You are on page 1of 16

Ageing Research Reviews 33 (2017) 89–104

Contents lists available at ScienceDirect

Ageing Research Reviews


journal homepage: www.elsevier.com/locate/arr

Review

Role of the mitochondrial DNA replication machinery in


mitochondrial DNA mutagenesis, aging and age-related diseases
Karen L. DeBalsi, Kirsten E. Hoff, William C. Copeland ∗
Genome Integrity and Structural Biology Laboratory, National Institute of Environmental Health Sciences, National Institutes of Health, Research Triangle
Park, NC 27709, USA

a r t i c l e i n f o a b s t r a c t

Article history: As regulators of bioenergetics in the cell and the primary source of endogenous reactive oxygen species
Received 24 February 2016 (ROS), dysfunctional mitochondria have been implicated for decades in the process of aging and age-
Received in revised form 19 April 2016 related diseases. Mitochondrial DNA (mtDNA) is replicated and repaired by nuclear-encoded mtDNA
Accepted 19 April 2016
polymerase ␥ (Pol ␥) and several other associated proteins, which compose the mtDNA replication
Available online 30 April 2016
machinery. Here, we review evidence that errors caused by this replication machinery and failure to
repair these mtDNA errors results in mtDNA mutations. Clonal expansion of mtDNA mutations results
Keywords:
in mitochondrial dysfunction, such as decreased electron transport chain (ETC) enzyme activity and
Mitochondrial DNA mutations
MtDNA replication
impaired cellular respiration. We address the literature that mitochondrial dysfunction, in conjunction
DNA polymerase gamma with altered mitochondrial dynamics, is a major driving force behind aging and age-related diseases.
POLG Additionally, interventions to improve mitochondrial function and attenuate the symptoms of aging are
Aging examined.
Age-related diseases Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
2. Mitochondrial biology and genetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3. MtDNA damage in aging: ROS versus replication errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4. Fidelity of DNA synthesis by pol ␥ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5. Exonuclease-deficient pol ␥ mouse links mtDNA mutations and deletions to aging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6. Age-related mitochondrial diseases associated with defective mtDNA replication machinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.1. DNA polymerase ␥ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.2. PolG2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.3. Twinkle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.4. Transcription Factor A of mitochondria (TFAM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.5. Mitochondrial genome maintenance exonuclease 1 (MGME1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.6. Ribonuclease H1 (RNase H1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
7. The contribution of clonal expansion of mtDNA mutations to mitochondrial dysfunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
8. Mitochondrial dynamics and aging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9. Interventions to attenuate the symptoms of aging. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .98
10. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

1. Introduction

∗ Corresponding author at: Genome Integrity and Structural Biology Laboratory,


Aging is a highly complex, progressive process characterized by
NIEHS, National Institutes of Health, 111 T.W. Alexander Dr., Bldg. 101, Rm. E316,
impairment in physiological function eventually leading to disease
Research Triangle Park, NC 27709, USA. and death. This impairment is the result of accumulated molec-
E-mail address: copelan1@niehs.nih.gov (W.C. Copeland). ular damage associated with metabolic activity and regulated by

http://dx.doi.org/10.1016/j.arr.2016.04.006
1568-1637/Published by Elsevier B.V.
90 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

genetic and environmental effects (Kirkwood, 2005). In the 1950’s, tinct, dynamic structures known as nucleoids (Garrido et al.,
Harmon first proposed that ROS generated by normal metabolism 2003). Depletion of mtDNA leads to dysfunctional mitochondria,
directly regulate the aging process (Harman, 1956). He later refined which can contribute to aging, disease and ultimately cell death.
his theory to include mitochondria as both the primary source Additionally, mtDNA point mutations and deletions can disrupt
and target of ROS, leading to the accumulation of defective mito- mitochondrial function if the percentage of mutant mtDNA per
chondria and the hindrance of functional replacements (Harman, cell reaches a threshold where not enough functional mitochon-
1972). Others specifically advocated that accumulation of ROS- dria remain to maintain proper energetic function (Wallace, 1999).
driven somatic mtDNA mutations, secondary to a lack of adequate Therefore, effective and faithful mtDNA replication is necessary for
repair mechanisms, is the main driving force behind aging (Fleming cell survival.
et al., 1982; Linnane et al., 1989). The mtDNA mutations lead to Of the 17 human cellular DNA polymerases, polymerase gamma
disruption of the electron transport chain (ETC), which generates (Pol ␥) is the only replicative polymerase known to function in
more ROS followed by additional mtDNA mutations, thus creating mitochondrial DNA replication and repair (Bebenek and Kunkel,
a “vicious” cycle of damage culminating in apoptosis (Mandavilli 2004; Ropp and Copeland, 1996; Sweasy et al., 2006). The Pol ␥
et al., 2002). holoenzyme is a heterotrimer consisting of a single 140 kDa cat-
Support for the classical Mitochondrial Free Radical Theory of alytic subunit (nuclear encoded by POLG at chromosomal locus
Aging (MFRTA) waned as new evidence linked ROS to both tox- 15q25) and a 55 kDa accessory subunit that forms a dimer (nuclear
icity and signaling pathways associated with longevity. Newer encoded by POLG2 at chromosomal locus 17q24.1). The catalytic
theories associate aging and mitochondrial dysfunction with clon- subunit has DNA polymerase, 3 → 5 exonuclease and 5 -dRP lyase
ally expanded mtDNA mutations caused by replication errors and activities (Graziewicz et al., 2006; Kaguni, 2004). The accessory sub-
altered mitochondrial dynamics. In particular, much of the mtDNA unit is required for enhanced DNA processivity by augmenting the
damage is linked with the nuclear-encoded proteins responsible affinity of the catalytic subunit for DNA (Lim et al., 1999; Young
for its replication. Here, we focus on aspects of the mitochondrial et al., 2015). The Pol ␥ holoenzyme functions in conjunction with
replication machinery that contribute to mtDNA mutagenesis and Twinkle, mtSSB, topoisomerase, mitochondrial RNA polymerase
aging. (POLRMT), RNase H1 and mitochondrial DNA ligase III (mtLigIII) to
faithfully replicate mtDNA (Fig. 2). Other proteins involved in main-
tenance of the mitochondrial genome include: (1) TFAM, which
2. Mitochondrial biology and genetics enhances mtDNA packaging and compaction (Ngo et al., 2014);
(2) transcription factor B of mitochondria (TFB2M); (3) Optic Atro-
Mitochondria are intracellular organelles that produce the phy 1 (OPA1), which plays a key role in mitochondrial fusion and
majority of adenosine triphosphate (ATP) in cells via the ETC mitophagy (Chen and Chan, 2005); (4) the RecB-type mitochondrial
and oxidative phosphorylation (OXPHOS). Mitochondria are also genome maintenance 5 → 3 exonuclease 1 (MGME1); (5) the RNA
involved in buffering cytosolic calcium, regulating metabolic and and DNA 5 flap endonuclease (FEN1) and (6) the helicase/nuclease
signaling pathways and controlling apoptosis through the mito- DNA2. MGME1 may participate in DNA repair, initiation of mtDNA
chondrial permeability transition pore (Wallace, 2005). replication, RNA primer removal or even the destructive elimina-
The ETC consists of four complexes located in the inner tion of improper mtDNA molecules, while the latter two proteins
mitochondrial membrane (Fig. 1). Energy from dietary reduc- are involved in mtDNA base excision repair (BER) (Copeland and
ing equivalents, nicotinamide adenine dinucleotide (NADH) and Longley, 2014).
flavine adenine dinucleotide (FADH2 ), flow down the ETC as elec-
trons that are used by Complexes I, III and IV to pump protons
across the mitochondrial inner membrane, generating a proton 3. MtDNA damage in aging: ROS versus replication errors
electrochemical gradient. Protons allowed back into the matrix
along this gradient are used by ATP synthase (Complex V) to drive Accumulations of mtDNA damage in the form of point muta-
synthesis of ATP from ADP and inorganic phosphate (Pi). ATP tions, deletions and depletions are correlated with aging and
is then translocated into the intermembrane space by adenine age-related diseases as observed in human skeletal muscle (Bua
nucleotide translocases (ANT1-4), which in conjunction with the et al., 2006; Fayet et al., 2002), liver (Yen et al., 1991), brain
voltage-dependent anion channel (VDAC) allow ATP into the cyto- (Corral-Debrinski et al., 1992a; Corral-Debrinski et al., 1994; Lin
sol (Wallace, 1997). Greater than 90% of the oxygen taken up into et al., 2002), heart (Corral-Debrinski et al., 1992b; Cortopassi and
the cell is consumed by the ETC. Premature electron leakage from Arnheim, 1990), skin fibroblasts (Michikawa et al., 1999) and
Complexes I and III convert 1–2% of the oxygen to ROS as superoxide colonic crypt stem cells (Taylor et al., 2003). Early reports com-
(O2 •− ). The superoxide anions are converted to hydrogen peroxide paring nucleotide substitutions in mtDNA from somatic tissues of
(H2 O2 ) by the superoxide dismutases (SODs). H2 O2 is then detoxi- different primates revealed a 10-fold higher rate of evolution for
fied either by peroxisomal catalase or the glutathione system. If not mtDNA relative to the nuclear genome (Brown et al., 1979; Brown
detoxified, H2 O2 is transformed into the hydroxyl radical (HO• ) via et al., 1982), implying a relatively high mutation rate for mtDNA.
the Fenton reaction (Murphy, 2009). More recently, estimates have increased the mutation rate from 10-
Human mtDNA is maternally inherited (Giles et al., 1980) and to 17-fold higher in mtDNA than nuclear genomic (nDNA) (Tuppen
exists as a circular, double-stranded genome organized into 16,569 et al., 2010). Therefore, it is essential to understand the origins of
base pairs. It encodes 37 mitochondrial genes: 22 transfer RNAs, 2 mutations in mtDNA.
mitochondrial ribosomal RNAs and 13 protein subunits of ETC com- It was proposed that the higher mutation rate in mtDNA is
plexes, with the exception of Complex II which is entirely nuclear caused by ROS because the mitochondrial genome is located near
encoded. There are several mtDNA copies per mitochondrion and the ETC in the inner mitochondrial membrane. These mutations
hundreds of mitochondria per cell (Miller et al., 2003). MtDNA can impact the ETC complexes, which leads to disruption of bioener-
exist as a pure population of wild type (homoplasmy) or a mix- getics and production of more ROS in a feedback loop (Kwong and
ture of wild type and mutant mtDNA, known as heteroplasmy. Sohal, 2000). Decreased protection of the mitochondrial genome
Proteins, such as transcription factor A of mitochondria (TFAM), due to both the lack of true histones and limited DNA repair mech-
mtDNA helicase Twinkle and mitochondrial single-stranded DNA anisms may also contribute to the higher mutation rate in mtDNA.
binding protein (mtSSB) can colocalize with mtDNA forming dis- Since high levels of ROS cause oxidative damage to cellular compo-
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 91

Fig. 1. The four complexes of the ETC and ATP synthase are located in the inner mitochondrial membrane. Electrons from NADH and FADH2 are transferred from Complex I
(NADH dehydrogenase) and Complex II (succinate dehydrogenase) to coenzyme Q (CoQ) before being transported to Complex III (cytochrome c reductase). From Complex III,
cytochrome c (Cyt c) delivers the electrons to Complex IV (cytochrome c oxidase), which ultimately reduces oxygen to water. This flow of electrons down the ETC generates
a proton electrochemical gradient, which is used by Complex V (ATP synthase) to synthesize ATP. ATP is then transported back into the intermembrane space by the ANT1-4
and beyond to the cytosol with a combination of ANT and VDAC. Protons can also enter the matrix through the uncoupling protein (UCP), which results in the uncoupling
or separation of electron transport through the ETC from ATP synthesis. Premature electron leakage from Complexes I and II form ROS as superoxide (O2 • − ), which can be
transformed into H2 O2 or the hydroxyl radical (HO• ). Subunits of each of the five complexes of the ETC, with the exception of Complex II, are encoded by both mtDNA and
nDNA as recounted in the table.

Fig. 2. A schematic diagram of the mtDNA replication machinery. POLRMT forms the RNA primer (jagged red line) needed to initiate DNA replication in conjunction with
TFAM (royal blue), TFB2 M (gray). After initiation, POLRMT, TFAM and TFB2 M separately leave the DNA and the RNA primer is degraded by RNase H1 (yellow). In a 5 to 3
direction, the Twinkle helicase (pink) unwinds dsDNA at the replication fork. The ssDNA is stabilized by mtSSB (light blue), while MGME1 (red) degrades ssDNA. MtLigIII
(white) seals the mtDNA nick. Nascent DNA (solid red line) is synthesized by Pol ␥ (green). Topoisomerases (brown) relieve the torsional tension in the DNA caused by
unwinding.

nents such as DNA, RNA, proteins and lipids, these observations The most common oxidative lesions in mtDNA are the G:C to
garnered support for the MFRTA (Chakravarti and Chakravarti, T:A transversion mutations caused by adenine pairing with the
2007; Cooke et al., 2003). guanine adduct 7,8-dihydro-8-oxo-deoxyguanosine (8-oxo-dG)
92 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

(Wang et al., 1998). Correlative evidence between increased lev- tive DNA polymerase known to exist in mammalian mitochondria,
els of 8-oxo-dG and age-related deletions in mtDNA from humans Pol ␥ is likely to produce these spontaneous errors. Comparison of
(Capel et al., 2005; Hayakawa et al., 1992; Hayakawa et al., 1991), mutation spectrum in mtDNA derived from in vivo sources with
mice (Vermulst et al., 2007), rats (Sawada and Carlson, 1987) and DNA copied in vitro by the highly purified human Pol ␥ reveals that
houseflies (Sohal and Sohal, 1991) has been reported. Conversely, over 85% of mutation detected in vivo could be recapitulated in vitro
G:C to T:A transversions are only a small fraction of the mutations by Pol ␥ (Zheng et al., 2006). This suggests that spontaneous repli-
found in aged Drosophila (Itsara et al., 2014) or humans (Kennedy cation errors by Pol ␥, and not ROS-induced damage, account for
et al., 2013; Zheng et al., 2006). The enzyme 8-oxoguanine DNA gly- the majority of base pair substitution mutations in mtDNA. Thus,
cosylase (OGG1) functions in the mitochondria to initiate removal understanding the fidelity of Pol ␥ is critical.
of the 8-oxo-dG from mtDNA, while 8-oxoGTPase (MTH1) elimi-
nates oxidized dGTP precursors to reduce their incorporation into
mtDNA. The ogg1 or ogg1/mth1 knock-out mice have 20-fold higher 4. Fidelity of DNA synthesis by pol ␥
levels of 8-oxoguanine in mtDNA (de Souza-Pinto et al., 2001) and
half the life expectancy relative to wild-type (WT) mice (Xie et al., The human catalytic subunit of Pol ␥ has high base substitution
2004), but no increase in mtDNA mutations (Halsne et al., 2012). fidelity over short repeat sequences (<3 nucleotides), as a misin-
Taken together, this suggests that ROS and 8-oxo-dG are not major sertion event only occurs approximately once in every 500,000
contributors to mtDNA mutations in WT organisms. base pairs synthesized (Longley et al., 2001). High fidelity is due
Antioxidant systems eliminate ROS. Therefore, if the MFRTA to high nucleotide selectivity and exonucleolytic proofreading. The
were sustainable then over-expressed antioxidant systems should intrinsic 3 → 5 exonuclease activity of Pol ␥ improves replica-
lead to greater longevity and fewer mtDNA mutations. This is tion fidelity at least 20-fold (Longley et al., 2001). Biochemical
the exact scenario found in transgenic mice over-expressing characterization of the human catalytic subunit found that 3 → 5
mitochondrially-targeted human catalase, which decreases H2 O2 exonuclease function could be eliminated by substituting D198 and
(Schriner et al., 2005). These same mice are also protected from E200 with alanine in the exonuclease domain (Longley et al., 1998).
mitochondrial dysfunction, decreased energy metabolism and To examine whether exonuclease function is critical to replication
lipid-induced insulin resistance in skeletal muscle (Lee et al., 2010). fidelity in vivo, several experiments were conducted. For example,
Interestingly, when human catalase is localized to peroxisomes or expression of an exonuclease-deficient Pol ␥ fusion protein in cul-
targeted to the nucleus in mice, no detectable reduction in mtDNA tured human cells resulted in the accumulation of mtDNA point
damage or significant extension of lifespan are observed (Schriner mutations (Spelbrink et al., 2000). Exonuclease-deficient Pol ␥ pan-
and Linford, 2006; Schriner et al., 2000). Over-expression of genes creatic islet cells caused impaired glucose tolerance, diabetes, and
involved in the maintenance of thiol redox homeostasis or co- increased apoptosis in ␤-cells (Bensch et al., 2007; Bensch et al.,
overexpression of Cu-Zn superoxide dismutase (Cu-ZnSOD) and 2009). In mice, transgenic overexpression of exonuclease-deficient
catalase increase the lifespan of Drosophila melanogaster by 33–48% Pol ␥ in cardiac tissue caused a greater than 23-fold increase in
(Orr et al., 2013; Orr and Sohal, 1994). When either Cu-ZnSOD mtDNA point mutations and detectable levels of large mtDNA
or catalase is over-expressed individually, only Cu-ZnSOD extends deletions (Zhang et al., 2000). These mice also presented with sec-
lifespan up to 48% while catalase has neutral or slightly negative ondary effects, such as increased apoptosis and cardiomyopathy
effects (Sun and Tower, 1999). However, when manganese super- (Zhang et al., 2005, 2003). However, the first strong experimental
oxide dismutase (MnSOD) and catalase are ectopically targeted to evidence demonstrating a causal relationship between aging and
the mitochondrial matrix of transgenic flies, lifespan is decreased the accumulation of mtDNA mutations/deletions secondary to an
up to 43% (Bayne et al., 2005). If overexpressing antioxidant systems exonuclease-deficient Pol ␥ was obtained with the creation of the
reduces ROS and leads to increased longevity, then the opposite mtDNA mutator mouse.
should hold true when these systems are knocked-out or deficient.
For example, mice deficient in either Cu-ZnSOD or MnSOD have
widespread oxidative damage and hepatocarcinogenesis or neu- 5. Exonuclease-deficient pol ␥ mouse links mtDNA
rodegeneration with reduced lifespan (Elchuri et al., 2005; Lebovitz mutations and deletions to aging
et al., 1996). On the other hand, inactivation of any of the five iso-
forms of SOD in Caenorhabditis elegans fails to decrease longevity, Independently developed by two labs, the mtDNA mutator
in spite of increased oxidative stress, and actually prolongs lifespan mouse has a homozygous knock-in that expresses a proofreading-
(Van Raamsdonk and Hekimi, 2012). Similarly, while SOD knock- deficient version of Pol ␥, resulting in elevated levels of mtDNA
out mice (Sod2−/− ) die within the first 10 days of life (Li et al., 1995), mutations (Kujoth et al., 2005; Trifunovic et al., 2004). The mice
heterozygous mice (Sod2+/− ) are viable with longevity and OXPHOS appear normal through adolescence but develop a progeroid
activity comparable to WT mice (Mansouri et al., 2006). This occurs phenotype including alopecia (hair loss) and graying, kyphosis (cur-
despite presenting with a premature aging (progeroid) pheno- vature of the spine), osteoporosis (loss of bone density), sarcopenia
type as evidenced by increased mitochondrial oxidative damage, (loss of muscle mass), weight loss, reduced subcutaneous fat,
cataracts and an impaired immune response (Van Remmen et al., presbycusis (hearing loss), reduced fertility, anemia, cardiac hyper-
2003). trophy (heart enlargement), and a reduced lifespan. Trifunovic et al.
The mixed results for whether ROS contributes to decreased (2004) also describe impaired OXPHOS with a reduction in ETC
or increased longevity may be explained by mitohormesis. In this enzyme activities and decreased ATP production.
theory, high levels of ROS promote mtDNA damage and the aged The progeroid phenotype and accumulated mutations of mtDNA
phenotype, while low levels of ROS are signaling molecules that mutator mice are not due to increased ROS production and oxida-
regulate stress response pathways and activate cellular processes tive stress (Trifunovic et al., 2005). Kujoth et al. (2005) found no
that enhance lifespan (Hekimi et al., 2011). As ROS gradually biomarkers of ROS induced damage to DNA, proteins or lipids, but
increase over time, so do the cumulative levels of mtDNA damage, report that tissue dysfunction is the result of apoptosis. In a follow-
with ROS eventually reaching unstoppable toxic concentrations up study to experimentally address the “vicious” cycle aspect of the
resulting in further mitochondrial dysfunction and aging. MFRTA, mouse embryonic fibroblasts from mtDNA mutator mice
Mutations can also arise as spontaneous errors of replication are shown to produce normal amounts of ROS and demonstrate no
during either DNA replication or repair events. As the only replica- increased sensitivity to oxidative stress (Trifunovic et al., 2005).
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 93

There are also no changes in the levels of antioxidants, protein 6. Age-related mitochondrial diseases associated with
carbonyls or aconitase enzyme activity in their tissues. defective mtDNA replication machinery
Despite only normal levels of ROS, mtDNA mutator mice do
accumulate mutations of different types: (1) a 3- to 8-fold increase 6.1. DNA polymerase 
in progressive, yet random, somatic point mutations; (2) circular
mtDNA with large deletions; and (3) multiple linear deleted mtDNA Several mitochondrial diseases have been linked to ineffective
fragments approximately 12 kb in length that are localized to the mtDNA replication by Pol ␥. To date, there are over 300 disease-
arc between the two origins of replication (Kujoth et al., 2005; associated mutations in POLG listed in the Human DNA Polymerase
Trifunovic et al., 2004). It remains highly contested as to which of Gamma Mutation Database (http://tools.niehs.nih.gov/polg). Some
these is the cause of the premature aging phenotype. associated diseases manifest in adulthood and include symptoms
Unlike the homozygous mtDNA mutator mice, mice heterozy- of aging, namely autosomal dominant and recessive forms of Pro-
gous for the Pol ␥ mutation have no progeroid phenotype or gressive External Ophthalmoplegia (adPEO/arPEO) and Parkinson’s
reduced lifespan (Kujoth et al., 2005; Trifunovic et al., 2004). The disease.
mutation load of these mice was originally reported to be indistin- PEO is associated with mtDNA point mutations and deletions
guishable from WT animals (Trifunovic et al., 2004) as determined (Hirano et al., 2001; Van Goethem et al., 2001, 2003b) first iden-
by PCR cloning and sequencing, a method that is limited by the error tified in an Italian family with heritable autosomal dominant PEO
frequency of the PCR polymerases, which is 0.13 mutations per kbp (adPEO) (Zeviani et al., 1989). PEO manifests during adulthood and
(Kujoth et al., 2005). However, using an adaptation of the more is characterized by bilateral ptosis and progressive weakening of
sensitive “random capture method,” which relies on PCR amplifi- the external eye muscles, proximal muscle weakness and wasting,
cation of single molecules for mutation detection but is not limited skeletal muscle with ragged red fibers, impaired ETC enzyme activ-
by polymerase fidelity, the heterozygous mice are found to have ity, exercise intolerance, sensory ataxic neuropathy and respiratory
a 500-fold higher mutational burden than age-matched wild-type failure due to muscle fatigue.
controls and a 30-fold increase over the oldest wild-type animals In 2001, a heterozygous mutation in Pol ␥ that substituted a
(Vermulst et al., 2007). Thus, at least at the level observed in the tyrosine with cysteine at codon 955 (Y955C) was first identified
heterozygous mice, mtDNA point mutations do not limit natural as a disease locus for adPEO, while three more missense muta-
lifespan. However, the higher mutation load observed by the ran- tions were associated with arPEO (Van Goethem et al., 2001). The
dom capture method (2000-fold higher in the homozygous mtDNA following year, several more Pol ␥ mutations were identified that
mutator mouse than WT) showed that the heteroplasmic level of cause both adPEO and arPEO (Lamantea et al., 2002). Four adPEO
mutations clearly exceeded the bioenergetic threshold to elicit the mutations (G923D, R943H, Y955C and A957S) were biochemically
premature aging phenotype. characterized (Graziewicz et al., 2004). Both R943H and Y955C
WT and heterozygous mice mainly accumulate deletions at alter direct interactions with incoming dNTPs resulting in severely
one hotspot between two 15-bp direct repeats, whereas homozy- decreased processivity and a 99% decrease in polymerase activity.
gous mutant mice accumulate deletions regardless of the presence The reduced catalytic efficiency and the profound stalling of DNA
of direct repeats (Vermulst et al., 2008). Excluding this hotspot, synthesis are predictive for the in vivo dysfunction presented by
homozygous mutant mice showed a 90-fold increase in circular PEO patients (Graziewicz et al., 2004). The Y955C mutation has a
mtDNA with large deletions, as compared to WT controls. Thus, wild-type catalytic rate, but a 45-fold decrease in binding affin-
the data suggest that deletions are the driving force behind the ity for the incoming dNTP. The proofreading deficient exonuclease
progeroid phenotype of the mtDNA mutator mouse (Vermulst domain causes a 10- to 100-fold increase in nucleotide misinser-
et al., 2008). However, these findings were contested because tion rates (Ponamarev et al., 2002). In a transgenic mouse model,
when employing Southern blots and long-range PCR, these circu- human Pol ␥ encoding Y955C was targeted to the heart. These mice
lar mtDNA molecules with large deletions are remarkably absent develop a molecular signature of PEO in the heart with cardiomy-
or present at extremely low levels, and as such, cannot be causal opathy, loss of mtDNA, increased 8-OHdG and premature death
for the advanced aging phenotype (Edgar et al., 2009). In fact, the (Lewis et al., 2007). The G923D and A957S mutations have 21% and
overall fraction of deletions in tissues from the mtDNA mutator 23%, respectively, of the catalytic activity of the wild-type enzyme,
mouse was only measured to be ∼1.0% (Kraytsberg et al., 2009). But which seems to correlate with the presentation of a milder clinical
according Vermulst et al. (2009), the random nature of the deletions phenotype (Graziewicz et al., 2004).
makes these methods insufficient for quantifying deletions. Thus, Most of the mutations in Pol ␥ are associated with arPEO. They
the number of deletions in the mtDNA mutator mouse continues are compound heterozygotes and often occur with common Pol
to be debated. ␥ mutations, such as A467T (Van Goethem et al., 2003a), W748S
Edgar et al. (2009) reported frequent point mutations in ETC (Van Goethem et al., 2004), R309L (Lamantea et al., 2002), G848S
genes in homozygous mtDNA mutator mice, while heterozygous (Lamantea et al., 2002), R627W (Van Goethem et al., 2003a), N846S
mice were not examined. The resulting amino acid substitutions (Van Goethem et al., 2003c) or T251I in cis with P587L (Lamantea
impair the stability of the ETC complexes, resulting in OXPHOS defi- et al., 2002; Scuderi et al., 2015), to name a few. Patients with arPEO
ciency. Therefore, they conclude that point mutations, particularly have the same clinical features as with adPEO, but with greater
in protein coding genes, are ultimately responsible for causing the variation in symptoms and time of onset.
progeroid phenotype in mtDNA mutator mice (Edgar et al., 2009). Parkinson’s disease (PD) is a progressive neurological disor-
Lastly, a third type of mutation found in the mtDNA mutator mouse der affecting greater than 1% of people over 60 years of age and
is the linear deleted mtDNA fragment. The deleted mtDNA was 4% of those over 80 years (Luo et al., 2015). It is characterized
comparable in all tissues and did not change over time. There- by the loss of dopaminergic neurons in the substantia nigra and
fore, these fragments are surmised to arise from stalled replication the formation of protein inclusions called Lewy bodies (Duvoisin,
due to Pol ␥ exonuclease deficiency and do not contribute to the 1992). However, the absence of Lewy bodies in the substantia nigra
progeroid phenotype (Kukat and Trifunovic, 2009). Hence, though has been found post-mortem in patients with Pol ␥ associated
it is agreed that mtDNA mutations contribute to the aging pheno- PD (Moslemi et al., 1999). Clinically, PD presents with motor and
type, it remains controversial as to whether these mutations are in non-motor symptoms. Motor dysfunctions include tremor, rigidity,
the form of point mutations, deletions or both. bradykinesia, postural instability and uncontrollable movements
while non-motor symptoms include sleep disturbances, depres-
94 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

sion, cognitive impairment and autonomic dysfunction (Jankovic, but the heterozygous knockout (+/−) mouse has a normal pheno-
2008). Aberrant mitochondrial functioning has been implicated in type (Humble et al., 2013). Currently, there are no mouse models
the pathogenesis of idiopathic PD with genetic analysis focusing on available for conditional knock-out or knockdown of POLG2. Knock-
age-related mtDNA mutations (Bender et al., 2006; Coxhead et al., down of porcine oocyte PolG2 causes an aging-like phenotype as
2015; Hudson et al., 2014; Keogh and Chinnery, 2015; Kraytsberg evidenced by decreased production of ATP and maturation promot-
et al., 2006). ing factor (MPF), reducing the fraction of oocytes that reach MII
Seven PEO families of various ethnic origins were reported to stage (Lee et al., 2015).
have Pol ␥ mutations with multiple mtDNA deletions (Luoma et al.,
2004) and reduced mtDNA copy number (Gui et al., 2015). The dom- 6.3. Twinkle
inant Y955C mutation and two new missense mutations, N468D
and A1105T, were discovered in these families (Luoma et al., 2004). Twinkle is a 72 kDA hexamer/heptamer that functions as a
Most of the women had early menopause or primary amenorrhea. 5 → 3 helicase for mtDNA replication (Korhonen et al., 2003). It
Similar results were obtained in a family with premature ovarian was classified as a helicase based on sequence similarity to bac-
failure associated with PEO and PD in which the diseases segregated teriophage T7 gene 4 primase-helicase (T7 gp4) (Spelbrink et al.,
with Y955C (Pagnamenta et al., 2006). A novel missense muta- 2001). Twinkle has two domains, a primase-like N-terminal domain
tion in Pol ␥ (A2492G) was identified in another family with PEO, and a C-terminal helicase domain (Shutt and Gray, 2006). The N-
PD, hypogonadism, hearing loss, muscle weakness and psychiatric terminal domain is important for ssDNA binding and stimulation
problems (Mancuso et al., 2004). However, it was later determined of the C-terminal helicase activity (Farge et al., 2007). Twinkle is
that the substitution is probably a neutral polymorphism as it has essential for mtDNA maintenance and copy number (Tyynismaa
a 1.1% frequency in controls (Luoma et al., 2007; Tiangyou et al., et al., 2004). It is involved in D-loop formation and localizes to the
2006). Two sisters with early-onset PD, peripheral neuropathy and nucleoid (Jemt et al., 2015; Milenkovic et al., 2013; Spelbrink et al.,
without PEO had a 37% decrease in mtDNA content with multiple 2001). Helicase activity along with strand annealing activity has
deletions and were found to be compound heterozygous for two been confirmed (Sen et al., 2012), but no primase activity has been
missense mutations (G737R and R853W) in Pol ␥ (Davidzon et al., detected to date.
2006). Several other point mutations in Pol ␥ have been identified Twinkle, encoded at chromosomal locus 10q24, was origi-
in patients diagnosed with a combination of PEO and PD, such as nally identified through sequencing analysis of a collection of PEO
heterozygous S511N with L463F (Hudson et al., 2007), homozygous patients with unidentified disease causality (Spelbrink et al., 2001).
W748S (Remes et al., 2008), heterozygous F943H (Brandon et al., In all, 31 dominant and 3 recessive mutations have been recognized,
2013), heterozygous P587L and W748S (Ylonen et al., 2013) and the majority of which cause age-related disease (Sanchez-Martinez
heterozygous H945L (Delgado-Alvarado et al., 2015). Lastly, Pol ␥ et al., 2012; Spelbrink et al., 2001). Most of the mutations cluster in
contains a polyglutamine tract in the N-terminal region encoded a region that is predicted to be the linker region (Sanchez-Martinez
by a CAG-repeat and rare variants in the CAG-repeat were found in et al., 2012). In vitro analysis on some of these mutations has had
patients with idiopathic PD (Luoma et al., 2007). However, a sim- mixed results concerning helicase activity despite some mutants
ilar analysis revealed no changes (Taanman and Schapira, 2005). displaying altered DNA binding affinity, nucleotide hydrolysis or
In support of the latter study, another group reported that there thermal stability (Holmlund et al., 2009; Korhonen et al., 2008;
is no contribution of common genetic variants of Pol ␥ contribut- Longley et al., 2010; Matsushima et al., 2008).
ing to the etiology of idiopathic PD (Tiangyou et al., 2006). Despite Overexpression of adPEO Twinkle mutations in HEK293 cells
these two confounding studies, the role of Pol ␥ in contributing to cause an accumulation of replication intermediates suggesting that
mtDNA mutations and reduced copy number thus impacting the these mutations are triggering replication stalling (Goffart et al.,
pathogenesis of PEO and PD patients appears convincing. 2009; Wanrooij et al., 2007). These mutants have decreased mtDNA
copy number and enlarged nucleoids. Taken together, these data
6.2. PolG2 support Twinkle’s role in mtDNA replication.
A homozygous Twinkle knock-out mouse is embryonic lethal at
PolG2 is a 110 kDa homodimeric accessory subunit for the cat- E8.5, but a conditional tissue-specific knockout in heart and skeletal
alytic subunit PolG (Carrodeguas et al., 2001) that functions to muscle is viable. This mouse has progressive mtDNA depletion lead-
enhance processivity of DNA replication by augmenting the binding ing to severely impaired mtDNA expression and respiratory chain
of PolG to DNA (Lim et al., 1999). Of the handful of heterozy- deficiency (Milenkovic et al., 2013). Conversely, systemic Twinkle
gous mutations identified in POLG2, G451E (Longley et al., 2006) overexpression in a transgenic mouse causes a significant increase
and R369G (Craig et al., 2012; Young et al., 2011) are linked to in mtDNA copy number (Tyynismaa et al., 2004); however, the mice
PEO. Homodimeric mutants have problems stimulating PolG pro- have enlarged nucleoid structures with impaired mtDNA replica-
cessivity, with G451E being more severely defective than R396G tion and transcription (Ylikallio et al., 2010). Depleting Twinkle
(Craig et al., 2012; Longley et al., 2006; Young et al., 2011). in Drosophila Schneider cells induces slow growth, limits viability
G451E homodimers also fail to enhance binding of PolG to DNA, and reduces mtDNA copy number 5-fold (Matsushima and Kaguni,
while R369G has weaker binding to PolG. Only heterozygotes have 2007). Overexpressing Twinkle in these same cells causes a modest
been clinically identified, and recent biochemical characteriza- increase in mtDNA copy number (Matsushima and Kaguni, 2007).
tion of heterodimers has found that the WT/G451E heterodimer Taken together, these data further indicate that Twinkle is essential
has significantly reduced stimulation of PolG processivity due to for mtDNA replication.
weaker binding to both PolG and DNA (Young et al., 2015). Thus MtDNA typically exists as monomeric double-stranded, cova-
G451E PolG2 is a dominant negative mutation causing adPEO. lently closed circles. In adult human hearts however, mtDNA has
The WT/R369G heterodimer functions like the wild-type enzyme been shown to exist as dimers and is organized into highly complex,
in vitro. Though WT/G451E and WT/R369G localize to mtDNA branched molecules (Pohjoismaki et al., 2009). This form of mtDNA
like WT PolG2, bioenergetic analysis of transfected HEK-293 cells has not been detected in human infants (Pohjoismaki et al., 2010) or
finds that both mutant cell lines have significantly reduced reserve found endogenously in other species (Pohjoismaki et al., 2009). Sys-
capacity (Young et al., 2015). Skeletal muscle with reduced reserve temic overexpression of Twinkle in mouse hearts causes an increase
capacity has muscle weakness and exercise intolerance, hallmarks in complex DNA forms and recombination junctions similar to adult
of PEO. A complete POLG2 knock-out is embryonic lethal in mice, human hearts (Pohjoismaki et al., 2009). Heart tissue samples from
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 95

patients harboring a Twinkle duplication mutation (amino acids embryonic lethal at E8.5, whereas the heterozygous mouse remains
352–364) that causes adPEO have mtDNA that is lacking in these viable with reduced mtDNA copy number, reduced mitochondrial
complex higher order structures (Pohjoismaki et al., 2010). This transcription and impaired ETC activity in the heart (Larsson et al.,
suggests Twinkle has an important role in maintenance of mtDNA 1998). Several conditional knock-out mice have been created tar-
higher order structures in adult human heart. geting cardiomyocytes (Li et al., 2000; Wang et al., 1999), skeletal
Mice overexpressing Twinkle are protected from left ventric- muscle cells (Wang et al., 1999; Wredenberg et al., 2002), pancre-
ular remodeling in the heart (Tanaka et al., 2013). This mouse atic ␤-cells (Silva et al., 2000), pyramidal neurons (Sorensen et al.,
attenuates VO-induced eccentric hypertrophy and improves car- 2001) and midbrain dopamine (DA) neurons (Ekstrand et al., 2007).
diac function while suppressing the production of mitochondrial For all the targeted cell types, TFAM depletion leads to mtDNA
ROS and subsequent redox-sensitive signals (Ikeda et al., 2015). depletion (not reported for midbrain DA neurons), mitochondrial
The Twinkle overexpression mouse was crossed with a manganese transcript depletion (not reported for midbrain DA neurons) and
superoxide dismutase heterozygous (MnSOD+/− ) knockout mouse respiratory chain deficiency, suggesting that TFAM is essential
(Pohjoismaki et al., 2013). The MnSOD+/− mouse has increased for mtDNA maintenance in vivo (Ekstrand et al., 2007; Li et al.,
oxidative damage leading to replication stalling and the forma- 2000; Silva et al., 2000; Sorensen et al., 2001; Wang et al., 1999;
tion of unicircular dimers in the heart (Li et al., 1995). Crossing Wredenberg et al., 2002). Both heart- and muscle-specific condi-
the MnSOD+/− mouse with the systemic Twinkle overexpression tional TFAM knock-out mice have cardiomyopathy with greater
mouse abolishes the MnSOD+/− phenotype (Pohjoismaki et al., severity in the heart-specific knock-out (Li et al., 2000; Wang et al.,
2013). Additionally, a transgenic knock-in mouse was created 1999). Skeletal muscle-specific TFAM knock-out mice have myopa-
expressing a dominant-negative K320E Twinkle in the heart (Baris thy with an abundance of COX-deficient red ragged fibers similar
et al., 2015). These mice develop ventricular arrhythmia with aging to mitochondrial disease patients (Wredenberg et al., 2002). TFAM
due to dysfunctional ETC enzyme activity. This suggests Twinkle deletion in pancreatic ␤-cells leads to diabetic mice (Silva et al.,
provides cardioprotection and could be a potential therapy for heart 2000). TFAM deletion in mouse pyramidal neurons causes mito-
failure patients. chondrial late-onset neurodegeneration (MILON) (Sorensen et al.,
Transgenic mice overexpressing dominant PEO Twinkle muta- 2001) whereas deletion in mouse midbrain DA neurons leads to PD
tions (either a point mutant A360T or an in-frame duplication of (Ekstrand et al., 2007). Heart failure, myopathy, diabetes, neurode-
amino acids 353–365) were created (Tyynismaa et al., 2005). The generation and PD are all age-related diseases strongly implicating
Twinkle duplication mouse, herein referred to as the deletor mouse, TFAM in aging.
presents with mitochondrial myopathy around 12 months of age Despite the potential for TFAM to play a role in PD, separate
(Tyynismaa et al., 2005). The phenotype of the deletor mouse is studies of 250 and 301 patients, respectively, identified no TFAM
strikingly similar to PEO patients, thus creating a PEO mouse model mutations (Alvarez et al., 2008; Bertram et al., 2007). To date, there
(Tyynismaa et al., 2010). The deletor mouse phenotypes can be are no confirmed TFAM clinical mutations, although the genetic
attenuated with a lipid-lowering agent, such as bezafibrate, how- variant rs2306604 may be linked to Alzheimer’s disease (Belin et al.,
ever this treatment causes severe hepatic side effects (Yatsuga and 2007) pending more comprehensive clinical and biochemical anal-
Suomalainen, 2012). The hepatic side effects are not expected in yses. However, TFAM has shown promise for being an anti-aging
humans as the dosage amounts are much lower. Thus, bezafibrate therapeutic agent. Studies of cardiomyocytes (Suarez et al., 2008),
has the potential to be a therapy for PEO patients (Yatsuga and SH-SY5Y neuroblastomas (Xu et al., 2009), and rat pancreatic islet
Suomalainen, 2012). cell lines (Gauthier et al., 2009) have been performed. TFAM over-
A conditional transgenic mouse overexpressing the PEO Twin- expression in cardiomyocytes improves mitochondrial function
kle duplication targeted to neurons has been generated (Song et al., when cells are exposed to hyperglycemic conditions (Suarez et al.,
2012). The mouse has neuron loss in the substantia nigra with 2008). Exogenous introduction of recombinant TFAM (rhTFAM)
significant mtDNA deletions and reduction of the protein Parkin to cardiomyocytes increases mtDNA copy number and attenuates
resulting in motor abnormality (Song et al., 2012). This mouse is hypertrophy, suggesting TFAM as a potential therapeutic agent to
used to study substantia nigra specific accumulation of mtDNA combat heart failure.
deletions in Parkinson’s disease. Mice heterozygous for Pdx1 (Pdx+/− ), a transcription factor
Moderate overexpression of Twinkle adPEO disease mutations, needed for pancreatic development and ␤-cell maturation/survival,
I334T or A442P, in Schneider cells allows for normal growth despite show diabetic symptoms and significantly reduced TFAM transcript
a 2-fold reduction in copy number, while other adPEO disease levels (Oliver-Krasinski and Stoffers, 2008). Rat pancreatic islet cells
mutations do not (Matsushima and Kaguni, 2007). However, over- transduced with a dominant-negative Pdx1 variant have the same
expression of A442 P in Drosophila melanogaster is lethal at the TFAM transcript reduction (Gauthier et al., 2009). TFAM overex-
pupal stage (Matsushima and Kaguni, 2007). MtDNA copy num- pression in rat pancreatic islet cells normalized Nd1 mRNA levels
ber is severely depleted along with mtDNA-encoded transcripts and restored glucose-induced ATP generation along with insulin
leading to reduced ETC Complex IV activity. secretion (Gauthier et al., 2009). This suggests TFAM could similarly
be a therapeutic agent for diabetes.
6.4. Transcription Factor A of mitochondria (TFAM) SH-SY5Y neuroblastoma cells were used to study TFAM’s role
in neurodegenerative disease. Overexpression of TFAM in these
TFAM, a member of the high mobility group (HMG) proteins cells provides protection against A␤1-42 induced mtDNA oxida-
(Parisi and Clayton, 1991), is a 25 kDa protein that binds upstream tive damage and MPP+ induced damage (Xu et al., 2009) suggesting
of both the heavy- and light-strand promoters of mtDNA (Fisher TFAM could also be used as a therapeutic agent in neurodegenera-
and Clayton, 1985; Fisher et al., 1987). TFAM was originally iden- tive disorders. To test this premise, SH-SY5Y neuroblastoma cybrids
tified as a regulator of mtDNA transcription (Fisher and Clayton, were created that mimic sporadic PD (Xu et al., 2009). These cells
1985). More recently TFAM has been identified as a core protein were then treated with protofection technology using a modified
of the mtDNA nucleoid (Kanki et al., 2004), a regulator of mtDNA rhTFAM that has an N-terminal 11 Arginine transduction domain
copy number (Ekstrand et al., 2004) and potentially a key factor in followed by a mitochondrial localization signal (Khan and Bennett,
mtDNA replication (Campbell et al., 2012). 2004). Uptake and mitochondrial localization of rhTFAM restored
Homozygous and heterozygous knock-outs of TFAM have been mitochondrial bioenergetics to impaired cells without altering nor-
created in mice (Larsson et al., 1998). The homozygous knock-out is
96 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

mal cells, further supporting the use of rhTFAM as a therapeutic In mice, an RNase H1 knockout is embryonic lethal at E8.5 when
agent (Iyer et al., 2009). mtDNA replication begins suggesting the mitochondrially located
Mice overexpressing human TFAM (hTFAM) are protected from RNase H1 is essential for development (Cerritelli et al., 2003). This is
memory impairment (Hayashi et al., 2008) and cardiac damage similar to the RNaseH2 null embryos that show lethality and a large
after myocardial infarction (Ikeuchi et al., 2005). Overexpres- number of ribonucleotides in the genomic DNA (Reijns et al., 2012).
sion increases mtDNA copy number, attenuates volume overload The RNaseH1 KO embryos also have a loss of mtDNA indicating
induced eccentric hypertrophy, improves cardiac function and sup- that RNase H1 has a vital role in mtDNA replication. In Drosophila
presses the production of mitoROS (Ikeda et al., 2015). However, no melanogaster, RNase H1 knockout is also lethal during the last larval
increases in ETC enzyme activities suggest a disassociation between instar stage, when rapid mtDNA replication is important for sur-
mtDNA copy number and the ETC (Ikeda et al., 2015; Ikeuchi et al., vival (Filippov et al., 2001). Knockdown of RNase H1 in human bone
2005). Another mouse model system crossed mice overexpress- osteosarcoma cells (143B TK− ) led to lower mtDNA copy number
ing TFAM with the mito-mice that have mutant mtDNA carrying a and complete stalling of replication (Ruhanen et al., 2011), sug-
deletion of 4696 bp leading to an overall increase in mtDNA con- gesting RNase H1 is vital for mtDNA replication. Mouse embryonic
tent and copy number without a decrease in the amount of deleted fibroblasts lacking RNaseH1 revealed retention of primers at OriH
mtDNA (Nishiyama et al., 2010). TFAM overexpression improved and OriL, which causes an obstacle to the replication machinery
the symptoms caused by mitochondrial deficiency signifying that upon the next round of replication (Holmes et al., 2015).
WT mtDNA is important for maintaining mitochondrial function
irrespective of mutant mtDNA proportions (Nishiyama et al., 2010).
However, Ylikallio et al. (2010) suggest caution should be taken 7. The contribution of clonal expansion of mtDNA
when considering the use of either recombinant TFAM and/or mutations to mitochondrial dysfunction
mtDNA as therapeutic agents since increased mtDNA copy number
causes nucleoid enlargement and defective replication and tran- If aging is caused by mtDNA mutations, then how do these muta-
scription. tions lead to dysfunction? The percentage of mutated mtDNA must
be between 60% and 90% in a given cell before dysfunction is mea-
6.5. Mitochondrial genome maintenance exonuclease 1 (MGME1) surable (Edgar and Trifunovic, 2009) (Fig. 3). Elevated heteroplasmy
can occur by clonal expansion of mutated mtDNA, mitotic seg-
MGME1 is a 37 kDa monomeric protein that is a regation, positive selection and/or random genetic drift. Genetic
mitochondrially-located, metal-dependent exonuclease (Szczesny drift has been proposed to be the more likely situation due to the
et al., 2013). MGME1 belongs to the PD-(D/E)XK nuclease super- relaxed replication of mtDNA, which is untethered to the cell cycle
family and is a close homologue of the bacterial RecB nuclease (Elson et al., 2001). However, reduced size gives deleted mtDNA a
(Kornblum et al., 2013; Szczesny et al., 2013). MGME1 selectively selective replication advantage over wild-type mtDNA, supporting
degrades ssDNA in both 3 → 5 and 5 → 3 directions, preferring positive selection (Diaz et al., 2002; Hayashi et al., 1991; Kowald
the latter (Kornblum et al., 2013; Szczesny et al., 2013). It can and Kirkwood, 2011).
degrade DNA flap and RNA-DNA flaps (Kornblum et al., 2013; Clonal expansion of mtDNA mutations has been described in
Szczesny et al., 2013), but not circular DNA (Kornblum et al., 2013). numerous tissues from aged individuals, such as skeletal muscle
HeLa cells and patient fibroblasts with MGME1 knocked-down (Fayet et al., 2002; Muller-Hocker et al., 1993), substantia nigra
have increased amounts and lengths of 7S mtDNA (Nicholls et al., neurons (Cottrell et al., 2001; Kraytsberg et al., 2006), heart (Muller-
2014). Conversely, patient fibroblasts that were transfected with Hocker, 1989) and colonic crypts (Taylor et al., 2003). Typically,
MGME1 have 7S mtDNA decreased in both amount and length, the clonal expansion of a mtDNA mutation results in highly focal
especially the 11 kb linear dsDNA fragments whose ends map to ETC defects, such as cytochrome c oxidase deficiency. For example,
the two classically derived replication origins, OH and OL . MGME1 the pattern of mosaic respiratory chain deficiency in the heart of
interacts directly with PolG suggesting a role in mtDNA replication, the mtDNA mutator mice is thought to indicate clonal expansion
likely providing 5 → 3 exonuclease activity (Nicholls et al., 2014). (Trifunovic et al., 2004). A similar pattern of focal ETC defects has
Three different families presenting with PEO, skeletal mus- also been reported in human brain (Taylor et al., 2014), colorec-
cle wasting/weakness, emaciation and respiratory distress bear tal epithelium (Greaves et al., 2014) and colonic crypt stem cells
MGME1 mutations (Kornblum et al., 2013). Two of these families (Taylor et al., 2003) and deemed to be the result of clonal expansion
(five affected individuals) have the same homozygous nonsense of pre-existing mutations.
mutation (c.456G > A, P.W152*) while a different sporadic case The timing of when mtDNA mutations occur remains debated
presents with homozygous c.698A > G, P.W233C point mutation. (Khrapko, 2011; Popadin et al., 2014). Somatic mtDNA mutations
The W152* nonsense mutation truncates MGME1 upstream of the were originally thought to occur late in life when the mtDNA
nuclease motif. W152* fibroblasts have no detectable full length is more prone to chemical damage and thereafter accumulate
or truncated MGME1. Protein folding appears to be impaired in exponentially (Khrapko, 2011). However, in order to reveal a patho-
W233C substituted MGME1. logical effect caused by clonal expansion, it was argued that these
mutations would have to occur early in life (Elson et al., 2001). For
6.6. Ribonuclease H1 (RNase H1) instance, the linear accumulation of mutations from mid-gestation
to late adult life as seen in the mtDNA mutator mice, along with
RNase H1, first identified in 1969 (Stein and Hausen, 1969), is a the same mutations between tissues, implies that the mutations
29 kDa monomeric protein that functions in both the nucleus and occurred extremely early in development (Trifunovic et al., 2004).
the mitochondria (Cerritelli et al., 2003). Alternative start codons This hypothesis was later confirmed with ultra-deep sequencing
permit translation of the mitochondrial (first AUG) and nuclear to study the genome-wide mtDNA mutation load in the mtDNA
(second AUG) forms of RNase H1 (Cerritelli and Crouch, 2009). mutator mouse (Ameur et al., 2011) and in human tissues (He et al.,
RNase H1 cleaves RNA/DNA hybrids containing at least four ribonu- 2010). Using Digital Deletion Detection (3D), a newer technology
cleotides (Ohtani et al., 1999). RNase H1 associates with the protein that allows for the study of rare, low occurring mtDNA deletions,
P32 suggesting RNase H1 involvement in mitochondrial pre-rRNA Taylor et al. (2014) report that although the total deletion load
processing (Wu et al., 2013). Mutations in the catalytic domain have increases in human brain with aging, the number and diversity
been identified in PEO patients (Reyes et al., 2015). of deletions remains constant. Therefore, the authors agree that
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 97

Fig. 3. Clonal expansion and heteroplasmy. Individual mitochondria contain many copies of mtDNA. During replication, a mutation and/or deletion can occur. This damage
accumulates by clonal expansion. When the fraction of mutated mtDNA relative to total mtDNA exceeds 60% depending on the tissue affected, dysfunction and disease
are measurable. The mitochondria are represented as the brown ovals, green circles are healthy mtDNA and red circles are damaged mtDNA. The yellow lightening bolt
represents damage occurring during replication and the blue circles represent the replication machinery.

mtDNA mutations associated with premature aging occur from Fusion and fission are opposing actions that work together to
early, pre-existing deletions and not de novo deletions (Taylor et al., maintain the number, size, shape and function of mitochondria
2014). Mutations that occur in early development also support (Chan, 2012). Fusion forms interconnected mitochondrial networks
the stem cell hypothesis in which mtDNA mutagenesis modu- that allow for the exchange of metabolites, proteins and DNA
lates somatic stem cell dysfunction. For instance, mtDNA mutator between mitochondria. There are three GTPases involved in fusion:
mice were found to have impaired neural and hematopoietic pro- the mitofusins, Mfn1 and Mfn2, for the outer mitochondrial mem-
genitor cells during embryogenesis, which resulted in decreased brane and OPA1 for the inner mitochondrial membrane. Depletion
self-renewal, quality and amount of quiescent cells and may con- of any of one these proteins results in severe dysfunction in res-
tribute to the progeroid phenotype (Ahlqvist et al., 2012). piratory capacity and fragmentation of mitochondria (Chen et al.,
Conversely, the spatial distribution of expanded mutations, 2005, 2010; Song et al., 2009), while homozygous knock-outs of
within or between tissues, implicates mutations later in life either Mfn1 or Mfn2 are embryonic lethal (Chen et al., 2003). Addi-
(Khrapko et al., 2004). Should mutations occur during embryo- tionally, conditional deletion of Mfn1 and Mfn2 in mouse skeletal
genesis, one would expect specific mutations to be more widely muscle correlates with increased accumulation of mtDNA point
distributed among tissues as seen in the mtDNA mutator mice. mutations and deletions (Chen et al., 2010). Fission segregates the
However, mutations limited to a local area, such as a single cell mitochondria into daughter cells during cell division and allows for
or a particular tissue, would be an indication of mutagenesis later the selective elimination of damaged mitochondria by mitophagy
in life. For example, Nicholas et al. (2009) describe that mtDNA (Chen and Chan, 2009; Twig et al., 2008). The dynamin-related pro-
deletions from two distinct but proximal areas of muscle from an tein Drp1 and its four potential receptors: Fis1, Mff, MiD49 and
82-year-old individual are completely different with no mutations MiD51/MIEF1 that recruit Drp1 to the outer mitochondrial mem-
in common. Within each area, the mutations are clonally expanded brane are critical for fission (Chan, 2012). Loss of function in these
but are of late origin since the mutations are not shared between proteins results in severe fission defects (Lee et al., 2004). Similar to
the same tissue (Nicholas et al., 2009). More recently, a combined fusion proteins, silencing Drp1 and Fis1 in human cells that are het-
scenario for mutations has been proposed in which mtDNA muta- eroplasmic for the pathological A3243 G mtDNA mutation results in
tions seen in aging tissue are a mixture of pre-existing and de novo increased mtDNA mutation loads as compared with untransformed
mutations that expand at different rates depending on cell type cells (Malena et al., 2009). As expected, ageing is associated with
and/or conditions placed upon the cells (Popadin et al., 2014). decreases in both fission and fusion (Bereiter-Hahn, 2014; Chauhan
et al., 2014; Jendrach et al., 2005; Seo et al., 2010).
As previously discussed, mtDNA mutations increase with age,
8. Mitochondrial dynamics and aging but heteroplasmy must exceed ∼60% in a given cell to have mea-
surable dysfunction. Clonal expansion to this level may be delayed
Mitochondria do not exist as solitary, rigidly structured via fusion-directed complementation, resulting in dilution of the
organelles, but rather are continuously remodeled through the mutated mtDNA through content mixing and, ultimately, protec-
dynamic processes of biogenesis, fusion, fission and mitophagy tion of mitochondrial function (Nakada et al., 2009; Seo et al., 2010).
(Chan, 2006, 2012). Mitochondrial biogenesis is needed to meet Support for inter-mitochondrial complementation as a protectant
the energy demands of the cell and to replace the damaged of respiratory function has been shown in human cells and in
mitochondria removed by mitophagy. Biogenesis is mediated by mice. In the first study, fusion with WT mtDNA rescued respiratory
the transcriptional co-activator, peroxisome-proliferator-activated enzyme activity in two cell lines rendered respiration-deficient by
receptor-␥ coactivator 1␣ (PGC-1␣) (Ventura-Clapier et al., 2008). pathogenic mitochondrial tRNA mutations (Ono et al., 2001). The
Age-related symptoms, particularly sarcopenia, have been linked to same group studied the pathogenesis of mutant mtDNA carrying
decreased expression of PGC-1␣ and hence, decreased mitochon- a 4696-basepair deletion in mice. Mixing with WT mtDNA pre-
drial biogenesis (Joseph et al., 2013; Wenz, 2011).
98 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

Fig. 4. Factors causal to aging. Spontaneous errors in the mitochondrial replication machinery cause mtDNA damage. Accumulations of this mtDNA damage occur via
clonal expansion until tissue specific heteroplasmy is reached and mitochondrial dysfunction is measurable. Increased mitochondrial dysfunction along with decreased
mitochondrial dynamics leads to aging and age-related diseases. Mitochondrial dysfunction may also cause increased apoptosis, which is associated with aging. At low
levels, ROS triggers stress response and biogenesis pathways and increases mitochondrial dynamics, which are protective and prevent aging. Caloric restriction and exercise
functions similarly. Lastly, high levels of ROS can cause damage to cellular components, such as DNA, RNA, protein and lipids, which may also contribute to aging.

served respiratory function until the mutated mtDNA exceeded 85% medium and maximum lifespan and delay the onset of age-related
(Nakada et al., 2001). In another study, Chen et al. (2010) knocked diseases in a variety of species, including yeast (Kaeberlein et al.,
out Mfn1 in the mtDNA mutator mouse to demonstrate the protec- 2007), C. elegans (Lee et al., 2006), rodents (Masoro, 2009) and Rhe-
tive effects of mitochondrial fusion on existing mtDNA mutations. sus monkeys (Colman et al., 2014).
Although the double knock-out leads to a synthetic neonatal lethal- Several mechanisms have been proposed to explain the effects
ity, mouse embryonic fibroblasts recover and show significantly of CR on extending longevity. Originally, it was shown that CR
reduced production of ATP due to severe disruption in ETC com- decreases ROS production in rodents, particularly at Complex I,
plexes. Mitochondrial fusion apparently allows the mtDNA mutator thereby preventing oxidative damage in mitochondria from heart
mouse to carry a higher mutational load without affecting respira- (Gredilla et al., 2002), liver (Lopez-Torres et al., 2002), skeletal
tory function (Chen et al., 2010). muscle (Lass et al., 1998) and brain (Sanz et al., 2005). However,
When fusion-directed complementation cannot decrease the mitohormesis predicts low levels of ROS, possibly due to increas-
mtDNA mutational load, severely damaged mitochondria may be ing the respiration rate, can activate systemic defense mechanisms
eliminated. In mammalian cells, Twig et al. (2008) describe fis- enhancing longevity (Hekimi et al., 2011). CR apparently increases
sion of individual mitochondria to form daughter organelles with oxidative metabolism, respiration and biogenesis through multi-
unequal mitochondrial membrane potentials (YMs). Daughter ple pathways. For example, CR induces mitochondrial biogenesis
mitochondria with high M undergo subsequent fusion. Daugh- and increases respiration through activation of sirtuin 1 (SIRT1), an
ter mitochondria with low M (depolarized) have reduced levels NAD+ -dependent histone deacetylase, and its downstream induc-
of OPA1. Selective fusion segregates them from the population, tion of PGC-1␣ (Chistiakov et al., 2014; Martin-Montalvo and de
after which they may be removed by mitophagy, thus maintaining a Cabo, 2013). CR likewise activates nutrient- or energy-sensing
healthy mitochondrial population and upholding the bioenergetic pathways such as 5 -adenosine monophosphate-activated protein
efficiency of the cell (Twig et al., 2008). As with the fission and fusion kinase (AMPK), which can then both activate PGC-1␣ and inacti-
processes, insufficient mitophagy and autophagy contribute to the vate the mechanistic target of rapamycin (mTOR) (Chistiakov et al.,
accumulation of mutant mtDNA and are a factor in aging and age- 2014; Ruetenik and Barrientos, 2015). For example, CR induced
related diseases (Gaziev et al., 2014; Green et al., 2011; Hubbard mitochondrial biogenesis in this manner and improved bioener-
et al., 2012; Rubinsztein et al., 2011). getic efficiency in a variety of cell lines (Lopez-Lluch et al., 2006).
CR-increased mitochondrial biogenesis improved mitochondrial
function and rescued some progeroid phenotypes in both humans
9. Interventions to attenuate the symptoms of aging and mtDNA mutator mice (Civitarese et al., 2007; Dillon et al.,
2012). CR has also been reported to enhance genomic stabil-
Although there is potential for the genetic manipulation of the
ity by reversing mtDNA damage (Cassano et al., 2004; Civitarese
proteins of the mtDNA replication machinery to be used as future
et al., 2007) and/or promoting mtDNA repair (Cabelof et al., 2003;
therapies for aging and age-related diseases, currently, the best-
Heydari et al., 2007).
known strategies for reducing the symptoms of aging caused by
Various types of exercise training verifiably improve mitochon-
mitochondrial dysfunction appear to be caloric restriction (CR) and
drial oxidative function in aging and mitochondrial myopathic
exercise. CR is the reduction in caloric intake by 10–40% while
muscles (Irving et al., 2015; Jeppesen et al., 2006; Short et al.,
providing essential nutrients and vitamins, thereby eschewing mal-
2004; Short et al., 2003; Wenz et al., 2009). As with CR, exer-
nutrition (Testa et al., 2014). CR has been shown to extend both
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 99

cise triggers mitochondrial biogenesis and increases ETC enzyme that high levels of ROS do not contribute to the aging pheno-
levels through upregulation of PGC-1␣ and/or TFAM in aged skele- type through oxidative damage to cellular components, such as
tal muscle (Barbieri et al., 2015; Kang et al., 2013). Exercise also proteins and lipids. However, ROS does not appear to be the ini-
influences mitochondrial dynamics by changing fusion and fis- tiating or critical factor. The theory of mitohormesis states that
sion protein expression and by promoting mitochondrial turnover low levels of ROS are important signaling molecules that regu-
through mitophagy (Drake et al., 2016). For instance, increased lev- late stress response pathways and enhance longevity. Therefore,
els of Drp1 were measured in skeletal muscle of mice after 60 min of clonally expanded mtDNA mutations caused by replication errors
endurance treadmill running, while Mnf1 and OPA1 did not signifi- lead to mitochondrial dysfunction. It is this process, in conjunction
cantly change (Pagano et al., 2014). However, acute exercise in rats with altered mitochondrial dynamics, that emerges as the driving
increased mRNA expression of Mnf1 and Mfn2 24 h post-exercise force behind aging (Fig. 4). Here, we showed that deficiency and/or
(Cartoni et al., 2005). In two human studies, Fis1 expression was dysfunction in many of the nuclear-encoded mtDNA replication
higher in regularly active individuals as compared to their seden- proteins, including Pol ␥, PolG2, Twinkle, TFAM, MGME1 and RNase
tary age-matched controls (Bori et al., 2012) and Fis1, Mfn1 and H1, results in age-related diseases and/or phenotypes both in vitro
Mfn2 proteins were upregulated after twelve weeks of aerobic and in vivo. Future studies focused on understanding the mecha-
exercise training in older men (Konopka et al., 2014). Furthermore, nisms of action are needed to effectively delay aging and prevent
mitophagy/autophagy markers were increased in normal mice fol- age-related diseases.
lowing short-term intense treadmill exercise (Grumati et al., 2011; Funding
Saleem et al., 2014). Similarly, resistance and endurance exercise This work was supported by the Intramural Research Program
training resulted in increased autophagy activity and reduced apo- of the NIH, National Institute of Environmental Health Sciences (ES
ptosis in aged skeletal muscle (Kim et al., 2013; Luo et al., 2013). 065078).
Even five months of endurance exercise in the mtDNA mutator
mouse moderated systemic apoptosis, thus helping to rescue the
Acknowledgments
progeroid phenotype in these mice (Safdar et al., 2011).
Of great importance is whether exercise can alter the aug-
The authors would like to thank Drs. Scott Lujan, Matthew
mented mtDNA mutation load observed in aging patients and/or
Longley and Rachel Krasich for critically reading and editing this
patients with mitochondrial myopathies. In the mtDNA mutator
manuscript and Ms. Maggie Humble for reviewing. We also thank
mouse, endurance exercise improves oxidative capacity, prevents
Dr. Scott Lujan for his expertise with 3-D graphics.
or repairs mtDNA depletion and mutations through tumor suppres-
sor protein p53 and increases lifespan (Safdar et al., 2011; Safdar
et al., 2015). However, others have reported that although oxidative References
capacity is improved, there are no changes in the mtDNA mutation
load (Jeppesen et al., 2009; Taivassalo et al., 2006). A promising Ahlqvist, K.J., Hamalainen, R.H., Yatsuga, S., Uutela, M., Terzioglu, M., Gotz, A.,
Forsstrom, S., Salven, P., Angers-Loustau, A., Kopra, O.H., Tyynismaa, H.,
alternative is gene shifting, the activation of satellite muscles cells Larsson, N.G., Wartiovaara, K., Prolla, T., Trifunovic, A., Suomalainen, A., 2012.
or dormant myofibers that do not harbor any mtDNA mutations Somatic progenitor cell vulnerability to mitochondrial DNA mutagenesis
to re-enter the cell cycle and fuse with existing mature myofibers underlies progeroid phenotypes in Polg mutator mice. Cell Metab. 15, 100–109.
Alvarez, V., Corao, A.I., Sanchez-Ferrero, E., De Mena, L., Alonso-Montes, C., Huerta,
and shift the heteroplasmy ratio in favor of WT over mutated C., Blazquez, M., Ribacoba, R., Guisasola, L.M., Salvador, C., Garcia-Castro, M.,
mtDNA (Taivassalo et al., 1999). This activation is accomplished Coto, E., 2008. Mitochondrial transcription factor A (TFAM) gene variation in
with either eccentric or concentric resistance exercise, as shown Parkinson’s disease. Neurosci. Lett. 432, 79–82.
Ameur, A., Stewart, J.B., Freyer, C., Hagstrom, E., Ingman, M., Larsson, N.G.,
in a patient with the G12315A tRNAleu(CUN) point mutation, where Gyllensten, U., 2011. Ultra-deep sequencing of mouse mitochondrial DNA:
resistance training significantly improved heteroplasmic ratios and mutational patterns and their origins. PLoS Genet. 7, e1002028.
increased the number of respiratory-competent fibers in mus- Barbieri, E., Agostini, D., Polidori, E., Potenza, L., Guescini, M., Lucertini, F.,
Annibalini, G., Stocchi, L., De Santi, M., Stocchi, V., 2015. The pleiotropic effect
cle fibers (Taivassalo et al., 1999). Resistance training in patients
of physical exercise on mitochondrial dynamics in aging skeletal muscle. Oxid.
with a large mtDNA deletion led to increased muscle strength, Med. Cell. Longevity 2015, 917085.
myofiber damage and regeneration, improved oxidative capacity, Baris, O.R., Ederer, S., Neuhaus, J.F., von Kleist-Retzow, J.C., Wunderlich, C.M., Pal,
M., Wunderlich, F.T., Peeva, V., Zsurka, G., Kunz, W.S., Hickethier, T., Bunck,
increased satellite cell portion (Murphy et al., 2008) and apparent
A.C., Stockigt, F., Schrickel, J.W., Wiesner, R.J., 2015. Mosaic deficiency in
mtDNA shifting (Spendiff et al., 2013). Preliminary data suggests mitochondrial oxidative metabolism promotes cardiac arrhythmia during
that mtDNA shifting also occurs in aging skeletal muscle following aging. Cell Metab. 21, 667–677.
resistance exercise training (Tarnopolsky, 2009). Bayne, A.C., Mockett, R.J., Orr, W.C., Sohal, R.S., 2005. Enhanced catabolism of
mitochondrial superoxide/hydrogen peroxide and aging in transgenic
Adherence to a strict CR regime and/or exercise training has Drosophila. Biochem. J 391, 277–284.
proven difficult in humans. Therefore, the use of CR mimetics, Bebenek, K., Kunkel, T.A., 2004. Functions of DNA polymerases. Adv. Protein Chem.
which target the same metabolic and stress pathways as CR and/or 69, 137–165.
Belin, A.C., Bjork, B.F., Westerlund, M., Galter, D., Sydow, O., Lind, C., Pernold, K.,
exercise, is now being examined. Examples include: SIRT-1 acti- Rosvall, L., Hakansson, A., Winblad, B., Nissbrandt, H., Graff, C., Olson, L., 2007.
vators, such as resveratrol and green tea polyphenols; glycolytic Association study of two genetic variants in mitochondrial transcription factor
inhibitors, such as 2-deoxy-d-glucose; insulin/AMPK activators, A (TFAM) in Alzheimer’s and Parkinson’s disease. Neurosci. Lett. 420, 257–262.
Bender, A., Krishnan, K.J., Morris, C.M., Taylor, G.A., Reeve, A.K., Perry, R.H., Jaros, E.,
such as metformin; and mTOR inhibitors, such as rapamycin, which Hersheson, J.S., Betts, J., Klopstock, T., Taylor, R.W., Turnbull, D.M., 2006. High
have been reviewed elsewhere (Ingram and Roth, 2015; Testa et al., levels of mitochondrial DNA deletions in substantia nigra neurons in aging and
2014). Parkinson disease. Nat. Genet. 38, 515–517.
Bensch, K.G., Degraaf, W., Hansen, P.A., Zassenhaus, H.P., Corbett, J.A., 2007. A
transgenic model to study the pathogenesis of somatic mtDNA mutation
accumulation in beta-cells. Diabetes Obes. Metab. 9 (Suppl. 2), 74–80.
10. Conclusion Bensch, K.G., Mott, J.L., Chang, S.W., Hansen, P.A., Moxley, M.A., Chambers, K.T., de
Graaf, W., Zassenhaus, H.P., Corbett, J.A., 2009. Selective mtDNA mutation
accumulation results in beta-cell apoptosis and diabetes development. Am. J.
Aging is a multifactorial, progressive, biological process with Physiol. 296, E672–680.
mitochondria playing a central role. The MFRTA, once at the fore- Bereiter-Hahn, J., 2014. Mitochondrial dynamics in aging and disease. Prog. Mol.
front of aging theory, has lost favor since its fundamental notion Biol. Transl. Sci. 127, 93–131.
Bertram, L., McQueen, M.B., Mullin, K., Blacker, D., Tanzi, R.E., 2007. Systematic
of a “vicious” cycle of ROS generating ETC disruption, and hence meta-analyses of Alzheimer disease genetic association studies: the AlzGene
more ROS, is based on anecdotal evidence. This does not imply database. Nat. Genet. 39, 17–23.
100 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

Bori, Z., Zhao, Z., Koltai, E., Fatouros, I.G., Jamurtas, A.Z., Douroudos, I.I., Terzis, G., Corral-Debrinski, M., Horton, T., Lott, M.T., Shoffner, J.M., McKee, A.C., Beal, M.F.,
Chatzinikolaou, A., Sovatzidis, A., Draganidis, D., Boldogh, I., Radak, Z., 2012. Graham, B.H., Wallace, D.C., 1994. Marked changes in mitochondrial DNA
The effects of aging, physical training, and a single bout of exercise on deletion levels in Alzheimer brains. Genomics 23, 471–476.
mitochondrial protein expression in human skeletal muscle. Exp. Gerontol. 47, Cortopassi, G.A., Arnheim, N., 1990. Detection of a specific mitochondrial DNA
417–424. deletion in tissues of older humans. Nucleic Acids Res. 18, 6927–6933.
Brandon, B.R., Diederich, N.J., Soni, M., Witte, K., Weinhold, M., Krause, M., Jackson, Cottrell, D.A., Blakely, E.L., Johnson, M.A., Ince, P.G., Borthwick, G.M., Turnbull, D.M.,
S., 2013. Autosomal dominant mutations in POLG and C10orf2: association 2001. Cytochrome c oxidase deficient cells accumulate in the hippocampus
with late onset chronic progressive external ophthalmoplegia and and choroid plexus with age. Neurobiol. Aging 22, 265–272.
Parkinsonism in two patients. J. Neurol. 260, 1931–1933. Coxhead, J., Kurzawa-Akanbi, M., Hussain, R., Pyle, A., Chinnery, P., Hudson, G.,
Brown, W.M., George, M.J., Wilson, A.C., 1979. Rapid evolution of animal 2016. Somatic mtDNA variation is an important component of Parkinson’s
mitochondrial DNA. Proc. Natl. Acad. Sci. U. S. A. 76, 1967–1971. disease. Neurobiol. Aging 38, e1-6.
Brown, W.M., Prager, E.M., Wang, A., Wilson, A.C., 1982. Mitochondrial DNA Craig, K., Young, M.J., Blakely, E.L., Longley, M.J., Turnbull, D.M., Copeland, W.C.,
sequences of primates: tempo and mode of evolution. J. Mol. Evol. 18, 225–239. Taylor, R.W., 2012. A p.R369 G POLG2 mutation associated with adPEO and
Bua, E., Johnson, J., Herbst, A., Delong, B., McKenzie, D., Salamat, S., Aiken, J.M., multiple mtDNA deletions causes decreased affinity between polymerase
2006. Mitochondrial DNA-deletion mutations accumulate intracellularly to gamma subunits. Mitochondrion 12, 313–319.
detrimental levels in aged human skeletal muscle fibers. Am. J. Hum. Genet. 79, Davidzon, G., Greene, P., Mancuso, M., Klos, K.J., Ahlskog, J.E., Hirano, M., Dimauro,
469–480. S., 2006. Early-onset familial parkinsonism due to POLG mutations. Ann.
Cabelof, D.C., Yanamadala, S., Raffoul, J.J., Guo, Z., Soofi, A., Heydari, A.R., 2003. Neurol. 59, 859–862.
Caloric restriction promotes genomic stability by induction of base excision de Souza-Pinto, N.C., Eide, L., Hogue, B.A., Thybo, T., Stevnsner, T., Seeberg, E.,
repair and reversal of its age-related decline. DNA Repair (Amst) 2, Klungland, A., Bohr, V.A., 2001. Repair of 8-oxodeoxyguanosine lesions in
295–307. mitochondrial DNA depends on the oxoguanine DNA glycosylase (OGG1) gene
Campbell, C.T., Kolesar, J.E., Kaufman, B.A., 2012. Mitochondrial transcription factor and 8-oxoguanine accumulates in the mitochondrial DNA of OGG1-defective
A regulates mitochondrial transcription initiation, DNA packaging, and mice. Cancer Res. 61, 5378–5381.
genome copy number. Biochim. Biophys. Acta 1819, 921–929. Delgado-Alvarado, M., de la Riva, P., Jimenez-Urbieta, H., Gago, B., Gabilondo, A.,
Capel, F., Rimbert, V., Lioger, D., Diot, A., Rousset, P., Mirand, P.P., Boirie, Y., Morio, Bornstein, B., Rodriguez-Oroz, M.C., 2015. Parkinsonism, cognitive deficit and
B., Mosoni, L., 2005. Due to reverse electron transfer, mitochondrial H2O2 behavioural disturbance caused by a novel mutation in the polymerase gamma
release increases with age in human vastus lateralis muscle although oxidative gene. J. Neurol. Sci. 350, 93–97.
capacity is preserved. Mech. Ageing Dev. 126, 505–511. Diaz, F., Bayona-Bafaluy, M.P., Rana, M., Mora, M., Hao, H., Moraes, C.T., 2002.
Carrodeguas, J.A., Theis, K., Bogenhagen, D.F., Kisker, C., 2001. Crystal structure and Human mitochondrial DNA with large deletions repopulates organelles faster
deletion analysis show that the accessory subunit of mammalian DNA than full-length genomes under relaxed copy number control. Nucleic Acids
polymerase gamma, Pol gamma B, functions as a homodimer. Mol. Cell 7, Res. 30, 4626–4633.
43–54. Dillon, L.M., Williams, S.L., Hida, A., Peacock, J.D., Prolla, T.A., Lincoln, J., Moraes,
Cartoni, R., Leger, B., Hock, M.B., Praz, M., Crettenand, A., Pich, S., Ziltener, J.L., Luthi, C.T., 2012. Increased mitochondrial biogenesis in muscle improves aging
F., Deriaz, O., Zorzano, A., Gobelet, C., Kralli, A., Russell, A.P., 2005. Mitofusins phenotypes in the mtDNA mutator mouse. Hum. Mol. Genet. 21, 2288–2297.
1/2 and ERRalpha expression are increased in human skeletal muscle after Drake, J.C., Wilson, R.J., Yan, Z., 2016. Molecular mechanisms for mitochondrial
physical exercise. J. Physiol. 567, 349–358. adaptation to exercise training in skeletal muscle. FASEB J. 30, 13–22.
Cassano, P., Lezza, A.M., Leeuwenburgh, C., Cantatore, P., Gadaleta, M.N., 2004. Duvoisin, R.C., 1992. Overview of Parkinson’s disease. Ann. N. Y. Acad. Sci. 648,
Measurement of the 4,834-bp mitochondrial DNA deletion level in aging rat 187–193.
liver and brain subjected or not to caloric restriction diet. Ann. N. Y. Acad. Sci. Edgar, D., Trifunovic, A., 2009. The mtDNA mutator mouse: dissecting
1019, 269–273. mitochondrial involvement in aging. Aging 1, 1028–1032.
Cerritelli, S.M., Crouch, R.J., 2009. Ribonuclease H: the enzymes in eukaryotes. FEBS Edgar, D., Shabalina, I., Camara, Y., Wredenberg, A., Calvaruso, M.A., Nijtmans, L.,
J. 276, 1494–1505. Nedergaard, J., Cannon, B., Larsson, N.G., Trifunovic, A., 2009. Random point
Cerritelli, S.M., Frolova, E.G., Feng, C., Grinberg, A., Love, P.E., Crouch, R.J., 2003. mutations with major effects on protein-coding genes are the driving force
Failure to produce mitochondrial DNA results in embryonic lethality in behind premature aging in mtDNA mutator mice. Cell Metab. 10, 131–138.
Rnaseh1 null mice. Mol. Cell 11, 807–815. Ekstrand, M.I., Falkenberg, M., Rantanen, A., Park, C.B., Gaspari, M., Hultenby, K.,
Chakravarti, B., Chakravarti, D.N., 2007. Oxidative modification of proteins: Rustin, P., Gustafsson, C.M., Larsson, N.G., 2004. Mitochondrial transcription
age-related changes. Gerontology 53, 128–139. factor A regulates mtDNA copy number in mammals. Hum. Mol. Genet. 13,
Chan, D.C., 2006. Mitochondria: dynamic organelles in disease, aging, and 935–944.
development. Cell 125, 1241–1252. Ekstrand, M.I., Terzioglu, M., Galter, D., Zhu, S., Hofstetter, C., Lindqvist, E., Thams,
Chan, D.C., 2012. Fusion and fission: interlinked processes critical for S., Bergstrand, A., Hansson, F.S., Trifunovic, A., Hoffer, B., Cullheim, S.,
mitochondrial health. Annu. Rev. Genet. 46, 265–287. Mohammed, A.H., Olson, L., Larsson, N.G., 2007. Progressive parkinsonism in
Chauhan, A., Vera, J., Wolkenhauer, O., 2014. The systems biology of mitochondrial mice with respiratory-chain-deficient dopamine neurons. Proc. Natl. Acad. Sci.
fission and fusion and implications for disease and aging. Biogerontology 15, U. S. A. 104, 1325–1330.
1–12. Elchuri, S., Oberley, T.D., Qi, W., Eisenstein, R.S., Jackson Roberts, L., Van Remmen,
Chen, H., Chan, D.C., 2005. Emerging functions of mammalian mitochondrial fusion H., Epstein, C.J., Huang, T.T., 2005. CuZnSOD deficiency leads to persistent and
and fission. Hum Mol Genet 14, R283–R289, Spec No. 2. widespread oxidative damage and hepatocarcinogenesis later in life. Oncogene
Chen, H., Chan, D.C., 2009. Mitochondrial dynamics–fusion, fission, movement, and 24, 367–380.
mitophagy–in neurodegenerative diseases. Hum. Mol. Genet. 18, R169–R176. Elson, J.L., Samuels, D.C., Turnbull, D.M., Chinnery, P.F., 2001. Random intracellular
Chen, H., Detmer, S.A., Ewald, A.J., Griffin, E.E., Fraser, S.E., Chan, D.C., 2003. drift explains the clonal expansion of mitochondrial DNA mutations with age.
Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are Am. J. Hum. Genet. 68, 802–806.
essential for embryonic development. J. Cell Biol. 160, 189–200. Farge, G., Pham, X.H., Holmlund, T., Khorostov, I., Falkenberg, M., 2007. The
Chen, H., Chomyn, A., Chan, D.C., 2005. Disruption of fusion results in accessory subunit B of DNA polymerase {gamma} is required for
mitochondrial heterogeneity and dysfunction. J. Biol. Chem. 280, mitochondrial replisome function. Nucleic Acids Res. 35, 902–911.
26185–26192. Fayet, G., Jansson, M., Sternberg, D., Moslemi, A.R., Blondy, P., Lombes, A., Fardeau,
Chen, H., Vermulst, M., Wang, Y.E., Chomyn, A., Prolla, T.A., McCaffery, J.M., Chan, M., Oldfors, A., 2002. Ageing muscle: clonal expansions of mitochondrial DNA
D.C., 2010. Mitochondrial fusion is required for mtDNA stability in skeletal point mutations and deletions cause focal impairment of mitochondrial
muscle and tolerance of mtDNA mutations. Cell 141, 280–289. function. Neuromuscul. Disord. 12, 484–493.
Chistiakov, D.A., Sobenin, I.A., Revin, V.V., Orekhov, A.N., Bobryshev, Y.V., 2014. Filippov, V., Filippov, M., Gill, S.S., 2001. Drosophila RNase H1 is essential for
Mitochondrial aging and age-related dysfunction of mitochondria. BioMed Res. development but not for proliferation. Mol. Genet. Genomics 265, 771–777.
Int. 2014, 238463. Fisher, R.P., Clayton, D.A., 1985. A transcription factor required for promoter
Civitarese, A.E., Carling, S., Heilbronn, L.K., Hulver, M.H., Ukropcova, B., Deutsch, recognition by human mitochondrial RNA polymerase. Accurate initiation at
W.A., Smith, S.R., Ravussin, E., Team, C.P., 2007. Calorie restriction increases the heavy- and light-strand promoters dissected and reconstituted in vitro. J.
muscle mitochondrial biogenesis in healthy humans. PLoS Med. 4, e76. Biol. Chem. 260, 11330–11338.
Colman, R.J., Beasley, T.M., Kemnitz, J.W., Johnson, S.C., Weindruch, R., Anderson, Fisher, R.P., Topper, J.N., Clayton, D.A., 1987. Promoter selection in human
R.M., 2014. Caloric restriction reduces age-related and all-cause mortality in mitochondria involves binding of a transcription factor to
rhesus monkeys. Nat. Commun. 5, 3557. orientation-independent upstream regulatory elements. Cell 50, 247–258.
Cooke, M.S., Evans, M.D., Dizdaroglu, M., Lunec, J., 2003. Oxidative DNA damage: Fleming, J.E., Miquel, J., Cottrell, S.F., Yengoyan, L.S., Economos, A.C., 1982. Is cell
mechanisms, mutation, and disease. FASEB J. 17, 1195–1214. aging caused by respiration-dependent injury to the mitochondrial genome?
Copeland, W.C., Longley, M.J., 2014. Mitochondrial genome maintenance in health Gerontology 28, 44–53.
and disease. DNA Repair (Amst) 19, 190–198. Garrido, N., Griparic, L., Jokitalo, E., Wartiovaara, J., Van Der Bliek, A.M., Spelbrink,
Corral-Debrinski, M., Horton, T., Lott, M.T., Shoffner, J.M., Beal, M.F., Wallace, D.C., J.N., 2003. Composition and dynamics of human mitochondrial nucleoids. Mol.
1992a. Mitochondrial DNA deletions in human brain: regional variability and Biol. Cell 14, 1583–1596.
increase with advanced age. Nat. Genet. 2, 324–329. Gauthier, B.R., Wiederkehr, A., Baquie, M., Dai, C., Powers, A.C., Kerr-Conte, J.,
Corral-Debrinski, M., Shoffner, J.M., Lott, M.T., Wallace, D.C., 1992b. Association of Pattou, F., MacDonald, R.J., Ferrer, J., Wollheim, C.B., 2009. PDX1 deficiency
mitochondrial DNA damage with aging and coronary atherosclerotic heart causes mitochondrial dysfunction and defective insulin secretion through
disease. Mutat. Res. 275, 169–180. TFAM suppression. Cell Metab. 10, 110–118.
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 101

Gaziev, A.I., Abdullaev, S., Podlutsky, A., 2014. Mitochondrial function and Humble, M.M., Young, M.J., Foley, J.F., Pandiri, A.R., Travlos, G.S., Copeland, W.C.,
mitochondrial DNA maintenance with advancing age. Biogerontology 15, 2013. Polg2 is essential for mammalian embryogenesis and is required for
417–438. mtDNA maintenance. Hum. Mol. Genet. 22, 1017–1025.
Giles, R.E., Blanc, H., Cann, H.M., Wallace, D.C., 1980. Maternal inheritance of Ikeda, M., Ide, T., Fujino, T., Arai, S., Saku, K., Kakino, T., Tyynismaa, H., Yamasaki, T.,
human mitochondrial DNA. Proc. Natl. Acad. Sci. U. S. A. 77, 6715–6719. Yamada, K., Kang, D., Suomalainen, A., Sunagawa, K., 2015. Overexpression of
Goffart, S., Cooper, H.M., Tyynismaa, H., Wanrooij, S., Suomalainen, A., Spelbrink, TFAM or twinkle increases mtDNA copy number and facilitates
J.N., 2009. Twinkle mutations associated with autosomal dominant progressive cardioprotection associated with limited mitochondrial oxidative stress. PLoS
external ophthalmoplegia lead to impaired helicase function and in vivo One 10, e0119687.
mtDNA replication stalling. Hum. Mol. Genet. 18, 328–340. Ikeuchi, M., Matsusaka, H., Kang, D., Matsushima, S., Ide, T., Kubota, T., Fujiwara, T.,
Graziewicz, M.A., Longley, M.J., Bienstock, R.J., Zeviani, M., Copeland, W.C., 2004. Hamasaki, N., Takeshita, A., Sunagawa, K., Tsutsui, H., 2005. Overexpression of
Structure-function defects of human mitochondrial DNA polymerase in mitochondrial transcription factor a ameliorates mitochondrial deficiencies
autosomal dominant progressive external ophthalmoplegia. Nat. Struct. Mol. and cardiac failure after myocardial infarction. Circulation 112, 683–690.
Biol. 11, 770–776. Ingram, D.K., Roth, G.S., 2015. Calorie restriction mimetics: can you have your cake
Graziewicz, M.A., Longley, M.J., Copeland, W.C., 2006. DNA polymerase gamma in and eat it, too? Ageing Res. Rev. 20, 46–62.
mitochondrial DNA replication and repair. Chem. Rev. 106, 383–405. Irving, B.A., Lanza, I.R., Henderson, G.C., Rao, R.R., Spiegelman, B.M., Nair, K.S., 2015.
Greaves, L.C., Nooteboom, M., Elson, J.L., Tuppen, H.A., Taylor, G.A., Commane, D.M., Combined training enhances skeletal muscle mitochondrial oxidative capacity
Arasaradnam, R.P., Khrapko, K., Taylor, R.W., Kirkwood, T.B., Mathers, J.C., independent of age. J. Clin. Endocrinol. Metab. 100, 1654–1663.
Turnbull, D.M., 2014. Clonal expansion of early to mid-life mitochondrial DNA Itsara, L.S., Kennedy, S.R., Fox, E.J., Yu, S., Hewitt, J.J., Sanchez-Contreras, M.,
point mutations drives mitochondrial dysfunction during human ageing. PLoS Cardozo-Pelaez, F., Pallanck, L.J., 2014. Oxidative stress is not a major
Genet. 10, e1004620. contributor to somatic mitochondrial DNA mutations. PLoS Genet. 10,
Gredilla, R., Lopez-Torres, M., Barja, G., 2002. Effect of time of restriction on the e1003974.
decrease in mitochondrial H2O2 production and oxidative DNA damage in the Iyer, S., Thomas, R.R., Portell, F.R., Dunham, L.D., Quigley, C.K., Bennett Jr., J.P., 2009.
heart of food-restricted rats. Microsc. Res. Tech. 59, 273–277. Recombinant mitochondrial transcription factor A with N-terminal
Green, D.R., Galluzzi, L., Kroemer, G., 2011. Mitochondria and the mitochondrial transduction domain increases respiration and mitochondrial
autophagy-inflammation-cell death axis in organismal aging. Science 333, gene expression. Mitochondrion 9, 196–203.
1109–1112. Jankovic, J., 2008. Parkinson’s disease: clinical features and diagnosis. J. Neurol.
Grumati, P., Coletto, L., Schiavinato, A., Castagnaro, S., Bertaggia, E., Sandri, M., Neurosurg. Psychiatry 79, 368–376.
Bonaldo, P., 2011. Physical exercise stimulates autophagy in normal skeletal Jemt, E., Persson, O., Shi, Y., Mehmedovic, M., Uhler, J.P., Davila Lopez, M., Freyer, C.,
muscles but is detrimental for collagen VI-deficient muscles. Autophagy 7, Gustafsson, C.M., Samuelsson, T., Falkenberg, M., 2015. Regulation of DNA
1415–1423. replication at the end of the mitochondrial D-loop involves the helicase
Gui, Y.X., Xu, Z.P., Lv, W., Zhao, J.J., Hu, X.Y., 2015. Evidence for polymerase gamma, TWINKLE and a conserved sequence element. Nucleic Acids Res. 43,
POLG1 variation in reduced mitochondrial DNA copy number in Parkinson’s 9262–9275.
disease. Parkinsonism Relat. Disord. 21, 282–286. Jendrach, M., Pohl, S., Voth, M., Kowald, A., Hammerstein, P., Bereiter-Hahn, J.,
Halsne, R., Esbensen, Y., Wang, W., Scheffler, K., Suganthan, R., Bjoras, M., Eide, L., 2005. Morpho-dynamic changes of mitochondria during ageing of human
2012. Lack of the DNA glycosylases MYH and OGG1 in the cancer prone double endothelial cells. Mech. Ageing Dev. 126, 813–821.
mutant mouse does not increase mitochondrial DNA mutagenesis. DNA Repair Jeppesen, T.D., Schwartz, M., Olsen, D.B., Wibrand, F., Krag, T., Duno, M., Hauerslev,
(Amst) 11, 278–285. S., Vissing, J., 2006. Aerobic training is safe and improves exercise capacity in
Harman, D., 1956. Aging: a theory based on free radical and radiation chemistry. J. patients with mitochondrial myopathy. Brain 129, 3402–3412.
Gerontol. 11, 298–300. Jeppesen, T.D., Duno, M., Schwartz, M., Krag, T., Rafiq, J., Wibrand, F., Vissing, J.,
Harman, D., 1972. The biologic clock: the mitochondria? J. Am. Geriatr. Soc. 20, 2009. Short- and long-term effects of endurance training in patients with
145–147. mitochondrial myopathy. Eur. J. Neurol. 16, 1336–1339.
Hayakawa, M., Torii, K., Sugiyama, S., Tanaka, M., Ozawa, T., 1991. Age-associated Joseph, A.M., Adhihetty, P.J., Wawrzyniak, N.R., Wohlgemuth, S.E., Picca, A., Kujoth,
accumulation of 8-hydroxydeoxyguanosine in mitochondrial DNA of human G.C., Prolla, T.A., Leeuwenburgh, C., 2013. Dysregulation of mitochondrial
diaphragm. Biochem. Biophys. Res. Commun. 179, 1023–1029. quality control processes contribute to sarcopenia in a mouse model of
Hayakawa, M., Hattori, K., Sugiyama, S., Ozawa, T., 1992. Age-associated oxygen premature aging. PLoS One 8, e69327.
damage and mutations in mitochondrial DNA in human hearts. Biochem. Kaeberlein, M., Burtner, C.R., Kennedy, B.K., 2007. Recent developments in yeast
Biophys. Res. Commun. 189, 979–985. aging. PLoS Genet. 3, e84.
Hayashi, J., Ohta, S., Kikuchi, A., Takemitsu, M., Goto, Y., Nonaka, I., 1991. Kaguni, L.S., 2004. DNA polymerase gamma, the mitochondrial replicase. Annu.
Introduction of disease-related mitochondrial DNA deletions into HeLa cells Rev. Biochem 73, 293–320.
lacking mitochondrial DNA results in mitochondrial dysfunction. Proc. Natl. Kang, C., Chung, E., Diffee, G., Ji, L.L., 2013. Exercise training attenuates
Acad. Sci. U. S. A. 88, 10614–10618. aging-associated mitochondrial dysfunction in rat skeletal muscle: role of
Hayashi, Y., Yoshida, M., Yamato, M., Ide, T., Wu, Z., Ochi-Shindou, M., Kanki, T., PGC-1alpha. Exp. Gerontol. 48, 1343–1350.
Kang, D., Sunagawa, K., Tsutsui, H., Nakanishi, H., 2008. Reverse of Kanki, T., Nakayama, H., Sasaki, N., Takio, K., Alam, T.I., Hamasaki, N., Kang, D.,
age-dependent memory impairment and mitochondrial DNA damage in 2004. Mitochondrial nucleoid and transcription factor A. Ann. N. Y. Acad. Sci.
microglia by an overexpression of human mitochondrial transcription factor a 1011, 61–68.
in mice. J. Neurosci. 28, 8624–8634. Kennedy, S.R., Salk, J.J., Schmitt, M.W., Loeb, L.A., 2013. Ultra-sensitive sequencing
He, Y., Wu, J., Dressman, D.C., Iacobuzio-Donahue, C., Markowitz, S.D., Velculescu, reveals an age-related increase in somatic mitochondrial mutations that are
V.E., Diaz Jr., L.A., Kinzler, K.W., Vogelstein, B., Papadopoulos, N., 2010. inconsistent with oxidative damage. PLoS Genet. 9, e1003794.
Heteroplasmic mitochondrial DNA mutations in normal and tumour cells. Keogh, M.J., Chinnery, P.F., 2015. Mitochondrial DNA mutations in
Nature 464, 610–614. neurodegeneration. Biochim. Biophys. Acta 1847, 1401–1411.
Hekimi, S., Lapointe, J., Wen, Y., 2011. Taking a good look at free radicals in the Khan, S.M., Bennett Jr., J.P., 2004. Development of mitochondrial gene replacement
aging process. Trends Cell Biol. 21, 569–576. therapy. J. Bioenerg. Biomembr. 36, 387–393.
Heydari, A.R., Unnikrishnan, A., Lucente, L.V., Richardson, A., 2007. Caloric Khrapko, K., Ebralidse, K., Kraytsberg, Y., 2004. Where and when do somatic
restriction and genomic stability. Nucleic Acids Res. 35, 7485–7496. mtDNA mutations occur? Ann. N. Y. Acad. Sci. 1019, 240–244.
Hirano, M., Marti, R., Ferreiro-Barros, C., Vila, M.R., Tadesse, S., Nishigaki, Y., Khrapko, K., 2011. The timing of mitochondrial DNA mutations in aging. Nat.
Nishino, I., Vu, T.H., 2001. Defects of intergenomic communication: autosomal Genet. 43, 726–727.
disorders that cause multiple deletions and depletion of mitochondrial DNA. Kim, Y.A., Kim, Y.S., Oh, S.L., Kim, H.J., Song, W., 2013. Autophagic response to
Semin. Cell Dev. Biol. 12, 417–427. exercise training in skeletal muscle with age. J. Physiol. Biochem. 69, 697–705.
Holmes, J.B., Akman, G., Wood, S.R., Sakhuja, K., Cerritelli, S.M., Moss, C., Kirkwood, T.B., 2005. Understanding the odd science of aging. Cell 120, 437–447.
Bowmaker, M.R., Jacobs, H.T., Crouch, R.J., Holt, I.J., 2015. Primer retention Konopka, A.R., Suer, M.K., Wolff, C.A., Harber, M.P., 2014. Markers of human
owing to the absence of RNase H1 is catastrophic for mitochondrial DNA skeletal muscle mitochondrial biogenesis and quality control: effects of age
replication. Proc. Natl. Acad. Sci. U. S. A. 112, 9334–9339. and aerobic exercise training. J. Gerontol. A Biol. Sci. Med. Sci. 69, 371–378.
Holmlund, T., Farge, G., Pande, V., Korhonen, J., Nilsson, L., Falkenberg, M., 2009. Korhonen, J.A., Gaspari, M., Falkenberg, M., 2003. TWINKLE Has 5’ −> 3’ DNA
Structure-function defects of the twinkle amino-terminal region in progressive helicase activity and is specifically stimulated by mitochondrial
external ophthalmoplegia. Biochim. Biophys. Acta 1792, 132–139. single-stranded DNA-binding protein. J. Biol. Chem. 278, 48627–48632.
Hubbard, V.M., Valdor, R., Macian, F., Cuervo, A.M., 2012. Selective autophagy in Korhonen, J.A., Pande, V., Holmlund, T., Farge, G., Pham, X.H., Nilsson, L.,
the maintenance of cellular homeostasis in aging organisms. Biogerontology Falkenberg, M., 2008. Structure-function defects of the TWINKLE linker region
13, 21–35. in progressive external ophthalmoplegia. J. Mol. Biol. 377, 691–705.
Hudson, G., Schaefer, A.M., Taylor, R.W., Tiangyou, W., Gibson, A., Venables, G., Kornblum, C., Nicholls, T.J., Haack, T.B., Scholer, S., Peeva, V., Danhauser, K.,
Griffiths, P., Burn, D.J., Turnbull, D.M., Chinnery, P.F., 2007. Mutation of the Hallmann, K., Zsurka, G., Rorbach, J., Iuso, A., Wieland, T., Sciacco, M., Ronchi,
linker region of the polymerase gamma-1 (POLG1) gene associated with D., Comi, G.P., Moggio, M., Quinzii, C.M., DiMauro, S., Calvo, S.E., Mootha, V.K.,
progressive external ophthalmoplegia and Parkinsonism. Arch. Neurol. 64, Klopstock, T., Strom, T.M., Meitinger, T., Minczuk, M., Kunz, W.S., Prokisch, H.,
553–557. 2013. Loss-of-function mutations in MGME1 impair mtDNA replication and
Hudson, G., Gomez-Duran, A., Wilson, I.J., Chinnery, P.F., 2014. Recent cause multisystemic mitochondrial disease. Nat. Genet. 45, 214–219.
mitochondrial DNA mutations increase the risk of developing common Kowald, A., Kirkwood, T.B., 2011. Evolution of the mitochondrial fusion-fission
late-onset human diseases. PLoS Genet. 10, e1004369. cycle and its role in aging. Proc. Natl. Acad. Sci. U. S. A. 108, 10237–10242.
102 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

Kraytsberg, Y., Kudryavtseva, E., McKee, A.C., Geula, C., Kowall, N.W., Khrapko, K., mitochondrial biogenesis and bioenergetic efficiency. Proc. Natl. Acad. Sci. U. S.
2006. Mitochondrial DNA deletions are abundant and cause functional A. 103, 1768–1773.
impairment in aged human substantia nigra neurons. Nat. Genet. 38, 518–520. Lopez-Torres, M., Gredilla, R., Sanz, A., Barja, G., 2002. Influence of aging and
Kraytsberg, Y., Simon, D.K., Turnbull, D.M., Khrapko, K., 2009. Do mtDNA deletions long-term caloric restriction on oxygen radical generation and oxidative DNA
drive premature aging in mtDNA mutator mice? Aging Cell 8, 502–506. damage in rat liver mitochondria. Free Radic. Biol. Med. 32, 882–889.
Kujoth, G.C., Hiona, A., Pugh, T.D., Someya, S., Panzer, K., Wohlgemuth, S.E., Hofer, Luo, L., Lu, A.M., Wang, Y., Hong, A., Chen, Y., Hu, J., Li, X., Qin, Z.H., 2013. Chronic
T., Seo, A.Y., Sullivan, R., Jobling, W.A., Morrow, J.D., Van Remmen, H., Sedivy, resistance training activates autophagy and reduces apoptosis of muscle cells
J.M., Yamasoba, T., Tanokura, M., Weindruch, R., Leeuwenburgh, C., Prolla, T.A., by modulating IGF-1 and its receptors, Akt/mTOR and Akt/FOXO3a signaling in
2005. Mitochondrial DNA mutations, oxidative stress, and apoptosis in aged rats. Exp. Gerontol. 48, 427–436.
mammalian aging. Science 309, 481–484. Luo, Y., Hoffer, A., Hoffer, B., Qi, X., 2015. Mitochondria: a therapeutic target for
Kukat, A., Trifunovic, A., 2009. Somatic mtDNA mutations and aging–facts and parkinson’s disease? Int. J. Mol. Sci. 16, 20704–20730.
fancies. Exp. Gerontol. 44, 101–105. Luoma, P., Melberg, A., Rinne, J.O., Kaukonen, J.A., Nupponen, N.N., Chalmers, R.M.,
Kwong, L.K., Sohal, R.S., 2000. Age-related changes in activities of mitochondrial Oldfors, P.A., Rautakorpi, I., Peltonen, P.L., Majamaa, P.K., Somer, H.,
electron transport complexes in various tissues of the mouse. Arch. Biochem. Suomalainen, A., 2004. Parkinsonism, premature menopause, and
Biophys. 373, 16–22. mitochondrial DNA polymerase gamma mutations: clinical and molecular
Lamantea, E., Tiranti, V., Bordoni, A., Toscano, A., Bono, F., Servidei, S., genetic study. Lancet 364, 875–882.
Papadimitriou, A., Spelbrink, H., Silvestri, L., Casari, G., Comi, G., Zeviani, M., Luoma, P.T., Eerola, J., Ahola, S., Hakonen, A.H., Hellstrom, O., Kivisto, K.T., Tienari,
2002. Mutations of mitochondrial DNA polymerase gamma are a frequent P.J., Suomalainen, A., 2007. Mitochondrial DNA polymerase gamma variants in
cause of autosomal dominant or recessive Progressive External idiopathic sporadic Parkinson disease. Neurology 69, 1152–1159.
Ophthalmoplegia. Ann. Neurol. 52, 211–219. Malena, A., Loro, E., Di Re, M., Holt, I.J., Vergani, L., 2009. Inhibition of
Larsson, N.G., Wang, J., Wilhelmsson, H., Oldfors, A., Rustin, P., Lewandoski, M., mitochondrial fission favours mutant over wild-type mitochondrial DNA.
Barsh, G.S., Clayton, D.A., 1998. Mitochondrial transcription factor A is Hum. Mol. Genet. 18, 3407–3416.
necessary for mtDNA maintenance and embryogenesis in mice. Nat. Genet. 18, Mancuso, M., Filosto, M., Oh, S.J., DiMauro, S., 2004. A novel polymerase gamma
231–236. mutation in a family with ophthalmoplegia, neuropathy, and Parkinsonism.
Lass, A., Sohal, B.H., Weindruch, R., Forster, M.J., Sohal, R.S., 1998. Caloric restriction Arch. Neurol. 61, 1777–1779.
prevents age-associated accrual of oxidative damage to mouse skeletal muscle Mandavilli, B.S., Santos, J.H., Van Houten, B., 2002. Mitochondrial DNA repair and
mitochondria. Free Radic. Biol. Med. 25, 1089–1097. aging. Mutat. Res. 509, 127–151.
Lebovitz, R.M., Zhang, H., Vogel, H., Cartwright Jr., J., Dionne, L., Lu, N., Huang, S., Mansouri, A., Muller, F.L., Liu, Y., Ng, R., Faulkner, J., Hamilton, M., Richardson, A.,
Matzuk, M.M., 1996. Neurodegeneration, myocardial injury, and perinatal Huang, T.T., Epstein, C.J., Van Remmen, H., 2006. Alterations in mitochondrial
death in mitochondrial superoxide dismutase-deficient mice. Proc. Natl. Acad. function, hydrogen peroxide release and oxidative damage in mouse hind-limb
Sci. U. S. A. 93, 9782–9787. skeletal muscle during aging. Mech. Ageing Dev. 127, 298–306.
Lee, Y.J., Jeong, S.Y., Karbowski, M., Smith, C.L., Youle, R.J., 2004. Roles of the Martin-Montalvo, A., de Cabo, R., 2013. Mitochondrial metabolic reprogramming
mammalian mitochondrial fission and fusion mediators Fis1, Drp1, and Opa1 induced by calorie restriction. Antioxid. Redox Signal. 19, 310–320.
in apoptosis. Mol. Biol. Cell 15, 5001–5011. Masoro, E.J., 2009. Caloric restriction-induced life extension of rats and mice: a
Lee, G.D., Wilson, M.A., Zhu, M., Wolkow, C.A., de Cabo, R., Ingram, D.K., Zou, S., critique of proposed mechanisms. Biochim. Biophys. Acta 1790, 1040–1048.
2006. Dietary deprivation extends lifespan in Caenorhabditis elegans. Aging Matsushima, Y., Kaguni, L.S., 2007. Differential phenotypes of active site and
Cell 5, 515–524. human adPEO mutations in drosophila mitochondrial DNA helicase expressed
Lee, H.Y., Choi, C.S., Birkenfeld, A.L., Alves, T.C., Jornayvaz, F.R., Jurczak, M.J., Zhang, in schneider cells. J. Biol. Chem. 282, 9436–9444.
D., Woo, D.K., Shadel, G.S., Ladiges, W., Rabinovitch, P.S., Santos, J.H., Petersen, Matsushima, Y., Farr, C.L., Fan, L., Kaguni, L.S., 2008. Physiological and biochemical
K.F., Samuel, V.T., Shulman, G.I., 2010. Targeted expression of catalase to defects in carboxyl-terminal mutants of mitochondrial DNA helicase. J. Biol.
mitochondria prevents age-associated reductions in mitochondrial function Chem. 283, 23964–23971.
and insulin resistance. Cell Metab. 12, 668–674. Michikawa, Y., Mazzucchelli, F., Bresolin, N., Scarlato, G., Attardi, G., 1999.
Lee, S.K., Zhao, M.H., Zheng, Z., Kwon, J.W., Liang, S., Kim, S.H., Kim, N.H., Cui, X.S., Aging-dependent large accumulation of point mutations in the human mtDNA
2015. Polymerase subunit gamma 2 affects porcine oocyte maturation and control region for replication. Science 286, 774–779.
subsequent embryonic development. Theriogenology 83, 121–130. Milenkovic, D., Matic, S., Kuhl, I., Ruzzenente, B., Freyer, C., Jemt, E., Park, C.B.,
Lewis, W., Day, B.J., Kohler, J.J., Hosseini, S.H., Chan, S.S.L., Green, E., Haase, C.P., Falkenberg, M., Larsson, N.G., 2013. TWINKLE is an essential mitochondrial
Keebaugh, E., Long, R., Ludaway, T., Russ, R., Steltzer, J., Tioleco, N., Santoianni, helicase required for synthesis of nascent D-loop strands and complete mtDNA
R., Copeland, W.C., 2007. MtDNA depletion, oxidative stress, cardiomyopathy, replication. Hum. Mol. Genet. 22, 1983–1993.
and death from transgenic cardiac targeted human mutant polymerase Miller, F.J., Rosenfeldt, F.L., Zhang, C., Linnane, A.W., Nagley, P., 2003. Precise
gamma. Lab. Invest. 87, 326–335. determination of mitochondrial DNA copy number in human skeletal and
Li, Y., Huang, T.T., Carlson, E.J., Melov, S., Ursell, P.C., Olson, J.L., Noble, L.J., cardiac muscle by a PCR-based assay: lack of change of copy number with age.
Yoshimura, M.P., Berger, C., Chan, P.H., Wallace, D.C., Epstein, C.J., 1995. Dilated Nucleic Acids Res. 31, e61.
cardiomyopathy and neonatal lethality in mutant mice lacking manganese Moslemi, A.R., Melberg, A., Holme, E., Oldfors, A., 1999. Autosomal dominant
superoxide dismutase. Nat. Genet. 11, 376–381. progressive external ophthalmoplegia: distribution of multiple mitochondrial
Li, H., Wang, J., Wilhelmsson, H., Hansson, A., Thoren, P., Duffy, J., Rustin, P., DNA deletions. Neurology 53, 79–84.
Larsson, N.G., 2000. Genetic modification of survival in tissue-specific Muller-Hocker, J., Seibel, P., Schneiderbanger, K., Kadenbach, B., 1993. Different
knockout mice with mitochondrial cardiomyopathy. Proc. Natl. Acad. Sci. U. S. in situ hybridization patterns of mitochondrial DNA in cytochrome c
A. 97, 3467–3472. oxidase-deficient extraocular muscle fibres in the elderly. Virchows Arch. A
Lim, S.E., Longley, M.J., Copeland, W.C., 1999. The mitochondrial p55 accessory 422, 7–15.
subunit of human DNA polymerase gamma enhances DNA binding, promotes Muller-Hocker, J., 1989. Cytochrome-c-oxidase deficient cardiomyocytes in the
processive DNA synthesis, and confers N-ethylmaleimide resistance. J. Biol. human heart–an age-related phenomenon: a histochemical ultracytochemical
Chem. 274, 38197–38203. study. Am. J. Pathol. 134, 1167–1173.
Lin, M.T., Simon, D.K., Ahn, C.H., Kim, L.M., Beal, M.F., 2002. High aggregate burden Murphy, J.L., Blakely, E.L., Schaefer, A.M., He, L., Wyrick, P., Haller, R.G., Taylor, R.W.,
of somatic mtDNA point mutations in aging and Alzheimer’s disease brain. Turnbull, D.M., Taivassalo, T., 2008. Resistance training in patients with single,
Hum. Mol. Genet. 11, 133–145. large-scale deletions of mitochondrial DNA. Brain 131, 2832–2840.
Linnane, A.W., Marzuki, S., Ozawa, T., Tanaka, M., 1989. Mitochondrial DNA Murphy, M.P., 2009. How mitochondria produce reactive oxygen species. Biochem.
mutations as an important contributor to ageing and degenerative diseases. J. 417, 1–13.
Lancet 1, 642–645. Nakada, K., Inoue, K., Ono, T., Isobe, K., Ogura, A., Goto, Y.I., Nonaka, I., Hayashi, J.I.,
Longley, M.J., Prasad, R., Srivastava, D.K., Wilson, S.H., Copeland, W.C., 1998. 2001. Inter-mitochondrial complementation: mitochondria-specific system
Identification of 5’-deoxyribose phosphate lyase activity in human DNA preventing mice from expression of disease phenotypes by mutant mtDNA.
polymerase gamma and its role in mitochondrial base excision repair in vitro. Nat. Med. 7, 934–940.
Proc. Natl. Acad. Sci. U. S. A. 95, 12244–12248. Nakada, K., Sato, A., Hayashi, J., 2009. Mitochondrial functional complementation
Longley, M.J., Nguyen, D., Kunkel, T.A., Copeland, W.C., 2001. The fidelity of human in mitochondrial DNA-based diseases. Int. J. Biochem. Cell Biol. 41, 1907–1913.
DNA polymerase gamma with and without exonucleolytic proofreading and Ngo, H.B., Lovely, G.A., Phillips, R., Chan, D.C., 2014. Distinct structural features of
the p55 accessory subunit. J. Biol. Chem. 276, 38555–38562. TFAM drive mitochondrial DNA packaging versus transcriptional activation.
Longley, M.J., Clark, S., Yu Wai Man, C., Hudson, G., Durham, S.E., Taylor, R.W., Nat. Commun. 5, 3077.
Nightingale, S., Turnbull, D.M., Copeland, W.C., Chinnery, P.F., 2006. Mutant Nicholas, A., Kraytsberg, Y., Guo, X., Khrapko, K., 2009. On the timing and the
POLG2 disrupts DNA polymerase gamma subunits and causes progressive extent of clonal expansion of mtDNA deletions: evidence from single-molecule
external ophthalmoplegia. Am. J. Hum. Genet. 78, 1026–1034. PCR. Exp. Neurol. 218, 316–319.
Longley, M.J., Humble, M.M., Sharief, F.S., Copeland, W.C., 2010. Disease variants of Nicholls, T.J., Zsurka, G., Peeva, V., Scholer, S., Szczesny, R.J., Cysewski, D., Reyes, A.,
the human mitochondrial DNA helicase encoded by C10orf2 differentially alter Kornblum, C., Sciacco, M., Moggio, M., Dziembowski, A., Kunz, W.S., Minczuk,
protein stability, nucleotide hydrolysis and helicase activity. J. Biol. Chem. 285, M., 2014. Linear mtDNA fragments and unusual mtDNA rearrangements
29690–29702. associated with pathological deficiency of MGME1 exonuclease. Hum. Mol.
Lopez-Lluch, G., Hunt, N., Jones, B., Zhu, M., Jamieson, H., Hilmer, S., Cascajo, M.V., Genet. 23, 6147–6162.
Allard, J., Ingram, D.K., Navas, P., de Cabo, R., 2006. Calorie restriction induces Nishiyama, S., Shitara, H., Nakada, K., Ono, T., Sato, A., Suzuki, H., Ogawa, T., Masaki,
H., Hayashi, J., Yonekawa, H., 2010. Over-expression of Tfam improves the
K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104 103

mitochondrial disease phenotypes in a mouse model system. Biochem. Sawada, M., Carlson, J.C., 1987. Changes in superoxide radical and lipid peroxide
Biophys. Res. Commun. 401, 26–31. formation in the brain, heart and liver during the lifetime of the rat. Mech.
Ohtani, N., Haruki, M., Morikawa, M., Crouch, R.J., Itaya, M., Kanaya, S., 1999. Ageing Dev. 41, 125–137.
Identification of the genes encoding Mn2+-dependent RNase HII and Schriner, S.E., Linford, N.J., 2006. Extension of mouse lifespan by overexpression of
Mg2+-dependent RNase HIII from Bacillus subtilis: classification of RNases H catalase. Age 28, 209–218.
into three families. Biochemistry 38, 605–618. Schriner, S.E., Ogburn, C.E., Smith, A.C., Newcomb, T.G., Ladiges, W.C., Dolle, M.E.,
Oliver-Krasinski, J.M., Stoffers, D.A., 2008. On the origin of the beta cell. Genes Dev. Vijg, J., Fukuchi, K., Martin, G.M., 2000. Levels of DNA damage are unaltered in
22, 1998–2021. mice overexpressing human catalase in nuclei. Free Radic. Biol. Med. 29,
Ono, T., Isobe, K., Nakada, K., Hayashi, J.I., 2001. Human cells are protected from 664–673.
mitochondrial dysfunction by complementation of DNA products in fused Schriner, S.E., Linford, N.J., Martin, G.M., Treuting, P., Ogburn, C.E., Emond, M.,
mitochondria. Nat. Genet. 28, 272–275. Coskun, P.E., Ladiges, W., Wolf, N., Van Remmen, H., Wallace, D.C., Rabinovitch,
Orr, W.C., Sohal, R.S., 1994. Extension of life-span by overexpression of superoxide P.S., 2005. Extension of murine life span by overexpression of catalase targeted
dismutase and catalase in Drosophila melanogaster. Science 263, 1128–1130. to mitochondria. Science 308, 1909–1911.
Orr, W.C., Radyuk, S.N., Sohal, R.S., 2013. Involvement of redox state in the aging of Scuderi, C., Borgione, E., Castello, F., Lo Giudice, M., Santa Paola, S., Giambirtone, M.,
Drosophila melanogaster. Antioxid. Redox Signal. 19, 788–803. Di Blasi, F.D., Elia, M., Amato, C., Citta, S., Gagliano, C., Barbarino, G., Vitello,
Pagano, A.F., Py, G., Bernardi, H., Candau, R.B., Sanchez, A.M., 2014. Autophagy and G.A., Musumeci, S.A., 2015. The in cis T251I and P587L POLG1 base changes:
protein turnover signaling in slow-twitch muscle during exercise. Med. Sci. description of a new family and literature review. Neuromuscul. Disord. 25,
Sports Exerc. 46, 1314–1325. 333–339.
Pagnamenta, A.T., Taanman, J.W., Wilson, C.J., Anderson, N.E., Marotta, R., Duncan, Sen, D., Nandakumar, D., Tang, G.Q., Patel, S.S., 2012. Human mitochondrial DNA
A.J., Bitner-Glindzicz, M., Taylor, R.W., Laskowski, A., Thorburn, D.R., Rahman, helicase TWINKLE is both an unwinding and annealing helicase. J. Biol. Chem.
S., 2006. Dominant inheritance of premature ovarian failure associated with 287, 14545–14556.
mutant mitochondrial DNA polymerase gamma. Hum. Reprod. 21, 2467–2473. Seo, A.Y., Joseph, A.M., Dutta, D., Hwang, J.C., Aris, J.P., Leeuwenburgh, C., 2010.
Parisi, M.A., Clayton, D.A., 1991. Similarity of human mitochondrial transcription New insights into the role of mitochondria in aging: mitochondrial dynamics
factor 1 to high mobility group proteins. Science 252, 965–969. and more. J. Cell Sci. 123, 2533–2542.
Pohjoismaki, J.L., Goffart, S., Tyynismaa, H., Willcox, S., Ide, T., Kang, D., Short, K.R., Vittone, J.L., Bigelow, M.L., Proctor, D.N., Rizza, R.A., Coenen-Schimke,
Suomalainen, A., Karhunen, P.J., Griffith, J.D., Holt, I.J., Jacobs, H.T., 2009. J.M., Nair, K.S., 2003. Impact of aerobic exercise training on age-related changes
Human heart mitochondrial DNA is organized in complex catenated networks in insulin sensitivity and muscle oxidative capacity. Diabetes 52, 1888–1896.
containing abundant four-way junctions and replication forks. J. Biol. Chem. Short, K.R., Vittone, J.L., Bigelow, M.L., Proctor, D.N., Nair, K.S., 2004. Age and
284, 21446–21457. aerobic exercise training effects on whole body and muscle protein
Pohjoismaki, J.L., Goffart, S., Taylor, R.W., Turnbull, D.M., Suomalainen, A., Jacobs, metabolism. Am. J. Physiol. 286, E92–101.
H.T., Karhunen, P.J., 2010. Developmental and pathological changes in the Shutt, T.E., Gray, M.W., 2006. Twinkle, the mitochondrial replicative DNA helicase,
human cardiac muscle mitochondrial DNA organization, replication and copy is widespread in the eukaryotic radiation and may also be the mitochondrial
number. PLoS One 5, e10426. DNA primase in most eukaryotes. J. Mol. Evol. 62, 588–599.
Pohjoismaki, J.L., Williams, S.L., Boettger, T., Goffart, S., Kim, J., Suomalainen, A., Silva, J.P., Kohler, M., Graff, C., Oldfors, A., Magnuson, M.A., Berggren, P.O., Larsson,
Moraes, C.T., Braun, T., 2013. Overexpression of twinkle-helicase protects N.G., 2000. Impaired insulin secretion and beta-cell loss in tissue-specific
cardiomyocytes from genotoxic stress caused by reactive oxygen species. Proc. knockout mice with mitochondrial diabetes. Nat. Genet. 26, 336–340.
Natl. Acad. Sci. U. S. A. 110, 19408–19413. Sohal, R.S., Sohal, B.H., 1991. Hydrogen peroxide release by mitochondria increases
Ponamarev, M.V., Longley, M.J., Nguyen, D., Kunkel, T.A., Copeland, W.C., 2002. during aging. Mech. Ageing Dev. 57, 187–202.
Active site mutation in DNA polymerase gamma associated with progressive Song, Z., Ghochani, M., McCaffery, J.M., Frey, T.G., Chan, D.C., 2009. Mitofusins and
external ophthalmoplegia causes error-prone DNA synthesis. J. Biol. Chem. OPA1 mediate sequential steps in mitochondrial membrane fusion. Mol. Biol.
277, 15225–15228. Cell 20, 3525–3532.
Popadin, K., Safdar, A., Kraytsberg, Y., Khrapko, K., 2014. When man got his mtDNA Song, L., Shan, Y., Lloyd, K.C., Cortopassi, G.A., 2012. Mutant twinkle increases
deletions? Aging Cell 13, 579–582. dopaminergic neurodegeneration, mtDNA deletions and modulates Parkin
Reijns, M.A., Rabe, B., Rigby, R.E., Mill, P., Astell, K.R., Lettice, L.A., Boyle, S., Leitch, expression. Hum. Mol. Genet. 21, 5147–5158.
A., Keighren, M., Kilanowski, F., Devenney, P.S., Sexton, D., Grimes, G., Holt, I.J., Sorensen, L., Ekstrand, M., Silva, J.P., Lindqvist, E., Xu, B., Rustin, P., Olson, L.,
Hill, R.E., Taylor, M.S., Lawson, K.A., Dorin, J.R., Jackson, A.P., 2012. Enzymatic Larsson, N.G., 2001. Late-onset corticohippocampal neurodepletion
removal of ribonucleotides from DNA is essential for Mammalian genome attributable to catastrophic failure of oxidative phosphorylation in MILON
integrity and development. Cell 149, 1008–1022. mice. J. Neurosci. 21, 8082–8090.
Remes, A.M., Hinttala, R., Karppa, M., Soini, H., Takalo, R., Uusimaa, J., Majamaa, K., Spelbrink, J.N., Toivonen, J.M., Hakkaart, G.A., Kurkela, J.M., Cooper, H.M., Lehtinen,
2008. Parkinsonism associated with the homozygous W748S mutation in the S.K., Lecrenier, N., Back, J.W., Speijer, D., Foury, F., Jacobs, H.T., 2000. In vivo
POLG1 gene. Parkinsonism Relat. Disord. 14, 652–654. functional analysis of the human mitochondrial DNA polymerase POLG
Reyes, A., Melchionda, L., Nasca, A., Carrara, F., Lamantea, E., Zanolini, A., Lamperti, expressed in cultured human cells. J. Biol. Chem. 275, 24818–24828.
C., Fang, M., Zhang, J., Ronchi, D., Bonato, S., Fagiolari, G., Moggio, M., Ghezzi, D., Spelbrink, J.N., Li, F.Y., Tiranti, V., Nikali, K., Yuan, Q.P., Tariq, M., Wanrooij, S.,
Zeviani, M., 2015. RNASEH1 mutations impair mtDNA replication and cause Garrido, N., Comi, G., Morandi, L., Santoro, L., Toscano, A., Fabrizi, G.M., Somer,
adult-Onset mitochondrial encephalomyopathy. Am. J. Hum. Genet. 97, H., Croxen, R., Beeson, D., Poulton, J., Suomalainen, A., Jacobs, H.T., Zeviani, M.,
186–193. Larsson, C., 2001. Human mitochondrial DNA deletions associated with
Ropp, P.A., Copeland, W.C., 1996. Cloning and characterization of the human mutations in the gene encoding twinkle, a phage T7 gene 4-like protein
mitochondrial DNA polymerase, DNA polymerase gamma. Genomics 36, localized in mitochondria. Nat. Genet. 28, 223–231.
449–458. Spendiff, S., Reza, M., Murphy, J.L., Gorman, G., Blakely, E.L., Taylor, R.W., Horvath,
Rubinsztein, D.C., Marino, G., Kroemer, G., 2011. Autophagy and aging. Cell 146, R., Campbell, G., Newman, J., Lochmuller, H., Turnbull, D.M., 2013.
682–695. Mitochondrial DNA deletions in muscle satellite cells: implications for
Ruetenik, A., Barrientos, A., 2015. Dietary restriction, mitochondrial function and therapies. Hum. Mol. Genet. 22, 4739–4747.
aging: from yeast to humans. Biochim. Biophys. Acta 1847, 1434–1447. Stein, H., Hausen, P., 1969. Enzyme from calf thymus degrading the RNA moiety of
Ruhanen, H., Ushakov, K., Yasukawa, T., 2011. Involvement of DNA ligase III and DNA-RNA Hybrids: effect on DNA-dependent RNA polymerase. Science 166,
ribonuclease H1 in mitochondrial DNA replication in cultured human cells. 393–395.
Biochim. Biophys. Acta 1813, 2000–2007. Suarez, J., Hu, Y., Makino, A., Fricovsky, E., Wang, H., Dillmann, W.H., 2008.
Safdar, A., Bourgeois, J.M., Ogborn, D.I., Little, J.P., Hettinga, B.P., Akhtar, M., Alterations in mitochondrial function and cytosolic calcium induced by
Thompson, J.E., Melov, S., Mocellin, N.J., Kujoth, G.C., Prolla, T.A., Tarnopolsky, hyperglycemia are restored by mitochondrial transcription factor A in
M.A., 2011. Endurance exercise rescues progeroid aging and induces systemic cardiomyocytes. Am. J. Physiol. Cell Physiol. 295, C1561–1568.
mitochondrial rejuvenation in mtDNA mutator mice. Proc. Natl. Acad. Sci. U. S. Sun, J., Tower, J., 1999. FLP recombinase-mediated induction of Cu/Zn-superoxide
A. 108, 4135–4140. dismutase transgene expression can extend the life span of adult Drosophila
Safdar, A., Khrapko, K., Flynn, J.M., Saleem, A., De Lisio, M., Johnston, A.P., melanogaster flies. Mol. Cell. Biol. 19, 216–228.
Kratysberg, Y., Samjoo, I.A., Kitaoka, Y., Ogborn, D.I., Little, J.P., Raha, S., Parise, Sweasy, J.B., Lauper, J.M., Eckert, K.A., 2006. DNA polymerases and human diseases.
G., Akhtar, M., Hettinga, B.P., Rowe, G.C., Arany, Z., Prolla, T.A., Tarnopolsky, Radiat. Res. 166, 693–714.
M.A., 2015. Exercise-induced mitochondrial p53 repairs mtDNA mutations in Szczesny, R.J., Hejnowicz, M.S., Steczkiewicz, K., Muszewska, A., Borowski, L.S.,
mutator mice. Skeletal Muscle 6, 7. Ginalski, K., Dziembowski, A., 2013. Identification of a novel human
Saleem, A., Carter, H.N., Hood, D.A., 2014. p53 is necessary for the adaptive changes mitochondrial endo-/exonuclease Ddk1/c20orf72 necessary for maintenance
in cellular milieu subsequent to an acute bout of endurance exercise. Am. J. of proper 7S DNA levels. Nucleic Acids Res. 41, 3144–3161.
Physiol. Cell Physiol. 306, C241–249. Taanman, J.W., Schapira, A.H., 2005. Analysis of the trinucleotide CAG repeat from
Sanchez-Martinez, A., Calleja, M., Peralta, S., Matsushima, Y., Hernandez-Sierra, R., the DNA polymerase gamma gene (POLG) in patients with Parkinson’s disease.
Whitworth, A.J., Kaguni, L.S., Garesse, R., 2012. Modeling pathogenic mutations Neurosci. Lett. 376, 56–59.
of human twinkle in Drosophila suggests an apoptosis role in response to Taivassalo, T., Fu, K., Johns, T., Arnold, D., Karpati, G., Shoubridge, E.A., 1999. Gene
mitochondrial defects. PLoS One 7, e43954. shifting: a novel therapy for mitochondrial myopathy. Hum. Mol. Genet. 8,
Sanz, A., Caro, P., Ibanez, J., Gomez, J., Gredilla, R., Barja, G., 2005. Dietary restriction 1047–1052.
at old age lowers mitochondrial oxygen radical production and leak at complex Taivassalo, T., Gardner, J.L., Taylor, R.W., Schaefer, A.M., Newman, J., Barron, M.J.,
I and oxidative DNA damage in rat brain. J. Bioenerg. Biomembr. 37, 83–90. Haller, R.G., Turnbull, D.M., 2006. Endurance training and detraining in
104 K.L. DeBalsi et al. / Ageing Research Reviews 33 (2017) 89–104

mitochondrial myopathies due to single large-scale mtDNA deletions. Brain Vermulst, M., Wanagat, J., Kujoth, G.C., Bielas, J.H., Rabinovitch, P.S., Prolla, T.A.,
129, 3391–3401. Loeb, L.A., 2008. DNA deletions and clonal mutations drive premature aging in
Tanaka, A., Ide, T., Fujino, T., Onitsuka, K., Ikeda, M., Takehara, T., Hata, Y., Ylikallio, mitochondrial mutator mice. Nat. Genet. 40, 392–394.Vermulst, M., Wanagat,
E., Tyynismaa, H., Suomalainen, A., Sunagawa, K., 2013. The overexpression of J., Loeb, L.A., 2009. On mitochondria, mutations, and methodology. Cell Metab.
twinkle helicase ameliorates the progression of cardiac fibrosis and heart 10, 437.
failure in pressure overload model in mice. PLoS One 8, e67642. Wallace, D.C., 1997. Mitochondrial DNA in aging and disease. Sci. Am. 277, 40–47.
Tarnopolsky, M.A., 2009. Mitochondrial DNA shifting in older adults following Wallace, D.C., 1999. Mitochondrial diseases in man and mouse. Science 283,
resistance exercise training. Applied Physiol. Nutr. Metab. 34, 348–354. 1482–1488.
Taylor, R.W., Barron, M.J., Borthwick, G.M., Gospel, A., Chinnery, P.F., Samuels, D.C., Wallace, D.C., 2005. A mitochondrial paradigm of metabolic and degenerative
Taylor, G.A., Plusa, S.M., Needham, S.J., Greaves, L.C., Kirkwood, T.B., Turnbull, diseases, aging, and cancer: a dawn for evolutionary medicine. Annu. Rev.
D.M., 2003. Mitochondrial DNA mutations in human colonic crypt stem cells. J. Genet. 39, 359–407.
Clin. Invest. 112, 1351–1360. Wang, D., Kreutzer, D.A., Essigmann, J.M., 1998. Mutagenicity and repair of
Taylor, S.D., Ericson, N.G., Burton, J.N., Prolla, T.A., Silber, J.R., Shendure, J., Bielas, oxidative DNA damage: insights from studies using defined lesions. Mutat. Res.
J.H., 2014. Targeted enrichment and high-resolution digital profiling of 400, 99–115.
mitochondrial DNA deletions in human brain. Aging Cell 13, 29–38. Wang, J., Wilhelmsson, H., Graff, C., Li, H., Oldfors, A., Rustin, P., Bruning, J.C., Kahn,
Testa, G., Biasi, F., Poli, G., Chiarpotto, E., 2014. Calorie restriction and dietary C.R., Clayton, D.A., Barsh, G.S., Thoren, P., Larsson, N.G., 1999. Dilated
restriction mimetics: a strategy for improving healthy aging and longevity. cardiomyopathy and atrioventricular conduction blocks induced by
Curr. Pharm. Des. 20, 2950–2977. heart-specific inactivation of mitochondrial DNA gene expression. Nat. Genet.
Tiangyou, W., Hudson, G., Ghezzi, D., Ferrari, G., Zeviani, M., Burn, D.J., Chinnery, 21, 133–137.
P.F., 2006. POLG1 in idiopathic Parkinson disease. Neurology 67, 1698–1700. Wanrooij, S., Goffart, S., Pohjoismaki, J.L., Yasukawa, T., Spelbrink, J.N., 2007.
Trifunovic, A., Wredenberg, A., Falkenberg, M., Spelbrink, J.N., Rovio, A.T., Bruder, Expression of catalytic mutants of the mtDNA helicase twinkle and polymerase
C.E., Bohlooly, Y.M., Gidlof, S., Oldfors, A., Wibom, R., Tornell, J., Jacobs, H.T., POLG causes distinct replication stalling phenotypes. Nucleic Acids Res. 35,
Larsson, N.G., 2004. Premature ageing in mice expressing defective 3238–3251.
mitochondrial DNA polymerase. Nature 429, 417–423. Wenz, T., Diaz, F., Hernandez, D., Moraes, C.T., 2009. Endurance exercise is
Trifunovic, A., Hansson, A., Wredenberg, A., Rovio, A.T., Dufour, E., Khvorostov, I., protective for mice with mitochondrial myopathy. J. Appl. Physiol. 106,
Spelbrink, J.N., Wibom, R., Jacobs, H.T., Larsson, N.G., 2005. Somatic mtDNA 1712–1719.
mutations cause aging phenotypes without affecting reactive oxygen species Wenz, T., 2011. Mitochondria and PGC-1alpha in aging and age-associated
production. Proc. Natl. Acad. Sci. U. S. A. 102, 17993–17998. diseases. J. Aging Res. 2011, 810619.
Tuppen, H.A., Blakely, E.L., Turnbull, D.M., Taylor, R.W., 2010. Mitochondrial DNA Wredenberg, A., Wibom, R., Wilhelmsson, H., Graff, C., Wiener, H.H., Burden, S.J.,
mutations and human disease. Biochim. Biophys. Acta 1797, 113–128. Oldfors, A., Westerblad, H., Larsson, N.G., 2002. Increased mitochondrial mass
Twig, G., Hyde, B., Shirihai, O.S., 2008. Mitochondrial fusion, fission and autophagy in mitochondrial myopathy mice. Proc. Natl. Acad. Sci. U. S. A. 99,
as a quality control axis: the bioenergetic view. Biochim. Biophys. Acta 1777, 15066–15071.
1092–1097. Wu, H., Sun, H., Liang, X., Lima, W.F., Crooke, S.T., 2013. Human RNase H1 is
Tyynismaa, H., Sembongi, H., Bokori-Brown, M., Granycome, C., Ashley, N., Poulton, associated with protein P32 and is involved in mitochondrial pre-rRNA
J., Jalanko, A., Spelbrink, J.N., Holt, I.J., Suomalainen, A., 2004. Twinkle helicase processing. PLoS One 8, e71006.
is essential for mtDNA maintenance and regulates mtDNA copy number. Hum. Xie, Y., Yang, H., Cunanan, C., Okamoto, K., Shibata, D., Pan, J., Barnes, D.E., Lindahl,
Mol. Genet. 13, 3219–3227. T., McIlhatton, M., Fishel, R., Miller, J.H., 2004. Deficiencies in mouse Myh and
Tyynismaa, H., Mjosund, K.P., Wanrooij, S., Lappalainen, I., Ylikallio, E., Jalanko, A., Ogg1 result in tumor predisposition and G to T mutations in codon 12 of the
Spelbrink, J.N., Paetau, A., Suomalainen, A., 2005. Mutant mitochondrial K-ras oncogene in lung tumors. Cancer Res. 64, 3096–3102.
helicase Twinkle causes multiple mtDNA deletions and a late-onset Xu, S., Zhong, M., Zhang, L., Wang, Y., Zhou, Z., Hao, Y., Zhang, W., Yang, X., Wei, A.,
mitochondrial disease in mice. Proc. Natl. Acad. Sci. U. S. A. 102, 17687–17692. Pei, L., Yu, Z., 2009. Overexpression of Tfam protects mitochondria against
Tyynismaa, H., Carroll, C.J., Raimundo, N., Ahola-Erkkila, S., Wenz, T., Ruhanen, H., beta-amyloid-induced oxidative damage in SH-SY5Y cells. FEBS J. 276,
Guse, K., Hemminki, A., Peltola-Mjosund, K.E., Tulkki, V., Oresic, M., Moraes, 3800–3809.
C.T., Pietilainen, K., Hovatta, I., Suomalainen, A., 2010. Mitochondrial myopathy Yatsuga, S., Suomalainen, A., 2012. Effect of bezafibrate treatment on late-onset
induces a starvation-like response. Hum. Mol. Genet. 19, 3948–3958. mitochondrial myopathy in mice. Hum. Mol. Genet. 21, 526–535.
Van Goethem, G., Dermaut, B., Lofgren, A., Martin, J.J., Van Broeckhoven, C., 2001. Yen, T.C., Su, J.H., King, K.L., Wei, Y.H., 1991. Ageing-associated 5 kb deletion in
Mutation of POLG is associated with progressive external ophthalmoplegia human liver mitochondrial DNA. Biochem. Biophys. Res. Commun. 178,
characterized by mtDNA deletions. Nat. Genet. 28, 211–212. 124–131.
Van Goethem, G., Martin, J.J., Dermaut, B., Lofgren, A., Wibail, A., Ververken, D., Ylikallio, E., Tyynismaa, H., Tsutsui, H., Ide, T., Suomalainen, A., 2010. High
Tack, P., Dehaene, I., Van Zandijcke, M., Moonen, M., Ceuterick, C., De Jonghe, P., mitochondrial DNA copy number has detrimental effects in mice. Hum. Mol.
Van Broeckhoven, C., 2003a. Recessive POLG mutations presenting with Genet. 19, 2695–2705.
sensory and ataxic neuropathy in compound heterozygote patients with Ylonen, S., Ylikotila, P., Siitonen, A., Finnila, S., Autere, J., Majamaa, K., 2013.
progressive external ophthalmoplegia. Neuromuscul. Disord. 13, 133–142. Variations of mitochondrial DNA polymerase gamma in patients with
Van Goethem, G., Martin, J.J., Van Broeckhoven, C., 2003b. Progressive external Parkinson’s disease. J. Neurol. 260, 3144–3149.
ophthalmoplegia characterized by multiple deletions of mitochondrial DNA: Young, M.J., Longley, M.J., Li, F.Y., Kasiviswanathan, R., Wong, L.J., Copeland, W.C.,
unraveling the pathogenesis of human mitochondrial DNA instability and the 2011. Biochemical analysis of human POLG2 variants associated with
initiation of a genetic classification. Neuromol. Med. 3, 129–146. mitochondrial disease. Hum. Mol. Genet. 20, 3052–3066.
Van Goethem, G., Schwartz, M., Lofgren, A., Dermaut, B., Van Broeckhoven, C., Young, M.J., Humble, M.M., DeBalsi, K.L., Sun, K.Y., Copeland, W.C., 2015. POLG2
Vissing, J., 2003c. Novel POLG mutations in progressive external disease variants: analyses reveal a dominant negative heterodimer, altered
ophthalmoplegia mimicking mitochondrial neurogastrointestinal mitochondrial localization and impaired respiratory capacity. Hum. Mol.
encephalomyopathy. Eur. J. Hum. Genet. 11, 547–549. Genet. 24, 5184–5197.
Van Goethem, G., Luoma, P., Rantamaki, M., Al Memar, A., Kaakkola, S., Hackman, Zeviani, M., Servidei, S., Gellera, C., Bertini, E., DiMauro, S., DiDonato, S., 1989. An
P., Krahe, R., Lofgren, A., Martin, J.J., De Jonghe, P., Suomalainen, A., Udd, B., Van autosomal dominant disorder with multiple deletions of mitochondrial DNA
Broeckhoven, C., 2004. POLG mutations in neurodegenerative disorders with starting at the D-loop region. Nature 339, 309–311.
ataxia but no muscle involvement. Neurology 63, 1251–1257. Zhang, D., Mott, J.L., Chang, S.W., Denniger, G., Feng, Z., Zassenhaus, H.P., 2000.
Van Raamsdonk, J.M., Hekimi, S., 2012. Superoxide dismutase is dispensable for Construction of transgenic mice with tissue-specific acceleration of
normal animal lifespan. Proc. Natl. Acad. Sci. U. S. A. 109, 5785–5790. mitochondrial DNA mutagenesis. Genomics 69, 151–161.
Van Remmen, H., Ikeno, Y., Hamilton, M., Pahlavani, M., Wolf, N., Thorpe, S.R., Zhang, D., Mott, J.L., Farrar, P., Ryerse, J.S., Chang, S.W., Stevens, M., Denniger, G.,
Alderson, N.L., Baynes, J.W., Epstein, C.J., Huang, T.T., Nelson, J., Strong, R., Zassenhaus, H.P., 2003. Mitochondrial DNA mutations activate the
Richardson, A., 2003. Life-long reduction in MnSOD activity results in mitochondrial apoptotic pathway and cause dilated cardiomyopathy.
increased DNA damage and higher incidence of cancer but does not accelerate Cardiovasc. Res. 57, 147–157.
aging. Physiol. Genomics 16, 29–37. Zhang, D., Mott, J.L., Chang, S.W., Stevens, M., Mikolajczak, P., Zassenhaus, H.P.,
Ventura-Clapier, R., Garnier, A., Veksler, V., 2008. Transcriptional control of 2005. Mitochondrial DNA mutations activate programmed cell survival in the
mitochondrial biogenesis: the central role of PGC-1alpha. Cardiovasc. Res. 79, mouse heart. Am. J. Physiol. Heart Circ. Physiol. 288, H2476–2483.
208–217. Zheng, W., Khrapko, K., Coller, H., Thilly, W.G., Copeland, W.C., 2006. Origins of
Vermulst, M., Bielas, J.H., Kujoth, G.C., Ladiges, W.C., Rabinovitch, P.S., Prolla, T.A., human mitochondrial point mutations as DNA polymerase gamma-mediated
Loeb, L.A., 2007. Mitochondrial point mutations do not limit the natural errors. Mutat. Res. 599, 11–20.
lifespan of mice. Nat. Genet. 39, 540–543.

You might also like