ArticlePDF Available

Role of Mitochondria in Human Aging

Authors:
  • National Yang Ming Chiao Tung University

Abstract and Figures

Mitochondria are the major intracellular source and target sites of reactive oxygen species (ROS) that are continually generated as by-products of aerobic metabolism in animal and human cells. It has been demonstrated that mitochondrial respiratory function declines with age in various human tissues and that a defective respiratory chain results in enhanced production of ROS and free radicals in mitochondria. On the other hand, accumulating evidence now indicates that lipid peroxidation, protein modification and mitochondrial DNA (mtDNA) muutation are concurrently increased during aging. On the basis of these observations and the fact that the rate of cellular production of superoxide anions and hydrogen peroxide increases with age, it has recently been postulated that oxidative stress is a major contributory factor in the aging process. A causal relationship between oxidative modification and mutation of mtDNA, mitochondrial dysfunction and aging has emerged, although some details have remained unsolved. In this article, the role of mitochondria in the human aging process is reviewed on the basis of recent findings gathered from our and other laboratories.
Content may be subject to copyright.
~rnal of
Biomedical
Science
Meeting Report
The Twelfth Joint Annual Conference on Biomedical Sciences, Taipei, April 1%20, 1997
I
J Biomed Sci 1997;4:319-326
Received: June 6, 1997
Accepted: August 15, 1997
Department of Biochemistry and
Center for Cellular and Molecular Biology,
National Yang-Ming University,
Taipei, Taiwan
Role of Mitochondria in
Human Aging
Key Words Abstract
Oxidative damage Mitochondria are the major intracellular source and target sites of reactive
Reactive oxygen species oxygen species (ROS) that are continually generated as by-products of aerobic
Mitochondria metabolism in animal and human cells. It has been demonstrated that mito-
Mitochondrial DNA chondrial respiratory function declines with age in various human tissues and
Mutation that a defective respirators" chain results in enhanced production of ROS and
Aging free radicals in mitochondria. On the other hand, accumulating evidence now
indicates that lipid peroxidation, protein modification and mitochondrial
DNA (mtDNA) mutation are concurrently increased during aging. On the
basis of these observations and the fact that the rate of cellular production of
superoxide anions and hydrogen peroxide increases with age, it has recently
been postulated that oxidative stress is a major contributory factor in the aging
process. A causal relationship between oxidative modification and mutation
ofmtDNA, mitochondrial dysfunction and aging has emerged, although some
details have remained unsolved. In this article, the role of mitochondria in the
human aging process is reviewed on the basis of recent findings gathered from
our and other laboratories.
***~e*samoea*oe,eee,ee
Introduction
Aging is a multifactorial biological process, which is
accompanied by a general decline in biochemical and
physiological functions that leads to the decreased ability
of an individual to respond to a wide range of stresses or
challenges and increased susceptibility to age-associated
diseases and death.
In 1956, Harman [20] first proposed that free radicals
are likely the key factor involved in the aging process. The
main concept was that the accumulation of free-radical-
elicited oxidative damage to macromolecules is a major
contributor to aging. Subsequently, he extended the idea
to suggest that mitochondria are the major target of free-
radical attack that leads to human aging [ 19]. In the past
two decades, the flee-radical theory of aging has been
widely tested and gained great support from molecular
and cellular biological research of aging. Miquel et al. [38,
39] provided great support to this notion by showing that
mitochondrial DNA (mtDNA) damage and lipofuscin
pigment formation in animal tissues are concurrently
increased during aging. In light of the fact that mitochon-
dria are the major intracellular source and vulnerable tar-
get of reactive oxygen species (ROS) and free radicals [10,
43], Linnane et al. [37] further hypothesized that accumu-
lation of somatic mutations in mtDNA is a major contri-
butor to human aging and degenerative diseases. This
modem hypothesis of aging focused on enhanced produc-
tion of ROS and accumulation of mtDNA mutations in
mitochondria of postmitotic cells upon human aging.
KARG E R
Fax +41 61 306 1234
E-Mail karger@karger.ch
www. karger, corn
© 1997 National Science CounciI, ROC
S. Karger AG, Basel
1021-7770/97/0046-0319515.00/0
This article is also accessible online at:
http://BioMedNet,com / karger
Professor Yau-Huei WeiDepartment of Biochemistry, School of Life Science
National Yang-Ming University
Taipei, Taiwan 112 (ROC)
Tel. +886 2 826 71 18, fax +886 2 826 48 43, e-mail joeman@dns.ym.edu.tw
Fig. 1.
Mitochondrial theory of aging and age-related degenera-
tive diseases. The electron transport system in the mitochondrial
inner membrane, which is composed of mtDNA-encoded protein
subunits and nuclear DNA-encoded subunits, is actively involved in
the adenosine-triphosphate-synthesis-coupled respiration that con-
sumes about 95% of the oxygen uptake of the tissue cells. A fraction
of the oxygen is incompletely reduced by 1-electron transfer (mostly
via ubisemiquinone) to generate ROS and organic free radicals,
which may cause oxidative damage and mutation of the nearby
mtDNA molecules that are attached, at least transiently, to the inner
membrane. The oxidatively modified and mutated mtDNA are tran-
scribed and translated to produce defective protein subunits that are
assembled to form defective respiratory enzymes. The impaired elec-
tron transport chain not only works less efficiently in adenosine tri-
phosphate synthesis, but also generates more ROS, which will further
enhance the oxidative damage to various biomolecules in mitochon-
dria. This vicious cycle operates in an age-dependent manner and
results in the widely observed age-related accumulation of oxidative
damage and mutation of mtDNA, which ultimately leads to a pro-
gressive decline in the bioenergetic function of tissue cells in the
aging process. On the other hand, free-radical scavenger systems,
DNA repair systems and mitochondrial turnover for removal of the
oxidative damages by ROS and free radicals become less efficient in
the aging process. Therefore, the accumulation of oxidatively dam-
aged and mutated mtDNA and defective mitochondria and pertur-
bation of mitochondrial turnover act synergistically to cause the gen-
eral decline of biochemical and physiological functions of tissue cells
in the aging process of the human.
In this article, we review recent findings of the oxida-
tive-stress-elicited oxidative damage and mutation of
mtDNA and mitochondrial function decline during hu-
man aging and then discuss the pivotal role of mitochon-
dria in the aging process.
Oxidative Stress and Aging
Mitochondria are the intracellular organelles responsi-
ble for adenosine triphosphate synthesis through the cou-
pling of oxidative phosphorylation to respiration in hu-
man and animal cells [48, 50]. Under normal physiologi-
cal conditions, about 1-5 % of oxygen consumed by mito-
chondria is converted to superoxide anions, hydrogen
peroxide and other ROS [10, 51]. It has been established
that several sites of the respiratory chain are involved in
the generation of ROS [61, 70], and that ROS and free
radicals (e.g. ubisemiquinone and flavosemiquinone) are
continually generated and maintained at a relatively high
steady-state level in mitochondria. It was recently esti-
mated that one normal rat liver mitochondrion can pro-
duce about 3 x
107
superoxide anions in a day [10]. In
fact, within a certain concentration range, ROS assume
important physiological functions such as oxygen burst of
neutrophils and smooth muscle relaxation. In addition,
ROS and free radicals have been demonstrated to act as a
secondary messenger to activate the transcription factors
including NF-~zB and AP-1 [32, 54]. However, an excess
of ROS is harmful to cells [51, 68]. To cope with the ROS,
human cells express antioxidant enzymes including man-
ganese superoxide dismutase (MnSOD), copper/zinc su-
peroxide dismutase (Cu/ZnSOD), glutathione peroxidase
and catalase. MnSOD and Cu/ZnSOD convert superox-
ide anions to hydrogen peroxide, which is then trans-
tbrmed to water by glutathione peroxidase or by catalase.
320 J Biomed Sci 1997;4:319-326 Lee/Wei
Although these enzymes, together with other antioxi-
dants, can dispose of ROS and free radicals, a fraction of
them may escape these cellular defense mechanisms and
cause damage to cellular constituents including DNA,
RNA, proteins and lipids [4, 51 ]. In fact, the concentra-
tion of 8-hydroxy-2'-deoxyguanosine (8-OH-dG), a spe-
cific product of oxidative damage to DNA, has been
shown to increase with age in the tissues of mammals and
insects [2, 17, 52, 57]. Agarwal and Sohat [2] demon-
strated that hyperoxia and X-ray irradiation result in a
substantial increase in the level of 8-OH-dG in the
exposed body site of houseflies. Moreover, it was ob-
served that the specific contents of 8-OH-dG in the flies
overexpressing Cu/ZnSOD and catalase were much lower
than those of the wild-type flies [56]. Furthermore, caloric
restriction was demonstrated to extend the average and
the maximum life spans and to concurrently decrease the
age-related accumulation of 8-OH-dG in various tissues
of the mice [57]. These observations suggest that oxida-
tive damage to DNA plays an important role in aging.
On the other hand, aging-associated accumulation of
inactivated or modified proteins, such as enzymes that
are partially denatured, oxidized and catalytically inac-
tive, has been demonstrated in living organisms including
nematodes, flies and humans [58, 63, 64]. The accumula-
tion of protein carbonyls and the loss of glucose-6-phos-
phate dehydrogenase activity, used as indicators of pro-
tein oxidative damage, have been observed to increase
with age [56, 58, 64]. The intracellular levels of proteolytic
enzymes that hydrolyze oxidatively modified proteins are
insufficient for effective disposal of the aging-associated
increase to aberrant proteins [64]. In the muscle of
Dro-
sophila melanogaster,
the induction of heat shock protein
(hsp) 70 was shown to be responsive to aging [71]. The
same muscle-specific induction of hsp70 was also ob-
served in young flies with mutations in the Cu/ZnSOD or
catalase gene. hsp70 is normally induced in response to
heat and other stresses, and apparently functions to pro-
mote renaturation and folding of proteins, prevent further
protein aggregation and denaturation and facilitate pro-
teolysis of degraded proteins [36, 71]. Several lines of
research have suggested that aging-specific hsp70 expres-
sion may be a result of oxidative damage to proteins [36].
On the other hand, it was shown that the activities of pro-
teases involved in the degradation of aberrant proteins is
markedly decreased in aged animal tissues [33]. It has
long been proposed that pel~urbation of proteolysis is
involved in the formation oflipofuscin and possibly in the
manifestation of animal aging [30, 31]. In fact, lipofuscin
and tipofuscin-like secondary lysosomes are accumulated
in various animal tissues with age [24, 30]. The fluores-
cent pigments are thought to result from cross-linking
between oxidatively modified proteins and lipid peroxi-
dation products, which are concurrently increased during
aging [38, 39]. These observations, together with the
recent findings of wide-spread aging-associated mtDNA
mutations [34, 68, 74], have led us to propose that oxida-
tive damage to cellular constituents and their accumula-
tion with age are the major contributors to the aging pro-
cess (fig. 1).
It has also been observed that the fruit flies with homo-
zygous mutations in either the Cu/ZnSOD or catalase
gene exhibit increased sensitivity to oxidative stress and
have a reduced viability and life span [18, 49]. Because
glutathione peroxidase is absent in
D. melanogaster, Cu/
ZnSOD and catalase thus provide the major enzymatic
antioxidant defenses [59]. Flies that overexpress Cu/
ZnSOD alone or in combination with the overexpression
of catalase were found to exhibit increased resistance to
oxidative stress and have significantly less oxidative dam-
age to proteins and a longer life span [44, 45, 56]. It is
generally accepted that the activities and capacities of
antioxidant systems of tissue cells are declining with age,
leading to the gradual loss of pro-oxidant/antioxidant bal-
ance and accumulation of oxidative damage in the aging
process. These observations provide further support of
the notion that oxidative stress plays an important role in
the aging process.
Oxidative Damage to Mitochondria in Aging
Since mitochondria are the major intracellular source
of ROS, they are thus subjected to direct attacks of ROS
in animal and human cells. It has been recently demon-
strated that the rate of production of superoxide anions
and hydrogen peroxide in mitochondria increases with
age in several mammalian and insect tissues [51, 60-63].
The increase in hydrogen peroxide production ofD.
mela-
nogaster
under aging-elicited oxidative stress was demon-
strated to be related to the oxidative damage to mtDNA
and membrane lipids of mitochondria [60]. Sohal et al.
[63] further demonstrated that the average life span of
dipteran flies is inversely correlated with the rate of mito-
chondrial production of superoxide anions and hydrogen
peroxide and with the level of protein carbonyls in the
tissue cells. Moreover, the age-related increase in the rate
of generation of mitochondriat hydrogen peroxide was
observed to decrease by 40% in the fruit flies overexpress-
ing Cu/ZnSOD and catalase as compared with the wild-
Mitochondrial Role in Aging J Biomed Sci 1997;4:319-326 321
type flies [56]. Therefore, the rate of hydrogen peroxide
release by mitochondria is an important determinant of
the oxidative damage sustained by mitochondria. Ames et
al. [4] first demonstrated that oxidative damage to
mtDNA is much more extensive than that to nuclear
DNA. The specific content of 8-OH-dG of mtDNA was
about 16 times higher than that of nuclear DNA in the
liver of 3-month-old rats. Furthermore, the 8-OH-dG
content in liver mtDNA of the 24-month-old rat was 3
times higher than that of the 3-month-old rat [ 17]. More-
over, the levels of oxidative stress and oxidatively modi-
fied proteins and lipid peroxides in mitochondria have
been shown to increase with age [3, 7, 28, 58]. In addition,
the 8-OH-dG contents in the mtDNA of human dia-
phragm, heart muscle and brain tissues were found to
increase in an age-dependent manner [21, 22]. It has also
been found that mitochondrial glutathione is markedly
oxidized with aging in the rat and mouse [6]. The ratio
between the oxidized and reduced glutathione rises with
age in the liver, kidney and brain of these animals. In the
same study, the 8-OH-dG content of mtDNA was also
found to increase with age in the rat and mouse. Addition-
ally, oral administration ofantioxidants protected against
both glutathione oxidation and mtDNA damage in rats
and mice. These observations suggest a close relationship
between oxidative stress, indicated by glutathione oxida-
tion, and mtDNA damage during the aging process.
Mitochondrial DNA Mutations in Human Aging
Each human and animal cell contains several hundred
to more than a thousand mitochondria, each carrying 2-
10 copies ofmtDNA. Human mtDNA is a 16,569-bp cir-
cular double-stranded DNA molecule [5]. This extrachro-
mosomal genome contains genes coding for 13 polypep-
tides involved in respiration and oxidative phosphoryla-
tion and 2 ribosomal RNA and a set of 22 transfer RNA
(tRNA) that are essential for protein synthesis in mito-
chondria [5, 50]. mtDNA is a naked compact DNA mole-
cule without protective histones and replicates rapidly
without efficient proofreading and DNA repair systems
[12, 14]. It is transiently attached to the mitochondrial
inner membrane, in which a considerable amount of ROS
is continually produced by the respirator-5, chain [27, 28].
These characteristics have rendered mtDNA vulnerable
to attacks by ROS and free radicals generated by the elec-
tron leak of the respiratory chain of mitochondria [46, 67,
681.
In the past 8 years, a number of point mutations, dele-
tions and tandem duplications of mtDNA have been
found in various tissues of aged individuals [13, 15, 29,
34, 35, 42, 69, 72, 74, 77, 78]. These mutant mtDNAs
usually coexist with the wild-type mtDNA within a cell
(heteroplasmy), and the degree ofheteroplasmy often var-
ies in different tissues of the same individual [34, 77, 79].
It has been welt established that many of these mtDNAs-
mutations accumulate with age in postmitotic tissues of
the human [15, 34, 35, 68, 72, 77]. Some of these aging-
associated mtDNA mutations were originally observed in
the affected tissues of patients with mitochondrial dis-
eases. The most common mtDNA mutation is the 4,977-
bp deletion, with a 13-bp direct repeat flanking the 5'- and
3'-end breakpoints at nucleotide position (np) 8,470/
8,482 and np 13,447/13,459, respectively [55, 74]. This
mtDNA deletion was first observed in the muscle of
patients with mitochondrial myopathies, including
chronic progressive external ophthalmoplegia, Kearns-
Sayre syndrome and Pearsons' syndrome [25, 55]. Multi-
ple large-scale deletions of mtDNA have also been found
in various tissues of elderly subjects [67, 68, 72, 77, 79].
Two point mutations at np 3,243 and np 8,344 of
mtDNA, which are, respectively, associated with mito-
chondrial myopathy, encephalopathy, lactic acidosis and
stroke-like episodes (MELAS) syndrome and myoclonic
epilepsy and ragged red fibers (MERRF) syndrome [66]
have been also detected in the muscle of aged individuals
[42, 78]. Additionally, 6 different types of tandem dupli-
cations were found in the D loop region of mtDNA from
the brain, muscle, liver and skin tissues of normal sub-
jects; the incidence and abundance of these tandemly
duplicated mtDNA are increased with age [35, 69].
It is important to note that the proportions of these
mutant mtDNAs in aging human tissues rarely- exeed 1%
[15, 34, 35, 68, 72, 77]. The type and relative proportion
of the aging-associated mutant mtDNA are usually deter-
mined by the choice of primers and PCR conditions.
Recently, a more detailed detection system was designed
for extensive screening ofmtDNA deletions in human tis-
sues [23, 75]. By using 180 kinds of PCR primer pairs, this
system enables one to detect all the possible mtDNA with
deletions over 500 bp. Hayakawa et al. [23] applied this
system to analyze mtDNA from normal hearts of human
subjects of various ages. They observed an extensive frag-
mentation of mtDNA into minicircles with different
sizes. The incidence and abundance of the mutant
mtDNA were found to increase with age and to correlate
well with the oxidative damage to mtDNA [23, 75]. It is
worth mentioning that the aforementioned mutation and
322 J Biomed Sci 1997;4:319-326 Lee/Wei
oxidative damage of mtDNA represent only the tip of the
iceberg of all the mtDNA damages occurring in the aging
process [68].
Although the proportion of the mutant mtDNA was
found to correlate with the 8-OH-dG content of mtDNA
[22, 75], it is poorly understood how oxidative stress or
ROS cause mtDNA mutations. Adachi et al. [1] have
recently demonstrated that ROS may cause large-scale
deletion ofmtDNA in animals. They detected a 4-kb dele-
tion of mtDNA in the heart of Balb/c mice that had
received chronic intraperitoneal injection of doxorubicin,
which is known to induce cardiomyopathy and elicit pro-
found lipid peroxidation of heart mitochondria [ 1]. More-
over, they found that administration of coenzyme Qi0 (a
free-radical scavenger) to the mice could effectively pre-
prevent the mtDNA deletion and decrease the lipid per-
oxide contents of the heart mitochondria. This finding
provides first direct evidence to support the notion that
ROS and free radicals are involved in the large-scale dele-
tion of mtDNA. Although it remains to be established
how mtDNA mutations are initiated or promoted by ROS
and free radicals, recent molecular biological studies have
provided useful information to better understand the
mechanisms of mtDNA mutations.
Sequence analysis of the reported deletions of human
mtDNA revealed that they occurred more frequently
between the origins of replication of the H and L strands
[55, 69] and caused a loss or truncation of genes encoding
tRNA and mRNA that are essential for the proper func-
tioning of mitochondria [68]. The breakpoints of many of
these mtDNA deletions are flanked by direct-repeat se-
quences. Slipped mispairing during DNA replication be-
tween direct repeat sequences [55], homologous replica-
tion [76] and topoisomerase II cleavage [8] have been
suggested to be the possible mechanisms for mtDNA dele-
tions. Presumably due to their stem loop structures, the
mitochondrial tRNA genes are thought to be hot spots for
point mutations [66]. Indeed, more than 10 different
point mutations have been found in the tP, aNA Leu(UUR)
gene [53]. On the other hand, the start sites and the inser-
tion sites of the tandem duplications in the D loop region
of mtDNA have been found to be localized in the regions
containing either a poly C run or a direct repeat sequence
[35, 69]. Moreover, certain regions of mtDNA have been
demonstrated to be particularly sensitive to oxidative
insult of ROS and are prone to mutation [26]. The puta-
tive hot spots for oxidative modification and mutation of
mtDNA could be near or at the unusual structures includ-
ing bent, antibent and non-B DNA sequences in human
mtDNA [26]. These observations suggest that the unusual
structure and/or nucleotide sequence of human mtDNA
are the important factors involved in aging-associated
mtDNA mutations. In addition, it was hypothesized that
genotoxic intermediates of lipid peroxidation may play a
role in eliciting age-associated DNA mutations [28]. The
region of mtDNA that is attached to the ROS-generating
sites in the mitochondrial inner membrane should be
more susceptible to oxidative damage, strand breakage
and mutation [28]. Furthermore, ROS-induced mutagen-
esis has been observed to be DNA polymerase specific
[ 16]. Thus, it is possible that the frequency of occurrence
and the type of mtDNA mutation are determined, at least
in part, by the interaction between mitochondrial DNA
polymerase and the DNA molecules that bear the ROS-
induced oxidative damage during DNA replication.
In addition, several mtDNA mutations have been
reported to occur more frequently in sun-exposed skin at
relatively high levels [47, 72]. This observation suggests
that free radicals generated by environmental insult (e.g.
sunshine, air pollutants and cigarette smoke) may also
play an important role in the induction of mtDNA muta-
tions during the aging process.
Mitochondrial Respiratory Function Declines
with Age
It was first demonstrated in 1989 that the respiratory
function of mitochondria gradually declines with age in
the human liver [73] and skeletal muscle [65], respective-
ly. This phenomenon has been confirmed by several
investigators [48]. The respiratory control, oxidative
phosphorylation efficiency, the rates of resting (state 4)
and adenosine-diphosphate-stimulated (state 3) respira-
tion and the activities of the respiratory enzyme com-
plexes all decline with age in various human tissues [48,
65, 73]. In addition, the number of skeletal and heart mus-
cle fibers deficient in cytochrome c oxidase was found to
increase with age [40, 41]. Since the age-dependent de-
cline of the glutamate-malate-supported respiration was
found to be more dramatic than that of the succinate-sup-
ported respiration, we conjectured that mutation(s) in the
7 genes of NADH dehydrogenase encoded by mtDNA
may be involved in this aging-associated respiratory func-
tion decline. We quickly confirmed this idea by showing
that both the frequency of occurrence and abundance of
the 4,977-bp deleted mtDNA increases with age in the liv-
er [34, 74] and many other tissues [15, 34, 69, 72]. It was
recently observed that the extent of mtDNA mutation
strongly correlates with the progressive decrease in cyto-
Mitochondrial Role in Aging J Biomed Sci 1997;4:319-326 323
chrome c oxidase activity in aging human muscle [48].
Because mtDNA has very little redundancy and high
information density, the large-scale deletions often cause
the removal or truncation of multiple structural genes and
tRNA genes and thereby lead to multiple respiratory-
chain deficiencies. In addition, we recently found several
tandem duplications in the D loop of human mtDNA [35,
69], which contains the replication origin OH and two
transcriptional promoters for each strand of mtDNA [5,
11]. The D loop region is the only and most important
control region in human mtDNA. Therefore, any type of
mutation in the regulatory elements in the D loop of
mtDNA can cause alterations of mtDNA replication and
transcription. In addition, oxidative modification to the
nucleobases could also elicit the errors of mtDNA replica-
tion and gene expression. Therefore, accumulation of oxi-
datively damaged and mutated mtDNA may contribute
to the age-dependent progressive decline of respiratory
function especially in postmitotic cells [4, 39, 48, 67, 68].
On the other hand, it has been reported that the steady-
state levels of mitochondrial transcripts are significantly
reduced during aging of
D. melanogaster;
these changes
correlate very well with the life span of the insect [9]. This
decline in the expression of mitochondrial genes might be
in part caused by damage to mtDNA. Interestingly, it was
recently found that not only cytochrome c oxidase, which
contains subunits encoded by the mitochondrial genome,
but also glutamate dehydrogenase, a nuclear DNA-en-
coded enzyme present in the mitochondrial matrix, grad-
ually lose enzymatic activity during the aging of
D. mela-
nogaster
[9]. These observations suggest that although
mtDNA is more vulnerable to oxidative damage, some
age-related defects in the nuclear genome may also be
involved in aging.
However, it is important to note that the proportions
of the age-related mtDNA mutations in various human
tissues are not so high as those seen in the target tissues of
patients with mitochondrial myopathies [67]. It thus ap-
pears difficult for us to comprehend any significant dele-
terious effect on mitochondrial functions exerted by such
low proportions of the mutant mtDNA in human tissues.
The age-dependent decline of mitochondrial respiratory
function may also be due to the direct ROS damage to
proteins, aside from the deleterious effects of mutation or
oxidative damage to mtDNA. Moreover, it is possible
that all the mutations and oxidative damages to mtDNA
impair, in a synergistic manner, the function of the elec-
tron transport chain and elicit a profound increase in the
rate of ROS generation. A broad spectrum of oxidative
damage and mutation of mtDNA may be effected through
a recently proposed vicious cycle in the aging process [68].
However, it is worth noting that a clear causal relation-
ship between oxidative modification and mutation of
mtDNA, mitochondrial dysfunction and aging remains to
be established.
Concluding Remarks
Mitochondria are responsible for the supply of meta-
bolic energy and are also the main intraceltular source and
target of ROS and free radicals, which are generated as
by-products in the respiratory chain. In the mammalian
cells, the proper assembly and function of mitochondria
are effected through the coordination between gene prod-
ucts encoded by mitochondrial and nuclear genomes [50].
Communication between the nucleus and the mitochon-
drion is essential for delicate regulation of the synthesis of
proteins in the cytosol and their import into mitochon-
dria. ROS and some metabolites that regulate the activa-
tion of specific transcription factors, which may exert
their functions in the nucleus, have been proposed to be
the signals tbr communication between the mitochon-
drion and the nucleus [50, 76]. Besides the effects of the
nuclear genome on the expression of mitochondrial genes,
the mitochondrial genome can also affect the expression
ofnuclar gene-encoded mitochondrial proteins [50]. Oxy-
gen concentration, exercise and hormone levels have been
demonstrated to be able to regulate the mRNA level of
cytochrome c oxidase in the mammal [50, 51]. Therefore,
mitochondria may act as a sensor in regulating energy
metabolism and the release of ROS in response to extra-
cellular stimuli. Moreover, within a certain concentration
range, ROS and free radicals may act as a secondary mes-
senger in some signal transduction pathways [32, 54].
Normally, the overproduced ROS can be scavenged by
enzymatic and nonenzymatic antioxidant systems to pre-
vent deleterious oxidative damage. However, as a result
of an aging-associated increase in ROS generation in the
respiratory chain and a decrease in the intracellular con-
centrations of antioxidants and activities of free-radical-
scavenging enzymes, the age-related elevation of ROS and
oxidative stress is harmful to the cell [51 ]. As the major
intraceltular source ofROS, mitochondria are particularly
vulnerable to oxidative damage. Experimental data from
our and other laboratories have provided ample evidence
to support the notion that mutation and oxidative damage
to mtDNA and mitochondrial respiratory function de-
cline are important contributors to human aging [68].
Although a causal relationship between oxidative modifi-
324 J Biomed Sci 1997;4:319-326 Lee/Wei
cation and mutation of mtDNA, mitochondrial dysfunc-
tion and aging has emerged, the detailed mechanism by
which these molecular and biochemical events cause
aging remains to be established. Understanding of the
age-dependent changes of the structure and function of
mitochondria in the aging process should be of prime
importance in unraveling the molecular basis of aging in
the coming years.
Acknowledgments
Part of the work described in this article was supported by
research grants from the National Science Council (NSC86-2314-
B010-090) and the Department of Health, Executive Yuan, ROC.
Y.-H.W. wishes to express his appreciation of considerable financial
support from the National Science Council in the course of studying
mitochondrial role in human aging.
References
1 Adachi K, Fujiura Y, Mayumi F, Nozuhara A,
Sugiu Y, Sakanashi T, Hidaka T, Toshima H.
A deletion of mitochondrial DNA in murine
doxorubicin-induced card••toxicity. Biochem
Biophys Res Commun 195:945-951; 1993.
2 Agarwal S, Sohal RS. DNA oxidative damage
and life expectancy in houseflies, Proc Natl
Acad Sci USA 91:12332-12335;1994.
3 Agarwal S, Sohal RS. Differential oxidative
damage to mitochondrial proteins during
aging. Mech Ageing Dev 85:55-63; 1995.
4 Ames BN, Shigenaga MK, Hagen TM. Oxi-
dants, ant•oxidants, and the degenerative dis-
eases of aging. Proe Natl Acad Sci USA 90:
7915-7922; 1993.
5 Anderson S, Bankier AT, Barrell BG, De
Bruijn MHL, Coulson AR, Drouin J, Eperon
IC, Nierlich DP, Roe BA, Sanger F, Schreier
PH, Smith AJH, Staden R, Young IG. Se-
quence and organization of the human mito-
chondrial genome. Nature 290:457-465; 1981.
6 Asuncion JGDL, Millan A, Pla R, Bruseghini
L, Esteras A, Pallardo FV, Sastre J, Vina J.
Mitochondrial glutathione oxidation correlates
with age-associated oxidative damage to mito-
chondrial DNA. FASEB J 10:333-338;1996.
7 Bandy B, Day•son AJ. Mitochondrial muta-
tions may increase oxidative stress: Implica-
tions for carcinogenesis and aging. Free Radic
Biol Med 8:5213-5239; 1990.
8 Blok RB, Thorburn DR, Thompson GN, Dahl
HHM. A topoisomerase II cleavage site is asso-
ciated with a novel mitochondrial DNA dele-
tion. Hum Genet 95:75-81; 1995.
9 Calleja M, Pena P, Ugalde C, Ferreiro C, Mar-
co R, Garesse R. Mitochondrial DNA remains
intact during Drosophila aging, but the levels of
mitochondrial transcripts are significantly re-
duced. J BM Chem 268:18891-18897;1993.
10 Chance B, Sies H, Boveris A. Hydmperoxide
metabolism in mammalian organs. Physiol
Rev 59:527-605;1979.
11 Clayton DA. Replication of animal mitochon-
drial DNA. Cell 28:693-705;1982.
12 Clayton DA, Doda JN, Friedberg EC. The
absence of a pyrimidine dimer repair mecha-
nism in mammalian mitochondria. Proc Natl
Acad Sci USA 71:2777-2781;1974.
13 Cortopassi GA, Arnheim N. Detection of a spe-
cific mitochondrial DNA deletion in tissues of
older humans. Nucleic Acids Res 18:6927-
6933;1990.
14 Driggers WJ, Grishko VI, LeDoux SP, Wilson
GL. Defective repair of oxidative damage in
the mitochondrial DNA of a xeroderma pig-
mentosum group A cell line. Cancer Res 56:
1262-1266;1996.
15 Fahn H J, Wang LS, Hsieh RH, Chang SC, Kao
SH, Hnang MH, We• YH. Age-related 4,977 bp
deletion in human lung mitochondrial DNA.
Am J Respir Crit Care Med 154:1141-1145;
1996.
16 Feig DI, Loeb LA. Oxygen radical induced mu-
tagenesis is DNA potymerase specific. J Mol
Biol 235:33-41; 1994.
17 Fraga CG, Shigenaga MK, Park JW, Degan P,
Ames BN. Oxidative damage to DNA during
aging: 8-hydroxy-2"-deoxyguanosine in rat or-
gan DNA and urine. Proc Natl Acad Sci USA
87:4533-4537;1990.
18 Griswold CM, Matthews AL, Bewley KE, Ma-
haffey JW. Molecular characterization and res-
cue of acatalasemic mutants of
Drosophila me-
lanogaster.
Genetics 134:781-788; 1993.
19 Harman D. The biological clock: The mito-
chondria? J Am Geriatr Soc 20:145-147; 1972.
20 Harman D. The aging process. Proc Natl Acad
Sci USA 78:7124-7128;1981.
21 Hayakawa M, Torii K, Sugiyama S, Tanaka M,
Ozawa T. Age-associated accumulation of 8-
hydroxydeoxyguanosine in mitochondrial
DNA of human diaphragm. Biochem Biophys
Res Commun 179:1023-1029;1991.
22 Hayakawa M, Hattori K, Sugiyama S, Ozawa
T. Age-associated oxygen damage and muta-
tions in mitochondrial DNA in human heart.
Biochem Biophys Res Commun 189:979-985;
1992.
23 Hayakawa M, Katsumata K, Yoneda M, Tana-
ka M, Sugiyama S, Ozawa T. Age-related exten-
sive fragmentation of mitochondrial DNA into
minicircles. Biochem Biophys Res Commun
226:369-377; 1996.
24 Hayashida M, Yu BP, Masoro EJ, Iwasaki K,
Ikeda T. An electron microscopic examination
of age-related changes in the rat kidney: The
influence of diet. Exp Gerontol 21:535-553;
1986.
25 Holt IJ, Harding AE, Morgan-Hughes JA. De-
letions of mitochondrial DNA in patients with
mitochondrial myopathies. Nature 331:717-
719;1988.
26 Hou JH, We• YH. The unusual structures of
the hot-regions flanking large-scale deletions in
human mitochondrial DNA. Biochem J 318:
1065-1070;1996.
27 Hruszkewycz AM. Evidence for mitochondrial
DNA damage by lipid peroxidation. Biochem
Biophys Res Commun 153:191-197; 1988.
28 Hruszkewycz AM. Lipid peroxidation and
mtDNA degeneration. A hypothesis. Mutat
Res 275:243-248;1992.
29 Ikebe S, Tanaka M, Ohno K, Sat• W, Hattori
K, Kondo T, Mizuno Y, Ozawa T. Increase of
deleted mitochondrial DNA in the striatum in
Parkinson's disease and senescence. Biochem
Biophys Res Commun 170:1044-1048;1990.
30 Ivy GO, Roopsingh R, Kanai S, Ohta M, Sato
Y, Kitani K. Leupeptin causes an accumula-
tion of lipofuscin-like substances and other
signs of aging in kidneys of young rats: Further
evidence for the protease inhibitor model of
aging. Ann N Y Acad Sci 786:12-23; 1996.
31 Ivy GO, Schottler F, Wenzel J, Baudry M,
Lynch G. Inhibitors of lysosomal enzymes: Ac-
cumulation of lipofuscin-like dense bodies in
the brain. Science 226:985-987; 1984.
32 Lander HM. An essential role for free radicals
and derived species in signal transduction. FA-
SEBJ 11:118-124;1997.
33 Lavie L, Reznick AZ, Gershon D. Decreased
protein and puromycinylpeptide degradation
in livers of senescent mice. Murat Res 275:
217-225;1992.
34 Lee HC, Pang CY, Hsu HS, We• YH. Differ-
ential accumulations of 4,977 bp deletion in
mitochondrial DNA of various tissues in hu-
man ageing. Biochim Biophys Acta 1226:37-
43; 1994.
35 Lee HC, Pang CY, Hsu HS, We• YH. Ageing-
associated tandem duplications in the D-loop
of mitochondrial DNA of human muscle.
FEBS Lett 354:79-83;1994.
36 Lindquist S, Craig EA. The heat-shock pro-
teins. Annu Rev Genet 22:631-677; 1988.
37 Linnane AW, Marzuki S, Ozawa T, Tanaka M.
Mitochondrial DNA mutations as an impor-
tant contributor to ageing and degenerative dis-
ease. Lancet i:642-645;1989.
Mitochondrial Role in Aging J Biomed Sci 1997;4:319-326 325
38 Miquel J. An integrated theory of aging as the
result of mitochondrial DNA mutation in dif-
ferentiated cells. Arch Gerontol Geriatr 12:99-
t17;1991.
39 Miquel J, Economos JE, Johnson JE Jr. Mito-
chondrial rote in cell ageing. Exp Gerontoi 15:
575-591;1980.
40 Mfiller-HScker J. Cytochrome-c-oxidase defi-
cient cardiomyocytes in the human heart- and
age-related phenomenon. A histochemical ul-
tracytochemical study. Am J Pathol 134:1167-
1173;1989.
41 Miiller-HScker J. Cytochrome c oxidase defi-
cient fibers in the limb muscle and diaphragm
of man without muscular disease: An age-relat-
ed alteration. J Neurol Sci 100:14-21;1990.
42 Mtinscher C, Rieger T, Miiller-HScker J, Kad-
enbach B. The point mutation of mitochon-
drial DNA characteristic for MERRF disease is
found also in healthy people of different ages.
FEBS Lett 317:27-30; 1993.
43 Nohl H, Hegner D. Do mitochondria produce
oxygen radicals in vivo? Eur J Biochem 82:
563-567;1978.
44 Orr WC, Sohal RS. Effects of Cu-Zn superox-
ide dismutase overexpression on life span and
resistance to oxidative stress in transgenic
Dro-
sophila melanogaster.
Arch Biochem Biophys
301:34-40; 1993.
45 Orr WC, Sohal RS. Extension of life-span by
overexpression of superoxide dismutase and
catalase in
Drosophila rnelanogaster.
Science
263:i128-1130;1994.
46 Ozawa T, Sahashi K, Nakase Y, Chance B.
Extensive tissue oxygenation associated with
mitochondrial DNA mutations. Biochem Bio-
phys Res Commun 213:432-438;1995.
47 Pang CY, Lee HC, Yang JH, Wei YH. Human
skin mitochondrial DNA deletions associated
with light exposure. Arch Biochem Biophys
312:534-538;1994.
48 Papa S. Mitochondrial oxidative phosphoryla-
tion changes in the life span. Molecular aspects
and physiopathological implications. Biochim
Biophys Acta 1276:87-105; 1996.
49 Philtips JP, Cmnpbell SD, Michaud D, Char-
bonneau M, Hilliker AJ. Null mutation of cop-
per/zinc superoxide dismutase in Drosophila
confer hypersensitivity to paraquat and re-
duced longevity. Proc Natl Acad Sci USA 86:
2761-2765; 1989.
50 Poyton RO. Crosstalk between nuclear and mi-
tochondrial genomes. Annu Rev Biochem 65:
563-607;1996.
51 Richter C, Gogvadze V, Laffranchi R, Schlap-
bach R, Schnizer M, Suter M, Walter P, Yaffee
M. Oxidants in mitochondria: From physiolo-
gy to disease. Biochim Biophys Acta 1271:67-
74;1995.
52 Richter C, Park JW, Ames BN. Normal oxida-
tive damage to mitochondrial and nuclear
DNA is extensive. Proc Natl Acad Sei USA 85:
6465-6467; 1988.
53 Schon EA, Hirano M, DiMauro S. Mitochon-
drial encephalomyopathies: Clinical and mo-
lecular analysis. J Bioenerg Biomembr 26:291-
299;t994.
54 Sen CK, Packer L. Antioxidant and redox regu-
lation of gene transcription. FASEB J 10:709-
720;1996.
55 Shoffner JM, Lott MT, Voljavec AS, Soueidan
SA, Costigan DA, Wallace DC. Spontaneous
Kearns-Sayre/chronic progressive external
ophthalmoplegia plus syndrome associated
with a mitochondrial DNA deletion: A slip-
replication model and metabolic therapy. Proc
Natl Acad Sci USA 86:7952-7956; 1989.
56 Sohal RS, Agarwal A, Agarwal S, Orr WC.
Simultaneous overexpression of copper and
zine-containing superoxide dismutase and cat-
alase retards age-related oxidative damage and
increases metabolic potential in
Drosophila
melanogaster.
J Biol Chem 270:15671-15674;
1995.
57 Sohal RS, Agarwal S, Candas M, Forster M, Lal
H. Effect of age and caloric restriction on DNA
oxidative damage in different tissues of
C57BL/6 mice. Mech Ageing Dev 76:215-224;
1994.
58 Sohal RS, Agarwal S, Dubey A, Orr WC. Pro-
tein oxidative damage is associated with life
expectancy of houseflies. Proc Natl Acad Sci
USA 90:7255-7259;1993.
59 Sohal RS, Arnold L, Orr WC. Effect of age on
superoxide dismutase, catalase, glutathione re-
duetase, inorganic peroxides, TBA-reactive
material, GSH/GSSG, NADPH/NADP ÷ and
NADH/NAD ÷ in
Drosophila melanogaster.
Mech Ageing Dev 56:223-235; 1990.
60 Sohal RS, Dubey A. Mitochondrial oxidative
damage, hydrogen peroxide release, and aging.
Free Radic Biol Med 16:621-626; 1994.
61 Sohal RS, Ku HH, Agarwal S, Forster MJ, Lal
H. Oxidative damage, mitochondrial oxidant
generation and antioxidant defenses during
aging and in response to food restriction in the
mouse. Mech Ageing Dev 74:121 - t 33; 1994.
62 Sohal RS, Sohal BH. Hydrogen peroxide re-
lease by mitoehondria increases during aging.
Mech Ageing Dev 57:187-202; 1991.
63 Sohal RS, Sohal BH, Orr WC. Mitochondrial
superoxide and hydrogen peroxide generation,
protein oxidative damage and longevity in dif-
ferent species of flies. Free Radic Biol Med 19:
499-504;1995.
64 Stadtman ER. Protein oxidation and aging.
Science 257:1220-1224; 1992.
65 Trounce I, Byrne E, Marzuki S. Decline in skel-
etal muscle mitochondrial respiratory" chain
function: Possible factor in ageing. Lancet i:
637-639;1989.
66 Wallace DC. Diseases of the mitochondrial
DNA. Annu Rev Biochem 61:1175-1212;
1992.
67 Wei YH. Mitochondrial DNA alterations as
ageing-associated molecular events. Mutat Res
275:145-155;1992.
68 Wei YH. Oxidative stress and mitochondrial
DNA mutations in human aging. Pmc Soc Exp
Biol Med 1997, in press.
69 Wei YH, Pang CY, You B J, Lee HC. Tandem
duplications and large-scale deletions of mito-
chondrial DNA are early molecular events of
human aging process. Ann N Y Acad Sci 786:
82-101;1996.
70 Wei YH, Scholes CP, King TE. Ubisemiqui-
none radicals from the cytochmme
b-cl
com-
plex of mitochondrial electron transport chain
- demonstration of QP-S radical formation.
Biochem Biophys Res Commun 99:1411-
1419;1981.
71 Wheeler JC, Bieschke ET, Tower J. Muscle-
specific expression of Drosophila hsp70 in re-
sponse to aging and oxidative stress. Proc Natl
Acad Sci USA 92:10408-10412; 1995.
72 Yang JH, Lee HC, Wei YH. Photoageing-asso-
ciated mitochondrial DNA length mutations in
human skin. Arch Dermatol Res 287:641-648;
1995.
73 Yen TC, Chen SH, King KL, Wei YH. Liver
mitochondrial respiratory functions decline
with age. Biochem Biophys Res Commun 65:
994-1003; 1989.
74 Yen TC, Su JH, King KL, Wei YH. Ageing-
associated 5 kb deletion in human liver mito-
chondrial DNA. Biochem Biophys Res Com-
mun 178:124-131;1991.
75 Yoneda M, Katsumata K, Hayakawa M, Tana-
ka M, Ozawa T. Oxygen stress induces an
apoptotic cell death associated with fragmenta-
tion of mitochondrial genome. Biochem Bio-
phys Res Commun 209:723-729;1995.
76 Zeviani M. Nucleus-driven mutations of hu-
man mitochondrial DNA. J Inherit Metab Dis
15:456-471; 1992.
77 Zhang C, Baumer A, Maxwell R J, Linnane
AW, Nagley P. Multiple mitochondrial DNA
deletions in an eiderly human individual.
FEBS Lett 297:34-38;1992.
78 Zhang C, Linnane AW, Nagley P. Occurrence
of a particular base substitution (3243 A to G)
in mitochondrial DNA of tissues of ageing hu-
mans. Biochem Biophys Res Commun 195:
1104-1110;1993.
79 Zhang C, Liu VWS, Nagley P. Gross mosaic
pattern of mitochondrial DNA deletions in
skeletal muscle tissues of an individual adult
human subject. Biochem Biophys Res Com-
mun 233:56-60;1997.
326 J Biomed Sci 1997;4:319-326 Lee/Wel
... CoQ10 is also an important endogenous antioxidant [27,28] that is transported in association with cholesterol [29]. Just as CoQ10 supplementation can effectively bypass mitochondrial "respiratory chain" defects [30], improving cell energy, so too can the lowering of CoQ10 with statins unmask (precipitate) mitochondrial compromise [31]-impairing cell energy and increasing free radical release [32,33]. These effects on mitochondrial energy production and impaired antioxidation contribute to statin AEs, including but not limited to risk of muscle symptoms, new-onset diabetes, and cognitive impairment [20,21,34,35]. ...
... This prooxidant predominance is tied to statin adverse effects [102], tied in turn to older age and metabolic syndrome factors [31]. It is in groups with worse mitochondrial status-e.g., those with older age or metabolic syndrome features-that CoQ10 withdrawal due to statins unmasks mitochondrial impairment [30] and increases free radical release [32,33]-offering prospects for statins' prooxidant properties to override statins' antioxidant mechanisms. ...
... Thus, in obese, diabetic, hypertensive, and/or older patients-patients at risk for serious COVID-19 outcomes [17] that might be deemed to most merit adjunctive off-label statin therapy-statins may be adverse not just to oxidative stress but, in consequence, to downstream effects of oxidative stress that may worsen COVID-19 sequelae as well. One relevant effect (and cause) of oxidative stress is mitochondrial dysfunction [33], for which adverse portent is magnified in patients with hypoxemia. Another, as above, is apoptosis [83], with resultant triggering of inflammation [40] and coagulation activation [40], two serious adverse prognostic indicators in COVID-19 [50]. ...
Article
Full-text available
Statins have been widely advocated for use in COVID-19 based on large favorable observational associations buttressed by theoretical expected benefits. However, past favorable associations of statins to pre-COVID-19 infection outcomes (also buttressed by theoretical benefits) were unsupported in meta-analysis of RCTs, RR = 1.00. Initial RCTs in COVID-19 appear to follow this trajectory. Healthy-user/tolerator effects and indication bias may explain these disparities. Moreover, cholesterol drops in proportion to infection severity, so less severely affected individuals may be selected for statin use, contributing to apparent favorable statin associations to outcomes. Cholesterol transports fat-soluble antioxidants and immune-protective vitamins. Statins impair mitochondrial function in those most reliant on coenzyme Q10 (a mevalonate pathway product also transported on cholesterol)—i.e., those with existing mitochondrial compromise, whom data suggest bear increased risks from both COVID-19 and from statins. Thus, statin risks of adverse outcomes are amplified in those patients at risk of poor COVID-19 outcomes—i.e., those in whom adjunctive statin therapy may most likely be given. High reported rates of rhabdomyolysis in hospitalized COVID-19 patients underscore the notion that statin-related risks as well as benefits must be considered. Advocacy for statins in COVID-19 should be suspended pending clear evidence of RCT benefits, with careful attention to risk modifiers.
... Mitochondria are a leading target of ROS 88,89 due to proximity to ROS production (much of which occurs in mitochondria 88,90,91 ), such that mtDNA mutate at 10-1000x the rate of nuclear DNA 92,93 . Since all mtDNA genes are germane to oxidative phosphorylation [94][95][96] , mtDNA damage commonly hampers mitochondrial respiratory chain function, which in turn further impairs cell energy production and often further increases ROS release 86,87 104,105 , which may be magnified in settings of low mt EN production 106,107 .) ...
... Mitochondria are a leading target of ROS 88,89 due to proximity to ROS production (much of which occurs in mitochondria 88,90,91 ), such that mtDNA mutate at 10-1000x the rate of nuclear DNA 92,93 . Since all mtDNA genes are germane to oxidative phosphorylation [94][95][96] , mtDNA damage commonly hampers mitochondrial respiratory chain function, which in turn further impairs cell energy production and often further increases ROS release 86,87 104,105 , which may be magnified in settings of low mt EN production 106,107 .) ...
Article
Full-text available
Background Overlapping chronic multisymptom illnesses (CMI) include Chronic Fatigue Syndrome (CFS), fibromyalgia, irritable bowel syndrome, multiple chemical sensitivity, and Gulf War illness (GWI), and subsets of autism spectrum disorder (ASD). GWI entails a more circumscribed set of experiences that may provide insights of relevance to overlapping conditions. Objectives To consolidate evidence regarding a role for oxidative stress and mitochondrial dysfunction (OSMD), as primary mediators in CMI, using GWI as a departure point. Methods Exposure relations, character, timecourse and multiplicity of symptoms, and objective correlates of GWI are compared to expectation for OSMD. Objective correlates of OSMD in GWI and overlapping conditions are examined. Discussion OSMD is an expected consequence of known GWI exposures; is compatible with symptom characteristics observed; and accords with objective markers and health conditions linked to GWI, extending to autoimmune disease and infection. Emergent triangulating evidence directly supports OSMD in multisymptom “overlap” CMI conditions, with similarities to, and diagnosed at elevated rates in, GWI, suggesting a common role in each. Conclusions GWI is compatible with a paradigm by which uncompensated exposure to oxidative/nitrative stressors accompanies and triggers mitochondrial dysfunction, cell energy compromise, and multiple downstream effects such as vulnerability to autoantibodies. This promotes a profile of protean symptoms with variable latency emphasizing but not confined to energy-demanding post-mitotic tissues, according with (and accounting for) known properties of multisystem overlap conditions. This advances understanding of GWI; health conditions attending GWI at elevated rates; and overlap conditions like CFS and ASD, providing prospects for vulnerability assessment, mitigation of progression, treatment, and future prevention – with implications germane to additive and excessive environmental oxidative stressor exposures in the civilian setting.
... These include genomic instability, attrition in telomeres, loss of proteostasis, deregulated nutrient sensing, altered intercellular communication, cellular senescence, stem cell exhaustion, epigenetic alterations and mitochondrial dysfunction [1]. Ample evidence exists for a role for mitochondrial dysfunction in ageing [2][3][4][5]. In ageing and under stressful conditions, neurons exhibit an increase in oxidative phosphorylation (OXPHOS) by elevating glycolysis in astrocytes to meet the increased neuronal demand for lactate [6,7]. ...
Article
Full-text available
Mitochondrial dysfunction including deficits of mitophagy is seen in aging and neurodegenerative disorders including Alzheimer’s disease (AD). Apart from traditionally targeting amyloid beta (Aβ), the main culprit in AD brains, other approaches include investigating impaired mitochondrial pathways for potential therapeutic benefits against AD. Thus, a future therapy for AD may focus on novel candidates that enhance optimal mitochondrial integrity and turnover. Bioactive food components, known as nutraceuticals, may serve as such agents to combat AD. Urolithin A is an intestinal microbe-derived metabolite of a class of polyphenols, ellagitannins (ETs). Urolithin A is known to exert many health benefits. Its antioxidant, anti-inflammatory, anti-atherogenic, anti-Aβ, and pro-mitophagy properties are increasingly recognized. However, the underlying mechanisms of urolithin A in inducing mitophagy is poorly understood. This review discusses the mitophagy deficits in AD and examines potential molecular mechanisms of its activation. Moreover, the current knowledge of urolithin A is discussed, focusing on its neuroprotective properties and its potential to induce mitophagy. Specifically, this review proposes potential mechanisms by which urolithin A may activate and promote mitophagy.
... "Aging or senescence" refers to the gradual degenerative changes in the body (that occur as one ages), manifested as abnormal levels of functional metabolism, a decrease in the ability to adapt to the external environment, and a decline in the ability to resist stimulation damage [1]. According to WHO statistics, the proportion of the global population aged 60 and older is expected to increase from 900 million in 2015 to 1.4 billion in 2030, 2.1 billion in 2050, and exceed 3.2 billion in 2100. ...
Article
Full-text available
This study investigated the antioxidant effects of whey protein peptide on learning and memory in aging C57BL/6N mice. A total of 72 SPF male C57BL/6N mice were used. Twelve mice were randomly selected as the control group, and the other mice were intraperitoneally injected with D-galactose (100 mg/kg body weight for 6 weeks), during which, the mice in the control group were intraperitoneally injected with the same amount of normal saline. After 6 weeks, the blood was taken from the epicanthus and the serum MDA level was measured, according to which, the mice were randomly divided into the model control group, the whey protein group (1.5 g/kg body weight), and three Whey protein peptide (WHP) intervention groups (0.3 g/kg body weight, 1.5 g/kg body weight, 3.0 g/kg body weight). The water solution of the test sample was administered by oral gavage every day. The intervention period was 30 days, during which, the model control group, the whey protein group, and the whey protein peptide group continued receiving intraperitoneal injections of D-galactose, while the control group continued receiving intraperitoneal injections of normal saline. After the intervention, behavioral experiments were conducted in the following order: open field test, water maze test, and new object recognition test. After the behavioral experiment, the morphology of hippocampal formation was observed by HE staining and TUNEL labeling. Oxidative stress-related indexes in the serum, liver, and brain were detected. Expression levels of the cholinergic system-related enzymes and proinflammatory cytokines in brain tissue were detected. Western blot was used to detect the expression of synaptic plasticity-related proteins in the mouse brain. The results showed that WHP could significantly improve the accumulation of MDA and PC, increase the activities of SOD and GSH-Px, resist oxidative stress injury, and enhance the potential of endogenous antioxidant defense mechanisms. WHP can significantly improve the decline of aging-related spatial exploration, body movement, and spatial and non-spatial learning/memory ability. Its specific mechanism may be related to reducing the degeneration of hippocampal nerve cells, reducing the apoptosis of nerve cells, improving the activity of AChE, reducing the expression of inflammatory factors (TNF-α and IL-1β) in brain tissue, reducing oxidative stress injury, and improving the expression of p-CaMKⅡ and BDNF synaptic plasticity protein. These results indicate that WHP can improve aging-related oxidative stress, as well as learning and memory impairment.
... As a result, impaired mitochondrial function inevitably leads to a pathological state, ranging from subtle alterations in neuronal function to cell death and neurodegeneration. As mentioned before, an age-dependent decrease of brain bioenergetics metabolism together with an impaired redox homeostasis occur during the aging process in individuals with a history of a normal and healthy life (13,40,66,103). ...
Article
Full-text available
Age-related neurodegenerative diseases such as Alzheimeŕs disease (AD) are distressing conditions causing countless levels of suffering for which treatment is often insufficient or inexistent. Considered to be the most common cause of dementia and an incurable, progressive neurodegenerative disorder, the intricate pathogenic mechanisms of AD continues to be revealed and, consequently, an effective treatment is yet to be developed. Among the diverse hypothesis that have been proposed to explain AD pathogenesis, the one concerning mitochondrial dysfunction has raised as one of the most discussed with an actual acceptance in the field. It posits that manipulating mitochondrial function and understanding the deficits that result in mitochondrial injury may help to control and/or limit the development of AD. To achieve such goal, the concept of mitochondrial medicine places itself as a promising gathering of strategies to directly manage the major insidious disturbances of mitochondrial homeostasis as well as attempts to directly or indirectly manage its consequences in the context of AD. The aim of this review is to summarize the evolution that occurred from the establishment of mitochondrial homeostasis perturbation as masterpieces in AD pathogenesis up until the development of mitochondrial medicine. Following a brief glimpse in the past and current hypothesis regarding the triad of aging, mitochondria and AD, this manuscript will address the major mechanisms currently believed to participate in above mentioned events. Both pharmacological and lifestyle interventions will also be reviewed as AD-related mitochondrial therapeutics. This article is protected by copyright. All rights reserved.
Article
Full-text available
We evaluated whether veterans with Gulf War illness (VGWI) report greater ionizing radiation adverse effects (RadAEs) than controls; whether radiation-sensitivity is tied to reported chemical-sensitivity; and whether environmental exposures are apparent risk factors for reported RadAEs (rRadAEs). 81 participants (41 VGWI, 40 controls) rated exposure to, and rRadAEs from, four radiation types. The relations of RadAE-propensity (defined as the ratio of rRadAEs to summed radiation exposures) to Gulf War illness (GWI) presence and severity, and to reported chemical-sensitivity were assessed. Ordinal logistic regression evaluated exposure prediction of RadAE-propensity in the full sample, in VGWI, and stratified by age and chemical-sensitivity. RadAE-propensity was increased in VGWI (vs. controls) and related to GWI severity (p < 0.01) and chemical-sensitivity (p < 0.01). Past carbon monoxide (CO) exposure emerged as a strong, robust predictor of RadAE-propensity on univariable and multivariable analyses (p < 0.001 on multivariable assessment, without and with adjustment for VGWI case status), retaining significance in age-stratified and chemical-sensitivity-stratified replication analyses. Thus, RadAE-propensity, a newly-described GWI-feature, relates to chemical-sensitivity, and is predicted by CO exposure—both features reported for nonionizing radiation sensitivity, consistent with shared mitochondrial/oxidative toxicity across radiation frequencies. Greater RadAE vulnerability fits an emerging picture of heightened drug/chemical susceptibility in VGWI.
Article
Full-text available
A third of 1990-1 Gulf-deployed personnel developed drug/chemical-induced multisymptom illness, “Gulf War illness” (GWI). Veterans with GWI (VGWI) report increased drug/exposure adverse effects (AEs). Using previously collected data from a case-control study, we evaluated whether the fraction of exposures that engendered AEs (“AE Propensity”) is increased in VGWI (it was); whether AE Propensity is related to self-rated “chemical sensitivity” (it did); and whether specific exposures “predicted” AE Propensity (they did). Pesticides and radiation exposure were significant predictors, with copper significantly “protective”—in the total sample (adjusted for GWI-status) and separately in VGWI and controls, on multivariable regression. Mitochondrial impairment and oxidative stress (OS) underlie AEs from many exposures irrespective of nominal specific mechanism. We hypothesize that mitochondrial toxicity and interrelated OS from pesticides and radiation position people on the steep part of the curve of mitochondrial impairment and OS versus symptom/biological disruption, amplifying impact of new exposures. Copper, meanwhile, is involved in critical OS detoxification processes.
Article
Lipid peroxidation is generally accepted to increase during ageing. We focused on stable lipid peroxidation end products, lipophilic fluorescent pigments (LFP). Their role in ageing in animal/human model has thoroughly been investigated. Changes in characteristics and quantity of LFP during tobacco leaf ageing were studied. We evaluated effect of altered ageing rate on LFP characteristics and accumulation. We employed plants with modulated cytokinins (CK) levels as they are plant hormones controlling development and ageing. Tobacco transgenic plants with senescence-induced CK up-regulation (SAG-ipt plants) displayed rejuvenation of senescent leaves. Slowed ageing was achieved in plants with constitutively down-regulated CK (CKX plants). Both 3D and synchronous fluorescence spectra of LFP indicate more marked differences of fluorophores´ compositions between two different control tobacco cultivars than between control and its transgenic plant. Major fluorophore amount increased with leaf age in both control and CKX plants, in SAG-ipt tobacco its quantity declined in the oldest leaves. The postponed leaf ageing in CKX plants resulted in a diminution of LFP. Higher LFP level in SAG-ipt plants compared with their control leaves of the same age was observed. LFP composition as well as relative abundance of individual LFP might be characteristic for each tobacco cultivar and depend on leaf age.
Article
Full-text available
Importance: A "mystery" illness striking US and Canadian diplomats to Cuba (and now China) "has confounded the FBI, the State Department and US intelligence agencies." Sonic explanations for the so-called "health attacks" have long dominated media reports, propelled by peculiar sounds heard and auditory symptoms experienced. Sonic mediation was justly rejected by experts. We assessed whether pulsed radiofrequency/microwave radiation (RF/MW) exposure can accommodate reported facts in diplomats, including unusual ones. Observations: 1. Noises: Chirping, ringing or grinding noises were heard at night, during episodes reportedly triggering health problems, by many diplomats. Pulsed RF/MW engenders just these "sounds" via the "Frey effect." Ability to hear the sounds depends on high frequency hearing and low ambient noise. "Sounds" differ by head dimensions. 2. Signs/symptoms: Hearing loss and tinnitus are prominent in affected diplomats - and in RF/MW-affected individuals. Each of protean symptoms that diplomats report, also affect persons reporting symptoms from RF/MW: Sleep problems, headaches, and cognitive problems dominate in both groups. Sensations of pressure or vibration figure in each. Both encompass vision, balance and speech problems, and nosebleeds. Brain injury and brain swelling are reported in both. 3. Mechanisms: Oxidative stress provides a documented mechanism of RF/MW injury compatible with reported signs and symptoms; sequelae of endothelial dysfunction (yielding blood flow compromise), membrane damage, blood brain barrier disruption, mitochondrial injury, apoptosis, and autoimmune triggering afford downstream mechanisms, of varying persistence, that merit investigation. 4. Of note, microwaving of the US embassy in Moscow is historically documented. Conclusions and relevance: Reported facts appear consistent with RF/MW as the source of injury in Cuba diplomats. Non-diplomats citing symptoms from RF/MW, often with an inciting pulsed-RF/MW exposure, report compatible health conditions. Under the RF/MW hypothesis, lessons learned for diplomats and for RF/MW-affected "civilians" may each aid the other.
Chapter
Plasma contains cholesteryl, glyceryl, and phosphoglyceryl esters of fatty acids as major lipid components. These compounds are carried in plasma and provide a source of metabolic fuel that drives all biochemical reactions required for normal cellular activity and they provide structural material for cell membranes and lipoproteins. These compounds also include molecules that signal receptor activation on the cell surface, triggering subsequent signaling cascades through various protein kinases and phospholipases. This chapter provides a brief summary of plasma lipid composition and the main chemical and biochemical methods of detecting and quantifying the major primary and secondary products of oxidation of plasma lipoproteins and presents a discussion of the physicochemical, chemical, and metabolic activities of the major individual molecular species or small groups thereof. It also provides attention to the possible involvement of oxo-lipids in metabolic syndromes and disease conditions, as well as prevention of oxidation. The nature and extent of plasma lipid autoxidation depends on the fatty acid composition of glycerolipids, glycerophospholipids, and cholesteryl esters of plasma lipo-proteins. The fatty acid composition of the lipoproteins varies with the diet and to a lesser extent with the age, sex, and general health of the individual.
Article
Full-text available
The complete sequence of the 16,569-base pair human mitochondrial genome is presented. The genes for the 12S and 16S rRNAs, 22 tRNAs, cytochrome c oxidase subunits I, II and III, ATPase subunit 6, cytochrome b and eight other predicted protein coding genes have been located. The sequence shows extreme economy in that the genes have none or only a few noncoding bases between them, and in many cases the termination codons are not coded in the DNA but are created post-transcriptionally by polyadenylation of the mRNAs.
Article
This review focuses on molecular mechanisms that underlie the communication between the nuclear and mitochondrial genomes in eukaryotic cells. These genomes interact in at least two ways. First, they contribute essential subunit polypeptides to important mitochondrial proteins; second, they collaborate in the synthesis and assembly of these proteins. The first type of interaction is important for the regulation of oxidative energy production. Isoforms of the nuclear-coded subunits of cytochrome c oxidase affect the catalytic functions of its mitochondrially coded subunits. These isoforms are differentially regulated by environmental and developmental signals and probably allow tissues to adjust their energy production to different energy demands. 'Ihe second type of interaction requires the bidirectional flow of information between the nucleus and the mitochondrion. Communication from the nucleus to the mitochondrion makes use of proteins that are translated in the cytosol and imported by the mitochondrion. Communication from the mitochondrion to the nucleus involves metabolic signals and one or more signal transduction pathways that function across the inner mitochondrial membrane. An understanding of both types of interaction is important for an understanding of OXPHOS diseases and aging.
Article
Oxidative damage to DNA can be caused by excited oxygen species, which are produced by radiation or are by-products of aerobic metabolism. The oxidized base, 8-hydroxydeoxyguanosine (oh8dG), 1 of approximately 20 known radiation damage products, has been assayed in the DNA of rat liver. oh8dG is present at a level of 1 per 130,000 bases in nuclear DNA and 1 per 8000 bases in mtDNA. Mitochondria treated with various prooxidants have an increased level of oh8dG. The high level of oh8dG in mtDNA may be caused by the immense oxygen metabolism, relatively inefficient DNA repair, and the absence of histones in mitochondria. It may be responsible for the observed high mutation rate of mtDNA.
Article
It has recently been suggested that mitochondrial DNA (mtDNA) mutations are important contributors to human ageing and degenerative diseases. Using PCR techniques, we demonstrated three types of mtDNA length mutations, a 4977 bp deletion, a 7436 bp deletion and tandem duplications, in normal human skin tissues. We found that these mutations started to appear in the third decade of life, and the age at which the mutations could be detected in sun-exposed skin was usually younger than in non-exposed skin. Moreover, the incidences of these deletions and tandem duplications of mtDNA in sun-exposed skin were all significantly higher than those in non-exposed skin (Pr=0.4, P