ArticlePDF Available

Abstract and Figures

Aging is characterized by a progressive loss of muscle mass and muscle strength. Declines in skeletal muscle mitochondria are thought to play a primary role in this process. Mitochondria are the major producers of reactive oxygen species, which damage DNA, proteins, and lipids if not rapidly quenched. Animal and human studies typically show that skeletal muscle mitochondria are altered with aging, including increased mutations in mitochondrial DNA, decreased activity of some mitochondrial enzymes, altered respiration with reduced maximal capacity at least in sedentary individuals, and reduced total mitochondrial content with increased morphological changes. However, there has been much controversy over measurements of mitochondrial energy production, which may largely be explained by differences in approach and by whether physical activity is controlled for. These changes may in turn alter mitochondrial dynamics, such as fusion and fission rates, and mitochondrially induced apoptosis, which may also lead to net muscle fiber loss and age-related sarcopenia. Fortunately, strategies such as exercise and caloric restriction that reduce oxidative damage also improve mitochondrial function. While these strategies may not completely prevent the primary effects of aging, they may help to attenuate the rate of decline.
Content may be subject to copyright.
Hindawi Publishing Corporation
Journal of Aging Research
Volume 2012, Article ID 194821, 20 pages
doi:10.1155/2012/194821
Review Article
Skeletal Muscle Mitochondria and Aging: A Review
Courtney M. Peterson, Darcy L. Johannsen, and Eric Ravussin
Department of John S. Mclhenny Skeletal Muscle Physiology, Pennington Biomedical Research Center,
6400 Perkins Road, Baton Rouge, LA 70808, USA
Correspondence should be addressed to Darcy L. Johannsen, darcy.johannsen@pbrc.edu
Received 23 March 2012; Accepted 21 May 2012
Academic Editor: Holly M. Brown-Borg
Copyright © 2012 Courtney M. Peterson et al. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.
Aging is characterized by a progressive loss of muscle mass and muscle strength. Declines in skeletal muscle mitochondria are
thought to play a primary role in this process. Mitochondria are the major producers of reactive oxygen species, which damage
DNA, proteins, and lipids if not rapidly quenched. Animal and human studies typically show that skeletal muscle mitochondria
are altered with aging, including increased mutations in mitochondrial DNA, decreased activity of some mitochondrial enzymes,
altered respiration with reduced maximal capacity at least in sedentary individuals, and reduced total mitochondrial content
with increased morphological changes. However, there has been much controversy over measurements of mitochondrial energy
production, which may largely be explained by dierences in approach and by whether physical activity is controlled for. These
changes may in turn alter mitochondrial dynamics, such as fusion and fission rates, and mitochondrially induced apoptosis, which
may also lead to net muscle fiber loss and age-related sarcopenia. Fortunately, strategies such as exercise and caloric restriction that
reduce oxidative damage also improve mitochondrial function. While these strategies may not completely prevent the primary
eects of aging, they may help to attenuate the rate of decline.
1. Introduction
Around the fourth decade of life, both muscle mass and
strength begin to decline [1], and these declines accelerate
with advancing age [2].Thelossofmusclemassoccursat
arateofjustunder1%peryear[3] and appears to be an
unavoidable consequence of aging, although it can be slowed
by exercise, especially resistance training [46]. A significant
concern is that as one ages, changes in muscle mass and
strength tend to be dissociated. Data from the Baltimore
Longitudinal Study of Aging [7] and the Health ABC study
[3] showed using DXA and CT that muscle strength declined
three times faster than muscle mass, suggesting a decrease
in muscle “quality.” This posits that along with an overall
reduction in tissue mass, changes are occurring within the
skeletal muscle to aect strength. Changes such as accumula-
tion of intra- and extra-myocellular lipids, improper folding
of structural and contractile proteins, and mitochondrial
dysfunction are thought to occur with age and are the topic
of intense scrutiny [810].
Dysfunctional mitochondria in particular are thought
to play a key role in muscle function decline, as the
mitochondria are the main producers of both cellular energy
and free radicals. Alterations in mitochondria have been
noted in aging, including decreased total volume, increased
oxidative damage, and reduced oxidative capacity. These
biochemical and bioenergetic changes are accompanied by
perturbations in cellular dynamics, such as a decrease in
mitochondrial biogenesis and an increase in mitochondrially
mediated apoptosis (Figure 1). These changes may underlie
not only a loss of muscle quality with age, but also other
common age-associated pathologies such as ectopic lipid
infiltration, systemic inflammation, and insulin resistance
[10]. In this paper, we examine the evidence for age-related
changes in skeletal muscle mitochondria, with a focus on
studies conducted in humans.
1.1. Overview of Mitochondria. Mitochondria are double-
membrane encoded organelles with their own genome that
2 Journal of Aging Research
Muscle
Mass
Function
Mitochondria
Number
Morphology changes
DNA mutations
Biogenesis
Oxidative capacity ?
Autophagy
Apoptosis
Oxidative
stress
ROS
Lipid/protein
damage
Antioxidant
system
Aged skeletal muscle
?
Exercise
CR
CR mimetics
Antioxidants
Figure 1: This cartoon describes the changes in skeletal muscle with aging on the right side of the figure. Both the mass and function
of skeletal muscle are decreased in elderly people. Furthermore, at the mitochondrial level, the number of mitochondria is decreased in
parallel with changes in mitochondrial morphology. Mitochondrial DNA, oxidative capacity, biogenesis, and autophagy are all decreased
in conjunction with an increased number of DNA mutations and increased levels of apoptosis. Finally, oxidative stress is increased in the
muscles of elderly people in association with cellular lipid, protein, and DNA damage. The bottom left of the cartoon shows that exercise,
caloric restriction, caloric restriction mimetics, and antioxidants can all delay the aging of skeletal muscle.
consume oxygen and substrates to generate the vast majority
of ATP while producing reactive oxygen species in the
process. They also participate in a wide range of other
cellular processes, including signal transduction, cell cycle
regulation, oxidative stress, thermogenesis, and apoptosis.
In doing so, they are highly dynamic organelles that are
continuously remodeling through biogenesis, fission, fusion,
and autophagy, thus responding to and modulating cellular
dynamics. For instance, they undergo biogenesis to meet
increased energy demands in response to exercise, and they
ensure cellular quality by initiating an apoptotic program to
remove defective cells.
Mitochondria transduce energy from substrates through
the tricarboxylic acid (TCA) cycle and the electron trans-
port system (ETS) to generate ATP. The ETS consists of
multipolypeptide complexes (I–V) embedded in the inner
mitochondrial membrane (IMM) that receive electrons from
reducing equivalents NADH and FADH2, generated by
dehydrogenase activity in the TCA cycle. The electrons are
transferred along the complexes with O2serving as the
final acceptor at complex IV [11](Figure 2). The reduction
potential (propensity to accept an electron) increases along
the chain of complexes, and the energy generated is sucient
to drive the translocation of hydrogen ions across the IMM.
This creates a proton gradient and membrane potential
(collectively termed the proton motive force) that drives
the synthesis of ATP as protons flow back to the matrix
via complex V (ATP synthase). This process is also called
oxidative phosphorylation (OXPHOS). However, the ETS is
not a perfectly ecient system, and significant (and highly
variable) proton “leak” occurs by the movement of hydrogen
ions back into the matrix space that is not mediated through
complex V. In this manner, proton transfer can be uncoupled
from ATP phosphorylation, and this ineciency contributes
to the demand for reducing equivalents. Mitochondria are
therefore thought to have a much greater capacity to generate
ATP than what is usually required [12].
Because of this metabolic latitude, many assume that
mild impairment in mitochondrial function per se does
not cause cellular disturbances associated with aging and
with chronic diseases such as insulin resistance and type 2
diabetes [12]. Indeed, whether mitochondrial dysfunction
is a cause or a consequence of cellular impairment and
the aging process is a subject of intense debate. Moreover,
the definition of mitochondrial dysfunction itself has been
the subject of controversy. For example, alterations in
mitochondrial mRNA transcripts may not result in changes
in protein levels, so it is not wholly clear whether this
state—whether compensatory or not—can be counted as a
disruption in normal function [13].
In this paper, we focus on the aging of skeletal muscle
(SM) mitochondria, with an eye towards the putative role
of mitochondria in SM aging. We look at several dierent
aspects of mitochondrial function—which we define to mean
any process that involves the mitochondria—and explore
the evidence for dysfunction, which we define to mean any
deviation from normal function. Section 2 examines the
impact of aging on the biochemical and bioenergetic path-
ways in SM mitochondria. Section 3 explores the changes
in mitochondrial and cellular dynamics, which may further
exacerbate the age-related decline in muscular function. The
review concludes in Section 4 with an overview of strategies
to attenuate the aging of SM mitochondria.
2. Biochemical and Bioenergetic
Aging of Mitochondria
During the aging process, mitochondria are characterized by
changes in oxidative stress, a decay in mitochondrial DNA,
Journal of Aging Research 3
Intermembrane space
4H+4H+2H+
3H+
ATPADP
+Pi
FADH2FAD+
+2H
+
+2H
+
NAD+
NADH
+H
+
Matrix H2O
2H+
Intermem
b
rane s
p
ace
4H
+
4
H
+
2H
+
3H
+
ATPAD
P
+ P
i
FADH
FA
D
+
+
2H
+
+
2
H
+
N
AD+
N
A
D
H
+
H
+
Matrix
H
O
2H
+
+ 1/2 O2
(a)
Drp1
Fis1
Completion of
fission
(1) Fission
(division)
(2) Fusion
Opa1
Fusion of inner
membrane
Docking of
Mfn1/2
Constriction
of inner and
outer
membrane
Adapted from Dimmer and Scorrano, 2007
(b)
Figure 2: Mitochondrial processes are both static (a) and dynamic (b). (a) depicts the classical movement of electrons along complexes
I–IV embedded in the inner mitochondrial membrane with the generation of a proton gradient (membrane potential). The proton gradient
causes hydrogen ions to flow back into the mitochondrial matrix via complex V (ATP synthase), producing ATP in the process. (b) depicts
the processes of mitochondrial fusion and fission. Mitochondria can undergo constriction and division (1), mediated by Drp1, which bonds
and localizes to the constriction site via an interaction with the receptor-like protein Fis1 (2). During fusion, a tether of the Mfn1/2 to
collateral mitochondrial Mfn1/2 conjoins the outer membranes. Opa1, an inner membrane GTPase protein, facilitates the fusion of the
inner membrane, cristae formation, and unifying of compartments.
a reduction in some enzyme activities, and alterations in
mitochondrial respiration. These biochemical changes are
accompanied by phenotypic changes in the mitochondria
themselves. In old age, a significant proportion of the
mitochondrial organelles are abnormally enlarged and more
roundedinshape(reviewedin[14,15]). An increasing
proportion of them are also depolarized or nonfunctional,
perhaps indicative of defects in mitochondrial turnover
[1517]. Within the mitochondria themselves, shortened
mitochondrial cristae and vacuolization of the matrix are
apparent, which lead to homogenization of the materials
within the mitochondrial compartments [14,18]. Coupled
with these morphological changes, the density of mitochon-
dria in SM substantially drops [13,1921], as shown for
example, by electron microscopy in the vastus lateralis muscle
of people over 60 years of age [20].
2.1. Increased Oxidative Damage: ROS Production and Scav-
enging. Within the cell, the mitochondria constitute the
major source of reactive oxygen species (ROS). Mitochon-
drial complexes I and III are the main sites of superoxide
generation and contribute the most to ROS production [22].
The reactive oxygen species, including O2
and H2O2,can
cause oxidative damage to surrounding structures and the
particularly vulnerable mtDNA, which is in close proximity
to the primary site of ROS production. Oxidation by ROS
results in the synthesis of faulty proteins, oxidized lipids,
and mtDNA mutations, which may lead to cellular and
mitochondrial dysfunction. These processes are implicated
in the mitochondrial theory of aging [23,24], which holds
that the accumulation of ROS damage over time leads to
age-associated mitochondrial impairment. In general, ROS
production is found to be increased in aged muscle in
both the subsarcolemmal and intermyofibrillar pools of
mitochondria [25] (though some report a decrease in ROS
production with age [26]). This increase in ROS production
is associated with oxidation of ETS complex V, leading
to decreased ATP production [27,28], increased levels
of 8-oxodeoxyguanosine (8-oxoG) from DNA oxidation
4 Journal of Aging Research
[25,29], increased levels of protein carbonyls, and increased
nitration [30]. In particular, proteomic studies have found
increased nitration of complex II and altered carbonylation
of complex I, complex V, and isocitrate dehydrogenase
(reviewed in [31]).
Cumulative oxidative damage may in part be attributed
to a reduction in ETS activity that would extend the
length of time that electrons remain at complexes I and
III, increasing the potential for donation of electrons to
oxygen [32]. In theory, it could also be attributed to
reduced activity of antioxidant defenses, including man-
ganese superoxide dismutase (MnSOD), catalase (CAT), and
glutathione peroxidase (GPx). These enzymes work together
to convert O2
to H2O2, which is then further reduced
to H2O. Data on antioxidant enzyme activity and aging is
mixed, with some studies showing increased activity [33
35] while others show reduced enzyme activity [3638].
We recently found in a group of elderly adults aged 70–
84 years that the expression of total superoxide dismutase
(SOD) was unchanged compared to that of a group of
young (20–34 years old) BMI-matched individuals; however,
urinary isoprostanes (a marker of whole-body oxidative
damage) was increased by 38% [10]. Furthermore, TCA
cycle activity measured as citrate synthase (CS) activity
was 19% lower, and mitochondrial capacity was 17% lower
in the elderly, although mitochondrial content (OXPHOS
protein) was unchanged. Safdar et al. recently showed that
the protein content of MnSOD (mitochondria-specific) was
significantly reduced in active and inactive older adults
compared to their younger counterparts; however, MnSOD
activity in the older subjects who were recreationally active
was similar to that of young subjects [39]. Together, these
data suggest that oxidative stress is increased with age
without an accompanying increase in antioxidant activity
and may be linked with mitochondrial dysfunction. Exercise
and caloric restriction may alleviate the age-associated
oxidative stress by upregulating antioxidant enzyme activity
even in the face of reduced antioxidant protein content
(Figure 3).
If oxidative stress contributes to mitochondrial dys-
function, can mitochondrial dysfunction be reversed by
changing the cellular oxidative status? A recent study
attempted to answer this question using a transgenic
mouse model to express human catalase targeted to the
mitochondria (MCAT) [40]. Mitochondrial function was
assessed both in vitr o and in vivo in wildtype young
(3–6 months old) and older (15–18 months old) lean
healthy mice versus age-matched MCAT transgenic lit-
termates. Whereas the older wildtype mice exhibited all
the “usual” deleterious metabolic impairments including
increased oxidative damage, reduced mitochondrial con-
tent (30%) and function, increased intramuscular lipid
(70%), and marked muscle insulin resistance (35%), the
older MCAT mice resembled the young mice and demon-
strated none of the age-associated impairments. These data
suggest that by increasing ROS scavenging and reducing
the oxidized state of the cell, many age-related deficits can
be prevented.
2.2. Increased Uncoupling? Another mechanism to reduce
oxidative stress may be mediated by attenuation of the proton
motive force, thereby decreasing the potential for electrons
to form oxygen radicals. This could be accomplished by
increasing respiration or by increasing proton leak. Rodent
studies on age-associated changes in proton leak are mixed.
Iossa et al. [41] showed that proton leak was decreased
with age, resulting in increased mitochondrial eciency.
Conversely, reduced coupling was found in aged mice using
an in vivo approach to measure the P/O ratio [42]. Data in
aged humans are scarce. We recently found that the in vivo
P/O ratio determined by ATP turnover (demand) and oxygen
uptake of the vastus lateralis was 21% lower in elderly
compared to young adults [10], suggesting that less ATP
was produced per oxygen uptake in the elderly adults. This
lends support to earlier work by Amara et al. [43] showing
that mild uncoupling protects against age-related declines
in mitochondrial function and cellular ATP concentration.
Furthermore, we found that a lower P/O ratio was associated
with lower SOD activity, indicating a possible link between
cellular antioxidant activity and mitochondrial coupling. In
particular, increased oxidative stress with nonupregulated
antioxidant defense may induce mitochondrial uncoupling
in order to reduce the oxidative potential. This potential
adaptive mechanism should be further explored.
2.3. Decay of Mitochondrial DNA. Age appears to aect
both mitochondrial DNA (mtDNA) content and integrity.
Mammalian cells typically contain on the order of 1,000–
10,000 copies of mtDNA, which code for less than 10% of all
mitochondrial proteins; the rest is encoded by nuclear DNA
[44,45]. At 16.6 kbp long, each double-stranded circular
mtDNA molecule encodes 13 genes involved in the ETS
(subunits of complexes I, III, IV, and V), 22 tRNAs, and
2rRNAs[46]. A majority of studies [4750], though not
all [51], report that mtDNA copy number decreases with
age in human SM and may be caused at least in part by
oxidative damage, as the age-associated decline in mtDNA
copy number tends to be greater in more oxidative fibers
[52].
The decline in mtDNA is also accompanied by an
increase in mtDNA damage, particularly deletions and
oxidative lesions, but also point mutations, tandem duplica-
tions, and rearrangements (reviewed in [53,54]; in particular
[3638,48,5564]). mtDNA is particularly susceptible to
oxidative damage because it is near the major site of ROS
production, lacks protective histones, and has weaker DNA
repair mechanisms [65,66]. One study found that deletions
aect up to 70% of mtDNA molecules in the SM of
individuals aged 80 years and older [56]. However, there are
disagreements over mtDNA mutation frequencies (reviewed
in [54]) and to what extent ROS production is responsible;
there is increasing evidence that perhaps a majority of
mutations are due to the inherent error rate of mtDNA
polymerase gamma (Pol gamma) [67].
ThefunctionalimpactofmtDNAdamageisstillamatter
of debate both in general (reviewed in [54,67,68]) and
in the context of sarcopenia (reviewed in [69]). Is mtDNA
Journal of Aging Research 5
Mitochondria in young muscle
ADP ATP
Little oxidative damage
Mitoch.
efficiency
Oxidative
capacity
+++ Good Good
Mitoch.
number
III
III IV
V
Q
C
O2H2O
(a)
Mitochondria in older muscle
ADP ATP
ROS
Ox. damage
Ox. damage
III
III IV
V
Q
C
O2H2O
(b)
Mitochondria in older muscle + antioxidant mechanism
ADP ATP
ROS Oxidative
damage
III
III IV
V
Q
C
O2H2O
(c)
Figure 3: The electron transport system (ETS) of the inner mitochondrial membrane is the primary site of reactive oxygen species (ROS)
production and therefore the main source of oxidative stress (damage to proteins, lipids, and DNA) in the mitochondria and in the cell.
Free radical superoxide anions (O2
) are generated when electrons are donated from complexes I and III of the ETS to O2instead of the
appropriate ETS subunit. 2–4% of total oxygen consumption may go toward the production of ROS instead of energy as ATP. Scavenging
enzymes represent an important mitochondrial defense mechanism against oxidative stress by neutralizing O2
within the mitochondrial
matrix (superoxide dismutase; MnSOD =SOD2) and catalyzing the reduction of mitochondrial SOD2-generated H2O2to nontoxic H2Oin
the mitochondria and the cell (glutathione peroxidase and catalase). Mitochondria in young muscle (a) are numerous and ecient. With
age (b), muscle mitochondria become less numerous and seem to develop impaired function associated with reduced oxidative capacity.
Through lifestyle changes such as exercise, and caloric restriction, and caloric restriction mimetics, we hypothesize that antioxidative enzymes
are upregulated, and that most of the above impairments in aged muscle may be improved (c).
damage a consequence or a cause of the aging process? On
one hand, the onset of a mitochondrial decline in energy
production occurs before mtDNA mutations are often
detectable [70]. On the other hand, several studies in humans
reveal strong correlations between mtDNA mutation rates
and bioenergetic deficiency (typically complex IV) or muscle
fiber atrophy [43,55,62,63,7174]. In addition, specific
point mutations cluster at mtDNA replication control sites,
which may reduce gene transcription and be at least partially
responsible for declining levels of protein synthesis with
age [60,75]. Other evidence comes from mtDNA mutator
mice, which are genetically engineered to have defects in
the proofreading function of Pol gamma. They accumulate
mtDNA mutations at an accelerated rate, have more abnor-
mal mitochondria, and exhibit premature aging, sarcopenia,
and reduced lifespan [7678].
2.4. Alterations in Mitochondrial mRNA and Protein Levels.
The expression of most mitochondrial genes, including
cytochrome c [79], does not change with age. However,
declines in gene transcripts are observed for several of
the polypeptide components of complexes I, IV, and V,
whereas for complexes II and III, declines are evident for
6 Journal of Aging Research
only a couple components with most being unchanged
[73,8084]. In addition, the transcripts of some nuclear
genes encoding mitochondrial regulatory proteins, a few
TCA proteins (including one transcript variant of citrate
synthase), and some glycolysis enzymes appear to be altered
in aged human SM [48,52,80,83,84]. Many similar
changes in gene expression (or lack thereof) have been
observed in monkeys and rats [52,8587]. However, there
are discrepancies among reports, which are at least in part
explained by muscle-specific dierences in gene expression
[52,88]. Whether these alterations in gene transcripts of
energy-producing pathways aect muscle aging is unclear.
While much evidence points to a functional role, Giresi et
al. found that the key genes associated with sarcopenia are
involved in inflammation, apoptosis, and regulating mRNA
splicing, not in OXPHOS [89].
Of course, age-related alterations in mRNA levels may
not induce similar changes in protein abundances, which are
controlled by the balance between synthesis and degradation
[13,90]. Both mitochondrial protein synthesis [75]and
proteolysis via ubiquitin-proteasome and lysosomal degra-
dation systems (reviewed in [91,92]) are known to decline
with age. The majority of mitochondrial proteins, including
polypeptide components of the ETS complexes, have not
been found to change with age [10,93]. Notable exceptions
include complex II, which tends to increase with age at least
in rodents [86,94,95], and complex IV, which tends to
decrease with age in both humans [50,96,97] and rodents
[86,98]. This is supported by observations of increased
complex IV-deficient and complex II-hyperactive fibers with
age [99101]. For complexes I, III, and V, proteomic analysis
shows that the occurrence of abnormal polypeptides tends to
increase with age in both human [102] and rodent SM [86,
94,98,103], but is decreased in human studies using other
approaches [48,50,96]. Proteins involved in glycolysis tend
to be unchanged or reduced with age, which is consistent
with the observed shift from glycolytic to more oxidative
metabolism, while the data on TCA proteins (in particular,
citrate synthase and isocitrate dehydrogenase) is mixed [48,
50,86,9496,98,102,103].
2.5. A Decline in Mitochondrial Energy Production? It is
unclear whether these age-related changes in turn aect
mitochondrial energy production. When physical activity is
not a criterion in subject selection, substantial declines in
enzyme activity are often found to occur with age. Most such
studies in human SM report that complex I and complex IV
activities decrease substantially, perhaps because these two
complexeshavemoreoftheirsubunitsencodedbythemore
vuinerable mtDNA than the other complexes [1,21,56,58,
75,79,90,97,104,105]. Similar results have been reported
in rodents and dogs [2,5,52,106112], though there have
been exceptions both in humans [93,113,114] and rodents
[86,94,109]. The activity of complex II, which is encoded
entirely by the nuclear genome, appears not to change with
age in either human [1,56,64,113,114] or animal [2,
86,108] SM, with a few exceptions [39,79,94,105]. Data
regarding complex III activity is both less available and more
mixed, with some studies reporting no change in human
[113,114] or animal [2,108] SM, but others reporting a
decrease in activity [5,56]. Interestingly, one study found
evidence of an age-related decline in complex III activity in
females but not males [90]. Finally, the activities of enzymes
involved in the TCA cycle and glycolytic pathways tend to
be unchanged or decline with age, and in some cases (e.g.,
citrate synthase), the trend is not clear [10,21,26,39,48,56,
75,93,96,97,104,105,111113,115120]. Some of these
discrepancies may be explained by the dierential impact of
aging on dierent muscles. For example, one study in rats
found a decrease in complex IV activity in the lateral but
not medial gastrocnemius [52], and a study in humans found
that citrate synthase activity was negatively correlated with
age in the gastrocnemius but not the vastus lateralis [121].
Other inconsistencies may be explained by dierences in
isolation or techniques, or in the normalization of enzymatic
activity. For example, Picard et al. found that measurements
on isolated mitochondria tend to exaggerate the declines in
mitochondrial function in comparison to measurements in
intact mitochondria in permeabilized myofibers [86].
Enzymatic changes may in turn aect mitochondrial
respiration and ATP flux. Short et al. found that the maximal
capacity for ATP synthesis drops by 8% per decade or
5% when normalizing to mitochondrial protein content
[48]. Similar results concerning mitochondrial respiration
were reported by Trounce et al. [122]. But other human
studies have found no evidence of a decline in mitochondrial
respiration [26,93,115,117], and there is no consistent
pattern regarding how the respiratory control ratio—the
ratio of states III and IV respiration activities—changes with
age. Similarly, there are conflicting reports on fatty acid-
linked mitochondrial respiration including reduced carnitine
palmitoyltransferase-dependent and -independent pathways
with age (particularly nonphosphorylating respiration [48,
86,108]), while others see no age-related changes [26,115,
117]. In vivo techniques have also been used to measure
oxidative capacity. In vivo measurements of SM oxidative
activity are typically performed using phosphorous magnetic
resonance spectroscopy (31P-MRS) to probe the kinetics
of phosphocreatine and its recovery time following muscle
contraction. Some in vivo studies in older humans [10,20,
118,120], but not all [119,123,124], and at least one in mice
[125] have found evidence of reductions in maximal ATP
flux. In particular, Conley et al. found that elderly adults have
a 50% reduction in oxidative capacity per muscle volume
and a 30% reduction per mitochondrial volume [20]. An
MRS study on basal ATP flux in humans, however, found no
decline in flux with aging [43].
However, it has become increasingly clear that most
of the declines in mitochondrial function attributed to
chronological age are instead a result of physical inactivity.
When physical activity levels are matched between young
and old subjects or physical activity is otherwise taken
into account, most studies find no age-related changes in
mitochondrial enzyme activities, mitochondrial respiration,
or ATP flux [13,39,50,101,126129]. Interestingly,
those studies that do report age-related declines even after
matching for activity levels tend to involve sedentary young
Journal of Aging Research 7
and old subjects [104,114,130], indicating that declines in
function with aging may occur predominantly in sedentary
individuals. In particular, in ex vivo studies, the activities
of ETS enzymes and citrate synthase and mitochondrial
respiration show strong dependences on physical activity
level that are independent of age [101,131,132]. Similar
results are also reported in mice [98,133]. In vivo MRS
studies on activity-matched subjects also tend to find no
evidence of a change in mitochondrial oxidative capacity—
in this case, maximal ATP flux—between young and old
subjects [134136]. However, one in vivo MRS study in
activity-matched sedentary individuals did report a reduc-
tion in basal oxidation and phosphorylation [130]. Also,
glycolytic flux is lower in older activity-matched people
[135,136], and a recent MRS study showed that oxidative
capacity does indeed change in some muscles with age but
also demonstrated that physical activity is intimately linked
with oxidative capacity [128]. Interestingly, there is evidence
that physical activity can reverse the age-related declines in
most but not all mitochondrial markers of energy production
[50,132], which while encouraging, nonetheless indicates
that there are residual declines in a small subset of markers
that cannot be completely prevented.
3. Age-Related Changes in
Mitochondrial Dynamics
3.1. Decreased Mitochondrial Biogenesis. Once thought of
as relatively static round organelles, mitochondria are now
recognized as highly dynamic, existing in networks that are
constantly being remodeled by biogenesis, fusion and fission,
and degradative processes such as autophagy. Through these
dynamics, they both respond to and drive cellular processes,
including apoptosis, whose dysregulation is thought to be a
key factor in sarcopenia.
Mitochondrial biogenesis is the expansion of existing
mitochondrial content—whether through growth of the
mitochondrial network (increase in mitochondrial mass)
or division of preexisting mitochondria (increase in mito-
chondrial number; also discussed later in this paper). It
is triggered when the energy demand exceeds respiratory
capacity—in particular, in response to exercise, stress,
hypoxia, nutrient availability, hormones (including insulin),
ROS production, and temperature (reviewed in [137,138]).
Once biogenesis is triggered, the nuclear genome produces
mitochondrial regulatory factors, which are then imported
into the mitochondria to initiate replication and transcrip-
tion of mtDNA, ultimately resulting in expansion of the
mitochondrial network. The initial triggers are thought to
arise from signaling cascades involving the energy sen-
sor AMP kinase (AMPK) and/or alterations in Ca2+ flux
and protein kinases such as calcium/calmodulin-dependent
kinases (CAMKs), protein kinase C (PKC), and p38 MAPK
(particularly in response to muscle contraction) (reviewed
in [138]). These signals in turn induce the expression of the
peroxisome proliferator-activated receptor (PPAR) gamma
coactivator (PGC-1) family of cotranscription factors, par-
ticularly PGC-1α, which is also directly activated by AMPK
[139] and p38 MAPK [140142], as well as many key
signaling molecules like SIRT1 [143,144] and transducers of
regulated CREB (cAMP response element-binding protein)-
binding proteins (TORCs) [145].
As the master regulator of biogenesis, PGC-1αcoordi-
nates and cooperates with multiple cotranscription factors—
including PPARs, myocyte enhancing factors (MEFs), and
CREB—to induce the transcription of nuclear genes encod-
ing mitochondrial proteins [146,147]. Most importantly,
PGC-1αactivates the nuclear respiratory factors 1 and
2 (Nrf-1, Nrf-2) on the promoters [148,149], thereby
driving the transcription of an even greater assortment of
nuclear-encoded mitochondrial proteins (reviewed in [137,
138,150]), including mitochondrial transcription factor A
(Tfam). Tfam controls mtDNA transcription and content
and organizes mtDNA into nucleoid-like structures, which
are thought to maintain mtDNA integrity [151]. The Tfam
precursor protein, along with numerous other mitochon-
drial precursor proteins, are targeted to the mitochondria
by chaperones and then imported via the mitochondrial
protein import machinery (reviewed in [152155]). Import
is accomplished through a set of aqueous pores formed from
translocases of the outer membrane (TOMs; e.g., Tom20)
and translocases of the inner membrane (TIMs) through
pathways that depend on whether the polypeptide is destined
for the outer mitochondrial membrane (OMM), IMM,
or mitochondrial matrix. Once imported, the precursor
proteins are modified, folded, and assembled into their final
form. In particular, Tfam can then act on mtDNA driving its
replication through Pol gamma and its transcription through
mitochondrial RNA polymerase [156]. Finally, the translated
ETS polypeptides from the nuclear and mitochondrial
genomes are assembled into multisubunit enzymes.
With increasing age, the density of mitochondria in
SM drops substantially [13,1921,25], suggesting an
overall decline in mitochondrial biogenesis. Moreover, there
is an impairment in AMPK-stimulated biogenesis in old
age [157]; however, the reasons why are largely unknown.
Declining PGC-1αlevels could explain the reduction in
mitochondrial biogenesis, as overexpression of PGC-1αin
the SM of aged mice improved oxidative capacity, suppressed
mitochondrial degradation, and prevented muscle atrophy
[158]. These improvements were accompanied by attenua-
tion of the age-related increase in inflammatory cytokines
and by prevention of insulin resistance [158]. However,
measurements of PGC-1αin aged SM are not definitive:
one study found a decline in protein abundance in rats
[25], and some studies of gene expression have reported
reduced expression [96,112,157], while others found no
change [10,159]. Similarly, reports on the gene and protein
expressions of both Nrf-1 and Tfam are conflicting [25,50,
80,96,159,160]; however, preliminary evidence suggests
that Nrf-1 binding to the Tfam promoter appears to increase
in old age [160].
3.2. Changes in Fission and Fusion? Mitochondrial dynamics
are also influenced by the balance between fission and
8 Journal of Aging Research
fusion (Figure 2(b)). Fission, or division, of mitochon-
dria is required to transmit mitochondria among dividing
cells and to meet increased ATP demands. Fission also
plays a key role in maintaining mitochondrial quality and
mtDNA integrity, as it allows dysfunctional mitochondria
to be severed from the network and to be removed by
autophagy [161,162]; indeed, mitochondria excised by
fission often have a lower membrane potential, a target
for autophagy [16]. In mammals, fission is known to
be orchestrated by two proteins: Fis1 and the GTPase
dynamin-related protein Drp1. Fis1, which localizes in
the OMM, recruits the cytosolic protein Drp1 [162,163].
Once recruited to the OMM, Drp1 wraps around and
constricts the membrane, initiating fission (reviewed in
[164]). One study found that activation of the fission
machinery was sucient to induce muscular atrophy [165],
while another study found that in yeast, increased fission
leads to a shortened lifespan and a greater sensitivity to
ROS-induced apoptosis [166]. Because Fis1 protein levels
areelevatedinagingratSM[94] and Drp1 transcripts
tend to be lower in older humans [21], age-associated
increases in fission may indeed contribute to sarcopenia.
On the other hand, suppression of Fis1 or Drp1 produces
elongated mitochondrial networks and a senescent pheno-
type but increases ROS production and mtDNA damage
[163,167,168].
The counterpart of fission is fusion, which unifies
mitochondria. Fusion allows mixing of mitochondrial com-
partments, facilitating equilibration of OXPHOS proteins,
energy exchange, and complementation of the mitochondrial
genome [169,170]. In aging, this mixing is thought to
prevent mutations in mtDNA from resulting in respiratory
dysfunction; however, it also permits the accumulation of
mutated mtDNA that might otherwise be removed via fission
and autophagy [170,171]. For these reasons, fusion plays a
role in regulating mtDNA integrity and respiratory function
[172174]. In humans, fusion is known to be controlled by
optic atrophy 1 (Opa1) and the two GTPases mitofusin 1 and
2 (the isoforms Mfn1 and Mfn2). Mfn1 and Mfn2 are located
in the OMM, where they promote tethering and fusion
[162], while Opa1 facilitates fusion from its localization
on the IMM (reviewed in [164]). Opa1 also assists in
regulating degradative processes: it regulates apoptosis by
keeping the inner mitochondrial cristae junctions tight
to prevent cytochrome c release, which triggers apoptosis
[175,176],anditscleavagemaybeinvolvedinflagging
dysfunctional mitochondria for autophagic removal [176].
One study reported that Mfn2 gene expression was lower in
the SM of older humans [21]. Interestingly, muscle-specific
Mfn1- and Mfn2-knockout mice experience enhanced mito-
chondrial proliferation and increased mutations in and
depletion of mtDNA; these changes occur in parallel with
accelerated muscle loss [177]. A mutation in the other key
fusion protein, Opa1, is associated with reduced oxidative
phosphorylation and ATP production in human SM [178].
Thus, age-associated changes in the dynamical remodel-
ing processes of fission and fusion likely aect mtDNA
integrity, respiratory function, ROS production, and cellular
senescence.
3.3. Alterations in Mitochondrial Turnover. Mitochondrial
turnover is executed predominantly by the autophagy-
lysosome system, a cellular housekeeping system that
degrades mitochondria as well as other cellular compo-
nents. Removal of mitochondria through the autophagy-
lysosome system is known as “mitophagy.” In mitophagy,
dysfunctional mitochondria are recognized and engulfed
in a double-membrane structure called a phagophore or
preautophagosome. Once the engulfing process is complete,
vesicles called autophagosomes form. The autophagosomes
then fuse with the lysosome, producing autolysosomes,
and the contents are hydrolytically degraded and recycled
(reviewed in [16,179,180]). The process is mediated through
a collection of several autophagy gene (Atg) products. In
yeast, the recognition process is enacted through Atg32,
a mitophagy-specific receptor on the OMM [181], and
through other key players, such as Atg8 and its activator
Atg7. Mitophagy in mammals has been less well char-
acterized, but the mammalian homologues to Atg32 and
Atg8 are believed to be Nix and LC3, respectively [180,
182]. In addition, mitophagy in mammalian cells may
be carried out through ubiquitination of OMM proteins,
followed by recognition via an LC3 complex (reviewed in
[180]).
Relative to other mitochondrial dynamics, there is less
known about the role of mitophagy in the aging of SM.
Evidence suggests that mitophagy selectively removes defec-
tive mitochondria that are depolarized or produce excessive
ROS (reviewed in [16,17,179,183]). In corroboration,
suppression of autophagy results in increased ROS pro-
duction, reduced oxygen consumption, and higher mtDNA
mutation rates [184,185]. With age, autophagy has been
found to decline both in general (reviewed in [186]) and
in the SM of aged rats [187]. While the repercussions are
still unknown, mitophagy has been negatively correlated
with oxidative damage and apoptosis [187], suggesting
that reduced autophagy rates may contribute to muscular
dysfunction. This is confirmed by studies on muscle-specific
Atg7 knockout mice, who accumulate abnormal mitochon-
dria, have lower resting oxygen consumption, and experience
increased oxidative stress and higher rates of apoptosis;
these mice also suer from muscle atrophy, weakness, and
myofibril degeneration [188,189]. Furthermore, studies in a
few species indicate that enhanced autophagy may increase
lifespan (reviewed in [190]).
Mitochondrial quality control and degradation are also
modulated by the ubiquitin-proteasome system, which
removes oxidized proteins and short-lived proteins (reviewed
in [191]). Evidence from studies in mammals has suggested
that ubiquitin-proteasome activity declines with age in
SM and may contribute to muscular atrophy (reviewed
in [91,192]). However, aging may dierentially aect the
components of the ubiquitin-proteasome system [193], as
some genes involved—including some proteasome subunits
and ubiquitin-specific proteases—are expressed at higher
levels in SM from older humans, yet others are unchanged
or decreased with age [49,84,94,194196]. Finally, changes
in proteasomal activity may be fiber type-specific [197]:
in particular, one study in humans reported that ubiquitin
Journal of Aging Research 9
protein levels increase in fast-twitch muscle fibers, which
may explain why type II fibers atrophy faster with age [198].
3.4. Increased Mitochondria-Mediated Apoptosis. Mitochon-
dria also respond to and modulate cellular dynamics through
apoptotic signaling, which is activated when either OXPHOS
or their redox potential is disrupted, or in response to
proapoptotic signals (reviewed in [199]). Mitochondria
can induce apoptosis through either caspase-dependent
or caspase-independent signaling mechanisms (reviewed
in [199201]). Caspase-dependent signaling is dependent
on the release of cytochrome c from within the mito-
chondria, which triggers a cascade of actions by cysteine
proteases called caspases that results in apoptosis. In brief,
cytochrome c and other proapoptotic factors are released
from the mitochondria. Once released, cytochrome c joins
with apoptotic protease activating factor-1 (Apaf-1) and
procaspase-9 to form a complex called the apoptosome. The
apoptosome then cleaves and activates procaspase-9, which
acts on caspase-3. Caspase-3 in turn activates a caspase-
activated DNase (CAD) to degrade DNA and initiates
cellular degradation. In the caspase-independent pathway,
which is also known as “mitoptosis,” the mitochondria
release apoptosis-inducing factor (AIF) and endonuclease
G (EndoG), which then induce chromatin condensation
and DNA fragmentation. There are multiple points in these
pathways that are regulated by pro- and anti-apoptotic
proteins. In particular, the release of apoptotic triggers
appears to be modulated through two mechanisms: (1)
the balance of proapoptotic (e.g., Bax) and anti-apoptotic
proteins (e.g., Bcl-2), particularly from the Bcl-2 family,
which control OMM stability and form the mitochondrial
apoptosis-induced channel (MAC), and (2) the mitochon-
drial permeability transition pore (mPTP). One example of
the direct connection between mitochondrial and cellular
dynamics is that mitochondrial fragmentation occurs around
the time that proapoptotic factors are released from the
mitochondria, and this step is contingent upon increased
fission through Drp1 as well as a block in mitochondrial
fusion (reviewed in [202]).
Apoptosis increases significantly with age and likely
contributes to sarcopenia and other age-associated declines
(reviewed in [201,203,204]). Though it is dicult to
prove this conclusively, many studies have correlated rates
of apoptosis with markers of sarcopenia or SM function
(reviewed in [201,203,204]). In humans, the percent of
apoptotic cells tends to increase with age, though generally
more so in type II fibers [194,205,206]. The weight of
evidence from both human and animal studies suggests that
the caspase-independent pathway is upregulated with age,
while the caspase-dependent pathway is not (reviewed in
[201]). In particular, one study found an increase in AIF
gene transcripts in SM from older people, but no change in
Bax, Bcl-2, or caspase-3 expression [207], and another study
in humans reported no change in caspase-3 or -7 activity
[194]. This is also supported by animal studies, which have
found that the mPTP becomes more susceptible towards
being opened [25,208,209] and that the levels and activities
of caspase-independent apoptotic proteins AIF and EndoG
increase [25,112,210213] with age. Recent evidence in rats
suggests that apoptotic susceptibilities and markers are age-
and fiber type-specific, which may explain some of the mixed
results, particularly in regard to the caspase-dependent
pathway [25,214216]. Interestingly, a study showed that
disuse atrophy increased caspase-3 activity in young rats but
not old and dramatically increased EndoG levels in old rats
but not young, indicating that older SM likely responds to
apoptotic stimuli through dierent signaling pathways than
younger SM [212].
4. Strategies to Attenuate
Mitochondrial Aging
4.1. Exercise. Exercise training has long been known to
induce mitochondrial biogenesis, upregulate SM gene
expression and protein synthesis, and increase SM oxidative
capacity [217,218]. It is apparent from previous research that
physical activity decreases during aging [219]. Therefore, it
remains unclear whether the abnormalities in mitochondrial
function are a primary eect of aging or are due to
the associated decline in physical activity. This issue is
compounded by the cross-sectional nature of some studies
without objective control for activity level [130,220222].
For example, we recently found that mitochondrial capacity
was reduced in elderly subjects compared to their young
BMI-matched counterparts [10]. In this study, we included
only sedentary individuals, defined as less than 2 hours
of intentional physical activity per week. Despite careful
screening for activity levels using questionnaires, we found
that daily physical activity was significantly lower than
reported in the elderly group. This suggests that independent
of exercise training, simply living an active lifestyle may have
a significant impact on mitochondrial function; however,
we cannot determine the cause and eect due to the cross-
sectional nature of the analysis.
There is strong evidence that exercise training can
improve SM mitochondrial function in elderly adults [84,
96,97,223225] and may also protect against age-associated
apoptosis [226]. Short et al. found that 4 months of
aerobic exercise in older adults increased protein synthe-
sis, mitochondrial enzyme activity (citrate synthase and
cytochrome c oxidase), and expression of genes involved in
mitochondrial content and biogenesis to levels similar to
those in younger adults [97,227]. In a separate study, 12
weeks of aerobic exercise training increased mitochondrial
content and activity, particularly in the subsarcolemmal
fraction [224]. Exercise has also been shown to increase
the activity of antioxidant enzymes and heat shock proteins
[228], potentially reducing ROS production and decreasing
the potential for mitochondrial oxidative damage thought
to occur during aging. Indeed, Parise et al. showed that
resistance exercise in older adults increased antioxidant
content (but not activity), decreased oxidative damage (8-
OHdG), and increased complex IV activity [229,230].
Despite these significant dierences in mitochondrial
parameters, most data show that some impairment remains.
10 Journal of Aging Research
For example, the usual age-associated decline in mitochon-
drial oxidative capacity was absent in older adults who were
chronically endurance trained. However, chronic exercise
did not completely restore the expression of mitochondrial
proteins, mtDNA content, and mitochondrial transcription
factors to the levels of younger subjects, suggesting a
persisting, independent eect of age [50]. Supporting this,
Melov et al. used an “omics” approach to show that after
6 months of exercise training, the transcriptional signature
of aging was substantially but incompletely reversed back to
the transcriptome of younger adults [84]. In all, the results
to date indicate that exercise can help to attenuate age-
associated changes in SM mitochondria. However, it does
not completely prevent these changes. The data available
are rather limited and apply mostly to short-term exercise
interventions (i.e., 12–24 weeks) rather than chronic exercise.
In addition, the impact of active lifestyle changes—for exam-
ple, decreasing sedentary time and increasing “nonexercise”
activity—on mitochondrial function in elderly adults may
be significant [231] and needs further investigation, as this
may be a more logical approach rather than prescribing an
exercise regimen.
4.2. Caloric Restriction. Caloric restriction, which typically
involves consuming 20–40% fewer calories than normal, also
preserves mitochondrial health and attenuates SM decline
with age. Caloric restriction (CR) is recognized as the most
robust intervention that retards both primary aging (natural
age-related deterioration) and secondary aging (accelerated
aging due to disease and negative lifestyle behaviors), thereby
increasing both median and maximal lifespan in many
species. While CR studies in primates and humans are
largely ongoing, studies in rodents consistently show that CR
extends maximum life span by up to 50% and reduces the
incidence of many age-associated diseases, including cancer
and metabolic diseases (reviewed in [232]). Preliminary
evidence in rhesus monkeys indicates similar eects can be
expected in primates [231].
The benefits ascribed to CR are believed to be due in large
part to reductions in oxidative stress (reviewed in [233]). In
primates, a decade-long CR intervention resulted in marked
decreases in oxidative damage to lipids and proteins [234]; in
aged rats, CR also reduces ROS production [159,235,236].
As a result, aged CR animals exhibit fewer mtDNA and
nuclear DNA mutations and less oxidative damage to SM
mitochondria than their ad libitum-fed counterparts [99,
236240]. Some of these findings have been replicated in the
still ongoing CR trial in humans, dubbed CALERIE, which
reported that CR subjects had less mtDNA damage and
more mtDNA content than controls [241]. Microarray and
RT-PCR experiments confirm that CR increases transcripts
of genes involved in ROS scavenging, including SOD and
GPx, and decreases transcripts from stress response genes
[196,242].
Calorie restriction appears to modulate mitochondrial
eciency, content, and function. CR lowers energy expendi-
ture in animals and humans by producing mitochondria that
consumelessoxygenyetareabletomaintainnormallevels
of ATP production (reviewed in [232]; in particular, [159,
237,241243]). It is generally believed that this energetic
adaptation is mediated via decreased proton leak, which
has been confirmed in rodent studies, and that decreased
proton leak is in turn enabled by the shift to a less oxidative
milieu [159,237,244]. In terms of mitochondrial content
and function, CR does not aect the gene expression, protein
level, or activity of citrate synthase, nor the activities of other
TCA proteins [111,112,241,242,245]. However, CR may
aect some ETS enzymes. CR reduced the age-associated
accumulation of complex IV-negative and complex II-
hyperactive fibers in rats [99] and in rhesus monkeys [100].
Only complex IV activity, however, is consistently responsive
to CR, showing increased activity in comparison to that in
aged rodents and usually similar activity to their younger
counterparts [110112,246]. Nonetheless, CR may prevent
the decline in the activities of complexes I–III in some
muscles [111]. However, a study in rhesus monkeys found
that about 9 years of CR only attenuated the decline in
gene expression of the complex II iron-protein subunit [85],
and a 14-week CR intervention where the rats were fed
at 70% of ad libitum levels found no increases in any
mitochondrial gene transcripts or proteins [245]. CR rodents
did have fewer ETS abnormalities, but the abnormalities
were otherwise similar to controls, suggesting that CR only
aects the onset of fiber atrophy [240,247]. However, in
primate studies, CR did not alter the number of fibers with
ETS abnormalities, but nonetheless preserved muscle fiber
[100,248].
CR also aects mitochondrial dynamics. In rodent
SM, CR also increases mitochondrial biogenesis relative to
controls by slowing the decline in PGC-1αgene expression
with age, which may be at least partially responsible for
maintaining oxidative capacity in aged CR animals [110,
112]. The CALERIE study in humans also reported increased
transcripts from several genes involved in mitochondrial
biogenesis, including PGC-1α, Tfam, and SIRT1 [241]. It
is not clear, however, whether CR aects the mitochondrial
metabolic regulator AMPK, as there have been mixed results
[249,250]. In addition, CR attenuated the decline in
mitophagy in the SM of old rodents [187] and the decline
in ubiquitin-proteasome activity in monkeys [251]. It also
reduced apoptosis susceptibility by promoting a remodeling
of caspases and caspase-related proteins to favor decreased
likelihood of cytochrome c release from the mitochondria
[187,252254]. These findings are particularly important
as mitochondrial dysfunction and apoptosis have been
proposed as key mediators of sarcopenia.
Combined, these results suggest that CR reduces oxida-
tive stress and remodels mitochondrial dynamics to promote
the production of more fuel-ecient mitochondria that
producelessROSandtofavortheremovalofdysfunctional
mitochondria [187,243,255]. Regardless of the underlying
mechanism, CR has repeatedly been shown in rodents to
either partially or fully oppose the hallmarks of sarcopenia,
including the age-related declines in muscle force, fiber cross-
sectional area, fiber number, and fiber type (reviewed in
[99,247,254,256258]). Similar findings have been reported
in CR studies in rhesus monkeys [100,248].
Journal of Aging Research 11
4.3. Mimetics. Currently, there is strong interest in using a
CR- or exercise-mimetic to attenuate age-related mitochon-
drial dysfunction. The best known of these is resveratrol
(3,5,4,9-trihydroxystilbene), a phytoalexin that is abundant
in red wine. In the context of secondary aging, resver-
atrol’s eects on mitochondria and SM have been well
studied. In rodent models involving metabolic pathology,
resveratrol improves mtDNA copy number and function,
increases mitochondrial biogenesis, improves exercise capac-
ity and motor function, and mitigates metabolic dysfunction
[259261]. On a molecular level, resveratrol induces an
increase in the expression of PGC-1α,Tfam,andUCP3,
and increases SIRT1, AMPK, and PGC-1αactivation [259,
260]. In humans, a recent 30-day study of resveratrol
supplementation in 11 obese but metabolically healthy men
reported improved mitochondrial respiration in the presence
of fatty acid-derived substrate, increased AMPK and citrate
synthase activity, and higher SIRT1 and PGC-1αprotein
levels; however, no changes in mitochondrial content were
observed [262]. Interestingly, while resveratrol was originally
identified as a potent activator of SIRT1, there is controversy
surrounding its mode of action, and it may exert much of its
eects through AMPK [260,263,264].
Rodent studies consistently show improved mitochon-
drial health with resveratrol (reviewed in [265]). For
example, in senescence-accelerated prone mice, those given
resveratrol had higher physical endurance, maximal force
contraction, and oxygen consumption, and higher levels
of transcripts from PGC-1αand ETS genes [266]. Aged
rats given resveratrol exhibited similar responses; they also
displayed reduced levels of oxidative stress but no changes
in most apoptotic markers, indicating the improvements
were mediated through changes in redox status and not
apoptotic pathways [267]. Like CR, resveratrol’s eects on
mitochondrial biogenesis are believed to be mediated in
large part via reductions in mitochondrial ROS production
and via upregulation of fatty acid catabolism coupled with
a downregulation in fatty acid synthesis [137]. However,
one recent study of long-term resveratrol supplementation
in mice did not find any improvement in the age-related
declines in PGC-1α, mitochondrial content, muscle mass,
and maximal isometric force production; though resveratrol
did preserve type II fiber contractile function and reduce
oxidative stress [268].
5. Conclusions
In summary, animal and human data consistently show that
skeletal muscle mitochondria are altered in aging, includ-
ing increased mutations in mitochondrial DNA, decreased
expression of some mitochondrial proteins, reduced enzyme
activity and altered respiration with reduced maximal capac-
ity in sedentary adults, and reduced total mitochondrial
content with increased morphological changes. Since the
primary role of mitochondria is to produce ATP to maintain
the energy status of the cell, shifts in respiratory activity and
capacity can lower the membrane potential, reduce cellular
ATP concentration, and signal cellular apoptotic events.
Increased apoptosis without correspondingly increased pro-
tein synthesis will eventually lead to net muscle fiber loss.
All of these factors probably contribute to age-associated
sarcopenia, and mounting evidence suggests that most
of these age-related changes can be either prevented or
attenuated through increased physical activity.
There is some thought that the accumulation of oxidative
damage caused by long-term ROS production is responsible
for these changes with age. Indeed, blocking ROS formation
by the targeted expression of antioxidant enzymes amelio-
rates age-associated dysfunction and returns mitochondrial
parameters to those of young animals (Figure 3). Whether
this is also true for humans is unknown; however, strategies
that improve mitochondrial function, such as exercise and
caloric restriction, also reduce ROS production and increase
antioxidant defenses. Exercise is also known to be a powerful
stimulant of mitochondrial biogenesis and oxidative capac-
ity, although exercise training in elderly adults probably does
not completely reverse the primary eects of aging. However,
exercise training—even adopting active lifestyle habits—
may clearly reduce the rate of mitochondrial decline and
attenuate the aging phenotype. Whether CR- and exercise-
mimetics, including resveratrol, work as eectively remains
to be determined and is an area of active research.
In conclusion, there is clearly a need for more research in
this field and particularly to:
(i) figure out which age-related changes are universal
and which depend on physical activity or lifestyle
habits or gender;
(ii) unravel how cells compensate for oxidative damage
and mitochondrial dysfunction;
(iii) elucidate how fusion, fission, and autophagy work
together to ensure quality control and to remove
defective mitochondria; and
(iv) determine which mitochondrial declines can be
slowed down (or even reversed) by healthier lifestyles.
We expect this to continue be a fruitful and exciting area of
study in years to come.
References
[1] W.R.Frontera,V.A.Hughes,K.J.Lutz,andW.J.Evans,“A
cross-sectional study of muscle strength and mass in 45- to
78-yr-old men and women,Journal of Applied Physiology,
vol. 71, no. 2, pp. 644–650, 1991.
[2] V. A. Hughes, W. R. Frontera, M. Wood et al., “Longitudinal
muscle strength changes in older adults: influence of muscle
mass, physical activity, and health,The Journals of Gerontol-
ogy A, vol. 56, no. 5, pp. B209–B217, 2001.
[3] B.H.Goodpaster,S.W.Park,T.B.Harrisetal.,“Theloss
of skeletal muscle strength, mass, and quality in older adults:
the Health, Aging and Body Composition Study,Journals of
Gerontology A, vol. 61, no. 10, pp. 1059–1064, 2006.
[4] M.A.Fiatarone,E.F.ONeill,N.D.Ryan,K.M.Clements,
G. R. Solares, and M. E. Nelson, “Exercise training and
nutritional supplementation for physical frailty in very
elderly people,The New England journal of medicine, vol.
330, no. 25, pp. 1769–1775, 1994.
12 Journal of Aging Research
[5] B. H. Goodpaster, P. Chomentowski, B. K. Ward et al.,
“Eects of physical activity on strength and skeletal muscle
fat infiltration in older adults: a randomized controlled trial,
Journal of Applied Physiology, vol. 105, no. 5, pp. 1498–1503,
2008.
[6] M. E. Nelson, M. A. Fiatarone, C. M. Morganti, I. Trice,
R.A.Greenberg,andW.J.Evans,“Eects of high-intensity
strength training on multiple risk factors for osteoporotic
fractures: a randomized controlled trial,Journal of the
American Medical Association, vol. 272, no. 24, pp. 1909–
1914, 1994.
[7] E. J. Metter, N. Lynch, R. Conwit, R. Lindle, J. Tobin,
and B. Hurley, “Muscle quality and age: cross-sectional and
longitudinal comparisons,The Journals of Gerontology A,
vol. 54, no. 5, pp. B207–B218, 1999.
[8] M.G.Cree,B.R.Newcomer,C.S.Katsanosetal.,“Intramus-
cular and liver triglycerides are increased in the elderly,The
Journal of Clinical Endocrinology and Metabolism, vol. 89, no.
8, pp. 3864–3871, 2004.
[9] A. R. Hipkiss, “Mitochondrial dysfunction, proteotoxicity,
and aging: causes or eects, and the possible impact of
NAD+-controlled protein glycation,Advances in Clinical
Chemistry, vol. 50, pp. 123–150, 2010.
[10] D. L. Johannsen, K. E. Conley, S. Bajpeyi et al., “Ectopic
lipid accumulation and reduced glucose tolerance in elderly
adults are accompanied by altered skeletal muscle mito-
chondrial activity,The Journal of Clinical Endocrinology and
Metabolism, vol. 97, no. 1, pp. 242–250, 2012.
[11] M. Saraste, “Oxidative phosphorylation at the fin de siecle,
Science, vol. 283, no. 5407, pp. 1488–1493, 1999.
[12] J. O. Holloszy, “Skeletal muscle,“mitochondrial deficiency”
does not mediate insulin resistance,The American Journal
of Clinical Nutrition, vol. 89, no. 1, pp. 463S–466S, 2009.
[13] B. F. Miller, M. M. Robinson, M. D. Bruss, M. Hellerstein, and
K. L. Hamilton, “A comprehensive assessment of mitochon-
drial protein synthesis and cellular proliferation with age and
caloric restriction,Aging Cell, vol. 11, no. 1, pp. 150–161,
2012.
[14] M. K. Shigenaga, T. M. Hagen, and B. N. Ames, “Oxidative
damage and mitochondrial decay in aging,Proceedings of the
National Academy of Sciences of the United States of America,
vol. 91, no. 23, pp. 10771–10778, 1994.
[15] A. Terman, T. Kurz, M. Navratil, E. A. Arriaga, and U.
T. Brunk, “Mitochondrial turnover and aging of long-lived
postmitotic cells: the mitochondrial-lysosomal axis theory of
aging,Antioxidants and Redox Signaling,vol.12,no.4,pp.
503–535, 2010.
[16] G. Twig, B. Hyde, and O. S. Shirihai, “Mitochondrial
fusion, fission and autophagy as a quality control axis: the
bioenergetic view,Biochimica et Biophysica Acta, vol. 1777,
no. 9, pp. 1092–1097, 2008.
[17] G. Twig, A. Elorza, A. J. Molina et al., “Fission and selective
fusion govern mitochondrial segregation and elimination by
autophagy,The EMBO Journal, vol. 27, no. 2, pp. 433–446,
2008.
[18] E. Beregi, O. Regius, T. Huttl, and Z. Gobl, “Age-related
changes in the skeletal muscle cells,Zeitschrift fur Gerontolo-
gie, vol. 21, no. 2, pp. 83–86, 1988.
[19] P. Poggi, C. Marchetti, and R. Scelsi, “Automatic morpho-
metric analysis of skeletal muscle fibers in the aging man,
Anatomical Record, vol. 217, no. 1, pp. 30–34, 1987.
[20] K. E. Conley, S. A. Jubrias, and P. C. Esselman, “Oxidative
capacity and ageing in human muscle,The Journal of
Physiology, vol. 5261, part 1, pp. 203–210, 2000.
[21] J. D. Crane, M. C. Devries, A. Safdar, M. J. Hamadeh,
and M. A. Tarnopolsky, “The eect of aging on human
skeletal muscle mitochondrial and intramyocellular lipid
ultrastructure,The Journals of Gerontology A,vol.65,no.2,
pp. 119–128, 2010.
[22] S. Nakamura, T. Takamura, N. Matsuzawa-Nagata et al.,
“Palmitate induces insulin resistance in H4IIEC3 hepatocytes
through reactive oxygen species produced by mitochondria,
The Journal of Biological Chemistry, vol. 284, no. 22, pp.
14809–14818, 2009.
[23] D. Harman, “Aging: a theory based on free radical and radi-
ation chemistry.,Journal of Gerontology, vol. 11, no. 3, pp.
298–300, 1956.
[24] D. Harman, “Free radical theory of aging: an update: increas-
ing the functional life span,Annals of the New York Academy
of Sciences, vol. 1067, pp. 10–21, 2006.
[25] B. Chabi, V. Ljubicic, K. J. Menzies, J. H. Huang, A. Saleem,
and D. A. Hood, “Mitochondrial function and apoptotic
susceptibility in aging skeletal muscle,Aging Cell, vol. 7, no.
1, pp. 2–12, 2008.
[26] E. H¨
utter, M. Skovbro, B. Lener et al., “Oxidative stress and
mitochondrial impairment can be separated from lipofuscin
accumulation in aged human skeletal muscle,Aging Cell, vol.
6, no. 2, pp. 245–256, 2007.
[27] A. Mansouri, F. L. Muller, Y. Liu et al., “Alterations in mito-
chondrial function, hydrogen peroxide release and oxidative
damage in mouse hind-limb skeletal muscle during aging,
Mechanisms of Ageing and Development, vol. 127, no. 3, pp.
298–306, 2006.
[28] C. S. Yarian, I. Rebrin, and R. S. Sohal, “Aconitase and
ATP synthase are targets of malondialdehyde modification
and undergo an age-related decrease in activity in mouse
heart mitochondria,Biochemical and Biophysical Research
Communications, vol. 330, no. 1, pp. 151–156, 2005.
[29] V. Pesce, A. Cormio, F. Fracasso et al., “Age-related mitochon-
drial genotypic and phenotypic alterations in human skeletal
muscle,” Free Radical Biology and Medicine, vol. 30, no. 11,
pp. 1223–1233, 2001.
[30] M. F. Beal, “Oxidatively modified proteins in aging and
disease,Free Radical Biology & Medicine,vol.32,no.9,pp.
797–803, 2002.
[31] L. Staunton, K. O’Connell, and K. Ohlendieck, “Proteomic
profiling of mitochondrial enzymes during skeletal muscle
aging,Journal of Aging Research, vol. 2011, Article ID
908035, 9 pages, 2011.
[32] Y. Kushnareva, A. N. Murphy, and A. Andreyev, “Complex I-
mediated reactive oxygen species generation: modulation by
cytochrome c and NAD(P)+oxidation-reduction state,The
Biochemical Journal, vol. 368, part 2, pp. 545–553, 2002.
[33] E.Barreiro,C.Coronell,B.Lavi
˜
na, A. Ram´
ırez-Sarmiento,
M. Orozco-Levi, and J. Gea, “Aging, sex dierences, and
oxidative stress in human respiratory and limb muscles,Free
Radical Biology and Medicine, vol. 41, no. 5, pp. 797–809,
2006.
[34] L. L. Ji, D. Dillon, and E. Wu, “Alteration of antioxidant
enzymes with aging in rat skeletal muscle and liver,The
American Journal of Physiology, vol. 258, no. 4, part 2, pp.
918–923, 1990.
[35] T. A. Luhtala, E. B. Roecker, T. Pugh, R. J. Feuers, and
R. Weindruch, “Dietary restriction attenuates age-related
increases in rat skeletal muscle antioxidant enzyme activi-
ties,Journals of Gerontology, vol. 49, no. 5, pp. B231–B238,
1994.
Journal of Aging Research 13
[36] T. Armeni, G. Principato, J. L. Quiles, C. Pieri, S. Bompadre,
and M. Battino, “Mitochondrial dysfunctions during aging:
vitamin E deficiency or caloric restriction—two dierent
ways of modulating stress,Journal of Bioenergetics and
Biomembranes, vol. 35, no. 2, pp. 181–191, 2003.
[37] J. Ren, Q. Li, S. Wu, S. Y. Li, and S. A. Babcock, “Cardiac
overexpression of antioxidant catalase attenuates aging-
induced cardiomyocyte relaxation dysfunction,Mechanisms
of Ageing and Development, vol. 128, no. 3, pp. 276–285, 2007.
[38] E. Xia, G. Rao, H. Van Remmen, A. R. Heydari, and A.
Richardson, “Activities of antioxidant enzymes in various tis-
sues of male Fischer 344 rats are altered by food restriction,
Journal of Nutrition, vol. 125, no. 2, pp. 195–201, 1995.
[39] A.Safdar,M.J.Hamadeh,J.J.Kaczor,S.Raha,J.Debeer,and
M. A. Tarnopolsky, “Aberrant mitochondrial homeostasis in
the skeletal muscle of sedentary older adults,PLoS ONE, vol.
5, no. 5, p. e10778, 2010.
[40]H.Y.Lee,C.S.Choi,A.L.Birkenfeld,T.C.Alves,F.R.
Jornayvaz, and M. J. Jurczak, “Targeted expression of catalase
to mitochondria prevents age-associated reductions in mito-
chondrial function and insulin resistance,Cell Metabolism,
vol. 12, no. 6, pp. 668–674, 2010.
[41] S. Iossa, M. P. Mollica, L. Lionetti, R. Crescenzo, R. Tasso,
and G. Liverini, “A possible link between skeletal muscle
mitochondrial eciency and age-induced insulin resistance,
Diabetes, vol. 53, no. 11, pp. 2861–2866, 2004.
[42] D. J. Marcinek, K. A. Schenkman, W. A. Ciesielski, D. Lee,
and K. E. Conley, “Reduced mitochondrial coupling in vivo
alters cellular energetics in aged mouse skeletal muscle,The
Journal of Physiology, vol. 569, part 2, pp. 467–473, 2005.
[43] C. E. Amara, E. G. Shankland, S. A. Jubrias, D. J. Marcinek,
M. J. Kushmerick, and K. E. Conley, “Mild mitochondrial
uncoupling impacts cellular aging in human muscles in vivo,”
Proceedings of the National Academy of Sciences of the United
States of America, vol. 104, no. 3, pp. 1057–1062, 2007.
[44] M. F. Lopez, B. S. Kristal, E. Chernokalskaya, A. Lazarev,
A. I. Shestopalov, A. Bogdanova et al., “High-throughput
profiling of the mitochondrial proteome using anity frac-
tionation and automation,Electrophoresis, vol. 21, no. 16,
pp. 3427–3440, 2000.
[45] S. Calvo, M. Jain, X. Xie, S. A. Sheth, B. Chang, O.
A. Goldberger et al., “Systematic identification of human
mitochondrial disease genes through integrative genomics,
Nature Genetics, vol. 38, no. 5, pp. 576–582, 2006.
[46]S.Anderson,A.T.Bankier,B.G.Barrell,M.H.deBruijn,
A. R. Coulson, J. Drouin et al., “Sequence and organization
of the human mitochondrial genome,Nature, vol. 290, no.
5806, pp. 457–465, 1981.
[47] E. V. Menshikova, V. B. Ritov, L. Fairfull, R. E. Ferrell, D.
E. Kelley, and B. H. Goodpaster, “Eects of exercise on
mitochondrial content and function in aging human skeletal
muscle,” Journals of Gerontology A, vol. 61, no. 6, pp. 534–540,
2006.
[48] K. R. Short, M. L. Bigelow, J. Kahl, R. Singh, J. Coenen-
Schimke, S. Raghavakaimal et al., “Decline in skeletal muscle
mitochondrial function with aging in humans. .,Proceedings
of the National Academy of Sciences of the United States of
America, vol. 102, no. 15, pp. 5618–5623, 2005.
[49] S. Welle, K. Bhatt, B. Shah, N. Needler, J. M. Delehanty, and
C. A. Thornton, “Reduced amount of mitochondrial DNA in
aged human muscle,Journal of Applied Physiology, vol. 94,
no. 4, pp. 1479–1484, 2003.
[50] I. R. Lanza, D. K. Short, K. R. Short, S. Raghavakaimal, R.
Basu, M. J. Joyner et al., “Endurance exercise as a counter-
measure for aging,Diabetes, vol. 57, no. 11, pp. 2933–2942,
2008.
[51] A. Barrientos, J. Casademont, F. Cardellach et al., “Qualita-
tive and quantitative changes in skeletal muscle mtDNA and
expression of mitochondrial-encoded genes in the human
aging process,Biochemical and Molecular Medicine, vol. 62,
no. 2, pp. 165–171, 1997.
[52] R. Barazzoni, K. R. Short, and K. S. Nair, “Eects of aging
on mitochondrial DNA copy number and cytochrome c
oxidase gene expression in rat skeletal muscle, liver, and
heart,Journal of Biological Chemistry, vol. 275, no. 5, pp.
3343–3347, 2000.
[53] D.McKenzie,E.Bua,S.McKiernan,Z.Cao,J.Wanagat,andJ.
M. Aiken, “Mitochondrial DNA deletion mutations: a causal
role in sarcopenia,European Journal of Biochemistry, vol.
269, no. 8, pp. 2010–2015, 2002.
[54] K. Khrapko and J. Vijg, “Mitochondrial DNA mutations and
aging: devils in the details?” Tren ds i n Ge ne ti cs, vol. 25, no. 2,
pp. 91–98, 2009.
[55] E. Bua, J. Johnson, A. Herbst et al., “Mitochondrial DNA-
deletion mutations accumulate intracellularly to detrimental
levels in aged human skeletal muscle fibers,American
Journal of Human Genetics, vol. 79, no. 3, pp. 469–480, 2006.
[56] B. Chabi, B. Mousson de Camaret, A. Chevrollier, S.
Boisgard, and G. Stepien, “Random mtDNA deletions and
functional consequence in aged human skeletal muscle,
Biochemical and Biophysical Research Communications, vol.
332, no. 2, pp. 542–549, 2005.
[57] S. Melov, J. M. Shoner, A. Kaufman, and D. C. Wallace,
“Marked increase in the number and variety of mitochon-
drial DNA rearrangements in aging human skeletal muscle,
Nucleic Acids Research, vol. 23, no. 20, pp. 4122–4126, 1995.
[58] J. M. Cooper, V. M. Mann, and A. H. V. Schapira, “Analyses of
mitochondrial respiratory chain function and mitochondrial
DNA deletion in human skeletal muscle: eect of ageing,
Journal of the Neurological Sciences, vol. 113, no. 1, pp. 91–
98, 1992.
[59] F. Pallotti, X. Chen, E. Bonilla, and E. A. Schon, “Evidence
that specific mtDNA point mutations may not accumulate
in skeletal muscle during normal human aging,American
Journal of Human Genetics, vol. 59, no. 3, pp. 591–602, 1996.
[60] Y. Wang, Y. Michikawa, C. Mallidis, Y. Bai, L. Woodhouse, K.
E. Yarasheski et al., “Muscle-specific mutations accumulate
with aging in critical human mtDNA control sites for
replication,Proceedings of the National Academy of Sciences
of the United States of America, vol. 98, no. 7, pp. 4022–4027,
2001.
[61] Y. Michikawa, F. Mazzucchelli, N. Bresolin, G. Scarlato,
and G. Attardi, “Aging-dependent large accumulation of
point mutations in the human mtDNA control region for
replication,Science, vol. 286, no. 5440, pp. 774–779, 1999.
[62] J. Wanagat, Z. Cao, P. Pathare, and J. M. Aiken, “Mito-
chondrial DNA deletion mutations colocalize with segmental
electron transport system abnormalities, muscle fiber atro-
phy, fiber splitting, and oxidative damage in sarcopenia,The
FASEB Journal, vol. 15, no. 2, pp. 322–332, 2001.
[63] K. Torii, S. Sugiyama, M. Tanaka, K. Takagi, Y. Hanaki, K. Iida
et al., “Aging-associated deletions of human diaphragmatic
mitochondrial DNA,American Journal of Respiratory Cell
and Molecular Biology, vol. 6, no. 5, pp. 543–549, 1992.
14 Journal of Aging Research
[64] P. Fattoretti, J. Vecchiet, G. Felzani et al., “Succinic dehydro-
genase activity in human muscle mitochondria during aging:
a quantitative cytochemical investigation,Mechanisms of
Ageing and Development, vol. 122, no. 15, pp. 1841–1848,
2001.
[65] V. A. Bohr, T. Stevnsner, and N. C. de Souza-Pinto, “Mito-
chondrial DNA repair of oxidative damage in mammalian
cells,” Gene, vol. 286, no. 1, pp. 127–134, 2002.
[66] F. M. Yakes and B. Van Houten, “Mitochondrial DNA
damage is more extensive and persists longer than nuclear
DNA damage in human cells following oxidative stress,
Proceedings of the National Academy of Sciences of the United
States of America, vol. 94, no. 2, pp. 514–519, 1997.
[67] N. G. Larsson, “Somatic mitochondrial DNA mutations in
mammalian aging,Annual Review of Biochemistry, vol. 79,
pp. 683–706, 2010.
[68] C. Meissner, “Mutations of mitochondrial DNA—cause or
consequence of the ageing process?” ZGerontolGeriatr, vol.
40, no. 5, pp. 325–333, 2007.
[69] A. Hiona and C. Leeuwenburgh, “The role of mitochondrial
DNA mutations in aging and sarcopenia: implications for the
mitochondrial vicious cycle theory of aging,” Experimental
Gerontology, vol. 43, no. 1, pp. 24–33, 2008.
[70] K. E. Conley, D. J. Marcinek, and J. Villarin, “Mitochondrial
dysfunction and age,Current Opinion in Clinical Nutrition
and Metabolic Care, vol. 10, no. 6, pp. 688–692, 2007.
[71] E.J.Brierley,M.A.Johnson,R.N.Lightowlers,O.F.James,
and D. M. Turnbull, “Role of mitochondrial DNA mutations
in human aging: implications for the central nervous system
and muscle,Annals of Neurology, vol. 43, no. 2, pp. 217–223,
1998.
[72] V. Pesce, A. Cormio, F. Fracasso et al., “Age-related mitochon-
drial genotypic and phenotypic alterations in human skeletal
muscle,” Free Radical Biology and Medicine, vol. 30, no. 11,
pp. 1223–1233, 2001.
[73] G. Fayet, M. Jansson, D. Sternberg et al., “Ageing muscle:
clonal expansions of mitochondrial DNA point mutations
and deletions cause focal impairment of mitochondrial
function,Neuromuscular Disorders, vol. 12, no. 5, pp. 484–
493, 2002.
[74] H. C. Lee, C. Y. Pang, H. S. Hsu, and Y. H. Wei, “Ageing-
associated tandem duplications in the D-loop of mitochon-
drial DNA of human muscle,FEBS Letters, vol. 354, no. 1,
pp. 79–83, 1994.
[75]O.E.Rooyackers,D.B.Adey,P.A.Ades,andK.S.Nair,
“Eect of age on in vivo rates of mitochondrial protein syn-
thesis in human skeletal muscle,Proceedings of the National
Academy of Sciences of the United States of America, vol. 93,
no. 26, pp. 15364–15369, 1996.
[76] A. Trifunovic, A. Wredenberg, M. Falkenberg et al., “Prema-
ture ageing in mice expressing defective mitochondrial DNA
polymerase,Nature, vol. 429, no. 6990, pp. 417–423, 2004.
[77] G. C. Kujoth, A. Hiona, T. D. Pugh et al., “Medicine: mito-
chondrial DNA mutations, oxidative stress, and apoptosis in
mammalian aging,Science, vol. 309, no. 5733, pp. 481–484,
2005.
[78] A. Hiona, A. Sanz, G. C. Kujoth et al., “Mitochondrial DNA
mutations induce mitochondrial dysfunction, apoptosis and
sarcopenia in skeletal muscle of mitochondrial DNA mutator
mice,PLoS ONE, vol. 5, no. 7, Article ID e11468, 2010.
[79] D. Booli, S. C. Scacco, R. Vergari, G. Solarino, G. Santacroce,
and S. Papa, “Decline with age of the respiratory chain
activity in human skeletal muscle,Biochimica et Biophysica
Acta, vol. 1226, no. 1, pp. 73–82, 1994.
[80] S. Welle, K. Bhatt, and C. A. Thornton, “High-abundance
mRNAs in human muscle: comparison between young and
old,Journal of Applied Physiology, vol. 89, no. 1, pp. 297–304,
2000.
[81] S. Welle, A. I. Brooks, J. M. Delehanty et al., “Skeletal muscle
gene expression profiles in 20–29 year old and 65–71 year old
women,Experimental Gerontology, vol. 39, no. 3, pp. 369–
377, 2004.
[82]J.M.Zahn,R.Sonu,H.Vogel,E.Crane,K.Mazan-
Mamczarz, and R. Rabkin, “Transcriptional profiling of aging
in human muscle reveals a common aging signature,PLoS
Genetics, vol. 2, no. 7, p. e115, 2006.
[83] S. Welle, A. I. Brooks, J. M. Delehanty, N. Needler, and C.
A. Thornton, “Gene expression profile of aging in human
muscle,” Physiol Genomics, vol. 14, no. 2, pp. 149–159, 2003.
[84] S. Melov, M. A. Tarnopolsky, K. Beckman, K. Felkey, and
A. Hubbard, “Resistance exercise reverses aging in human
skeletal muscle,PLoS ONE, vol. 2, no. 5, p. e465, 2007.
[85] T. Kayo, D. B. Allison, R. Weindruch, and T. A. Prolla, “Influ-
ences of aging and caloric restriction on the transcriptional
profile of skeletal muscle from rhesus monkeys,Proceedings
of the National Academy of Sciences of the United States of
America, vol. 98, no. 9, pp. 5093–5098, 2001.
[86] M. Picard, D. Ritchie, K. J. Wright et al., “Mitochondrial
functional impairment with aging is exaggerated in isolated
mitochondria compared to permeabilized myofibers,Aging
Cell, vol. 9, no. 6, pp. 1032–1046, 2010.
[87] J. P. de Magalhaes, J. Curado, and G. M. Church, “Meta-
analysis of age-related gene expression profiles identifies
common signatures of aging,Bioinformatics, vol. 25, no. 7,
pp. 875–881, 2009.
[88] C. N. Lyons, O. Mathieu-Costello, and C. D. Moyes, “Regula-
tion of skeletal muscle mitochondrial content during aging,
Journals of Gerontology A, vol. 61, no. 1, pp. 3–13, 2006.
[89] P. G. Giresi, E. J. Stevenson, J. Theilhaber et al., “Identifi-
cation of a molecular signature of sarcopenia,Physiological
Genomics, vol. 21, pp. 253–263, 2005.
[90] V. K. Mootha, J. Bunkenborg, J. V. Olsen et al., “Integrated
analysis of protein composition, tissue diversity, and gene
regulation in mouse mitochondria,Cell, vol. 115, no. 5, pp.
629–640, 2003.
[91] L. Combaret, D. Dardevet, D. B´
echet, D. Taillandier, L.
Mosoni, and D. Attaix, “Skeletal muscle proteolysis in aging,
Current Opinion in Clinical Nutrition and Metabolic Care, vol.
12, no. 1, pp. 37–41, 2009.
[92] B. A. Irving, M. M. Robinson, and K. S. Nair, “Age eect on
myocellular remodeling: response to exercise and nutrition
in humans,Ageing Research Reviews, vol. 11, no. 3, pp. 374–
389, 2012.
[93] P. Gianni, K. J. Jan, M. J. Douglas, P. M. Stuart, and M.
A. Tarnopolsky, “Oxidative stress and the mitochondrial
theory of aging in human skeletal muscle,Experimental
Gerontology, vol. 39, no. 9, pp. 1391–400, 2004.
[94] K. O’Connell and K. Ohlendieck, “Proteomic DIGE analysis
of the mitochondria-enriched fraction from aged rat skeletal
muscle,” Proteomics, vol. 9, no. 24, pp. 5509–5524, 2009.
[95] I. Piec, A. Listrat, J. Alliot, C. Chambon, R. G. Taylor, and D.
Bechet, “Dierential proteome analysis of aging in rat skeletal
muscle,” The FASEB Journal, vol. 19, no. 9, pp. 1143–1145,
2005.
[96] S. Ghosh, R. Lertwattanarak, N. Lefort, M. Molina-Carrion,
J. Joya-Galeana, and B. P. Bowen, “Reduction in reactive
oxygen species production by mitochondria from elderly
Journal of Aging Research 15
subjects with normal and impaired glucose tolerance,Dia-
betes, vol. 60, no. 8, pp. 2051–2060, 2011.
[97] K. R. Short, J. L. Vittone, M. L. Bigelow et al., “Impact of
aerobic exercise training on age-related changes in insulin
sensitivity and muscle oxidative capacity,Diabetes, vol. 52,
no. 8, pp. 1888–1896, 2003.
[98] R. M. P. Alves, R. Vitorino, P. Figueiredo, J. A. Duarte, R.
Ferreira, and F. Amado, “Lifelong physical activity modula-
tion of the skeletal muscle mitochondrial proteome in mice,
Journals of Gerontology A, vol. 65, no. 8, pp. 832–842, 2010.
[99] L. E. Aspnes, C. M. Lee, R. Weindruch, S. S. Chung, E. B.
Roecker, and J. M. Aiken, “Caloric restriction reduces fiber
loss and mitochondrial abnormalities in aged rat muscle,
The FASEB Journal, vol. 11, no. 7, pp. 573–581, 1997.
[100] S. H. McKiernan, R. J. Colman, E. Aiken et al., “Cel-
lular adaptation contributes to calorie restriction-induced
preservation of skeletal muscle in aged rhesus monkeys,
Experimental Gerontology, vol. 47, no. 3, pp. 229–236, 2012.
[101] E. J. Brierley, M. A. Johnson, O. F. James, and D. M. Turnbull,
“Eects of physical activity and age on mitochondrial
function,Monthly Journal of the Association of Physicians,
vol. 89, no. 4, pp. 251–258, 1996.
[102] C. Gelfi, A. Vigano, M. Ripamonti et al., “The human muscle
proteome in aging,Journal of Proteome Research, vol. 5, no.
6, pp. 1344–1353, 2006.
[103] P. Donoghue, L. Staunton, E. Mullen, G. Manning, and K.
Ohlendieck, “DIGE analysis of rat skeletal muscle proteins
using nonionic detergent phase extraction of young adult
versus aged gastrocnemius tissue,Journal of Proteomics, vol.
73, no. 8, pp. 1441–1453, 2010.
[104] M. Tonkonogi, M. Fernstr¨
om, B. Walsh et al., “Reduced
oxidative power but unchanged antioxidative capacity in
skeletal muscle from aged humans,Pflugers Archiv European
Journal of Physiology, vol. 446, no. 2, pp. 261–269, 2003.
[105] A. R. Coggan, R. J. Spina, D. S. King et al., “Histochemical
and enzymatic comparison of the gastrocnemius muscle of
young and elderly men and women,Journals of Gerontology,
vol. 47, no. 3, pp. B71–B76, 1992.
[106] G. Lenaz, C. Bovina, C. Castelluccio et al., “Mitochondrial
complex I defects in aging,Molecular and Cellular Biochem-
istry, vol. 174, no. 1-2, pp. 329–333, 1997.
[107] S. Kumaran, M. Subathra, M. Balu, and C. Panneerselvam,
Age-associated decreased activities of mitochondrial elec-
tron transport chain complexes in heart and skeletal muscle:
role of L-carnitine, Chemico-Biological Interactions, vol. 148,
no. 1-2, pp. 11–18, 2004.
[108] J. Kerner, P. J. Turkaly, P. E. Minkler, and C. L. Hoppel, “Aging
skeletal muscle mitochondria in the rat: decreased uncou-
pling protein-3 content,American Journal of Physiology, vol.
281, no. 5, pp. E1054–E1062, 2001.
[109] A. Mansouri, F. L. Muller, Y. Liu et al., “Alterations in mito-
chondrial function, hydrogen peroxide release and oxidative
damage in mouse hind-limb skeletal muscle during aging,
Mechanisms of Ageing and Development, vol. 127, no. 3, pp.
298–306, 2006.
[110] R. T. Hepple, D. J. Baker, M. McConkey, T. Murynka, and R.
Norris, “Caloric restriction protects mitochondrial function
with aging in skeletal and cardiac muscles, Rejuvenation
Research, vol. 9, no. 2, pp. 219–222, 2006.
[111] R. T. Hepple, D. J. Baker, J. J. Kaczor, and D. J. Krause, “Long-
term caloric restriction abrogates the age-related decline in
skeletal muscle aerobic function,The FASEB Journal, vol. 19,
no. 10, pp. 1320–1322, 2005.
[112] D. J. Baker, A. C. Betik, D. J. Krause, and R. T. Hepple, “No
decline in skeletal muscle oxidative capacity with aging in
long-term calorically restricted rats: eects are independent
of mitochondrial DNA integrity,Journals of Gerontology A,
vol. 61, no. 7, pp. 675–684, 2006.
[113] F. Capel, V. Rimbert, D. Lioger et al., “Due to reverse electron
transfer, mitochondrial H2O2release increases with age in
human vastus lateralis muscle although oxidative capacity is
preserved,Mechanisms of Ageing and Development, vol. 126,
no. 4, pp. 505–511, 2005.
[114] O. Pastoris, F. Boschi, M. Verri et al., “The eects of
aging on enzyme activities and metabolite concentrations in
skeletal muscle from sedentary male and female subjects,
Experimental Gerontology, vol. 35, no. 1, pp. 95–104, 2000.
[115] U. F. Rasmussen, P. Krustrup, M. Kjær, and H. N. Rasmussen,
“Experimental evidence against the mitochondrial theory of
aging A study of isolated human skeletal muscle mitochon-
dria,Experimental Gerontology, vol. 38, no. 8, pp. 877–886,
2003.
[116] L. Larsson, B. Sjodin, and J. Karlsson, “Histochemical
and biochemical changes in human skeletal muscle with
age in sedentary males, age 22–65 yrs,” Acta Physiologica
Scandinavica, vol. 103, no. 1, pp. 31–39, 1978.
[117] U. F. Rasmussen, P. Krustrup, M. Kjaer, and H. N. Ras-
mussen, “Human skeletal muscle mitochondrial metabolism
in youth and senescence: no signs of functional changes
in ATP formation and mitochondrial oxidative capacity,
Pflugers Archiv European Journal of Physiology, vol. 446, no.
2, pp. 270–278, 2003.
[118] K. K. McCully, R. A. Fielding, W. J. Evans, J. S. Leigh, and
J. D. Posner, “Relationships between in vivo and in vitro
measurements of metabolism in young and old human calf
muscles,Journal of Applied Physiology, vol. 75, no. 2, pp.
813–819, 1993.
[119] P. D. Chilibeck, C. R. McCreary, G. D. Marsh et al., “Eval-
uation of muscle oxidative potential by 31P-MRS during
incremental exercise in old and young humans,European
Journal of Applied Physiology and Occupational Physiology,
vol. 78, no. 5, pp. 460–465, 1998.
[120] K. K. McCully, M. A. Forciea, L. M. Hack et al., “Muscle
metabolism in older subjects using 31P magnetic resonance
spectroscopy,Canadian Journal of Physiology and Pharma-
cology, vol. 69, no. 5, pp. 576–580, 1991.
[121] J. A. Houmard, M. L. Weidner, K. E. Gavigan, G. L. Tyndall,
M. S. Hickey, and A. Alshami, “Fiber type and citrate
synthase activity in the human gastrocnemius and vastus
lateralis with aging,” Journal of Applied Physiology, vol. 85, no.
4, pp. 1337–1341, 1998.
[122] I. Trounce, E. Byrne, and S. Marzuki, “Decline in skeletal
muscle mitochondrial respiratory chain function: possible
factor in ageing,The Lancet, vol. 1, no. 8639, pp. 637–639,
1989.
[123] K. Schunk, M. Pitton, C. D¨
uber,W.Kersjes,S.Schadmand-
Fischer, and M. Thelen, “Dynamic phosphorus-31 magnetic
resonance spectroscopy of the quadriceps muscle: eects of
age and sex on spectroscopic results,Investigative Radiology,
vol. 34, no. 2, pp. 116–125, 1999.
[124] D. J. Taylor, G. J. Kemp, C. H. Thompson, and G. K. Radda,
Ageing: eects on oxidative function of skeletal muscle in
vivo,” Molecular and Cellular Biochemistry, vol. 174, no. 1-2,
pp. 1321–1322, 1997.
[125] D. J. Marcinek, K. A. Schenkman, W. A. Ciesielski, D. Lee,
and K. E. Conley, “Reduced mitochondrial coupling in vivo
16 Journal of Aging Research
alters cellular energetics in aged mouse skeletal muscle,The
Journal of Physiology, vol. 569, part 2, pp. 467–473, 2005.
[126] A. Barrientos, J. Casademont, A. R¨
otig et al., “Absence of
relationship between the level of electron transport chain
activities and aging in human skeletal muscle,Biochemical
and Biophysical Research Communications, vol. 229, no. 2, pp.
536–539, 1996.
[127] E. J. Brierly, M. A. Johnson, A. Bowman et al., “Mitochon-
drial function in muscle from elderly athletes,Annals of
Neurology, vol. 41, no. 1, pp. 114–116, 1997.
[128] R. G. Larsen, D. M. Callahan, S. A. Foulis, and J. A. Kent-
Braun, “Age-related changes in oxidative capacity dier
between locomotory muscles and are associated with phys-
ical activity behavior,Applied Physiology, Nutrition and
Metabolism, vol. 37, no. 1, pp. 88–99, 2012.
[129] D. L. Waters, W. M. Brooks, C. R. Qualls, and R. N.
Baumgartner, “Skeletal muscle mitochondrial function and
lean body mass in healthy exercising elderly,” Mechanisms of
Ageing and Development, vol. 124, no. 3, pp. 301–309, 2003.
[130] K. F. Petersen, D. Befroy, S. Dufour et al., “Mitochondrial
dysfunction in the elderly: possible role in insulin resistance,
Science, vol. 300, no. 5622, pp. 1140–1142, 2003.
[131] D. N. Proctor, W. E. Sinning, J. M. Walro, G. C. Sieck, and
P. W. R. Lemon, “Oxidative capacity of human muscle fiber
types: eects of age and training status,Journal of Applied
Physiology, vol. 78, no. 6, pp. 2033–2038, 1995.
[132] V. Rimbert, Y. Boirie, M. Bedu, J. F. Hocquette, P. Ritz, and
B. Morio, “Muscle fat oxidative capacity is not impaired
by age but by physical inactivity: association with insulin
sensitivity,The FASEB Journal, vol. 18, no. 6, pp. 737–739,
2004.
[133] P. A. Figueiredo, S. K. Powers, R. M. Ferreira, F. Amado,
H. J. Appell, and J. A. Duarte, “Impact of lifelong sedentary
behavior on mitochondrial function of mice skeletal muscle,
The Journals of Gerontology A, vol. 64, no. 9, pp. 927–939,
2009.
[134] J. A. Kent-Braun and A. V. Ng, “Skeletal muscle oxidative
capacity in young and older women and men,Journal of
Applied Physiology, vol. 89, no. 3, pp. 1072–1078, 2000.
[135] I. R. Lanza, D. E. Befroy, and J. A. Kent-Braun, “Age-related
changes in ATP-producing pathways in human skeletal
muscle in vivo,” Journal of Applied Physiology, vol. 99, no. 5,
pp. 1736–1744, 2005.
[136] I. R. Lanza, R. G. Larsen, and J. A. Kent-Braun, “Eects of
old age on human skeletal muscle energetics during fatiguing
contractions with and without blood flow,The Journal of
Physiology, vol. 583, part 3, pp. 1093–1105, 2007.
[137] G. Lopez-Lluch, P. M. Irusta, P. Navas, and R. de Cabo,
“Mitochondrial biogenesis and healthy aging,Experimental
Gerontology, vol. 43, no. 9, pp. 813–819, 2008.
[138] D. A. Hood, I. Irrcher, V. Ljubicic, and A. M. Joseph,
“Coordination of metabolic plasticity in skeletal muscle,The
Journal of Experimental Biology, vol. 209, part 2, pp. 2265–
2275, 2006.
[139] S. Jager, C. Handschin, J. St-Pierre, and B. M. Spiegelman,
AMP-activated protein kinase (AMPK) action in skeletal
muscle via direct phosphorylation of PGC-1alpha,Proceed-
ings of the National Academy of Sciences of the United States of
America, vol. 104, no. 29, pp. 12017–12022, 2007.
[140] P. Puigserver, J. Rhee, J. Lin et al., “Cytokine stimulation of
energy expenditure through p38 MAP kinase activation of
PPARgamma coactivator-1,Molecular Cell, vol. 8, no. 5, pp.
971–982, 2001.
[141] D. Knutti, D. Kressler, and A. Kralli, “Regulation of the
transcriptional coactivator PGC-1 via MAPK-sensitive inter-
action with a repressor,Proceedings of the National Academy
of Sciences of the United States of America, vol. 98, no. 17, pp.
9713–9718, 2001.
[142] T. Akimoto, S. C. Pohnert, P. Li et al., “Exercise stimulates
Pgc-1alpha transcription in skeletal muscle through activa-
tion of the p38 MAPK pathway,The Journal of Biological
Chemistry, vol. 280, no. 20, pp. 19587–19593, 2005.
[143] J. T. Rodgers, C. Lerin, W. Haas, S. P. Gygi, B. M. Spiegelman,
and P. Puigserver, “Nutrient control of glucose homeostasis
through a complex of PGC-1αand SIRT1,Nature, vol. 434,
no. 7029, pp. 113–118, 2005.
[144] Z. Gerhart-Hines, J. T. Rodgers, O. Bare et al., “Metabolic
control of muscle mitochondrial function and fatty acid
oxidation through SIRT1/PGC-1alpha,The EMBO Journal,
vol. 26, no. 7, pp. 1913–1923, 2007.
[145] Z. Wu, X. Huang, Y. Feng et al., “Transducer of reg-
ulated CREB-binding proteins (TORCs) induce PGC-1α
transcription and mitochondrial biogenesis in muscle cells,
Proceedings of the National Academy of Sciences of the United
States of America, vol. 103, no. 39, pp. 14379–14384, 2006.
[146] P. Puigserver, Z. Wu, C. W. Park, R. Graves, M. Wright, and
B. M. Spiegelman, “A cold-inducible coactivator of nuclear
receptors linked to adaptive thermogenesis,Cell, vol. 92, no.
6, pp. 829–839, 1998.
[147] P. Puigserver, G. Adelmant, Z. Wu et al., “Activation of PPARγ
coactivator-1 through transcription factor docking,Science,
vol. 286, no. 5443, pp. 1368–1371, 1999.
[148] J. V. Virbasius, C. A. Virbasius, and R. C. Scarpulla,
“Identity of GABP with NRF-2, a multisubunit activator of
cytochrome oxidase expression, reveals a cellular role for
and ETS domain activator of viral promoters,Genes and
Development, vol. 7, no. 3, pp. 380–392, 1993.
[149] C. M. A. Virbasius, J. V. Virbasius, and R. C. Scarpulla, “NRF-
1, an activator involved in nuclear-mitochondrial interac-
tions, utilizes a new DNA-binding domain conserved in a
family of developmental regulators,Genes and Development,
vol. 7, no. 12 A, pp. 2431–2445, 1993.
[150] R. C. Scarpulla, “Nuclear control of respiratory chain
expression by nuclear respiratory factors and PGC-1-related
coactivator,Annals of the New York Academy of Sciences, vol.
1147, pp. 321–334, 2008.
[151] J. V. Virbasius and R. C. Scarpulla, “Activation of the human
mitochondrial transcription factor A gene by nuclear respi-
ratory factors: a potential regulatory link between nuclear
and mitochondrial gene expression in organelle biogenesis,
Proceedings of the National Academy of Sciences of the United
States of America, vol. 91, no. 4, pp. 1309–1313, 1994.
[152] D. A. Hood and A. M. Joseph, “Mitochondrial assembly:
protein import,The Proceedings of the Nutrition Society, vol.
63, no. 2, pp. 293–300, 2004.
[153] F. Diaz and C. T. Moraes, “Mitochondrial biogenesis and
turnover,Cell Calcium, vol. 44, no. 1, pp. 24–35, 2008.
[154] D. A. Hood, P. J. Adhihetty, M. Colavecchia et al., “Mitochon-
drial biogenesis and the role of the protein import pathway,
Medicine and Science in Sports and Exercise,vol.35,no.1,pp.
86–94, 2003.
[155] W. Neupert and J. M. Herrmann, “Translocation of proteins
into mitochondria,Annual Review of Biochemistry, vol. 76,
pp. 723–749, 2007.
[156] N. Gleyzer, K. Vercauteren, and R. C. Scarpulla, “Control of
mitochondrial transcription specificity factors (TFB1M and
Journal of Aging Research 17
TFB2M) by nuclear respiratory factors (NRF-1 and NRF-
2) and PGC-1 family coactivators,Molecular and Cellular
Biology, vol. 25, no. 4, pp. 1354–1366, 2005.
[157] R. M. Reznick, H. Zong, J. Li et al., “Aging-associated reduc-
tions in AMP-activated protein kinase activity and mito-
chondrial biogenesis,Cell Metabolism, vol. 5, no. 2, pp. 151–
156, 2007.
[158] T. Wenz, S. G. Rossi, R. L. Rotundo, B. M. Spiegelman, and
C. T. Moraes, “Increased muscle PGC-1αexpression protects
from sarcopenia and metabolic disease during aging,Pro-
ceedings of the National Academy of Sciences of the United
States of America, vol. 106, no. 48, pp. 20405–20410, 2009.
[159] L. Bevilacqua, J. J. Ramsey, K. Hagopian, R. Weindruch, and
M. E. Harper, “Eects of short- and medium-term calorie
restriction on muscle mitochondrial proton leak and reactive
oxygen species production,American Journal of Physiology,
vol. 286, no. 5, pp. E852–E861, 2004.
[160] A. M. S. Lezza, V. Pesce, A. Cormio et al., “Increased
expression of mitochondrial transcription factor A and
nuclear respiratory factor-1 in skeletal muscle from aged
human subjects,FEBS Letters, vol. 501, no. 1–3, pp. 74–78,
2001.
[161] A. E. Frazier, C. Kiu, D. Stojanovski, N. J. Hoogenraad, and
M. T. Ryan, “Mitochondrial morphology and distribution in
mammalian cells,Biological Chemistry, vol. 387, no. 12, pp.
1551–1558, 2006.
[162] G. Benard and M. Karbowski, “Mitochondrial fusion and
division: regulation and role in cell viability,Seminars in Cell
& Developmental Biology, vol. 20, no. 3, pp. 365–374, 2009.
[163] D. Stojanovski, O. S. Koutsopoulos, K. Okamoto, and M.
T. Ryan, “Levels of human Fis1 at the mitochondrial outer
membrane regulate mitochondrial morphology,Journal of
Cell Science, vol. 117, part 7, pp. 1201–1210, 2004.
[164] D. C. Chan, “Mitochondrial fusion and fission in mammals,
Annual Review of Cell and Developmental Biology, vol. 22, pp.
79–99, 2006.
[165] V. Romanello, E. Guadagnin, L. Gomes et al., “Mitochondrial
fission and remodelling contributes to muscle atrophy,” The
EMBO Journal, vol. 29, no. 10, pp. 1774–1785, 2010.
[166] C. Q. Scheckhuber, R. A. Wanger, C. A. Mignat, and H. D.
Osiewacz, “Unopposed mitochondrial fission leads to severe
lifespan shortening,Cell Cycle, vol. 10, no. 18, pp. 3105–
3110, 2011.
[167] S. Lee, S. Y. Jeong, W. C. Lim et al., “Mitochondrial fission
and fusion mediators, hFis1 and OPA1, modulate cellular
senescence,Journal of Biological Chemistry, vol. 282, no. 31,
pp. 22977–22983, 2007.
[168] E. Smirnova, D. L. Shurland, S. N. Ryazantsev, and A. M.
van der Bliek, “A human dynamin-related protein controls
the distribution of mitochondria,The Journal of Cell Biology,
vol. 143, no. 2, pp. 351–358, 1998.
[169] T. Ono, K. Isobe, K. Nakada, and J. I. Hayashi, “Human cells
are protected from mitochondrial dysfunction by comple-
mentation of DNA products in fused mitochondria,Nature
Genetics, vol. 28, no. 3, pp. 272–275, 2001.
[170] A. Sato, K. Nakada, and J. I. Hayashi, “Mitochondrial
dynamics and aging: mitochondrial interaction preventing
individuals from expression of respiratory deficiency caused
by mutant mtDNA,Biochimica et Biophysica Acta, vol. 1763,
no. 5-6, pp. 473–481, 2006.
[171] A. Kowald and T. B. Kirkwood, “Evolution of the mitochon-
drial fusion-fission cycle and its role in aging,Proceedings of
the National Academy of Sciences of the United States of
America, vol. 108, no. 25, pp. 10237–10242, 2011.
[172] S. P. Kirkwood, E. A. Munn, and G. A. Brooks, “Mito-
chondrial reticulum in limb skeletal muscle,The American
Journal of Physiology, vol. 3, part 1, pp. C395–C402, 1986.
[173] H. Chen, A. Chomyn, and D. C. Chan, “Disruption of fusion
results in mitochondrial heterogeneity and dysfunction,The
Journal of Biological Chemistry, vol. 280, no. 28, pp. 26185–
26192, 2005.
[174] D. Bach, S. Pich, F. X. Soriano et al., “Mitofusin-2 deter-
mines mitochondrial network architecture and mitochon-
drial metabolism: a novel regulatory mechanism altered in
obesity,Journal of Biological Chemistry, vol. 278, no. 19, pp.
17190–17197, 2003.
[175] S. Cipolat, T. Rudka, D. Hartmann et al., “Mitochondrial
rhomboid PARL regulates cytochrome c release during apop-
tosis via OPA1-dependent cristae remodeling,Cell, vol. 126,
no. 1, pp. 163–175, 2006.
[176] S. Duvezin-Caubet, R. Jagasia, J. Wagener et al., “Proteolytic
processing of OPA1 links mitochondrial dysfunction to alter-
ations in mitochondrial morphology,Journal of Biological
Chemistry, vol. 281, no. 49, pp. 37972–37979, 2006.
[177] H. Chen, M. Vermulst, Y. E. Wang et al., “Mitochondrial
fusion is required for mtdna stability in skeletal muscle and
tolerance of mtDNA mutations,Cell, vol. 141, no. 2, pp.
280–289, 2010.
[178] R. Lodi, C. Tonon, M. L. Valentino et al., “Deficit of in vivo
mitochondrial ATP production in OPA1-related dominant
optic atrophy,Annals of Neurology, vol. 56, no. 5, pp. 719–
723, 2004.
[179] I. Kim, S. Rodriguez-Enriquez, and J. J. Lemasters, “Selective
degradation of mitochondria by mitophagy,Archives of
Biochemistry and Biophysics, vol. 462, no. 2, pp. 245–253,
2007.
[180] K. Wang and D. J. Klionsky, “Mitochondria removal by auto-
phagy,Autophagy, vol. 7, no. 3, pp. 297–300, 2011.
[181] K. Okamoto, N. Kondo-Okamoto, and Y. Ohsumi, “Mito-
chondria-anchored receptor Atg32 mediates degradation of
mitochondria via selective autophagy,Developmental Cell,
vol. 17, no. 1, pp. 87–97, 2009.
[182] T. Kanki, “Nix, a receptor protein for mitophagy in mam-
mals,Autophagy, vol. 6, no. 3, pp. 433–435, 2010.
[183] Y. Gu, C. Wang, and A. Cohen, “Eect of IGF-1 on the bal-
ance between autophagy of dysfunctional mitochondria and
apoptosis,FEBS Letters, vol. 577, no. 3, pp. 357–360, 2004.
[184] Y. Zhang, H. Qi, R. Taylor, W. Xu, L. F. Liu, and S.
Jin, “The role of autophagy in mitochondria maintenance:
characterization of mitochondrial functions in autophagy-
deficient S. cerevisiae strains,Autophagy, vol. 3, no. 4, pp.
337–346, 2007.
[185] G. Cavallini, A. Donati, M. Taddei, and E. Bergamini, “Evi-
dence for selective mitochondrial autophagy and failure in
aging,Autophagy, vol. 3, no. 1, pp. 26–27, 2007.
[186] A. M. Cuervo, E. Bergamini, U. T. Brunk, W. Dr¨
oge, M.
Ffrench, and A. Terman, “Autophagy and aging: the impor-
tance of maintaining “clean” cells,Autophagy, vol. 1, no. 3,
pp. 131–140, 2005.
[187] S. E. Wohlgemuth, A. Y. Seo, E. Marzetti, H. A. Lees, and C.
Leeuwenburgh, “Skeletal muscle autophagy and apoptosis
during aging: eects of calorie restriction and life-long
exercise,Experimental Gerontology, vol. 45, no. 2, pp. 138–
148, 2010.
[188] E. Masiero and M. Sandri, “Autophagy inhibition induces
atrophy and myopathy in adult skeletal muscles,Autophagy,
vol. 6, no. 2, pp. 307–309, 2010.
18 Journal of Aging Research
[189] J. J. Wu, C. Quijano, E. Chen et al., “Mitochondrial dysfunc-
tion and oxidative stress mediate the physiological impair-
ment induced by the disruption of autophagy,Aging, vol. 1,
no. 4, pp. 425–437, 2009.
[190] T. Vellai, K. Takacs-Vellai, M. Sass, and D. J. Klionsky, “The
regulation of aging: does autophagy underlie longevity?”
Trends in Cell Biology, vol. 19, no. 10, pp. 487–494, 2009.
[191] E. B. Taylor and J. Rutter, “Mitochondrial quality control
by the ubiquitin-proteasome system,Biochemical Society
Tra ns ac tions, vol. 39, no. 5, pp. 1509–1513, 2011.
[192] P. Low, “The role of ubiquitin-proteasome system in ageing,
General and Comparative Endocrinology, vol. 172, no. 1, pp.
39–43, 2011.
[193] M. Altun, H. C. Besche, H. S. Overkleeft et al., “Muscle wast-
ing in aged, sarcopenic rats is associated with enhanced
activity of the ubiquitin proteasome pathway,Journal of
Biological Chemistry, vol. 285, no. 51, pp. 39597–39608, 2010.
[194] S. A. Whitman, M. J. Wacker, S. R. Richmond, and M.
P. Godard, “Contributions of the ubiquitin-proteasome
pathway and apoptosis to human skeletal muscle wasting
with age,Pflugers Archiv European Journal of Physiology, vol.
450, no. 6, pp. 437–446, 2005.
[195] V. J. Dalbo, M. D. Roberts, S. E. Hassell, R. D. Brown, and
C. M. Kerksick, “Eects of age on serum hormone con-
centrations and intramuscular proteolytic signaling before
and after a single bout of resistance training,The Journal
of Strength & Conditioning Research, vol. 25, no. 1, pp. 1–9,
2011.
[196] C. K. Lee, R. G. Klopp, R. Weindruch, and T. A. Prolla, “Gene
expression profile of aging and its retardation by caloric
restriction,Science, vol. 285, no. 5432, pp. 1390–1393, 1999.
[197] K. H. Strucksberg, K. Tangavelou, R. Schr¨
oder, and C. S.
Clemen, “Proteasomal activity in skeletal muscle: a matter of
assay design, muscle type, and age,Analytical Biochemistry,
vol. 399, no. 2, pp. 225–229, 2010.
[198] D. Cai, K. K. Lee, M. Li, M. K. Tang, and K. M. Chan,
“Ubiquitin expression is up-regulated in human and rat
skeletal muscles during aging,Archives of Biochemistry and
Biophysics, vol. 425, no. 1, pp. 42–50, 2004.
[199] D. R. Green and J. C. Reed, “Mitochondria and apoptosis,
Science, vol. 281, no. 5381, pp. 1309–1312, 1998.
[200] S. Y. Jeong and D. W. Seol, “The role of mitochondria in
apoptosis,BMB Reports, vol. 41, no. 1, pp. 11–22, 2008.
[201] E. Marzetti, J. C. Hwang, H. A. Lees et al., “Mitochondrial
death eectors: relevance to sarcopenia and disuse muscle
atrophy,Biochimica et Biophysica, vol. 1800, no. 3, pp. 235–
244, 2010.
[202] D. F. Suen, K. L. Norris, and R. J. Youle, “Mitochondrial
dynamics and apoptosis,Genes & Development, vol. 22, no.
12, pp. 1577–1590, 2008.
[203] A. J. Dirks and C. Leeuwenburgh, “The role of apoptosis in
age-related skeletal muscle atrophy,” Sports Medicine, vol. 35,
no. 6, pp. 473–483, 2005.
[204] E. Marzetti and C. Leeuwenburgh, “Skeletal muscle apopto-
sis, sarcopenia and frailty at old age,Experimental Gerontol-
ogy, vol. 41, no. 12, pp. 1234–1238, 2006.
[205] L. T. Malmgren, C. E. Jones, and L. M. Bookman, “Muscle
fiber and satellite cell apoptosis in the aging human thy-
roarytenoid muscle: a stereological study with confocal laser
scanning microscopy,Otolaryngology, vol. 125, no. 1, pp. 34–
39, 2001.
[206] T. Koseki, N. Inohara, S. Chen, and G. Nunez, “ARC, an
inhibitor of apoptosis expressed in skeletal muscle and heart
that interacts selectively with caspases,Proceedings of the
National Academy of Sciences of the United States of America,
vol. 95, no. 9, pp. 5156–5160, 1998.
[207] S. Y. Park, H. Y. Kim, J. H. Lee, K. H. Yoon, M. S. Chang,
and S. K. Park, “The age-dependent induction of apoptosis-
inducing factor (AIF) in the human semitendinosus skeletal
muscle,” Cellular and Molecular Biology Letters, vol. 15, no. 1,
pp. 1–12, 2010.
[208] J. Tamilselvan, G. Jayaraman, K. Sivarajan, and C. Pan-
neerselvam, “Age-dependent upregulation of p53 and
cytochrome c release and susceptibility to apoptosis in skele-
tal muscle fiber of aged rats: role of carnitine and lipoic acid,
Free Radical Biology and Medicine, vol. 43, no. 12, pp. 1656–
1669, 2007.
[209] A. Y. Seo, J. Xu, S. Servais et al., “Mitochondrial iron
accumulation with age and functional consequences,Aging
Cell, vol. 7, no. 5, pp. 706–716, 2008.
[210] A. Dirks and C. Leeuwenburgh, “Apoptosis in skeletal muscle
with aging,” American Journal of Physiology Regulatory, vol.
282, no. 2, pp. 519–527, 2002.
[211] E. Marzetti, S. E. Wohlgemuth, H. A. Lees, H. Y. Chung, S.
Giovannini, and C. Leeuwenburgh, “Age-related activation
of mitochondrial caspase-independent apoptotic signaling
in rat gastrocnemius muscle,Mechanisms of Ageing and
Development, vol. 129, no. 9, pp. 542–549, 2008.
[212] C. Leeuwenburgh, C. M. Gurley, B. A. Strotman, and E. E.
Dupont-Versteegden, “Age-related dierences in apoptosis
with disuse atrophy in soleus muscle,American Journal of
Physiology, vol. 288, no. 5, pp. R1288–R1296, 2005.
[213] P. M. Siu, E. E. Pistilli, and S. E. Alway, “Apoptotic responses
to hindlimb suspension in gastrocnemius muscles from
young adult and aged rats,American Journal of Physiology,
vol. 289, no. 4, pp. R1015–R1026, 2005.
[214] K. M. Rice and E. R. Blough, “Sarcopenia-related apoptosis is
regulated dierently in fast- and slow-twitch muscles of the
aging F344/N ×BN rat model,Mechanisms of Ageing and
Development, vol. 127, no. 8, pp. 670–679, 2006.
[215] E. M. McMillan and J. Quadrilatero, “Dierential apoptosis-
related protein expression, mitochondrial properties, pro-
teolytic enzyme activity, and DNA fragmentation between
skeletal muscles,American Journal of Physiology, vol. 300, no.
3, pp. R531–R543, 2011.
[216] P. J. Adhihetty, V. Ljubicic, K. J. Menzies, and D. A.
Hood, “Dierential susceptibility of subsarcolemmal and
intermyofibrillar mitochondria to apoptotic stimuli,Amer-
ican Journal of Physiology, vol. 289, no. 4, pp. C994–C1001,
2005.
[217] J. O. Holloszy, “Biochemical adaptations in muscle. Eects
of exercise on mitochondrial oxygen uptake and respiratory
enzyme activity in skeletal muscle,Journal of Biological
Chemistry, vol. 242, no. 9, pp. 2278–2282, 1967.
[218] J. O. Holloszy, “Adaptations of muscular tissue to training,
Progress in Cardiovascular Diseases, vol. 18, no. 6, pp. 445–
458, 1976.
[219] D. L. Johannsen, J. P. DeLany, M. I. Frisard et al., “Physical
activity in aging: comparison among young, aged, and
nonagenarian individuals,Journal of Applied Physiology, vol.
105, no. 2, pp. 495–501, 2008.
[220] A. R. Coggan, R. J. Spina, D. S. King et al., “Histochemical
and enzymatic comparison of the gastrocnemius muscle of
young and elderly men and women,Journals of Gerontology,
vol. 47, no. 3, pp. B71–B76, 1992.
[221] J. M. Cooper, V. M. Mann, and A. H. Schapira, “Analyses of
mitochondrial respiratory chain function and mitochondrial
DNA deletion in human skeletal muscle: eect of ageing,
Journal of Aging Research 19
Journal of the Neurological Sciences, vol. 113, no. 1, pp. 91–
98, 1992.
[222] K. R. Short, M. L. Bigelow, J. Kahl et al., “Decline in skeletal
muscle mitochondrial function with aging in humans,
Proceedings of the National Academy of Sciences, vol. 102, no.
15, pp. 5618–5623, 2005.
[223] S. A. Jubrias, P. C. Esselman, L. B. Price, M. E. Cress, and
K. E. Conley, “Large energetic adaptations of elderly muscle
to resistance and endurance training,Journal of Applied
Physiology, vol. 90, no. 5, pp. 1663–1670, 2001.
[224] E. V. Menshikova, V. B. Ritov, L. Fairfull, R. E. Ferrell, D.
E. Kelley, and B. H. Goodpaster, “Eects of exercise on
mitochondrial content and function in aging human skeletal
muscle,” The Journals of Gerontology A, vol. 61, no. 6, pp. 534–
540, 2006.
[225] K. R. Short, “Mitochondrial ATP measurements,American
Journal of Physiology Regulatory, vol. 287, no. 1, pp. R243–
R245, 2004.
[226] W. Song, H. B. Kwak, and J. M. Lawler, “Exercise training
attenuates age-induced changes in apoptotic signaling in rat
skeletal muscle,Antioxidants and Redox Signaling, vol. 8, no.
3-4, pp. 517–528, 2006.
[227] K. R. Short, J. L. Vittone, M. L. Bigelow, D. N. Proctor, and
K. S. Nair, “Age and aerobic exercise training eects on whole
body and muscle protein metabolism,American Journal of
Physiology, vol. 286, no. 1, pp. 92–101, 2004.
[228] C. Leeuwenburgh, R. Fiebig, R. Chandwaney, and Ji Li Li,
Aging and exercise training in skeletal muscle: responses
of glutathione and antioxidant enzyme systems,American
Journal of Physiology, vol. 267, no. 2, part 2, pp. R439–R445,
1994.
[229] G. Parise, A. N. Brose, and M. A. Tarnopolsky, “Resistance
exercise training decreases oxidative damage to DNA and
increases cytochrome oxidase activity in older adults,Exper-
imental Gerontology, vol. 40, no. 3, pp. 173–180, 2005.
[230] G. Parise, S. M. Phillips, J. J. Kaczor, and M. A. Tarnopolsky,
Antioxidant enzyme activity is up-regulated after unilateral
resistance exercise training in older adults,Free Radical
Biology and Medicine, vol. 39, no. 2, pp. 289–295, 2005.
[231] R. J. Colman, R. M. Anderson, S. C. Johnson et al., “Caloric
restriction delays disease onset and mortality in rhesus
monkeys,Science, vol. 325, no. 5937, pp. 201–204, 2009.
[232] J. R. Speakman and S. E. Mitchell, “Caloric restriction,
Molecular Aspects of Medicine, vol. 32, no. 3, pp. 159–221,
2011.
[233] B. J. Merry, “Oxidative stress and mitochondrial function
with aging—the eects of calorie restriction,” Aging Cell, vol.
3, no. 1, pp. 7–12, 2004.
[234] T. A. Zainal, T. D. Oberley, D. B. Allison, L. I. Szweda, and
R. Weindruch, “Caloric restriction of rhesus monkeys lowers
oxidative damage in skeletal muscle,The FASEB Journal, vol.
14, no. 12, pp. 1825–1836, 2000.
[235] B. Drew, P. A. Dirks, C. Selman et al., “Eects of aging
and caloric restriction on mitochondrial energy production
in gastrocnemius muscle and heart,American Journal of
Physiology, vol. 284, no. 2, pp. R474–R480, 2003.
[236] A. Lass, B. H. Sohal, R. Weindruch, M. J. Forster, and R. S.
Sohal, “Caloric restriction prevents age-associated accrual of
oxidative damage to mouse skeletal muscle mitochondria,
Free Radical Biology and Medicine, vol. 25, no. 9, pp. 1089–
1097, 1998.
[237] L. Bevilacqua, J. J. Ramsey, K. Hagopian, R. Weindruch, and
M. E. Harper, “Long-term caloric restriction increases UCP3
content but decreases proton leak and reactive oxygen species
production in rat skeletal muscle mitochondria,American
Journal of Physiology, vol. 289, no. 3, pp. E429–E438, 2005.
[238] M. L. Hamilton, H. Van Remmen, J. A. Drake et al., “Does
oxidative damage to DNA increase with age?” Proceedings
of the National Academy of Sciences of the United States of
America, vol. 98, no. 18, pp. 10469–10474, 2001.
[239] F. Usuki, A. Yasutake, F. Umehara, and I. Higuchi, “Beneficial
eects of mild lifelong dietary restriction on skeletal muscle:
prevention of age-related mitochondrial damage, morpho-
logical changes, and vulnerability to a chemical toxin,Acta
Neuropathologica, vol. 108, no. 1, pp. 1–9, 2004.
[240] C. M. Lee, L. E. Aspnes, S. S. Chung, R. Weindruch,
and J. M. Aiken, “Influences of caloric restriction on age-
associated skeletal muscle fiber characteristics and mitochon-
drial changes in rats and mice,Annals of the New York
Academy of Sciences, vol. 854, pp. 182–191, 1998.
[241] A. E. Civitarese, S. Carling, L. K. Heilbronn et al., “Calorie
restriction increases muscle mitochondrial biogenesis in
healthy humans,PLoS Medicine, vol. 4, no. 3, p. e76, 2007.
[242] R. Sreekumar, J. Unnikrishnan, A. Fu et al., “Eects of caloric
restriction on mitochondrial function and gene transcripts
in rat muscle,American Journal of Physiology, vol. 283, no. 1,
pp. E38–E43, 2002.
[243] G. Lopez-Lluch, N. Hunt, B. Jones et al., “Calorie restric-
tion induces mitochondrial biogenesis and bioenergetic
eciency,Proceedings of the National Academy of Sciences of
the United States of America, vol. 103, no. 6, pp. 1768–1773,
2006.
[244] S. B. Lal, J. J. Ramsey, S. Monemdjou, R. Weindruch, and
M. E. Harper, “Eects of caloric restriction on skeletal
muscle mitochondrial proton leak in aging rats,Journals of
Gerontology A, vol. 56, no. 3, pp. B116–B122, 2001.
[245] C. R. Hancock, D. H. Han, K. Higashida, S. H. Kim, and J.
O. Holloszy, “Does calorie restriction induce mitochondrial
biogenesis? A reevaluation,The FASEB Journal, vol. 25, no.
2, pp. 785–791, 2011.
[246] V. G. Desai, R. Weindruch, R. W. Hart, and R. J. Feuers,
“Influences of age and dietary restriction on gastrocnemius
electron transport system activities in mice,Archives of
Biochemistry and Biophysics, vol. 333, no. 1, pp. 145–151,
1996.
[247] E. Bua, S. H. McKiernan, and J. M. Aiken, “Calorie restriction
limits the generation but not the progression of mitochon-
drial abnormalities in aging skeletal muscle,The FASEB
Journal Biology, vol. 18, no. 3, pp. 582–584, 2004.
[248] S. H. McKiernan, R. J. Colman, M. Lopez et al., “Caloric
restriction delays aging-induced cellular phenotypes in rhe-
sus monkey skeletal muscle,Experimental Gerontology, vol.
46, no. 1, pp. 23–29, 2011.
[249] A. A. Gonzalez, R. Kumar, J. D. Mulligan, A. J. Davis, R.
Weindruch, and K. W. Saupe, “Metabolic adaptations to
fasting and chronic caloric restriction in heart, muscle, and
liver do not include changes in AMPK activity,American
Journal of Physiology, vol. 287, no. 5, pp. E1032–E1037, 2004.
[250] K. Shinmura, K. Tamaki, and R. Bolli, “Short-term caloric
restriction improves ischemic tolerance independent of
opening of ATP-sensitive K+ channels in both young and
aged hearts,Journal of Molecular and Cellular Cardiology,
vol. 39, no. 2, pp. 285–296, 2005.
[251] Z. Q. Wang, Z. E. Floyd, J. Qin et al., “Modulation of skeletal
muscle insulin signaling with chronic caloric restriction in
cynomolgus monkeys,Diabetes, vol. 58, no. 7, pp. 1488–
1498, 2009.
20 Journal of Aging Research
[252] A. J. Dirks and C. Leeuwenburgh, “Aging and lifelong calorie
restriction result in adaptations of skeletal muscle apoptosis
repressor, apoptosis-inducing factor, X-linked inhibitor of
apoptosis, caspase-3, and caspase-12,Free Radical Biology
and Medicine, vol. 36, no. 1, pp. 27–39, 2004.
[253] E. Marzetti, J. M. Lawler, A. Hiona, T. Manini, A. Y. Seo, and
C. Leeuwenburgh, “Modulation of age-induced apoptotic
signaling and cellular remodeling by exercise and calorie
restriction in skeletal muscle,Free Radical Biology and
Medicine, vol. 44, no. 2, pp. 160–168, 2008.
[254] T. Phillips and C. Leeuwenburgh, “Muscle fiber specific
apoptosis and TNF-alpha signaling in sarcopenia are attenu-
ated by life-long calorie restriction,The FASEB Journal, vol.
19, no. 6, pp. 668–670, 2005.
[255] H. Y. Lee, C. S. Choi, A. L. Birkenfeld et al., “Targeted expres-
sion of catalase to mitochondria prevents age-associated
reductions in mitochondrial function and insulin resistance,
Cell Metabolism, vol. 12, no. 6, pp. 668–674, 2010.
[256] E. Marzetti, H. A. Lees, S. E. Wohlgemuth, and C. Leeuwen-
burgh, “Sarcopenia of aging: underlying cellular mechanisms
and protection by calorie restriction, BioFactors, vol. 35, no.
1, pp. 28–35, 2009.
[257] A. M. Payne, S. L. Dodd, and C. Leeuwenburgh, “Life-long
calorie restriction in Fischer 344 rats attenuates age-related
loss in skeletal muscle-specific force and reduces extracellular
space,Journal of Applied Physiology, vol. 95, no. 6, pp. 2554–
2562, 2003.
[258] S. H. McKiernan, E. Bua, J. McGorray, and J. Aiken, “Early-
onset calorie restriction conserves fiber number in aging rat
skeletal muscle,The FASEB Journal, vol. 18, no. 3, pp. 580–
581, 2004.
[259] M. Lagouge, C. Argmann, Z. Gerhart-Hines et al., “Resver-
atrol improves mitochondrial function and protects against
metabolic disease by activating SIRT1 and PGC-1α,” Cell, vol.
127, no. 6, pp. 1109–1122, 2006.
[260] J. H. Um, S. J. Park, H. Kang et al., “AMP-activated protein
kinase-deficient mice are resistant to the metabolic eects of
resveratrol,Diabetes, vol. 59, no. 3, pp. 554–563, 2010.
[261] J. A. Baur, K. J. Pearson, N. L. Price et al., “Resveratrol
improves health and survival of mice on a high-calorie diet,
Nature, vol. 444, no. 7117, pp. 337–342, 2006.
[262] S. Timmers, E. Konings, L. Bilet et al., “Calorie restriction-
like eects of 30 days of resveratrol supplementation on
energy metabolism and metabolic profile in obese humans,
Cell Metabolism, vol. 14, no. 5, pp. 612–622, 2011.
[263] M. Pacholec, J. E. Bleasdale, B. Chrunyk et al., “SRT1720,
SRT2183, SRT1460, and resveratrol are not direct activators
of SIRT1,The Journal of Biological Chemistry, vol. 285, no.
11, pp. 8340–8351, 2010.
[264] J. L. Barger, T. Kayo, J. M. Vann et al., “A low dose of dietary
resveratrol partially mimics caloric restriction and retards
aging parameters in mice,PLoS ONE, vol. 3, no. 6, p. e2264,
2008.
[265] Z. Ungvari, W. E. Sonntag, R. de Cabo, J. A. Baur, and A.
Csiszar, “Mitochondrial protection by resveratrol,Exercise
and Sport Sciences Reviews, vol. 39, no. 3, pp. 128–132, 2011.
[266] T. Murase, S. Haramizu, N. Ota, and T. Hase, “Suppression
of the aging-associated decline in physical performance by
a combination of resveratrol intake and habitual exercise in
senescence-accelerated mice,” Biogerontology, vol. 10, no. 4,
pp. 423–434, 2009.
[267] J. R. Jackson, M. J. Ryan, Y. Hao, and S. E. Alway, “Mediation
of endogenous antioxidant enzymes and apoptotic signaling
by resveratrol following muscle disuse in the gastrocnemius
muscles of young and old rats,American Journal of Physiol-
ogy, vol. 299, no. 6, pp. R1572–R1581, 2010.
[268] J. R. Jackson, M. J. Ryan, and S. E. Alway, “Long-term
supplementation with resveratrol alleviates oxidative stress
but does not attenuate sarcopenia in aged mice,The Journals
of Gerontology A, vol. 66, no. 7, pp. 751–764, 2011.
... Exercise leads to an increase in muscle strength and function, which are closely correlated with mitochondrial biogenesis and function [5,9]. PGC-1α is a master regulator of exercise's effects, controlling the expression of genes related to energy metabolism and mitochondrial biogenesis. ...
... Mitochondria, the energy-producing organelles, play a crucial role in both myoblast differentiation and muscle function [9]. Based on this, we decided to evaluate the effect of GTAE on mitochondrial mass and function. ...
Article
Full-text available
The decline in the function and mass of skeletal muscle during aging or other pathological conditions increases the incidence of aging-related secondary diseases, ultimately contributing to a decreased lifespan and quality of life. Much effort has been made to surmise the molecular mechanisms underlying muscle atrophy and develop tools for improving muscle function. Enhancing mitochondrial function is considered critical for increasing muscle function and health. This study is aimed at evaluating the effect of an aqueous extract of Gloiopeltis tenax (GTAE) on myogenesis and muscle atrophy caused by dexamethasone (DEX). The GTAE promoted myogenic differentiation, accompanied by an increase in peroxisome proliferator-activated receptor γ coactivator α (PGC-1α) expression and mitochondrial content in myoblast cell culture. In addition, the GTAE alleviated the DEX-mediated myotube atrophy that is attributable to the Akt-mediated inhibition of the Atrogin/MuRF1 pathway. Furthermore, an in vivo study using a DEX-induced muscle atrophy mouse model demonstrated the efficacy of GTAE in protecting muscles from atrophy and enhancing mitochondrial biogenesis and function, even under conditions of atrophy. Taken together, this study suggests that the GTAE shows propitious potential as a nutraceutical for enhancing muscle function and preventing muscle wasting.
... Mitochondria are crucial for carrying out a set of processes other than energy production in the cell, including calcium and iron buffering, amino acid and lipid metabolism, thermogenesis, apoptosis, and reactive oxygen species (ROS) signaling [93][94][95]. Age-related biochemical and bioenergetic changes, including reduction of mitochondrial volume and oxidative capacity and production of high levels of ROS, have been observed and associated with perturbations in mitochondrial dynamics, biogenesis, and apoptosis [93][94][95]. ...
... Mitochondria are crucial for carrying out a set of processes other than energy production in the cell, including calcium and iron buffering, amino acid and lipid metabolism, thermogenesis, apoptosis, and reactive oxygen species (ROS) signaling [93][94][95]. Age-related biochemical and bioenergetic changes, including reduction of mitochondrial volume and oxidative capacity and production of high levels of ROS, have been observed and associated with perturbations in mitochondrial dynamics, biogenesis, and apoptosis [93][94][95]. ...
Article
Full-text available
Sarcopenia has a complex pathophysiology that encompasses metabolic dysregulation and muscle ultrastructural changes. Among the drivers of intracellular and ultrastructural changes of muscle fibers in sarcopenia, mitochondria and their quality control pathways play relevant roles. Mononucleated muscle stem cells/satellite cells (MSCs) have been attributed a critical role in muscle repair after an injury. The involvement of mitochondria in supporting MSC-directed muscle repair is unclear. There is evidence that a reduction in mitochondrial biogenesis blunts muscle repair, thus indicating that the delivery of functional mitochondria to injured muscles can be harnessed to limit muscle fibrosis and enhance restoration of muscle function. Injection of autologous respiration-competent mitochondria from uninjured sites to damaged tissue has been shown to reduce infarct size and enhance cell survival in preclinical models of ischemia–reperfusion. Furthermore, the incorporation of donor mitochondria into MSCs enhances lung and cardiac tissue repair. This strategy has also been tested for regeneration purposes in traumatic muscle injuries. Indeed, the systemic delivery of mitochondria promotes muscle regeneration and restores muscle mass and function while reducing fibrosis during recovery after an injury. In this review, we discuss the contribution of altered MSC function to sarcopenia and illustrate the prospect of harnessing mitochondrial delivery and restoration of MSCs as a therapeutic strategy against age-related sarcopenia.
... Thus, early cancer detection is suggested to be one of the best strategies for cancer prevention [71]. Exploring this possibility would be complex given muscle weakness and altered mitochondrial functions could occur in other health conditions such as ageing and muscle disuse [72,73]. These findings also position mitochondrial reprogramming as a potential therapeutic target in pre-atrophy weakness and cachexia during metastatic ovarian cancer. ...
... The main free radicals are usually formed by in chain reaction. Table 1 lists some examples of free radicals: reactive Oxygen species (ROS) and reactive Nitrogen species (RNS), how they are formed and their form of action in a living organism (Peterson et al., 2012;Chen and Zweier, 2014;Deboer, 2015). ...
Article
Full-text available
Este artigo tem como objetivo discutir os conceitos básicos da capacidade/atividade antioxidante nos alimentos e do equilíbrio oxidativo necessário à saúde. A linguagem neste artigo é mantida básica o suficiente para que o público em geral possa entender. O artigo fornece informações condensadas sobre esta área como parte de um esforço mais amplo para popularizar a ciência. O artigo começa discutindo os conceitos básicos de antioxidantes e radicais livres desde os conceitos de química geral. As reações de redução da oxidação são essenciais para a vida, mas também produzem radicais livres nocivos, é a dicotomia básica a ser discutida. O trabalho apresenta então a importância da alimentação no equilíbrio oxidativo do metabolismo humano e como esse equilíbrio é necessário para a manutenção e promoção da saúde, principalmente reduzindo o risco de doenças crônicas não transmissíveis. Por fim, o artigo discute o papel dos antioxidantes no antienvelhecimento e na proteção do DNA.
... The defects in the oxidative phosphorylation system assembly on the cristae of mitochondria give rise to numerous serious human diseases 8 . Being the primary manufacturers of both cellular energy and free radicals, dysfunctional mitochondria are particularly thought to induce a considerable decline in critical muscle function 9 . Moreover, as the biological effect of mitochondria is linked to aging, mitochondrial dysfunction in the electron transport chain involving free radical generation has been implicated in the aging process. ...
Article
Full-text available
Mitochondria are essential organelles in cellular energy metabolism and other cellular functions. Mitochondrial dysfunction is closely linked to cellular damage and can potentially contribute to the aging process. The purpose of this study was to investigate the subcellular structure of mitochondria and their activities in various cellular environments using super-resolution stimulated emission depletion (STED) nanoscopy. We examined the morphological dispersion of mitochondria below the diffraction limit in sub-cultured human primary skin fibroblasts and mouse skin tissues. Confocal microscopy provides only the overall morphology of the mitochondrial membrane and an indiscerptible location of nucleoids within the diffraction limit. Conversely, super-resolution STED nanoscopy allowed us to resolve the nanoscale distribution of translocase clusters on the mitochondrial outer membrane and accurately quantify the number of nucleoids per cell in each sample. Comparable results were obtained by analyzing the translocase distribution in the mouse tissues. Furthermore, we precisely and quantitatively analyzed biomolecular distribution in nucleoids, such as the mitochondrial transcription factor A (TFAM), using STED nanoscopy. Our findings highlight the efficacy of super-resolution fluorescence imaging in quantifying aging-related changes on the mitochondrial sub-structure in cells and tissues.
... The mechanisms that result in muscle ageing are not fully clear, and it is likely that numerous, interrelated processes drive the overall process. Some of the proposed mechanisms include reduced satellite cell number (Shefer et al. 2006), abnormal proteasomal degradation pathways (Fernando et al. 2019), increased ROS production (Palomero et al. 2013), mitochondrial dysfunction (Peterson et al. 2012), increased inflammation (Dalle et al. 2017) and abnormal expression long-noncoding RNAs (Marttila et al. 2020). ...
Preprint
Full-text available
The mechanisms underlying skeletal muscle ageing, whilst poorly understood, are thought to involve dysregulated micro (mi)RNA expression. Using young and aged rat skeletal muscle tissue, we applied high-throughput RNA sequencing to comprehensively study alterations in miRNA expression occurring with age, as well as the impact of caloric restriction (CR) on these changes. Furthermore, the function of the proteins targeted by these age- and CR-associated miRNAs was ascertained. Numerous known and novel age-associated miRNAs were identified of which CR normalised 45.5% to youthful levels. Our results suggested miRNAs upregulated with age to downregulated proteins involved in muscle tissue development and metabolism, as well as longevity pathways, such as AMPK and autophagy. Furthermore, our results found miRNAs downregulated with age to upregulate pro-inflammatory proteins, particularly those involved in innate immunity and the complement and coagulation cascades. Interestingly, CR was particularly effective at normalising miRNAs upregulated with age, rescuing their associated protein coding genes but was less effective at rescuing anti-inflammatory miRNAs downregulated with age. Lastly, the effects of a specific miRNA, miR-96-5p, identified by our analysis to be upregulated with age, were studied in culture C2C12 myoblasts. We demonstrated miR-96-5p to decrease cell viability and markers of mitochondrial biogenesis, myogenic differentiation and autophagy. Overall, our results provide useful information regarding how miRNA expression changes in skeletal muscle, as well as the consequences of these changes and how they are ameliorated by CR.
... To start, age appeared to trigger mitochondrial biogenesis in female rats due to their elevated PGC1α levels ( Fig. 2A), which might reflect an endeavor to replace dysfunctional non-intact mitochondria, resulting in the unchanged CS activity observed. Nonetheless, an age-related decline in mitochondrial density due to a reduced biogenesis is pointed out by the literature [50]. ...
Article
Sarcopenia is associated with reduced quality of life and premature mortality. The sex disparities in the processes underlying sarcopenia pathogenesis, which include mitochondrial dysfunction, are ill-understood and can be decisive for the optimization of sarcopenia-related interventions. To improve the knowledge regarding the sex differences in skeletal muscle aging, the gastrocnemius muscle of young and old female and male rats was analyzed with a focus on mitochondrial remodeling through the proteome profiling of mitochondria-enriched fractions. To the best of our knowledge, this is the first study analyzing sex differences in skeletal muscle mitochondrial proteome remodeling. Data demonstrated that age induced skeletal muscle atrophy and fibrosis in both sexes. In females, however, this adverse skeletal muscle remodeling was more accentuated than in males and might be attributed to an age-related reduction of 17beta-estradiol signaling through its estrogen receptor alpha located in mitochondria. The females-specific mitochondrial remodeling encompassed increased abundance of proteins involved in fatty acid oxidation, decreased abundance of the complexes subunits, and enhanced proneness to oxidative posttranslational modifications. This conceivable accretion of damaged mitochondria in old females might be ascribed to low levels of Parkin, a key mediator of mitophagy. Despite skeletal muscle atrophy and fibrosis, males maintained their testosterone levels throughout aging, as well as their androgen receptor content, and the age-induced mitochondrial remodeling was limited to increased abundance of pyruvate dehydrogenase E1 component subunit beta and electron transfer flavoprotein subunit beta. Herein, for the first time, it was demonstrated that age affects more severely the skeletal muscle mitochondrial proteome of females, reinforcing the necessity of sex-personalized approaches towards sarcopenia management, and the inevitability of the assessment of mitochondrion-related therapeutics. Open access | Full article here: https://doi.org/10.1016/j.freeradbiomed.2024.04.005
Thesis
Full-text available
Doktoravhandling, med artikler (Annet vitenskapelig) Fritextbeskrivning Abstract [en] Statins (3-hydroxy-3-methylglutaryl-CoA reductase, HMGCR, inhibitors) comprise the gold standard for the management of hypercholesterolaemia and prevention of cardiovascular disease (CVDs). However, they are accompanied by potential adverse effects, notably muscle pain and sleep disturbance. These side effects can significantly impact patient adherence to statin therapy and thus increase the risk for CVDs. Despite extensive research, the underlying mechanisms of statin-associated myopathy and sleep disturbance are poorly understood. In Paper I, we conducted a cross-sectional cohort study to investigate the association between statin use and genetic variants for HMGCR with the risk for insomnia and chronotype using UK biobank cohort data. Statin use, insomnia and chronotype were assessed by a self-report touchscreen questionnaire. Statin treatment was associated with an increased risk of insomnia compared to controls, while genetic variants for HMGCR inhibition were associated with a reduced risk for insomnia. No association with late evening chronotype were observed with statin use or genetic variants for HMGCR. In Paper II, we employed Drosophila melanogaster to examine the effect of statins and the role of central inhibition of Hmgcr on sleep behaviour. Flies were treated with fluvastatin for five days and Hmgcr was knocked down in pan neurons and pars intercerebralis (PI), equivalent to the mammalian hypothalamus. Sleep patterns were recorded and analysed. Pan-neuronal- as well as PI inhibition of Hmgcr recapitulates fluvastatin-induced enhanced sleep latency and reduced sleep duration. In Paper III, we deciphered the underlying mechanisms for statin-induced myopathy using D. melanogaster. We found that fluvastatin treatment induced muscular damage, mitochondrial phenotypes, lowered locomotion, reduced climbing activity and was associated with lipotoxicity, impaired muscle differentiation and regeneration, and lowered expression of skeletal muscle chloride channels. Interestingly, selective inhibition of skeletal muscle chloride channels recapitulates fluvastatin-induced myofibrillar damage and lowered climbing activity, while selective Hmgcr inhibition in the skeletal muscles recapitulates fluvastatin-induced mitochondrial round-shape and reduced locomotion activity. In Paper IV, we explored the sequential events of myofibril damage and mitochondrial phenotypes associated with fluvastatin and examined whether inhibition of Hmgcr in the skeletal muscles recapitulates fluvastatin effects on mitochondrial respiratory parameters using D. melanogaster. Acute fluvastatin treatment was associated with reduced mitochondrial content and roundness of the mitochondria without noticeable myofibrillar damage. Intriguingly, chronic fluvastatin treatment was associated with stronger mitochondrial phenotypes along with severe myofibrillar damage, which suggests that mitochondrial phenotypes precede myofibrillar damage. Moreover, selective Hmgcr inhibition did not impact mitochondrial respiratory functions.
Article
Soft drink consumption has become a highly controversial public health issue. Given the pattern of consumption in China, sugar-sweetened beverage is the main type of soft drink consumed. Due to containing high levels of fructose, a soft drink may have a deleterious effect on handgrip strength (HGS) due to oxidative stress, inflammation and insulin resistance. However, few studies show an association between soft drink consumption and HGS in adults. We aimed to investigate the association between soft drink consumption and longitudinal changes in HGS among a Chinese adult population. A longitudinal population-based cohort study (5-year follow-up, median: 3·66 years) was conducted in Tianjin, China. A total of 11 125 participants (56·7 % men) were enrolled. HGS was measured using a handheld digital dynamometer. Soft drink consumption (mainly sugar-containing carbonated beverages) was measured at baseline using a validated FFQ. ANCOVA was used to evaluate the association between soft drink consumption and annual change in HGS or weight-adjusted HGS. After adjusting for multiple confounding factors, the least square means (95 % CI) of annual change in HGS across soft drink consumption frequencies were −0·70 (–2·49, 1·09) for rarely drinks, −0·82 (–2·62, 0·97) for < 1 cup/week and −0·86 (–2·66, 0·93) for ≥ 1 cup/week (Pfor trend < 0·05). Likewise, a similar association was observed between soft drink consumption and annual change in weight-adjusted HGS. The results indicate that higher soft drink consumption was associated with faster HGS decline in Chinese adults.
Preprint
Full-text available
Objectives: A high proportion of women with advanced epithelial ovarian cancer (EOC) experience weakness and cachexia. This relationship is associated with increased morbidity and mortality. EOC is the most lethal gynecological cancer, yet no preclinical cachexia model has demonstrated the combined hallmark features of metastasis, ascites development, muscle loss and weakness in adult immunocompetent mice. Methods: Here, we evaluated a new model of ovarian cancer-induced cachexia with the advantages of inducing cancer in adult immunocompetent C57BL/6J mice through orthotopic injections of EOC cells in the ovarian bursa. We characterized the development of metastasis, ascites, muscle atrophy, muscle weakness, markers of inflammation, and mitochondrial stress in the tibialis anterior (TA) and diaphragm ~45, ~75 and ~90 days after EOC injection. Results: Primary ovarian tumour sizes were progressively larger at each time point while robust metastasis, ascites development, and reductions in body, fat and muscle weights occurred by 90 Days. There were no changes in certain inflammatory (TNFa61537;), atrogene (MURF1 and Atrogin) or GDF15 markers within both muscles whereas IL-6 was increased at 45 and 90 Day groups in the diaphragm. TA weakness in 45 Day preceded atrophy and metastasis that were observed later (75 and 90 Day, respectively). The diaphragm demonstrated both weakness and atrophy in 45 Day. In both muscles, this pre-metastatic muscle weakness corresponded with considerable reprogramming of gene pathways related to mitochondrial bioenergetics as well as reduced functional measures of mitochondrial pyruvate oxidation and creatine-dependent ADP/ATP cycling as well as increased reactive oxygen species emission (hydrogen peroxide). Remarkably, muscle force per unit mass at 90 days was partially restored in the TA despite the presence of atrophy and metastasis. In contrast, the diaphragm demonstrated progressive weakness. At this advanced stage, mitochondrial pyruvate oxidation in both muscles exceeded control mice suggesting an apparent metabolic super-compensation corresponding with restored indices of creatine-dependent adenylate cycling. Conclusion: This mouse model demonstrates the concurrent development of cachexia and metastasis that occurs in women with EOC. The model provides physiologically relevant advantages of inducing tumour development within the ovarian bursa in immunocompetent adult mice. Moreover, the model reveals that muscle weakness in both TA and diaphragm precedes metastasis while weakness also precedes atrophy in the TA. An underlying mitochondrial bioenergetic stress corresponded with this early weakness. Collectively, these discoveries can direct new research towards the development of therapies that target pre-atrophy and pre-metastatic weakness during EOC in addition to therapies targeting cachexia.
Article
Full-text available
Morphometry and oxidative capacity of slow-twitch (type I) and fast-twitch (type IIa and IIb) muscle fibers obtained from vastus lateralis needle biopsies were compared between younger (21-30 yr) and older (51-62 yr) normal fit (maximal O2 uptake = 47.0 vs. 32.3 ml.kg-1.min-1) and endurance-trained (66.3 vs. 52.7 ml.kg-1.min-1) men (n = 6/group). The older groups had smaller type IIa (31%) and IIb (40%) fiber areas and fewer capillaries surrounding these fibers than did younger groups. The reduced type II fiber areas and capillary contacts associated with aging were also observed in the older trained men. However, the capillary supply per unit type II fiber area was not affected by age but was enhanced by training. Additionally, on the basis of quantitative histochemical analysis, succinate dehydrogenase activities of type IIa fibers in the older trained men [4.07 +/- 0.68 (SD) mmol.min-1.l-1] were similar to those observed in younger trained men (4.00 +/- 0.48 mmol.min-1.l-1) and twofold higher than in older normal fit men (2.01 +/- 0.65 mmol.min-1.l-1; age x fitness interaction, P < 0.05). Type I muscle fibers were unaffected by age but were larger and had more capillary contacts and higher succinate dehydrogenase activities in the trained groups. The findings of this study suggest that aging results in a decrease in type II fiber size and oxidative capacity in healthy men and that this latter effect can be prevented by endurance training. Conclusions regarding the effects of age and training status on muscle capillarization depend largely on how these data are expressed.
Article
Full-text available
The recently discovered aging-dependent large accumulation of point mutations in the human fibroblast mtDNA control region raised the question of their occurrence in postmitotic tissues. In the present work, analysis of biopsied or autopsied human skeletal muscle revealed the absence or only minimal presence of those mutations. By contrast, surprisingly, most of 26 individuals 53 to 92 years old, without a known history of neuromuscular disease, exhibited at mtDNA replication control sites in muscle an accumulation of two new point mutations, i.e., A189G and T408A, which were absent or marginally present in 19 individuals younger than 34 years. These two mutations were not found in fibroblasts from 22 subjects 64 to 101 years of age (T408A), or were present only in three subjects in very low amounts (A189G). Furthermore, in several older individuals exhibiting an accumulation in muscle of one or both of these mutations, they were nearly absent in other tissues, whereas the most frequent fibroblast-specific mutation (T414G) was present in skin, but not in muscle. Among eight additional individuals exhibiting partial denervation of their biopsied muscle, four subjects >80 years old had accumulated the two muscle-specific point mutations, which were, conversely, present at only very low levels in four subjects ≤40 years old. The striking tissue specificity of the muscle mtDNA mutations detected here and their mapping at critical sites for mtDNA replication strongly point to the involvement of a specific mutagenic machinery and to the functional relevance of these mutations.
Article
Recent studies have demonstrated the need for complementing cellular genomic information with specific information on expressed proteins, or proteomics, since the correlation between the two is poor. Typically, proteomic information is gathered by analyzing samples on two‐dimensional gels with the subsequent identification of specific proteins of interest by using trypsin digestion and mass spectrometry in a process termed peptide mass fingerprinting. These procedures have, as a rule, been labor‐intensive and manual, and therefore of low throughput. The development of automated proteomic technology for processing large numbers of samples simulataneously has made the concept of profiling entire proteomes feasible at last. In this study, we report the initiation of the (eventual) complete profile of the rat mitochondrial proteome by using high‐throughput automated equipment in combination with a novel fractionation technique using minispin affinity columns. Using these technologies, approximately one hundred proteins could be identified in several days. In addition, separate profiles of calcium binding proteins, glycoproteins, and hydrophobic or membrane proteins could be generated. Because mitochondrial dysfunction has been implicated in numerous diseases, such as cancer, Alzheimer's disease and diabetes, it is probable that the identification of the majority of mitochondrial proteins will be a beneficial tool for developing drug and diagnostic targets for associated diseases.
Article
Recent studies have demonstrated the need for complementing cellular genomic information with specific information on expressed proteins, or proteomics, since the correlation between the two is poor. Typically, proteomic information is gathered by analyzing samples on two-dimensional gels with the subsequent identification of specific proteins of interest by using trypsin digestion and mass spectrometry in a process termed peptide mass fingerprinting. These procedures have, as a rule, been labor-intensive and manual, and therefore of low throughput. The development of automated proteomic technology for processing large numbers of samples simulataneously has made the concept of profiling entire proteomes feasible at last. In this study, we report the initiation of the (eventual) complete profile of the rat mitochondrial proteome by using high-throughput automated equipment in combination with a novel fractionation technique using minispin affinity columns. Using these technologies, approximately one hundred proteins could be identified in several days. In addition, separate profiles of calcium binding proteins, glycoproteins, and hydrophobic or membrane proteins could be generated. Because mitochondrial dysfunction has been implicated in numerous diseases, such as cancer, Alzheimer's disease and diabetes, it is probable that the identification of the majority of mitochondrial proteins will be a beneficial tool for developing drug and diagnostic targets for associated diseases.
Article
It has been suggested that a decline in skeletal muscle oxidative capacity is a general consequence of aging in humans. However, previous studies have not always controlled for the effects of varying levels of physical activity on muscle oxidative capacity. To test the hypothesis that, when matched for comparable habitual physical activity levels, there would be no age-related decline in the oxidative capacity of a locomotor muscle, the postexercise recovery time of phosphocreatine was compared in the tibialis anterior muscle of young [n = 19; 33.8 +/- 4.8 (SD) yr] and older [n = 18; 75.5 +/- 4.5 yr] healthy women and men of similar, relatively low, activity levels. The intramuscular metabolic measurements were accomplished by using phosphorus magnetic resonance spectroscopy. The results indicate that there was no age effect on the postexercise recovery time of phosphocreatine recovery. thus supporting the stated hypothesis. These data suggest that there is no requisite decline in skeletal muscle oxidative capacity with aging in humans, at least through the seventh decade.
Article
Eukaryotic cells maintain the overall shape of their mitochondria by balancing the opposing processes of mitochondrial fusion and fission. Unbalanced fission leads to mitochondrial fragmentation, and unbalanced fusion leads to mitochondrial elongation. Moreover, these processes control not only the shape but also the function of mitochondria. Mitochondrial dynamics allows mitochondria to interact with each other; without such dynamics, the mitochondrial population consists of autonomous organelles that have impaired function. Key components of the mitochondrial fusion and fission machinery have been identified, allowing initial dissection of their mechanisms of action. These components play important roles in mitochondrial function and development as well as programmed cell death. Disruption of the fusion machinery leads to neurodegenerative disease.
Article
Progressive damage to mitochondrial DNA (mtDNA) during life is thought to contribute to aging processes. However, this idea has been difficult to reconcile with the small fraction of mtDNA so far found to be altered. Here, examination of mtDNA revealed high copy point mutations at specific positions in the control region for replication of human fibroblast mtDNA from normal old, but not young, individuals. Furthermore, in longitudinal studies, one or more mutations appeared in an individual only at an advanced age. Some mutations appeared in more than one individual. Most strikingly, a T414G transversion was found, in a generally high proportion (up to 50 percent) of mtDNA molecules, in 8 of 14 individuals above 65 years of age (57 percent) but was absent in 13 younger individuals.