell 1996;85:27–37. [50] Wei W, Hemmer RM, Sedivy JM. Role of p14(ARF) in replicative and induced senescence of human fibroblasts. Mol Cell Biol 2001;21:6748–57. [51] Westhoff JH, Hilgers KF, Steinbach MP, Hartner A, Klanke B, Amann K, et al. Hypertension induces somatic cellular senescence in rats and humans by induction of cell cycle inhibitor p16INK4a. Hypertension 2008;52:123–9. [52] Cesselli D, Beltrami AP, D'Aurizio F, Marcon P, Bergamin N, Toffoletto B, et al. Effects of age and heart failure on human cardiac stem cell function. Am J Pathol 2011;179: 349–66. [53] Leri A, Franco S, Zacheo A, Barlucchi L, Chimenti S, Limana F, et al. Ablation of telomerase and telomere loss leads to cardiac dilatation and heart failure associated with p53 upregulation. EMBO J 2003;22:131–9. [54] Chang E, Harley CB. Telomere length and replicative aging in human vascular tissues. Proc Natl Acad Sci U S A 1995;92:11190–4.

U

535 536 537 538 539 540 541 Q10 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 Q11 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620

A.K. Biala et al. / Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

E

6

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706

A.K. Biala et al. / Journal of Molecular and Cellular Cardiology xxx (2015) xxx–xxx

[98] Johnson D, Allman E, Nehrke K. Regulation of acid–base transporters by reactive oxygen species following mitochondrial fragmentation. Am J Physiol Cell Physiol 2012;302:C1045–54. [99] Wohlgemuth SE, Seo AY, Marzetti E, Lees HA, Leeuwenburgh C. Skeletal muscle autophagy and apoptosis during aging: effects of calorie restriction and life-long exercise. Exp Gerontol 2010;45:138–48. [100] Dutta D, Calvani R, Bernabei R, Leeuwenburgh C, Marzetti E. Contribution of impaired mitochondrial autophagy to cardiac aging: mechanisms and therapeutic opportunities. Circ Res 2012;110:1125–38. [101] Joseph AM, Adhihetty PJ, Wawrzyniak NR, Wohlgemuth SE, Picca A, Kujoth GC, et al. Dysregulation of mitochondrial quality control processes contribute to sarcopenia in a mouse model of premature aging. PLoS One 2013;8:e69327. [102] Dorn GW. Nix nought nothing: fairy tale or real deal. J Mol Cell Cardiol 2010. [103] 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. [104] 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. [105] Dorn GW, Kirshenbaum LA. Cardiac reanimation: targeting cardiomyocyte death by BNIP3 and NIX/BNIP3L. Oncogene 2008;27(Suppl. 1):S158–67. [106] Dhingra R, Margulets V, Chowdhury SR, Thliveris J, Jassal D, Fernyhough P, et al. Bnip3 mediates doxorubicin-induced cardiac myocyte necrosis and mortality through changes in mitochondrial signaling. Proc Natl Acad Sci U S A 2014;111: E5537–44. [107] Chen Y, Dorn GW. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science 2013;340:471–5. [108] Jin SM, Youle RJ. The accumulation of misfolded proteins in the mitochondrial matrix is sensed by PINK1 to induce PARK2/Parkin-mediated mitophagy of polarized mitochondria. Autophagy 2013;9:1750–7. [109] Bhandari P, Song M, Chen Y, Burelle Y, Dorn GW. Mitochondrial contagion induced by parkin deficiency in Drosophila hearts and its containment by suppressing mitofusin. Circ Res 2014;114:257–65.

N C O

R

R

E

C

T

E

D

P

R O

O

F

[86] Chen H, Detmer SA, Ewald AJ, Griffin EE, Fraser SE, Chan DC. Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development. J Cell Biol 2003;160:189–200. [87] Chen H, Chan DC. Physiological functions of mitochondrial fusion. Ann N Y Acad Sci 2010;1201:21–5. [88] Olichon A, Guillou E, Delettre C, Landes T, Arnaune-Pelloquin L, Emorine LJ, et al. Mitochondrial dynamics and disease, OPA1. Biochim Biophys Acta 1763;2006: 500–9. [89] Ishihara N, Nomura M, Jofuku A, Kato H, Suzuki SO, Masuda K, et al. Mitochondrial fission factor Drp1 is essential for embryonic development and synapse formation in mice. Nat Cell Biol 2009;11:958–66. [90] Alavi MV, Bette S, Schimpf S, Schuettauf F, Schraermeyer U, Wehrl HF, et al. A splice site mutation in the murine Opa1 gene features pathology of autosomal dominant optic atrophy. Brain 2007;130:1029–42. [91] Dorn GW, Clark CF, Eschenbacher WH, Kang MY, Engelhard JT, Warner SJ, et al. MARF and Opa1 control mitochondrial and cardiac function in Drosophila. Circ Res 2011;108:12–7. [92] Zhao L, Zou X, Feng Z, Luo C, Liu J, Li H, et al. Evidence for association of mitochondrial metabolism alteration with lipid accumulation in aging rats. Exp Gerontol 2014;56:3–12. [93] Bhandari P, Song M, Chen Y, Burelle Y, Dorn GW. Mitochondrial contagion induced by Parkin deficiency in Drosophila hearts and its containment by suppressing mitofusin. Circ Res 2014;114:257–65. [94] Park YY, Lee S, Karbowski M, Neutzner A, Youle RJ, Cho H. Loss of MARCH5 mitochondrial E3 ubiquitin ligase induces cellular senescence through dynaminrelated protein 1 and mitofusin 1. J Cell Sci 2010;123:619–26. [95] AU D-MM, AU DI, AU SL AUDP, AU LJ, AU MM, et al. TI - zeta PKC induces phosphorylation and inactivation of I kappa B-alpha, in vitro; 1997 8(None specified). [96] Giedt RJ, Yang C, Zweier JL, Matzavinos A, Alevriadou BR. Mitochondrial fission in endothelial cells after simulated ischemia/reperfusion: role of nitric oxide and reactive oxygen species. Free Radic Biol Med 2012;52:348–56. [97] Figge MT, Reichert AS, Meyer-Hermann M, Osiewacz HD. Deceleration of fusion– fission cycles improves mitochondrial quality control during aging. PLoS Comput Biol 2012;8:e1002576.

U

707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 773

7

Please cite this article as: Biala AK, et al, Mitochondrial dynamics: Orchestrating the journey to advanced age, J Mol Cell Cardiol (2015), http:// dx.doi.org/10.1016/j.yjmcc.2015.04.015

741 742 743 744 745 746 747 748 749 750 751 752 753 Q12 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772

Mitochondrial dynamics: Orchestrating the journey to advanced age.

Aging is a degenerative process that unfortunately is an inevitable part of life and risk factor for cardiovascular disease including heart failure. A...
686KB Sizes 0 Downloads 6 Views