Vous êtes sur la page 1sur 11

Mitochondrial Control of Cellular Life, Stress, and Death

Lorenzo Galluzzi, Oliver Kepp, Christina Trojel-Hansen and Guido Kroemer


Circ Res. 2012;111:1198-1207
doi: 10.1161/CIRCRESAHA.112.268946
Circulation Research is published by the American Heart Association, 7272 Greenville Avenue, Dallas, TX 75231
Copyright 2012 American Heart Association, Inc. All rights reserved.
Print ISSN: 0009-7330. Online ISSN: 1524-4571

The online version of this article, along with updated information and services, is located on the
World Wide Web at:
http://circres.ahajournals.org/content/111/9/1198

Permissions: Requests for permissions to reproduce figures, tables, or portions of articles originally published in
Circulation Research can be obtained via RightsLink, a service of the Copyright Clearance Center, not the
Editorial Office. Once the online version of the published article for which permission is being requested is located,
click Request Permissions in the middle column of the Web page under Services. Further information about this
process is available in thePermissions and Rights Question and Answer document.
Reprints: Information about reprints can be found online at:
http://www.lww.com/reprints
Subscriptions: Information about subscribing to Circulation Research is online at:
http://circres.ahajournals.org//subscriptions/

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Review
This Review is the first in an ongoing series on Mitochondria in Health and Disease, which includes the following articles:

Mitochondrial Control of Cellular Life, Stress, and Death


Mitochondria and Mitophagy: The Yin and Yang of Cell Death Control
Mitochondria as a Drug Target in Ischemic Heart Disease and Cardiomyopathy
Physiological Roles of the Permeability Transition Pore
Exercising Mitochondria for Cardioprotection
Mitochondria and Endothelial Function
Mitochondria and ROS Generation
Guido Kroemer & Lorenzo Galluzzi, Guest Editors

Mitochondrial Control of Cellular Life, Stress, and Death


Lorenzo Galluzzi, Oliver Kepp, Christina Trojel-Hansen, Guido Kroemer

Abstract: Since the discovery that mitochondrial membrane permeabilization represents a critical step in the
regulation of intrinsic apoptosis, mitochondria have been viewed as pluripotent organelles, controlling cell death
as well as several aspects of cell survival. Mitochondria constitute the most prominent source of ATP and are
implicated in multiple anabolic and catabolic circuitries. In addition, mitochondria coordinate cell-wide stress
responses, such as autophagy, and control nonapoptotic cell death routines, such as regulated necrosis. Thus,
mitochondria seem to regulate a continuum of cellular functions, spanning from physiological metabolism to
stress responses and death. The involvement of mitochondria in both vital and lethal processes is crucial for both
embryonic and postembryonic development, as well as for the maintenance of adult tissue homeostasis. In line with
this notion, primary mitochondrial defects or alterations in the signaling pathways that converge on or emanate
from mitochondria underpin a large number of human diseases, including premature aging, neurodegenerative
disorders, cardiovascular disorders, and cancer. Here, we provide an overview of the molecular mechanisms that
enable mitochondria to sustain cell survival, coordinate stress responses, and mediate cell death, linking these
pathwayswhenever relevantto cardiovascular health and disease. (Circ Res. 2012;00:1198-1207.)
Key Words: caspases cytochrome c inflammasome innate immunity reactive oxygen species

30 years later, these efforts culminated in Mitchells hypothesis of chemiosmosis, postulating that the energy required for
the mitochondrial synthesis of ATP would be stored in the
form of a proton electrochemical gradient across the mitochondrial inner membrane.3,4
Developmental instances of programmed cell death (PCD)
have been observed as early as in the 2nd century AD, when
Galen of Pergamum noted the transitory state of the fetal arterial duct (while actually ignoring that he was observing cells,
because the cell theory was formulated by Carl Vogt only in

itochondria began to attract attention from the scientific community at the end of the 19th century,
when they were known as bioblasts.1 Very soon, however,
the term bioblast was abandoned in favor of the current
etymology, attributable to the threadlike appearance of mitochondria throughout spermatogenesis (mitos=thread,
chondros=granule).2 Pioneering attempts to isolate mitochondria were undertaken in the early 1930s, eliciting an intense
wave of investigation aimed at characterizing the biochemical
reactions that underlie cellular respiration.3 Approximately

Original received March 8, 2012; revision received April 23, 2012; accepted April 23, 2012. In July 2012, the average time from submission to first
decision for all original research papers submitted to Circulation Research was 11.2 days.
From the INSERMU848, Villejuif, France (O.K., C.T-H., G.K.); Institut Gustave Roussy, Villejuif, France (L.G., O.K., C.T-H.); Universit Paris Sud, Le
Kremlin-Bictre, France (O.K., C.T-H.); Metabolomics Platform, Institut Gustave Roussy, Villejuif, France (G.K.); Centre de Recherche des Cordeliers,
Paris, France (G.K.); Ple de Biologie, Hpital Europen Georges Pompidou, AP-HP, Paris, France (G.K.); and Universit Paris Descartes, Sorbonne Paris
Cit, Paris, France (L.G., G.K.).
Correspondence to Guido Kroemer, INSERM, U848, Institut Gustave Roussy, Pavillon de Recherche 1, 39 rue Camille Desmoulins, F-94805 Villejuif,
France. E-mail kroemer@orange.fr
2012 American Heart Association, Inc.
Circulation Research is available at http://circres.ahajournals.org

DOI: 10.1161/CIRCRESAHA.112.268946

1198

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Galluzzi et al Mitochondria Regulate Cellular Life and Death 1199

Non-standard Abbreviations and Acronyms


IMS
MOMP
OM
PCD
PTPC
RIPK1
ROS

mitochondrial intermembrane space


mitochondrial outer membrane permeabilization
outer membrane
programmed cell death
permeability transition pore complex
receptor-interacting protein kinase-1
reactive oxygen species

1842). Nevertheless, the concept of PCD has been theorized


as late as in 1964, thanks to the milestone work of Sir Richard
Lockshin,5 and the implication of mitochondria in the control
of cellular demise has not emerged until the 1990s,69 when
the morphological manifestations of a particular form of
PCD, apoptosis, were known10,11 but the underlying molecular
mechanisms were largely ignored.12,13 Since then, great efforts
have been made toward a more refined understanding of the
molecular circuitries that bridge mitochondria and the regulation of stress responses at the cell-wide level.14 This crosstalk may culminate in the re-establishment of homeostasis or,
when stress is excessive and the associated damage is beyond
recovery, in the activation of several mechanisms of cellular
demise, including apoptosis and regulated necrosis.12,15
In essence, mitochondria occupy a key position in the management of several cellular functions, including physiological metabolism, stress responses, and cell death. Intriguingly,
several mitochondrial components that operate as cell death
regulators also exert cell deathunrelated functions.16,17 This
might reflect the evolutionary origin of mitochondria (which,
according to the endosymbiotic hypothesis, have evolved
from proteobacteria engulfed by protoeukaroytic cells)18 or
their need to centralize a wide array of molecular functions
within a relatively small number of proteins.17
Here, we provide an overview of the metabolic circuitries
whereby mitochondria sustain the survival of eukaryotic cells,
the signaling pathways by which mitochondria coordinate the
cell response to stress, as well as the molecular mechanisms
that bridge mitochondrial functions to the execution of PCD,
focusing, whenever appropriate, on the relevance of such
pathways for cardiovascular health and disease.

Mitochondrial Role in Metabolism


The major contribution of mitochondria to the cellular metabolism is based on their capacity to generate ATP in a highly efficient, O2-dependent fashion, starting from the glycolytic product
pyruvate.4,19,20 Thus, in the presence of oxygen, pyruvate (rather
than being converted into lactate and secreted, as it occurs during anaerobic glycolysis) is imported into the mitochondrial
matrix and feeds the Krebs cycle, which is also known as tricarboxylic acid or citric acid cycle. The Krebs cycle not only
generates energized nucleotides (de facto slightly increasing the
bioenergetic efficiency of anaerobic glycolysis) but also drives
the accumulation of reduced NADH into the mitochondrial matrix.19 Through a concerted series of redox reactions catalyzed
by 4 multisubunit enzymes embedded in the mitochondrial inner membrane (ie, respiratory complexes IIV) and 2 diffusible factors (ie, cytochrome c, coenzyme Q10) that function as

electron shuttles within the mitochondrial intermembrane space


(IMS), NADH is oxidized by O2 concomitant with the extrusion
of protons into the IMS. This biochemical cascade is known
as oxidative phosphorylation and generates the electrochemical
gradient that is used by the F1FO-ATP synthase to generate the
bulk of intracellular ATP stores.4,20
However, the metabolic functions of mitochondria are
not limited to those ensured by the Krebs cycle and oxidative
phosphorylation. The compartmentalized ultrastructure of
mitochondria provides an optimal microenvironment for
many other biosynthetic and catabolic pathways. The first is
oxidation. Once fatty acids are imported into mitochondria by
the fatty acyl-coenzyme A synthetase and the carnitine shuttle, 4
distinct enzymatic activities ensure their degradation into acetylcoenzyme A units, which feed the Krebs cycle.21 Importantly,
oxidation constitutes the most prominent source of energy in
the adult mammalian heart, whereas glycolysis predominates
in the fetal heart as well as in pathological conditions, such as
hypertension, hypertrophy, and ischemia.22 The second is heme
biosynthesis. The Krebs cycle provides succinylcoenzyme A for
the generation of D-aminolevulinic acid (the precursor of heme),
and the 3 final steps of the cascade are catalyzed by mitochondrial
enzymes.23 The third is steroidogenesis. Mitochondria host part
of the mevalonate pathway, as well as 3 enzymes that participate
in the conversion of cholesterol into mineralcorticoids,
androgens, and estrogens.24 The fourth is amino acid metabolism.
On deamination by the urea cycle (also known as KrebsHenseleit cycle, which ensures ammonium detoxification),
amino acids can enter the Krebs cycle, either directly or at the
level of pyruvate.25 The fifth is the assembly of Fe/S clusters.
Mitochondrial enzymes and scaffold proteins catalyze the
generation of Fe/S clusters, which are then exported into the
cytosol and incorporated into acceptor proteins.26 The sixth is
gluconeogenesis. Pyruvate carboxylase, the enzyme that converts
pyruvate into oxaloacetate, is a mitochondrial enzyme, whereas
phosphoenolpyruvic acid carboxykinase, which catalyzes the
conversion of oxaloacetate into phosphoenolpyruvate, is evenly
distributed between mitochondria and the cytosol.27,28 The seventh
is ketogenesis. In liver cells, when the rates of oxidation are
high or when the activity of the Krebs cycle is reduced (because
of the shortage of some intermediates), acetyl-coenzyme A
is transformed in ketone bodies (ie, acetoacetate, acetone, and
hydroxybutyrate) entirely within mitochondria.27 Of note,
in ketotic conditions (ie, when the circulating levels of ketone
bodies are high, such as during fasting), acetoacetate, acetone,
and hydroxybutyrate can be used efficiently by the heart for
energy production.29 The eighth is Ca2+ storage. Together with
the endoplasmic reticulum, mitochondria constitute the major
storage compartment for intracellular Ca2+ ions. Ca2+ fluxes not
only regulate the bioenergetic and anabolic metabolism but also
impact on the activation of cell death pathways.30,31
Altogether, these examples underscore the notion that
mitochondria are critical for multiple facets of the metabolism
of the cell (Figure 1).

Mitochondrial Control of Stress Responses


In adverse conditions, cells activate a wide array of mechanisms to re-establish homeostasis and to repair stress-induced molecular damage.14 Mitochondria are involved in the

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

1200Circulation ResearchOctober 12, 2012

Figure 1. Bioenergetic functions of mitochondria. Mitochondria host, as a whole or in part, critical bioenergetic circuitries, including
the tricarboxylic acid (TCA) cycle, oxidative phosphorylation, the urea cycle, oxidation, and ketogenesis, as shown in this simplified
diagram. Please note that several intermediate reactions and all transporters have been purposely omitted. In the healthy adult heart,
oxidation constitutes the most prominent source of acetyl-CoA molecules for the generation of ATP via the TCA cycle and oxidative
phosphorylation. Conversely, glycolysis predominates in the fetal heart as well as during pathological states, such as hypertension,
hypertrophy, and ischemia. Of note, in ketotic conditions (ie, when the circulating levels of acetoacetate, acetone, or hydroxybutyrate
are high), the heart can efficiently use such liver-derived ketone bodies, rather than fatty acids, for energy production. -KGDH indicates
-ketoglutarate dehydrogenase; ACDH, acyl-CoA dehydrogenase; ACO, aconitase; ALD, aldolase; ASL, argininosuccinate lyase; ASS,
argininosuccinate synthetase; -HDBDH, -hydroxybutyrate dehydrogenase; CoA, coenzyme A; Cyt c, cytochrome c; ECH, enoylCoA hydratase; ENO, enolase; G3PDH, glyceraldehyde-3-phosphate dehydrogenase; HCDH, 3-hydroxyacyl-CoA dehydrogenase; HK,
hexokinase; HMG-CoA, 3-hydroxy-3-methylglutaryl-CoA; HMGCL, HMG-CoA lyase; HMGCS, HMG-CoA synthase; IDH, isocitrate
dehydrogenase; IMS, mitochondrial intermembrane space; MDH, malate dehydrogenase; OTC, ornithine transcarbamylase; PDH,
pyruvate dehydrogenase; PFK1, phosphofructokinase-1; PGI, phosphoglucose isomerase; PGK, phosphoglycerate kinase; PGM,
phosphoglycerate mutase; PK, pyruvate kinase; SDH, succinate dehydrogenase; Q10, Q10 coenzyme. (Illustration credit: Ben Smith.)

response of cells to a wide number of perturbations, including


oxidative stress and pathogen invasion. In addition, mitochondria play a critical role during autophagy in that they both can
be the substrate of autophagic degradation (mitophagy)32 or
can be specifically spared by the autophagic machinery to ensure energy production during stress responses.33,34
In the majority of eukaryotic cells, the mitochondrial
respiratory chain, notably at the level of complex I and complex
III, constitutes the most prominent source of intracellular
reactive oxygen species (ROS), even in baseline conditions.35
Thus, ROS constitute a normal side product of respiration, they
are required for several intracellular signaling pathways,36 and
are normally handled by an entire battery of mitochondrial,
cytosolic, and peroxisomal antioxidant systems, including
the following: superoxide dismutases, which catalyze the

dismutation of superoxide anions (O2) into O2 and hydrogen


peroxide (H2O2); catalases, which catalyze the decomposition
of H2O2 into H2O and O2; peroxiredoxins, which reduce H2O2 as
well as other oxidized species such as organic hydroperoxides
and peroxynitrites; thioredoxin, acting as an efficient reducing
agent, thanks to a highly conserved CXXC motif that can be
restored by thioredoxin reductase; glutathione, a tripeptide
with intrinsic antioxidant properties that can be recycled on
reduction by glutathione reductase; glutathione peroxidase,
an enzyme that uses the reducing equivalents of glutathione
for converting H2O2 into H2O and organic hydroperoxides
into the corresponding alcohols; and, simply, thiol-containing
proteins, which are endowed with the capacity to buffer ROS
by forming oxidized disulfide bonds.37 Extramitochondrial
sources of ROS include xanthine oxidase, which catalyzes

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Galluzzi et al Mitochondria Regulate Cellular Life and Death 1201


the oxidation of hypoxanthine to xanthine (1 step in the
catabolism of purines),38 multicomponent nicotinamide
adenine dinucleotide phosphate (NADPH) oxidases, such as
NOX1,39 as well as enzymes of the cytochrome P450 family.40
In the cardiovascular system, extramitochondrial sources
of ROS account for a consistent proportion (if not for the
majority) of the intracellular ROS content.41
As a consequence of oxidative stress, be it endogenous (ie, attributable to self-generated ROS) or exogenous (ie, attributable
to extracellular ROS), mitochondria undergo several functional
adaptations. For instance, in response to ROS, mitochondrial
thiols become rapidly oxidized, a modification that influences
the activity of various mitochondrial proteins like the adenine
nucleotide translocase (ANT).42,43 ANT normally mediates the
exchange of ADP and ATP between the cytosol and the mitochondrial matrix.44 This enzymatic activity gets impaired on
oxidization, resulting in intramitochondrial ADP shortage and
inhibited F1FO-ATP synthase.45 Furthermore, several proteomic
changes are rapidly triggered in mitochondria by oxidative
stress, including the upregulation of (part of) the antioxidant
systems listed above as well as of several molecular chaperons
that counteract ROS-induced protein unfolding.46,47 Altogether,
these adaptations maximize the ROS-buffering capacity of mitochondria (and of other cellular compartments), thus allowing
the re-establishment of homeostasis in response to mild oxidative stress. However, if oxidative stress persists and the associated molecular damage (ie, lipid peroxidation, protein misfolding,
mitochondrial DNA mutations) is beyond recovery, then mitochondria can translate such an adaptive response into the
activation of cell death. Several stimuli can trigger ROS overgeneration in cells from the cardiovascular system, including
hyperglycemia,48 hypoxia,49 and the presence of inflammatory
cytokines, such as tumor necrosis factor-.50 Importantly, such
a deregulated ROS production contributes to the pathogenesis
of a plethora of cardiovascular disorders, even in the absence of
massive cell death induction. The diseases include, but are not
limited to, congestive heart failure, left ventricular hypertrophy,
coronary heart disease, and cardiac arrhythmias.51
In response to infection by viruses, intracellular bacteria, and
some eukaryotic parasites, cells activate a series of mechanisms
that aim at clearing the invading pathogen and alerting neighboring uninfected cells. These defense strategies belong to innate immunity, because they are not based on the recognition of
specific antigens by the immune system, but rather of pathogenassociated molecular patterns, such as double-stranded RNA or
lipopolysaccharide.52 Pathogen-associated molecular patterns
are normally detected by so-called pattern recognition receptors, including plasma membrane-embedded and intracellular
members of the Toll-like receptor and NOD superfamilies,53,54
as well as melanoma differentiation-associated 5 (MDA-5) and
retinoic acid-inducible gene 1 (RIG-I), 2 cooperating intracellular sensors of viral RNA.55 Most often, activated pattern recognition receptors signal via various intermediates, including
myeloid differentiation primary response gene 88 (MYD88), to
transcription factors such as interferon regulatory factor (IRF)-3,
and nuclear factor-B (NF-B), underlying the synthesis of
high amounts of antiviral interferons, notably interferon .56,57
Mitochondria play a multifaceted and critical role in the
elicitation of innate immune responses.58 First, mitochondria

provide a scaffold for RIG-I signaling to NF-B via the mitochondrial antiviral signaling protein (MAVS),59 translocase
of outer mitochondrial membrane 70 (TOM70) (a translocase
of the mitochondrial outer membrane [OM]) appears to be required for optimal IRF-3 activation by MAVS,60 and mitofusin
2 (a mitochondrial protein regulating fusion) is a direct MAVS
inhibitor.61 In line with these notions, the knockdown of MAVS
and TOM70 by RNA interference abolished the Sendai virus
induced activation of NF-B and IRF-3 in human embryonic
kidney (HEK)-293T cells and allowed for viral replication.59,60
Second, several other components of the innate immune system either translocate to mitochondria on activation (such as
TNFR-associated factor 6 [TRAF-6], during the macrophagic
response to engulfed bacteria, and immunity-related GTPase
family, M [IRGM], which drives a mitochondrial fissiondependent autophagic pathway for the clearance of invading
bacteria)62,63 or are localized at mitochondria in physiological
conditions, such as the pattern recognition receptor NLR family
member X1 (NLRX1).64 Third, mitochondrial DNA has been
recently shown to be required for the optimal activation of the
NLR family, pyrin domain containing 3 (NLRP3) inflammasome (a supramolecular platform catalyzing the maturation of
proinflammatory interleukins, such as interleukin-1 and interleukin-18) in response to lipopolysaccharide.65,66 Fourth, mitochondrial alarmins, including mitochondrial DNA and formyl
peptides, constitute prominent damage-associated molecular
patterns and, hence, can activate strong inflammatory reactions once released into the extracellular space as a result of
cell death.67 Interestingly, IMS proteins that, once released in
the cytosol, trigger the execution phase of apoptosis (eg, cytochrome c) might be viewed as intracellular alarmins and, hence,
as part of a cell-intrinsic innate mechanism that ensures the
elimination of cells with damaged mitochondria.68 Intriguingly,
mice lacking NLRP3 are more susceptible to infection by endocarditis-associated bacteria, such as group B Streptococci.69
Furthermore, the inflammasome appears to be required for
maintaining intestinal bacteria, including potentially cardiotoxic Enterococci, under control.70 Nevertheless, a direct link
between mitochondria as regulators of innate immunity and
cardiovascular diseases has not yet been unveiled.
Macroautophagy (often and hereafter referred to as
autophagy) is a catabolic pathway leading to the lysosomal
degradation of cytoplasmic structures (including organelles,
portions of the cytoplasm, and invading pathogens).71,72 Baseline
levels of autophagy ensure the turnover of old, ectopic, and
damaged intracellular components.73 In addition, autophagy
gets upregulated as a cytoprotective mechanism in response to a
wide array of stressful conditions, including nutrient deprivation,
hypoxia, and the presence of toxic xenobiotics, such as anticancer
agents.14 Autophagy is involved in several physiological processes,
including development and differentiation.74 Furthermore,
defects in the autophagic machinery play a pathological role
in various settings, such as aging, neurodegenerative disorders,
and cancer.73,75,76 The relationship between mitochondria and the
stress response centered around autophagy is dual (Figure 2).
First, mitochondria can be the selective target of autophagy
(which in this case is called mitophagy).32 Mitophagy is particularly relevant for the disposal of damaged mitochondria,
which are potentially dangerous for the cells in that they often

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

1202Circulation ResearchOctober 12, 2012


hypoxia-induced mitophagy as mediated by the AMP-activated
protein kinase (AMPK),8284 although the underlying molecular
mechanisms have not yet been entirely elucidated. In conditions of nutrient deprivation, a pool of mitochondria elongates
is spared from degradation and sustains cell viability during
the autophagic response.33,34 Finally, it recently has been shown
that 0 yeast cells (ie, yeast cells lacking mitochondrial DNA
because of prolonged culture in ethidium bromide) exhibit defects in both the autophagic flux and the induction of autophagic genes,85 further strengthening the notion that mitochondria
and autophagy are intimately interconnected.

Mitochondrial Regulation of Cell Death

Figure 2. Dual cross-talk between autophagy and


mitochondria. On one hand, damaged or partially permeabilized
mitochondria are selectively degraded by the autophagic
machinery, a physiological mechanism that is critical for the
maintenance of intracellular homeostasis. The removal of damaged
(and, hence, potentially harmful) mitochondria by mitophagy has
recently been shown to be critical for the cardioprotective effects
of ischemic preconditioning as well as for the avoidance for chronic
cardiac disorders. On the other hand, mitochondria participate
in stress-induced autophagy (at least) by 3 mechanisms: (1) by
contributing lipids and proteins to forming phagophores; (2) by
providing proautophagic reactive oxygen species (ROS); and (3) by
elongating, avoiding autophagic degradation, and preserving ATP
production during the stress response. DRP1 indicates dynaminrelated protein 1. (Illustration credit: Ben Smith.)

overproduce ROS (in the absence of ATP synthesis) and are


prone to become fully permeabilized (hence leading to the
release of cell death inducers).13,75 In line with this notion,
Pink1/ mice (lacking a mitochondrial serine/threonine kinase involved in mitophagy) develop left ventricular dysfunction and evidence of pathological cardiac hypertrophy early
in life.77 Furthermore, mitophagy recently has been shown to
underlie the cardioprotective effects of ischemic preconditioning.78,79 Taken together, these observations suggest that the
clearance of damaged (and hence potentially harmful) mitochondria by mitophagy is critical for the avoidance of both
chronic and acute cardiovascular disorders.
Second, mitochondria actively contribute to the autophagic
response by multiple mechanisms. Mitochondrial membranes
(together with the endoplasmic reticulum) are believed to be
the source of the phagophore, ie, the first organelle that forms
to mediate autophagic sequestration.80 Mitochondrial ROS
appear to be required for some forms of autophagy,81 notably

According to currently accepted models, PCD can be executed by multiple signaling pathways, differing from each
other not only in their molecular constituents but also in their
morphological manifestations, biochemical features, and
immunological consequences at the organismal level.1012
The best-characterized among these pathways are intrinsic
apoptosis, extrinsic apoptosis, and regulated necrosis. The
mitochondrial control of extrinsic apoptosis is limited, ie, mitochondria provide an amplificatory loop for caspase activation but are not strictly required for the execution of cell death.
Conversely, mitochondrial permeabilization is indispensable
for intrinsic apoptosis and for many instances of regulated necrosis to proceed to completion (Figure 3).12,13,15
Extrinsic apoptosis is ignited by either the ligand-induced
oligomerization of transmembrane proteins of the death receptor
superfamily (eg, FAS/CD95, tumor necrosis factor receptor
1)86 or when the extracellular concentrations of specific trophic
factors (eg, netrin-1) fall below a certain threshold, leading to
the transduction of lethal signals by the so-called dependence
receptors (eg, unc-5 homolog A [C. elegans] [UNC5A], deleted in
colon cancer [DCC]).87 On oligomerization, the cytoplasmic tails
of death receptors drive the assembly of a multiprotein complex
known as death-inducing signaling complex, eventually leading
to the activation of the proapoptotic caspase-8caspase-3
cascade.88 In some cell types such as lymphocytes, known as type
I cells, the caspase cascade executes cell death in the absence
of primary mitochondrial involvement.89,90 On the contrary, in
so-called type II cells (including hepatocytes), active caspase-8
catalyzes the proteolytic cleavage of the B-cell lymphoma-2
(BCL-2) family member BID, leading to the generation of a
mitochondrion-permeabilizing peptide known as truncated
BID.91,92 In line with this notion, FAS-induced hepatocellular
apoptosis is inhibited in Bid/ mice,93 demonstrating that
mitochondrial OM permeabilization (MOMP), besides
constituting a point of no return during intrinsic apoptosis,
provides a contribution to some instances of extrinsic apoptosis
(ie, as triggered by death receptors in type II cells) that cannot be
ensured by the sole caspase-8caspase-3 cascade. Conversely,
so far no obvious links between dependence receptorignited
extrinsic apoptosis, which appears to be transduced by the
adaptor protein DRAL and caspase-9,94 and mitochondria have
been described.
Intrinsic apoptosis can be initiated by a wide variety of
intracellular perturbations (eg, DNA damage, oxidative stress,
cytosolic Ca2+ overload, endoplasmic reticulum stress).13
Despite the heterogeneous signaling pathways that sense such

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Galluzzi et al Mitochondria Regulate Cellular Life and Death 1203


Figure 3. Mitochondrial regulation of cell death.
Mitochondria have been shown to control (at least) 3
distinct cell death modalities: (A) extrinsic apoptosis;
(B) intrinsic apoptosis; and (C) regulated necrosis.
A, Extrinsic apoptosis can be triggered by the
ligation of death receptors, such as the tumor
necrosis factor- (TNF) receptor-1 (TNFR1). This
leads to the assembly of an intracellular multiprotein
complex, including (among other components,
which we purposely omitted for the sake of
simplicity) pro-caspase-8, the receptor-interacting
protein kinase 1 (RIPK1), and the adaptor protein
FAS-associated with a death domain (FADD).
Within such a complex, caspase-8 undergoes
spatial proximity-induced autoactivation and,
hence, becomes able to cleave several substrates,
including RIPK1, its homolog RIPK3, and caspase-3,
which coordinates the execution of cell death.
B, Intrinsic apoptosis involves a central step
of regulation that is mediated by mitochondria.
Thus, in response to multiple distinct intracellular
stress conditions, mitochondrial membranes can
become permeabilized because of the pore-forming
activity of proapoptotic members of the B-cell
lymphoma-2 (Bcl-2) protein family, such as BAX
and BAK. Alternatively, mitochondria can lose their
structural integrity after the so-called mitochondrial
permeability transition (MPT), a phenomenon that
is initiated at the mitochondrial inner membrane.
In both cases, permeabilized mitochondria allow
for the release of potentially toxic proteins into
the cytoplasm, including cytochrome c (Cyt c).
Together with dATP and the adaptor protein
apoptotic peptidase-associated factor 1 (APAF-1),
cytoplasmic Cyt c drives the assembly of the socalled apoptosome, a supramolecular complex for
the activation of pro-caspase-9. Eventually, active
caspase-9 proteolytically processes pro-caspase-3,
resulting in the execution of apoptotic cell death.
In some (but not all) cell types, extrinsic and
intrinsic apoptosis are interconnected by the BCL-2
homology 3 (BH3)-only protein BID, which can be
converted by caspase-8 into a mitochondrionpermeabilizing fragment. C, One particular instance
of regulated necrosis, also known as necroptosis, is triggered by TNFR1 ligation when caspases (in particular caspase-8) are inactive (for
instance because of the presence of the caspase inhibitor Z-VAD-fmk). In these conditions, RIPK1 and RIPK3 engage in physical and
functional interactions with mixed lineage kinase domain-like (MLKL) to form a multiprotein complex called necrosome. The necrosome
stimulates regulated necrosis at the mitochondrial level by inhibiting adenine nucleotide transferase (ANT), by exacerbating glutaminolysis
(and, hence, inducing the overgeneration of reactive oxygen species [ROS]), and by promoting mitochondrial fragmentation. Casp indicates
caspase; DRP1, dynamin-related protein 1; GLUD1, glutamate dehydrogenase 1; MOMP, mitochondrial outer membrane permeabilization;
PGAM5, phosphoglycerate mutase family member 5; tBID, truncated BID. (Illustration credit: Ben Smith.)

perturbations (whose detailed description goes beyond the


scope of this review), they are all connected to a mitochondrionorchestrated control mechanism.95 Thus, when proapoptotic
signals prevail, the OM loses its integrity, resulting in the
immediate cessation of mitochondrial ATP synthesis as well
as in the release of several potentially toxic IMS proteins into
the cytosol. MOMP can originate at the OM because of the
pore-forming activity of proapoptotic proteins of the BCL-2
family, such as BAX and BAK.96 Alternatively, MOMP can be
secondary to an abrupt increase in the permeability to small
solutes of the mitochondrial inner membrane, a phenomenon
that is known as mitochondrial permeability transition.97 In
the past, mitochondrial permeability transition was believed
to derive from the opening of the so-called permeability
transition pore complex, a supramolecular channel assembled
at the junctions between the mitochondrial inner membrane
and the OM.97 However, so far gene knockout experiments
in mice failed to attribute a critical role in the regulation of

intrinsic apoptosis to putative components of the permeability


transition pore complex,98100 casting doubts on the actual
pathophysiological relevance of the permeability transition
pore complex. One notable exception is constituted by
cyclophilin D, whose absence has been demonstrated to limit
(necrotic, not apoptotic) cell death in multiple in vivo settings,
encompassing cardiac ischemia/reperfusion injury.101,102
Irrespective of the initiating pathway, MOMP results in
the activation of caspase-dependent and caspase-independent
mechanisms for the execution of cell death, the former including the well-known cytochrome ctriggered, APAF-1dependent and dATP-dependent assembly of the apoptosome,
a supramolecular platform for the activation of the caspase9caspase-3 cascade.103,104 A precise description of the molecular determinants whereby MOMP engages cell death
executioner mechanisms is beyond the scope of the present review and can be found elsewhere.13 Still, it should be noted that
all instances of intrinsic apoptosis rely on MOMP and, hence,

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

1204Circulation ResearchOctober 12, 2012


are strictly controlled by mitochondria,12 and the execution of
intrinsic apoptosis often ensues an unsuccessful attempt of
cells to reestablish homeostasis via cell-wide stress responses
that are often orchestrated around mitochondria.14
Although suspicions of the existence of regulated necrosis
have permeated cell death research since the late 1990s,105,106
it was only in 2005 that this notion had gained full consensus, thanks to the discovery of a chemical inhibitor of necrosis with potential therapeutic applications, ie, necrostatin-1.107
Necrostatin-1, which specifically inhibits the enzymatic activity receptor-interacting protein kinase 1 (RIPK1),108 has
constituted an invaluable tool for the elucidation of the signaling pathways that underlie a specific instance of regulated
necrosis, ie, necroptosis.15,109 During the past few years, it has
become clear that necroptosis (which was previously known
to follow tumor necrosis factor receptor-1 ligation in the presence of chemical or genetic inhibitors of caspases)105 almost
always relies on the activation of RIPK1 and the functional and
physical interaction of RIPK1 with its homolog RIPK3.110112
Until very recently, however, the signaling pathways engaged
by the supramolecular complex containing RIPK1 and RIPK3
(the so-called necrosome) to execute regulated necrosis were
largely unknown, as were the molecular cascades leading nonnecroptotic instances of regulated necrosis,113 although several
hints pointed to the involvement of mitochondria.15 In particular, it had been observed that regulated necrosis: requires ROS
(of both mitochondrial and extramitochondrial origin)39,114116
and entails the accumulation of oxidized intermediates (such
as membrane-damaging lipid hydroperoxides),117 as well as
of redox-active compounds (including advanced glycation
end products);118 can proceed along with the inhibition of ANT;
can involve the overactivation of bioenergetic pathways, such
as glutaminolysis (which is partially catalyzed within mitochondria);112,119 and can be mediated by a (perhaps calpain-dependent) signaling cascade that, in response to DNA damage,
drives the release of apoptosis-inducing factor from the IMS, in
turn resulting in the overactivation of the DNA damagerelated
enzyme poly(ADP-ribose) polymerase (PARP).120122 Very recently, Wang et al provided further insights into the proximal
signaling cascade that is elicited by the necrosome123 and demonstrated that the mitochondrial phosphatase phosphoglycerate mutase family member 5 (PGAM5) is also is implicated
in multiple distinct pathways of regulated necrosis, because
it recruits and activates the mitochondrial fragmenting factor
dynamin-related protein 1.124 These observations lend further
support to the notion that mitochondria exert a key role in the
control of regulated necrosis.
Until recently, apoptosis was considered as an important
component in the pathogenesis of myocardial infarction and
heart failure,125 although it was known that the majority of
cardiomyocytes respond to ischemic conditions by undergoing necrosis. Presumably, this was attributable to the fact
that necrosis was believed to be uncontrollable and, hence,
virtually useless from a therapeutic point of view. Along with
the recognition of the wide pathophysiological relevance of
regulated necrosis, the role of apoptosis in the pathogenesis
of myocardial infarction and heart failure is being reconsidered.126 Murine models of ischemia/reperfusion demonstrate
that the pharmacological or genetic inhibition of proteins

involved in regulated necrosis, including cyclophilin,101,102


RIPK1,127 and PARP,128 provides consistent cardioprotective
effects. Such a conceptual shift has been paralleled by the discovery of human cardiac stem cells,129,130 de facto revolutionizing the conception of the heart as a merely postmitotic organ
and paving new ways for the development of cardioregenerative strategies.

Conclusion
As overviewed in this article, throughout evolution
mitochondria have changed their status from (symbiotic)
parasites to essential regulators of metabolism, stress
responses, and cell death. Intriguingly, multiple mitochondrial
constituents (encompassing respiratory chain proteins and
interactors, such as cytochrome c and apoptosis-inducing
factor (AIF), putative components of the permeability
transition pore complex like adenine nucleotide translocase,
as well as several other proteins of the IMS) exert multiple
independent activities and de facto regulate a continuum of
physiological (metabolic) and pathological (homeostatic or
lethal) processes.16,17 Thus, with a relatively small number of
proteins, mitochondria operate as highly specialized cellular
sentinels that continuously monitor the conditions of the cell
to finely adapt its metabolic activity, to coordinate adaptive
responses to stress, and to decide on its final and inexorable
fate. Not bad for an intracellular bug.

Sources of Funding
The authors are supported by the Ligue contre le Cancer (quipe labellise), AXA Chair for Longevity Research, Cancrople Ile-deFrance, Institut National du Cancer Fondation Bettencourt-Schueller,
Fondation de France, Fondation pour la Recherche Mdicale,
Agence National de la Recherche and the European Commission
(Apo-Sys, ArtForce, ChemoRes. Death-Train), and the LabEx
Immuno-Oncology.

Disclosures
None.

References
1. Altmann R. Die Elementarorganismen Und Ihre Beziehungen Zu Den
Zellen. Leipzig, Germany: Veit ; 1890.
2. Benda C. ber die Spermatogenese der Vertebraten und hheren
Evertebraten, II Theil: die Histiogenese der Spermien. Arch Anat
Physiol. 1898;73:393398.
3. Ernster L, Schatz G. Mitochondria: a historical review. J Cell Biol.
1981;91:227s255s.
4. Mitchell P. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature. 1961;191:144148.
5. Lockshin RA, Williams CM. Programmed cell death II. Endocrine potentiation of the breakdown of the intersegmental muscles of silkmoths.
J Insect Physiol. 1964;10:643649.
6. Sentman CL, Shutter JR, Hockenbery D, Kanagawa O, Korsmeyer SJ.
bcl-2 inhibits multiple forms of apoptosis but not negative selection in
thymocytes. Cell. 1991;67:879888.
7. Zamzami N, Marchetti P, Castedo M, Decaudin D, Macho A, Hirsch
T, Susin SA, Petit PX, Mignotte B, Kroemer G. Sequential reduction of mitochondrial transmembrane potential and generation of
reactive oxygen species in early programmed cell death. J Exp Med.
1995;182:367377.
8. Zamzami N, Marchetti P, Castedo M, Zanin C, Vayssire JL, Petit PX,
Kroemer G. Reduction in mitochondrial potential constitutes an early
irreversible step of programmed lymphocyte death in vivo. J Exp Med.
1995;181:16611672.

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Galluzzi et al Mitochondria Regulate Cellular Life and Death 1205


9. Liu X, Kim CN, Yang J, Jemmerson R, Wang X. Induction of apoptotic
program in cell-free extracts: requirement for dATP and cytochrome c.
Cell. 1996;86:147157.
10. Galluzzi L, Maiuri MC, Vitale I, Zischka H, Castedo M, Zitvogel L,
Kroemer G. Cell death modalities: classification and pathophysiological
implications. Cell Death Differ. 2007;14:12371243.
11. Kroemer G, Galluzzi L, Vandenabeele P, et al; Nomenclature Committee
on Cell Death 2009. Classification of cell death: recommendations of
the Nomenclature Committee on Cell Death 2009. Cell Death Differ.
2009;16:311.
12. Galluzzi L, Vitale I, Abrams JM, et al. Molecular definitions of cell death
subroutines: recommendations of the Nomenclature Committee on Cell
Death 2012. Cell Death Differ. 2012;19:107120.
13. Kroemer G, Galluzzi L, Brenner C. Mitochondrial membrane permeabilization in cell death. Physiol Rev. 2007;87:99163.
14. Kroemer G, Mario G, Levine B. Autophagy and the integrated stress
response. Mol Cell. 2010;40:280293.
15. Vandenabeele P, Galluzzi L, Vanden Berghe T, Kroemer G. Molecular
mechanisms of necroptosis: an ordered cellular explosion. Nat Rev Mol
Cell Biol. 2010;11:700714.
16. Galluzzi L, Joza N, Tasdemir E, Maiuri MC, Hengartner M, Abrams
JM, Tavernarakis N, Penninger J, Madeo F, Kroemer G. No death
without life: vital functions of apoptotic effectors. Cell Death Differ.
2008;15:11131123.
17. Galluzzi L, Kepp O, Trojel-Hansen C, Kroemer G. Non-apoptotic functions of apoptosis-regulatory proteins. EMBO Rep. 2012;13:322330.
18. Sagan L. On the origin of mitosing cells. J Theor Biol. 1967;14:255274.
19. Fernie AR, Carrari F, Sweetlove LJ. Respiratory metabolism: glycolysis,
the TCA cycle and mitochondrial electron transport. Curr Opin Plant
Biol. 2004;7:254261.
20. Reichert AS, Neupert W. Mitochondriomics or what makes us breathe.
Trends Genet. 2004;20:555562.
21. Houten SM, Wanders RJ. A general introduction to the biochemistry of mitochondrial fatty acid -oxidation. J Inherit Metab Dis.
2010;33:469477.
22. Huss JM, Kelly DP. Mitochondrial energy metabolism in heart failure: a
question of balance. J Clin Invest. 2005;115:547555.
23. Ajioka RS, Phillips JD, Kushner JP. Biosynthesis of heme in mammals.
Biochim Biophys Acta. 2006;1763:723736.
24. Miller WL. Role of mitochondria in steroidogenesis. Endocr Dev.
2011;20:119.
25. Watford M. The urea cycle: a two-compartment system. Essays Biochem.
1991;26:4958.
26. Lill R, Mhlenhoff U. Iron-sulfur-protein biogenesis in eukaryotes.
Trends Biochem Sci. 2005;30:133141.
27. Michal G. Biochemical Pathways: an Atlas of Biochemistry and
Molecular Biology. New York, NY: Wiley; 1999.
28. Voet D, Voet JG. Biochemistry. New York, NY: John Wiley; 1995.
29. Kodde IF, van der Stok J, Smolenski RT, de Jong JW. Metabolic and genetic regulation of cardiac energy substrate preference. Comp Biochem
Physiol, Part A Mol Integr Physiol. 2007;146:2639.
30. Campanella M, Pinton P, Rizzuto R. Mitochondrial Ca2+ homeostasis in
health and disease. Biol Res. 2004;37:653660.
31. Rizzuto R, Pozzan T. Microdomains of intracellular Ca2+: molecular determinants and functional consequences. Physiol Rev. 2006;86:369408.
32. Youle RJ, Narendra DP. Mechanisms of mitophagy. Nat Rev Mol Cell
Biol. 2011;12:914.
33. Galluzzi L, Kepp O, Kroemer G. Mitochondrial dynamics: a strategy for
avoiding autophagy. Curr Biol. 2011;21:R478R480.
34. Gomes LC, Di Benedetto G, Scorrano L. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat
Cell Biol. 2011;13:589598.
35. Wei Y, Scholes CP, King TE. Ubisemiquinone radicals from the cytochrome b-c1 complex of the mitochondrial electron transport chain
demonstration of QP-S radical formation. Biochem Biophys Res
Commun. 1981;99:14111419.
36. Dalton TP, Shertzer HG, Puga A. Regulation of gene expression by reactive oxygen. Annu Rev Pharmacol Toxicol. 1999;39:67101.
37. Balaban RS, Nemoto S, Finkel T. Mitochondria, oxidants, and aging.
Cell. 2005;120:483495.
38. Harrison R. Structure and function of xanthine oxidoreductase: where
are we now? Free Radic Biol Med. 2002;33:774797.
39. Yazdanpanah B, Wiegmann K, Tchikov V, Krut O, Pongratz C, Schramm
M, Kleinridders A, Wunderlich T, Kashkar H, Utermhlen O, Brning

JC, Schtze S, Krnke M. Riboflavin kinase couples TNF receptor 1 to


NADPH oxidase. Nature. 2009;460:11591163.
40. Danielson PB. The cytochrome P450 superfamily: biochemistry, evolution and drug metabolism in humans. Curr Drug Metab.
2002;3:561597.
41. Dikalov S, Griendling KK, Harrison DG. Measurement of reactive oxygen species in cardiovascular studies. Hypertension. 2007;49:717727.
42. Pestana CR, Silva CH, Uyemura SA, Santos AC, Curti C. Impact of adenosine nucleotide translocase (ANT) proline isomerization on Ca2+induced cysteine relative mobility/mitochondrial permeability transition
pore. J Bioenerg Biomembr. 2010;42:329335.
43. Costantini P, Belzacq AS, Vieira HL, Larochette N, de Pablo MA,
Zamzami N, Susin SA, Brenner C, Kroemer G. Oxidation of a critical thiol residue of the adenine nucleotide translocator enforces Bcl2-independent permeability transition pore opening and apoptosis.
Oncogene. 2000;19:307314.
44. Brenner C, Subramaniam K, Pertuiset C, Pervaiz S. Adenine nucleotide
translocase family: four isoforms for apoptosis modulation in cancer.
Oncogene. 2011;30:883895.
45. Temkin V, Huang Q, Liu H, Osada H, Pope RM. Inhibition of ADP/ATP
exchange in receptor-interacting protein-mediated necrosis. Mol Cell
Biol. 2006;26:22152225.
46. Sharma S, Dewald O, Adrogue J, Salazar RL, Razeghi P, Crapo JD,
Bowler RP, Entman ML, Taegtmeyer H. Induction of antioxidant gene
expression in a mouse model of ischemic cardiomyopathy is dependent
on reactive oxygen species. Free Radic Biol Med. 2006;40:22232231.
47. Singh M, Sharma H, Singh N. Hydrogen peroxide induces apoptosis in HeLa cells through mitochondrial pathway. Mitochondrion.
2007;7:367373.
48. Ye G, Metreveli NS, Donthi RV, Xia S, Xu M, Carlson EC, Epstein PN.
Catalase protects cardiomyocyte function in models of type 1 and type 2
diabetes. Diabetes. 2004;53:13361343.
49. Duranteau J, Chandel NS, Kulisz A, Shao Z, Schumacker PT. Intracellular
signaling by reactive oxygen species during hypoxia in cardiomyocytes.
J Biol Chem. 1998;273:1161911624.
50. Dhingra S, Sharma AK, Arora RC, Slezak J, Singal PK. IL-10 attenuates
TNF-alpha-induced NF kappaB pathway activation and cardiomyocyte
apoptosis. Cardiovasc Res. 2009;82:5966.
51. Afanasev I. ROS and RNS signaling in heart disorders: could antioxidant treatment be successful? Oxid Med Cell Longev. 2011;2011:293769.
52. Medzhitov R, Janeway CA Jr. Decoding the patterns of self and nonself
by the innate immune system. Science. 2002;296:298300.
53. Meylan E, Tschopp J, Karin M. Intracellular pattern recognition receptors in the host response. Nature. 2006;442:3944.
54. Takeuchi O, Akira S. Pattern recognition receptors and inflammation.
Cell. 2010;140:805820.
55. Kato H, Takeuchi O, Sato S, et al. Differential roles of MDA5 and RIG-I
helicases in the recognition of RNA viruses. Nature. 2006;441:101105.
56. Honda K, Yanai H, Negishi H, Asagiri M, Sato M, Mizutani T, Shimada
N, Ohba Y, Takaoka A, Yoshida N, Taniguchi T. IRF-7 is the master
regulator of type-I interferon-dependent immune responses. Nature.
2005;434:772777.
57. Wang J, Basagoudanavar SH, Wang X, Hopewell E, Albrecht R,
Garca-Sastre A, Balachandran S, Beg AA. NF-kappa B RelA subunit is
crucial for early IFN-beta expression and resistance to RNA virus replication. J Immunol. 2010;185:17201729.
58. West AP, Shadel GS, Ghosh S. Mitochondria in innate immune responses. Nat Rev Immunol. 2011;11:389402.
59. Seth RB, Sun L, Ea CK, Chen ZJ. Identification and characterization
of MAVS, a mitochondrial antiviral signaling protein that activates NFkappaB and IRF 3. Cell. 2005;122:669682.
60. Liu XY, Wei B, Shi HX, Shan YF, Wang C. Tom70 mediates activation of interferon regulatory factor 3 on mitochondria. Cell Res.
2010;20:9941011.
61. Yasukawa K, Oshiumi H, Takeda M, Ishihara N, Yanagi Y, Seya T,
Kawabata S, Koshiba T. Mitofusin 2 inhibits mitochondrial antiviral signaling. Sci Signal. 2009;2:ra47.
62. West AP, Brodsky IE, Rahner C, Woo DK, Erdjument-Bromage H,
Tempst P, Walsh MC, Choi Y, Shadel GS, Ghosh S. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS.
Nature. 2011;472:476480.
63. Singh SB, Ornatowski W, Vergne I, Naylor J, Delgado M, Roberts E,
Ponpuak M, Master S, Pilli M, White E, Komatsu M, Deretic V. Human
IRGM regulates autophagy and cell-autonomous immunity functions
through mitochondria. Nat Cell Biol. 2010;12:11541165.

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

1206Circulation ResearchOctober 12, 2012


64. Arnoult D, Soares F, Tattoli I, Castanier C, Philpott DJ, Girardin SE.
An N-terminal addressing sequence targets NLRX1 to the mitochondrial
matrix. J Cell Sci. 2009;122:31613168.
65. Kepp O, Galluzzi L, Kroemer G. Mitochondrial control of the NLRP3
inflammasome. Nat Immunol. 2011;12:199200.
66. Nakahira K, Haspel JA, Rathinam VA, Lee SJ, Dolinay T, Lam HC,
Englert JA, Rabinovitch M, Cernadas M, Kim HP, Fitzgerald KA, Ryter
SW, Choi AM. Autophagy proteins regulate innate immune responses
by inhibiting the release of mitochondrial DNA mediated by the NALP3
inflammasome. Nat Immunol. 2011;12:222230.
67. Zhang Q, Itagaki K, Hauser CJ. Mitochondrial DNA is released by shock
and activates neutrophils via p38 map kinase. Shock. 2010;34:5559.
68. Arnoult D, Soares F, Tattoli I, Girardin SE. Mitochondria in innate immunity. EMBO Rep. 2011;12:901910.
69. Costa A, Gupta R, Signorino G, Malara A, Cardile F, Biondo C, Midiri
A, Galbo R, Trieu-Cuot P, Papasergi S, Teti G, Henneke P, Mancuso G,
Golenbock DT, Beninati C. Activation of the NLRP3 inflammasome by
group B streptococci. J Immunol. 2012;188:19531960.
70. Zitvogel L, Kepp O, Galluzzi L, Kroemer G. Inflammasomes in carcinogenesis and anticancer immune responses. Nat Immunol. 2012;13:
343351.
71. Levine B, Mizushima N, Virgin HW. Autophagy in immunity and inflammation. Nature. 2011;469:323335.
72. Mizushima N, Levine B, Cuervo AM, Klionsky DJ. Autophagy fights
disease through cellular self-digestion. Nature. 2008;451:10691075.
73. Levine B, Kroemer G. Autophagy in the pathogenesis of disease. Cell.
2008;132:2742.
74. Mizushima N, Levine B. Autophagy in mammalian development and
differentiation. Nat Cell Biol. 2010;12:823830.
75. Green DR, Galluzzi L, Kroemer G. Mitochondria and the autophagyinflammation-cell death axis in organismal aging. Science. 2011;333:
11091112.
76. Rubinsztein DC, Mario G, Kroemer G. Autophagy and aging. Cell.
2011;146:682695.
77. Billia F, Hauck L, Konecny F, Rao V, Shen J, Mak TW. PTEN-inducible
kinase 1 (PINK1)/Park6 is indispensable for normal heart function. Proc
Natl Acad Sci USA. 2011;108:95729577.
78. Hoshino A, Matoba S, Iwai-Kanai E, Nakamura H, Kimata M, Nakaoka
M, Katamura M, Okawa Y, Ariyoshi M, Mita Y, Ikeda K, Ueyama T,
Okigaki M, Matsubara H. p53-TIGAR axis attenuates mitophagy
to exacerbate cardiac damage after ischemia. J Mol Cell Cardiol.
2012;52:175184.
79. Huang C, Andres AM, Ratliff EP, Hernandez G, Lee P, Gottlieb RA.
Preconditioning involves selective mitophagy mediated by Parkin and
p62/SQSTM1. PLoS ONE. 2011;6:e20975.
80. Tooze SA, Yoshimori T. The origin of the autophagosomal membrane.
Nat Cell Biol. 2010;12:831835.
81. Scherz-Shouval R, Elazar Z. Regulation of autophagy by ROS: physiology and pathology. Trends Biochem Sci. 2011;36:3038.
82. Emerling BM, Weinberg F, Snyder C, Burgess Z, Mutlu GM, Viollet B,
Budinger GR, Chandel NS. Hypoxic activation of AMPK is dependent
on mitochondrial ROS but independent of an increase in AMP/ATP ratio.
Free Radic Biol Med. 2009;46:13861391.
83. Mihaylova MM, Shaw RJ. The AMPK signalling pathway coordinates cell growth, autophagy and metabolism. Nat Cell Biol.
2011;13:10161023.
84. Egan DF, Shackelford DB, Mihaylova MM, et al. Phosphorylation of
ULK1 (hATG1) by AMP-activated protein kinase connects energy sensing to mitophagy. Science. 2011;331:456461.
85. Graef M, Nunnari J. Mitochondria regulate autophagy by conserved signalling pathways. EMBO J. 2011;30:21012114.
86. Wajant H. The Fas signaling pathway: more than a paradigm. Science.
2002;296:16351636.
87. Mehlen P. Dependence receptors: the trophic theory revisited. Sci Signal.
2010;3:pe47.
88. Muzio M, Chinnaiyan AM, Kischkel FC, ORourke K, Shevchenko A,
Ni J, Scaffidi C, Bretz JD, Zhang M, Gentz R, Mann M, Krammer PH,
Peter ME, Dixit VM. FLICE, a novel FADD-homologous ICE/CED-3like protease, is recruited to the CD95 (Fas/APO-1) deathinducing signaling complex. Cell. 1996;85:817827.
89. Barnhart BC, Alappat EC, Peter ME. The CD95 type I/type II model.
Semin Immunol. 2003;15:185193.
90. Scaffidi C, Fulda S, Srinivasan A, Friesen C, Li F, Tomaselli KJ, Debatin
KM, Krammer PH, Peter ME. Two CD95 (APO-1/Fas) signaling pathways. EMBO J. 1998;17:16751687.

91. Li H, Zhu H, Xu CJ, Yuan J. Cleavage of BID by caspase 8 mediates the mitochondrial damage in the Fas pathway of apoptosis. Cell.
1998;94:491501.
92. Luo X, Budihardjo I, Zou H, Slaughter C, Wang X. Bid, a Bcl2 interacting
protein, mediates cytochrome c release from mitochondria in response to
activation of cell surface death receptors. Cell. 1998;94:481490.
93. Yin XM, Wang K, Gross A, Zhao Y, Zinkel S, Klocke B, Roth KA,
Korsmeyer SJ. Bid-deficient mice are resistant to Fas-induced hepatocellular apoptosis. Nature. 1999;400:886891.
94. Mille F, Thibert C, Fombonne J, Rama N, Guix C, Hayashi H, Corset V,
Reed JC, Mehlen P. The Patched dependence receptor triggers apoptosis
through a DRAL-caspase-9 complex. Nat Cell Biol. 2009;11:739746.
95. Green DR, Kroemer G. The pathophysiology of mitochondrial cell
death. Science. 2004;305:626629.
96. Tait SW, Green DR. Mitochondria and cell death: outer membrane permeabilization and beyond. Nat Rev Mol Cell Biol. 2010;11:621632.
97. Brenner C, Grimm S. The permeability transition pore complex in cancer
cell death. Oncogene. 2006;25:47444756.
98. Baines CP, Kaiser RA, Sheiko T, Craigen WJ, Molkentin JD. Voltagedependent anion channels are dispensable for mitochondrial-dependent
cell death. Nat Cell Biol. 2007;9:550555.
99. Galluzzi L, Kroemer G. Mitochondrial apoptosis without VDAC. Nat
Cell Biol. 2007;9:487489.
100. Kokoszka JE, Waymire KG, Levy SE, Sligh JE, Cai J, Jones DP,
MacGregor GR, Wallace DC. The ADP/ATP translocator is not essential for the mitochondrial permeability transition pore. Nature.
2004;427:461465.
101. Baines CP, Kaiser RA, Purcell NH, Blair NS, Osinska H, Hambleton
MA, Brunskill EW, Sayen MR, Gottlieb RA, Dorn GW, Robbins J,
Molkentin JD. Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature. 2005;434:658662.
102. Schinzel AC, Takeuchi O, Huang Z, Fisher JK, Zhou Z, Rubens J, Hetz
C, Danial NN, Moskowitz MA, Korsmeyer SJ. Cyclophilin D is a component of mitochondrial permeability transition and mediates neuronal cell death after focal cerebral ischemia. Proc Natl Acad Sci USA.
2005;102:1200512010.
103. Kroemer G, Martin SJ. Caspase-independent cell death. Nat Med.
2005;11:725730.
104. Li P, Nijhawan D, Budihardjo I, Srinivasula SM, Ahmad M, Alnemri
ES, Wang X. Cytochrome c and dATP-dependent formation of Apaf1/caspase-9 complex initiates an apoptotic protease cascade. Cell.
1997;91:479489.
105. Vercammen D, Beyaert R, Denecker G, Goossens V, Van Loo G,
Declercq W, Grooten J, Fiers W, Vandenabeele P. Inhibition of caspases
increases the sensitivity of L929 cells to necrosis mediated by tumor
necrosis factor. J Exp Med. 1998;187:14771485.
106. Vercammen D, Brouckaert G, Denecker G, Van de Craen M, Declercq
W, Fiers W, Vandenabeele P. Dual signaling of the Fas receptor: initiation of both apoptotic and necrotic cell death pathways. J Exp Med.
1998;188:919930.
107. Degterev A, Huang Z, Boyce M, Li Y, Jagtap P, Mizushima N, Cuny GD,
Mitchison TJ, Moskowitz MA, Yuan J. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for ischemic brain injury. Nat
Chem Biol. 2005;1:112119.
108. Degterev A, Hitomi J, Germscheid M, Chen IL, Korkina O, Teng X,
Abbott D, Cuny GD, Yuan C, Wagner G, Hedrick SM, Gerber SA,
Lugovskoy A, Yuan J. Identification of RIP1 kinase as a specific cellular
target of necrostatins. Nat Chem Biol. 2008;4:313321.
109. Galluzzi L, Vanden Berghe T, Vanlangenakker N, Buettner S, Eisenberg
T, Vandenabeele P, Madeo F, Kroemer G. Programmed necrosis from
molecules to health and disease. Int Rev Cell Mol Biol. 2011;289:135.
110. Cho YS, Challa S, Moquin D, Genga R, Ray TD, Guildford M, Chan
FK. Phosphorylation-driven assembly of the RIP1-RIP3 complex
regulates programmed necrosis and virus-induced inflammation. Cell.
2009;137:11121123.
111. He S, Wang L, Miao L, Wang T, Du F, Zhao L, Wang X. Receptor interacting protein kinase-3 determines cellular necrotic response to TNFalpha. Cell. 2009;137:11001111.
112. Zhang DW, Shao J, Lin J, Zhang N, Lu BJ, Lin SC, Dong MQ, Han J.
RIP3, an energy metabolism regulator that switches TNF-induced cell
death from apoptosis to necrosis. Science. 2009;325:332336.
113. Zong WX, Ditsworth D, Bauer DE, Wang ZQ, Thompson CB. Alkylating
DNA damage stimulates a regulated form of necrotic cell death. Genes
Dev. 2004;18:12721282.

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Galluzzi et al Mitochondria Regulate Cellular Life and Death 1207


114. Goossens V, Stang G, Moens K, Pipeleers D, Grooten J. Regulation
of tumor necrosis factor-induced, mitochondria- and reactive oxygen
species-dependent cell death by the electron flux through the electron
transport chain complex I. Antioxid Redox Signal. 1999;1:285295.
115. Kim YS, Morgan MJ, Choksi S, Liu ZG. TNF-induced activation of the
Nox1 NADPH oxidase and its role in the induction of necrotic cell death.
Mol Cell. 2007;26:675687.
116. Schulze-Osthoff K, Bakker AC, Vanhaesebroeck B, Beyaert R, Jacob
WA, Fiers W. Cytotoxic activity of tumor necrosis factor is mediated by
early damage of mitochondrial functions. Evidence for the involvement
of mitochondrial radical generation. J Biol Chem. 1992;267:53175323.
117. Suffys P, Beyaert R, De Valck D, Vanhaesebroeck B, Van Roy F, Fiers W.
Tumour-necrosis-factor-mediated cytotoxicity is correlated with phospholipase-A2 activity, but not with arachidonic acid release per se. Eur J
Biochem. 1991;195:465475.
118. Van Herreweghe F, Mao J, Chaplen FW, Grooten J, Gevaert K,
Vandekerckhove J, Vancompernolle K. Tumor necrosis factor-induced modulation of glyoxalase I activities through phosphorylation
by PKA results in cell death and is accompanied by the formation
of a specific methylglyoxal-derived AGE. Proc Natl Acad Sci USA.
2002;99:949954.
119. Goossens V, Grooten J, Fiers W. The oxidative metabolism of glutamine. A modulator of reactive oxygen intermediate-mediated cytotoxicity of tumor necrosis factor in L929 fibrosarcoma cells. J Biol Chem.
1996;271:192196.
120. Cao G, Xing J, Xiao X, Liou AK, Gao Y, Yin XM, Clark RS, Graham
SH, Chen J. Critical role of calpain I in mitochondrial release of
apoptosis-inducing factor in ischemic neuronal injury. J Neurosci.
2007;27:92789293.
121. Wang Y, Kim NS, Haince JF, Kang HC, David KK, Andrabi SA, Poirier
GG, Dawson VL, Dawson TM. Poly(ADP-ribose) (PAR) binding to

apoptosis-inducing factor is critical for PAR polymerase-1-dependent


cell death (parthanatos). Sci Signal. 2011;4:ra20.
122. Wang Y, Kim NS, Li X, Greer PA, Koehler RC, Dawson VL, Dawson
TM. Calpain activation is not required for AIF translocation in PARP-1dependent cell death (parthanatos). J Neurochem. 2009;110:687696.
123. Sun L, Wang H, Wang Z, He S, Chen S, Liao D, Wang L, Yan J,
Liu W, Lei X, Wang X. Mixed lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell.
2012;148:213227.
124. Wang Z, Jiang H, Chen S, Du F, Wang X. The mitochondrial phosphatase PGAM5 functions at the convergence point of multiple necrotic
death pathways. Cell. 2012;148:228243.
125. Whelan RS, Kaplinskiy V, Kitsis RN. Cell death in the pathogenesis of heart disease: mechanisms and significance. Annu Rev Physiol.
2010;72:1944.
126. Kung G, Konstantinidis K, Kitsis RN. Programmed necrosis, not apoptosis, in the heart. Circ Res. 2011;108:10171036.
127. Smith CC, Davidson SM, Lim SY, Simpkin JC, Hothersall JS, Yellon
DM. Necrostatin: a potentially novel cardioprotective agent? Cardiovasc
Drugs Ther. 2007;21:227233.
128. Fiorillo C, Ponziani V, Giannini L, Cecchi C, Celli A, Nassi N, Lanzilao
L, Caporale R, Nassi P. Protective effects of the PARP-1 inhibitor
PJ34 in hypoxic-reoxygenated cardiomyoblasts. Cell Mol Life Sci.
2006;63:30613071.
129. Torella D, Ellison GM, Mndez-Ferrer S, Ibanez B, Nadal-Ginard B.
Resident human cardiac stem cells: role in cardiac cellular homeostasis
and potential for myocardial regeneration. Nat Clin Pract Cardiovasc
Med. 2006;3 Suppl 1:S813.
130. Behfar A, Terzic A. Derivation of a cardiopoietic population from human mesenchymal stem cells yields cardiac progeny. Nat Clin Pract
Cardiovasc Med. 2006;3:S78S82.

Downloaded from http://circres.ahajournals.org/ by guest on August 11, 2014

Vous aimerez peut-être aussi