ISSN 0006-2979, Biochemistry (Moscow), 2023, Vol. 88, Nos. 12-13, pp. 1997-2006 © Pleiades Publishing, Ltd., 2023.
Published in Russian in Biokhimiya, 2023, Vol. 88, No. 12, pp. 2387-2398.
1997
REVIEW
Role of Mitochondrial DNA in Yeast Replicative Aging
Aglaia V. Azbarova
1,2
and Dmitry A. Knorre
1,a
*
1
Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, 119991 Moscow, Russia
2
Faculty of Bioengineering and Bioinformatics, Lomonosov Moscow State University, 119991 Moscow, Russia
a
e-mail: knorre@belozersky.msu.ru
Received August 23, 2023
Revised November 13, 2023
Accepted November 15, 2023
AbstractDespite the diverse manifestations of aging across different species, some common aging features and underlying
mechanisms are shared. In particular, mitochondria appear to be among the most vulnerable systems in both metazoa and
fungi. In this review, we discuss how mitochondrial dysfunction is related to replicative aging in the simplest eukaryotic
model, the baker’s yeast Saccharomyces cerevisiae. We discuss a chain of events that starts from asymmetric distribution
of mitochondria between mother and daughter cells. With age, yeast mother cells start to experience a decrease in mito-
chondrial transmembrane potential and, consequently, a decrease in mitochondrial protein import efficiency. This induces
mitochondrial protein precursors in the cytoplasm, the loss of mitochondrial DNA (mtDNA), and at the later stages
cell death. Interestingly, yeast strains without mtDNA can have either increased or decreased lifespan compared to the
parental strains with mtDNA. The direction of the effect depends on their ability to activate compensatory mechanisms
preventing or mitigating negative consequences of mitochondrial dysfunction. The central role of mitochondria in yeast
aging and death indicates that it is one of the most complex and, therefore, deregulation-prone systems in eukaryotic cells.
DOI: 10.1134/S0006297923120040
Keywords: yeast, development, aging, mitochondrial DNA, mitochondrial dysfunction
Abbreviations: ERC, extrachromosomal rDNA circle; mtDNA, mitochondrial DNA; OxPhos, oxidative phosphorylation;
RLS,replicative lifespan.
* To whom correspondence should be addressed.
INTRODUCTION
Aging is the process characterized by a decrease in
fertility and an increase in mortality over time [1], a phe-
nomenon that inevitably emerges in the process of living
systems evolution. This is attributed to the fact that nat-
ural selection diminishes in the post-production period,
as it primarily affects genes and traits associated with
reproduction. Consequently, as an organism generates
more offspring, the impact of natural selection on traits
(and the corresponding genes) that emerge after repro-
duction begins to wane [2, 3]. This is applicable to mi-
croorganisms as well, including bacteria and unicellular
fungi. Bacterial and yeast species that divide asymmetri-
cally– those that produce a mother cell and a daughter
cell of differing sizes– usually exhibit aging [4-7].
The baker’s yeast, Saccharomyces cerevisiae, is the
most extensively studied unicellular organism where ag-
ing has been described [8-10]. Investigations into yeast
aging mechanisms have expanded our understanding of
the fundamental principles of aging. Moreover, yeast
aging studies have facilitated the identification of targe-
table systems that, upon intervention, can lead to lifes-
pan extension. A significant advantage of baker’s yeast
over other model organisms is the high proliferation
rate of yeast cells. Additionally, it is relatively easy to
produce genetically stable mutant yeast cell lines, which
are instrumental in identifying gene and protein func-
tions [11]. Presently, several collections of yeast mutant
strains are available, these collections have been utilized
to conduct genetic screenings aimed at discov ering lon-
gevity genes [12, 13]. For example, caloric restriction
dietary mimetics, including resveratrol, were identified
using the yeast aging model [14]. The efficacy of these
mimetics was later confirmed in animal studies [15].
A screening of mutations altering S. cerevisiae lifespan
revealed 238 genes whose deletion increases it [16].
These genes comprised orthologs of genes whose mu-
AZBAROVA, KNORRE1998
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
Fig. 1. Two scenarios of replicative aging in yeast. Approximate replicative ages are based on the literature cited in the text.
tations altered the lifespan of the nematodes Caenor-
habditis elegans [16]. The results of this screening were
enriched with genes encoding tricarboxylic acid cycle
enzymes and proteins required for mitochondrial pro-
tein translation, including mitochondrial ribosome pro-
tein components [16]. In a separate study, the authors
demonstrated that deleting yeast orthologs of nematode
aging genes increased, rather than decreased, the lifes-
pan of yeast [17]. Taken together, these results suggest
a conservation of aging mechanisms among eukaryotes.
The aging process of baker’s yeast S. cerevisiae is
typically defined not by chronological time but by the
number of daughter cells a mother cell generates. There-
fore, it is usually referred to as replicative aging [18]
(Fig.1). Accordingly, the number of daughter cells that
a mother cell can produce before its death is termed the
“replicative lifespan”(RLS). A single yeast mother cell
can only generate a limited number of daughter cells,
usually ranging between 15 and 25 [19]. Meanwhile, the
first aging manifestation in the yeast mother cell, such as
changes in the expression of stress response genes, oc-
curs after the formation of the first few buds [20].
As yeast cells age, they undergo deregulation in
gene expression and protein biogenesis. This process
results in disproportion of subunits in protein com-
plexes, such as vacuolar ATPase, in replicatively old
yeast cells [21]. In recent years, advancements in mi-
crofluidics techniques have facilitated the examination
of replicative aging at the individual cell level [22,23].
The aging trajectory of each individual cell is predeter-
mined during the initial cell cycles, leading to one of
two scenarios. Though both scenarios inevitably result
in cell death, the causes of cell viability loss differ be-
tween them [24].
The first scenario is characterized by the loss of
chromatin silencing. This is evidenced by the elevated
expression of rDNA-GFP, a GFP gene integrated into the
non-transcribed rRNA spacer region. Yeast cells aging
through this first scenario maintain a relatively consis-
tent cell cycle duration throughout their lifetime up to
their late life. However, during the last few cell cycles,
which precedes death, such yeast cells form elongated
daughter cells, as depicted in Fig.1.
The second scenario correlates with an increased cell
cycle length during early life stages and the production
of rounded daughter cells. Moreover, in yeast cells aged
under the second scenario, a deficiency of heme is ob-
served [25]. Given that the final stages of heme synthesis
occur within mitochondria, and a lack of heme leads to
mitochondrial dysfunction [26], the aging via this sce-
nario is linked to mitochondrial dysfunction. The death
of yeast cells is preceded by a decrease in Leu4-GFP
levels. Leu4p, or alpha-isopropylmalate synthase, is a
crucial part of the leucine biosynthesis pathway; hence,
a reduction in Leu4p triggers leucine deficiency in older
cells. The second aging scenario is linked to increased
expression of HSP104, a gene that encodes the cytosolic
heat shock protein, Hsp104p [27]. Hsp104p binds and
temporarily stores mitochondrial protein precursors in
the cytoplasm [28], whereas increased concentration of
Hsp104p is a hallmark of the accumulation of unfolded
proteins [29]. Recently, the information about the de-
scribed above scenarios inspired the creation of a genetic
oscillator designed to make yeast cells alternate between
these two aging trajectories, thereby boosting their lifes-
pan by 82% [30].
Replicative aging in yeast can be considered as a
deterministic developmental program that progresses
ROLE OF MITOCHONDRIA IN YEAST AGING 1999
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
along one of at least two possible trajectories (Fig.1).
Deregulation of certain cellular systems restrict yeast
lifespan while concurrently the same systems contrib-
ute to the longevity of multicellular organisms [17]. One
of these yeast aging trajectories is likely associated with
mitochondrial dysfunction. In this review, we discuss
the causal relationship between mitochondrial dysfunc-
tion and other replicative aging manifestations in yeast.
MITOCHONDRIAL DYSFUNCTION
IN YEAST REPLICATIVE AGING
The division of baker’s yeast cells is highly asym-
metrical, resulting in mother and daughter cells of dif-
fering sizes and varying concentrations of certain pro-
teins [31]. This protein content asymmetry is primarily
due to diffusion limitations, which restrict the transport
of large cellular structures, such as large protein aggre-
gates and the organelles harboring them, between the
cells [32]. Moreover, the protein aggregates and cell or-
ganelles are actively transported and can be selectively
anchored in either the mother or daughter cell cortex
[33,34]. The transport of mitochondria from mother to
daughter cells is facilitated by the actin cytoskeleton and
the myosin protein, Myo2p [35]. These cytoskeleton
components anchor the mitochondria with the mean of
mitochondrial outer membrane protein, Mmr1p [36].
Concurrently, the Num1p and Mfb1p proteins ensure
that a portion of the mitochondria remain at the pole of
the mother cell [37, 38]. This active transport and se-
lective retention of mitochondria make it possible to
distribute them between the mother and daughter cells.
This process primarily sends the most functional mito-
chondria to the daughter cell, while ensuring that the
mother cell retains a portion of functional mitochon-
dria, along with the damaged ones [39].
The asymmetric distribution of mitochondria be-
tween mother and daughter cells results in accumula-
tion of mitochondria with oxidized matrix molecules in
the mother cell compared to the mitochondria passed
on to the daughter cell. This is evidenced by the redox-
sensitive fluorescent proteins targeted towards the mito-
chondrial matrix [40, 41]. If the mitochondria are het-
erogeneous, the daughter cell is more likely to inherit
functional mitochondria, while the mother cell retains
the non-functional ones [39]. However, the mechanisms
behind the selectivity of mitochondrial anchoring and
transport remain unclear.
Mitochondrial dysfunction begins to manifest in the
mother cell after several consecutive rounds of asymmet-
rical divisions. Specifically, during the first ten cell cy-
cles, transmembrane potential of the mother cell mito-
chondria decreases [42], and the mitochondrial network
collapses [42, 43]. Factors such as protonophores [44] and
oxidative stress [45, 46] can induce mitochondrial frag-
mentation. This leads to the assumption that age-related
mitochondrial fragmentation is induced by age-depen-
dent mitochondrial depolarization. Studies have shown
that the deletion of DNM1, a gene that encodes the
Dnm1p protein, which mediates mitochondrial fission,
extends the replicative lifespan of yeast [47]. Thus, it can
be concluded that mitochondrial fragmentation plays
acausative role in detrimental processes in aging cells,
and it cannot be considered as a mere side effect of the
depolarization.
Furthermore, the age-dependent decrease in mi-
tochondrial transmembrane potential(ΔΨ) contributes
to a diminished efficiency of protein import into mito-
chondria [48]. While protein import through the outer
membrane does not depend on the ΔΨ, protein trans-
location through the inner membrane stops when ΔΨ is
dissipated [48].
The inhibition of protein import into mitochon-
dria has two unfavorable outcomes for the cell. Firstly,
it can cause a deficiency in proteins vital for mitochon-
drial DNA(mtDNA) replication, as well as mitochon-
drial transcription and translation. It is noteworthy that
yeast aging correlates with a drop in Mip1p levels, a mi-
tochondrial DNA polymerase, an enzyme indispensable
for mtDNA replication [49]. Interestingly, increased ex-
pression of the TOM70 gene, encoding a component of
the translocase of the outer mitochondrial membrane
(TOM) complex, may partially offset this effect by ac-
tivating protein import [49]. This finding suggests that
compromised protein import is a contributing factor to
replicative aging in yeast. Eventually, the inhibition of
mitochondrial protein import culminates in the loss of
mtDNA in replicatively aged cells [50] (Fig.1).
Secondly, the inhibition of protein import into
mitochondria results in the accumulation of the mito-
chondrial protein precursors in the cytoplasm. These
precursors are known to be toxic to the cell. For exam-
ple, the expression of “clogger” proteins, which block
the mitochondrial protein import system, inhibits the
growth of yeast cells [28]. Although S. cerevisiae cells can
survive without mtDNA and oxidative phosphorylation
(OxPhos) by relying on glycolysis, the deletion of TIM-
(mitochondrial inner membrane translocase) or TOM-
complex genes are usually lethal [51].
We propose that these two unfavorable phenome-
na– a reduction in the concentration of essential pro-
teins in the mitochondria and proteotoxic stress induced
by mitochondrial protein precursors in the cytosol
contribute to an extended cell cycle duration and, even-
tually, cell death. To our knowledge, there are no direct
experimental assessments of the mitochondrial contri-
bution to these two yeast aging scenarios. However, the
second scenario is associated with mitochondrial net-
work fragmentation [25] and a decrease in mitochon-
drial membrane potential, as evidenced by a reduction
in DiOC6 staining of aging cells [27]. This implies that
AZBAROVA, KNORRE2000
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
mitochondrial dysfunction is more likely associated with
the second scenario.
BIDIRECTIONAL EFFECT
OF MITOCHONDRIAL DNA DEPLETION
ON YEAST REPLICATIVE LIFESPAN
The mitochondrion, a semi-autonomous cellular
organelle found in most eukaryotic species, retains its
own DNA. This DNA encodes components of mito-
chondrial translation systems, as well as some respirato-
ry chain proteins [52]. Therefore, mutations in mtDNA
or its complete disappearance (denoted as rho
0
) result in
the loss of mitochondrial capacity to perform OxPhos.
In certain laboratory strains, such as YPK9, the
elimination of mitochondrial DNA (rho
0
mutation) trig-
gers an increase in RLS [53-57]. However, this effect is
not universal; in some strains, the rho
0
mutations have
the opposite effect [58], or no effect [53]. An example
of this can be observed in the W303-1A strain, which
is frequently used for studying OxPhos in yeast [53].
Furthermore, removal of mtDNA in strains based on
BY4742– the genetic background for yeast mutant col-
lections– does not induce an increase in RLS [59,60].
We propose that the contrasting effects of the rho
0
mutation can be attributed to the varying responses
of different strains to mitochondrial dysfunction [53].
Thecomplete depletion of mitochondrial DNA in yeasts
alters both gene expression levels and protein concen-
trations [61-63]. Notably, rho
0
mutation activates retro-
grade signaling cascade, which in turn activates genes
that encode glycolysis and the glyoxylate cycle enzymes
[61, 64]. This process enables cells to adapt their met-
abolic functions to conditions where OxPhos is unfeasi-
ble and reactions linked to respiration, such as succinate
oxidation, are blocked.
At the same time, laboratory strains can react differ-
ently to the loss of mtDNA. For instance, the W303-1A
strain, following the complete depletion of mtDNA,
does not increase the expression of the CIT2 gene, en-
coding citrate-synthase localized in peroxisomes [65].
Conversely, YPK9 rho
0
cells, which exhibit a high RLS,
upregulate this gene [53]. Furthermore, different labora-
tory strains can vary in rDNA array length [66], which
is positively correlated with yeast RLS [67]. This cor-
relation is attributed to the regulation of the histone
deacetylase gene SIR2. Sir2p limits the formation of
ex trachromosomal rDNA circles (ERCs), identified as
one of the factors contributing to aging. Concur rently,
upstream activation factors (UAFs) bind to chromo-
somal rDNA and, when not bound, inhibit Sir2p. As
a result, strains with a shorter rDNA array length pos-
sess a higher number of free UAFs, which inhibit Sir2p,
induce ERC formation. ERCs accumulate in older
cells, disrupting chromatin silencing and increasing
the probability of cell death with age, leading to de-
creased RLS [67].
We suggest that the lifespan of yeast strains with
shorter rDNA array lengths is primarily constrained by
the activity of SIR2, the formation of ERCs, and the
prevalence of aging through the second scenario. There-
fore, mitochondrial dysfunction does not pose a limita-
tion to the RLS in these strains. Conversely, in strains
with a longer rDNA array, aging predominantly proceeds
via the second scenario, which is linked to mitochon-
drial function. In this context, mitochondrial dysfunc-
tion accelerates the aging process and shortens RLS.
The strains’ variability in the consequences of
mtDNA depletion could also be explained by the dy-
namics of the cell adaptation to mtDNA loss, a pro-
cess requiring tenth of generations [68]. Notably, new-
ly formed rho
0
cells exhibit a reduced lifespan, while
adapted to the absence of mtDNA rho
0
cells demon-
strate an extended lifespan in comparison to the paren-
tal strain[68]. It could be hypothesized that this adap-
tation process unfolds at different speeds and follows
distinct pathways in various laboratory strains. This, in
turn, contributes to the observed variability in the im-
pact of the rho
0
mutation on RLS.
The loss of mtDNA could have differential effects
on the two aging scenarios previously outlined. While
itmay accelerate aging in one scenario, it could poten-
tially inhibit the aging process in the other. For instance,
in the context of the second scenario, the rho
0
mutation
could help yeast cells in pre-adapting to the detrimen-
tal processes of aging, such as disruption of heme bio-
genesis. The variable proportions of cells in laboratory
strains, which are presumed to age following one sce-
nario or another, may contribute to the uncertainty re-
garding the effect of mtDNA loss on RLS.
THE ROLE OF NUCLEAR-ENCODED
MITOCHONDRIAL PROTEINS
IN YEAST REPLICATIVE LONGEVITY
The yeast mitochondrial proteome comprises ~900
proteins, the vast majority of these proteins are encoded
within the nucleus and are imported into the mitochon-
dria from the cytoplasm [69]. Therefore, mitochondrial
dysfunction can arise not only from the loss of mtDNA
but also from the inactivation of nuclear genes that en-
code these proteins. In a study involving whole-genome
sequencing of an entire yeast knockout collection, it was
revealed that the deletion of 129 out of roughly 5000
nonessential genes leads to the loss of mtDNA [70].
Intriguingly, several of these genes were also identified in
a screen designed to uncover genes whose deletion results
in an increased RLS of yeast [16]. This overlap is not like-
ly to be random and suggests an interconnection between
mitochondrial DNA maintenance and RLS (Fig.2).
ROLE OF MITOCHONDRIA IN YEAST AGING 2001
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
Fig. 2. The Venn diagram illustrates the intersection of two gene sets in yeast S.cerevisiae. One set includes genes whose deletion leads to mtDNA
loss (Puddu etal. [70]), the other set comprises genes whose deletion results in an increased replicative lifespan (McCormick etal. [16]). The prob-
ability of a random overlap of 17 or more genes between these two samples is less than 0.1%.
However, it is important to note that not all genes,
whose deletion results in mtDNA loss, necessarily lead
to an increased lifespan. Among the genes encoding for
mitochondrial proteins, whose deletion concurrently
enhances yeast RLS, one category stands out: the genes
encoding components of the mitochondrial ribosome.
The Venn diagram depicted in Fig. 2 illustrates that
the deletion of MRPL33, MRPL40, MRPL49, IMG1,
and IMG2 genes, which encode mitochondrial ribo-
somal proteins, leads to an increase in RLS. Further-
more, the deletion of another ribosomal protein gene,
MRPL25, renders cell resistance to pro-oxidants and
increases RLS, with the maximum lifespan of the de-
letion strain reaching 60 [71]. Additionally, the deletion
of the SOV1 gene, which is essential for mitochondrial
ribosome assembly, has also been observed to increase
yeast RLS [72].
The mitochondrial ribosome cannot be assembled
irrespective of the reasons causing mtDNA loss because
mtDNA encodes mitochondrial rRNA subunits [73],
making it impossible for mitochondrial ribosomes to
assemble in the absence of mtDNA. Thus, what makes
protein components of mitochondrial ribosomes special
remains uncertain. However, the import mechanism of
mitochondrial ribosome proteins (MRPs) diverges from
that of other mitochondrial proteins. A significant num-
ber of MRPs lack a distinctive mitochondrial address
sequence, a phenomenon attributed to the conservative
nature of their function, including the N-terminal re-
gion of these proteins [74]. Thus, deleting MRP genes
could affect the mitochondrial import machinery in a
unique way, distinct from the effect caused by the de-
letion of other mitochondrial proteins harboring a con-
ventional N-end mitochondrial address. For instance,
the deletion of MRP genes could potentially unburden
a specific set of accessory TOM proteins. An alternative
explanation for the unique phenotype associated with
MRP deletion could be that the absence of rRNA in the
mitochondria of rho
0
cells can increase toxicity of some
other MRP proteins, consequently accelerating the aging
process in yeast mother cells. Caballero etal. suggested
that the disappearance of certain mitochondrial trans-
lation machinery components could result in acquiring
novel signaling functions by the remaining components.
These functions could be related to chromatin silencing,
potentially leading to an extended lifespan [72]. This
hypothesis aligns with observations made regarding the
Cbs2p protein, which possesses dual cytosol and mito-
chondrial localization. This protein appears to be es-
sential for the life-extension effect of the mitochondrial
translation activator gene, CBS1, deletion [72].
While the deletion of certain nuclear-encoded mi-
tochondrial protein genes increases RLS, the deletion
of some other genes substantially diminishes it. For in-
stance, deleting the MIP1 gene, which encodes mito-
chondrial DNA polymerase, results in a reduced yeast
lifespan [72]. MtDNA replication is impossible without
Mip1p. Similarly, a decrease in RLS is also noted upon
the deletion of the COX4, COX7 genes [72, 75], and cyto-
chrome c heme lyase gene CYC3 [76]. COX genes encode
complexIV subunits of the respiratory chain. Their ab-
sence disrupts OxPhos, thereby rendering the utilization
AZBAROVA, KNORRE2002
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
of non-fermentable carbon sources impossible [77].
Thedeletion of TOM70, a gene that encodes a compo-
nent of the TOM complex, reduces RLS [49]. Similarly,
RLS is decreased upon the deletion of PIM1, a gene en-
coding mitochondrial matrix protease [78]. We propose
that mitochondrial dysfunction may accelerate aging by
inflating the number of cells that age by the second sce-
nario. Meanwhile, the second scenario is characterized
by a shortened yeast RLS, in comparison to the first
scenario, which is associated with deregulation of chro-
matin silencing [25].
Mitochondrial dysfunction can be induced by mi-
tochondrial inhibitors such as antimycin A or oligomy-
cin D, as well as protonophores, which dissipate mi-
tochondrial membrane potential. These interventions,
similar to the rho
0
mutation, exhibit bidirectional effects
on RLS that can vary based on other factors [76,79].
For instance, on the one hand, the “weak” protono-
phore dinitrophenol, when applied at a concentration
of 5 mM, has been shown to slightly increase RLS [79].
This effect was likely mediated through the activation
of the Rtg mitochondria-to-nucleus signaling pathway.
Meanwhile, it has been shown that protonophores ac-
tivate the PDR5 gene, a known target of the Rtg path-
way[80], and the Rtg pathway can trigger adaptive re-
sponses that extend the RLS of yeast [81].
On the other hand, the “strong” protonophore cy-
anide-p-trifluoromethoxy phenylhydrazone, FCCP, re-
sults in a significant decrease in RLS of BY4741 cells[82].
How can protonophores contribute to this decrease in
RLS? The diminished mitochondrial transmembrane
potential, induced by protonophores, can exhibit selec-
tive toxicity towards cells with limited capacity to gen-
erate ΔΨ. Indeed, the protonophores are particularly
toxic to rho
0
cells, where the ΔΨ is generated by the ad-
enine nucleotide translocator, which appears to be not
very effective [83, 84]. As the replicative age of a yeast
cell advances, the likelihood of total loss of mtDNA and
respiratory activity elevates. Therefore, high concentra-
tions of protonophores might be selectively detrimental
to older cells relative to younger ones, thereby reducing
RLS [83,84].
CONCLUSIONS
We hypothesize the following chain of events in
aging yeast cells. The asymmetric distribution of mi-
tochondria results in the accumulation of dysfunction-
al mitochondria in the mother cell. As a result, mito-
chondrial dysfunction emerges during the early stages of
replicative aging. With age, the yeast cells experience a
deficiency in nucleus-encoded mitochondrial proteins.
This deficiency subsequently induces mutations or a
complete loss of mtDNA. At the same time, toxic mito-
chondrial protein precursors are accumulated within the
cytoplasm. Furthermore, as yeast cells age, they increas-
ingly depend on oxidative metabolism, in contrast to
younger cells that rely more heavily on glycolysis. Yeast
replicative aging is associated with upregulated pentose
phosphate pathway, tricarboxylic acid cycle and glycer-
ol biosynthesis pathways, alongside with a decrease in
cellular ATP concentration [85]. Thus, while mitochon-
drial dysfunction increases with age, the sequential met-
abolic adjustments in the cell render it increasingly re-
liant on OxPhos. This seeming paradox may contribute
to a diminished growth rate of the old cells, ultimately
leading to cell death.
The notion that mitochondrial dysfunction decreas-
es viability of older cells implies that cells with initially
compromised OxPhos may have a shortened lifespan.
Furthermore, the way how yeast cells adapt to mito-
chondrial dysfunction, such as activating mitochondria-
to-nucleus pathway targets, seems to be a crucial fac-
tor for yeast longevity. Taken together, the yeast aging
model reveals a vulnerable system within eukaryotic
cells, which appears susceptible to deregulation. This
system, the coordination mechanism between the mito-
chondria and the nucleus, is an intricate machinery that
has evolved over a billion years since the symbiogenesis
of archaea and mitochondrial ancestors.
Contributions. A.A. and D.K. contributed to the
analysis of the literature and the writing of the text.
D.K. prepared the illustrations.
Funding. This work was supported by the Russian
Science Foundation, grant no.22-14-00108.
Ethics declarations. The authors declare no conflict
of interest. This article does not describe any research
involving humans or animals as subjects.
REFERENCES
1. Kirkwood, T. B., and Austad, S. N. (2000) Why do we
age? Nature, 408, 233-238, doi:10.1038/35041682.
2. Flatt, T., and Partridge, L. (2018) Horizons in the evo-
lution of aging, BMC Biol., 16, 93, doi: 10.1186/s12915-
018-0562-z.
3. Wensink, M. J., and Cohen, A. A. (2021) Thedanaid the-
ory of aging, Front. Cell Dev. Biol., 9, 671208, doi:10.3389/
fcell.2021.671208.
4. Mortimer, R. K., and Johnston, J. R. (1959) Life span
of individual yeast cells, Nature, 183, 1751-1752, doi:
10.1038/1831751a0.
5. Barker, M. G., and Walmsley, R. M. (1999) Replicative
ageing in the fission yeast Schizosaccharomyces pombe,
Yeast, 15, 1511-1518, doi: 10.1002/(SICI)1097-0061
(199910)15:14<1511::AID-YEA482>3.0.CO;2-Y.
6. Ackermann, M., Stearns, S. C., and Jenal, U. (2003)
Senescence in a bacterium with asymmetric division,
Science, 300, 1920, doi:10.1126/science.1083532.
ROLE OF MITOCHONDRIA IN YEAST AGING 2003
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
7. Bouklas, T., and Fries, B. C. (2015) Aging as an emer-
gent factor that contributes to phenotypic variation in
Cryptococcus neoformans, Fungal Genet. Biol., 78, 59-64,
doi:10.1016/j.fgb.2014.10.004.
8. Laun, P., Bruschi, C. V., Dickinson, J. R., Rinnerthal-
er,M., Heeren, G., Schwimbersky, R., Rid, R., and Bre-
itenbach, M. (2007) Yeast mother cell-specific ageing, ge-
netic (in)stability, and the somatic mutation theory of age-
ing, Nucleic Acids Res., 35, 7514-7526, doi:10.1093/nar/
gkm919.
9. Barros, M. H., da Cunha, F. M., Oliveira, G. A., Tahara,
E.B., and Kowaltowski, A.J. (2010) Yeast as a model to
study mitochondrial mechanisms in ageing, Mech. Ageing
Dev., 131, 494-502, doi:10.1016/j.mad.2010.04.008.
10. Jazwinski, S. M., Jiang, J. C., and Kim, S. (2018) Ad-
aptation to metabolic dysfunction during aging: making
the best of a bad situation, Exp. Gerontol., 107, 87-90,
doi:10.1016/j.exger.2017.07.013.
11. Botstein, D., and Fink, G. R. (2011) Yeast: an experi-
mental organism for 21st Century biology, Genetics, 189,
695-704, doi:10.1534/genetics.111.130765.
12. Campos, S. E., and DeLuna, A. (2019) Functional ge-
nomics of dietary restriction and longevity in yeast, Mech.
Ageing Dev., 179, 36-43, doi:10.1016/j.mad.2019.02.003.
13. Novarina, D., Janssens, G. E., Bokern, K., Schut, T.,
van Oerle, N. C., Kazemier, H. G., Veenhoff, L. M., and
Chang, M. (2020) Agenome-wide screen identifies genes
that suppress the accumulation of spontaneous mutations
in young and aged yeast cells, Aging Cell, 19, e13084,
doi:10.1111/acel.13084.
14. Howitz, K. T., Bitterman, K. J., Cohen, H. Y., Lamming,
D. W., Lavu, S., Wood, J. G., Zipkin, R. E., Chung, P.,
Kisielewski, A., Zhang, L.-L., Scherer, B., and Sinclair,
D. A. (2003) Small molecule activators of sirtuins extend
Saccharomyces cerevisiae lifespan, Nature, 425, 191-196,
doi:10.1038/nature01960.
15. Okamoto, N., Sato, Y., Kawagoe, Y., Shimizu, T., and
Kawamura, K. (2022) Short-term resveratrol treatment
restored the quality of oocytes in aging mice, Aging, 14,
5628-5640, doi:10.18632/aging.204157.
16. McCormick, M. A., Delaney, J. R., Tsuchiya, M.,
Tsuchiyama, S., Shemorry, A., Sim, S., Chou, A. C.-Z.,
Ahmed,U., Carr, D., Murakami, C. J., Schleit, J., Sut-
phin, G. L., Wasko, B. M., Bennett, C. F., Wang, A. M.,
Olsen, B., Beyer, R. P., Bammler, T. K., Prunkard, D.,
Johnson, S. C., Pennypacker, J. K., An, E., Anies, A.,
Castanza, A. S., Choi, E., Dang, N., Enerio, S., Fletch-
er,M., Fox, L., Goswami, S., Higgins, S. A., Holmberg,
M. A., Hu, D., Hui, J., Jelic, M., Jeong, K.-S., John-
ston, E., Kerr, E. O., Kim, J., Kim, D., Kirkland, K.,
Klum, S., Kotireddy, S., Liao, E., Lim, M., Lin, M. S.,
Lo, W. C., Lockshon, D., Miller, H. A., Moller, R. M.,
Muller, B., Oakes, J., Pak, D. N., Peng, Z. J., Pham,
K. M., Pollard, T. G., Pradeep, P., Pruett, D., Rai, D.,
Robison, B., Rodriguez, A. A., Ros, B., Sage, M., Singh,
M. K., Smith, E. D., Snead, K., Solanky, A., Spec-
tor, B. L., Steffen, K. K., Tchao, B. N., Ting, M. K.,
Vander Wende, H., Wang, D., Welton, K. L., Westman,
E. A., Brem, R. B., Liu, X. G., Suh, Y., Zhou, Z., Kae-
berlein,M., and Kennedy, B. K. (2015) Acomprehensive
analysis of replicative lifespan in 4,698 single-gene dele-
tion strains uncovers conserved mechanisms of aging, Cell.
Metab., 22, 895-906, doi:10.1016/j.cmet.2015.09.008.
17. Smith, E. D., Tsuchiya, M., Fox, L. A., Dang, N.,
Hu,D., Kerr, E. O., Johnston, E. D., Tchao, B. N., Pak,
D. N., Welton, K. L., Promislow, D. E. L., Thomas, J.H.,
Kaeberlein, M., and Kennedy, B. K. (2008) Quantitative
evidence for conserved longevity pathways between di-
vergent eukaryotic species, Genome Res., 18, 564-570,
doi:10.1101/gr.074724.107.
18. He, C., Zhou, C., and Kennedy, B. K. (2018) Theyeast
replicative aging model, Biochim. Biophys. Acta Mol. Basis
Dis., 1864, 2690-2696, doi:10.1016/j.bbadis.2018.02.023.
19. Ashrafi, K., Sinclair, D., Gordon, J. I., and Guarente, L.
(1999) Passage through stationary phase advances repli-
cative aging in Saccharomyces cerevisiae, Proc. Natl. Acad.
Sci. USA, 96, 9100-9105, doi:10.1073/pnas.96.16.9100.
20. Knorre, D. A., Azbarova, A. V., Galkina, K. V., Feniouk,
B. A., and Severin, F. F. (2018) Replicative aging as a
source of cell heterogeneity in budding yeast, Mech. Ageing
Dev., 176, 24-31, doi:10.1016/j.mad.2018.09.001.
21. Janssens, G. E., Meinema, A. C., González, J., Wolters,
J. C., Schmidt, A., Guryev, V., Bischoff, R., Wit, E. C.,
Veenhoff, L. M., and Heinemann, M. (2015) Protein bio-
genesis machinery is a driver of replicative aging in yeast,
Elife, 4, e08527, doi:10.7554/eLife.08527.
22. Orner, E. P., Zhang, P., Jo, M. C., Bhattacharya, S.,
Qin, L., and Fries, B. C. (2019) High-throughput yeast
aging analysis for Cryptococcus (HYAAC) microfluidic de-
vice streamlines aging studies in Cryptococcus neoformans,
Commun. Biol., 2, 256, doi:10.1038/s42003-019-0504-5.
23. Yu, R., Jo, M. C., and Dang, W. (2020) Measuring the
replicative lifespan of Saccharomyces cerevisiae using the
HYAA microfluidic platform, Methods Mol. Biol., 2144,
1-6, doi:10.1007/978-1-0716-0592-9_1.
24. Jin, M., Li, Y., O’Laughlin, R., Bittihn, P., Pillus, L.,
Tsimring, L. S., Hasty, J., and Hao, N. (2019) Diver-
gent aging of isogenic yeast cells revealed through sin-
gle-cell phenotypic dynamics, Cell Syst., 8, 242-253.e3,
doi:10.1016/j.cels.2019.02.002.
25. Li, Y., Jiang, Y., Paxman, J., O’Laughlin, R., Klepin,S.,
Zhu, Y., Pillus, L., Tsimring, L. S., Hasty, J., and Hao,N.
(2020) A programmable fate decision landscape under-
lies single-cell aging in yeast, Science, 369, 325-329,
doi:10.1126/science.aax9552.
26. Atamna, H., Killilea, D. W., Killilea, A. N., and Ames,
B.N. (2002) Heme deficiency may be a factor in the mi-
tochondrial and neuronal decay of aging, Proc. Natl. Acad.
Sci. USA, 99, 14807-14812, doi:10.1073/pnas.192585799.
27. Xie, Z., Zhang, Y., Zou, K., Brandman, O., Luo, C.,
Ouyang, Q., and Li, H. (2012) Molecular phenotyp-
ing of aging in single yeast cells using a novel micro-
AZBAROVA, KNORRE2004
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
fluidic device, Aging Cell, 11, 599-606, doi: 10.1111 /
j.1474-9726.2012.00821.x.
28. Krämer, L., Dalheimer, N., Räschle, M., Storchová, Z.,
Pielage, J., Boos, F., and Herrmann, J. M. (2023) Mi-
toStores: chaperone-controlled protein granules store
mitochondrial precursors in the cytosol, EMBO J., 42,
e112309, doi:10.15252/embj.2022112309.
29. Seppä, L., Hänninen, A.-L., and Makarow, M. (2004) Up-
regulation of the Hsp104 chaperone at physiological tem-
perature during recovery from thermal insult, Mol. Micro-
biol., 52, 217-225, doi:10.1111/j.1365-2958.2003.03959.x.
30. Zhou, Z., Liu, Y., Feng, Y., Klepin, S., Tsimring, L. S.,
Pillus, L., Hasty, J., and Hao, N. (2023) Engineering lon-
gevity-design of a synthetic gene oscillator to slow cellu-
lar aging, Science, 380, 376-381, doi: 10.1126/science.
add7631.
31. Yang, J., McCormick, M. A., Zheng, J., Xie, Z., Tsuchi-
ya, M., Tsuchiyama, S., El-Samad, H., Ouyang, Q.,
Kaeberlein, M., Kennedy, B. K., and Li, H. (2015) Sys-
tematic analysis of asymmetric partitioning of yeast pro-
teome between mother and daughter cells reveals “aging
factors” and mechanism of lifespan asymmetry, Proc.
Natl. Acad. Sci. USA, 112, 11977-11982, doi:10.1073/pnas.
1506054112.
32. Zhou, C., Slaughter, B. D., Unruh, J. R., Guo, F., Yu, Z.,
Mickey, K., Narkar, A., Ross, R. T., McClain, M., and
Li, R. (2014) Organelle-based aggregation and retention
of damaged proteins in asymmetrically dividing cells, Cell,
159, 530-542, doi:10.1016/j.cell.2014.09.026.
33. Hill, S. M., Hanzén, S., and Nyström, T. (2017) Restrict-
ed access: spatial sequestration of damaged proteins during
stress and aging, EMBO Rep., 18, 377-391, doi:10.15252/
embr.201643458.
34. Kumar, S., de Boer, R., and van der Klei, I. J. (2018) Yeast
cells contain a heterogeneous population of peroxisomes
that segregate asymmetrically during cell division, J.Cell
Sci., 131, jcs207522, doi:10.1242/jcs.207522.
35. Förtsch, J., Hummel, E., Krist, M., and Westermann,B.
(2011) Themyosin-related motor protein Myo2 is an es-
sential mediator of bud-directed mitochondrial move-
ment in yeast, J. Cell Biol., 194, 473-488, doi: 10.1083/
jcb.201012088.
36. Itoh, T., Toh-E, A., and Matsui, Y. (2004) Mmr1p is a
mitochondrial factor for Myo2p-dependent inheritance of
mitochondria in the budding yeast, EMBOJ., 23, 2520-
2530, doi:10.1038/sj.emboj.7600271.
37. Klecker, T., Scholz, D., Förtsch, J., and Westermann,B.
(2013) The yeast cell cortical protein Num1 integrates
mitochondrial dynamics into cellular architecture, J.Cell
Sci., 126, 2924-2930, doi:10.1242/jcs.126045.
38. Pernice, W. M., Vevea, J. D., and Pon, L. A. (2016) Arole
for Mfb1p in region-specific anchorage of high-function-
ing mitochondria and lifespan in Saccharomyces cerevisiae,
Nat. Commun., 7, 10595, doi:10.1038/ncomms10595.
39. Pernice, W. M., Swayne, T. C., Boldogh, I. R., and Pon,
L. A. (2017) Mitochondrial tethers and their impact on
lifespan in budding yeast, Front. Cell. Dev. Biol., 5, 120,
doi:10.3389/fcell.2017.00120.
40. McFaline-Figueroa, J. R., Vevea, J., Swayne, T. C.,
Zhou, C., Liu, C., Leung, G., Boldogh, I. R., and Pon,
L. A. (2011) Mitochondrial quality control during inher-
itance is associated with lifespan and mother-daughter
age asymmetry in budding yeast, Aging Cell,
10, 885-895,
doi:10.1111/j.1474-9726.2011.00731.x.
41. Manzano-López, J., Matellán, L., Álvarez-Llamas, A.,
Blanco-Mira, J.C., and Monje-Casas, F. (2019) Asym-
metric inheritance of spindle microtubule-organizing
centres preserves replicative lifespan, Nat. Cell. Biol., 21,
952-965, doi:10.1038/s41556-019-0364-8.
42. Hughes, A. L., and Gottschling, D. E. (2012) An early
age increase invacuolar pH limits mitochondrial function
and lifespan in yeast, Nature, 492, 261-265, doi:10.1038/
nature11654.
43. Azbarova, A. V., Galkina, K. V., Sorokin, M. I., Severin,
F.F., and Knorre, D.A. (2017) Thecontribution of Sac-
charomyces cerevisiae replicative age to the variations in the
levels of Trx2p, Pdr5p, Can1p and Idh isoforms, Sci. Rep.,
7, 13220, doi:10.1038/s41598-017-13576-w.
44. Galkina, K. V., Zyrina, A. N., Golyshev, S. A., Kash-
ko, N. D., Markova, O. V., Sokolov, S. S., Severin,
F. F., and Knorre, D. A. (2020) Mitochondrial dynam-
ics in yeast with repressed adenine nucleotide translo-
cator AAC2, Eur.J. Cell Biol., 99, 151071, doi:10.1016/j.
ejcb.2020.151071.
45. Nguyen, T. T. M., Iwaki, A., Ohya, Y., and Izawa, S. (2014)
Vanillin causes the activation of Yap1 and mitochondrial
fragmentation in Saccharomyces cerevisiae, J.Biosci. Bio-
eng., 117, 33-38, doi:10.1016/j.jbiosc.2013.06.008.
46. Rogov, A. G., Ovchenkova, A. P., Goleva, T. N., Kireev,
I.I., and Zvyagilskaya, R.A. (2017) New yeast models for
studying mitochondrial morphology as affected by oxida-
tive stress and other factors, Anal. Biochem., 552, 24-29,
doi:10.1016/j.ab.2017.04.003.
47. Scheckhuber, C. Q., Erjavec, N., Tinazli, A., Hamann,A.,
Nyström, T., and Osiewacz, H.D. (2007) Reducing mito-
chondrial fission results in increased life span and fitness
of two fungal ageing models, Nat. Cell Biol., 9, 99-105,
doi:10.1038/ncb1524.
48. Poveda-Huertes, D., Taskin, A. A., Dhaouadi, I., Myke-
tin, L., Marada, A., Habernig, L., Büttner, S., and Vögtle,
F.-N. (2021) Increased mitochondrial protein import and
cardiolipin remodelling upon early mtUPR, PLoS Genet.,
17, e1009664, doi:10.1371/journal.pgen.1009664.
49. Liu, Q., Chang, C. E., Wooldredge, A. C., Fong, B., Ken-
nedy, B.K., and Zhou, C. (2022) Tom70-based transcrip-
tional regulation of mitochondrial biogenesis and aging,
Elife, 11, e75658, doi:10.7554/eLife.75658.
50. Veatch, J. R., McMurray, M. A., Nelson, Z. W., and
Gottschling, D. E. (2009) Mitochondrial dysfunction
leads to nuclear genome instability via an iron-sulfur
cluster defect, Cell, 137, 1247-1258, doi: 10.1016/j.cell.
2009.04.014.
ROLE OF MITOCHONDRIA IN YEAST AGING 2005
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
51. Giaever, G., Chu, A. M., Ni, L., Connelly, C., Riles, L.,
ronneau, S., Dow, S., Lucau-Danila, A., Anderson,K.,
André, B., Arkin, A. P., Astromoff, A., El-Bakkoury, M.,
Bangham, R., Benito, R., Brachat, S., Campanaro, S.,
Curtiss, M., Davis, K., Deutschbauer, A., Entian, K.-D.,
Flaherty, P., Foury, F., Garfinkel, D. J., Gerstein, M.,
Gotte, D., Güldener, U., Hegemann, J. H., Hempel, S.,
Herman, Z., Jaramillo, D. F., Kelly, D. E., Kelly, S. L.,
Kötter, P., LaBonte, D., Lamb, D. C., Lan, N., Li-
ang,H., Liao, H., Liu, L., Luo, C., Lussier, M., Mao,R.,
Menard,P., Ooi, S. L., Revuelta, J. L., Roberts, C. J.,
Rose, M., Ross-Macdonald, P., Scherens, B., Schim-
mack, G., Shafer, B., Shoemaker, D. D., Sookhai-
Mahadeo, S., Storms, R. K., Strathern, J. N., Val-
le, G., Voet, M., Volckaert, G., Wang, C.-Y., Ward, T.
R., Wilhelmy, J., Winzeler, E. A., Yang, Y., Yen, G.,
Youngman, E., Yu, K., Bussey, H., Boeke, J. D., Sny-
der, M., Philippsen, P., Davis, R. W., and Johnston, M.
(2002) Functional profiling of the Saccharomyces cer-
evisiae genome, Nature, 418, 387-391, doi: 10.1038/
nature00935.
52. Janouškovec, J., Tikhonenkov, D. V., Burki, F., Howe,
A. T., Rohwer, F. L., Mylnikov, A. P., and Keeling, P. J.
(2017) Anew lineage of eukaryotes illuminates early mito-
chondrial genome reduction, Curr. Biol., 27, 3717-3724.e5,
doi:10.1016/j.cub.2017.10.051.
53. Kirchman, P. A., Kim, S., Lai, C. Y., and Jazwinski, S.M.
(1999) Interorganelle signaling is a determinant of longev-
ity in Saccharomyces cerevisiae, Genetics, 152, 179-190,
doi:10.1093/genetics/152.1.179.
54. Borghouts, C., Benguria, A., Wawryn, J., and Jazwins-
ki, S. M. (2004) Rtg2 protein links metabolism and ge-
nome stability in yeast longevity, Genetics, 166, 765-777,
doi:10.1093/genetics/166.2.765.
55. Woo, D. K., and Poyton, R. O. (2009) Theabsence of a
mitochondrial genome in rho
0
yeast cells extends lifespan
independently of retrograde regulation, Exp. Gerontol.,
44, 390-397, doi:10.1016/j.exger.2009.03.001.
56. Miceli, M. V., Jiang, J. C., Tiwari, A., Rodriguez-
Quiñones, J.F., and Jazwinski, S.M. (2011) Loss of mi-
tochondrial membrane potential triggers the retrograde
response extending yeast replicative lifespan, Front. Genet.,
2, 102, doi:10.3389/fgene.2011.00102.
57. Jiang, J. C., Stumpferl, S. W., Tiwari, A., Qin, Q., Ro-
driguez-Quiñones, J. F., and Jazwinski, S. M. (2016)
Identification of the target of the retrograde response
that mediates replicative lifespan extension in Saccha-
romyces cerevisiae, Genetics, 204, 659-673, doi: 10.1534/
genetics.116.188086.
58. Heeren, G., Jarolim, S., Laun, P., Rinnerthaler, M.,
Stolze, K., Perrone, G. G., Kohlwein, S. D., Nohl, H.,
Dawes, I. W., and Breitenbach, M. (2004) The role of
respiration, reactive oxygen species and oxidative stress
in mother cell-specific ageing of yeast strains defective in
the RAS signalling pathway, FEMS Yeast Res., 5, 157-167,
doi:10.1016/j.femsyr.2004.05.008.
59. Kaeberlein, M., Hu, D., Kerr, E. O., Tsuchiya, M.,
Westman, E. A., Dang, N., Fields, S., and Kennedy,
B. K. (2005) Increased life span due to calorie restric-
tion in respiratory-deficient yeast, PLoS Genet., 1, e69,
doi:10.1371/journal.pgen.0010069.
60. Kaeberlein, M., Kirkland, K. T., Fields, S., and Kennedy,
B.K. (2005) Genes determining yeast replicative life span
in a long-lived genetic background, Mech. Ageing Dev.,
126, 491-504, doi:10.1016/j.mad.2004.10.007.
61. Traven, A., Wong, J. M., Xu, D., Sopta, M., and Ingles,
C. J. (2001) Interorganellar communication. Altered nu-
clear gene expression profiles in a yeast mitochon-
drial DNA mutant, J. Biol. Chem., 276, 4020-4027,
doi:10.1074/jbc.M006807200.
62. Epstein, C. B., Waddle, J. A., Hale, W., Davé, V., Thorn-
ton, J., Macatee, T. L., Garner, H. R., and Butow, R. A.
(2001) Genome-wide responses to mitochondrial dys-
function, Mol. Biol. Cell, 12, 297-308, doi: 10.1091/
mbc.12.2.297.
63. Liu, S., Liu, S., He, B., Li, L., Li, L., Wang, J., Cai, T.,
Chen, S., and Jiang, H. (2021) OXPHOS deficiency ac-
tivates global adaptation pathways to maintain mito-
chondrial membrane potential, EMBO Rep., 22, e51606,
doi:10.15252/embr.202051606.
64. Liu, Z., and Butow, R. A. (1999) Atranscriptional switch
in the expression of yeast tricarboxylic acid cycle genes in
response to a reduction or loss of respiratory function, Mol.
Cell. Biol., 19, 6720-6728, doi:10.1128/MCB.19.10.6720.
65. Starovoytova, A. N., Sorokin, M. I., Sokolov, S. S., Sev-
erin, F.F., and Knorre, D.A. (2013) Mitochondrial sig-
naling in Saccharomyces cerevisiae pseudohyphae forma-
tion induced by butanol, FEMS Yeast Res., 13, 367-374,
doi:10.1111/1567-1364.12039.
66. Kwan, E. X., Wang, X. S., Amemiya, H. M., Brewer, B.J.,
and Raghuraman, M.K. (2016) rDNA copy number vari-
ants are frequent passenger mutations in Saccharomyces
cerevisiae deletion collections and denovo transformants,
G3, 6, 2829-2838, doi:10.1534/g3.116.030296.
67. Hotz, M., Thayer, N. H., Hendrickson, D. G., Schinski,
E.L., Xu, J., and Gottschling, D.E. (2022) rDNA array
length is a major determinant of replicative lifespan in bud-
ding yeast, Proc. Natl. Acad. Sci. USA,
119, e2119593119,
doi:10.1073/pnas.2119593119.
68. Garcia, E. J., de Jonge, J. J., Liao, P.-C., Stivison, E.,
Sing, C. N., Higuchi-Sanabria, R., Boldogh, I. R., and
Pon, L. A. (2019) Reciprocal interactions between mtDNA
and lifespan control in budding yeast, Mol. Biol. Cell, 30,
2943-2952, doi:10.1091/mbc.E18-06-0356.
69. Morgenstern, M., Stiller, S. B., Lübbert, P., Peikert, C.D.,
Dannenmaier, S., Drepper, F., Weill, U., Höß, P., Feuer-
stein, R., Gebert, M., Bohnert, M., van der Laan, M.,
Schuldiner, M., Schütze, C., Oeljeklaus, S., Pfanner,N.,
Wiedemann, N., and Warscheid, B. (2017) Definition
of a high-confidence mitochondrial Proteome at quan-
titative scale, Cell Rep., 19, 2836-2852, doi: 10.1016/
j.celrep.2017.06.014.
AZBAROVA, KNORRE2006
BIOCHEMISTRY (Moscow) Vol. 88 Nos. 12-13 2023
70. Puddu, F., Herzog, M., Selivanova, A., Wang, S., Zhu,J.,
Klein-Lavi, S., Gordon, M., Meirman, R., Millan-
Zambrano, G., Ayestaran, I., Salguero, I., Sharan, R.,
Li, R., Kupiec, M., and Jackson, S. P. (2019) Genome
architecture and stability in the Saccharomyces cerevisiae
knockout collection, Nature, 573, 416-420, doi:10.1038/
s41586-019-1549-9.
71. Heeren, G., Rinnerthaler, M., Laun, P., von Seyerl, P.,
Kössler, S., Klinger, H., Hager, M., Bogengruber, E.,
Jarolim, S., Simon-Nobbe, B., Schüller, C., Carmona-
Gutierrez, D., Breitenbach-Koller, L., Mück, C., Jansen-
Dürr, P., Criollo, A., Kroemer, G., Madeo, F., and Breit-
enbach, M. (2009) Themitochondrial ribosomal protein
of the large subunit, Afo1p, determines cellular longevity
through mitochondrial back-signaling via TOR1, Aging,
1, 622-636, doi:10.18632/aging.100065.
72. Caballero, A., Ugidos, A., Liu, B., Öling, D., Kvint,K.,
Hao, X., Mignat, C., Nachin, L., Molin, M., and
Nyström, T. (2011) Absence of mitochondrial translation
control proteins extends life span by activating sirtuin-de-
pendent silencing, Mol. Cell, 42, 390-400, doi: 10.1016/
j.molcel.2011.03.021.
73. Foury, F., Roganti, T., Lecrenier, N., and Purnelle, B.
(1998) The complete sequence of the mitochondrial ge-
nome of Saccharomyces cerevisiae, FEBS Lett., 440,
325-331, doi:10.1016/S0014-5793(98)01467-7.
74. Bykov, Y. S., Flohr, T., Boos, F., Zung, N., Herrmann,
J.M., and Schuldiner, M. (2022) Widespread use of un-
conventional targeting signals in mitochondrial ribo-
some proteins, EMBO J., 41, e109519, doi: 10.15252/
embj.2021109519.
75. Piper, P. W., Jones, G. W., Bringloe, D., Harris, N.,
MacLean, M., and Mollapour, M. (2002) The short-
ened replicative life span of prohibitin mutants of yeast
appears to be due to defective mitochondrial segregation
in old mother cells, Aging Cell, 1, 149-157, doi:10.1046/
j.1474-9728.2002.00018.x.
76. Yi, D.-G., Hong, S., and Huh, W.-K. (2018) Mitochon-
drial dysfunction reduces yeast replicative lifespan by
elevating RAS-dependent ROS production by the ER-
localized NADPH oxidase Yno1, PLoS One, 13, e0198619,
doi:10.1371/journal.pone.0198619.
77. Stenger, M., Le, D. T., Klecker, T., and Westermann,B.
(2020) Systematic analysis of nuclear gene function in
respiratory growth and expression of the mitochon-
drial genome in S. cerevisiae, Microb. Cell, 7, 234-249,
doi:10.15698/mic2020.09.729.
78. Erjavec, N., Bayot, A., Gareil, M., Camougrand, N.,
Nystrom, T., Friguet, B., and Bulteau, A.-L. (2013) De-
letion of the mitochondrial Pim1/Lon protease in yeast
results in accelerated aging and impairment of the pro-
teasome, Free Radic. Biol. Med., 56, 9-16, doi: 10.1016/
j.freeradbiomed.2012.11.019.
79. Barros, M. H., Bandy, B., Tahara, E. B., and Kowaltows-
ki, A.J. (2004) Higher respiratory activity decreases mi-
tochondrial reactive oxygen release and increases life span
in Saccharomyces cerevisiae, J. Biol. Chem., 279, 49883-
49888, doi:10.1074/jbc.M408918200.
80. Galkina, K. V., Finkelberg, J. M., Markova, O. V., Az-
barova, A. V., Banerjee, A., Kumari, S., Sokolov, S.S.,
Severin, F. F., Prasad, R., and Knorre, D. A. (2020) Pro-
tonophore FCCP provides fitness advantage to PDR-
deficient yeast cells, J.Bioenerg. Biomembr., 52, 383-395,
doi:10.1007/s10863-020-09849-1.
81. Jiang, J. C., Jaruga, E., Repnevskaya, M. V., and Jazwins-
ki, S. M. (2000) An intervention resembling caloric re-
striction prolongs life span and retards aging in yeast,
FASEBJ., 14, 2135-2137, doi:10.1096/fj.00-0242fje.
82. Stöckl, P., Zankl, C., Hütter, E., Unterluggauer, H.,
Laun,P., Heeren, G., Bogengruber, E., Herndler-Brand-
stetter, D., Breitenbach, M., and Jansen-Dürr, P. (2007)
Partial uncoupling of oxidative phosphorylation in-
duces premature senescence in human fibroblasts and
yeast mother cells, Free Radic. Biol. Med., 43, 947-958,
doi:10.1016/j.freeradbiomed.2007.06.005.
83. Dupont, C.-H., Mazat, J. P., and Guerin, B. (1985)
The role of adenine nucleotide translocation in the en-
ergization of the inner membrane of mitochondria iso-
lated from ϱ
+
and ϱ
o
strains of Saccharomyces cerevisi-
ae, Biochem. Biophys. Res. Commun., 132, 1116-1123,
doi:10.1016/0006-291X(85)91922-9.
84. Antonenko, Y. N., Avetisyan, A. V., Cherepanov, D.A.,
Knorre, D. A., Korshunova, G. A., Markova, O. V.,
Ojovan, S. M., Perevoshchikova, I. V., Pustovidko, A.V.,
Rokitskaya, T. I., Severina, I. I., Simonyan, R. A., Smir-
nova, E. A., Sobko, A. A., Sumbatyan, N. V., Sever-
in, F. F., and Skulachev, V. P. (2011) Derivatives of
rhodamine19 as mild mitochondria-targeted cationic un-
couplers, J.Biol. Chem., 286, 17831-17840, doi:10.1074/
jbc.M110.212837.
85. Leupold, S., Hubmann, G., Litsios, A., Meinema, A.C.,
Takhaveev, V., Papagiannakis, A., Niebel, B., Jans-
sens, G., Siegel, D., and Heinemann, M. (2019) Sac-
charomyces cerevisiae goes through distinct metabolic
phases during its replicative lifespan, Elife, 8, e41046,
doi:10.7554/eLife.41046.
Publisher’s Note. Pleiades Publishing remains neu-
tral with regard to jurisdictional claims in published
maps and institutional affiliations.