Sie sind auf Seite 1von 14

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56

AUTOPHAGY IN HEALTH AND DISEASE: V. Mitophagy as a Way of Life


Roberta A. Gottlieb, Raquel S. Carreira
Abstract Our understanding of autophagy has expanded greatly in recent years, largely due to the identification of the many genes involved in the process, and to the development of better methods to monitor the process, such as GFP-LC3 to visualize autophagosomes in vivo. A number of groups have demonstrated a tight connection between autophagy and mitochondrial turnover. Mitochondrial quality control is the process whereby mitochondria undergo successive rounds of fusion and fission with a dynamic exchange of components in order to segregate functional and damaged elements. Removal of the mitochondrion that contains damaged components is accomplished via autophagy (mitophagy). Mitophagy also serves to eliminate the subset of mitochondria producing the most reactive oxygen species, and episodic removal of mitochondria will reduce the oxidative burden, thus linking the mitochondrial free radical theory of aging with longevity achieved through caloric restriction. Mitophagy must be balanced by biogenesis to meet tissue energy needs, but the system is tunable and highly dynamic. This process is of greatest importance in long-lived cells such as cardiomyocytes, neurons, and memory T cells. Autophagy is known to decrease with age, and the failure to maintain mitochondrial quality control through mitophagy may explain why the heart, brain, and components of the immune system are most vulnerable to dysfunction as organisms age. Autophagy machinery and regulation Many reviews have covered the details of autophagy in recent years (32, 72); the key elements are depicted in Fig. 1. For the purpose of this discussion it is important to recognize that there are a growing number of adaptor proteins that serve to recruit the autophagosome to its target. For instance, p62 binds to ubiquitinated proteins and links autophagosome membranes through an interaction with LC3 (44). Ubiquitin ligases restricted to specific subcellular locations, such as MULAN and Parkin on mitochondria, may serve to target select organelles for autophagic destruction, although whether this is a selective process, and how selection is accomplished, remain to be elucidated (28, 58). Chaperone-mediated autophagy, cytoplasm to vacuole targeting, and macroautophagy are examples of selective forms of autophagy. Macroautophagy (hereafter referred to as autophagy) includes the selective elimination of mitochondria (mitophagy), endoplasmic reticulum (reticulophagy), peroxisomes (pexophagy), ribosomes (ribophagy), granules (crinophagy), and pathogens (xenophagy) (1). Cytosol, cytoskeleton, nuclei (nucleophagy), and protein aggregates (aggrephagy) can also be removed by autophagy (85). The most universal stressor to induce autophagy is starvation. Although a number of pathways may be involved, AMP-activated kinase (AMPK) is clearly an important element. AMPK exerts a negative regulatory effect on mTOR, thereby releasing this powerful brake to the induction of autophagy (20). In addition to the loss of an inhibitory signal, AMPK triggers SIRT1-dependent deacetylation of the transcriptional regulators PGC-1alpha and FOXO1, culminating in the transcriptional modulation of mitochondrial biogenesis (7). Thus autophagy and mitochondrial biogenesis are coordinately regulated. Autophagy is also induced by oxidative stress, and amino acid restriction has been shown to trigger production of reactive oxygen species (ROS). Chen et al. showed that the ROS originate from mitochondria (10), and Elazar identified Atg4 as the component of the autophagy machinery that responds to ROS (90). While this is a global and nonspecific response, we hypothesize that ubiquitin ligases associated with the mitochondrial outer membrane might respond to the local production of ROS, thereby conferring specificity to the process. Autophagy is also induced by failure of the ubiquitin-proteasome system (UPS) (51). Based on the extremely rapid turnover of key rate-limiting components of the autophagy machinery, even transient disruption of UPSmediated protein degradation may be sufficient to upregulate autophagy (27). The UPS is sensitive to oxidative stress and therefore will amplify the autophagic response to ROS (111). Interestingly, both Beclin1 and Atg5 contain a Bcl-2-binding domain that resembles the BH3 domain of many of the pro-apoptotic members of the Bcl-2 family. Beclin1 binding by Bcl-2 is reported to negatively regulate autophagy (76), and displacement of Bcl-2 by BH3-only proteins can stimulate autophagy (24, 63). The Bcl-2-

Formatted: Numbering: Continuous

57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112

binding domain appears to be necessary for autophagy (5). Given the importance of Bcl-2 family members in regulating mitochondrial integrity, it seems reasonable to hypothesize that Bcl-2 proteins might also govern selective mitophagy. Mitochondria and their dark side While mitochondria perform essential functions for the cell, notably ATP production via oxidative phosphorylation, heme biosynthesis, and calcium homeostasis, their continued existence within the cell may sometimes seem like a pact with the devil. Damage to the mitochondrial outer membrane leads to release of cytochrome c, triggering caspase activation and apoptosis. More catastrophic stresses can lead to pathologic opening of the mitochondrial permeability transition pore (MPTP), accompanied by transient but massive release of ROS and calcium (4, 112). This can trigger neighboring mitochondria to do the same, culminating in activation of calcium-dependent proteases (calpains) and lipases (cPLA2), as well as ROS-activated iPLA2, which together ensure the necrotic destruction of the cell (64). However, even when mitochondria are functionally normal, 1-2% of the oxygen they consume is converted to superoxide and then to hydrogen peroxide. Akin to cars that emit smog even when idling, mitochondria produce small amounts of superoxide even when ATP production is minimal. Damaged but still functional mitochondria might release up to ten-fold more hydrogen peroxide (30). In a mitochondria-rich, metabolically active organ like the heart, this will impose a substantial oxidative burden over time. Given the mitochondrial half-life of two weeks and a cardiomyocyte lifespan of many decades, the organism is confronted with a sizable challenge to deal with the cumulative oxidative damage in its longest-lived tissues. Oxidative stress will cause DNA damage, and mitochondrial DNA repair enzymes are less efficient (55). A defective version of mitochondrial DNA polymerase gamma (involved in synthesis and repair) introduces a high frequency of mutations in the mitochondrial genome. A mouse model expressing the mutant polymerase gives rise to a phenotype of accelerated aging (104). Mitophagy as a general feature of autophagy Mitochondrial integrity is essential to cellular homeostasis. If the energetic demand is low, excess mitochondria are unnecessary and may generate excessive reactive oxygen species. If they become uncoupled, they can actually consume ATP (106). Therefore, their elimination by autophagy is an efficient cytoprotective response. Furthermore, damaged or unstable mitochondria may release ROS, cytochrome c, AIF, SMAC/DIABLO, and other apoptosis-promoting factors which would promote damage to neighboring mitochondria and the entire cell (15). In this setting, one can imagine that elimination of a few unstable and potentially dangerous mitochondria by autophagy will preserve the integrity of the remaining population and maintain cellular homeostasis. In fact, autophagic removal of mitochondria has been shown to occur following both induction (107) and inhibition of apoptosis (108), suggesting that mitophagy is required for cell survival following mitochondria injury. The first description of mitochondria within an autophagosome in mammalian cells was made in 1957, in kidneys of newborn mice (13). In the heart, the first report was in 1980, following perfusion of rabbit hearts under hypoxic conditions (19). At this time the authors were already suggesting that autophagy of mitochondria was important in the repair of damaged cardiac cells (19). Since then, mitophagy has been shown to have a role in mitochondrial and cellular quality control (23, 74, 79, 101) (8), in cellular differentiation [differentiation of reticulocytes and T lymphocytes (73, 89, 91, 110)], and in disease [Parkinson Disease (70)], and therefore mitophagy has been the center of growing attention. Signals for mitophagy In budding yeast, mitophagy requires Atg11, the selective autophagy-specific factor, but not Atg19, the Cvt cargo receptor, showing that mitophagy is a selective process (46). Kissova et al. (49, 50) reported that the mitochondrial outer membrane protein Uth1, in coordination with Atg proteins, mediates the selective removal of mitochondria, and the inactivation of the UTH1 gene eliminates selective mitophagy. Aup1, a putative protein phosphatase localized in the mitochondrial intermembrane space, was also suggested to be critical for mitophagy (94). However, Okamoto et al. (75) using post-log phase yeast cells studied under conditions favoring oxidative phosphorylation did not confirm the need for Aup1 and Uth1 for the selective elimination of mitochondria. Okamoto et al. (75) together with Kanki et al. (47) suggested the existence of a new

113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168

mitochondrial outer membrane protein necessary for the selective autophagic degradation of mitochondria Atg32. Atg32 mutant cells are defective in mitophagy, but competent in bulk autophagy and the Cvt pathway, which indicates that Atg32 is specific for mitophagy (47, 75). Atg32 serves as a mitochondrial surface receptor and binds to Atg11 and Atg8 to direct selective mitophagy . The mammalian homologs for Atg32 and UTH1 have not been identified. The selective removal of mitochondria also appears to occur in mammals, mediated by the mitochondria permeability transition pore (8, 23, 87, 101) and apoptotic proteins, such as Bnip3 (Bcl-2 and adenovirus E1B 19 kDa-interacting protein 3) and NIX (Bnip3-like protein X) (73, 89, 91, 110). The mitochondrial permeability transition pore (MPTP) The pathological role of the MPTP is well known [for review see (33, 36, 84)]. Opening of the MPTP (this phenomenon is termed the mitochondrial permeability transition or MPT) is a common response to ischemiareperfusion injury, particularly to stresses such as ROS and calcium overload. MPT makes the inner mitochondrial membrane permeable to solutes of up to 1,500 Da (2), resulting in depolarization due to dissipation of the electrochemical gradient, which in turn causes ATP synthase to operate in reverse, consuming ATP (106). However, recent evidence suggests a physiological role for the MPTP in: i) calcium homeostasis (2, 43); ii) cardioprotection against ischemia-reperfusion (IR) injury provided by ischemic and pharmacological preconditioning and postconditioning (39, 59); and iii) mitochondrial removal by autophagy (23, 87, 101). Mitochondria are known to be degraded by the autophagosomal-lysosomal pathway, but the basis on which individual mitochondria are selected for autophagy is unknown. Lemasters group (23, 87) demonstrated that induction of autophagy in rat hepatocytes by serum deprivation and glucagon causes an increase of spontaneously depolarizing mitochondria, and these mitochondria are sequestered by autophagosomes. The authors concluded that the MPTP is responsible for mitochondrial depolarization and that it leads to mitochondrial sequestration into autophagosomes. Similar results were obtained using nicotinamide-treated human fibroblasts (45). Nicotinamide-activated autophagy selectively removes mitochondria with low membrane potential, which is attenuated by treatment with cyclosporin A (45). Twig and colleagues (101) showed that in pancreatic beta cells, fission generates asymmetric daughter mitochondria: one subpopulation has increased membrane potential and high probability of fusion, while the other has decreased potential and reduced probability of fusion (101). Dysfunctional mitochondria are excluded from subsequent rounds of fission and fusion, and eventually are removed by autophagy. The authors argued against the MPTP as the cause of the depolarization because the depolarization was observed even in the presence of 1 M cyclosporin A. In cardiac cells, starvation-induced autophagy has been shown to cause mitochondrial depolarization which is prevented by cyclosporin A (8), indicating the MPTP is involved in starvation-induced mitochondrial depolarization. We have also shown that cyclophilin D, a component of the MPTP , is required for cardiac mitochondrial removal by autophagy induced by starvation (8). Cardiomyocytes from cyclophilin D-deficient mice do not upregulate autophagy when subjected to starvation, in contrast to cardiomyocytes from wild-type mice (8). Cyclophilin D and the MPTP are implicated in the initiation of autophagy and removal of mitochondria in starvation-induced autophagy and may represent a homeostatic mechanism for cellular and mitochondrial quality control. Apoptotic proteins Bnip3 and Bnip3-like (Bnip3L), also known as Nix, are proteins with homology to Bcl-2 in the BH3 domain, which induce both cell death and autophagy. Bnip3 and Nix can induce mitophagy by triggering mitochondrial depolarization, which is known to induce mitochondrial removal by autophagy (23, 101). Mitochondrial depolarization has been shown to be both sensitive and insensitive to MPTP inhibitors (80), and therefore there is a debate whether the induction of autophagy by Nix and Bnip3 is through the pore. Autophagy is induced as a protective response to Bnip3-mediated apoptotic signaling (Figure 2) (37, 38, 66). Bnip3 is activated following ischemia-reperfusion and is responsible for loss of cardiac myocytes. It functions as a redox sensor where increased oxidative stress induces homodimerization and activation of Bnip3 (54). It has been reported that Bnip3 mediates cell death via opening of the MPTP (53, 86, 102). However, it was recently described that in isolated heart mitochondria Bnip3-mediated mitochondrial permeabilization, swelling and cytochrome c release is insensitive to cyclosporin A, an inhibitor of the MPTP and independent of cyclophilin D, an essential component of the MPTP (80). Moreover, Bnip3 causes mitochondrial matrix

169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224

remodeling and large amplitude swelling of the inner membrane, which leads to disassembly of OPA1 complexes and release from the mitochondria (80). Given the wide array of consequences of Bnip3 activation, it may induce autophagy by several mechanisms, including via mitochondrial depolarization and opening of the MPTP or through interference with the fusion-fission machinery. NIX is required for programmed mitophagy during reticulocyte maturation (89, 91, 110). The maturation process includes elimination of membrane bound organelles, such as mitochondria (31, 52). Mitochondria are cleared from reticulocytes through an autophagy-related process with the major difference that the contents of the autophagic vacuole are not recycled but are eliminated by exocytosis (31, 40). Novak et al. (73) showed that in murine reticulocytes, Nix functions as a selective autophagy receptor that binds to LC3/GABARAP proteins. Autophagy machinery is still functional in Nix-deficient reticulocytes, but mitochondria are not engulfed by the autophagosome. The authors speculate that Nix-dependent recruitment of autophagosomeassociated LC3/GABARAP proteins to mitochondria might mediate membrane tethering and/or hemifusion of mitochondria with autophagosomes (73). This could be the reason why mitochondria align to autophagosomes, but are not engulfed in Nix-deficient reticulocytes (73). NIX-deficient mice exhibit defective erythroid development, mild-to-moderate anemia, reticulocytosis, and an increase in splenic erythropoiesis (21, 89, 91). Interestingly, knockout of autophagy-specific genes (Ulk1, Atg5 and Atg7) does not prevent mitophagy during reticulocyte maturation, suggesting there is an alternative pathway for mitochondrial removal (56, 65, 110). Mitochondrial turnover and role of bouts of autophagy Mitochondrial biogenesis consists of the growth and division of pre-existing mitochondria. According to the endosymbiotic theory, mitochondria descend from a proteobacteria endosymbiont that became established in a eukaryotic host cell. Given their bacterial origin, mitochondria have their own genome, and mitochondrial proteins are encoded by both nuclear and mitochondrial genomes (103). Mitochondrial biogenesis requires the coordination of several distinct processes: i) inner and outer mitochondrial membrane synthesis; ii) synthesis of mitochondrial proteins; iii) synthesis and import of proteins encoded by the nuclear genome; iv) replication of mitochondrial DNA; and v) mitochondrial fusion and fission. Mitochondrial turnover comprises mitophagy followed by biogenesis. Mitochondria were proposed to turn over as a unit, since inner membrane proteins such as cytochrome aa3, b and c, the inner membrane lipid cardiolipin, and mitochondrial DNA in the matrix all have the same half life (29, 81, 82). The mitochondrial halflive values found in the literature are very disparate. Reports from the early 1970s suggest that under normal conditions, mitochondria of non-proliferating tissues (e.g., brain, heart, kidney and liver) turn over with a halflife of 1025 days (69, 77), while others suggest ~5-6 days for rat heart and liver (82). The methods used to determine mitochondrial turnover, usually electron microscopy and radioactive-labeled components (81), present several problems. Electron microscopy, which only shows a snapshot of a section of the cell, also fails to provide information about mitochondria viability mitochondria might have normal shape, but not be functional and therefore will be targeted for removal (100). The rate of decay of radioactively labeled components of the mitochondria is used as an indicator of mitochondrial destruction (81). However, some of the isotopic precursors used in such experiments can be reutilized for resynthesis of macromolecules, and the decay rates will vary depending on whether the isotope is reutilized or not (81). Given the technical limitations, new methods to determine mitochondrial turnover are warranted; progress on this question awaits the development of better approaches. At present, the rate of mitochondrial turnover is not definitively known. Mitochondrial fusion, fission, autophagy and biogenesis Another limitation of the techniques referred in the previous section is that they do not reflect mitochondrial dynamics. Mitochondria are under constant cycles of fission and fusion; if subjected to stress they may be removed by autophagy, and biogenesis might increase to meet energetic demand (42). The removal of mitochondria under stress conditions can serve to supply the cell with amino acids, rid the cells of damaged mitochondria that are deleterious, or prevent cell death (3). A drastic depletion of mitochondriaa purge might be useful to eliminate mtDNA copies that contain enough mutations to interfere with function. Such a bottleneck has been proposed to explain the observation that of all the mtDNA copies present in the germ-cell precursor, only a fraction will be amplified to generate the approximately 105 mtDNA copies present in the mature oocyte. Studies with the polymerase gamma mutator mouse show that mutations that do not result in

225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280

amino acid changes are tolerated, while mutations causing amino acid changes are strongly underrepresented in the protein-coding genes, providing incontrovertible evidence that mtDNA is subject to strong purifying selection in the maternal germ line, and hints at a fitness test of the mitochondria that express proteins encoded by the mitochondrial genome. This purifying selection of functional mtDNA genomes will decrease the frequency of transmission of mutated genomes to the offspring (93). While this has been clearly demonstrated in the oocyte, it is unknown whether such a purge might take place in somatic cells to maintain fitness of the mitochondrial genome. We recently showed that in cell culture, an episode of starvation results in significant depletion of mitochondria (8). Whether this process is followed by mitochondrial repopulation with the most functional mtDNA genomes and improved mitochondrial function remains to be determined. However, studies of fasting and refeeding can result in myocardial dysfunction (78); if mitochondrial purging occurs during the fasting episodes, it may be maladaptive. The removal of mitochondria usually needs to be compensated by mitochondrial biogenesis or it will become detrimental (11). For example, in the liver during subacute sepsis, mitochondrial function is impaired and mitochondria are removed by autophagy. In this model, clearance of mitochondria is followed by mitochondrial replenishment (16). Mitochondrial biogenesis in the presence of defective mitochondrial function is also observed in type 1 diabetes (92). In hearts of OVE26 mice, a chronic model of type 1 diabetes, there is upregulation of mitochondrial proteins, increase in mitochondrial area and number and mitochondrial DNA (92). Despite the higher number of mitochondria, their function is impaired. Similar results are observed in insulinresistant mice (22). It is possible that increased biogenesis is a compensatory mechanism for defective mitochondrial function. Biogenesis in the absence of balanced mitophagy to remove defective mitochondria may be maladaptive. Further studies are warranted to clarify this. In most situations, increased mitochondrial biogenesis is beneficial, such as in caloric restriction and aging, exercise, amyotrophic lateral sclerosis (ALS) treatment, neonatal cardiomyocytes response to LPS, and renal proximal tubular cells subjected to oxidative stress. In skeletal muscles, caloric restriction enhances mitochondrial protein turnover by enhancing mitochondrial degradation via autophagy and by stimulating mitochondrial biogenesis. Caloric restriction activates sirtuin 1 (SIRT1) which activates autophagy, via deacetylating autophagy proteins (57), and mitochondrial biogenesis, via activation of peroxisome proliferatoractivated receptor gamma cofactor-1alpha (PGC-1alpha) (12, 35, 62). The net result is that caloric restriction ensures good mitochondrial function with aging by promoting mitochondrial turnover. A similar mechanism is observed with exercise (60, 61, 68, 105). Skeletal muscle biopsies of humans performing high-intensity interval training showed an increase in SIRT1, nuclear PGC-1alpha and mitochondrial transcription factor A, which lead to an increase in skeletal muscle mitochondria and improved exercise performance (60, 61). Biopsies performed in older men show that even with aging, exercise enhances mitochondrial respiratory chain activity and mitochondrial DNA, which is likely related to increases in mitochondria biogenesis (68). ALS is characterized by the presence of intracellular aggregates in the motor neurons, mitochondrial dysfunction, and deficiency of autophagy (88). A clinical study demonstrated that lithium slows the progression of ALS in patients by activating autophagy, thereby leading to the removal of defective mitochondria and protein aggregates, and by stimulating mitochondrial biogenesis (25, 26). In brain and heart, hypoxic preconditioning was shown to stimulate mitochondrial biogenesis via increase of PGC-1 alpha, nuclear respiratory factor-1, and mitochondrial transcription factor A (34, 67). In neonatal rat cardiomyocytes the protective response to LPS consists of the rapid removal of damaged organelles by autophagy and replacement of dysfunctional mitochondria by mitochondrial biogenesis. Hickson-Bick et al. (41) reported that LPS treatment in neonatal cardiomyocytes increases the transcription of mitochondrial transcription factor A, nuclear accumulation of redox-sensitive nuclear respiratory factor 1, and expression of PGC-1. The dose of LPS used did not induce apoptosis or necrosis in these cells. However, when autophagy was blocked, ROS production was increased and apoptosis was induced, showing that autophagy plays a protective role in LPS-induced injury (109). In renal proximal tubular cells subjected to oxidative stress there is a decrease in mitochondrial function and an increase in mitophagy (83). In these cells, following injury, there is an increase in PGC-1alpha which induces mitochondrial biogenesis (83). The authors observed upregulation of PGC-1alpha from day 1 to 3 following injury, which was consistent with the initiation of the recovery of mitochondrial function. At day 4, PGC-1alpha levels returned to basal values concomitant with the restoration of mitochondrial function (83). This illustrates a negative feedback mechanism that prevents excessive mitochondrial formation once mitochondrial function is restored. It also represents a fine balance between mitophagy and mitochondrial biogenesis, where a decrease in mitochondria number would lead to energetic deficiency, but also an excess of mitochondria could be

281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327

deleterious, by increasing the formation of ROS and leading to oxidative stress. Autophagy and biogenesis work as intertwined mechanisms that assure mitochondrial quality and cellular homeostasis. What happens when mitochondrial turnover is inadequate? Proliferating cells can remove biological wastes by cell division; however cardiac myocytes cannot because they are terminally differentiated and maintain cellular homeostasis only by the activation of degrading pathways, namely, macroautophagy, microautophagy and chaperone-mediated autophagy (97). Thus, the accumulation of defective mitochondria and lysosomes in aged myocytes is a reflection of inefficient autophagy (99). Senescent myocytes are characterized by the existence of giant mitochondria originating from oxidative damage followed by inefficient mitochondrial DNA repair and autophagy (9, 14, 95, 98). Terman and Brunk suggested that autophagic engulfment of large mitochondria requires more energy, and consequently, is less efficient (96). In an ideal situation of mitochondrial turnover, mitochondrial recycling should provide for removal of damaged mitochondria and their replacement by normal, replicating mitochondria (6, 95). However, this is not always the case and the accumulation of defective mitochondria in post-mitotic cells is frequently observed (6). de Grey (17, 18) suggested that the lower respiratory rate of defective mitochondria would be accompanied by less oxidative damage than that caused by normal mitochondria, which might make them less prone to autophagy. Nekhaeva and collaborators (71) showed that in humans, mitochondrial fission can be increased by certain mutations in the mitochondrial DNA, resulting in the replacement of normal mitochondria by mutated ones which are less susceptible to oxidative damage and potentially less vulnerable to autophagy. With increasing age, lysosomal activity decreases and autophagy is inefficient due to the accumulation of lipofuscin, a brown granular pigment that consists of cross-linked lipids and proteins produced during lysosomal digestion (48, 97). In the aging heart, the gradual inhibition of autophagy is at least in part caused by the intralysosomal accumulation of lipofuscin (96). The impairment or suppression of autophagy plays a critical role in the development of aging-related disorders in the heart. Therapeutic approaches to increase autophagy and mitophagy may prove to be cardioprotective. Life in the balance, longevity the goal Self-eating, recycling, cash-for-your-clunkers: Trade up to the mitochondrial equivalent Prius. The road to rejuvenation is paved with destruction, For clearing the rubble precedes reconstruction. But remember that lifes circular dance Depends on opposite forces in balance: Excess destruction, too much biogenesis, Brings heart failure, cancer, or neurodegeneris.

Acknowledgements This work was supported by NIH R01 HL60590 and NIH R01 AG33283 (RAG), and a postdoctoral fellowship from the Portuguese Foundation for Science and Technology of Portugal (SFRH/BPD/40454/2007) (RSC).

328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383

References 1. Beau I, Esclatine A, and Codogno P. Lost to translation: when autophagy targets mature ribosomes. Trends Cell Biol 18: 311-314, 2008. 2. Bernardi P. Mitochondrial transport of cations: channels, exchangers, and permeability transition. Physiol Rev 79: 1127-1155, 1999. 3. Boya P, Gonzalez-Polo RA, Casares N, Perfettini JL, Dessen P, Larochette N, Metivier D, Meley D, Souquere S, Yoshimori T, Pierron G, Codogno P, and Kroemer G. Inhibition of macroautophagy triggers apoptosis. Mol Cell Biol 25: 1025-1040, 2005. 4. Brady NR, Hamacher-Brady A, Westerhoff HV, and Gottlieb RA. A wave of reactive oxygen species (ROS)-induced ROS release in a sea of excitable mitochondria. Antioxid Redox Signal 8: 1651-1665, 2006. 5. Brady NR, Hamacher-Brady A, Yuan H, and Gottlieb RA. The autophagic response to nutrient deprivation in the HL-1 cardiac myocyte is modulated by Bcl-2 and sarco/endoplasmic reticulum calcium stores. FEBS Journal 274: 3184-3197, 2007. 6. Brunk UT and Terman A. The mitochondrial-lysosomal axis theory of aging: accumulation of damaged mitochondria as a result of imperfect autophagocytosis. Eur J Biochem 269: 1996-2002, 2002. 7. Canto C, Jiang LQ, Deshmukh AS, Mataki C, Coste A, Lagouge M, Zierath JR, and Auwerx J. Interdependence of AMPK and SIRT1 for Metabolic Adaptation to Fasting and Exercise in Skeletal Muscle. Cell Metab 11: 213-219. 8. Carreira RS, Lee Y, Ghochani M, Gustafsson AB, and Gottlieb RA. Cyclophilin D is required for mitochondrial removal by autophagy in cardiac cells. Autophagy In press., 2010. 9. Chan DC. Mitochondria: dynamic organelles in disease, aging, and development. Cell 125: 1241-1252, 2006. 10. Chen Y, Azad MB, and Gibson SB. Superoxide is the major reactive oxygen species regulating autophagy. Cell Death Differ 16: 1040-1052, 2009. 11. Cherra SJ and Chu CT. Autophagy in neuroprotection and neurodegeneration: A question of balance. Future Neurol 3: 309-323, 2008. 12. Civitarese AE, Carling S, Heilbronn LK, Hulver MH, Ukropcova B, Deutsch WA, Smith SR, and Ravussin E. Calorie restriction increases muscle mitochondrial biogenesis in healthy humans. PLoS Med 4: e76, 2007. 13. Clark SL, Jr. Cellular differentiation in the kidneys of newborn mice studies with the electron microscope. J Biophys Biochem Cytol 3: 349-362, 1957. 14. Coleman R, Silbermann M, Gershon D, and Reznick AZ. Giant mitochondria in the myocardium of aging and endurance-trained mice. Gerontology 33: 34-39, 1987. 15. Crompton M. The mitochondrial permeability transition pore and its role in cell death. Biochem J 341 ( Pt 2): 233-249, 1999. 16. Crouser ED, Julian MW, Huff JE, Struck J, and Cook CH. Carbamoyl phosphate synthase-1: a marker of mitochondrial damage and depletion in the liver during sepsis. Crit Care Med 34: 2439-2446, 2006. 17. de Grey AD. A proposed refinement of the mitochondrial free radical theory of aging. Bioessays 19: 161-166, 1997. 18. de Grey AD. The reductive hotspot hypothesis of mammalian aging: membrane metabolism magnifies mutant mitochondrial mischief. Eur J Biochem 269: 2003-2009, 2002. 19. Decker RS and Wildenthal K. Lysosomal alterations in hypoxic and reoxygenated hearts. I. Ultrastructural and cytochemical changes. Am J Pathol 98: 425-444, 1980. 20. Deshmukh AS, Treebak JT, Long YC, Viollet B, Wojtaszewski JF, and Zierath JR. Role of adenosine 5'-monophosphate-activated protein kinase subunits in skeletal muscle mammalian target of rapamycin signaling. Mol Endocrinol 22: 1105-1112, 2008. 21. Diwan A, Koesters AG, Odley AM, Pushkaran S, Baines CP, Spike BT, Daria D, Jegga AG, Geiger H, Aronow BJ, Molkentin JD, Macleod KF, Kalfa TA, and Dorn GW, 2nd. Unrestrained erythroblast development in Nix-/- mice reveals a mechanism for apoptotic modulation of erythropoiesis. Proc Natl Acad Sci U S A 104: 6794-6799, 2007. 22. Duncan JG, Fong JL, Medeiros DM, Finck BN, and Kelly DP. Insulin-resistant heart exhibits a mitochondrial biogenic response driven by the peroxisome proliferator-activated receptor-alpha/PGC-1alpha gene regulatory pathway. Circulation 115: 909-917, 2007. 23. Elmore SP, Qian T, Grissom SF, and Lemasters JJ. The mitochondrial permeability transition initiates autophagy in rat hepatocytes. FASEB J 15: 2286-2287, 2001.

384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438

24. Erlich S, Mizrachy L, Segev O, Lindenboim L, Zmira O, Adi-Harel S, Hirsch JA, Stein R, and Pinkas-Kramarski R. Differential interactions between Beclin 1 and Bcl-2 family members. Autophagy 3: 561568, 2007. 25. Fornai F, Longone P, Cafaro L, Kastsiuchenka O, Ferrucci M, Manca ML, Lazzeri G, Spalloni A, Bellio N, Lenzi P, Modugno N, Siciliano G, Isidoro C, Murri L, Ruggieri S, and Paparelli A. Lithium delays progression of amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A 105: 2052-2057, 2008. 26. Fornai F, Longone P, Ferrucci M, Lenzi P, Isidoro C, Ruggieri S, and Paparelli A. Autophagy and amyotrophic lateral sclerosis: The multiple roles of lithium. Autophagy 4: 527-530, 2008. 27. Gao Z, Gammoh N, Wong PM, Erdjument-Bromage H, Tempst P, and Jiang X. Processing of autophagic protein LC3 by the 20S proteasome. Autophagy 6: 126-137. 28. Geisler S, Holmstrom KM, Skujat D, Fiesel FC, Rothfuss OC, Kahle PJ, and Springer W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat Cell Biol 12: 119-131. 29. Goffart S, von Kleist-Retzow JC, and Wiesner RJ. Regulation of mitochondrial proliferation in the heart: power-plant failure contributes to cardiac failure in hypertrophy. Cardiovasc Res 64: 198-207, 2004. 30. Grivennikova VG, Kareyeva AV, and Vinogradov AD. What are the sources of hydrogen peroxide production by heart mitochondria? Biochim Biophys Acta. 31. Gronowicz G, Swift H, and Steck TL. Maturation of the reticulocyte in vitro. J Cell Sci 71: 177-197, 1984. 32. Gustafsson AB and Gottlieb RA. Autophagy in ischemic heart disease. Circ Res 104: 150-158, 2009. 33. Gustafsson AB and Gottlieb RA. Heart mitochondria: gates of life and death. Cardiovasc Res 77: 334-343, 2008. 34. Gutsaeva DR, Carraway MS, Suliman HB, Demchenko IT, Shitara H, Yonekawa H, and Piantadosi CA. Transient hypoxia stimulates mitochondrial biogenesis in brain subcortex by a neuronal nitric oxide synthase-dependent mechanism. J Neurosci 28: 2015-2024, 2008. 35. Haigis MC and Guarente LP. Mammalian sirtuins--emerging roles in physiology, aging, and calorie restriction. Genes Dev 20: 2913-2921, 2006. 36. Halestrap AP and Pasdois P. The role of the mitochondrial permeability transition pore in heart disease. Biochim Biophys Acta 1787: 1402-1415, 2009. 37. Hamacher-Brady A, Brady NR, Gottlieb RA, and Gustafsson AB. Autophagy as a protective response to Bnip3-mediated apoptotic signaling in the heart. Autophagy 2: 307-309, 2006. 38. Hamacher-Brady A, Brady NR, Logue SE, Sayen MR, Jinno M, Kirshenbaum LA, Gottlieb RA, and Gustafsson AB. Response to myocardial ischemia/reperfusion injury involves Bnip3 and autophagy. Cell Death Differ 14: 146-157, 2007. 39. Hausenloy D, Wynne A, Duchen M, and Yellon D. Transient mitochondrial permeability transition pore opening mediates preconditioning-induced protection. Circulation 109: 1714-1717, 2004. 40. Heynen MJ, Tricot G, and Verwilghen RL. Autophagy of mitochondria in rat bone marrow erythroid cells. Relation to nuclear extrusion. Cell Tissue Res 239: 235-239, 1985. 41. Hickson-Bick DL, Jones C, and Buja LM. Stimulation of mitochondrial biogenesis and autophagy by lipopolysaccharide in the neonatal rat cardiomyocyte protects against programmed cell death. J Mol Cell Cardiol 44: 411-418, 2008. 42. Hyde BB, Twig G, and Shirihai OS. Organellar vs cellular control of mitochondrial dynamics. Semin Cell Dev Biol, 2010. 43. Ichas F and Mazat JP. From calcium signaling to cell death: two conformations for the mitochondrial permeability transition pore. Switching from low- to high-conductance state. Biochim Biophys Acta 1366: 33-50, 1998. 44. Ichimura Y, Kominami E, Tanaka K, and Komatsu M. Selective turnover of p62/A170/SQSTM1 by autophagy. Autophagy 4: 1063-1066, 2008. 45. Kang HT and Hwang ES. Nicotinamide enhances mitochondria quality through autophagy activation in human cells. Aging Cell 8: 426-438, 2009. 46. Kanki T and Klionsky DJ. Mitophagy in yeast occurs through a selective mechanism. J Biol Chem 283: 32386-32393, 2008. 47. Kanki T, Wang K, Cao Y, Baba M, and Klionsky DJ. Atg32 is a mitochondrial protein that confers selectivity during mitophagy. Dev Cell 17: 98-109, 2009. 48. Keller JN, Dimayuga E, Chen Q, Thorpe J, Gee J, and Ding Q. Autophagy, proteasomes, lipofuscin, and oxidative stress in the aging brain. Int J Biochem Cell Biol 36: 2376-2391, 2004.

439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494

49. Kissova I, Deffieu M, Manon S, and Camougrand N. Uth1p is involved in the autophagic degradation of mitochondria. J Biol Chem 279: 39068-39074, 2004. 50. Kissova I, Salin B, Schaeffer J, Bhatia S, Manon S, and Camougrand N. Selective and nonselective autophagic degradation of mitochondria in yeast. Autophagy 3: 329-336, 2007. 51. Korolchuk VI, Menzies FM, and Rubinsztein DC. Mechanisms of cross-talk between the ubiquitinproteasome and autophagy-lysosome systems. FEBS Lett, 2009. 52. Koury MJ, Koury ST, Kopsombut P, and Bondurant MC. In vitro maturation of nascent reticulocytes to erythrocytes. Blood 105: 2168-2174, 2005. 53. Kubasiak LA, Hernandez OM, Bishopric NH, and Webster KA. Hypoxia and acidosis activate cardiac myocyte death through the Bcl-2 family protein BNIP3. Proc Natl Acad Sci U S A 99: 12825-12830, 2002. 54. Kubli DA, Quinsay MN, Huang C, Lee Y, and Gustafsson AB. Bnip3 functions as a mitochondrial sensor of oxidative stress during myocardial ischemia and reperfusion. Am J Physiol Heart Circ Physiol 295: H2025-2031, 2008. 55. Kujoth GC, Bradshaw PC, Haroon S, and Prolla TA. The role of mitochondrial DNA mutations in mammalian aging. PLoS Genet 3: e24, 2007. 56. Kundu M, Lindsten T, Yang CY, Wu J, Zhao F, Zhang J, Selak MA, Ney PA, and Thompson CB. Ulk1 plays a critical role in the autophagic clearance of mitochondria and ribosomes during reticulocyte maturation. Blood 112: 1493-1502, 2008. 57. Lee IH, Cao L, Mostoslavsky R, Lombard DB, Liu J, Bruns NE, Tsokos M, Alt FW, and Finkel T. A role for the NAD-dependent deacetylase Sirt1 in the regulation of autophagy. Proc Natl Acad Sci U S A 105: 3374-3379, 2008. 58. Li W, Bengtson MH, Ulbrich A, Matsuda A, Reddy VA, Orth A, Chanda SK, Batalov S, and Joazeiro CA. Genome-wide and functional annotation of human E3 ubiquitin ligases identifies MULAN, a mitochondrial E3 that regulates the organelle's dynamics and signaling. PLoS One 3: e1487, 2008. 59. Lim SY, Davidson SM, Hausenloy DJ, and Yellon DM. Preconditioning and postconditioning: the essential role of the mitochondrial permeability transition pore. Cardiovasc Res 75: 530-535, 2007. 60. Little JP, Safdar A, Cermak NM, Tarnopolsky MA, and Gibala MJ. Acute endurance exercise increases the nuclear abundance of PGC-1{alpha} in trained human skeletal muscle. Am J Physiol Regul Integr Comp Physiol. 61. Little JP, Safdar AS, Wilkin GP, Tarnopolsky MA, and Gibala MJ. A practical model of low-volume high-intensity interval training induces mitochondrial biogenesis in human skeletal muscle: potential mechanisms. J Physiol. 62. Lopez-Lluch G, Hunt N, Jones B, Zhu M, Jamieson H, Hilmer S, Cascajo MV, Allard J, Ingram DK, Navas P, and de Cabo R. Calorie restriction induces mitochondrial biogenesis and bioenergetic efficiency. Proc Natl Acad Sci U S A 103: 1768-1773, 2006. 63. Maiuri MC, Le Toumelin G, Criollo A, Rain JC, Gautier F, Juin P, Tasdemir E, Pierron G, Troulinaki K, Tavernarakis N, Hickman JA, Geneste O, and Kroemer G. Functional and physical interaction between Bcl-X(L) and a BH3-like domain in Beclin-1. EMBO J 26: 2527-2539, 2007. 64. Mancuso DJ, Abendschein DR, Jenkins CM, Han X, Saffitz JE, Schuessler RB, and Gross RW. Cardiac ischemia activates calcium-independent phospholipase A2beta, precipitating ventricular tachyarrhythmias in transgenic mice: rescue of the lethal electrophysiologic phenotype by mechanism-based inhibition. J Biol Chem 278: 22231-22236, 2003. 65. Matsui M, Yamamoto A, Kuma A, Ohsumi Y, and Mizushima N. Organelle degradation during the lens and erythroid differentiation is independent of autophagy. Biochem Biophys Res Commun 339: 485-489, 2006. 66. Mazure NM and Pouyssegur J. Atypical BH3-domains of BNIP3 and BNIP3L lead to autophagy in hypoxia. Autophagy 5: 868-869, 2009. 67. McLeod CJ, Jeyabalan AP, Minners JO, Clevenger R, Hoyt RF, Jr., and Sack MN. Delayed ischemic preconditioning activates nuclear-encoded electron-transfer-chain gene expression in parallel with enhanced postanoxic mitochondrial respiratory recovery. Circulation 110: 534-539, 2004. 68. Menshikova EV, Ritov VB, Fairfull L, Ferrell RE, Kelley DE, and Goodpaster BH. Effects of exercise on mitochondrial content and function in aging human skeletal muscle. J Gerontol A Biol Sci Med Sci 61: 534-540, 2006. 69. Menzies RA and Gold PH. The turnover of mitochondria in a variety of tissues of young adult and aged rats. J Biol Chem 246: 2425-2429, 1971.

495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549

70. Narendra D, Tanaka A, Suen DF, and Youle RJ. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol 183: 795-803, 2008. 71. Nekhaeva E, Bodyak ND, Kraytsberg Y, McGrath SB, Van Orsouw NJ, Pluzhnikov A, Wei JY, Vijg J, and Khrapko K. Clonally expanded mtDNA point mutations are abundant in individual cells of human tissues. Proc Natl Acad Sci U S A 99: 5521-5526, 2002. 72. Nishida K, Kyoi S, Yamaguchi O, Sadoshima J, and Otsu K. The role of autophagy in the heart. Cell Death Differ 16: 31-38, 2009. 73. Novak I, Kirkin V, McEwan DG, Zhang J, Wild P, Rozenknop A, Rogov V, Lohr F, Popovic D, Occhipinti A, Reichert AS, Terzic J, Dotsch V, Ney PA, and Dikic I. Nix is a selective autophagy receptor for mitochondrial clearance. EMBO Rep 11: 45-51. 74. Nowikovsky K, Reipert S, Devenish RJ, and Schweyen RJ. Mdm38 protein depletion causes loss of mitochondrial K+/H+ exchange activity, osmotic swelling and mitophagy. Cell Death Differ 14: 1647-1656, 2007. 75. Okamoto K, Kondo-Okamoto N, and Ohsumi Y. Mitochondria-anchored receptor Atg32 mediates degradation of mitochondria via selective autophagy. Dev Cell 17: 87-97, 2009. 76. Pattingre S, Tassa A, Qu X, Garuti R, Liang XH, Mizushima N, Packer M, Schneider MD, and Levine B. Bcl-2 antiapoptotic proteins inhibit beclin 1-dependent autophagy. Cell 122: 927-939, 2005. 77. Pfeifer U. Inhibition by insulin of the formation of autophagic vacuoles in rat liver. A morphometric approach to the kinetics of intracellular degradation by autophagy. J Cell Biol 78: 152-167, 1978. 78. Pinotti MF, Leopoldo AS, Silva MD, Sugizaki MM, do Nascimento AF, Lima-Leopoldo AP, Aragon FF, Padovani CR, and Cicogna AC. A comparative study of myocardial function and morphology during fasting/refeeding and food restriction in rats. Cardiovasc Pathol, 2009. 79. Priault M, Salin B, Schaeffer J, Vallette FM, di Rago JP, and Martinou JC. Impairing the bioenergetic status and the biogenesis of mitochondria triggers mitophagy in yeast. Cell Death Differ 12: 16131621, 2005. 80. Quinsay MN, Lee Y, Rikka S, Sayen MR, Molkentin JD, Gottlieb RA, and Gustafsson AB. Bnip3 mediates permeabilization of mitochondria and release of cytochrome c via a novel mechanism. J Mol Cell Cardiol, 2009. 81. Rabinowitz M and Swift H. Mitochondrial nucleic acids and their relation to the biogenesis of mitochondria. Physiol Rev 50: 376-427, 1970. 82. Rabinowitz M and Zak R. Mitochondria and cardiac hypertrophy. Circ Res 36: 367-376, 1975. 83. Rasbach KA and Schnellmann RG. Signaling of mitochondrial biogenesis following oxidant injury. J Biol Chem 282: 2355-2362, 2007. 84. Rasola A and Bernardi P. The mitochondrial permeability transition pore and its involvement in cell death and in disease pathogenesis. Apoptosis 12: 815-833, 2007. 85. Ravikumar B, Acevedo-Arozena A, Imarisio S, Berger Z, Vacher C, O'Kane CJ, Brown SD, and Rubinsztein DC. Dynein mutations impair autophagic clearance of aggregate-prone proteins. Nat Genet 37: 771-776, 2005. 86. Regula KM, Ens K, and Kirshenbaum LA. Inducible expression of BNIP3 provokes mitochondrial defects and hypoxia-mediated cell death of ventricular myocytes. Circ Res 91: 226-231, 2002. 87. Rodriguez-Enriquez S, Kim I, Currin RT, and Lemasters JJ. Tracker dyes to probe mitochondrial autophagy (mitophagy) in rat hepatocytes. Autophagy 2: 39-46, 2006. 88. Ross CA and Poirier MA. Protein aggregation and neurodegenerative disease. Nat Med 10 Suppl: S10-17, 2004. 89. Sandoval H, Thiagarajan P, Dasgupta SK, Schumacher A, Prchal JT, Chen M, and Wang J. Essential role for Nix in autophagic maturation of erythroid cells. Nature 454: 232-235, 2008. 90. Scherz-Shouval R, Shvets E, Fass E, Shorer H, Gil L, and Elazar Z. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. Embo J 26: 1749-1760, 2007. 91. Schweers RL, Zhang J, Randall MS, Loyd MR, Li W, Dorsey FC, Kundu M, Opferman JT, Cleveland JL, Miller JL, and Ney PA. NIX is required for programmed mitochondrial clearance during reticulocyte maturation. Proc Natl Acad Sci U S A 104: 19500-19505, 2007. 92. Shen X, Zheng S, Thongboonkerd V, Xu M, Pierce WM, Jr., Klein JB, and Epstein PN. Cardiac mitochondrial damage and biogenesis in a chronic model of type 1 diabetes. Am J Physiol Endocrinol Metab 287: E896-905, 2004. 93. Stewart JB, Freyer C, Elson JL, Wredenberg A, Cansu Z, Trifunovic A, and Larsson NG. Strong purifying selection in transmission of mammalian mitochondrial DNA. PLoS Biol 6: e10, 2008.

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

94. Tal R, Winter G, Ecker N, Klionsky DJ, and Abeliovich H. Aup1p, a yeast mitochondrial protein phosphatase homolog, is required for efficient stationary phase mitophagy and cell survival. J Biol Chem 282: 5617-5624, 2007. 95. Terman A. Garbage catastrophe theory of aging: imperfect removal of oxidative damage? Redox Rep 6: 15-26, 2001. 96. Terman A and Brunk UT. Autophagy in cardiac myocyte homeostasis, aging, and pathology. Cardiovasc Res 68: 355-365, 2005. 97. Terman A and Brunk UT. Oxidative stress, accumulation of biological 'garbage', and aging. Antioxid Redox Signal 8: 197-204, 2006. 98. Terman A, Dalen H, Eaton JW, Neuzil J, and Brunk UT. Mitochondrial recycling and aging of cardiac myocytes: the role of autophagocytosis. Exp Gerontol 38: 863-876, 2003. 99. Terman A, Gustafsson B, and Brunk UT. The lysosomal-mitochondrial axis theory of postmitotic aging and cell death. Chem Biol Interact 163: 29-37, 2006. 100. Tolkovsky AM, Xue L, Fletcher GC, and Borutaite V. Mitochondrial disappearance from cells: a clue to the role of autophagy in programmed cell death and disease? Biochimie 84: 233-240, 2002. 101. Twig G, Elorza A, Molina AJ, Mohamed H, Wikstrom JD, Walzer G, Stiles L, Haigh SE, Katz S, Las G, Alroy J, Wu M, Py BF, Yuan J, Deeney JT, Corkey BE, and Shirihai OS. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J 27: 433-446, 2008. 102. Vande Velde C, Cizeau J, Dubik D, Alimonti J, Brown T, Israels S, Hakem R, and Greenberg AH. BNIP3 and genetic control of necrosis-like cell death through the mitochondrial permeability transition pore. Mol Cell Biol 20: 5454-5468, 2000. 103. Ventura-Clapier R, Garnier A, and Veksler V. Transcriptional control of mitochondrial biogenesis: the central role of PGC-1alpha. Cardiovasc Res 79: 208-217, 2008. 104. Vermulst M, Wanagat J, Kujoth GC, Bielas JH, Rabinovitch PS, Prolla TA, and Loeb LA. DNA deletions and clonal mutations drive premature aging in mitochondrial mutator mice. Nat Genet 40: 392-394, 2008. 105. Vina J, Gomez-Cabrera MC, Borras C, Froio T, Sanchis-Gomar F, Martinez-Bello VE, and Pallardo FV. Mitochondrial biogenesis in exercise and in ageing. Adv Drug Deliv Rev 61: 1369-1374, 2009. 106. Weiss JN, Korge P, Honda HM, and Ping P. Role of the mitochondrial permeability transition in myocardial disease. Circ Res 93: 292-301, 2003. 107. Xue L, Fletcher GC, and Tolkovsky AM. Autophagy is activated by apoptotic signalling in sympathetic neurons: an alternative mechanism of death execution. Mol Cell Neurosci 14: 180-198, 1999. 108. Xue L, Fletcher GC, and Tolkovsky AM. Mitochondria are selectively eliminated from eukaryotic cells after blockade of caspases during apoptosis. Curr Biol 11: 361-365, 2001. 109. Yuan H, Perry CN, Huang C, Iwai-Kanai E, Carreira RS, Glembotski CC, and Gottlieb RA. LPSinduced autophagy is mediated by oxidative signaling in cardiomyocytes and is associated with cytoprotection. Am J Physiol Heart Circ Physiol 296: H470-479, 2009. 110. Zhang J, Randall MS, Loyd MR, Dorsey FC, Kundu M, Cleveland JL, and Ney PA. Mitochondrial clearance is regulated by Atg7-dependent and -independent mechanisms during reticulocyte maturation. Blood 114: 157-164, 2009. 111. Zhang X, Zhou J, Fernandes AF, Sparrow JR, Pereira P, Taylor A, and Shang F. The proteasome: a target of oxidative damage in cultured human retina pigment epithelial cells. Invest Ophthalmol Vis Sci 49: 3622-3630, 2008. 112. Zorov DB, Filburn CR, Klotz LO, Zweier JL, and Sollott SJ. Reactive oxygen species (ROS)-induced ROS release: a new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J Exp Med 192: 1001-1014, 2000.

599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626

Figure Legends Fig. 1. Overview of autophagy. Cellular stresses, including reactive oxygen or reactive nitrogen species (ROS, RNS), or nutritional/energetic stress activate AMPK and inhibit mTOR to signal autophagy through Atg1. Once initiated, Beclin-1 and VPS34, a Class I PI3 kinase, trigger downstream events leading to activation of the first of two ubiquitin-like pathways. First, Atg12 is activated similarly to ubiquitin by Atg7, an E1-like enzyme, which then transfers it to Atg10, an E2-like enzyme, which then conjugates it to Lysine 130 of Atg5. The Atg5/Atg12 conjugate then complexes with a homodimer of Atg16, and the complex assembles on a membrane structure termed the phagophore. This event is a prerequisite for the second ubiquitin-like pathway, which involves the cleavage of the terminal Cys residue of Atg8 (LC3) by Atg4, a cysteine protease, which exposes a terminal glycine residue. Atg4 is redox-regulated and appears to upregulate autophagy in the face of oxidative stress. Atg4 in conjunction with Atg7 facilitates conjugation of LC3 onto phosphatidylethanolamine in the lipid bilayer of the membrane. The cup-shaped phagophore is recruited to engulf targets via adaptor proteins such as p62, which binds ubiquitinated protein aggregates and LC3. The growing ends of the phagophore eventually meet and fuse to enclose the target within the double-membrane structure of the autophagosome. At this point Atg12/5/16L are released, and the autophagosome fuses with a lysosome; fusion and degradation of contents requires acidification, which is mediated by the vacuolar proton ATPase (VPATPase); lysosomal enzymes such as cathepsins, and lipases have pH optima well below 6.0. The entire process of formation and destruction occurs on a time-scale of minutes. Fig. 2. Mitophagy. The murine atrial myocyte-derived cell line, HL-1, was transfected with mitochondriatargeted red fluorescent protein (mitoDsRed) and LC3 fused to green fluorescent protein (GFP-LC3). Cells were induced to undergo autophagy and imaged by deconvolution fluorescence microscopy. The raw, unprocessed image of a region of a cell is shown at top right. After deconvolution and thresholding, it can be seen that mitochondria are present within autophagosomes.

Das könnte Ihnen auch gefallen