You are on page 1of 14

REVIEWS

Mitonuclear communication in
homeostasis and stress
Pedro M. Quirós*, Adrienne Mottis* and Johan Auwerx
Abstract | Mitochondria participate in crucial cellular processes such as energy harvesting and
intermediate metabolism. Although mitochondria possess their own genome — a vestige of their
bacterial origins and endosymbiotic evolution — most mitochondrial proteins are encoded in the
nucleus. The expression of the mitochondrial proteome hence requires tight coordination
between the two genomes to adapt mitochondrial function to the ever-changing cellular milieu.
In this Review, we focus on the pathways that coordinate the communication between
mitochondria and the nucleus during homeostasis and mitochondrial stress. These pathways
include nucleus-to‑mitochondria (anterograde) and mitochondria-to‑nucleus (retrograde)
communication, mitonuclear feedback signalling and proteostasis regulation, the integrated
stress response and non-cell-autonomous communication. We discuss how mitonuclear
communication safeguards cellular and organismal fitness and regulates lifespan.

Endosymbionts Mitochondria are derived from α‑proteobacteria that response’ that signals to the nucleus to alter the expression
Organisms that live within were engulfed by the precursor of modern eukaryotic cells of nuclear genes to modify cellular function and repro‑
another organism in a before evolving as endosymbionts over millions of years. gramme cell metabolism. The integration of these antero­
symbiotic relationship.
These organelles have maintained some of their ances‑ grade (from nucleus to mitochondria) and retrograde
Electron transport chain
tral bacterial characteristics, such as a circular genome (from mitochondria to nucleus) signals — also known
(ETC). A group of protein and the capacity to produce ATP — mito­chondria are as mitonuclear communication — constitutes a robust
complexes that pass electrons the core of energy metabolism within the cell1. During network that helps cells to maintain homeostasis under
from one to another via redox evolution, however, many of the proteo­bacterial genes basal conditions and enables their adaptation to various
reactions coupled with the
were pro­gressively transferred to the nuclear genome, stressors. However, mitonuclear communication is most
transfer of protons across the
inner mitochondrial and mitochondria acquired new components and func‑ often bidirectional and of a hormetic nature, combining
membrane, creating a proton tions from the host cell, resulting in profound changes in anterograde and retrograde signals. Thus, retrograde sig‑
gradient that drives both the mitochondrial and the nuclear genome and pro‑ nals that are generated in reaction to some mitochondrial
ATP synthesis.
teome2. Out of more than 1,200 proteins that are present stressors trigger particular nuclear responses, which will
Intermediate metabolism
in mitochondria, only 13 are encoded by the mammalian induce the speci­fic expression of certain mitochondrial
The intermediate steps by mitochondrial DNA (mtDNA) — 12 in Caenorhabditis proteins; in turn, these proteins will resolve the original
which nutrient molecules or elegans3,4. These 13 proteins constitute important compo‑ perturbations in the mitochondria through ‘mitonuclear
foodstuffs are metabolized nents of all complexes of the electron transport chain (ETC) feedback’. Depending on the severity and nature of the
intracellularly.
except for complex II, the components of which are exclu‑ initial stress signal, mitonuclear communication can
sively encoded within the nuclear genome. Therefore, result in an ‘integrated stress response’ (ISR). The ISR
the nucleus and mito­chondria must continuously can also be induced by other stress signals, such as those
­coordinate the transcription and translation, as well as originating from endoplasmic reticulum (ER) stress,
Laboratory for Integrative
the t­ ranslocation and import, of ­mitochondrial proteins3. triggering a global cellular response by decreasing pro‑
and Systems Physiology,
Ecole Polytechnique Fédérale Mitochondria are not only at the heart of cellular tein synthesis. Moreover, mitonuclear stress can also trig‑
de Lausanne (EPFL), energy harvesting, however, they also regulate many ger a non-cell-autonomous response, which modulates
CH‑1015, Lausanne, aspects of intermediate metabolism, calcium buffering and the function of distant cells to facilitate an organismal
Switzerland. processes such as apoptosis. Consequently, mitochon‑ response or adaptation to stress (FIG. 1).
*These authors contributed
equally to this article.
drial function is under tight nuclear control, through Interest in understanding these complex mito­nuclear
Correspondence to J.A. so‑called ‘anterograde regulation’, which can decrease responses has increased during the past decade, mainly
admin.auwerx@epfl.ch or increase mitochondrial activity, as well as promoting because of their key involvement in health and lifespan.
doi:10.1038/nrm.2016.23 mito­chondrial biogenesis, depending on cellular needs. In this Review, we discuss the functional interaction
Published online 9 March 2016 Conversely, mitochondria can generate a ‘retrograde between mitochondria and the nucleus, with a particular

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 213


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Anterograde Retrograde mitochondrial transcription and translation are encoded


regulation response in the nucleus, and many of them are regulated by NRFs.
Mitochondrial transcription is controlled by mito‑
Nucleus
chondrial transcription and termination factors that
collabor­ate with the mitochondrial RNA polymerase to
Mitochondrion generate two polycistronic mitochondrial RNAs9, which
are translated by mitochondrial ribosomes in a process
Extracellular that is tightly regulated by nuclear-encoded factors and
communication Mitonuclear
feedback integrated into cellular processes10.
ER Certain nuclear receptors can also activate the expres‑
Integrated
sion of nuclear-encoded mitochondrial proteins. For
stress response example, some peroxisome proliferator-activated recep‑
tors (PPARs), such as PPARδ, stimulate the expression
Figure 1 | Mitonuclear communication. Mitochondria and
Nature Reviews | Molecular Cell
the nucleus communicate closely with each other. From a Biology of enzymes that are involved in mitochondrial fatty acid
‘mitocentric’ point of view, as illustrated here, signals sent oxidation, especially in the heart and muscle, in which this
from the nucleus to mitochondria constitute anterograde process is key for generating ATP11,12. Other PPARs have
regulation, whereas those sent by mitochondria to the similar functions in other tissues (reviewed in REF. 13).
nucleus are defined as the retrograde response. Although The oestrogen-related receptors (ERRs) ERRα, β and γ are
mitonuclear feedback responses can be orchestrated associated with the expression of nuclear-encoded mito‑
bidirectionally, they usually originate from mitochondria chondrial proteins that are involved in the tricarboxylic
and consequently induce a nuclear response that is acid (TCA) cycle, OXPHOS and fatty acid oxidation14,15.
specifically geared towards the mitochondria. The Fine-tuning of the action of these transcription fac‑
integrated stress response is a general cellular stress
tors is coordinated by co-activators or co-­repressors16.
pathway that controls cytosolic protein synthesis; this stress
response can be triggered in mitochondria, in the
Within the context of mitochondria, PPARγ c­ o‑
endoplasmic reticulum (ER) and in the cytosol. Mitonuclear activator 1α (PGC1α), PGC1β and the PGC-related
communication can also be extracellular, as mitochondria co-­activator PRC are known to activate the NRF, ERR
can send extracellular cues (known as mitokines), which and PPAR factors to induce mitochondrial bio­genesis17,18,
affect nuclear regulation in a non-cell-autonomous manner. whereas ­co-repressors such as nuclear receptor co-­
repressor 1 or r­ eceptor-interacting protein 140 can inhibit
their function19,20.
emphasis on how this interaction is affected by cellu‑ These anterograde signals are activated by upstream
lar stress. We also describe the different pathways that sensors that detect changes in metabolic conditions.
coordi­nate mitonuclear communication and how this A decrease in ATP synthesis, as seen during exercise and
process influences cellular function and organismal caloric restriction, increases the AMP/ATP ratio and acti‑
homeostasis and lifespan. vates AMP-activated protein kinase (AMPK), which, in
turn, increases cellular NAD+ levels. Increased NAD+
Anterograde regulation levels lead to activation of sirtuin 1 (SIRT1)21,22 — a
Mitochondrial function is controlled by the nucleus NAD+-dependent deacetylase that positively regulates
through anterograde signals, which allows these PGC1α — and subsequent activation of mitochondrial
organelles to adapt to the cellular milieu. This regu‑ energy metabolism and biogenesis23. Increases in levels
lation is based mainly on the expression of nuclear-­ of calcium ions (Ca2+) during exercise activate PGC1α
encoded mitochondrial proteins that induce mtDNA — either directly through Ca2+/calmodulin-­dependent
gene expression, and on several transcription factors protein kinase type IV (CaMKIV)24 or through the
and co‑­regulators that regulate the expression of the activation of AMPK by Ca2+/calmodulin-­dependent
­nuclear-encoded mitochondrial proteome (FIG. 2). protein kinase kinase‑β (CaMKKβ) 25 — and the
Calcium buffering The expression of nuclear-encoded mitochondrial myocyte-­specific enhancer factor 2A, promoting mito‑
A mechanism that regulates proteins is predominantly mediated by two nuclear tran‑ chondrial bio­genesis. Cold stress also activates PGC1α
calcium ion (Ca2+) scription factors: nuclear respiratory factor 1 (NRF1) through cAMP-­dependent protein kinase (also known
concentrations within and GA‑binding protein‑α (GABPα; also known as as protein kinase A (PKA)) and the cAMP response
the cytoplasm.
NRF2α). Both transcription factors, but mainly NRF1, ­element-binding protein (CREB) (FIG. 2).
Oxidative phosphorylation control the expression of genes encoding cytochrome c In addition to its role in homeostasis, anterograde
(OXPHOS). A biochemical and the vast majority of nuclear-encoded subunits that communication can also signal the existence of nuclear
pathway within mitochondria are involved in oxidative phosphorylation (OXPHOS), stress to mitochondria, thereby bringing about a reduc‑
that generates ATP through
as well as proteins implicated in mtDNA replication, tion in mitochondrial metabolism to attenuate the
the oxidation of nutrients.
transcription and translation. They also control the consequences of nuclear DNA damage. For instance,
Nuclear receptors expression of mitochondrial ribosomal proteins and artificially inducing telomere dysfunction in mice is
Ligand-dependent tRNA synthetases, rate-limiting haem biosynthetic associated with impaired mitochondrial function and
transcription factors that are enzymes, ion channel proteins and various compo‑ diminished biogenesis through the repression of PGC1α-
characterized by a central
DNA-binding domain and
nents of the protein import machinery 5–8. The expres‑ and PGC1β‑driven programmes by p53  (REF.  26) .
a carboxy-terminal sion of mitochondrial-encoded proteins also depends Activation of poly(ADP-ribose) polymerase 1 (PARP1)
ligand-binding domain. on nuclear regulation, as all of the factors that activate and PARP2 in response to DNA strand breaks likewise

214 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Exercise, Cold Exercise


Energetic stress responses. The retrograde stress
Ca2+
CR response was initially described in yeast, in which it
regulates carbon and nitrogen metabolism31. Yeast
PARPi AMPK CAMKKβ cAMP Ca2+ contains the retrograde response genes RTG1, RTG2
and RTG3, which induce different metabolic enzymes
DNA
and thereby activate alternative metabolic pathways to
PARP1 NAD NR PKA counteract mitochondrial dysfunction32; these include
damage
peroxisomal anaplerotic reactions such as the glyoxy-
SIRT1 NCOR1 late cycle33. Independent of the induction of the RTG
CREB CAMKIV Calcineurin genes, other pathways, including the carbon catabolite-­
RIP140
derepressing protein kinase (also known as Snf1; the
p53 PGC1α yeast orthologue of AMPK) pathway, the TORC1
Telomere pathway and the Sir2 pathway, can also be activated in
dysfunction
response to energetic stress in yeast to prevent abnormal
histone deacetylation and reprogramme metabolism,
PGC1α PGC1α PGC1α Calcineurin to regulate protein synthesis, and to improve protein
PPARs NRFs ERRs MEF2 homeostasis, respectively 34–36.
Although the absence of clear RTG orthologues
in other species hindered the simple generalization
• Mitochondrial biogenesis
• OXPHOS of the retrograde signal, some pathways are acti‑
• TCA vated in a similar fashion after mitochondrial stress,
• Fatty acid oxidation
Nucleus testify­ing to functional conservation. In C. elegans,
mutations in the ETC subunit-encoding genes iron–
Figure 2 | Anterograde regulation. Different extracellular stimuli
Nature Reviews can regulate
| Molecular Cell Biology
mitochondrial function through nuclear regulation. Exercise and caloric restriction (CR) sulfur protein 1 (isp‑1) and clk‑1 activate AMPK sub‑
activate AMP-activated protein kinase (AMPK), which, in turn, activates sirtuin 1 (SIRT1) unit α2 (REF. 37), which transduces changes in cellular
through the regulation of NAD+ levels. NAD+ can also be modulated by NAD+ precursors energy levels resulting from mitochondrial stress38.
such as nicotinamide riboside (NR) or by inhibitors of poly(ADP-ribose) polymerase 1 Defects in the TCA cycle also trigger the expression
(PARP1), termed PARPi, which are activated upon DNA damage. SIRT1 activates PPAR-γ of glyoxy­late cycle genes such as gei‑7 — an isocitrate
co-activator 1α (PGC1α), which, through interactions with peroxisome proliferator- lyase/malate synthase — through mechanisms that are
activated receptors (PPARs), nuclear respiratory factors (NRFs) or oestrogen-related ill-defined but that are most likely to involve AMP-
receptors (ERRs), induces the expression of mitochondrial genes. PGC1α signalling is activated kinase 2 (aak‑2), mediating the transition
antagonized by the co-repressors nuclear receptor co-repressor 1 (NCOR1) and to mitochondria-independent energy production39,40.
receptor-interacting protein 140 (RIP140), or by p53 in response to telomere dysfunction.
Administration of α‑ketoglutarate — a key TCA cycle
Cold and exercise can also activate the expression of mitochondrial genes by activating
PGC1α through the cAMP–protein kinase A (PKA)– cAMP response element-binding intermediate — also mediates cellular adaptations,
protein (CREB) pathway, or through Ca2+ signalling by Ca2+/calmodulin-dependent leading to lifespan ­extension through TOR, aak‑2 and
protein kinase kinase‑β (CAMKKβ) or Ca2+/calmodulin-dependent protein kinase type IV daf‑16 (REF. 41).
(CAMKIV). Additionally, Ca2+ induces mitochondrial gene expression by activating In mammals, the retrograde response has also
myocyte-specific enhancer factor (MEF2) through the phosphatase calcineurin. been linked to mTOR and AMPK; however, as mam‑
OXPHOS, oxidative phosphorylation; TCA, tricarboxylic acid. malian cells lack the glyoxylate cycle, they use differ‑
ent anaplerotic reactions to adapt their metabolism to
manage energy deficits. A decrease in mito­chondrial
represses mitochondrial function by depleting cellular ATP synthesis stimulates AMPK, promoting the acti‑
NAD+ levels27 and consequently inhibiting the activity of vation of PGC1α, which stimulates mitochondrial
SIRT1 (REF. 28). Treatment with NAD+ precursors, such energy metabolism and biogenesis23 (FIG. 3). AMPK
as nicotinamide riboside, or with PARP inhibitors can also triggers the mitochondrial quality control system,
Metabolic reprogramming
The set of changes in cellular restore NAD+ levels and activate an anterograde signal which regulates mitochondrial dynamics and induces
metabolism that allow through SIRT1–PGC1α29 (FIG. 2). mitophagy42. Similarly, reduced mTOR activity, such
quiescent cells to proliferate; as under energetic stress (for example, during exer‑
it is considered to be a Retrograde response cise and nutrient limitation), facilitates mitochondrial
hallmark of cancer.
Mitochondria generate a wide range of retrograde retro­grade signalling, whereas its activation inhibits
Reactive oxygen species signals, through which they regulate various cellular the retrograde response43.
(ROS). Reactive molecules and organismal activities and protect against mito‑
generated in cells by the chondrial dysfunction by activating the expression Ca2+-dependent stress responses. Mitochondria are
reduction of oxygen with a
of nuclear genes involved in metabolic reprogramming essential to regulate the levels of intracellular Ca 2+
single electron (superoxide),
two electrons (hydrogen or stress defence30. The retrograde response exists in (REF. 44). Different mitochondrial stressors, such as the
peroxide) or three electrons all organisms, but the regulation and nature of the loss of (or mutations in) mtDNA, disruption of ETC
(hydroxyl radical). pathways involved can vary. Despite the pleiotropic complexes and OXPHOS, or treatment with ionophores,
nature of the retrograde signals, these pathways can trigger the loss of mitochondrial membrane potential and
Anaplerotic reactions
Reactions that replenish
be classified into energetic stress responses, Ca2 +- the subsequent release of Ca2+ into the cytoplasm45–48.
intermediates of dependent responses and reactive oxygen species (ROS) Elevated levels of free cytosolic Ca 2+ activate the
metabolic pathways. stress responses, depending on the trigger. phosphatase calcineurin, which activates the nuclear

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 215


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

PGC1α AMPK Low ATP

JNK ROS
Mitochondrion
• Mitochondrial biogenesis ETC
NFE2L2
• Mitochondrial quality control mtDNA
KEAP1
OXPHOS
mtDNA defects
dysfunction
Antioxidant response NFE2L2 Ca2+
degradation

NF-κB Loss of ∆ψm


• Calcium metabolism genes
• Glycolytic genes Calcineurin Ca2+ CAMKKβ

NFATC

Nucleus
PKC JNK p38 CAMKIV

EGR1 ATF2 CEBPδ CHOP CREB

Figure 3 | The retrograde response. Defects in oxidative phosphorylation (OXPHOS) and mitochondrial DNA (mtDNA),
Nature Reviews | Molecular Cell Biology
caused by damage in the electron transport chain (ETC) and by mutations, respectively, activate a retrograde response to
the nucleus that can be triggered by a decrease in the level of ATP, increased signalling by reactive oxygen species (ROS)
or the release of Ca2+ from mitochondria. Low ATP levels activate AMP-activated protein kinase (AMPK), which stimulates
mitochondrial biogenesis and quality control. An increase in ROS also activates anterograde regulation through AMPK
or the c‑Jun N‑terminal kinase (JNK) pathway, by activating PPARγ co-activator 1α (PGC1α). Increased levels of ROS inhibit
the kelch-like ECH-associated protein 1 (KEAP1)‑mediated proteasomal degradation of nuclear factor erythroid 2‑related
factor 2 (NFE2L2) and facilitates the translocation of NFE2L2 to the nucleus and the subsequent activation of an
antioxidant response. Loss of the mitochondrial membrane potential (Δψm) results in the release of Ca2+ from
mitochondria, inducing the expression of genes for calcium metabolism and glycolysis through two mechanisms. First,
calcineurin translocates to the nucleus with nuclear factor-κB (NF‑κB), which can also be activated by ROS, and with
nuclear factor of activated T cells (NFATC). Alternatively, Ca2+ can activate several kinases, such as protein kinase C (PKC),
JNK–p38 and Ca2+/‌calmodulin-dependent protein kinase type IV (CAMKIV), which, in turn, activate different transcription
Glyoxylate cycle factors such as early growth response protein 1 (EGR1), ATF2, cAMP response element-binding protein (CREB),
An anabolic pathway, and a CCAAT/enhancer-binding protein‑δ (CEBPδ) and CEBP homologous protein (CHOP). The release of Ca2+ can also activate
variation of the tricarboxylic
anterograde regulation through Ca2+/calmodulin-­dependent protein kinase kinase‑β (CAMKKβ).
acid cycle, that converts two
acetyl-CoA molecules to
dicarboxylic acid (succinate);
this cycle is not present factor‑κB (NF‑κB) p105 subunit and nuclear factor ROS-dependent responses. Mitochondrial ROS are
in mammals. of activated T cells (NFATC)49–51 (FIG. 3). Both of these gener­ated through aerobic metabolism, often result‑
transcription factors then translocate to the nucleus, ing from defective ETC, and they function directly by
Mitophagy
where they promote the synthesis of proteins involved regulating redox biology and as signalling molecules
A selective form of autophagy
that is responsible for in Ca 2+ transport and storage 50,52, and of glycolytic in numerous cellular processes, under normal as well
the elimination of and gluconeogenic enzymes 52. Elevated Ca 2+ levels as stress conditions55,56. In C. elegans, mutation in clk‑1
defective mitochondria. also directly activate different Ca2+-regulated kinases, — a mitochondrial hydrolase that is required for syn‑
such as CAMKIV, Ca2+-dependent protein kinase C, thesizing the ETC component ubiquinone57 — inhibits
Ionophores
Molecules that bind to ions
c‑Jun N-terminal kinase (JNK) and p38 MAPK; respiration and increases ROS levels, which stabilize
and allow their passage these, in turn, stimulate different transcription fac‑ and activate hypoxia-inducible factor 1 (HIF‑1), a tran‑
across a membrane or tors such as CREB, early growth response protein 1 scription factor that regulates adaptation to low oxy‑
a lipophilic phase. (EGR1), cAMP-dependent transcription factor ATF2, gen levels58. ROS also signal through the orthologue of
CCAAT/enhancer-binding protein‑δ (CEBPδ) and stress-­inducible p38 MAPK, PMK‑1, which phosphory­
Mitochondrial membrane
potential CEBP homologous protein (CHOP)45,49 (FIG. 3). Which lates and activates the transcription factor skinhead 1
An electrochemical potential of these transcription factors is activated depends on (SKN‑1), thereby promoting its nuclear localization59–61.
formed by chemiosmosis that the cell type and the activating stimulus; furthermore, These mechanisms induce the expression of adaptive
results from the proton activation not only facilitates mitochondrial adaptation ROS defence genes, such as superoxide dismutase and
gradient generated across the
inner mitochondrial membrane
but also leads to pleiotropic responses affecting calcium catalase60. CLK‑1 itself can also function as a messen‑
by the mitochondrial metabolism, insulin signalling, glucose metabolism ger of mitochondrial stress, as certain CLK‑1 isoforms
respiratory chain. and cell proliferation24,25,48–54. translocate to the nucleus following ROS formation

216 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

and induce a protective programme, while inhibiting and of metabolic genes to adapt to the provoking stress.
the mitochondrial unfolded protein response (UPRmt) The UPRmt is triggered by mitochondrial proteotoxic
by binding chromatin, a property that is conserved in stresses, such as the accumulation of unfolded proteins,
mammalian cells62. impairment of the protein quality control system, mito-
In Drosophila melanogaster, disruption of complex I of nuclear imbalance or inhibition of the ETC74. The UPRmt
the ETC triggers a signalling cascade that involves ROS- has been characterized most fully in C. elegans, in which
mediated dimerization of the D. melanogaster homologue it can be induced using RNA-mediated interference
of the mammalian kinase ASK1, which promotes Jnk (RNAi) directed against some nuclear-encoded mito‑
signalling to the forkhead box O (Foxo) transcription chondrial proteins, such as cytochrome c oxidase (cco‑1),
factor and activates the cyclin-dependent kinase inhib‑ a mitochondrial protein quality-control protease, spastic
itor Dacapo63. In parallel, increased AMP levels arising paraplegia 7 (spg‑7); or several mitochondrial ribosomal
from a decrease in mitochondrial ATP synthesis activate proteins, typified by mitochondrial ribo­somal protein S5
AMPK and p53, leading to the down­regulation of cyclin (mrps‑5)75,76 (see below for more details about mrps‑5).
E1 and subsequent G1 arrest as part of the G1–S cell cycle Furthermore, doxycycline and chloram­phenicol —
checkpoint63. Through the Jnk pathway, ROS also activate which are antibiotics that target bacterial as well as
the UPRmt in flies with muscle-­specific depletion of the mitochondrial translation owing to the evolutionary
complex I subunit NADH–ubiquinone oxidoreductase conservation of the machinery — activate the UPRmt by
75 kDa (Nd75)64 (see below). depleting mitochondrial-encoded ETC subunits, simi‑
In mammals, increases in ROS to levels that are not lar to mtDNA depletion by ethidium bromide75. PARP
deleterious to cell function induce a retrograde signal to inhibitors or NAD+ precursors that activate SIRT1 and
activate detoxification enzymes and antioxidant proteins mitochondrial function, as well as the longevity com‑
in mitochondria and the cytosol65–68. This activation is pounds resveratrol and rapamycin, also activate the
mediated by the binding of transcription factors such as UPRmt owing to a mitonuclear imbalance, which in
nuclear factor erythroid 2‑related factor 2 (NFE2L2; also this case is generated by enhancing the production of
known as NRF2, but not to be confused with GABPα mtDNA-encoded ETC proteins77–79.
mentioned above) to antioxidant response elements69 The UPRmt in C. elegans involves the digestion by the
(FIG. 3). Increased levels of mitochondrial ROS were also matrix protease CLPP of unfolded or unassembled mito‑
reported to activate NF‑κB, thereby promoting cellular chondrial proteins into peptides, which are transported
proliferation and survival in cancer cells51. ROS can also into the cytoplasm by the HAF‑1 transporter 80 (FIG. 4a).
induce mitochondrial biogenesis and the expression of Through a poorly understood mechanism, the cyto‑
genes involved in OXPHOS by promoting JNK–PGC1α plasmic accumulation of these peptides induces a tran‑
signalling 70, and can promote mitochondria fuel switching scriptional response coordinated by ATFS‑1 (activating
by mediating complex II phosphorylation71 and meta‑ transcription factor associated with stress 1). ATFS‑1
bolic reprogramming through mitochondrial uncoupling possesses a mitochondrial targeting sequence as well as
and AMPK activation72. a nuclear localization signal. Under normal conditions, it
constantly shuttles to mitochondria, where it is degraded
Mitochondria fuel switching Mitonuclear feedback and proteostasis by the LON protease (LONP); however, in response to
Adaptations in mitochondrial A well-conserved protein quality control system, com‑ mitochondrial stress, ATFS‑1 import to mito­chondria
function and electron transport prising mainly chaperones and proteases encoded in is attenuated, causing it to localize in the nucleus
chain organization as a result of the nucleus, exists in mitochondria to maintain mito­ instead81. Here, ATFS‑1 and two other factors, DVE‑1
changes in the availability
of the fuel source.
chondrial proteostasis73. These proteins participate in and ubiquitin-­like 5 (UBL‑5), induce the expression of
the folding, assembly and turnover of mitochondrial several genes that are involved in mitochondrial quality
Mitochondrial uncoupling proteins under both normal and stress conditions. control and cellular metabolism to restore proteostasis.
The dissociation of Stresses that exceed the capacity of this protein quality These include the mitochondrial chaperones heat shock
mitochondrial respiration from
control system are sensed by mitochondria and com‑ protein 6 (hsp‑6), hsp‑60 and DNaJ domain 10 (dnj‑10);
ATP generation, characterized
by increased permeability of municated to the nucleus to promote the expression of the intermembrane ATPases associated with diverse
the inner mitochondrial these quality control components, as well as other com‑ cellu­lar activities (iAAA) protease ymel‑1; the mito‑
membrane to protons and pensatory genes that have been implicated in restoring chondrial fission factor drp‑1; glycolytic genes such as
subsequent dissipation of mito­chondrial homeostasis through bidirectional com‑ gpd‑2; detoxification genes such as the transcription fac‑
mitochondrial
membrane potential.
munication between the two organelles. Depending on tor skn‑1; and the core components of the translocase of
the nature of the stressor, three different mitonuclear the inner membrane 23 (TIM23) complex, tim‑23 and
Mitonuclear imbalance proteostasis responses have been described: the UPRmt, tim‑17 (REFS 80–84) (FIG. 4a). During stress, ATFS‑1 also
Mitonuclear imbalance occurs which activates the expression of proteases, chaper‑ limits the expression of other mitochondrial genes in the
when oxidative
ones and other stress response genes; proteolytic stress nucleus, such as those encoding TCA cycle enzymes and
phosphorylation subunits
encoded by both responses, which specifically induce the expression ETC subunits, by repressing their promoters84. To ensure
mitochondrial and nuclear of some mitochondrial proteases; and the heat shock that the expression of mitochondrial and nuclear ETC
DNA fail to assemble together response, which activates mitochondrial chaperones. subunits is coordinated under stress conditions, cer‑
in precise stoichiometric ratios tain splice variants of ATFS‑1 are specifically imported
to ensure proper mitochondrial
function, owing to an increase
The UPRmt. The UPRmt is a protective transcriptional into mitochondria during stress (through an unknown
or decrease of subunits from response that promotes the expression of mitochondrial mechanism), where they repress the expression of
one of the origins. proteostasis genes to stabilize mitochondrial function, mtDNA-encoded ETC subunits84 (FIG. 4a).

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 217


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

The UPRmt has also been described in D. melano‑ an internal deletion (ΔOTC), which causes mitochon‑
gaster, although it has been less extensively studied in drial protein overload, induces the expression of the
this model. Overexpression of a mutated form of the D. melanogaster orthologues of the chaperones HSP60
mitochondrial ornithine transcarbamylase containing and mtHSP70 (REF. 85). Mild mitochondrial distress in

a Caenorhabditis elegans
Mitochondrion Nucleus
? • OXPHOS
OXPHOS ?
ATFS-1 ? ATFS-1 • TCA genes
ATFS-1
mtDNA

? • Glycolysis genes
LONP ? ATFS-1
ATFS-1 • OXPHOS assembly

CLPP ? DVE-1
HAF-1 DVE-1 Mitochondrial
UBL-5 ATFS-1 UBL-5 chaperones
and proteases

b Mammals
UPRmt • CHOP
JNK2 JUN • CEBPβ

CEBPβ
CHOP
Mitochondrial
chaperones
and proteases
• HSP60 MURE CHOP MURE
• HSP10
• CLPP
? • Mitophagy
SIRT3 FOXO3A FOXO3A • ROS detoxification

Proteolytic stress responses ? • LONP


• AFG3l2
• SPG7
StAR StAR • YME1l1

Heat stress response HSF1 BRG1 Mitochondrial


SSBP1 SSBP1 SSBP1 HSF1 and cytosolic
chaperones

Figure 4 | Mitonuclear feedback. a | In worms, the CLPP-mediated cleavage of unfolded proteins in mitochondria initiates
Nature Reviews | Molecular Cell Biology
mitochondrial unfolded protein response (UPRmt) signalling. The efflux of short peptides through the mitochondrial
transporter HAF‑1 somehow inhibits mitochondrial protein import and, consequently, the transcription factor ATFS‑1
(activating transcription factor associated with stress 1), which is normally imported into mitochondria and degraded by the
LON protease (LONP), translocates into the nucleus. Here, together with ubiquitin-like 5 (UBL‑5) and DVE‑1, it induces
the expression of UPRmt target genes — namely, those encoding chaperones and proteases — to restore proteostasis.
ATFS‑1 can also positively regulate glycolysis and assembly of the oxidative phosphorylation (OXPHOS) system and
negatively regulate the expression of tricarboxylic acid (TCA) cycle and OXPHOS genes. In response to mitochondrial stress,
splice variants of ATFS‑1 accumulate in mitochondria and repress the expression of mitochondrial OXPHOS genes.
b | In mammals, the activation of JUN by c-Jun N-terminal kinase 2 (JNK2) leads to the induction of CEBP homologous protein
(CHOP) and CCAAT/enhancer-binding protein-β (CEBPβ), which heterodimerize and activate UPRmt genes. Mitochondrial
proteotoxic stress also activates sirtuin 3 (SIRT3), which deacetylates forkhead box transcription factor FOXO3A, causing its
translocation to the nucleus to induce the expression of reactive oxygen species (ROS) detoxification and mitophagy genes.
The import of steroidogenic acute regulatory protein (StAR) from the mitochondrial outer membrane into the matrix
induces, through an unknown mechanism, the expression of several proteases, which subsequently clear StAR from
mitochondria. Heat causes single-stranded DNA-binding protein 1 (SSBP1) to translocate into the nucleus, in association
with heat shock factor 1 (HSF1), to bind the chromatin modifier BRG1 and induce the expression of mitochondrial
and cytosolic chaperones. AFG3L2, AFG3‑like protein 2; HSP, heat shock protein; mtDNA, mitochondrial DNA;
MURE, mitochondrial unfolded response element; SPG7, spastic paraplegia 7; YME1L1, YME1-like protein 1.

218 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

muscle cells arising from the muscle-specific knockdown correlates with the expression of Cox5b/COX5B and
of Nd75 from complex I in D. melanogaster also induces Spg7/SPG7 — the mouse/human orthologues of worm
the UPRmt, together with a mitohormesis response64 cco‑1 and spg‑7 — suggesting that the low abundance of
(see below). these genes triggers the UPRmt in mammals as it does
Several lines of evidence suggest a certain degree of in C. elegans 99.
conservation of the existence of the UPRmt in vertebrates,
including in humans. The first line of evidence comes Proteolytic stress responses. In certain contexts, mito‑
from studies using cell lines. In monkey kidney COS7 chondrial proteotoxicity can lead to a specific proteolytic
cells, the mitochondrial protein overload generated by response and, in the case of the accumulation of unfolded
ΔOTC causes the activation of CHOP and CEBPβ86,87. proteins within the intermembrane space, genes other
Heterodimers of these transcription factors bind to con‑ than those involved in the canonical UPRmt are activated
served mitochondrial unfolded response elements and to resolve the stress. Overexpression of a mutant form
induce the expression of UPRmt genes, including those of endonuclease G reportedly causes stress in the inter­
encoding the chaperones HSP60, HSP10 and mtDNAJ; membrane space, leading to the production of ROS and
the proteases YME1L1, CLPP and PMPCB; the mito‑ phosphorylation of AKT (also known as protein kinase
chondrial import complex subunit TIM17A; and some B)100; consequently, the intermembrane space Ser protease
mitochondrial enzymes, such as thioredoxin 2 (REF. 88). HTRA2 (also known as OMI) and NRF1 are induced,
Knockdown of mitochondrial Leu-rich PPR motif-­ ­ultimately leading to increased proteasome activity 100.
containing protein, which impairs complex IV, acti‑ The steroidogenic acute regulatory protein (StAR)
vates the UPRmt by inducing mitonuclear imbalance in overload response (SOR) is another form of proteo‑
a neuroblastoma cell line and in worms89. Independent lytic stress and occurs in the context of the regulation
of CHOP, the mammalian mitochondrial deacetylase of steroido­genesis in the adrenal cortex and gonads101.
SIRT3 is also activated following mitochondrial proteo‑ StAR contributes to steroid synthesis by transferring
toxic stress; SIRT3 coordinates an antioxidant response cholesterol through the outer mitochondrial mem‑
and mitophagy 90 (FIG. 4b). Overexpression of CLPX, the brane to the inner mitochondrial membrane, where it is
regulatory subunit of complex CLPXP, is reportedly converted into steroids. StAR accumulation and activ‑
sufficient to induce the UPRmt in C2C12 myoblasts91. ity are prevented by its import into the mitochondrial
Inhibition of HSP90 chaperones in the mitochondria of matrix 102, which is followed by sequential degradation by
a human glioblastoma cell line triggers the UPRmt and ­mitochondrial proteases101 (FIG. 4b).
autophagy, potentially by inducing CHOP and CEBPβ
and repressing NF-κB92. Besides the induction of UPRmt, Heat shock response. Heat stress constitutes a threat to
over­expression of ΔOTC also leads to the activation of proteins in all cellular compartments103, but it can also
mitophagy, suggesting communication between the elicit a specific mitochondrial signal to the nucleus,
­different ­mitochondrial quality control systems93. causing the expression of the mitochondrial chaperones
Further evidence comes from several mouse models HSP10 and HSP60, in addition to regular heat shock pro‑
with mitochondrial dysfunction, in which the UPRmt is teins104. In cultured cells, heat causes the translocation of
induced, indicating that it seems to exert homeostatic the mitochondrial single-stranded DNA-binding protein 1
functions in mammals. For instance, deficiency of aspar‑ (SSBP1) in a complex with heat shock factor 1 (HSF1)
tyl tRNA synthetase in mouse skeletal and heart muscle into the nucleus, where, by recruiting the chromatin-­
causes the loss of mitochondrial proteostasis, which, in remodelling factor BRG1, it induces the expression of
the heart, activates the UPRmt and promotes mitochon‑ mitochondrial and cytoplasmic–nuclear chaperones that
drial biogenesis94. Cardiomyocyte-specific deletion of are essential for maintaining the mitochondrial membrane
the mitofusins in mice prevents mitochondrial fusion, potential and for survival following heat shock (FIG. 4b).
thereby impairing mitophagy and causing the accumu‑
Mitohormesis
lation of dysfunctional mitochondria and activation of Integrated stress response
A process in which low levels of
mitochondrial stress cause a the UPRmt (REF. 95). A mutation in the gene encoding the Mitochondrial stress can also activate the ISR, which is
protective cellular response mitochondrial helicase Twinkle, which causes multiple a general cellular response that modulates global pro‑
that reduces susceptibility to mtDNA deletions and late-onset mitochondrial disease tein synthesis (FIG. 5). This response is induced by dif‑
disease and potentially in mice, induces a progressive OXPHOS deficiency ferent stress stimuli, including the ER unfolded protein
promotes longevity.
accompanied by an increase in HSP60, mtHSP70 and response (UPRer), the UPRmt, oxidative stress, nutrient
Mitofusins CLPP protein levels in muscle96. The histone deacetylase deprivation, the presence of viral double-stranded RNA
High-molecular-mass GTPases SIRT7, which has emerged as a master regulator of or haem deficiencies105,106. The key component of this
that are essential for mitochondrial homeostasis97, can also promote, in con‑ stress response is the α-subunit of eukaryotic trans­lation
mitochondrial fusion.
junction with NRF1, the regenerative capacity of aged initiation factor 2 (eIF2α). Phosphorylation of eIF2α
Mouse BXD genetic haematopoietic stem cells through a mechanism that by several different kinases — including general con‑
reference population potentially involves the UPRmt (REF. 98). Finally, another trol non-derepressible 2 (GCN2), PKR-like ER kinase,
A family of recombinant inbred indication in support of conservation of the UPRmt in protein kinase double stranded RNA-dependent or
mouse strains, which is mammals comes from studies in the mouse BXD genetic haem-regulated inhibitor — globally inhibits protein
currently the largest and
reference population, in which the expression of the proto­ synthesis but concurrently facilitates the specific expres‑
best-characterized mouse
genetic reference population, typical UPRmt gene network, comprising mammalian sion of stress-response genes107 such as ATF4 (REF. 108).
composed of ~160 lines. orthologues of six worm UPRmt regulators, negatively ATF4, in turn, induces the expression of a wide range

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 219


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Mitochondrion Nucleus

Amino acid ER Haem deficiency,


starvation stress dsRNA heavy metals
ETC
• CHOP
ROS GCN2 PERK PKR HRI • GADD34
• ATF3
ATF4 • TRIB3
P
eIF2α eIF2α • BIP

Cell survival Apoptosis


Protein Translation of
translation ATF4

Cytosolic
mPOS, homeostasis
UPRam genes
MTS
UPRam

Proteasome

Figure 5 | The integrated stress response. The central regulation node of the integrated stress response (ISR) is the
eukaryotic translation initiation factor eIF2α, which, when phosphorylated, inhibitsNature Reviews
cytosolic | Molecular
translation. Cell Biology
Four different
kinases are known to phosphorylate eIF2α in response to various stresses: general control non-derepressible 2 (GCN2) is
activated by amino acid starvation; PKR-like ER‑kinase (PERK) is activated following endoplasmic reticulum (ER) stress;
protein kinase double stranded RNA-dependent (PKR) is activated by double-stranded RNA (dsRNA) following viral
infection; and haem-regulated inhibitor (HRI) is activated by heavy metals and haem deficiency. Defects in the electron
transport chain (ETC), reactive oxygen species (ROS) and mitochondrial proteotoxic stresses can activate GCN2, PERK or
HRI, depending on the context. Phosphorylation of eIF2α promotes the selective translation of the transcription factor ATF4,
which, in turn, promotes the expression of CEBP homologous protein (CHOP), growth arrest and DNA damage-inducible
protein 34 (GADD34), ATF3, tribbles homologue 3 (TRIB3) and immunoglobulin heavy chain-binding protein (BIP), as well as
other transcription factors, to restore cellular homeostasis. However, apoptosis can also ensue in the case of irreversible
cellular damage. In yeast, under conditions of mitochondrial stress, when protein import to mitochondria is slowed, newly
translated mitochondrial polypeptides accumulate in the cytosol, inducing mitochondrial precursor over-accumulation
stress (mPOS). This accumulation blocks translation and triggers an unfolded protein response activated by mis-targeting of
proteins (UPRam), which activates the proteasomal degradation pathway and induces the expression of gene sets that are
involved in restoring cellular homeostasis. MTS, mitochondrial targeting sequence.

of stress proteins, such as CHOP, growth arrest and motif protein113. Tetracyclines such as doxycycline,
DNA damage-­inducible protein 34 (GADD34), ATF3, which induce mitonuclear imbalance and mitochondrial
immuno­globulin heavy chain-binding protein (BIP; also dysfunction75,114, can also trigger the ATF4‑mediated
known as GRP78) or tribbles homologue 3 (TRIB3), to expression of stress genes 115. The accumulation of
restore proper c­ ellular function105,109,110. unfolded proteins in the inner mitochondrial membrane
The ISR and the phosphorylation of eIF2α as its pivot that is generated by the loss of HTRA2 in the brain also
are highly conserved. In C. elegans, mutations in several activates a CHOP-dependent ISR, protecting against
ETC genes, such as clk‑1 or isp‑1, induce ROS-mediated neuronal cell death116. In cultured cells, the outcome of
phosphorylation of eIF2α by the kinase GCN‑2, thereby the link between mitochondrial stress and the ISR can
reducing cytosolic translation and the generation of new be protective or deleterious, probably depending on the
mitochondrial proteins111. Activation of the ISR in this intensity and duration of the mitochondrial insults. In
context seems to be crucial, as loss of function of GCN2 certain cells, inhibition of some mitochondrial ETC
Mitokines affects mitochondrial activity and sensitizes cells to subunits activates an ATF4‑dependent ISR, promot‑
Signals released from a cell or
tissue in response to
mitochondrial stress, with detrimental consequences111. ing cell viability and metabolic reprogramming 117,118;
mitochondrial stress to In mammals, different mitochondrial stressors can however, the mitochondrial-dependent activation of
modulate cellular or activate the ISR. Arsenic-mediated mitochondrial the ISR is deleterious in other cell types, inhibiting cell
organismal homeostasis stress, through eIF2α phosphorylation, impairs the ­proliferation or altering lipid metabolism119,120.
and longevity.
function of the mitochondrial protein import complex Finally, beyond the prototypical ISR, other mech­
Ketogenesis and activates a gene expression programme for main‑ anisms that control and decrease global protein synthe‑
A metabolic pathway in the taining mitochondrial proteostasis in mammals, as sis can be activated under conditions of mitochondrial
liver that generates ketones well as in C. elegans 112. Altered or impaired expression stress, especially in yeast. In this organism, mito­
using acetyl-CoA, which of mtDNA in different mammalian cell types activates chondrial stress generated by loss of mitochondrial
provides an important fuel
source for the brain and other
the GCN2‑mediated ISR, triggering the expression of membrane potential or by defects in proteins of the mito‑
tissues during long-term several different stress proteins, such as CHOP, TRIB3, chondrial import machinery decreases mito­chondrial
fasting. ATF3 and hairy/enhancer-of‑split related with YRPW protein import, resulting in the aberrant cytosolic

220 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

accumulation of newly synthesized unfolded mitochon‑ Mitochondrial stress modulates lifespan


drial peptides121,122. This so‑called mitochondrial precur‑ Mitochondrial dysfunction constitutes a hallmark of
sor over-accumulation stress (mPOS) can not only lead and a driving force behind ageing 131, and mitochondrial
to mitochondria-mediated cell death, but can also miti­ insults underpin the development of several diseases132.
gate the stress by upregu­lating a large network of cyto‑ Conversely, subjecting mitochondria to specific stresses
solic genes to promote alternative pathways of protein extends lifespan across species in a manner dependent
synthesis and folding, as well as cell survival121 (FIG. 5). on mitonuclear communication. The most common
The mPOS response can also induce a UPR activated lifespan extender is attenuation of OXPHOS, which
by mis-targeting of proteins (UPRam), which protects the decreases respiration and concomitantly increases
cell against defects in mitochondrial import by ­inhibiting lifespan in different organisms (TABLE 1), although it is
protein synthesis and activating the proteasome122. often associated with slower development, reduced body
size and infertility 133,134. Several different, non-mutually
Stress and extracellular communication exclusive mechanisms have been proposed to explain
Mitochondrial stress signals not only activate different mitochondrial stress-dependent lifespan extension,
cellular responses but also trigger non-cell-autonomous including the retrograde stress response activated by
responses, implying that mitokines can be secreted from energy deficiency, ROS135 and Ca2+ signalling, as well as
cells with stressed mitochondria. In C. elegans, mito‑ by the UPRmt (REF. 76).
chondrial stress triggered by knockdown of ETC sub‑
units in a tissue-specific manner can be transduced to Energy regulation. In yeast, activation of the retro‑
distant tissues: when mitonuclear signalling was selec‑ grade response increases chronological lifespan, and
tively induced in neurons by neuron-specific knock‑ RTG2 and RTG3 are required to increase the chrono‑
down of cco‑1, the UPRmt was not only activated in the logical and replicative lifespan of respiration-deficient
nervous system but also in the intestine, with a beneficial strains136. Other retrograde pathways in yeast, such as
effect on the whole organism, including an increase in those involving TORC1 and transcription factor Sfp1
lifespan76. The origin of these inter-tissue cues is crucial pathways, and the activation of Sir2, also increase rep‑
for longevity, as the beneficial effects of UPRmt activa‑ licative lifespan, independently of RTGs35,36. Activation
tion are lost if the cues do not originate in the neurons76. of AMPK is required to extend the chronological
Similarly, in D. melanogaster, knockdown of ETC- lifespan of rho 0 cells 34. Consistent with its negative
encoding genes in neurons enhances longevity, implying effect on the retrograde response, disruption of TOR
the existence of systemic communication of an unknown signalling without a preceding mitochondrial stress
nature that affects the whole organism123. Likewise, mild also increases yeast replicative lifespan137. AMPK is also
ETC deficiency in fly muscles is communicated to other required for the increased longevity that is induced in
tissues through insulin signalling, which confirms the the worm isp‑1 and/or clk‑1 mutants37,138, as well as
non-cell-autonomous nature of the signal64. for the lifespan extension that is seen following loss of
In mammals, the mitochondria-mediated function of the worm p53 homologue cep‑1 (REF. 139) or
non-cell-autonomous response is thought to be regu­ putative homeobox protein 23 (REF. 140). Retrograde
lated either by mitochondria-derived peptides or by signalling through skn‑1 also governs longevity,
cellular hormones. Although different mitochondria-­ coordinat­ing the bimodal regulation of mitochondrial
derived secreted peptides have been described, further bio­genesis and mitophagy 141. Alterations in the TCA
mechanistic studies are needed to identify their exact cycle or supplementation with some of its intermedi‑
modes of action. Some mitochondria-derived peptides ates such as α‑ketoglutarate l­ ikewise increase longevity
are encoded by open reading frames identified in the in worms39,41 (TABLE 1).
mtDNA genome124. For example, the peptide humanin,
which is encoded by the 16S rRNA gene, is secreted from ROS signalling. Moderate levels of ROS are also c­ rucial
mitochondria and was reported to have cyto­protective transducers of mitochondrial stress that induce longev‑
properties in a model of Alzheimer disease125,126. Another ity in a range of organisms142. Following inhib­ition of
mitochondria-derived peptide is encoded within the respiration, ROS activates HIF‑1 (REF. 58) and increases
sequence encoding mitochondrial 12S rRNA and pro‑ lifespan in C. elegans 143. skn‑1 is also pivotal in trans­
motes AMPK activation in the muscle, thereby regu‑ ducing longevity cues that involve ROS signalling 60.
lating metabolic homeostasis and protecting against Low doses of toxic compounds that generate ROS,
insulin resistance and obesity 124. Fibroblast growth such as paraquat, arsenite or juglone, consistently
factor 21 (FGF21) is a fasting-regulated hormone with extend C. elegans lifespan61,144,145. This pro-longevity
pleiotropic actions in different tissues127. Interestingly, effect depends specifically on mitochondrial ROS, as
FGF21 functions as a potent mitokine and is secreted elevated levels of cytosolic ROS shorten lifespan146.
from skeletal muscle into the circulation in humans and ROS signalling is also required to induce longevity
mice with deficiencies in OXPHOS, in a manner medi‑ mediated by other types of mitochondrial communi‑
ated by the ATF4‑dependent ISR128–130. When it reaches cation: ROS are necessary for UPRmt cell-autonomous
the liver, FGF21 enhances ketogenesis and protects signalling in flies lacking ND75 in muscle cells and for
Rho0 cells
against diet-­induced obesity and insulin resistance by their increased longevity, whereas antioxidants reverse
Eukaryotic cells that lack enhancing β‑oxidation, lipolysis and browning of white longev­ity in some ETC worm mutants, such as clk‑1
mitochondrial DNA. adipose tissue130. and isp‑1 (REFS 64,135).

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 221


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Table 1 | Extended longevity triggered by mitochondrial stress


Function affected Organism Stressor Mediator or stress pathway Refs
Redox homeostasis Worm Paraquat, juglone or arsenite at low DAF‑16, SIR‑2.1, AAK‑2, ROS 61,144,
levels 145
TCA cycle Worm Malate or fumarate DAF‑16 39
α‑ketoglutarate TOR/LET‑363, AAK‑2, DAF‑16 41
Oxaloacetate AAK‑2, DAF‑16 168
ETC (complex I) Worm nuo‑2 depletion ROS 169
nuo‑6 mutation or depletion ROS 170
Fly Muscle-specific Nd75 depletion Non-cell-autonomous insulin 64
signalling
ETC (complex III) Worm cyc‑1 depletion ROS 169
isp‑1 mutation AAK‑2 133
isp‑1 depletion ROS 170
ETC (complex IV) Worm cco‑1 depletion UPRmt 76,169
cco‑1 intestinal- or neuronal-specific Non-cell-autonomous UPRmt 76
depletion
Mouse Surf1 knockout Calcium? 149
ETC (complex V) Worm atp‑3 depletion ROS 169
ETC (complexes I, II, Fly Depletion of complexes I, II, III, IV or V ? 123
III, IV or V)
Coenzyme Q Worm clk‑1 mutation ROS, HIF‑1, AAK‑2 58,134
biosynthesis
Fly Sbo (Coq2) depletion ? 171
Mouse Mclk1 heterozygosity ROS 147
Mitochondrial Worm CCCP ? 172
uncoupling
Mouse 2,4‑DNP ? 152
Mitochondrial Worm and mrps‑5 depletion or variation in Mrps5 UPRmt 75
translation mouse levels
Worm Doxycycline and chloramphenicol UPRmt 75
NAD levels
+
Worm Nicotinamide riboside SIR‑2.1, DAF‑16, UPR mt
77
Mitochondrial Fly Overexpression of Hsp22 ? 153
chaperones
Overexpression of Hsp60 in muscle ? 64
2,4‑DNP, 2,4‑dinitrophenylhydrazine; AAK‑2, AMP-activated kinase 2; atp‑3, ATP synthase subunit 3; CCCP, carbonyl cyanide m‑chloro-
phenylhydrazone; cco‑1, cytochrome c oxidase 1; cyc‑1, cytochrome c; ETC, electron transport chain; HIF‑1, hypoxia-inducible factor 1;
Hsp, heat shock protein; isp‑1, iron–sulfur protein 1; mrps‑5, mitochondrial ribosomal protein S5; nuo, NADH ubiquinone oxidoreductase;
ROS, reactive oxygen species; Surf1, surfeit gene 1; TCA, tricarboxylic acid; UPRmt, mitochondrial unfolded protein response.

Inactivation of some mouse orthologues that are The UPRmt. In worms, the longevity seen in loss-of-
mutated in long-lived worms and flies also increases function cco‑1 mutants depends on UBL‑5 and the
longevity (TABLE 1) through retrograde ROS or Ca2+ sig‑ induction of the UPRmt (REF. 76). Similarly, the extended
nalling. Mice heterozygous for a mutation in Mclk1 (also lifespan that is triggered by mitonuclear imbalance
known as Coq7), which encodes an enzyme involved induced by altered mitochondrial translation also
in coenzyme Q biosynthesis, have ETC defects and live depends on activation of the UPRmt (REF. 75). This mech‑
longer147,148. Similarly to flies, deficiency in the complex IV anism is evolutionarily conserved, as the expression of
assembly factor surfeit locus protein 1 (SURF1) also Mrps5 correlates inversely with longevity across the
increases longevity in mice149,150. The longevity induced mouse BXD genetic reference population, as it does
by mutations in Mclk1 and in Surf1 is thought to involve in C. elegans 75. C. elegans mutants of clk‑1 and isp‑1,
retrograde signals that depend on ROS and Ca2+, respec‑ which also display the ATFS‑1‑dependent UPRmt, have
tively 147,149; however, Surf1 mutants also show an increase an extended lifespan that is dependent on activation
in mitochondrial biogenesis, a tissue-specific induction of the ISR111. Of note, forced nuclear localization of
of the UPRmt and an antioxidant response mediated clk‑1, which functions as a suppressor of UPRmt gene
by NFE2L2 (REF. 151). Conversely, mild mitochondrial expression, abrogates the longevity of clk‑1 mutants62.
uncoupling, which enhances mitochondrial respiration UPRmt genes are activated in flies with extended lifespan
and decreases ROS production, also increases the lifespan that lack Nd75 specifically in muscle cells, and forced
of mice by modulating their metabolic conditions152. expression of Hsp60 in muscles is sufficient to extend

222 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Box 1 | Mitonuclear stress signalling and the immune response and powerful mechanism that cooperates with these
response mechanisms under normal and stress condi‑
Given their crucial functions in energy metabolism, mitochondria are the Achilles heel tions. Mitochondrial defects, especially decreases in ATP
of cellular activities, constituting a prime target for attacks by pathogens and toxins production, trigger a pleiotropic retrograde response
that target mitochondrial activity and integrity159. Furthermore, the proto-bacterial that has the main effect of activating alternative path‑
origins of mitochondria makes it tempting to speculate that cellular defence pathways
ways of energy generation and restoring mitochondrial
against pathogens have similarities with those that defend against mitochondrial
stress: when some molecules, such as mitochondrial DNA (mtDNA) and formyl peptides, homeostasis. Retrograde signalling to the nucleus and
are released from damaged mitochondria, the immune system can be activated as mitochondrial feedback occur in all eukaryotes; how‑
it is during bacterial infection. The mitochondrial unfolded protein response ever, the complexity of their regulation has evolved in
(UPRmt) is involved in both mitochondrial and immune surveillance, to resolve parallel with the complexity of organisms. Thus, a large
mitochondrial stress and infection alike. Several bacterial species encountered by number of components and pathways — which may be
Caenorhabditis elegans induce mitochondrial dysfunction and consequently activate either cell-autonomous or non-cell-autonomous, and
the UPRmt as a protective response160. Mutants in the UPRmt genes ubl‑5 tissue-specific — are associated with mitonuclear com‑
(ubiquitin-like 5), dve‑1 and atfs‑1 (activating transcription factor associated with munication in mammals. This plethora of mechanisms
stress 1) display weakened immunity and reduced survival against bacteria that are has hampered the study of mitonuclear communica‑
commonly pathogenic for C. elegans138,161. Following exposure to Pseudomonas
tion regulation and our understanding of the physio­
aeruginosa, ATFS‑1 induces the expression of genes encoding components of the
innate immunity response, such as lysozyme and antimicrobial peptides161. Immunity is logical and pathological consequences of the activation
also intimately linked to signalling by reactive oxygen species (ROS), as ROS activate of mitonuclear signalling. Nevertheless, mitonuclear
aak‑2 (AMP-activated kinase 2) and hif‑1 (hypoxia-inducible factor 1), leading to communication seems to be essential for maintaining
increased resistance of C. elegans to pathogenic strains of Escherichia coli138. certain physiological functions, such as the immune
In mammals, mitochondrial stress also elicits an enhanced immune response. response (BOX 1). Future studies using loss‑of‑function
Moderate mitochondrial DNA stress induced by loss of function of TFAM (transcription and gain‑of‑function in vivo models, as well as the analy‑
factor A, mitochondrial) activates a cytosolic antiviral signal that enhances the sis of different genetic populations155, including humans,
expression of a subset of interferon-responsive genes162. Mitochondria are also involved should help us to understand this complex regulation
in antiviral defence through the mitochondria-associated viral sensor, a mitochondrial in mammals.
outer membrane protein that activates an antiviral signalling response159. Mitochondrial-
The study of mitonuclear communication has
generated ROS contribute to the bactericidal activity of macrophages by activating cell
surface Toll-like receptors163. ROS also enhance immunity of mice heterozygous for a received great interest since the discovery that activa‑
mutation in the gene encoding mitochondrial 5‑demethoxyubiquinone hydroxylase tion of some stress response pathways promotes longev‑
— a protein involved in ubiquinone biosynthesis — against Salmonella enterica and ity. Non-deleterious defects in components of the ETC
tumour cell grafts, through the activation of HIF1α164,165. Cellular protection against activate a paradoxical beneficial response to promote
intracellular pathogens such as Mycobacterium spp. or S. enterica is furthermore health and extend lifespan, thereby strengthening and
dependent on mechanisms similar to mitophagy (the removal of defective expanding the mitohormesis theory 156. Additionally,
mitochondria)166. Conversely, extreme mitochondrial stress caused by TFAM deficiency the fact that this stress response can be regulated in a
in T cells causes lysosomal defects and an increased inflammatory response, which are non-cell-autonomous manner by mitokines opens up
recovered by restoring NAD+ levels and mitochondrial function167. a new field of investigation around these molecules,
which could function as pharmacological targets for
managing mitochondrial and ageing-related diseases.
lifespan64. Similarly, overexpression in D. melanogaster Although some hormones and peptides have been
of the chaperone Hsp22, which is induced upon oxi‑ described as potential mitokines, we expect that sev‑
dative stress, is sufficient to increase lifespan, which eral other molecules may function as mitokines. As
suggests a close ­connection between longevity and mitochondria regulate many cellular metabolic path‑
UPRmt signalling 153. ways, some intermediate metabolites could function
as mediators for both cell-autonomous and non-cell-­
Conclusion and future perspectives autonomous responses. Long non-coding RNAs,
Coordination between mitochondria and the nucleus microRNAs or different kinds of small RNAs could
is essential for maintaining cellular homeostasis and also mediate mitochondrial-mediated stress responses.
ensuring an effective response to stress. Mitochondrial Of note, some long non-coding RNAs, which can be
function is regulated by the nucleus through several exported to the cytosol to regulate protein synthesis,
transcription factors that activate the expression of the have been identified in mtDNA157. Similarly, mito‑
mitochondrial genome; through the transcriptional chondria-derived ­vesicles could be another potential
induction of several mitochondrial proteins encoded by ­communication signal158.
nuclear DNA; and by promoting mitochondrial adaptive In this Review, we have described the different path‑
pathways. Similarly, mitochondria can regulate nuclear ways that are governed by mitonuclear crosstalk under
function by controlling the levels of various metabo­ basal and stress conditions, providing an overview of
lites that affect gene expression, such as ATP, acetyl- these complex networks in various organisms. Although
CoA, NAD+ and ROS154. Mitochondria maintain their several evolutionarily conserved mechanisms of stress
Formyl peptides homeo­stasis by orchestrating different quality control response exist, future studies will be required to unravel
Small peptides in mitochondria mechanisms, including protein quality control pathways, the regulation and the molecular mechanisms behind
and bacteria with a formylated
amino-terminal Met and,
mitochondrial dynamics, mitophagy, mitochondrial these stress responses, as well as their physiological and
usually, a hydrophobic amino biogenesis and apoptosis1,73; however, crosstalk between pathological relevance, which may one day help us to
acid at the carboxy-terminal. mitochondria and the nucleus functions as a parallel design potential therapies for pervasive human diseases.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 223


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

1. Friedman, J. R. & Nunnari, J. Mitochondrial form 26. Sahin, E. et al. Telomere dysfunction induces metabolic 49. Biswas, G., Anandatheerthavarada, H. K., Zaidi, M.
and function. Nature 505, 335–343 (2014). and mitochondrial compromise. Nature 470, & Avadhani, N. G. Mitochondria to nucleus stress
2. Wallace, D. C. Mitochondria, bioenergetics, and 359–365 (2011). signaling: a distinctive mechanism of NFκB/Rel
the epigenome in eukaryotic and human evolution. A description of how telomere dysfunction activation through calcineurin-mediated inactivation
Cold Spring Harb. Symp. Quant. Biol. 74, 383–393 can initiate reprogramming of mitochondrial of IκBβ. J. Cell Biol. 161, 507–519 (2003).
(2009). regulation. 50. Biswas, G. et al. Retrograde Ca2+ signaling in C2C12
3. Pagliarini, D. J. et al. A mitochondrial protein 27. Luo, X. & Kraus, W. L. On PAR with PARP: cellular skeletal myocytes in response to mitochondrial genetic
compendium elucidates complex I disease biology. stress signaling through poly(ADP-ribose) and and metabolic stress: a novel mode of inter-organelle
Cell 134, 112–123 (2008). PARP‑1. Genes Dev. 26, 417–432 (2012). crosstalk. EMBO J. 18, 522–533 (1999).
4. Mercer, T. R. et al. The human mitochondrial 28. Bai, P. et al. PARP‑1 inhibition increases 51. Formentini, L., Sanchez-Arago, M., Sanchez-Cenizo, L.
transcriptome. Cell 146, 645–658 (2011). mitochondrial metabolism through SIRT1 activation. & Cuezva, J. M. The mitochondrial ATPase inhibitory
5. Dhar, S. S., Ongwijitwat, S. & Wong-Riley, M. T. Cell Metab. 13, 461–468 (2011). factor 1 triggers a ROS-mediated retrograde
Nuclear respiratory factor 1 regulates all ten nuclear- 29. Canto, C. et al. The NAD+ precursor nicotinamide prosurvival and proliferative response. Mol. Cell 45,
encoded subunits of cytochrome c oxidase in neurons. riboside enhances oxidative metabolism and protects 731–742 (2012).
J. Biol. Chem. 283, 3120–3129 (2008). against high-fat diet-induced obesity. Cell Metab. 15, 52. Amuthan, G. et al. Mitochondria-to‑nucleus stress
6. Virbasius, J. V. & Scarpulla, R. C. Activation of the 838–847 (2012). signaling induces phenotypic changes, tumor
human mitochondrial transcription factor A gene by 30. Jazwinski, S. M. The retrograde response: when progression and cell invasion. EMBO J. 20,
nuclear respiratory factors: a potential regulatory link mitochondrial quality control is not enough. 1910–1920 (2001).
between nuclear and mitochondrial gene expression Biochim. Biophys. Acta 1833, 400–409 (2013). 53. Lim, J. H. et al. Mitochondrial dysfunction induces
in organelle biogenesis. Proc. Natl Acad. Sci. USA 91, 31. Liu, Z. & Butow, R. A. Mitochondrial retrograde aberrant insulin signalling and glucose utilisation in
1309–1313 (1994). signaling. Annu. Rev. Genet. 40, 159–185 (2006). murine C2C12 myotube cells. Diabetologia 49,
7. Gleyzer, N., Vercauteren, K. & Scarpulla, R. C. Control 32. Sekito, T., Thornton, J. & Butow, R. A. Mitochondria- 1924–1936 (2006).
of mitochondrial transcription specificity factors to‑nuclear signaling is regulated by the subcellular 54. Guha, M., Fang, J. K., Monks, R., Birnbaum, M. J.
(TFB1M and TFB2M) by nuclear respiratory factors localization of the transcription factors Rtg1p and & Avadhani, N. G. Activation of Akt is essential for
(NRF‑1 and NRF‑2) and PGC‑1 family coactivators. Rtg3p. Mol. Biol. Cell 11, 2103–2115 (2000). the propagation of mitochondrial respiratory stress
Mol. Cell. Biol. 25, 1354–1366 (2005). 33. Jazwinski, S. M. & Kriete, A. The yeast retrograde signaling and activation of the transcriptional
8. Blesa, J. R., Prieto-Ruiz, J. A., Hernandez, J. M. response as a model of intracellular signaling of coactivator heterogeneous ribonucleoprotein A2.
& Hernandez-Yago, J. NRF‑2 transcription factor mitochondrial dysfunction. Front. Physiol. 3, 139 Mol. Biol. Cell 21, 3578–3589 (2010).
is required for human TOMM20 gene expression. (2012). 55. Schieber, M. & Chandel, N. S. ROS function in redox
Gene 391, 198–208 (2007). 34. Friis, R. M. et al. Rewiring AMPK and mitochondrial signaling and oxidative stress. Curr. Biol. 24,
9. Peralta, S., Wang, X. & Moraes, C. T. Mitochondrial retrograde signaling for metabolic control of aging R453–R462 (2014).
transcription: lessons from mouse models. and histone acetylation in respiratory-defective cells. 56. Shadel, G. S. & Horvath, T. L. Mitochondrial ROS
Biochim. Biophys. Acta 1819, 961–969 (2012). Cell Rep. 7, 565–574 (2014). signaling in organismal homeostasis. Cell 163,
10. Richter-Dennerlein, R., Dennerlein, S. & Rehling, P. 35. Heeren, G. et al. The mitochondrial ribosomal protein 560–569 (2015).
Integrating mitochondrial translation into the cellular of the large subunit, Afo1p, determines cellular 57. Miyadera, H. et al. Altered quinone biosynthesis in the
context. Nat. Rev. Mol. Cell Biol. 16, 586–592 longevity through mitochondrial back-signaling long-lived clk‑1 mutants of Caenorhabditis elegans.
(2015). via TOR1. Aging 1, 622–636 (2009). J. Biol. Chem. 276, 7713–7716 (2001).
11. Wang, Y. X. et al. Peroxisome-proliferator-activated 36. Caballero, A. et al. Absence of mitochondrial 58. Lee, S. J., Hwang, A. B. & Kenyon, C. Inhibition of
receptor δ activates fat metabolism to prevent obesity. translation control proteins extends life span by respiration extends C. elegans life span via reactive
Cell 113, 159–170 (2003). activating sirtuin-dependent silencing. Mol. Cell 42, oxygen species that increase HIF‑1 activity. Curr. Biol.
12. Tanaka, T. et al. Activation of peroxisome proliferator- 390–400 (2011). 20, 2131–2136 (2010).
activated receptor δ induces fatty acid β-oxidation in References 34–36 give examples of energetic Identification of a signalling mechanism involving
skeletal muscle and attenuates metabolic syndrome. retrograde signals in yeast involving the Ampk, ROS in lifespan extension following mild
Proc. Natl Acad. Sci. USA 100, 15924–15929 TOR and Sir2 pathways. respiration inhibition.
(2003). 37. Curtis, R., O’Connor, G. & DiStefano, P. S. Aging 59. Inoue, H. et al. The C. elegans p38 MAPK pathway
13. Michalik, L. et al. International Union of networks in Caenorhabditis elegans: AMP-activated regulates nuclear localization of the transcription
Pharmacology. LXI. Peroxisome proliferator-activated protein kinase (aak‑2) links multiple aging and factor SKN‑1 in oxidative stress response. Genes Dev.
receptors. Pharmacol. Rev. 58, 726–741 (2006). metabolism pathways. Aging Cell 5, 119–126 (2006). 19, 2278–2283 (2005).
14. Dufour, C. R. et al. Genome-wide orchestration of 38. Apfeld, J., O’Connor, G., McDonagh, T., 60. Zarse, K. et al. Impaired insulin/IGF1 signaling extends
cardiac functions by the orphan nuclear receptors DiStefano, P. S. & Curtis, R. The AMP-activated life span by promoting mitochondrial l‑proline
ERRα and γ. Cell Metab. 5, 345–356 (2007). protein kinase AAK‑2 links energy levels and insulin- catabolism to induce a transient ROS signal.
15. Alaynick, W. A. et al. ERRγ directs and maintains the like signals to lifespan in C. elegans. Genes Dev. 18, Cell Metab. 15, 451–465 (2012).
transition to oxidative metabolism in the postnatal 3004–3009 (2004). 61. Schmeisser, S. et al. Mitochondrial hormesis links
heart. Cell Metab. 6, 13–24 (2007). 39. Edwards, C. B., Copes, N., Brito, A. G., Canfield, J. low‑dose arsenite exposure to lifespan extension.
16. Mouchiroud, L., Eichner, L. J., Shaw, R. J. & Auwerx, J. & Bradshaw, P. C. Malate and fumarate extend Aging Cell 12, 508–517 (2013).
Transcriptional coregulators: fine-tuning metabolism. lifespan in Caenorhabditis elegans. PLoS ONE 8, 62. Monaghan, R. M. et al. A nuclear role for the
Cell Metab. 20, 26–40 (2014). e58345 (2013). respiratory enzyme CLK‑1 in regulating mitochondrial
17. Fernandez-Marcos, P. J. & Auwerx, J. Regulation of 40. Gallo, M., Park, D. & Riddle, D. L. Increased longevity stress responses and longevity. Nat. Cell Biol. 17,
PGC‑1α, a nodal regulator of mitochondrial of some C. elegans mitochondrial mutants explained 782–792 (2015).
biogenesis. Am. J. Clin. Nutr. 93, 884S–890S by activation of an alternative energy-producing Identification of a novel communication mechanism
(2011). pathway. Mech. Ageing Dev. 132, 515–518 (2011). involving the nuclear translocation of an isoform
18. Handschin, C. & Spiegelman, B. M. Peroxisome 41. Chin, R. M. et al. The metabolite α-ketoglutarate of CLK‑1, which regulates stress resistance
proliferator-activated receptor γ coactivator 1 extends lifespan by inhibiting ATP synthase and TOR. and lifespan.
coactivators, energy homeostasis, and metabolism. Nature 510, 397–401 (2014). 63. Owusu-Ansah, E., Yavari, A., Mandal, S.
Endocr. Rev. 27, 728–735 (2006). 42. Egan, D. F. et al. Phosphorylation of ULK1 (hATG1) & Banerjee, U. Distinct mitochondrial retrograde
19. Yamamoto, H. et al. NCoR1 is a conserved by AMP-activated protein kinase connects energy signals control the G1–S cell cycle checkpoint.
physiological modulator of muscle mass and oxidative sensing to mitophagy. Science 331, 456–461 (2011). Nat. Genet. 40, 356–361 (2008).
function. Cell 147, 827–839 (2011). 43. Lerner, C. et al. Reduced mammalian target of 64. Owusu-Ansah, E., Song, W. & Perrimon, N. Muscle
20. Seth, A. et al. The transcriptional corepressor RIP140 rapamycin activity facilitates mitochondrial retrograde mitohormesis promotes longevity via systemic
regulates oxidative metabolism in skeletal muscle. signaling and increases life span in normal human repression of insulin signaling. Cell 155, 699–712
Cell Metab. 6, 236–245 (2007). fibroblasts. Aging Cell 12, 966–977 (2013). (2013).
21. Canto, C. et al. AMPK regulates energy expenditure 44. Rizzuto, R., De Stefani, D., Raffaello, A. A description of non-cell-autonomous-mediated
by modulating NAD+ metabolism and SIRT1 activity. & Mammucari, C. Mitochondria as sensors and lifespan extension through ROS, the UPRmt and
Nature 458, 1056–1060 (2009). regulators of calcium signalling. Nat. Rev. Mol. insulin signalling following mitochondrial
22. Canto, C. et al. Interdependence of AMPK and SIRT1 Cell Biol. 13, 566–578 (2012). perturbation in D. melanogaster muscles.
for metabolic adaptation to fasting and exercise in 45. Arnould, T. et al. CREB activation induced by 65. Lu, W. et al. ZNF143 transcription factor mediates cell
skeletal muscle. Cell Metab. 11, 213–219 (2010). mitochondrial dysfunction is a new signaling pathway survival through upregulation of the GPX1 activity
23. Garcia-Roves, P. M., Osler, M. E., Holmstrom, M. H. that impairs cell proliferation. EMBO J. 21, 53–63 in the mitochondrial respiratory dysfunction.
& Zierath, J. R. Gain‑of‑function R225Q mutation in (2002). Cell Death Dis. 3, e422 (2012).
AMP-activated protein kinase γ3 subunit increases 46. Luo, Y., Bond, J. D. & Ingram, V. M. Compromised 66. Chen, X. L. & Kunsch, C. Induction of cytoprotective
mitochondrial biogenesis in glycolytic skeletal muscle. mitochondrial function leads to increased cytosolic genes through Nrf2/antioxidant response element
J. Biol. Chem. 283, 35724–35734 (2008). calcium and to activation of MAP kinases. Proc. Natl pathway: a new therapeutic approach for the
24. Wu, H. et al. Regulation of mitochondrial biogenesis Acad. Sci. USA 94, 9705–9710 (1997). treatment of inflammatory diseases. Curr. Pharm. Des.
in skeletal muscle by CaMK. Science 296, 349–352 47. Amuthan, G. et al. Mitochondrial stress-induced 10, 879–891 (2004).
(2002). calcium signaling, phenotypic changes and invasive 67. Kops, G. J. et al. Forkhead transcription factor
25. Woods, A. et al. Ca2+/calmodulin-dependent protein behavior in human lung carcinoma A549 cells. FOXO3a protects quiescent cells from oxidative stress.
kinase kinase-β acts upstream of AMP-activated Oncogene 21, 7839–7849 (2002). Nature 419, 316–321 (2002).
protein kinase in mammalian cells. Cell Metab. 2, 48. Srinivasan, S. et al. Disruption of cytochrome c 68. Tan, W. Q., Wang, K., Lv, D. Y. & Li, P. F. Foxo3a
21–33 (2005). oxidase function induces the Warburg effect and inhibits cardiomyocyte hypertrophy through
References 21–25 give examples of how different metabolic reprogramming. Oncogene http:// transactivating catalase. J. Biol. Chem. 283,
kinases signal to induce mitochondrial biogenesis. dx.doi.org/10.1038/onc.2015.227 (2015). 29730–29739 (2008).

224 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

69. Nguyen, T., Nioi, P. & Pickett, C. B. The 89. Kohler, F., Muller-Rischart, A. K., Conradt, B. triggers the activation of CHOP‑10 by a cell signaling
Nrf2‑antioxidant response element signaling pathway & Rolland, S. G. The loss of LRPPRC function induces dependent on the integrated stress response but not
and its activation by oxidative stress. J. Biol. Chem. the mitochondrial unfolded protein response. Aging 7, the mitochondrial unfolded protein response.
284, 13291–13295 (2009). 701–717 (2015). Mitochondrion 21, 58–68 (2015).
70. Chae, S. et al. A systems approach for decoding 90. Papa, L. & Germain, D. SirT3 regulates the 114. Moullan, N. et al. Tetracyclines disturb mitochondrial
mitochondrial retrograde signaling pathways. mitochondrial unfolded protein response. Mol. Cell. function across eukaryotic models: a call for caution
Sci. Signal. 6, rs4 (2013). Biol. 34, 699–710 (2014). in biomedical research. Cell Rep. 10, 1681–1691
71. Acin-Perez, R. et al. ROS-triggered phosphorylation 91. Al‑Furoukh, N. et al. ClpX stimulates the mitochondrial (2015).
of complex II by Fgr kinase regulates cellular unfolded protein response (UPRmt) in mammalian cells. 115. Bruning, A., Brem, G. J., Vogel, M. & Mylonas, I.
adaptation to fuel use. Cell Metab. 19, 1020–1033 Biochim. Biophys. Acta 1853, 2580–2591 (2015). Tetracyclines cause cell stress-dependent ATF4
(2014). 92. Siegelin, M. D. et al. Exploiting the mitochondrial activation and mTOR inhibition. Exp. Cell Res. 320,
72. Shi, S. Y. et al. DJ‑1 links muscle ROS production unfolded protein response for cancer therapy in mice 281–289 (2014).
with metabolic reprogramming and systemic energy and human cells. J. Clin. Invest. 121, 1349–1360 116. Moisoi, N. et al. Mitochondrial dysfunction triggered
homeostasis in mice. Nat. Commun. 6, 7415 (2015). (2011). by loss of HtrA2 results in the activation of a brain-
73. Quiros, P. M., Langer, T. & Lopez-Otin, C. New roles 93. Jin, S. M. & Youle, R. J. The accumulation of misfolded specific transcriptional stress response. Cell Death
for mitochondrial proteases in health, ageing and proteins in the mitochondrial matrix is sensed by Differ. 16, 449–464 (2009).
disease. Nat. Rev. Mol. Cell Biol. 16, 345–359 PINK1 to induce PARK2/Parkin-mediated mitophagy 117. Evstafieva, A. G. et al. A sustained deficiency of
(2015). of polarized mitochondria. Autophagy 9, 1750–1757 mitochondrial respiratory complex III induces an
74. Jovaisaite, V., Mouchiroud, L. & Auwerx, J. (2013). apoptotic cell death through the p53‑mediated
The mitochondrial unfolded protein response, 94. Dogan, S. A. et al. Tissue-specific loss of DARS2 inhibition of pro-survival activities of the activating
a conserved stress response pathway with activates stress responses independently of transcription factor 4. Cell Death Dis. 5, e1511
implications in health and disease. J. Exp. Biol. 217, respiratory chain deficiency in the heart. Cell Metab. (2014).
137–143 (2014). 19, 458–469 (2014). 118. Martinez-Reyes, I., Sanchez-Arago, M. & Cuezva, J. M.
75. Houtkooper, R. H. et al. Mitonuclear protein 95. Song, M., Mihara, K., Chen, Y., Scorrano, L. AMPK and GCN2–ATF4 signal the repression of
imbalance as a conserved longevity mechanism. & Dorn, G. W. 2nd. Mitochondrial fission and fusion mitochondria in colon cancer cells. Biochem. J. 444,
Nature 497, 451–457 (2013). factors reciprocally orchestrate mitophagic culling in 249–259 (2012).
Establishment of the concept of mitonuclear mouse hearts and cultured fibroblasts. Cell Metab. 21, 119. Silva, J. M., Wong, A., Carelli, V. & Cortopassi, G. A.
imbalance to explain the identification of Mrps5 as 273–285 (2015). Inhibition of mitochondrial function induces an
a mouse longevity gene and the pro-longevity effect 96. Khan, N. A. et al. Effective treatment of mitochondrial integrated stress response in oligodendroglia.
of mitochondrial translation inhibition. myopathy by nicotinamide riboside, a vitamin B3. Neurobiol. Dis. 34, 357–365 (2009).
76. Durieux, J., Wolff, S. & Dillin, A. The cell-non- EMBO Mol. Med. 6, 721–731 (2014). 120. Viader, A. et al. Aberrant Schwann cell lipid
autonomous nature of electron transport 97. Ryu, D. et al. A SIRT7‑dependent acetylation switch of metabolism linked to mitochondrial deficits leads
chain‑mediated longevity. Cell 144, 79–91 (2011). GABPβ1 controls mitochondrial function. Cell Metab. to axon degeneration and neuropathy. Neuron 77,
The first description of non-cell-autonomous- 20, 856–869 (2014). 886–898 (2013).
mediated longevity following mitochondrial stress. 98. Mohrin, M. et al. A mitochondrial UPR-mediated 121. Wang, X. & Chen, X. J. A cytosolic network
77. Mouchiroud, L. et al. The NAD+/sirtuin pathway metabolic checkpoint regulates hematopoietic stem suppressing mitochondria-mediated proteostatic
modulates longevity through activation of cell aging. Science 347, 1374–1377 (2015). stress and cell death. Nature 524, 481–484 (2015).
mitochondrial UPR and FOXO signaling. Cell 154, 99. Wu, Y. et al. Multilayered genetic and omics dissection 122. Wrobel, L. et al. Mistargeted mitochondrial proteins
430–441 (2013). of mitochondrial activity in a mouse reference activate a proteostatic response in the cytosol. Nature
78. Pirinen, E. et al. Pharmacological inhibition of population. Cell 158, 1415–1430 (2014). 524, 485–488 (2015).
poly(ADP-ribose) polymerases improves fitness and 100. Papa, L. & Germain, D. Estrogen receptor mediates References 121 and 122 identify a cytosolic stress
mitochondrial function in skeletal muscle. Cell Metab. a distinct mitochondrial unfolded protein response. response activated by mitochondrial stress.
19, 1034–1041 (2014). J. Cell Sci. 124, 1396–1402 (2011). 123. Copeland, J. M. et al. Extension of Drosophila life
79. Gariani, K. et al. Eliciting the mitochondrial unfolded 101. Bahat, A. et al. Transcriptional activation of LON gene span by RNAi of the mitochondrial respiratory chain.
protein response via NAD repletion reverses fatty liver by a new form of mitochondrial stress: a role for the Curr. Biol. 19, 1591–1598 (2009).
disease. Hepatology http://dx.doi.org/10.1002/ nuclear respiratory factor 2 in StAR overload response 124. Lee, C. et al. The mitochondrial-derived peptide
hep.28245 (2015). (SOR). Mol. Cell Endocrinol. 408, 62–72 (2015). MOTS‑c promotes metabolic homeostasis and reduces
References 21, 22, 28–29 and 77–79 describe 102. Bahat, A. et al. StAR enhances transcription of genes obesity and insulin resistance. Cell Metab. 21,
mechanisms of mitochondrial regulation involving encoding the mitochondrial proteases involved in its 443–454 (2015).
NAD+ and SIRT1. own degradation. Mol. Endocrinol. 28, 208–224 125. Lee, C., Yen, K. & Cohen, P. Humanin: a harbinger of
80. Haynes, C. M., Yang, Y., Blais, S. P., Neubert, T. A. (2014). mitochondrial-derived peptides? Trends Endocrinol.
& Ron, D. The matrix peptide exporter HAF‑1 signals 103. Akerfelt, M., Morimoto, R. I. & Sistonen, L. Heat shock Metab. 24, 222–228 (2013).
a mitochondrial UPR by activating the transcription factors: integrators of cell stress, development and 126. Hashimoto, Y. et al. A rescue factor abolishing
factor ZC376.7 in C. elegans. Mol. Cell 37, 529–540 lifespan. Nat. Rev. Mol. Cell Biol. 11, 545–555 neuronal cell death by a wide spectrum of familial
(2010). (2010). Alzheimer’s disease genes and Aβ. Proc. Natl Acad.
81. Nargund, A. M., Pellegrino, M. W., Fiorese, C. J., 104. Tan, K. et al. Mitochondrial SSBP1 protects cells from Sci. USA 98, 6336–6341 (2001).
Baker, B. M. & Haynes, C. M. Mitochondrial import proteotoxic stresses by potentiating stress-induced 127. Chau, M. D., Gao, J., Yang, Q., Wu, Z. & Gromada, J.
efficiency of ATFS‑1 regulates mitochondrial UPR HSF1 transcriptional activity. Nat. Commun. 6, 6580 Fibroblast growth factor 21 regulates energy
activation. Science 337, 587–590 (2012). (2015). metabolism by activating the AMPK–SIRT1–PGC‑1α
82. Benedetti, C., Haynes, C. M., Yang, Y., Harding, H. P. 105. Harding, H. P. et al. Regulated translation initiation pathway. Proc. Natl Acad. Sci. USA 107,
& Ron, D. Ubiquitin-like protein 5 positively regulates controls stress-induced gene expression in mammalian 12553–12558 (2010).
chaperone gene expression in the mitochondrial cells. Mol. Cell 6, 1099–1108 (2000). 128. Tyynismaa, H. et al. Mitochondrial myopathy induces
unfolded protein response. Genetics 174, 229–239 106. Harding, H. P. et al. An integrated stress response a starvation-like response. Hum. Mol. Genet. 19,
(2006). regulates amino acid metabolism and resistance to 3948–3958 (2010).
83. Haynes, C. M., Petrova, K., Benedetti, C., Yang, Y. oxidative stress. Mol. Cell 11, 619–633 (2003). 129. Suomalainen, A. et al. FGF‑21 as a biomarker for
& Ron, D. ClpP mediates activation of a mitochondrial 107. Donnelly, N., Gorman, A. M., Gupta, S. & Samali, A. muscle-manifesting mitochondrial respiratory chain
unfolded protein response in C. elegans. Dev. Cell 13, The eIF2α kinases: their structures and functions. deficiencies: a diagnostic study. Lancet Neurol. 10,
467–480 (2007). Cell. Mol. Life Sci. 70, 3493–3511 (2013). 806–818 (2011).
84. Nargund, A. M., Fiorese, C. J., Pellegrino, M. W., 108. Palam, L. R., Baird, T. D. & Wek, R. C. Phosphorylation 130. Kim, K. H. et al. Autophagy deficiency leads to
Deng, P. & Haynes, C. M. Mitochondrial and nuclear of eIF2 facilitates ribosomal bypass of an inhibitory protection from obesity and insulin resistance by
accumulation of the transcription factor ATFS‑1 upstream ORF to enhance CHOP translation. J. Biol. inducing Fgf21 as a mitokine. Nat. Med. 19, 83–92
promotes OXPHOS recovery during the UPRmt. Chem. 286, 10939–10949 (2011). (2013).
Mol. Cell 58, 123–133 (2015). 109. Novoa, I., Zeng, H., Harding, H. P. & Ron, D. References 124 and 128–130 report examples
References 80–84 unravel the molecular Feedback inhibition of the unfolded protein response of non-cell-autonomous signalling of mitochondrial
mechanisms of UPRmt regulation in worms. by GADD34‑mediated dephosphorylation of eIF2α. stress in mammals.
85. Pimenta de Castro, I. et al. Genetic analysis of J. Cell Biol. 153, 1011–1022 (2001). 131. Lopez-Otin, C., Blasco, M. A., Partridge, L.,
mitochondrial protein misfolding in Drosophila 110. Ohoka, N., Yoshii, S., Hattori, T., Onozaki, K. Serrano, M. & Kroemer, G. The hallmarks of aging.
melanogaster. Cell Death Differ. 19, 1308–1316 & Hayashi, H. TRB3, a novel ER stress-inducible gene, Cell 153, 1194–1217 (2013).
(2012). is induced via ATF4–CHOP pathway and is involved 132. Nunnari, J. & Suomalainen, A. Mitochondria: in
86. Horibe, T. & Hoogenraad, N. J. The chop gene in cell death. EMBO J. 24, 1243–1255 (2005). sickness and in health. Cell 148, 1145–1159 (2012).
contains an element for the positive regulation of the 111. Baker, B. M., Nargund, A. M., Sun, T. & Haynes, C. M. 133. Feng, J., Bussiere, F. & Hekimi, S. Mitochondrial
mitochondrial unfolded protein response. PLoS ONE Protective coupling of mitochondrial function and electron transport is a key determinant of life span
2, e835 (2007). protein synthesis via the eIF2α kinase GCN‑2. in Caenorhabditis elegans. Dev. Cell 1, 633–644
87. Zhao, Q. et al. A mitochondrial specific stress PLoS Genet. 8, e1002760 (2012). (2001).
response in mammalian cells. EMBO J. 21, 112. Rainbolt, T. K., Atanassova, N., Genereux, J. C. 134. Wong, A., Boutis, P. & Hekimi, S. Mutations in the
4411–4419 (2002). & Wiseman, R. L. Stress-regulated translational clk‑1 gene of Caenorhabditis elegans affect
The first description of the UPRmt. attenuation adapts mitochondrial protein import developmental and behavioral timing. Genetics 139,
88. Aldridge, J. E., Horibe, T. & Hoogenraad, N. J. through Tim17A degradation. Cell Metab. 18, 1247–1259 (1995).
Discovery of genes activated by the mitochondrial 908–919 (2013). 135. Yang, W. & Hekimi, S. A mitochondrial superoxide
unfolded protein response (mtUPR) and cognate 113. Michel, S., Canonne, M., Arnould, T. & Renard, P. signal triggers increased longevity in Caenorhabditis
promoter elements. PLoS ONE 2, e874 (2007). Inhibition of mitochondrial genome expression elegans. PLoS Biol. 8, e1000556 (2010).

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 17 | APRIL 2016 | 225


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

136. Kirchman, P. A., Kim, S., Lai, C. Y. & Jazwinski, S. M. 149. Dell’agnello, C. et al. Increased longevity and 164. Wang, D., Malo, D. & Hekimi, S. Elevated
Interorganelle signaling is a determinant of longevity refractoriness to Ca2+-dependent neurodegeneration mitochondrial reactive oxygen species generation
in Saccharomyces cerevisiae. Genetics 152, in Surf1 knockout mice. Hum. Mol. Genet. 16, affects the immune response via hypoxia-inducible
179–190 (1999). 431–444 (2007). factor‑1α in long-lived Mclk1+/– mouse mutants.
137. Kaeberlein, M. et al. Regulation of yeast replicative life 150. Zordan, M. A. et al. Post-transcriptional silencing J. Immunol. 184, 582–590 (2010).
span by TOR and Sch9 in response to nutrients. and functional characterization of the Drosophila 165. Wang, D. et al. An enhanced immune response of
Science 310, 1193–1196 (2005). melanogaster homolog of human Surf1. Genetics Mclk1+/– mutant mice is associated with partial
138. Hwang, A. B. et al. Feedback regulation via AMPK 172, 229–241 (2006). protection from fibrosis, cancer and the development
and HIF‑1 mediates ROS-dependent longevity in 151. Pulliam, D. A. et al. Complex IV‑deficient Surf1–/– mice of biomarkers of aging. PLoS ONE 7, e49606 (2012).
Caenorhabditis elegans. Proc. Natl Acad. Sci. USA initiate mitochondrial stress responses. Biochem. J. 166. Manzanillo, P. S. et al. The ubiquitin ligase parkin
111, E4458–E4467 (2014). 462, 359–371 (2014). mediates resistance to intracellular pathogens. Nature
139. Baruah, A. et al. CEP‑1, the Caenorhabditis elegans 152. Caldeira da Silva, C. C., Cerqueira, F. M., 501, 512–516 (2013).
p53 homolog, mediates opposing longevity outcomes Barbosa, L. F., Medeiros, M. H. & Kowaltowski, A. J. 167. Baixauli, F. et al. Mitochondrial respiration controls
in mitochondrial electron transport chain mutants. Mild mitochondrial uncoupling in mice affects energy lysosomal function during inflammatory T cell
PLoS Genet. 10, e1004097 (2014). metabolism, redox balance and longevity. Aging Cell responses. Cell Metab. 22, 485–498 (2015).
140. Walter, L., Baruah, A., Chang, H. W., Pace, H. M. 7, 552–560 (2008). References 160–167 report several examples of
& Lee, S. S. The homeobox protein CEH‑23 mediates References 147–152 give some examples of crosstalk between mitochondrial stress signalling
prolonged longevity in response to impaired mitochondrial stress affecting health and lifespan and immunity.
mitochondrial electron transport chain in C. elegans. in mammals. 168. Williams, D. S., Cash, A., Hamadani, L. & Diemer, T.
PLoS Biol. 9, e1001084 (2011). 153. Kim, H. J., Morrow, G., Westwood, J. T., Michaud, S. Oxaloacetate supplementation increases lifespan in
141. Palikaras, K., Lionaki, E. & Tavernarakis, N. & Tanguay, R. M. Gene expression profiling implicates Caenorhabditis elegans through an AMPK/FOXO-
Coordination of mitophagy and mitochondrial OXPHOS complexes in lifespan extension of flies dependent pathway. Aging Cell 8, 765–768 (2009).
biogenesis during ageing in C. elegans. Nature 521, over‑expressing a small mitochondrial chaperone, 169. Dillin, A. et al. Rates of behavior and aging specified
525–528 (2015). Hsp22. Exp. Gerontol. 45, 611–620 (2010). by mitochondrial function during development.
A description of how mitochondrial biogenesis and 154. Andreux, P. A., Houtkooper, R. H. & Auwerx, J. Science 298, 2398–2401 (2002).
mitophagy are coordinately regulated to determine Pharmacological approaches to restore mitochondrial A description of the longevity phenotypes of C.
lifespan in C. elegans. function. Nat. Rev. Drug Discov. 12, 465–483 (2013). elegans ETC mutants.
142. Ristow, M. Unraveling the truth about antioxidants: 155. Williams, E. G. & Auwerx, J. The convergence of 170. Yang, W. & Hekimi, S. Two modes of mitochondrial
mitohormesis explains ROS-induced health benefits. systems and reductionist approaches in complex trait dysfunction lead independently to lifespan extension in
Nat. Med. 20, 709–711 (2014). analysis. Cell 162, 23–32 (2015). Caenorhabditis elegans. Aging Cell 9, 433–447 (2010).
143. Zhang, Y., Shao, Z., Zhai, Z., Shen, C. & 156. Yun, J. & Finkel, T. Mitohormesis. Cell Metab. 19, 171. Liu, J. et al. Drosophila sbo regulates lifespan through
Powell‑Coffman, J. A. The HIF‑1 hypoxia-inducible 757–766 (2014). its function in the synthesis of coenzyme Q in vivo.
factor modulates lifespan in C. elegans. PLoS ONE 4, 157. Dietrich, A., Wallet, C., Iqbal, R. K., Gualberto, J. M. J. Genet. Genom. 38, 225–234 (2011).
e6348 (2009). & Lotfi, F. Organellar non-coding RNAs: emerging 172. Lemire, B. D., Behrendt, M., DeCorby, A. &
144. Heidler, T., Hartwig, K., Daniel, H. & Wenzel, U. regulation mechanisms. Biochimie 117, 48–62 Gaskova, D. C. elegans longevity pathways converge
Caenorhabditis elegans lifespan extension caused by (2015). to decrease mitochondrial membrane potential.
treatment with an orally active ROS-generator is 158. Sugiura, A., McLelland, G. L., Fon, E. A. & Mech. Ageing Dev. 130, 461–465 (2009).
dependent on DAF‑16 and SIR‑2.1. Biogerontology McBride, H. M. A new pathway for mitochondrial
11, 183–195 (2010). quality control: mitochondrial-derived vesicles. Acknowledgements
145. Lee, H. et al. The Caenorhabditis elegans AMP- EMBO J. 33, 2142–2156 (2014). The authors thank all members of the Auwerx laboratory for
activated protein kinase AAK‑2 is phosphorylated 159. Arnoult, D., Soares, F., Tattoli, I. & Girardin, S. E. their helpful comments on the manuscript and apologize for
by LKB1 and is required for resistance to oxidative Mitochondria in innate immunity. EMBO Rep. 12, the omission of relevant work owing to space constraints.
stress and for normal motility and foraging 901–910 (2011). P.M.Q. is supported by a long-term fellowship from the
behavior. J. Biol. Chem. 283, 14988–14993 160. Liu, Y., Samuel, B. S., Breen, P. C. & Ruvkun, G. European Molecular Biology Organization (EMBO; ALTF
(2008). Caenorhabditis elegans pathways that surveil and 480–2014). J.A. is the Nestlé Chair in Energy Metabolism,
146. Schaar, C. E. et al. Mitochondrial and cytoplasmic ROS defend mitochondria. Nature 508, 406–410 (2014). and the Auwerx laboratory is supported by grants from the
have opposing effects on lifespan. PLoS Genet. 11, 161. Pellegrino, M. W. et al. Mitochondrial UPR-regulated École Polytechnique Fédérale de Lausanne, the Swiss
e1004972 (2015). innate immunity provides resistance to pathogen National Science Foundation (31003A‑140780), the AgingX
147. Liu, X. et al. Evolutionary conservation of the infection. Nature 516, 414–417 (2014). programme of the Swiss Initiative for Systems Biology
clk‑1‑dependent mechanism of longevity: loss of 162. West, A. P. et al. Mitochondrial DNA stress primes (51RTP0‑151019), the Krebsforschung Schweiz/Swiss Cancer
mclk1 increases cellular fitness and lifespan in mice. the antiviral innate immune response. Nature 520, League (KFS‑3082‑02‑2013) and the US National Institutes
Genes Dev. 19, 2424–2434 (2005). 553–557 (2015). of Health (NIH; R01AG043930).
148. Lapointe, J. & Hekimi, S. Early mitochondrial 163. West, A. P. et al. TLR signalling augments macrophage
dysfunction in long-lived Mclk1+/– mice. J. Biol. Chem. bactericidal activity through mitochondrial ROS. Competing interest statement
283, 26217–26227 (2008). Nature 472, 476–480 (2011). The authors declare no competing interests.

226 | APRIL 2016 | VOLUME 17 www.nature.com/nrm


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like