ISSN 0006-2979, Biochemistry (Moscow), 2025, Vol. 90, No. 12, pp. 1919-1928 © The Author(s) 2025. This article is an open access publication.
Published in Russian in Biokhimiya, 2025, Vol. 90, No. 12, pp. 2021-2031.
1919
REVIEW
Mechanisms of Intracellular Selection
of Mitochondrial DNA
Georgii Muravyov
1,2
and Dmitry A. Knorre
2,a
*
1
Faculty of Bioengineering and Bioinformatics, Lomonosov Moscow State University,
119234 Moscow, Russia
2
A. N. Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University,
119234 Moscow, Russia
a
e-mail: knorre@belozersky.msu.ru
Received September 23, 2025
Revised November 15, 2025
Accepted November 20, 2025
AbstractEukaryotic cells contain multiple mitochondrial DNA (mtDNA) molecules. Heteroplasmy is coexis-
tence in the same cell of different mtDNA variants competing for cellular resources required for their repli-
cation. Here, we review documented cases of emergence and spread of selfish mtDNA (i.e., mtDNA that has
a selective advantage in a cell but decreases cell fitness) in eukaryotic species, from humans to baker’s yeast.
The review discusses hypothetical mechanisms enabling preferential proliferation of certain mtDNA variants
in heteroplasmy. We propose that selfish mtDNAs have significantly influenced the evolution of eukaryotes
and may be responsible for the emergence of uniparental inheritance and constraints on the mtDNA copy
number in germline cells.
DOI: 10.1134/S0006297925603296
Keywords: mitochondrial DNA, mtDNA quality control, mitophagy, heteroplasmy, intracellular selection, selfish
gene
* To whom correspondence should be addressed.
INTRODUCTION
Mitochondria are semi-autonomous organelles
with their own genome. Mitochondrial DNA (mtDNA)
has been found in a vast majority of eukaryotes, ex-
cept few species [1]. It encodes proteins necessary
for oxidative phosphorylation, as well as components
of the mitochondrial translation system [2]. Typical-
ly, a single eukaryotic cell contains several mtDNA
copies, although their exact number may depend on
the conditions, cell type, and prior cell history [3-5].
New mtDNA variants appear through mutations.
The state when a cell has several mtDNA variants is
called heteroplasmy [6]. During cell division, mtDNA
molecules are distributed randomly between the
new cells. After several (sometimes, many) genera-
tions, due to random genetic drift, the descendants
of a heteroplasmic cell retain only one of the mtDNA
variants, the state referred to as homoplasmy [7].
As a result, the emergence of heteroplasmy (due to
mutations or cell fusion) is balanced by random ge-
netic drift in the frequencies of mtDNA mitotypes,
leading to the preservation of one or another mtDNA
mitotype in the cell descendants (Fig. 1).
mtDNA is a subject of natural selection at multi-
ple organizational levels: molecular, organelle, cellu-
lar, or organismal [7]. Therefore, the selection process
can be multidirectional, i.e., the same mtDNA variant
might have a high fitness at one organizational level
and low fitness at another [7]. To describe selection
at the cellular level, it is important to introduce the
concept of mutant mtDNA pathogenicity threshold.
Cell phenotype is a complex function of mtDNA gen-
otypes, as it reflects manifestations of these geno-
types at the levels of transcription, translation, en-
zymatic activity, respiratory chain function, and cell
as a whole. However, when the proportion of mutant
mtDNA variants is less than a certain threshold for
a given heteroplasmic cell, harmful mutations may
not manifest themselves at the cellular or organismal
MURAVYOV, KNORRE1920
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
levels [8, 9], usually indicating an excess number of
mtDNA copies per cell. As a result, mtDNA variants
carrying mutations harmful to the cell and, at the
same time, providing an advantage in fitness at the
intracellular level, can avoid selection in individual
cells until their allele frequency reaches a specific
threshold. Such mutations can persist for a long time
or even gain an advantage in organisms and popula-
tions. Furthermore, in some cases, deleterious mtDNA
mutations harmful to the multicellular organism, may
gain a fitness advantage at the cellular level. This is
exemplified by the spread of colonic crypts contain-
ing cells with mutations in the mitochondrial cyto-
chrome oxidase CO1 gene [10].
mtDNA variants characterized by a high fitness
at the intracellular level but not beneficial (or even
deleterious) to the cell or entire organism fall un-
der the definition of “selfish genetic elements” [11].
In this work, we will refer to such mitotypes (mtDNA
variants) as selfish mtDNA. There have been numer-
ous descriptions of the emergence and spreading of
selfish mtDNA in both nature and artificial experi-
mental systems. In the next section, we discussed ex-
amples of intracellular selection of mtDNA enabling
proliferation of selfish mtDNA.
INTRACELLULAR SELECTION
AND SELFISH mtDNA IN NATURE
AND EXPERIMENTAL SYSTEMS
One evidence of mtDNA selection at the intracel-
lular level is clonal expansion of deletion-containing
mtDNA variants in the cells of postmitotic tissues of
multicellular animals. Clonal expansion is the process
where a mutated mtDNA molecule multiplies within
some cells, leading to a higher concentration of that
specific mtDNA, while in other cells, this mtDNA
variant is absent [12,  13]. Analysis of sequences of
mtDNA control regions in individual human somat-
ic cells has shown that proportion of mutant mtDNA
variants (i.e., those differing from the main mitotype
in a given organism) in some cells can significantly
increase with age [14]. In most cases, it is difficult to
distinguish whether the clonal expansion is caused
random stochastic drift or results from positive intra-
cellular selection. Nonetheless, the mutation patterns
found in individual cells from different tissues dif-
fer markedly, providing suggestive evidence for the
presence of positive intracellular selection. For exam-
ple, it has been shown that in cardiomyocytes, most
de  novo mutations occurred in the mtDNA genome
region with the coordinates 16,025-16,055  bp, where-
as in buccal epithelial cells, this region contained no
mutations [14]. This result suggests that either (1)the
spectrum of mutations varies greatly among differ-
ent tissues or (2) due to the intracellular selection,
some mtDNA variants undergo clonal expansion. Both
these processes – selection and genetic drift – have
been demonstrated in mice with artificially created
heteroplasmy. The animals showed an increase in the
frequency of certain mtDNA variants in some tissues,
but not in others [15]. In another study, sequencing
mtDNA from individual cells of aged (2-year-old) and
young mice revealed mutant mtDNA variants, the
proportion of which increased with age faster than
Fig. 1. Heteroplasmy arises as a result of mtDNA mutations in a cell or via cell fusion with other cells. Over generations,
random drift may cause descendants of the heteroplasmic cells to revert to homoplasmy. Mitochondria are not shown
in this figure, and it is assumed that mtDNAs constitute a single population within each cell.
MECHANISMS OF INTRACELLULAR SELECTION OF MITOCHONDRIAL DNA 1921
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
can be explained by random drift [16]. These muta-
tions, which were under positive selection in somat-
ic cells, were predominantly located in the mtDNA
control region containing the origin of replication.
Interestingly, in some cases, the relative content of
associated passenger mutations also increased [16].
Finally, comparison of frequencies of different
mtDNA variants in mothers and offspring suggested
the presence of both purifying [17, 18] and positive
[19, 20] selection of mtDNAs in mammalian germ-
lines. The direction of selection largely depends on
the frequencies of mtDNA alleles in germline cells
and specific mtDNA mitotypes (see discussion in [21,
22]). If a mutant mtDNA variant (mitotype) consti-
tutes a significant fraction of all molecules, the puri-
fying selection may occur at the whole-cell level. The
selection is enabled because, during the development
of follicular cells, the mtDNA diversity within indi-
vidual cells decreases due to the genetic bottleneck
effect (see review [23]), linking the mtDNA genotype
to the phenotype of the entire cell. At the same time,
both purifying selection against deleterious mtDNAs
with low allele frequencies and any kind of positive
selection are most likely restricted to the intracellular
level.
Selfish mtDNA is one of the challenges in the
mitochondrial replacement therapy. Mitochondrial
heteroplasmy is associated with a large number of
human hereditary diseases. Such diseases can be
transmitted from a mother to child, and their her-
itability depends on the ratio of mutant (pathogen-
ic) and normal mtDNA variants in the oocytes [24].
To prevent the spread of such diseases, scientists
have been actively developing the methods for mi-
tochondrial replacement therapy, leading to the first
live birth of a “three-parent” baby whose nuclear
DNA originated from the parents, while mtDNA was
from a donor (“third parent”) [25]. The transfer of
the spindle from the cytoplasm of the mothers cell,
who carried the pathogenic variant of mtDNA, into
the cytoplasm of the donor cell resulted in the do-
nors mtDNA constituting over 99% mtDNA in the de-
veloping embryo. However, in some cases, embryonic
stem cell lines derived by this method demonstrated
a gradual increase in the pathogenic variant of the
mothers mtDNA and loss of “healthy” donors mtDNA
[26], suggesting that the mutant mtDNA variant asso-
ciated with the disease had a higher intracellular fit-
ness compared to the donor mtDNA.
Examples of selfish mtDNA can be found among
invertebrates. For instance, in Caenorhabditis brigg-
sae, an mtDNA variant with a large deletion might
have an advantage in the cell, but was harmful at
the organismal level [27]. In Caenorhabditis elegans,
mtDNA variants with a large deletion have been
maintained in the state of heteroplasmy with the
wild-type mtDNA due to the effects of multidirectional
intra- and interorganismal selection. [28]. Remarkably,
mtDNA from one species can displace native mtDNA
in another species, as has been demonstrated in het-
eroplasmic fruit flies harboring two different mtDNAs
from two closely related species, Drosophila mauriti-
ana and Drosophila melanogaster [29].
Finally, an example of an organism in which self-
ish mtDNA variants can arise as a result of mutations
is baker’s yeast Saccharomyces cerevisiae. Yeast cells
inherit mtDNA from both parents (mating haploid
cells) [30]. If these parental haploid cells have dif-
ferent mitotypes, the progeny will have mitochondri-
al heteroplasmy. At the same time, some variants of
yeast mtDNA with extensive deletions (rho
) demon-
strated a bias in the mtDNA inheritance: crossing rho
cells with wild-type rho
+
cells led to the formation
of almost exclusively rho
diploid cells [31-33]. We
have recently shown that this bias can be explained
by both an increased number of mtDNA copies in
rho
cells and elevated intracellular fitness of rho
mtDNA variants compared to wild-type rho
+
mtDNA
variants [34]. Mutant mtDNA variants also emerged
and rapidly become fixed (remained the only vari-
ants) in yeast cells grown in the absence of selection
pressure at the whole-cell level, indicating the exis-
tence of intracellular selection favoring proliferation
of mtDNA variants with extensive deletions [35].
HYPOTHETICAL MECHANISMS
PROVIDING AN ADVANTAGE
TO SELFISH mtDNA IN THE CELL
While the presence of selfish mtDNAs in multi-
ple cell types and species has been experimentally
confirmed, the mechanisms that confer their compet-
itive advantage over wild-type DNA remain poorly
understood. Nonetheless, there are several hypothe-
ses, with a varying degree of indirect experimental
support, that aim to explain the differences in the in-
tracellular fitness between the wild-type and mutant
mtDNA mitotypes.
Increased replication initiation frequency. The
mechanism allowing one mtDNA variant to displace
another mtDNA variant may involve the differences in
the replication initiation frequency. This mechanism
implies that mutations in the mtDNA replication ori-
gin may result in the increased replication frequency
for one mtDNA variant compared to another (Fig.  2a).
Even a slight increase in the replication frequency, if
it occurs over many replication rounds, could lead to
the displacement of one mtDNA variant by another
during either individual development or across the
generations. In support of this assumption, it has
been found that the drivers for the age-associated
MURAVYOV, KNORRE1922
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
accumulation of mutant mtDNA are mtDNA variants
with mutations in the non-coding region containing
the OriH sequence responsible for the initiation of
mtDNA replication [16]. In D.  melanogaster fruit flies
with artificially created heteroplasmy, the “winning”
mtDNA variant was determined, at least in some cas-
es, by the sequence of the non-coding region [29].
However, the possibility of emergence of mtDNA
variants with “more active” replication origins leaves
open the question of why such variants had not ap-
peared and been fixed during the evolution. We sug-
gested that mutations increasing mtDNA fitness at
the intracellular level inevitably reduce the fitness
of these variants at the cellular or organismal lev-
els. For instance, an increased replication frequency
could reduce the transcription rate, since initiation
of one process delays the other [36]. Excessive repli-
cation initiation may increase the amount of mtDNA
per cell beyond the optimal levels, potentially causing
detrimental effects.
Advantages in the replication rate of mtDNA
molecules with large deletions. Another hypothet-
ically possible mechanism allowing some mtDNA
variants to have a higher intracellular fitness is the
difference in the duplication rate, as large deletions
reduce the time required for complete mtDNA rep-
lication (Fig. 2b). However, the key factor in this
case is the relation between the mtDNA duplication
time and replication initiation frequency, which de-
termines the proportion of replicating mtDNA mole-
cules at a given moment of time. If the fraction of
simultaneously replicating molecules is small, then
the time required for the replication of the entire
molecule should not have a significant effect on their
intracellular fitness. To illustrate this, let us consider
an extreme case, when there is only one functional
replisome in the mitochondria and the replication of
the next molecule does start until the replication of
the previous molecule is complete. Obviously, in this
case, the replication time will have no effect on the
relative intracellular fitness of the mtDNA molecules.
The elongation rate of human mitochondrial
DNA polymerase is 180-270 bases per minute [37,
38], suggesting that the replication of entire human
mtDNA molecule takes no more than 2.5 h. This is
expected to be true even considering that individual
Fig. 2. Hypothetical mechanisms allowing a particular variant of mtDNA to have a higher intracellular fitness despite the
presence of harmful mutations (see the text for discussion).
MECHANISMS OF INTRACELLULAR SELECTION OF MITOCHONDRIAL DNA 1923
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
mtDNA strands replicate asynchronously (see reviews
[39, 40]). At the same time, the half-life of mtDNA in
post-mitotic mammalian tissues is 1 to 3 weeks [41].
According to the results of computer modelling (tak-
ing into account several additional assumptions on
the mtDNA copy number in a cell), the time required
for the complete displacement of normal mtDNA by
deletion-carrying mtDNA molecules would be decades
[41]. Hence, it is unlikely that this mechanism has
a significant effect on the spread of mutant mtDNA
variants with deletions in postmitotic human tissues.
On the other hand, if mitochondrial biogenesis
and associated proliferation of mtDNA proceed rap-
idly, mutant deletion-carrying mtDNA variants might
gain an advantage. For example, when the content
of mtDNA was reduced by treatment with the DNA
intercalating agent ethidium bromide and then ethid-
ium bromide was removed, mtDNA variants with
large deletions repopulated the cells faster than the
full-length mtDNA molecules [42]. Following this line
of reasoning, mtDNA variants with deletions would
also gain an advantage in the case of elevated mtDNA
turnover, i.e., when the rates of mtDNA replication
and degradation increase simultaneously. Indeed,
mtDNA is actively degraded as a result of mitoph-
agy activation or mtDNA damage [3,  43]. Moreover,
mtDNA variants with deletions may gain a greater
relative advantage in species with a large mitochon-
drial genome and short cell cycle, which implies a
high proportion of replicating mtDNA molecules at
any given time. Such organisms include yeast species,
some of which have mitochondrial genomes exceed-
ing 100 kb [44].
Avoidance of mtDNA degradation. mtDNA is
constantly synthesized and degraded, even in postmi-
totic tissues [41]. mtDNA is also actively degraded in
germline cells, ensuring the mtDNA inheritance strict-
ly along the maternal line [45]. mtDNA degradation
occurs predominantly in the mitochondrial matrix
due to the activity of mitochondrial nucleases [46]
or during autophagy, along with the degradation of
mitochondria [47]. Theoretically, both these processes
can be selective towards some mtDNA variants com-
pared to others (Fig. 2, c, d). Let us consider these
processes separately.
Selective autophagy of mitochondria (mitophagy)
can ensuring the quality control of mtDNA, but this
requires specific mtDNA molecules to be “linked” to
protein variants encoded in these molecules [48,  49].
Such linkage is possible due to restrictions on the dif-
fusion of mtDNA-encoded complexes through the mi-
tochondrial network [50] or when the mitochondrial
dynamics processes (continuous mitochondrial fusion
and fission) are limited. For example, in Drosophi-
la, mutant mtDNA variants with harmful mutations
are eliminated via mitophagy during oogenesis [51].
This process is preceded by a decrease in the amount
of mitofusin protein responsible for the mitochon-
drial fusion, leading to the fragmentation of mito-
chondrial reticulum. It is important to note that this
mechanism takes place only in germline cells, but not
in somatic ones [51]. Therefore, selective mitophagy
in the absence of mitochondrial dynamics can lead
to the elimination of mutant mtDNA variants, thus
preventing their proliferation.
At present, no mechanism has been identified by
which selfish mtDNA could evade the control by mito-
phagy and gain advantage over the wild-type mtDNA.
However, it could be assumed that if an mtDNA vari-
ant encodes a factor that inhibits mitophagy or pro-
motes mitochondrial fusion, it can increase its own
intracellular fitness by evading the quality control
at the level of individual organelles (Fig.  2d). Despite
that mtDNA typically encodes a strictly defined set
of genes, mtDNA of some invertebrate species con-
tains ORFans – open reading frames that encode pro-
teins with no homology to any known protein, but
subjected to the purifying selection pressure [52,  53].
Although the functions of the encoded proteins re-
main unknown, we propose that they might modulate
the mitophagy or mitochondrial dynamics, thereby
contributing to the increased intracellular fitness of
mtDNA variants harboring them.
Another mechanism of mtDNA degradation is as-
sociated with the activity of nucleases located in the
mitochondrial matrix. The presence of double-strand
breaks in mtDNA triggers rapid degradation of lin-
ear mtDNA fragments, driven by the exonuclease
activities of mtDNA polymerase gamma and MGME1
nuclease [54]. A deficit in the exonuclease activity
can drive accumulation of deletion-carrying mtDNA
variants, as was found in mice with a homozygous
mutation in the mitochondrial mtDNA polymerase
gene (PolG
D257A/D257A
), leading to the disruption of its
exonuclease activity [55]. The degradation of male
mtDNA in Drosophila spermatids requires the presence
of another mitochondrial endonuclease, EndoG [56],
and exonuclease Poldip2 [46]. Finally, the mitochon-
drial genomes of some fungal species contain hom-
ing endonucleases capable of cleaving DNA and then
integrating their coding sequence at the site of the
double-strand break. This allows endonuclease genes
to spread in the population as a selfish element of
mitochondrial genome [57, 58]. Taken together, these
data suggest that mtDNA variants whose sequences
lack the sites for hydrolysis with mitochondrial nucle-
ases may gain an advantage under conditions when
mtDNA is actively degraded (Fig. 2c), e.g., as a result
of damage or during certain life cycle stages.
Survival of the weakest. In 1996, Aubrey de Grey
proposed that the wild-type mtDNA associated with
a fully functional respiratory chain might be more
MURAVYOV, KNORRE1924
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
susceptible to degradation than mutant mtDNA un-
able to provide an assembly of the functional respi-
ratory chain [59]. This hypothesis was named “sur-
vival of the weakest” (Fig.2e). Indeed, in some cases,
mitochondrial respiratory chain can be a source of
reactive oxygen species (ROS) [60, 61] that cause mu-
tations in mtDNA. This is consistent with the fact
that the C to T mutations typical of oxidative damage
in a single-stranded DNA, dominate in vertebrates
[62, 63]. At the same time, the damage to mtDNA
caused by hydrogen peroxide (one of ROS forms) can
induce mtDNA degradation [64].
However, it should be noted that this hypothe-
sis has not been confirmed experimentally, likely be-
cause there are other factors that can exert a strong
effect in the opposite direction. Indeed, the import
of mitochondrial proteins essential for mtDNA rep-
lication depends on the transmembrane potential at
the inner mitochondrial membrane and ATP content
in the mitochondrial matrix [65]. Therefore, mutant
variants unable to provide formation of the func-
tional respiratory chain and ATP synthase will be at
a disadvantage in this respect. Moreover, mitochon-
dria in cells actively fuse and divide, which equaliz-
es their content throughout the cell’s mitochondrial
network [66, 67] and makes it difficult to link spe-
cific mtDNA molecules to the areas of mitochondria
enriched with proteins encoded in them.
CONCLUSION
mtDNA selection occurs simultaneously at sever-
al organizational levels. Consequently, the emergence
of selfish mtDNA variants that have a replicative ad-
vantage but are detrimental to the cell or multicel-
lular organism is possible. Despite the mechanisms
that eliminate mutant mtDNA, selfish mtDNAs can
emerge in nature and experimental systems, that
might have an advantage due to a faster replication
or ability to evade degradation. Therefore, eukaryotic
organisms are under continuous pressure created by
the potential emergence of such mtDNA.
We propose that during the early stages of eu-
karyotic evolution, before the mitochondrial quali-
ty control systems evolved to suppress proliferation
of mutant mtDNA, selfish mtDNAs had posed a sig-
nificant challenge to eukaryotic cells. This pressure
has likely drove the evolution of cellular mecha-
nisms protecting against selfish mtDNA. Such mech-
anisms may include control of mtDNA copy number
in the germlines of multicellular animals [68], se-
lective mitophagy [69], and uniparental inheritance
of mtDNA in most eukaryotic species [70], which
prevents the horizontal transfer of selfish mtDNA
within populations.
Abbreviations
mtDNA mitochondrial DNA
Acknowledgments
This work is dedicated to the anniversary of Vladimir
Petrovich Skulachev, whose inexhaustible scientific
enthusiasm and original ideas have inspired the au-
thors of this review. We are grateful to Boris Alexan-
drovich Feniouk for discussing our work and provid-
ing critical comments.
Contributions
G.M. and D.K. analyzed the published articles and
wrote the manuscript; D.K. prepared the figures.
Funding
This work (all sections except “Avoidance of mtDNA
degradation”) was supported by the Russian Science
Foundation (project no.22-14-00108-П); “Avoidance of
mtDNA degradation” section was written as a part of
the State Assignment to the Lomonosov Moscow State
University (project no.AAAA-A19-119031390114-5).
Ethics approval and consent to participate
This work does not contain any studies involving
human or animal subjects.
Conflict of interest
The authors of this work declare that they have
noconflicts of interest.
Open access
This article is licensed under a Creative Commons At-
tribution 4.0 International License, which permits use,
sharing, adaptation, distribution, and reproduction
in any medium or format, as long as you give appro-
priate credit to the original author(s) and the source,
provide a link to the Creative Commons license, and
indicate if changes were made. The images or other
third party material in this article are included in the
article’s Creative Commons license, unless indicated
otherwise in a credit line to the material. If materi-
al is not included in the article’s Creative Commons
license and your intended use is not permitted by
statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit
http://creativecommons.org/licenses/by/4.0/.
REFERENCES
1. Roger, A. J., Muñoz-Gómez, S. A., and Kamikawa, R.
(2017) The origin and diversification of mitochondria,
Curr. Biol., 27, R1177-R1192, https://doi.org/10.1016/
j.cub.2017.09.015.
MECHANISMS OF INTRACELLULAR SELECTION OF MITOCHONDRIAL DNA 1925
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
2. Taanman, J. W. (1999) The mitochondrial genome:
structure, transcription, translation and replica-
tion, Biochim. Biophys. Acta, 1410, 103-123, https://
doi.org/10.1016/s0005-2728(98)00161-3.
3. Medeiros, T. C., Thomas, R. L., Ghillebert, R., and
Graef, M. (2018) Autophagy balances mtDNA syn-
thesis and degradation by DNA polymerase POLG
during starvation, J. Cell Biol., 217, 1601-1611, https://
doi.org/10.1083/jcb.201801168.
4. Zhang, Q., Wang, Z., Zhang, W., Wen, Q., Li, X.,
Zhou,J., Wu,X., Guo,Y., Liu,Y., Wei,C., Qian,W., and
Tian,Y. (2021) The memory of neuronal mitochondri-
al stress is inherited transgenerationally via elevated
mitochondrial DNA levels, Nat. Cell Biol., 23, 870-880,
https://doi.org/10.1038/s41556-021-00724-8.
5. Galeota-Sprung, B., Fernandez,A., and Sniegowski,P.
(2022) Changes to the mtDNA copy number during
yeast culture growth, R. Soc. Open Sci., 9, 211842,
https://doi.org/10.1098/rsos.211842.
6. De Giorgi, C., and Saccone, C. (1989) Mitochondri-
al genome in animal cells. Structure, organiza-
tion, and evolution, Cell Biophys., 14, 67-78, https://
doi.org/10.1007/BF02797392.
7. Rand, D.M. (2001) The units of selection on mitochon-
drial DNA, Annu Rev Ecol Syst, 32, 415-448, https://
doi.org/10.1146/annurev.ecolsys.32.081501.114109.
8. Rossignol, R., Faustin, B., Rocher, C., Malgat, M.,
Mazat, J.-P., and Letellier, T. (2003) Mitochondrial
threshold effects, Biochem. J., 370, 751-762, https://
doi.org/10.1042/BJ20021594.
9. Wallace, D.C. (1986) Mitotic segregation of mitochon-
drial DNAs in human cell hybrids and expression of
chloramphenicol resistance, Somat. Cell Mol. Genet.,
12, 41-49, https://doi.org/10.1007/BF01560726.
10. Greaves, L. C., Preston, S. L., Tadrous, P. J., Taylor,
R. W., Barron, M. J., Oukrif, D., Leedham, S. J.,
Deheragoda,M., Sasieni,P., Novelli, M.R., Jankowski,
J.A.Z., Turnbull, D.M., Wright, N.A., and McDonald,
S. A. C. (2006) Mitochondrial DNA mutations are
established in human colonic stem cells, and mu-
tated clones expand by crypt fission, Proc. Natl.
Acad. Sci. USA, 103, 714-719, https://doi.org/10.1073/
pnas.0505903103.
11. Ågren, J. A., and Clark, A. G. (2018) Selfish genet-
ic elements, PLoS Genet., 14, e1007700, https://
doi.org/10.1371/journal.pgen.1007700.
12. Khrapko, K., Bodyak, N., Thilly, W. G., van Orsouw,
N. J., Zhang, X., Coller, H. A., Perls, T. T., Upton, M.,
Vijg, J., and Wei, J. Y. (1999) Cell-by-cell scanning
of whole mitochondrial genomes in aged human
heart reveals a significant fraction of myocytes with
clonally expanded deletions, Nucleic Acids Res., 27,
2434-2441, https://doi.org/10.1093/nar/27.11.2434.
13. Taylor, S.D., Ericson, N.G., Burton, J.N., Prolla, T.A.,
Silber, J. R., Shendure, J., and Bielas, J. H. (2014)
Targeted enrichment and high-resolution digital
profiling of mitochondrial DNA deletions in human
brain, Aging Cell, 13, 29-38, https://doi.org/10.1111/
acel.12146.
14. Nekhaeva,E., Bodyak, N.D., Kraytsberg,Y., McGrath,
S. B., Van Orsouw, N. J., Pluzhnikov, A., Wei, J. Y.,
Vijg, J., and Khrapko, K. (2002) Clonally expanded
mtDNA point mutations are abundant in individual
cells of human tissues, Proc. Natl. Acad. Sci. USA, 99,
5521-5526, https://doi.org/10.1073/pnas.072670199.
15. Jenuth, J. P., Peterson, A. C., and Shoubridge, E. A.
(1997) Tissue-specific selection for different mtDNA
genotypes in heteroplasmic mice, Nat. Genet., 16,
93-95, https://doi.org/10.1038/ng0597-93.
16. Korotkevich, E., Conrad, D. N., Gartner, Z. J., and
O’Farrell, P. H. (2025) Selfish mutations promote
age-associated erosion of mtDNA integrity in mam-
mals, Nat. Commun., 16, 5435, https://doi.org/10.1038/
s41467-025-60477-y.
17. Stewart, J. B., Freyer, C., Elson, J. L., Wredenberg, A.,
Cansu, Z., Trifunovic, A., and Larsson, N.-G. (2008)
Strong purifying selection in transmission of mam-
malian mitochondrial DNA, PLoS Biol., 6, e10, https://
doi.org/10.1371/journal.pbio.0060010.
18. Wei, W., Tuna, S., Keogh, M. J., Smith, K. R., Aitman,
T. J., Beales, P. L., Bennett, D. L., Gale, D. P., Bitner-
Glindzicz, M.A.K., Black, G.C., Brennan,P., Elliott,P.,
Flinter, F. A., Floto, R. A., Houlden, H., Irving, M.,
Koziell, A., Maher, E. R., Markus, H. S., et al. (2019)
Germline selection shapes human mitochondri-
al DNA diversity, Science, 364, eaau6520, https://
doi.org/10.1126/science.aau6520.
19. Franco,M., Pickett, S.J., Fleischmann,Z., Khrapko,M.,
Cote-L’Heureux,A., Aidlen,D., Stein,D., Markuzon,N.,
Popadin, K., Braverman, M., Woods, D. C., Tilly, J. L.,
Turnbull, D. M., and Khrapko, K. (2022) Dynamics
of the most common pathogenic mtDNA variant
m.3243A > G demonstrate frequency-dependen-
cy in blood and positive selection in the germline,
Hum. Mol. Genet., 31, 4075-4086, https://doi.org/
10.1093/hmg/ddac149.
20. Zhang, H., Esposito, M., Pezet, M. G., Aryaman, J.,
Wei,W., Klimm,F., Calabrese,C., Burr, S.P., Macabelli,
C. H., Viscomi, C., Saitou, M., Chiaratti, M. R., Stew-
art, J. B., Jones, N., and Chinnery, P. F. (2021) Mito-
chondrial DNA heteroplasmy is modulated during
oocyte development propagating mutation transmis-
sion, Sci. Adv., 7, eabi5657, https://doi.org/10.1126/
sciadv.abi5657.
21. Franco, M., Cote-L’Heureux, A., Fleischmann, Z.,
Chen, Z., Khrapko, M., Vyshedskiy, B., Braver-
man, M., Popadin, K., Pickett, S., Woods, D. C., Tilly,
J. L., Turnbull, D., and Khrapko, K. (2023) A nov-
el, wave-shaped profile of germline selection of
pathogenic mtDNA mutations is discovered by by-
passing a classical statistical bias, bioRxiv, https://
doi.org/10.1101/2023.11.21.568140.
MURAVYOV, KNORRE1926
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
22. Fleischmann, Z., Cote-L’Heureux, A., Franco, M.,
Oreshkov, S., Annis, S., Khrapko, M., Aidlen, D.,
Popadin,K., Woods, D.C., Tilly, J.L., and Khrapko,K.
(2024) Reanalysis of mtDNA mutations of human
primordial germ cells (PGCs) reveals NUMT con-
tamination and suggests that selection in PGCs
may be positive, Mitochondrion, 74, 101817, https://
doi.org/10.1016/j.mito.2023.10.005.
23. Mishra, P., and Chan, D. C. (2014) Mitochondrial dy-
namics and inheritance during cell division, devel-
opment and disease, Nat. Rev. Mol. Cell Biol., 15,
634-646, https://doi.org/10.1038/nrm3877.
24. Taylor, R.W., and Turnbull, D.M. (2005) Mitochondri-
al DNA mutations in human disease, Nat. Rev. Genet.,
6, 389-402, https://doi.org/10.1038/nrg1606.
25. Zhang, J., Liu, H., Luo, S., Chavez-Badiola, A., Liu, Z.,
Yang, M., Munne, S., Konstantinidis, M., Wells, D.,
and Huang, T. (2016) First live birth using human
oocytes reconstituted by spindle nuclear transfer
for mitochondrial DNA mutation causing Leigh syn-
drome, Fertil. Steril., 106, e375-e376, https://doi.org/
10.1016/j.fertnstert.2016.08.004.
26. Kang, E., Wu, J., Gutierrez, N. M., Koski,A., Tippner-
Hedges,R., Agaronyan,K., Platero-Luengo,A., Martinez-
Redondo,P., Ma,H., Lee,Y., Hayama,T., van Dyken,C.,
Wang, X., Luo, S., Ahmed, R., Li, Y., Ji, D., Kayali, R.,
Cinnioglu,C., etal. (2016) Mitochondrial replacement
in human oocytes carrying pathogenic mitochon-
drial DNA mutations, Nature, 540, 270-275, https://
doi.org/10.1038/nature20592.
27. Clark, K. A., Howe, D. K., Gafner, K., Kusuma, D.,
Ping,S., Estes, S., and Denver, D.R. (2012) Selfish lit-
tle circles: transmission bias and evolution of large
deletion-bearing mitochondrial DNA in Caenorhabdi-
tis briggsae nematodes, PLoS One, 7, e41433, https://
doi.org/10.1371/journal.pone.0041433.
28. Gitschlag, B.L., Pereira, C.V., Held, J. P., McCandlish,
D. M., and Patel, M. R. (2024) Multiple distinct evo-
lutionary mechanisms govern the dynamics of self-
ish mitochondrial genomes in Caenorhabditis ele-
gans, Nat. Commun., 15, 8237, https://doi.org/10.1038/
s41467-024-52596-9.
29. Ma, H., and O’Farrell, P. H. (2016) Selfish drive can
trump function when animal mitochondrial genomes
compete, Nat. Genet., 48, 798-802, https://doi.org/
10.1038/ng.3587.
30. Linnane, A. W., Saunders, G. W., Gingold, E. B., and
Lukins, H. B. (1968) The biogenesis of mitochondria.
V. Cytoplasmic inheritance of erythromycin resistance
in Saccharomyces cerevisiae, Proc. Natl. Acad. Sci.
USA, 59, 903-910, https://doi.org/10.1073/pnas.59.3.903.
31. Ephrussi, B., de Margerie-Hottinguer, H., and
Roman, H. (1955) Suppressiveness: a new factor in
the genetic determinism of the synthesis of respira-
tory enzymes in yeast, Proc. Natl. Acad. Sci. USA, 41,
1065-1071, https://doi.org/10.1073/pnas.41.12.1065.
32. De Zamaroczy, M., Marotta, R., Faugeron-Fonty, G.,
Goursot,R., Mangin,M., Baldacci,G., and Bernardi,G.
(1981) The origins of replication of the yeast mito-
chondrial genome and the phenomenon of suppres-
sivity, Nature, 292, 75-78, https://doi.org/10.1038/
292075a0.
33. Karavaeva, I. E., Golyshev, S. A., Smirnova, E. A.,
Sokolov, S. S., Severin, F. F., and Knorre, D. A. (2017)
Mitochondrial depolarization in yeast zygotes inhib-
its clonal expansion of selfish mtDNA, J. Cell Sci., 130,
1274-1284, https://doi.org/10.1242/jcs.197269.
34. Kashko, N.D., Muravyov,G., Karavaeva,I., Glagoleva,
E. S., Logacheva, M. D., Garushyants, S. K., and
Knorre, D.A. (2025) Inheritance bias of deletion-har-
bouring mtDNA in yeast: The role of copy number
and intracellular selection, PLoS Genet., 21, e1011737,
https://doi.org/10.1371/journal.pgen.1011737.
35. Taylor, D.R., Zeyl,C., and Cooke,E. (2002) Conflicting
levels of selection in the accumulation of mitochon-
drial defects in Saccharomyces cerevisiae, Proc. Natl.
Acad. Sci. USA, 99, 3690-3694, https://doi.org/10.1073/
pnas.072660299.
36. Røyrvik, E. C., and Johnston, I. G. (2020) MtDNA se-
quence features associated with “selfish genomes”
predict tissue-specific segregation and reversion,
Nucleic Acids Res., 48, 8290-8301, https://doi.org/
10.1093/nar/gkaa622.
37. Clayton, D.A. (1982) Replication of animal mitochon-
drial DNA, Cell, 28, 693-705, https://doi.org/10.1016/
0092-8674(82)90049-6.
38. Korhonen, J. A., Pham, X. H., Pellegrini, M., and
Falkenberg, M. (2004) Reconstitution of a minimal
mtDNA replisome in vitro, EMBO J., 23, 2423-2429,
https://doi.org/10.1038/sj.emboj.7600257.
39. Yasukawa, T., and Kang, D. (2018) An overview of
mammalian mitochondrial DNA replication mech-
anisms, J. Biochem., 164, 183-193, https://doi.org/
10.1093/jb/mvy058.
40. Falkenberg, M., and Gustafsson, C. M. (2020) Mam-
malian mitochondrial DNA replication and mecha-
nisms of deletion formation, Crit. Rev. Biochem. Mol.
Biol., 55, 509-524, https://doi.org/10.1080/10409238.
2020.1818684.
41. Kowald,A., Dawson,M., and Kirkwood, T.B.L. (2014)
Mitochondrial mutations and ageing: can mitochon-
drial deletion mutants accumulate via a size based
replication advantage? J. Theor. Biol., 340, 111-118,
https://doi.org/10.1016/j.jtbi.2013.09.009.
42. Diaz, F., Bayona-Bafaluy, M. P., Rana, M., Mora, M.,
Hao, H., and Moraes, C. T. (2002) Human mitochon-
drial DNA with large deletions repopulates organelles
faster than full-length genomes under relaxed copy
number control, Nucleic Acids Res., 30, 4626-4633,
https://doi.org/10.1093/nar/gkf602.
43. Seshadri, A., and Badrinarayanan, A. (2025) Exonu-
clease action of replicative polymerase gamma drives
MECHANISMS OF INTRACELLULAR SELECTION OF MITOCHONDRIAL DNA 1927
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
damage-induced mitochondrial DNA clearance,
EMBO Rep., 26, 1385-1405, https://doi.org/10.1038/
s44319-025-00380-1.
44. Freel, K. C., Friedrich, A., and Schacherer, J. (2015)
Mitochondrial genome evolution in yeasts: an all-
encompassing view, FEMS Yeast Res., 15, fov023,
https://doi.org/10.1093/femsyr/fov023.
45. Sato, M., and Sato, K. (2013) Maternal inheritance
of mitochondrial DNA by diverse mechanisms to
eliminate paternal mitochondrial DNA, Biochim. Bio-
phys. Acta, 1833, 1979-1984, https://doi.org/10.1016/
j.bbamcr.2013.03.010.
46. Chen, Z., Zhang, F., Lee, A., Yamine, M., Wang, Z.-H.,
Zhang, G., Combs, C., and Xu, H. (2025) Mitochon-
drial DNA removal is essential for sperm develop-
ment and activity, EMBO J., 44, 1749-1773, https://
doi.org/10.1038/s44318-025-00377-5.
47. Lim, Y., Rubio-Peña, K., Sobraske, P. J., Molina,
P. A., Brookes, P. S., Galy, V., and Nehrke, K. (2019)
Fndc-1 contributes to paternal mitochondria elim-
ination in C. elegans, Dev. Biol., 454, 15-20, https://
doi.org/10.1016/j.ydbio.2019.06.016.
48. Busch, K. B., Kowald, A., and Spelbrink, J. N. (2014)
Quality matters: how does mitochondrial network
dynamics and quality control impact on mtDNA in-
tegrity? Philos. Trans. R. Soc. Lond. B Biol. Sci., 369,
20130442, https://doi.org/10.1098/rstb.2013.0442.
49. Knorre, D. A. (2020) Intracellular quality control of
mitochondrial DNA: evidence and limitations, Phi-
los. Trans. R. Soc. Lond. B Biol. Sci., 375, 20190176,
https://doi.org/10.1098/rstb.2019.0176.
50. Wilkens, V., Kohl, W., and Busch, K. (2013) Restrict-
ed diffusion of OXPHOS complexes in dynamic mi-
tochondria delays their exchange between cristae
and engenders a transitory mosaic distribution,
J. Cell Sci., 126, 103-116, https://doi.org/10.1242/
jcs.108852.
51. Lieber,T., Jeedigunta, S.P., Palozzi, J.M., Lehmann,R.,
and Hurd, T. R. (2019) Mitochondrial fragmentation
drives selective removal of deleterious mtDNA in
the germline, Nature, 570, 380-384, https://doi.org/
10.1038/s41586-019-1213-4.
52. Milani,L., Ghiselli,F., Maurizii, M.G., Nuzhdin, S.V.,
and Passamonti, M. (2014) Paternally transmitted
mitochondria express a new gene of potential vi-
ral origin, Genome Biol. Evol., 6, 391-405, https://
doi.org/10.1093/gbe/evu021.
53. Selifanova, M., Demianchenko, O., Noskova, E.,
Pitikov, E., Skvortsov, D., Drozd, J., Vatolkina, N.,
Apel, P., Kolodyazhnaya, E., Ezhova, M. A., Tzetlin,
A.B., Neretina, T.V., and Knorre, D.A. (2023) ORFans
in mitochondrial genomes of marine polychaete Poly-
dora, Genome Biol. Evol., 15, evad219, https://doi.org/
10.1093/gbe/evad219.
54. Peeva, V., Blei, D., Trombly, G., Corsi, S., Szukszto,
M. J., Rebelo-Guiomar, P., Gammage, P. A., Kudin,
A.P., Becker,C., Altmüller,J., Minczuk,M., Zsurka,G.,
and Kunz, W. S. (2018) Linear mitochondrial DNA
is rapidly degraded by components of the repli-
cation machinery, Nat. Commun., 9, 1727, https://
doi.org/10.1038/s41467-018-04131-w.
55. Nissanka,N., Bacman, S.R., Plastini, M.J., and Moraes,
C. T. (2018) The mitochondrial DNA polymerase
gamma degrades linear DNA fragments precluding
the formation of deletions, Nat. Commun., 9, 2491,
https://doi.org/10.1038/s41467-018-04895-1.
56. DeLuca, S. Z., and O’Farrell, P. H. (2012) Barri-
ers to male transmission of mitochondrial DNA in
sperm development, Dev. Cell, 22, 660-668, https://
doi.org/10.1016/j.devcel.2011.12.021.
57. Wolters, J.F., LaBella, A.L., Opulente, D.A., Rokas,A.,
and Hittinger, C. T. (2023) Mitochondrial genome di-
versity across the subphylum Saccharomycotina,
Front. Microbiol., 14, 1268944, https://doi.org/10.3389/
fmicb.2023.1268944.
58. Megarioti, A. H., and Kouvelis, V. N. (2020) The co-
evolution of fungal mitochondrial introns and their
homing endonucleases (GIY-YIG and LAGLIDADG),
Genome Biol. Evol., 12, 1337-1354, https://doi.org/
10.1093/gbe/evaa126.
59. De Grey, A. D. (1997) A proposed refinement of the
mitochondrial free radical theory of aging, Bioessays,
19, 161-166, https://doi.org/10.1002/bies.950190211.
60. Chernyak, B. V., Izyumov, D. S., Lyamzaev, K. G.,
Pashkovskaya, A. A., Pletjushkina, O. Y., Antonenko,
Y.N., Sakharov, D.V., Wirtz, K. W.A., and Skulachev,
V. P. (2006) Production of reactive oxygen spe-
cies in mitochondria of HeLa cells under oxida-
tive stress, Biochim. Biophys. Acta, 1757, 525-534,
https://doi.org/10.1016/j.bbabio.2006.02.019.
61. Skulachev, V. P. (2002) Programmed death phenome-
na: from organelle to organism, Ann. NY Acad. Sci.,
959, 214-237, https://doi.org/10.1111/j.1749-6632.2002.
tb02095.x.
62. Degtyareva, N. P., Saini, N., Sterling, J. F., Placentra,
V. C., Klimczak, L. J., Gordenin, D. A., and Doetsch,
P. W. (2019) Mutational signatures of redox stress
in yeast single-strand DNA and of aging in human
mitochondrial DNA share a common feature, PLoS
Biol., 17, e3000263, https://doi.org/10.1371/journal.
pbio.3000263.
63. Iliushchenko, D., Efimenko, B., Mikhailova, A. G.,
Shamanskiy, V., Saparbaev, M. K., Matkarimov, B. T.,
Mazunin, I., Voronka, A., Knorre, D., Kunz, W. S.,
Kapranov, P., Denisov, S., Fellay, J., Khrapko, K.,
Gunbin, K., and Popadin, K. (2025) Deciphering the
foundations of mitochondrial mutational spectra:
replication-driven and damage-induced signatures
across chordate classes, Mol. Biol. Evol., 42, msae261,
https://doi.org/10.1093/molbev/msae261.
64. Shokolenko, I., Venediktova, N., Bochkareva, A.,
Wilson, G. L., and Alexeyev, M. F. (2009) Oxidative
MURAVYOV, KNORRE1928
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
stress induces degradation of mitochondrial DNA,
Nucleic Acids Res., 37, 2539-2548, https://doi.org/
10.1093/nar/gkp100.
65. Wiedemann, N., and Pfanner, N. (2017) Mitochon-
drial machineries for protein import and assembly,
Annu. Rev. Biochem., 86, 685-714, https://doi.org/
10.1146/annurev-biochem-060815-014352.
66. Quintana-Cabrera, R., and Scorrano, L. (2023) Deter-
minants and outcomes of mitochondrial dynamics,
Mol. Cell, 83, 857-876, https://doi.org/10.1016/j.molcel.
2023.02.012.
67. Yang, L., Long, Q., Liu, J., Tang, H., Li, Y., Bao, F.,
Qin, D., Pei, D., and Liu, X. (2015) Mitochondrial fu-
sion provides an “initial metabolic complementation”
controlled by mtDNA, Cell. Mol. Life Sci., 72, 2585-
2598, https://doi.org/10.1007/s00018-015-1863-9.
68. Edwards, D. M., Røyrvik, E. C., Chustecki, J. M.,
Giannakis,K., Glastad, R.C., Radzvilavicius, A.L., and
Johnston, I. G. (2021) Avoiding organelle mutational
meltdown across eukaryotes with or without a
germline bottleneck, PLoS Biol., 19, e3001153,
https://doi.org/10.1371/journal.pbio.3001153.
69. Palozzi, J. M., Jeedigunta, S. P., Minenkova, A. V.,
Monteiro, V.L., Thompson, Z.S., Lieber,T., and Hurd,
T.R. (2022) Mitochondrial DNA quality control in the
female germline requires a unique programmed
mitophagy, Cell Metab., 34, 1809-1823.e6, https://
doi.org/10.1016/j.cmet.2022.10.005.
70. Birky, C. W., Jr. (2008) Uniparental inheritance of
organelle genes, Curr. Biol., 18, R692-R695, https://
doi.org/10.1016/j.cub.2008.06.049.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in published
maps and institutional affiliations. AI tools may have
been used in the translation or editing of this article.