ISSN 0006-2979, Biochemistry (Moscow), 2025, Vol. 90, No. 12, pp. 1929-1943 © The Author(s) 2025. This article is an open access publication.
Published in Russian in Biokhimiya, 2025, Vol. 90, No. 12, pp. 2032-2048.
1929
REVIEW
ATP in Mitochondria: Quantitative Measurement,
Regulation, and Physiological Role
Anna S. Lapashina
1,2
, Danila O. Tretyakov
1,2
, and Boris A. Feniouk
1,2,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,
119992 Moscow, Russia
a
e-mail: feniouk@belozersky.msu.ru
Received September 23, 2025
Revised November 18, 2025
Accepted November 20, 2025
AbstractOxidative phosphorylation in mitochondria is the main source of ATP in most eukaryotic cells.
Concentrations of ATP, ADP, and AMP affect numerous cellular processes, including macromolecule biosynthe-
sis, cell division, motor protein activity, ion homeostasis, and metabolic regulation. Variations in ATP levels
also influence concentration of free Mg
2+
, thereby extending the range of affected reactions. In the cytosol,
adenine nucleotide concentrations are relatively constant and typically are around 5  mM  ATP, 0.5  mM  ADP,
and 0.05 mM AMP. These concentrations are mutually constrained by adenylate kinases operating in the
cytosol and intermembrane space and are further linked to mitochondrial ATP and ADP pools via the ad-
enine nucleotide translocator. Quantitative data on absolute adenine nucleotide concentrations in the mito-
chondrial matrix are limited. Total adenine nucleotide concentration lies in the millimolar range, but the
matrix ATP/ADP ratio is consistently lower than the cytosolic ratio. Estimates of nucleotide fractions show
substantial variability (ATP 20-75%, ADP 20-70%, AMP 3-60%), depending on the organism and experimen-
tal conditions. These observations suggest that the ‘state 4’ – inhibition of oxidative phosphorylation in the
resting cells due to the low matrix ADP and elevated proton motive force that impedes respiratory chain
activity – is highly unlikely in vivo. In this review, we discuss proteins regulating ATP levels in mitochon-
dria and cytosol, consider experimental estimates of adenine nucleotide concentrations across a range of
biological systems, and examine the methods used for their quantification, with particular emphasis on the
genetically encoded fluorescent ATP sensors such as ATeam, QUEEN, and MaLion.
DOI: 10.1134/S0006297925603338
Keywords: mitochondria, ATP, ADP, ATeam, adenine nucleotide translocator (ANT), ATP synthase
* To whom correspondence should be addressed.
INTRODUCTION
Mitochondria are traditionally called the pow-
erhouses of eukaryotic cells, as one of their main
functions is ATP production through oxidative phos-
phorylation. Most biochemistry textbooks explain
that the majority of ATP is synthesized in the mito-
chondrial matrix from ADP and inorganic phosphate
by the H
+
-transporting ATP synthase (also known as
F
O
F
1
, ComplexV, F-ATPase, or F
1
F
O
-ATPase) located in
the inner mitochondrial membrane. ATP synthesis is
driven by the proton motive force (pmf) across the
inner mitochondrial membrane. The pmf is generated
by the respiratory-chain enzymes that oxidize NADH
and succinate produced in the tricarboxylic acid
(TCA) cycle, and, depending on the organism or tis-
sue, by the oxidation of additional substrates such as
glycerol-3-phosphate. The adenine nucleotide translo-
cator (ANT; also known as the ATP/ADP carrier, ATP/
ADP antiporter, AAC, or ATP/ADP translocase) exports
ATP from the matrix into the intermembrane space
in exchange for ADP, whereas inorganic phosphate
enters the matrix via the phosphate carrier (PiC,
encoded by the SLC25A3 gene in humans and MIR1
LAPASHINA et al.1930
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
in baker’s yeast). Nucleotide and phosphate exchange
between the cytosol and the intermembrane space
occurs through the mitochondrial porins – voltage-
dependent anion channels (VDAC) – located in the
outer mitochondrial membrane.
This simplified outline captures the basic prin-
ciples of oxidative phosphorylation, but it is insuffi-
cient for understanding the regulatory mechanisms
governing the process or clarifying how its dysfunc-
tion is related to various pathologies.
Decades of in  vitro research have generated an
extensive body of data on the functions of individual
proteins involved in oxidative phosphorylation and
the mechanisms that regulate them. Recent advanc-
es in microscopy have further accelerated structural
studies of mitochondria. Notably, cryo-electron mi-
croscopy has recently enabled high-resolution (2.5  Å)
visualization of mammalian respiratory-chain super-
complexes in their native cellular environment, re-
vealing previously unknown aspects of their organi-
zation and interactions [1]. This is only one of many
studies that have revealed the structure of mitochon-
drial proteins, their relative positions, dynamics, and
roles in the formation of cristae and contacts with
the endoplasmic reticulum. The body of experimental
data on the structure and localization of mitochondri-
al proteins in vivo is also growing rapidly.
Despite the considerable progress in mitochon-
drial biology, surprisingly little is known about the
precise conditions under which oxidative phosphory-
lation proceeds in vivo. We have very little quantita-
tive data on the absolute concentrations of ATP, ADP,
and AMP in the mitochondrial matrix, on how these
concentrations vary under different physiological and
pathological states, and on how such changes relate
to the pmf or to reactive oxygen species (ROS) pro-
duction. Because ATP and ADP are exchanged rapid-
ly between the matrix and the cytosol, mitochondrial
nucleotide concentrations modulate cytosolic pools
and are themselves shaped by them. Thus, shifts in
the intramitochondrial ATP/ADP/AMP ratio influence
both mitochondrial enzyme function and metabolic
processes in the cytosol.
Moreover, changes in ATP concentration in both
the cytosol and the mitochondrial matrix affect cel-
lular physiology by altering free magnesium levels.
ATP binds Mg
2+
roughly an order of magnitude more
strongly than ADP. In solutions with an ionic strength
of 0.1-0.15  M, the logarithms of the magnesium bind-
ing constants are approximately 4.0-4.6 for ATP,
3.1-3.5 for ADP, and 1.9-2.1 for AMP  [2]. Therefore,
ATP hydrolysis to ADP increases the concentration
of free magnesium, which in turn influences numer-
ous cellular processes, including those that are not
directly sensitive to the changes in nucleotide levels.
A notable example is chromatin condensation during
mitosis, which is triggered by the rise in cytosolic
free magnesium from ~0.5 to 1  mM. This increase oc-
curs in parallel with the drop in ATP concentration,
and it has been proposed that the ATP decline results
from the elevated ATP hydrolysis rate [3].
An example of a magnesium-dependent process
in mitochondria is the A/D transition of complex I
in the mammalian respiratory chain. Under hypoxic
conditions, complex I shifts from the active  (A) to the
de-activated  (D) state, and upon restoration of respi-
ration, the reverse transition normally occurs within
minutes. However, in the presence of 1-5 mM Mg
2+
,
this reactivation is slowed by several orders of mag-
nitude [4]. Given that the free magnesium concentra-
tion in the matrix exceeds that in the cytosol, this
phenomenon is likely to contribute to the regulation
of mitochondrial recovery from ischemia in  vivo.
In this review, we have aimed to summarize the
available literature on adenine nucleotide concen-
trations in the cytosol and mitochondrial matrix of
eukaryotic cells, as well as the methods used to mea-
sure ATP levels within mitochondria.
PROTEINS AFFECTING ATP CONCENTRATION
IN MITOCHONDRIA
The two main ‘players’ controlling the dynamics
of ATP concentration in the mitochondrial matrix are
ATP synthase [5] and adenine nucleotide transloca-
tor ANT [6], both located in the inner mitochondrial
membrane. Each of these proteins accounts for a sig-
nificant proportion of the total mitochondrial mem-
brane protein – up to 12% for ANT [6] and up to
20%for ATP synthase [7, 8]. Analysis of the mitochon-
drial proteome of the HEK293T cell line derived from
human embryonic kidneys showed that, in terms of
the number of molecules per cell, ANT (all isoforms)
is the most abundant mitochondrial protein, with
more than 8 million molecules per cell, while the
number of ATP synthase molecules reaches nearly
2 million per cell [9]. Only about 20 mitochondrial
proteins exceed ATP synthase in copy number. These
include, besides ANT, chaperones HSP60, HSP10, and
mt-HSP70, peroxiredoxin PRDX3; membrane chan-
nels such as the outer membrane porins VDAC1 and
VDAC3; several enzymes of the central carbon me-
tabolism such as malate dehydrogenase (MDH2) and
citrate synthase (CS) from the TCA cycle; phosphate
transporter SLC25A3; and several proteins involved
in mitochondrial translation and protein import [9].
ATP synthase can catalyze not only ATP synthe-
sis but also ATP hydrolysis. It thus links the activity
of the respiratory chain enzymes, membrane pro-
ton conductivity, and the ATP/ADP ratio in the ma-
trix. When pmf decreases due to increased proton
ATP IN MITOCHONDRIA 1931
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
conductivity or reduced activity of respiratory chain
enzymes, ATP synthase switches from synthesizing
to hydrolyzing ATP, pumping protons from the mito-
chondrial matrix into the intermembrane space and
preventing depolarization of the inner mitochon-
drial membrane. If pmf is absent, the enzyme may
exhibit high ATPase activity: it can hydrolyze from
several hundred to thousands of ATP molecules per
second [10].
Under the low-pmf conditions, the ATPase activ-
ity of mitochondrial ATP synthase is suppressed via
non-competitive inhibition by the MgADP complex
(the so-called ADP inhibition [10]) and by the regu-
latory protein IF1, which binds to the enzyme and
blocks its ATP hydrolysis activity without impeding
ATP synthesis [11]. In mammals, IF1 is believed to
prevent ATP depletion in the cell under the ener-
gy-deprived conditions, such as ischemia [11]. Studies
conducted by our group on bakers yeast have shown
that IF1 helps to resume division more quickly after
starvation [12]. Along with that, attenuated ADP inhi-
bition increased the growth rate in yeast lacking mi-
tochondrial DNA (rho
0
strain), which are incapable of
oxidative phosphorylation and instead synthesize ATP
in the cytosol via glycolysis [13]. Such cells lack a ful-
ly functional ATP synthase; however, their mitochon-
dria contain the hydrophilic F
1
subcomplex, which is
encoded by the nuclear genes and can hydrolyze but
not synthesize ATP. For the rho
0
cells, intensive ATP
hydrolysis in the mitochondrial matrix proves ad-
vantageous, as ANT exchanges ADP produced in the
matrix for cytoplasmic ATP, transferring one negative
charge from the cytosol to the matrix and thereby
maintaining pmf necessary for protein import and
several other mitochondrial functions. Thus, activities
of ATP synthase and its F
1
subcomplex, along with
regulation of these activities, are important factors
influencing mitochondrial ATP concentration and,
consequently, cell physiology.
ANT is highly specific for ATP and ADP. It can
transport some other nucleotides and small mole-
cules, but only at more than 100 times lower rates,
rendering such fluxes negligible under physiological
conditions. It is ANT that determines the specificity of
oxidative phosphorylation for ADP; ATP synthase can,
although with lower efficiency, phosphorylate other
nucleoside diphosphates, primarily GDP [14]. ANT is
an electrogenic carrier, and under normal conditions,
high pmf across the inner mitochondrial membrane
shifts the equilibrium toward the export of ATP from
mitochondria to the cytosol.
The ATP/ADP exchange is inhibited in the pres-
ence of magnesium because the nucleotides are trans-
ferred in their free form rather than as magnesium
complexes [15, 16] – a rare case for proteins that
bind nucleoside triphosphates.
Adenylate kinases also influence the concentra-
tions of ATP and other nucleotides in the cytosol and
mitochondria. These enzymes catalyze the reversible
transfer of a phosphate group between nucleoside
triphosphates (usually ATP or GTP) and AMP, pro-
ducing two nucleoside diphosphates. Currently, nine
isoforms of adenylate kinase have been characterized
in human cells ([17], see also a recent review [18] on
the physiological role of adenylate kinases in health
and disease). Here, we will consider only two mito-
chondrial enzymes: AK2 of the intermembrane space
and AK3 (GTP:AMP phosphotransferase) of the mito-
chondrial matrix. Another member of the adenylate
kinase family, AK4, is also localized in the mitochon-
drial matrix in some tissues but apparently lacks en-
zymatic activity [19,  20]; therefore, it is not discussed
in detail here. (It should be noted, however, that the
recombinant human AK4 with intact mitochondri-
al localization peptide expressed in Escherichia coli
cells demonstrated the ability to phosphorylate AMP
in the presence of ATP or GTP [21]). Despite its prob-
able lack of enzymatic activity in  vivo, AK4 is an im-
portant regulator of cellular processes that plays a
role in the oxidative stress response, participates in
hypoxic regulation of mitochondria, and is associated
with drug resistance, aggressive tumor progression,
and metastasis in several cancers, making it a prom-
ising diagnostic and therapeutic target [18].
Adenylate kinase AK2 interconverts ATP, ADP,
and AMP in the mitochondrial intermembrane space
and catalyzes the reaction (1):
NTP + AMP ⇆ ADP + NDP. (1)
Both purine and pyrimidine nucleoside triphos-
phates (NTPs) can act as phosphate donors, but only
AMP can be phosphorylated by AK2 [22]. Like the
vast majority of enzymes whose substrates are nu-
cleoside triphosphates, AK2 and all other adenylate
kinases work only with their magnesium complexes.
AMP, which has about two orders of magnitude low-
er affinity for magnesium ions than ATP, binds and
is phosphorylated without magnesium [23]. The ap-
parent equilibrium constant of the adenylate kinase
reaction varies between 0.4 and 1.3 depending on the
magnesium concentration, reaching a maximum at
approximately 0.5 mM of free magnesium [24]. AK2,
which is localized in the intermembrane space, and
cytosolic adenylate kinases establish a strict relation-
ship between the concentrations of cytoplasmic ade-
nine nucleotides (see the section ‘Adenine nucleotides
in the cytosol’ below). AK2 also plays an important
role in the dynamic disequilibrium of nucleotide com-
position between the cytosol and the intermembrane
space, converting AMP from the cytosol into ADP us-
ing ATP exported from the mitochondria by ANT [25].
LAPASHINA et al.1932
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
Fig. 1. Oxidative phosphorylation in human mitochondria. Proteins are labeled in bold: green – ATP synthase, cyan – phos-
phate transporter, red – respiratory chain enzymes, purple – adenine nucleotide translocator (ANT), yellow – adenylate
kinase AK2, brown – adenylate kinase AK3, gray – porin VDAC. Bold arrows indicate transport of compounds; thin arrows
indicate chemical transformations. The diagram depicts the situation when the respiratory chain generates a sufficiently
high proton motive force (pmf ) for ATP synthesis; pH is 8.0 in the matrix and 7.4 in both the cytosol and intermembrane
space. TCA cycle stands for tricarboxylic acid cycle.
Adenylate kinase AK3 (GTP:AMP phosphotrans-
ferase) is localized in the mitochondrial matrix [26]
and catalyzes the reaction (2):
GTP + AMP ⇆ ADP + GDP. (2)
This enzyme is present in mitochondria in a sig-
nificant amount: a proteomic analysis of the human
HEK293T cell line has shown that the number of AK3
molecules per cell is approximately 700,000 [9].
Functions of the proteins described above, as
well as the phosphate transporter SLC25A3 [27], are
shown schematically in Fig. 1.
Experimental studies from the 1970s and 1980s
indicate that the rate of oxidative phosphorylation
is determined by the ATP/ADP ratio in the inter-
membrane space [28, 29], and that the rate-limit-
ing process is ATP/ADP translocation through ANT
[30, 31]. However, a more detailed analysis has re-
vealed that ANT is only one of several key factors
constraining the rate of oxidative phosphorylation,
with its role being most significant when the respi-
ration rate is approximately 80% of the maximum
possible value [32].
Moreover, the ATP-Mg/P
i
carrier located in the in-
ner mitochondrial membrane also influences nucleo-
tide concentration dynamics in the mitochondrial ma-
trix. This protein facilitates electroneutral transport
of the MgATP
2−
complex and inorganic phosphate
through the inner mitochondrial membrane [33] (not
shown in Fig.  1). According to the proteomic study
mentioned above, the total number of all isoforms
of this membrane protein approaches 300,000 mole-
cules per cell [9].
Outside mitochondria, the phosphocreatine shut-
tle (and, in many invertebrates, the phosphoarginine
shuttle) significantly influences ATP concentration.
Mitochondrial creatine kinase (or arginine kinase), lo-
cated in the intermembrane space, uses ATP to phos-
phorylate creatine (or arginine). Cytosolic isoforms
of the same enzymes catalyze the reverse reaction,
regenerating ATP from phosphocreatine and ADP,
ATP IN MITOCHONDRIA 1933
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
and are associated with the sites of high energy con-
sumption, such as muscle fiber myosin, sarco/endo-
plasmic reticulum Ca
2+
-ATPase, plasma membrane
Na
+
/K
+
-ATPase, and other ATP-utilizing proteins.
However, a detailed examination of how these en-
zymes regulate cytosolic ATP concentrations lies be-
yond the scope of this review.
ADENINE NUCLEOTIDES IN THE CYTOSOL
As shown in Fig.  1, oxidative phosphorylation in
mitochondria results in the conversion of cytosolic
AMP, ADP, and inorganic phosphate to ATP. It should
be noted that, with high adenylate kinase activity,
the concentrations of ATP, ADP, and AMP will remain
close to the thermodynamic equilibrium. Thus, if
the rate of ATP consumption in the cell exceeds the
rate of its synthesis, ATP concentration will decrease
while ADP and AMP concentrations will increase.
Importantly, in this situation, the AMP concentration
relative to ATP will increase much more than the
ADP concentration [34]. Indeed, for the adenylate ki-
nase reaction (3):
[ATP] × [AMP]
[ADP]
2
= K, (3)
where [ATP], [ADP], and [AMP] are the equilibrium
concentrations of the corresponding nucleotides, and
K is the equilibrium constant, we have (4):
[ATP] × [AMP] = K × [ADP]
2
, (4)
and dividing both sides of the equation by [ATP]
2
, we
obtain (5):
[AMP]
[ATP]
(
[ADP]
[ATP]
)
2
, (5)
meaning the AMP/ATP ratio is proportional to the
square of the ADP/ATP ratio. Therefore, it is not sur-
prising that an increase in the AMP concentration sig-
nals energy deficiency in the cell. The hypothesis that
the AMP/ATP ratio regulates the activity of enzymes
involved in energy metabolism was first proposed by
Atkinson in 1964 during his study of yeast phospho-
fructokinase [35]. Atkinson later expanded this hypoth-
esis by introducing a variable reflecting the number
of high-energy phosphoanhydride bonds per adenine
nucleotide in the cell. Since an ATP molecule contains
two such bonds, and ADP contains one, this variable
is calculated by the formula (6):
2[ATP] + [ADP]
[ATP] + [ADP] + [AMP]
, (6)
ranging from zero (pure AMP; no ADP or ATP) to
two (pure ATP; no ADP or AMP). To normalize the
range to unity, Atkinson divided the expression by 2
and called the result the ‘adenylate energy charge’
(AEC) of the cell, which reflects the energy available
in the phosphoanhydride bonds of adenine nucleo-
tides for metabolic reactions [36]. Thus, cellular AEC
is calculated from intracellular concentrations of ATP,
ADP, and AMP using the formula (7):
AEC =
[ATP] + ½[ADP]
[ATP] + [ADP] + [AMP]
(7)
and ranges from zero to one.
Soon after its initial success, the AEC hypothesis
became less popular because the number of enzymes
directly regulated by adenine nucleotides was found
to be quite small. A revival occurred in the late 1980s
with the discovery of the AMP-activated protein ki-
nase (AMPK) cascade [37], which in response to an
increase in AMP initiates a series of reactions that
stimulate catabolism and suppress ATP-consuming
processes both on a short-term (via enzyme phos-
phorylation) and long-term (via gene expression) time
scale [38]. The subsequent extensive body of data on
AMPK activation during hypoxia, ischemia, muscle
activity, and other states of cellular energy depletion
provided a solid experimental basis for the role of
AEC in regulating cellular bioenergetics and broad-
ened its relevance beyond the strictly metabolic reg-
ulation [34].
As previously noted, due to the activity of ade-
nylate kinases, concentrations of adenine nucleotides
in the cytosol are tightly interconnected, governed
by thermodynamics of the adenylate kinase reaction.
Blair [24] calculated these concentrations and the free
magnesium levels equilibrated by adenylate kinase at
different AEC values under conditions close to those
in the cytosol (pH  7.5, 100  mM  K
+
; Fig.  2).
The data shown in Fig.  2 illustrate the connec-
tion between the ATP hydrolysis and the increase in
free magnesium concentration, as well as the relative
stability of ADP concentration across a wide range
of AEC values.
To date, a significant amount of experimental
data on nucleotide concentrations in the cytosol has
been accumulated. Most of these data are obtained
from measurements using whole cells and, strictly
speaking, reflect not the cytosolic concentrations but
the total, averaged concentrations in both cytosol
and mitochondria. However, since the volume of mi-
tochondria is generally much smaller than that of the
cytosol, the contribution of mitochondrial nucleotides
to the total pool is also small and, as a first approx-
imation, can be neglected, except for cardiomyocytes
and other cells with very high mitochondrial con-
tent. In a large-scale metabolomic study quantifying
the most abundant low-molecular-weight metabolites
in human cells and baker’s yeast, concentrations of
LAPASHINA et al.1934
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
Fig. 2. Relationship between adenylate energy charge (AEC)
and concentrations of free magnesium and adenine nucle-
otides in the presence of adenylate kinase. Data were tak-
en from [24] for the following conditions: total nucleotide
concentration 5mM, total magnesium concentration 5mM,
potassium concentration 100 mM, pH 7.5. The graph shows
total nucleotide concentrations (magnesium-bound plus free)
for ATP, ADP, and AMP, and concentration of free Mg
2+
.
adenine nucleotides in eukaryotic cells were reported
to fall within the following ranges: ATP, 1.4-7.0  mM;
ADP, 0.43-0.57  mM; and AMP, 0.036-0.103  mM [39].
These values are in good agreement with the majority
of previously obtained data [40, 41] and correspond
to an AEC value of about 0.95 (see Fig. 2).
ADENINE NUCLEOTIDES
IN MITOCHONDRIA
The interconversion of adenine nucleotides in
mitochondria is less understood than that in the cy-
tosol. As noted above, in mammals, the mitochondri-
al matrix apparently lacks adenylate kinase, which
could regenerate ADP from ATP and AMP. However,
AK3 is capable of converting AMP to ADP using GTP.
Mammalian mitochondria lack a carrier protein that
transports GTP across the inner membrane, so the
mitochondrial GTP and GDP pools can be considered
isolated from the cytosolic pool on a short timescale.
GTP in the matrix is synthesized by the succinyl-CoA
synthetase, an enzyme of the TCA cycle. Notably, ani-
mal mitochondria have two isoforms of this enzyme:
one synthesizes GTP and the other ATP [42]. The rel-
ative abundance of each isoform varies among dif-
ferent tissues.
GTP in mitochondria is mainly required for
RNA and protein synthesis, AMP rephosphorylation
by adenylate kinase AK3, and for conversion of ox-
aloacetate to phosphoenolpyruvate by mitochondrial
phosphoenolpyruvate carboxykinase (PEPCK-M) [43].
Nucleoside diphosphate kinases (NDPKs) consti-
tute another important class of enzymes influencing
the relationship between nucleotide concentrations.
They catalyze the exchange of phosphate groups
between various nucleoside di- and triphosphates,
maintaining thermodynamic equilibrium among ATP,
ADP, and other nucleotides in the cytosol. However,
in human cells, only one nucleoside diphosphate ki-
nase, namely NDPK D, has a mitochondrial target-
ing sequence [44, 45]. This enzyme is present in the
cell in significant amounts; proteomic studies of the
HEK293T cells report up to 200,000 NDPKD molecules
per cell [9]. Most experimental data suggest that it is
localized in the intermembrane space and is an-
chored to the inner mitochondrial membrane [46].
(Some studies, however, report nucleoside diphos-
phate kinase activity in the mitochondrial matrix of
vertebrate cells [47, 48]). Additional support for the
absence of NDPK in the matrix comes from the ex-
periments on baker’s yeast. Yeast normally possess-
es the mitochondrial GTP/GDP transporter, which is
absent in the vertebrate mitochondria. After genetic
deletion of this transporter, it was necessary to ex-
press human NDPK with an added matrix localiza-
tion signal to compensate for the deletion [49]. This
observation implies that yeast mitochondria cannot
produce GTP from ATP without NDPK artificially in-
troduced into the matrix.
We have not found studies that address the re-
lationship between guanine and adenine nucleotide
concentrations in the mitochondrial matrix, or that
report whether the adenylate kinase-driven thermo-
dynamic equilibrium observed in the cytosol is like-
wise sustained in the matrix.
Accurate quantification of nucleotide concentra-
tions in the mitochondrial matrix is crucial for eluci-
dating the mitochondrial contribution to the cellular
energy metabolism and its regulation. However, such
measurements are rather difficult, primarily because
of the substantial methodological difficulties. HPLC
and other analytical approaches provide highly ac-
curate measurements of total nucleotide content in
the sample. Yet, when applied to intact cells, they
cannot resolve relative contributions of the cytoso-
lic and mitochondrial pools. Since the mitochondrial
matrix usually occupies only a minor fraction of the
cellular volume, the resulting values predominantly
reflect cytosolic ATP, offering minimal information
about its concentration in the matrix.
Measuring nucleotide levels in the isolated mi-
tochondria provides only limited insight into in  vivo
concentrations of these nucleotides, as the isolation
process subjects mitochondria to substantial stress
and non-physiological conditions. The ATP/ADP/AMP
ratios measured in the isolated mitochondria vary
depending on biological source and experimental
ATP IN MITOCHONDRIA 1935
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
Table 1. Relative proportions of adenine nucleotides in the isolated mitochondrial preparations from bovine
heart and rat liver
Source ATP, % ADP, % AMP, % AEC Reference
Bovine heart mitochondria
12 30 58 0.27 [50]
39 33 28 0.56 [51]
Rat liver mitochondria
46 29 25 0.61 [52]
35 50 15 0.6 [16]
from 10 to 46 from 47 to 75 from 8 to 16 [53]
20 69 11 0.55 [31]
Note. Absolute nucleotide concentrations were not determined in these studies, with the exception of reference [31]. TheAEC
was calculated as (ATP + ½ADP)/(ATP + ADP + AMP).
conditions, yet they consistently diverge from the
cytosolic values and typically correspond to an AEC
level of approximately 0.6 or less (Table 1).
Furthermore, when isolated mitochondria are
incubated in the media containing phosphate and
magnesium, adenine nucleotides leak out rapidly.
This loss depends on the ATP/ADP/AMP ratio and
could reach as much as 75% of the total nucleotide
pool [52], implying that a substantial and variable
fraction of adenine nucleotides could be lost during
mitochondrial isolation. In light of these observations,
the relative nucleotide contents measured in the iso-
lated mitochondria seem to provide an unreliable ba-
sis for evaluating their matrix concentrations in vivo.
Nevertheless, as discussed below, available in  vivo es-
timates align reasonably well with the values shown
in Table 1.
One of the few studies that attempted to quantify
the in  vivo ratios of ATP, ADP, and their magnesium
complexes in the cytosol and mitochondria exam-
ined cambial cells of white maple (Acer pseudoplata-
nusL.)[54]. Using the
31
P-NMR spectra of living cells,
cell extracts, and isolated mitochondria, the authors
estimated nucleotide levels from the areas under the
peaks corresponding to the γ-phosphate of ATP and
the β-phosphate of ADP. The proportions of free nu-
cleotides and their Mg
2+
complexes were determined
from the chemical shifts(δ), which depend on pH and
Mg
2+
concentration. ATP concentrations in the cytosol
and mitochondrial matrix were reported to be simi-
lar – approximately 450  μM and 530  μM, respective-
ly – whereas ADP concentrations were about 30  μM
in the cytosol and 220  μM in the matrix. Thus, the
cytosolic ATP/ADP ratio was several-fold higher than
that in the matrix. It should be noted, however, that
the cytosolic ATP and ADP concentrations reported in
this study are far lower than those characteristic of
animal cells and yeast, in which ATP and ADP lev-
els are roughly 5  mM and 0.5 mM, respectively [39].
Moreover, the nucleotide concentrations reported in
the study were not validated by the independent
methods such as HPLC or luciferase assays, raising
concerns about absolute accuracy of these measure-
ments.
The authors also reported that most ATP exists as
a magnesium complex in both compartments – 88% in
the cytosol and 98% in the mitochondrial matrix [54].
In contrast, ADP was found to be predominantly free
in the cytosol (71%) but largely magnesium-bound in
the matrix (80%), implying that Mg
2+
concentration is
roughly an order of magnitude higher in the matrix
than in the cytosol (2.4  mM vs. 250  μM).
Cytosolic ATP levels in the white maple cambial
cells proved to be sensitive to metabolic context. They
rose severalfold after exposure to excess adenine and
decreased when cells were transferred into a phos-
phate-depleted medium or incubated with glycerol or
choline, which are rapidly phosphorylated. In con-
trast, the cytosolic ADP levels remained comparative-
ly stable. From this, the authors suggested that free
ADP in the cytosol acts as a central regulator of respi-
ration and cellular energy metabolism [54]. However,
considering the limited dynamic range of cytosolic
ADP permitted by adenylate kinase (Fig.  2) and the
well-established role of AMP as a principal metabolic
regulator, further experimental support might be nec-
essary to confirm this interpretation.
Finally, the authors of [54] argued that the com-
paratively low Mg
2+
concentration in the cytosol, rel-
ative to that of the mitochondrial matrix, favors the
presence of ADP in its free form, thereby promoting
its uptake by ANT and subsequent transport into the
matrix, where it participates in oxidative phosphory-
lation as MgADP.
The principal advantage of the NMR approach
used in this study is its ability to quantify adenine
LAPASHINA et al.1936
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
nucleotides in living cells. However, the method has
significant limitations, including technical complexity,
requirement for large amounts of biological material,
and low sensitivity to ADP and AMP.
Alternative approaches for assessing adenine
nucleotide levels in animal cells rely on rapid frac-
tionation to obtain a mitochondrial-enriched fraction
within seconds. One method, reported effective for
rat hepatocytes, uses brief exposure (20-40  s) to dig-
itonin, followed by rapid centrifugation (20  s) [55].
Digitonin selectively disrupts the sterol-rich plasma
membrane while leaving the sterol-poor inner mi-
tochondrial membrane intact, which allows, after
centrifugation, to obtain a supernatant containing
cytosolic nucleotides and a pellet containing in-
tact mitochondria. Using this technique, ATP, ADP,
and AMP proportions of 45-75%, 28-31%, and 3-21%
were reported in the mitochondria from rat hepa-
tocytes [55, 56].
A different rapid-fractionation technique subject-
ed rat liver cell suspensions to shear forces by pass-
ing them through a narrow metal needle, rupturing
the plasma membrane while leaving most mitochon-
dria intact [57]. The complete workflow – including
disruption and centrifugation through silicone oil into
fixative – required less than one minute. Using this
method, the ATP/ADP ratio was found to be ~7 in the
cytosolic fraction and close to 1 in the mitochondrial
fraction.
Another study from the mid-1970s reported an
intriguing and unconventional method for determin-
ing adenine nucleotide composition in rat liver mito-
chondria [58]. The technique involved rapid freezing
of the tissue, its homogenization in the frozen state,
lyophilization, and fractionation using a carbon tet-
rachloride–heptane solvent system. Despite appearing
at first unsuitable for separating membrane-bound
organelles, the procedure proved effective: electron
microscopy and enzyme assays confirmed good sep-
aration of cytosolic and mitochondrial fractions. The
authors found ATP, ADP, and AMP proportions of
roughly 85%, 14%, and 1% in the cytosol and 27%,
47%, and 26% in the mitochondria, respectively. Nota-
bly, these ratios correspond well to the matrix values
observed in the isolated mitochondria (Table  1) and
to the cytosolic concentrations determined by other
methods.
The reliability of this new method was further
verified on isolated rat hepatocytes by direct com-
parison between the digitonin permeabilization pro-
tocol and the freezing-lyophilization-organic solvent
fractionation. When both methods were applied to
the same hepatocyte sample, they yielded the same
values for mitochondrial adenine nucleotide content
within experimental error [59]. A related study ana-
lyzing pre- and postnatal rat liver mitochondria with
the organic solvent fractionation method reported the
following values for the matrix adenine nucleotides:
25%, 32%, and 43% for ATP, ADP, and AMP in adult
animals, and 21%, 19%, and 60% in the three-day-old
pups [60].
A combination of the rapid freezing, lyophiliza-
tion, and organic solvent fractionation method with
an ultra-fast (3-second) procedure for extracting rat
liver allowed estimation of the ATP/ADP ratio in mi-
tochondria in vivo. In the control group of rats, this
ratio was 0.92  ±  0.18; in the rats fasted for 48  h be-
fore liver extraction, it increased to 1.04  ±  0.03, and
in the rats subjected to anesthesia, it decreased to
0.55  ±  0.06 (pentobarbital) or 0.86  ±  0.04 (ketamine)
[61]. These data also agree with the experimental es-
timates of the ATP, ADP, and AMP ratios in the isolat-
ed rat liver mitochondria shown in Table  1.
Estimates of the ratio of ATP, ADP, and AMP con-
centrations in the mitochondrial matrix appear to
be reliable, since many experiments using different
objects and different methods yield similar results.
At the same time, the absolute values of these con-
centrations remain difficult to assess. The range of
absolute values reported in the literature is quite
broad (Table 2).
In the study [62], an attempt was made to quan-
titatively evaluate the content of a wide range of me-
tabolites in the mitochondrial matrix of HeLa cells.
To quickly and selectively separate mitochondria
from other cellular components, the authors used an
immunoaffinity method, capturing mitochondria by
an epitope of a recombinant protein localized in the
outer mitochondrial membrane. To quantify metabo-
lites in the mitochondrial preparations, the authors
employed liquid chromatography-mass spectrometry
(LC/MS) and estimated total mitochondrial matrix
volume per cell – required for calculating absolute
concentrations – using confocal microscopy. The ade-
nine nucleotide concentrations reported in this work
(Table  2, bottom row), however, raise substantial
concerns. Most experimental data, including mea-
surements obtained with fluorescent protein sensors
[66-69], consistently show that the ATP levels in the
matrix are in the millimolar range, nearly two orders
of magnitude higher than the 50  μM value reported
in [62]. A likely explanation is an error in the esti-
mation of mitochondrial volume, a parameter that is
difficult to determine precisely.
Taken together, the experimental evidence in-
dicates that in the cytosol of mammalian cells, ATP,
ADP, and AMP concentrations are constrained by the
adenylate kinase equilibrium and typically are esti-
mated as approximately 5  mM, 0.5  mM, and 0.05  mM,
respectively, yielding an AEC of about 0.95. In the
mitochondrial matrix, the total adenine nucleotide
concentration is likewise in the millimolar range,
ATP IN MITOCHONDRIA 1937
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
Table 2. Experimental estimates of absolute adenine nucleotide concentrations in the mitochondrial matrix
Source ATP, mM ADP, mM AMP, mM AEC Reference
Guinea pig heart 5.6-7.2 1.4-2.0 ND [63]
Rat liver 3.5 12.4 1.9 0.54 [31]
Rat liver (fasted rats) 5.2 7.8 1.5 0.63 [64]
Rat liver (fed rats) 1.4 7.8 3.8 0.41 [64]
Rat liver 12.5-16.4 12.0-14.1 ND [57]
Rat liver 10.4 5.9 4.3 0.65 [56]
Rat; renal proximal tubular cells 2.62 ND ND [65]
HeLa cells 0.05 0.07 0.02 0.61 [62]
Note. In the studies cited, concentrations were calculated using approximate estimates of matrix volume per milligram of
protein or per milligram of dry weight, except in the study [62], where the mitochondrial volume was derived from the
microscopy data. AEC denotes the adenylate energy charge; ND indicates cases where no data were available.
but the ATP/ADP/AMP ratios are more variable and
the AEC is lower – generally between 0.4 and 0.65.
It is widely assumed that in the resting state,
when cellular ATP demand is low, the adenylate ener-
gy charge in the mitochondrial matrix rises and ADP
becomes nearly fully phosphorylated to ATP. In this
scenario, ATP synthase is unable either to continue
ATP production or to consume pmf through proton
translocation coupled to ATP synthesis. As a result,
pmf increases and suppresses the activity of respira-
tory chain enzymes, producing the so-called ‘state  4’
[70], extensively characterized in vitro in isolated mi-
tochondria. Inhibition of respiration elevates the re-
duction state of the respiratory electron carriers and
increases the intracellular O
2
levels, thereby promot-
ing non-enzymatic formation of reactive oxygen spe-
cies (ROS) and increasing the risk of oxidative dam-
age [71]. Vladimir Skulachev proposed that, in order
to limit ROS production under these conditions, cells
increase proton conductivity of the inner mitochon-
drial membrane (‘mild uncoupling’), thereby lowering
pmf, stimulating respiration, and reducing O
2
concen-
tration [72].
The evidence summarized in this review, howev-
er, indicates that such complete depletion of matrix
ADP – and thus a true state  4 – is unlikely to occur
in vivo. All available data show that the matrix ADP
concentration remains appreciable and is consistently
higher than the cytosolic one, which itself remains
remarkably stable at ~0.5  mM – more than sufficient
to support continuous ATP synthesis by ATP synthase.
Accordingly, the processes grouped under the term
‘mild uncoupling’ likely serve a physiological role
distinct from the suppression of mitochondrial ROS
production in the resting state.
APPLICATION OF FLUORESCENT
PROTEIN SENSORS TO MEASURING
ATP CONCENTRATION IN MITOCHONDRIA
A powerful strategy for quantifying ATP levels
within both mitochondria and cytosol is the use of
fluorescent protein-based sensors whose spectral
characteristics shift upon ATP binding. The first such
sensor, ATeam, was developed by Hiromi Imamura
and colleagues [69]. ATeam comprises yellow and
cyan fluorescent proteins joined by a small protein
that undergoes a reversible conformational transi-
tion in response to ATP binding. This structural re-
arrangement decreases the distance between the flu-
orophores and enhances Förster resonance energy
transfer (FRET). Consequently, fluorescence spectral
measurements permit accurate monitoring of relative
abundance of the free versus ATP-bound forms of
the probe. A major advantage of this method is that
its readout is independent of probe concentration.
Application of ATeam in HeLa cells revealed mito-
chondrial ATP concentrations in the millimolar range,
although consistently lower than those observed in
the cytosol [69].
Further development of this approach substan-
tiated the initial observation and enabled real-time
assessment of ATP concentration changes within the
mitochondrial matrix in response to various stimuli
[66-68]. Parallel observations of a mitochondria-tar-
geted ATeam probe and its cytosolic counterpart
in cardiomyocytes demonstrated that hypoxia pro-
duces a substantial decline in the matrix ATP over
several hours, whereas cytosolic ATP levels remain
nearly constant. A similar pattern was observed un-
der normoxia following the addition of oligomycin,
LAPASHINA et al.1938
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
an ATP synthase inhibitor, although in this case the
signal changes were developing within minutes [67].
In another study, ATeam probes were employed
to analyze ATP concentration dynamics in the cyto-
sol and mitochondria of HeLa cells following glucose
depletion [73]. Under these conditions, cytosolic ATP
levels remained largely stable, whereas the mito-
chondrial matrix exhibited a transient elevation in
ATP followed by a marked decline. The authors pro-
posed that the initial rise in mitochondrial ATP level
results from the activity of hexokinases associated
with the outer mitochondrial membrane. Although
the hexokinase-catalyzed phosphorylation of glucose
to glucose-6-phosphate is generally regarded as ir-
reversible under physiological conditions, the au-
thors speculate that rapid mitochondrial ATP uptake
in the absence of cytosolic glucose could make the
reverse reaction feasible. This hypothesis, however,
requires further confirmation. Another study com-
bined the fluorescent dye TMRE with a mitochon-
dria-targeted ATeam probe to assess the correlation
between pmf and ATP levels in the neuronal mito-
chondria [74].
ATeam sensors have also been applied success-
fully to monitor mitochondrial ATP dynamics in in-
tact tissues and even whole organs. Analysis using
a mitochondria-targeted probe in the mouse hearts
demonstrated that empagliflozin, an inhibitor of the
sodium–glucose cotransporter 2 (SGLT2), elevates mi-
tochondrial ATP levels [75].
In our laboratory, an ATeam-derived sensor was
successfully employed to quantify ATP concentrations
in yeast cells and in isolated mitochondria. We select-
ed the yAT1.03 variant, optimized for expression in
Saccharomyces cerevisiae and exhibiting reduced pH
sensitivity relative to the original ATeam sensors [76].
We isolated mitochondria from the yeast expressing
this probe fused to the mitochondrial localization sig-
nal. Permeabilization with the channel-forming anti-
biotic alamethicin – which renders the inner mem-
brane permeable to low-molecular-weight solutes (up
to 1.5 kDa) but not to proteins [77] – enabled cali-
bration of the matrix yAT1.03 signal by adding ATP
at defined concentrations to the incubation medium.
These experiments demonstrated that ATP concen-
tration in the matrix of isolated yeast mitochondria
increases to millimolar levels upon the addition of
respiratory substrates [78]. This method also appears
promising for quantitative in  vitro studies of intra-
mitochondrial ATP concentration dynamics under
various conditions.
Besides the FRET-based ATeam family, additional
protein sensors comprising a single fluorescent pro-
tein with ATP-dependent spectral shifts have been de-
veloped. These ‘single-color’ sensors include QUEEN
[79] and MaLion [80]. QUEEN has been applied with
moderate success to quantify ATP levels in yeast mi-
tochondria [81], where – consistent with observations
in HeLa cells – mitochondrial ATP was found to be
slightly lower than cytosolic ATP. Another study em-
ploying QUEEN demonstrated that, in yeast, addition
of the protonophore FCCP (carbonyl cyanide p-trifluo-
romethoxyphenylhydrazone), which collapses the pmf
across the inner mitochondrial membrane, leads to
an increase in ATP within the matrix and a decrease
in the cytosol [82]. To account for this unexpected
behavior, the authors proposed that FCCP stimulates
ATP import from the cytosol into mitochondria via
the ATP-Mg
2+
/HPO
4
2−
carrier, although the presented
data provide insufficient evidence to support this
conclusion. In a separate investigation, use of the
mitochondria-targeted MaLion probe revealed that
when hepatocytes transition from glucose depriva-
tion to active glucose utilization, the mitochondrial
ATP concentration declines, implying a potential role
of the mitochondrial adenine-nucleotide dynamics
in regulating cellular metabolism [83].
Another family of fluorescent probes includes
the Perceval sensor and its analogs. These proteins
have very high affinity for both ATP and ADP, so in-
side the cells they are always bound to one of these
nucleotides. The spectral properties of Perceval dif-
fer depending on which nucleotide is bound, and
the overall fluorescence signal reliably reflects the
ATP/ADP ratio [84]. However, the signal of this probe
is sensitive to pH changes between 6 and 8, which
makes quantitative interpretation more difficult.
A later variant, PercevalHR, was developed with im-
proved spectral characteristics [85], but we found no
reports of its successful use for measuring ATP/ADP
ratios in mitochondria. Recently, a protein fluorescent
probe capable of detecting the GTP/GDP ratio in mi-
tochondria was developed [86].
Although the studies described above demon-
strate varying degrees of success in applying the
protein-based fluorescent sensors to mitochondrial
ATP measurements, several substantial methodologi-
cal limitations remain. All such probes are influenced
to some extent by pH and by ionic composition, dis-
play partial cross-reactivity with other nucleotides,
and exhibit temperature-dependent fluorescence be-
havior. Moreover, probes of the ATeam family are
subject to asynchronous maturation of their two
fluorescent proteins, potentially leading to the accu-
mulation of an unknown – yet possibly considerable
– population of non-functional probe species within
the cell. These aberrant molecules lack either the
FRET donor or acceptor and could, therefore, com-
promise the composite signal. Consequently, rigorous
calibration of the probe output directly in the exper-
imental sample (e.g., through permeabilization fol-
lowed by the addition of defined ATP concentrations)
ATP IN MITOCHONDRIA 1939
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
and cross-validation of ATP measurements using in-
dependent approaches (distinct fluorescent probes,
luciferin–luciferase assays, HPLC) are critically im-
portant.
CONCLUSION
When discussing adenine nucleotide levels in eu-
karyotic cells, it can be stated that under most condi-
tions, cytosolic ATP concentration lies in the millimo-
lar range, while the concentrations of ADP and AMP
are approximately 10-fold and 100-fold lower, respec-
tively. In the mitochondrial matrix, the proportion of
ATP is lower and the proportions of ADP and AMP
are higher compared with the cytosol; this indicates
that the so-called ‘state 4’, in which oxidative phos-
phorylation is inhibited in the resting cells due to the
lack of ADP in the matrix, is very unlikely to occur
in vivo under physiological conditions.
To clarify causal relationships between the ac-
tivities of bioenergetic enzymes, changes in pmf, and
adenine nucleotide concentrations in the mitochon-
drial matrix and cytosol, reliable methods are needed
for quantitative measurements of ATP concentrations
invivo. Fluorescent protein sensors are excellent tools
for monitoring dynamic changes in ATP concentration
in the cytosol and mitochondria, but they cannot pro-
vide precise values for absolute nucleotide concen-
trations because their calibration in  vivo is difficult.
Therefore, an optimal strategy is to combine dynamic
measurements obtained with several different fluores-
cent protein probes – calibrated using standard ATP
solutions after cell permeabilization – with tradition-
al nucleotide quantification methods (HPLC, lucifer-
in–luciferase assays) and newer analytical approaches
such as single-cell metabolomics, mass-spectrometry
imaging, and high-resolution respirometry within
the same experimental framework.
Abbreviations
AEC adenylate energy charge
AMPK AMP-activated protein kinase
FRET Förster resonance energy transfer
NDPK nucleoside diphosphate kinase
pmf proton motive force
TCA tricarboxylic acid cycle
Acknowledgments
The authors thank Dr. Dmitry Knorre for valuable
discussions and advice during the preparation of this
manuscript, and one of the reviewers for suggesting
that we address whether the low AEC values in the
matrix are compatible with the hypothesis of ‘mild un-
coupling’ as a protective mechanism against oxidative
stress.
Funding
The study was conducted under the State assign-
ment of Lomonosov Moscow State University
(AAAA-A19-119031390114-5) and financially supported
by the Russian Science Foundation (grant no. 20-14-
00268).
Ethics approval and consent to participate
This work does not contain any studies involving hu-
man and 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. Zheng, W., Chai, P., Zhu, J., and Zhang, K. (2024)
High-resolution in situ structures of mammalian
respiratory supercomplexes, Nature, 631, 232-239,
https://doi.org/10.1038/s41586-024-07488-9.
2. Burgess, D. R. (2004) NIST SRD 46. Critically Selected
Stability Constants of Metal Complexes: Version 8.0 for
Windows, National Institute of Standards and Tech-
nology [Internet], https://doi.org/10.18434/M32154.
3. Maeshima, K., Matsuda, T., Shindo, Y., Imamura, H.,
Tamura, S., Imai, R., Kawakami, S., Nagashima, R.,
Soga,T., Noji,H., Oka,K., and Nagai,T. (2018) A tran-
sient rise in free Mg
2+
ions released from ATP-Mg
hydrolysis contributes to mitotic chromosome con-
densation, Curr. Biol., 28, 444-451.e6, https://doi.org/
10.1016/j.cub.2017.12.035.
4. Babot, M., Birch, A., Labarbuta, P., and Galkin, A.
(2014) Characterisation of the active/de-active tran-
sition of mitochondrial complex I, Biochim. Biophys.
Acta, 1837, 1083-1092, https://doi.org/10.1016/j.bbabio.
2014.02.018.
LAPASHINA et al.1940
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
5. Zubareva, V. M., Lapashina, A. S., Shugaeva, T. E.,
Litvin, A.V., and Feniouk, B.A. (2020) Rotary ion-trans-
locating ATPases/ATP synthases: diversity, similar-
ities, and differences, Biochemistry (Moscow), 85,
1613-1630, https://doi.org/10.1134/S0006297920120135.
6. Klingenberg, M. (2008) The ADP and ATP trans-
port in mitochondria and its carrier, Biochim. Bio-
phys. Acta, 1778, 1978-2021, https://doi.org/10.1016/
j.bbamem.2008.04.011.
7. Penin, F., Deléage, G., Godinot, C., and Gautheron,
D. C. (1986) Efficient reconstitution of mitochondri-
al energy-transfer reactions from depleted mem-
branes and F1-ATPase as a function of the amount
of bound oligomycin sensitivity-conferring protein
(OSCP), Biochim. Biophys. Acta, 852, 55-67, https://
doi.org/10.1016/0005-2728(86)90056-3.
8. Matsuno-Yagi, A., and Hatefi, Y. (1988) Estimation of
the turnover number of bovine heart F0F1 complexes
for ATP synthesis, Biochemistry, 27, 335-340, https://
doi.org/10.1021/bi00401a050.
9. Morgenstern,M., Peikert, C.D., Lübbert,P., Suppanz,I.,
Klemm, C., Alka, O., Steiert, C., Naumenko, N.,
Schendzielorz,A., Melchionda,L., Mühlhäuser, W.W.D.,
Knapp, B., Busch, J. D., Stiller, S. B., Dannenmaier, S.,
Lindau, C., Licheva, M., Eickhorst, C., Galbusera, R.,
Zerbes, R.M., Ryan, M.T., Kraft,C., Kozjak-Pavlovic,V.,
Drepper,F., Dennerlein,S., Oeljeklaus,S., Pfanner,N.,
Wiedemann,N., and Warscheid,B. (2021) Quantitative
high-confidence human mitochondrial proteome
and its dynamics in cellular context, Cell Metab.,
33, 2464-2483.e18, https://doi.org/10.1016/j.cmet.
2021.11.001.
10. Lapashina, A. S., and Feniouk, B. A. (2018) ADP-In-
hibition of H+-FOF1-ATP Synthase, Biochemistry
(Moscow), 83, 1141-1160, https://doi.org/10.1134/
S0006297918100012.
11. Campanella, M., Casswell, E., Chong, S., Farah, Z.,
Wieckowski, M. R., Abramov, A. Y., Tinker, A., and
Duchen, M. R. (2008) Regulation of mitochondri-
al structure and function by the F1Fo-ATPase in-
hibitor protein, IF1, Cell Metab., 8, 13-25, https://
doi.org/10.1016/j.cmet.2008.06.001.
12. Galkina, K.V., Zubareva, V.M., Kashko, N.D., Lapashina,
A. S., Markova, O. V., Feniouk, B. A., and Knorre,
D. A. (2022) Heterogeneity of starved yeast cells in
IF1 levels suggests the role of this protein in vivo,
Front. Microbiol., 13, 816622, https://doi.org/10.3389/
fmicb.2022.816622.
13. Lapashina, A. S., Kashko, N. D., Zubareva, V. M.,
Galkina, K. V., Markova, O. V., Knorre, D. A., and
Feniouk, B. A. (2022) Attenuated ADP-inhibition of
FOF1 ATPase mitigates manifestations of mitochondri-
al dysfunction in yeast, Biochim. Biophys. Acta, 1863,
148544, https://doi.org/10.1016/j.bbabio.2022.148544.
14. Pfaff, E., and Klingenberg, M. (1968) Adenine nu-
cleotide translocation of mitochondria. 1. Specific-
ity and control, Eur. J. Biochem., 6, 66-79, https://
doi.org/10.1111/j.1432-1033.1968.tb00420.x.
15. Winkler, H. H., Bygrave, F. L., and Lehninger, A. L.
(1968) Characterization of the atractyloside-sensi-
tive adenine nucleotide transport system in rat liv-
er mitochondria, J. Biol. Chem., 243, 20-28, https://
doi.org/10.1016/s0021-9258(18)99320-8.
16. Matsuno-Yagi,A., and Hatefi,Y. (1990) Studies on the
mechanism of oxidative phosphorylation. Positive co-
operativity in ATP synthesis, J. Biol. Chem., 265, 82-88,
https://doi.org/10.1016/S0021-9258(19)40198-1.
17. Panayiotou, C., Solaroli, N., and Karlsson, A. (2014)
The many isoforms of human adenylate kinases,
Int. J. Biochem. Cell Biol., 49, 75-83, https://doi.org/
10.1016/j.biocel.2014.01.014.
18. Fujisawa,K. (2023) Regulation of adenine nucleotide
metabolism by adenylate kinase isozymes: physio-
logical roles and diseases, Int. J. Mol. Sci., 24, 5561,
https://doi.org/10.3390/ijms24065561.
19. Miyoshi, K., Akazawa, Y., Horiguchi, T., and Noma, T.
(2009) Localization of adenylate kinase 4 in mouse
tissues, Acta Histochem. Cytochem., 42, 55-64, https://
doi.org/10.1267/ahc.08012.
20. Liu,R., Ström, A.-L., Zhai,J., Gal,J., Bao,S., Gong,W.,
and Zhu, H. (2009) Enzymatically inactive adenylate
kinase 4 interacts with mitochondrial ADP/ATP trans-
locase, Int. J. Biochem. Cell Biol., 41, 1371-1380,
https://doi.org/10.1016/j.biocel.2008.12.002.
21. Panayiotou, C., Solaroli, N., Johansson, M., and
Karlsson, A. (2010) Evidence of an intact N-terminal
translocation sequence of human mitochondrial ad-
enylate kinase 4, Int. J. Biochem. Cell Biol., 42, 62-69,
https://doi.org/10.1016/j.biocel.2009.09.007.
22. Wilson, D. E., Povey, S., and Harris, H. (1976) Ad-
enylate kinases in man: evidence for a third lo-
cus, Ann. Hum. Genet., 39, 305-313, https://doi.org/
10.1111/j.1469-1809.1976.tb00134.x.
23. Nam, K., Thodika, A. R. A., Tischlik, S., Phoeurk, C.,
Nagy, T. M., Schierholz, L., Ådén, J., Rogne, P.,
Drescher,M., Sauer-Eriksson, A.E., and Wolf-Watz,M.
(2024) Magnesium induced structural reorganization
in the active site of adenylate kinase, Sci. Adv., 10,
eado5504, https://doi.org/10.1126/sciadv.ado5504.
24. Blair, J. M. (1970) Magnesium, potassium, and
the adenylate kinase equilibrium. Magnesium as
a feedback signal from the adenine nucleotide
pool, Eur. J. Biochem., 13, 384-390, https://doi.org/
10.1111/j.1432-1033.1970.tb00940.x.
25. Gellerich, F. N. (1992) The role of adenylate ki-
nase in dynamic compartmentation of adenine
nucleotides in the mitochondrial intermembrane
space, FEBS Lett., 297, 55-58, https://doi.org/10.1016/
0014-5793(92)80326-c.
26. Tomasselli, A. G., Schirmer, R. H., and Noda, L. H.
(1979) Mitochondrial GTP-AMP phosphotransfer-
ase. 1. Purification and properties, Eur. J. Biochem.,
ATP IN MITOCHONDRIA 1941
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
93, 257-262, https://doi.org/10.1111/j.1432-1033.1979.
tb12818.x.
27. Seifert, E.L., Ligeti,E., Mayr, J.A., Sondheimer,N., and
Hajnóczky, G. (2015) The mitochondrial phosphate
carrier: role in oxidative metabolism, calcium han-
dling and mitochondrial disease, Biochem. Biophys.
Res. Commun., 464, 369-375, https://doi.org/10.1016/
j.bbrc.2015.06.031.
28. Küster, U., Bohnensack,R., and Kunz,W. (1976) Con-
trol of oxidative phosphorylation by the extra-mito-
chondrial ATP/ADP ratio, Biochim. Biophys. Acta, 440,
391-402, https://doi.org/10.1016/0005-2728(76)90073-6.
29. Davis, E. J., and Davis-van Thienen, W. I. (1978) Con-
trol of mitochondrial metabolism by the ATP/ADP ra-
tio, Biochem. Biophys. Res. Commun., 83, 1260-1266,
https://doi.org/10.1016/0006-291x(78)91357-8.
30. Lemasters, J. J., and Sowers, A. E. (1979) Phosphate
dependence and atractyloside inhibition of mitochon-
drial oxidative phosphorylation. The ADP-ATP carrier
is rate-limiting, J. Biol. Chem., 254, 1248-1251, https://
doi.org/10.1016/s0021-9258(17)34194-7.
31. Slater, E. C., Rosing, J., and Mol, A. (1973) The phos-
phorylation potential generated by respiring mi-
tochondria, Biochim. Biophys. Acta, 292, 534-553,
https://doi.org/10.1016/0005-2728(73)90003-0.
32. Groen, A.K., Wanders, R.J., Westerhoff, H.V., van der
Meer,R., and Tager, J.M. (1982) Quantification of the
contribution of various steps to the control of mito-
chondrial respiration, J. Biol. Chem., 257, 2754-2757,
https://doi.org/10.1016/s0021-9258(19)81026-8.
33. Fiermonte,G., De Leonardis,F., Todisco,S., Palmieri,L.,
Lasorsa, F. M., and Palmieri, F. (2004) Identification
of the mitochondrial ATP-Mg/Pi transporter, J. Biol.
Chem., 279, 30722-30730, https://doi.org/10.1074/
jbc.M400445200.
34. Hardie, D.G., and Hawley, S.A. (2001) AMP-activated
protein kinase: the energy charge hypothesis revisit-
ed, Bioessays, 23, 1112-1119, https://doi.org/10.1002/
bies.10009.
35. Ramaiah, A., Hathaway, J. A., and Atkinson, D. E.
(1964) Adenylate as a metabolic regulator. Ef-
fect on yeast phosphofructokinase kinetics, J. Biol.
Chem., 239, 3619-3622, https://doi.org/10.1016/S0021-
9258(18)91181-6.
36. Atkinson, D.E. (1968) Energy charge of the adenylate
pool as a regulatory parameter. Interaction with feed-
back modifiers, Biochemistry, 7, 4030-4034, https://
doi.org/10.1021/bi00851a033.
37. Hardie, D.G., Carling,D., and Sim, A.T.R. (1989) The
AMP-activated protein kinase: a multisubstrate reg-
ulator of lipid metabolism, Trends Biochem. Sci., 14,
20-23, https://doi.org/10.1016/0968-0004(89)90084-4.
38. Hardie, D. G., Ross, F. A., and Hawley, S. A. (2012)
AMPK: a nutrient and energy sensor that maintains
energy homeostasis, Nat. Rev. Mol. Cell Biol., 13,
251-262, https://doi.org/10.1038/nrm3311.
39. Park, J. O., Rubin, S.A., Xu, Y.-F., Amador-Noguez,D.,
Fan,J., Shlomi,T., and Rabinowitz, J.D. (2016) Metab-
olite concentrations, fluxes and free energies imply
efficient enzyme usage, Nat. Chem. Biol., 12, 482-489,
https://doi.org/10.1038/nchembio.2077.
40. Gauthier, S., Coulpier, F., Jourdren, L., Merle, M.,
Beck, S., Konrad, M., Daignan-Fornier, B., and
Pinson, B. (2008) Co-regulation of yeast purine and
phosphate pathways in response to adenylic nu-
cleotide variations, Mol. Microbiol., 68, 1583-1594,
https://doi.org/10.1111/j.1365-2958.2008.06261.x.
41. Traut, T.W. (1994) Physiological concentrations of pu-
rines and pyrimidines, Mol. Cell Biochem., 140, 1-22,
https://doi.org/10.1007/BF00928361.
42. Johnson, J. D., Mehus, J. G., Tews, K., Milavetz, B. I.,
and Lambeth, D. O. (1998) Genetic evidence for the
expression of ATP- and GTP-specific succinyl-CoA syn-
thetases in multicellular eucaryotes, J. Biol. Chem.,
273, 27580-27586, https://doi.org/10.1074/jbc.273.
42.27580.
43. Stark, R., Pasquel, F., Turcu, A., Pongratz, R. L.,
Roden, M., Cline, G. W., Shulman, G. I., and Kibbey,
R. G. (2009) Phosphoenolpyruvate cycling via mito-
chondrial phosphoenolpyruvate carboxykinase links
anaplerosis and mitochondrial GTP with insulin
secretion, J. Biol. Chem., 284, 26578-26590, https://
doi.org/10.1074/jbc.M109.011775.
44. Milon, L., Meyer, P., Chiadmi, M., Munier, A.,
Johansson, M., Karlsson, A., Lascu, I., Capeau, J.,
Janin, J., and Lacombe, M. L. (2000) The human
nm23-H4 gene product is a mitochondrial nucleoside
diphosphate kinase, J. Biol. Chem., 275, 14264-14272,
https://doi.org/10.1074/jbc.275.19.14264.
45. Tokarska-Schlattner, M., Boissan, M., Munier, A.,
Borot, C., Mailleau, C., Speer, O., Schlattner, U., and
Lacombe, M.-L. (2008) The nucleoside diphosphate
kinase D (NM23-H4) binds the inner mitochondri-
al membrane with high affinity to cardiolipin and
couples nucleotide transfer with respiration, J. Biol.
Chem., 283, 26198-26207, https://doi.org/10.1074/
jbc.M803132200.
46. Boissan,M., Dabernat,S., Peuchant,E., Schlattner,U.,
Lascu,I., and Lacombe, M.-L. (2009) The mammalian
Nm23/NDPK family: from metastasis control to cilia
movement, Mol. Cell Biochem., 329, 51-62, https://
doi.org/10.1007/s11010-009-0120-7.
47. Muhonen, W. W., and Lambeth, D. O. (1995) The
compartmentation of nucleoside diphosphate kinase
in mitochondria, Comp. Biochem. Physiol. B Bio-
chem. Mol. Biol., 110, 211-223, https://doi.org/10.1016/
0305-0491(94)00123-c.
48. Lacombe, M.-L., Tokarska-Schlattner, M., Boissan, M.,
and Schlattner, U. (2018) The mitochondrial nucle-
oside diphosphate kinase (NDPK-D/NME4), a moon-
lighting protein for cell homeostasis, Lab Invest, 98,
582-588, https://doi.org/10.1038/s41374-017-0004-5.
LAPASHINA et al.1942
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
49. Gordon, D. M., Lyver, E. R., Lesuisse, E., Dancis, A.,
and Pain, D. (2006) GTP in the mitochondrial matrix
plays a crucial role in organellar iron homoeosta-
sis, Biochem. J., 400, 163-168, https://doi.org/10.1042/
BJ20060904.
50. Mansurova, S. E., Drobyshev, V. I., and Kulaev, I. S.
(1972) Nucleotides of beef heart mitochondria and
submitochondrial particles, J. Bioenerg., 3, 499-507,
https://doi.org/10.1007/BF01539058.
51. Brierley,G., and O’Brien, R. L. (1965) Compartmenta-
tion of heart mitochondria, J. Biol. Chem., 240, 4532-
4539, https://doi.org/10.1016/s0021-9258(18)97095-x.
52. Meisner,H., and Klingenberg,M. (1968) Efflux of ade-
nine nucleotides from rat liver mitochondria, J. Biol.
Chem., 243, 3631-3639, https://doi.org/10.1016/S0021-
9258(19)34186-9.
53. Haynes, R.C.,Jr., Picking, R.A., and Zaks, W.J. (1986)
Control of mitochondrial content of adenine nucleo-
tides by submicromolar calcium concentrations and
its relationship to hormonal effects, J. Biol. Chem.,
261, 16121-16125, https://doi.org/10.1016/s0021-9258
(18)66686-4.
54. Gout, E., Rébeillé, F., Douce, R., and Bligny, R. (2014)
Interplay of Mg2+, ADP, and ATP in the cytosol and
mitochondria: unravelling the role of Mg2+ in cell res-
piration, Proc. Natl. Acad. Sci. USA, 111, E4560-E4567,
https://doi.org/10.1073/pnas.1406251111.
55. Zuurendonk, P.F., and Tager, J.M. (1974) Rapid separa-
tion of particulate components and soluble cytoplasm
of isolated rat-liver cells, Biochim. Biophys. Acta, 333,
393-399, https://doi.org/10.1016/0005-2728(74)90022-x.
56. Akerboom, T.P.M., Bookelman,H., Zuurendonk, P.F.,
Meer, R., and Tager, J. M. (1978) Intramitochondrial
and extramitochondrial concentrations of adenine
nucleotides and inorganic phosphate in isolated he-
patocytes from fasted rats, Eur. J. Biochem., 84, 413-
420, https://doi.org/10.1111/j.1432-1033.1978.tb12182.x.
57. Tischler, M. E., Hecht,P., and Williamson, J.R. (1977)
Determination of mitochondrial/cytosolic metabolite
gradients in isolated rat liver cells by cell disrup-
tion, Arch. Biochem. Biophys., 181, 278-293, https://
doi.org/10.1016/0003-9861(77)90506-9.
58. Elbers, R., Heldt, H. W., Schmucker, P., Soboll, S.,
and Wiese, H. (1974) Measurement of the ATP/ADP
ratio in mitochondria and in the extramitochondri-
al compartment by fractionation of freeze-stopped
liver tissue in non-aqueous media, Hoppe Seylers Z.
Physiol. Chem., 355, 378-393, https://doi.org/10.1515/
bchm2.1974.355.1.378.
59. Sutton, R., and Pollak, J. K. (1978) The increasing
adenine nucleotide concentration and the mat-
uration of rat liver mitochondria during neona-
tal development, Differentiation, 12, 15-21, https://
doi.org/10.1111/j.1432-0436.1979.tb00985.x.
60. Schwenke, W. D., Soboll, S., Seitz, H. J., and Sies, H.
(1981) Mitochondrial and cytosolic ATP/ADP ratios
in rat liver in vivo, Biochem J, 200, 405-408, https://
doi.org/10.1042/bj2000405.
61. Soboll,S., Akerboom, T.P., Schwenke, W.D., Haase,R.,
and Sies, H. (1980) Mitochondrial and cytosolic ATP/
ADP ratios in isolated hepatocytes. A comparison of
the digitonin method and the non-aqueous fraction-
ation procedure, Biochem. J., 192, 951-954, https://
doi.org/10.1042/bj1920951.
62. Chen, W. W., Freinkman,E., Wang, T., Birsoy, K., and
Sabatini, D.M. (2016) Absolute quantification of ma-
trix metabolites reveals the dynamics of mitochon-
drial metabolism, Cell, 166, 1324-1337.e11, https://
doi.org/10.1016/j.cell.2016.07.040.
63. Soboll,S., and Bünger,R. (1981) Compartmentation of
adenine nucleotides in the isolated working guinea
pig heart stimulated by noradrenaline, Hoppe Seylers
Z. Physiol. Chem., 362, 125-132, https://doi.org/10.1515/
bchm2.1981.362.1.125.
64. Soboll, S., Scholz, R., and Heldt, H. W. (1978) Subcel-
lular metabolite concentrations. Dependence of mito-
chondrial and cytosolic ATP systems on the metabolic
state of perfused rat liver, Eur J Biochem, 87, 377-390,
https://doi.org/10.1111/j.1432-1033.1978.tb12387.x.
65. Pfaller,W., Guder, W.G., Gstraunthaler,G., Kotanko,P.,
Jehart, I., and Pürschel, S. (1984) Compartmentation
of ATP within renal proximal tubular cells, Biochim.
Biophys. Acta, 805, 152-157, https://doi.org/10.1016/
0167-4889(84)90162-9.
66. Yoshida,T., Alfaqaan,S., Sasaoka,N., and Imamura,H.
(2017) Application of FRET-based biosensor “ATeam”
for visualization of ATP levels in the mitochondri-
al matrix of living mammalian cells, Methods Mol.
Biol., 1567, 231-243, https://doi.org/10.1007/978-1-
4939-6824-4_14.
67. Kioka,H., Kato,H., Fujikawa,M., Tsukamoto,O., Suzu-
ki,T., Imamura,H., Nakano,A., Higo,S., Yamazaki,S.,
Matsuzaki,T., Takafuji,K., Asanuma,H., Asakura,M.,
Minamino,T., Shintani,Y., Yoshida,M., Noji,H., Kitaka-
ze,M., Komuro,I., Asano,Y., and Takashima,S. (2014)
Evaluation of intramitochondrial ATP levels identifies
G0/G1 switch gene 2 as a positive regulator of oxida-
tive phosphorylation, Proc. Natl. Acad. Sci. USA, 111,
273-278, https://doi.org/10.1073/pnas.1318547111.
68. Morciano,G., Imamura,H., Patergnani,S., Pedriali,G.,
Giorgi,C., and Pinton, P. (2020) Measurement of ATP
concentrations in mitochondria of living cells using
luminescence and fluorescence approaches, Meth-
ods Cell Biol., 155, 199-219, https://doi.org/10.1016/
bs.mcb.2019.10.007.
69. Imamura, H., Nhat, K. P. H., Togawa, H., Saito, K.,
Iino,R., Kato-Yamada,Y., Nagai,T., and Noji,H. (2009)
Visualization of ATP levels inside single living cells
with fluorescence resonance energy transfer-based
genetically encoded indicators, Proc. Natl. Acad.
Sci. USA, 106, 15651-15656, https://doi.org/10.1073/
pnas.0904764106.
ATP IN MITOCHONDRIA 1943
BIOCHEMISTRY (Moscow) Vol. 90 No. 12 2025
70. Gnaiger, E., Aasander Frostner, E., Abdul Karim, N.,
Abdel-Rahman, E. A., Abumrad, N. A., Acuna-Castro-
viejo,D., Adiele, R.C., etal. (2020) Mitochondrial phys-
iology, Bioenerg. Commun., https://doi.org/10.26124/
bec:2020-0001.v1.
71. Korshunov, S. S., Skulachev, V. P., and Starkov, A. A.
(1997) High protonic potential actuates a mechanism
of production of reactive oxygen species in mitochon-
dria, FEBS Lett., 416, 15-18, https://doi.org/10.1016/
s0014-5793(97)01159-9.
72. Skulachev, V. P. (1996) Role of uncoupled and
non-coupled oxidations in maintenance of safely
low levels of oxygen and its one-electron reductants,
Q. Rev. Biophys., 29, 169-202, https://doi.org/10.1017/
s0033583500005795.
73. Depaoli, M. R., Karsten, F., Madreiter-Sokolowski,
C. T., Klec, C., Gottschalk, B., Bischof, H., Eroglu, E.,
Waldeck-Weiermair,M., Simmen,T., Graier, W.F., and
Malli, R. (2018) Real-time imaging of mitochondrial
ATP dynamics reveals the metabolic setting of single
cells, Cell Rep., 25, 501-512.e3, https://doi.org/10.1016/
j.celrep.2018.09.027.
74. Suzuki, R., Hotta, K., and Oka, K. (2018) Transitional
correlation between inner-membrane potential and
ATP levels of neuronal mitochondria, Sci. Rep., 8,
2993, https://doi.org/10.1038/s41598-018-21109-2.
75. Choi, J., Matoba, N., Setoyama, D., Watanabe, D.,
Ohnishi, Y., Yasui, R., Kitai, Y., Oomachi, A.,
Kotobuki, Y., Nishiya, Y., Pieper, M. P., Imamura, H.,
Yanagita, M., and Yamamoto, M. (2023) The SGLT2
inhibitor empagliflozin improves cardiac energy
status via mitochondrial ATP production in diabetic
mice, Commun. Biol., 6, 278, https://doi.org/10.1038/
s42003-023-04663-y.
76. Botman,D., van Heerden, J.H., and Teusink,B. (2020)
An improved ATP FRET sensor for yeast shows het-
erogeneity during nutrient transitions, ACS Sens., 5,
814-822, https://doi.org/10.1021/acssensors.9b02475.
77. Gostimskaya, I.S., Grivennikova, V.G., Zharova, T.V.,
Bakeeva, L. E., and Vinogradov, A. D. (2003) In situ
assay of the intramitochondrial enzymes: use of ala-
methicin for permeabilization of mitochondria, Anal.
Biochem., 313, 46-52, https://doi.org/10.1016/s0003-
2697(02)00534-1.
78. Lapashina, A. S., Galkina, K. V., Markova, O. V.,
Knorre, D. A., and Feniouk, B.A. (2023) Observations
of ATP levels in mitochondria and whole yeast cells
using fluorescent probes, in Receptors and Intracellu-
lar Signaling Collected Articles (Berezhnov A. V., and
Zinchenko, V. P., eds) pp. 352-357 [in Russian].
79. Yaginuma,H., Kawai, S., Tabata, K.V., Tomiyama, K.,
Kakizuka, A., Komatsuzaki, T., Noji, H., and Imamu-
ra, H. (2014) Diversity in ATP concentrations in a
single bacterial cell population revealed by quanti-
tative single-cell imaging, Sci. Rep., 4, 6522, https://
doi.org/10.1038/srep06522.
80. Arai,S., Kriszt,R., Harada,K., Looi, L.-S., Matsuda,S.,
Wongso, D., Suo, S., Ishiura, S., Tseng, Y.-H.,
Raghunath, M., Ito, T., Tsuboi, T., and Kitaguchi, T.
(2018) RGB-color intensiometric indicators to visual-
ize spatiotemporal dynamics of ATP in single cells,
Angew. Chem. Int. Ed. Engl., 57, 10873-10878, https://
doi.org/10.1002/anie.201804304.
81. Takaine,M., Ueno,M., Kitamura,K., Imamura,H., and
Yoshida, S. (2019) Reliable imaging of ATP in living
budding and fission yeast, J. Cell Sci., 132, jcs230649,
https://doi.org/10.1242/jcs.230649.
82. Sawai, A., Taniguchi, T., Noguchi, K., Seike, T.,
Okahashi,N., Takaine,M., and Matsuda,F. (2025) ATP
supply from cytosol to mitochondria is an addition-
al role of aerobic glycolysis to prevent programmed
cell death by maintenance of mitochondrial mem-
brane potential, Metabolites, 15, 461, https://doi.org/
10.3390/metabo15070461.
83. Tsuno, S., Harada, K., Horikoshi, M., Mita, M.,
Kitaguchi, T., Hirai, M. Y., Matsumoto, M., and
Tsuboi, T. (2024) Mitochondrial ATP concentration
decreases immediately after glucose administration
to glucose-deprived hepatocytes, FEBS Open Bio, 14,
79-95, https://doi.org/10.1002/2211-5463.13744.
84. Berg, J., Hung, Y. P., and Yellen, G. (2009) A genet-
ically encoded fluorescent reporter of ATP:ADP ra-
tio, Nat. Methods, 6, 161-166, https://doi.org/10.1038/
nmeth.1288.
85. Tantama, M., Martínez-François, J. R., Mongeon, R.,
and Yellen, G. (2013) Imaging energy status in live
cells with a fluorescent biosensor of the intracellu-
lar ATP-to-ADP ratio, Nat. Commun., 4, 2550, https://
doi.org/10.1038/ncomms3550.
86. Zhang, M., Yang, B., Zhang, J., Song, Y., Wang, W.,
Li,N., Wang,Y., Li,W., and Wang,J. (2022) Monitoring
the dynamic regulation of the mitochondrial GTP-to-
GDP ratio with a genetically encoded fluorescent bi-
osensor, Angew. Chem. Int. Ed. Engl., 61, e202201266,
https://doi.org/10.1002/anie.202201266.
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.