ISSN 0006-2979, Biochemistry (Moscow), 2023, Vol. 88, No. 10, pp. 1513-1527 © Pleiades Publishing, Ltd., 2023.
Russian Text © The Author(s), 2023, published in Biokhimiya, 2023, Vol. 88, No. 10, pp. 1829-1846.
1513
REVIEW
Investigation of the Mechanism of Membrane Potential
Generation by Heme-Copper Respiratory Oxidases
in a Real Time Mode
S e r g e i A . S i l e t s k y
Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University,
119991 Moscow, Russia
e-mail: siletsky@belozersky.msu.ru
Received June 29, 2023
Revised August 15, 2023
Accepted August 15, 2023
AbstractHeme-copper respiratory oxidases are highly efficient molecular machines. These membrane enzymes catalyze
the final step of cellular respiration in eukaryotes and many prokaryotes: the transfer of electrons from cytochromes or
quinols to molecular oxygen and oxygen reduction to water. The free energy released in this redox reaction is converted
by heme-copper respiratory oxidases into the transmembrane gradient of the electrochemical potential of hydrogen ions
ΔμH
+
). Heme-copper respiratory oxidases have a unique mechanism for generating ΔμH
+
, namely, a redox-coupled proton
pump. A combination of direct electrometric method for measuring the kinetics of membrane potential generation with the
methods of prestationary kinetics and site-directed mutagenesis in the studies of heme-copper oxidases allows to obtain
a unique information on the translocation of protons inside the proteins in real time. The review summarizes the data of
studies employing time-resolved electrometry to decipher the mechanisms of functioning of these important bioenergetic
enzymes.
DOI: 10.1134/S0006297923100085
Keywords: bioenergetics, cytochrome oxidase, proteoliposomes, electrogenic proton transfer, ΔΨ generation, proton
pump, photoreduction, cytochrome aa
3
, kinetics, direct electrometric method, capacitive potentiometry, time resolution,
zinc ions
Abbreviations: ΔΨ, transmembrane electrical potential difference; ΔμH
+
,transmembrane gradient of the electrochemical poten-
tial of hydrogen ions; BNC,binuclear center; COX,cytochrome oxidase; F, O, O
H
, P, E, E
H
, and R,cytochrome c oxidase states;
P and N sides of the membrane, positively and negatively charged, respectively, aqueous phases separated by the coupling mem-
brane; PLS,proton loading site; Rubpy,tris(bipyridine)ruthenium complex.
INTRODUCTION.
HEME-COPPER RESPIRATORY OXIDASES
Terminal respiratory oxidases are key components
of respiratory chains of the mitochondria and most aer-
obic bacteria [1, 2]. These ubiquitous enzymes are at
the end of the respiratory chain and catalyze reduction
of molecular oxygen by cytochromec or ubiquinol. This
process is coupled with the formation of proton motive
force (PMF), a transmembrane gradient of the electro-
chemical potential of hydrogen ions (ΔμH
+
) [3-7] used as
an energy source for various energy-demanding cellular
processes, such as membrane transport, synthesis of bio-
molecules, movement of bacterial cells, etc.
Terminal respiratory oxidases are typically classified
into two superfamilies: heme-copper oxidases and cyto-
chromes bd [7]. A characteristic feature of heme-copper
oxidases is the presence of the oxygen-reducing catalyt-
ic site (the so-called binuclear center, BNC) formed by
closely located iron ion of the heme group and copper
ion. Mitochondrial cytochromec oxidase contains four
redox centers (Fig. 1). Its catalytic site is located inside
the enzyme hydrophobic core and consists of two redox
centers: high-spin heme a
3
iron and copper ion Cu
B
.
Four electrons supplied by cytochrome c through two
other redox centers (Cu
A
and low-spin heme a) sequen-
tially enter the BNC, which ultimately results in the re-
duction of one oxygen molecule to two water molecules.
SILETSKY1514
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 1. Scheme of cytochrome oxidase (COX) activity. Negatively and positively charged sides of the membrane face the mitochondrial matrix
and the intermembrane space, respectively. In the case of prokaryotic enzymes, the corresponding sides of the membrane face the cytoplasm and
the periplasmic space. Red and blue arrows indicate the transfer of electrons and protons, respectively.
In bacterial oxidases, the low-spin heme a can be re-
placed by heme b, and the high-spin heme a
3
can be re-
placed by hemes o
3
orb
3
. Some bacterial oxidases have
additional hemec centers.
The superfamily of heme-copper oxidases includes
three large groups: familiesA, B, and C [8]. Structur-
al modeling and comparative analysis of genomes have
identified genes of five new families of terminal oxidases
(D, E, F, G, and H) in archaebacteria [5]. FamilyA has
the largest number of representatives and contains the
most studied enzymes, including cytochrome oxidases
(COXs) from the bovine cardiac muscle mitochondria
and yeast, aa
3
-type COXs from Paracoccus denitrificans
and Rhodobacter sphaeroides, caa
3
oxidase from Thermus
thermophiles, and bo-type quinol oxidase from Esche-
richia coli [9-17]. A typical representative of the much
less studied familyB is ba
3
-type COX from T. thermo-
philus, which shows low amino acid sequence homology
with family A enzymes [18, 19]. The most evolutionary
distant family C includes cbb
3
oxidases from various or-
ganisms [20,21].
COX generates the PMF due to the vectorial (across
the membrane) delivery of electrons and protons neces-
sary for the catalytic reduction of oxygen to water in the
BNC. The so-called chemical (or substrate) protons are
transferred to the BNC from the internal aqueous phase
(negatively charged N side of the membrane), while elec-
trons are transferred from cytochromec from the posi-
tive outer side of the membrane (P side) (Fig. 1, [22]).
COX also transfers an average of 4 protons (pumped pro-
tons) from the N side to the P side per each reduced oxy-
gen molecule, i.e., functions as a proton pump [23, 24].
Therefore, the reaction catalyzed by COX is associated
with a charge separation corresponding to the directed
transmembrane transfer of 8 charges per each reduced
oxygen molecule and can be described by equation(1):
4 c
2+
+O
2
+8H
+
n
→4c
3+
+2H
2
O+4H
+
p
, (1)
where n is the mitochondrial matrix and p is the inter-
membrane space.
Generation of the membrane potential in a station-
ary mode (i.e., with multiple enzyme turnovers) by the
mitochondrial COX integrated into proteoliposomes was
demonstrated in the laboratory of Drachev in the early
experiments on the development of direct electrometric
method [25]. However, these studies failed to reveal the
electrogenic mechanism of membrane potential genera-
tion and formation of ΔμH
+
in the enzyme catalytic cycle.
Later advancements in the development of direct elec-
trometric method with the time resolution adequate for
studying the events of the catalytic cycle of photoactive
proteins [26] have been successfully used to investigate
the activity mechanisms of heme-copper oxidases. Since
these enzymes are not photoactivatable proteins, a new
methodological approach has been created to study their
electrogenic mechanisms with a high time resolution
(see below) based on a combination of direct electrom-
etry and special photoactivatable chemical reactions in-
duced by nanosecond laser pulses to trigger electron and
proton transfer in COX. This method allows to obtain
unique information on the structure and functioning
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1515
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
of the redox-dependent COX proton pump, as well as
on the coupling mechanism and stages of proton trans-
location in the protein in real time (prestationary mode)
[27-31].
THREE-DIMENSIONAL STRUCTURE OF COX
X-ray diffraction analysis has been used to estab-
lish the three-dimensional structures of several typical
heme-copper terminal oxidases of the A family, including
COX from the bovine heart mitochondria [32-35], bac-
terial COXs of the aa
3
type from P. denitrificans [36-38]
and R. sphaeroides [39, 40], and quinol oxidase bo
3
from
E. coli [41] and COX caa
3
from T. thermophiles [42].
Atomic structures have also been obtained for the mem-
bers of the B and C families, such as cytochrome ba
3
from T. thermophilus [19, 43] and COX cbb
3
from Pseudo-
monas shutzeri [21], respectively.
Heme-copper oxidases are characterized by the
largest central subunitI consisting of 12 transmembrane
α-helices. While bacterial COXs contain no more than
4subunits, COX from mammalian mitochondria is com-
posed of 13 subunits; its molecular weight (~200 kDa)
is approximately twice as large as that of bacterial en-
zymes. The three largest subunits of mitochondrial COX
are encoded by the mitochondrial genome. They form
the catalytic core of the enzyme and are homologous
to the three main subunits found in most typical pro-
karyotic family A aa
3
type COXs of the superfamily of
heme-copper oxidases.
All family A COXs contain two potential proton
pathways (from the internal aqueous phase towards the
BNC) that involve conserved proton-exchanging amino
acid residues and bound water molecules [19, 32, 36,
39, 42]. Proton pathway D (D channel) [30-32] leads
from the D132 residue through a chain of water mole-
cules and amino acid residues bound by hydrogen bonds
to the conserved E286 residue located 10-12 Å from
the BNC and 24-26 Å from D132 [24-26]. Water mole-
cules in the channel are stabilized through the hydrogen
bonds formed with highly conserved hydrophilic amino
acid residues (N139, N121, N207, S142, S200, S201, and
S197) [33]. Proton pathwayK (K channel) [44-46] is lo-
cated directly below the BNC and leads from the protein
surface on the membrane N side through T352, T359,
K362, and bound water to Y288 located near the BNC.
The entrance to the K channel is formed with the in-
volvement of the E101 residue of the second subunit I
[38, 47]. The chain of hydrogen bonds is interrupted be-
tween the K362 and T359 residues. It is believed that
changes in the K362 conformation restore the chain of
hydrogen bonds, ensuring controlled proton transfer
[38,48].
The hydrophilic domain, which is located above
the heme groups and Cu
B
, can theoretically provide the
exit of the pumped proton from the membrane outer
side to the aqueous phase [49]. This domain contains
a cluster of negatively charged amino acids (including
conserved D399 and D404 residues), several arginine
residues, propionate substituents of the heme, water
molecules, and bound redox-inactive metal atom (Mg
2+
in the mitochondrial COX that is partially replaced by
Mn
2+
in prokaryotic enzymes) [38]. The transfer of mo-
lecular oxygen in the BNC toward the active site occurs
through several hydrophobic channels in the middle part
of the membrane bilayer [32, 39, 50, 51]. The release of
water molecules formed in the BNC presumably takes
place at the membrane P side, above the heme groups,
in the area of contact between subunitsI andII, near the
Mg
2+
/Mn
2+
site [52].
COX CATALYTIC CYCLE INTERMEDIATES
The catalytic cycle of family A COXs is character-
ized by two phases, namely, oxidation and reduction
half-reactions (Fig. 2), each including two single-elec-
tron transitions [53]. During the reduction phase (O → E
and E R transitions), the first two electrons are trans-
ferred to the BNC; as a result, it acquires the ability to
bind molecular oxygen. The oxidative phase begins with
the interaction of reduced BNC (stateR; Fig. 2) with an
oxygen molecule with the formation of primary diatom-
ic oxygen adduct (state A; not shown in Fig. 2) [54].
The oxygen molecule is transferred through Cu
B
to the
central iron atom of the high-spin heme a
3
with the gen-
eration of the oxycomplex [55], which is a mixture of
the Fe
2+
–O
2
and Fe
3+
–O
2
states [56]. At the next stage,
the interatomic O–O bond is broken and the P
M
state is
formed (Fig. 2), which requires the transfer to O
2
of at
least one proton and four electrons from the active site
[57,58]. Two electrons are supplied during oxidation of
the heme a
3
Fe
2+
ion to the oxoferryl state Fe
4+
=O
2−
.
One electron comes from Cu
B
+
that is oxidized to Cu
B
2+
.
The fourth electron and the proton come from the closely
located conserved Y288 residue [59] that forms a cova-
lent bond with the histidine ligand of Cu
B
[33, 60, 61].
By donating an electron and a proton, Y288 is converted
into a radical [62].
Single-electron reduction of the P
M
intermediate
converts COX to the F state (transfer of the third electron
to Y288). Oxidation of the fourth cytochromec molecule
and transfer of the fourth electron to the BNC completes
the catalytic cycle with the formation of fully oxidized
O
H
state (Fig. 2). In the absence of electron donors, the
oxidized “unrelaxed” O
H
state spontaneously converts to
the oxidized stable state(O) within a few seconds. The O
and O
H
states differ in the electron affinity of their redox
centers, as well as in the ability for the transmembrane
proton transfer [11, 29, 63-66]. Using the treatment
with hydrogen peroxide or carbon monoxide, almost
SILETSKY1516
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 2. Catalytic cycle of familyA COX enzymes. Transitions R→O
H
and O
H
→R form, respectively, the oxidative and reduction phases of the
catalytic cycle. During the R→P
M
transition (red arrow), O
2
molecule binds to hemea
3
of the BNC in the completely reduced Fe
2+
Cu
B
1+
state,
the O–O bond breaks, and the hemea
3
oxoferryl complex Fe
4+
=O
2−
and Cu
B
1+
-O
H
is formed. Transitions P
M
→F, F→O
H
, O
H
→E
H
, and E
H
→R
(blue and blue-green arrows) involve sequential transfer of four electrons into the BNC and pumping of protons.
entire enzyme population can be transferred to the sta-
tionary states P
M
or F that correspond to the partial re-
duction of molecular oxygen in the BNC with two or
three electrons.
TIME-RESOLVED METHODS
FOR STUDYING CHARGE TRANSFER IN COX
In direct electrometry, proteoliposomes with a
protein under study are attached to an artificial mac-
roscopic membrane (collodium film impregnated with
a phospholipid solution in decane) and the transmem-
brane potential generation kinetics is recorded with a
submicrosecond time resolution using macroscopic AgCl
electrodes. The difference of the transmembrane elec-
trical potentials (ΔΨ) on the measured macroscopic
membrane increases proportionally to the generation
of electrical potential difference on the proteoliposome
membrane, which makes it possible to monitor the ki-
netics of electrogenic translocation of charges in the
studied protein with an adequate time resolution in the
prestationary mode [26, 67, 68].
The catalytic cycle of COX occurs on a millisec-
ond time scale and involves a sequence of intermediate
states of the oxygen-reducing BNC, as well as individual
electron and proton transfer events within transitions be-
tween the BNC intermediate states. All of these reactions
occur faster than the time resolution limits of fast mix-
ing methods. To study the catalytic cycle of COX with
a sufficient time resolution, oxidation and reduction re-
actions were induced with a nanosecond laser flash that
synchronizes the entire ensemble of enzyme molecules
(i.e., triggers simultaneous electron transfer by all mol-
ecules of the enzyme population). The first option (the
flow-flash method) involves studying the kinetics of ox-
idation of the completely reduced enzyme (R state) by
molecular oxygen in a single-turnover mode. To avoid
the stage of mixing with oxygen becoming the limiting
stage of the reaction, interaction with oxygen is triggered
by photolysis of the preformed COX complex with the
reduced BNC [69]. In the second case, electron is “in-
jected” using photoactivated tris(bipyridine) ruthenium
complex (Rubpy) into the enzyme that has previously
been transferred to a state characterized by a particular
extent of oxygen reduction in the BNC (O, P, or F) [70].
Although the second approach has been developed
more recently than the flow-flash method, it was the
first to be used in a combination with direct electrometry.
Itshould be noted that electron injection into COX offers
an important advantage, which is an ability to separate
the sets of elementary stages of charge transfer associated
with individual single-electron transitions in the catalyt-
ic cycle. In addition, COX oxidation by the oxygen mol-
ecule in the first approach occurs from the completely
reduced state R (in which, in addition to BNC, the input
redox centers are also in the reduced state), which is not
observed under normal conditions. This review presents
a detailed description of results obtained using electron
injection by the photoactivated Rubpy complex (results
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1517
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
of the studies using the flow-flash method can be found
in the brilliant reviews [71, 72]).
KINETICS OF MEMBRANE POTENTIAL
GENERATION BY MITOCHONDRIAL COX
For the first time, the stages of charge translocation
in the F → O transition of COX from bovine heart mito-
chondria were recorded by the joint efforts of the labo-
ratories headed by A. D. Kaulen and A. A. Konstantin-
ov [73, 74]. Electron transfer in the COX molecule was
triggered by a single electron injection from the Rubpy
complex induced with a laser flash. Later, the use of this
technique has been extended to investigation of the COX
molecules fixed in the P
M
state (the third electron in
the catalytic cycle; Fig. 2) and fully oxidized O state [63,
75, 76]. Several years after the time-resolved kinetics
of the membrane potential generation by COX had been
recorded at the Belozersky Institute of Physico-Chemi-
cal Biology, the Helsinki Bioenergetics Group developed
an approach that used a combination of time-resolved
electrometry and flow-flash method [77].
Electron injection into the F state (F O
H
tran-
sition) allowed to resolve three main components of
the membrane potential generation by family A COXs
[73,74]. The “fast” microsecond phase (~40 μs) reflect-
ed electron transfer from Cu
A
to heme a. The “medium”
(~1 ms) and “slow” (~4 ms) components corresponded
to the stages of vectorial proton transfer in the enzyme
caused by the electron transfer from heme a to the BNC.
Electron transfer between heme a and BNC normally
occurs along the membrane, i.e., is non-electrogen-
ic[22]. Electron injection from Rubpy into the P state
of mitochondrial COX (P F transition) was accompa-
nied by a set of the ΔΨ generation proton components
that had a similar amplitude, but were somewhat fast-
er [63], while photoreduction of oxidized COX (injec-
tion in the O state) was limited to the hemea reduction
by Cu
A
[63, 70, 76].
The total amplitudes of millisecond electrogenic
components in the kinetics of potential generation for
the P F and F → O stages were similar (~80%), which
led to the initial estimate of ~3 protons pumped through
the membrane. Or ~1.5 pumped protons at each of the
single-electron stages of the oxidative phase of the cata-
lytic cycle [63, 73]. The amplitudes in the presteady-state
measurements were estimated using the effect of mem-
brane potential on the steady-state redox equilibrium
between cytochrome c and heme a [78]. Based on this
effect, the transfer of an electron from Cu
A
to heme a
(20% of photoelectric response; the “fast” component) is
equivalent to the translocation of an elementary charge
by ½ the value of the membrane dielectric barrier. Hence,
the total amplitude of the millisecond proton phases
(80%) is equivalent to the transmembrane transloca-
tion of two total charges, which would correspond to the
transfer from the internal aqueous phase to the BNC of
one proton by ½ of membrane thickness for protonation
of the reduced oxygen atom (“substrate” proton) and
coupled translocation of two “pumped” protons: one
through the entire membrane and one more through ½
of its thickness [63, 73]. This assessment was consistent
with the earlier conclusions obtained in the studies of the
steady-state quasi-equilibrium reversal of the COX-cat-
alyzed reaction in the coupled mitochondria. It was as-
sumed that all four pumped protons can be transferred
through the membrane during the catalytic cycle oxida-
tive phase [79]. It was assumed that the reducing stage
of the photocycle does not possess a sufficient supply of
free energy and is not associated with the pumping of
protons, which also followed from relatively low values
of the BNC redox potentials under stationary conditions
(in the O state) [80].
KINETICS OF MEMBRANE POTENTIAL
GENERATION BY FAMILYA
BACTERIAL COX ENZYMES
The use of direct electrometric method for studying
the mechanism of COX activity has received further de-
velopment due to site-directed mutagenesis of family A
bacterial COXs, which are homologous to mitochondrial
COXs. The most studied enzymes of this group are bac-
terial COXs of the aa
3
type from R. sphaeroides [15, 28,
30, 31, 45, 81, 82] and P. denitrificans [76, 83-86].
The studies of the F O transition in the mutant
aa
3
-type COX from R. sphaeroides with the uncoupled
phenotype (N139D substitution [30, 87, 88]) have re-
sulted in the overestimation of the electrogenicity of
the single-electron transitions in the oxidative phase of
the COX catalytic cycle. The enzyme with the N139D
mutation in the D channel fully retained its oxygen re-
ductase activity in the steady-state measurements, but
lost the ability to acidify the outer medium in the prote-
oliposome suspension (i.e., to pump protons across the
membrane in the steady-state measurements). Gener-
ation of the membrane potential at the F O stage in
response to the electron injection in the mutant enzyme
reflected translocation of one elementary charge through
the membrane, namely, the transfer of an electron and
a proton from the opposite sides of the membrane to
the heme a
3
oxoferryl complex [30].
The ratio of the amplitudes of electrogenic compo-
nents of the electron transfer from the outer side of the
membrane [30, 31] and of the substrate proton trans-
fer from the internal aqueous phase to the BNC in the
N139D mutant allowed to estimate the electrogenic dis-
tance of heme a and BNC from the external aqueous
phase (~0.4 of the membrane dielectric thickness), which
was in good agreement with the structural data [32,36].
SILETSKY1518
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Accordingly, the relative amplitude of the electrogenic
phase, which reflects the transfer of the substrate proton
in the N139D mutant from the internal aqueous phase to
the BNC, was ~0.6 of the membrane dielectric thickness.
The total amplitude of the “medium” and “slow” elec-
trogenic phases in the P F and F O transitions in
the wild-type COX and mitochondrial COX is ~4 times
larger than the amplitude of the electrogenic phase of
electron transfer from the outer side of the membrane to
heme a [31, 63, 73], which is equivalent to the transfer of
~1.6 positive charge through the entire thickness of the
membrane dielectric. That is, in addition to the transfer
of one substrate proton from the internal aqueous phase
to the BNC (~0.6 of the membrane dielectric thickness),
additional electrogenic contribution to the single-elec-
tron transition in the wild-type COX, compared to the
N139D mutant, corresponds to the transfer of approx-
imately one proton through the membrane. In other
words, both single-electron transitions in the oxidative
phase of the COX catalytic cycle are associated with the
transfer of one substrate proton to the BNC and pump-
ing of no more than one proton through the membrane.
Similar estimates were obtained for the P F and F O
transitions in the reaction of fully reduced COX with ox-
ygen using the electrometric method [89] and by mea-
suring the protonation kinetics of a pH indicator [90].
Independently, Verkhovsky et al. [53] from the Hel-
sinki Bioenergetics Group demonstrated that the total
membrane potential generation in the reducing half-re-
action of the COX catalytic cycle can involve not only
the transfer of protons to the BNC, but also transmem-
brane pumping of protons. Later, the same group used a
variant of the flow-flash method to oxidize fully reduced
COX with an O
2
molecule in a single-turn mode followed
by the electron injection from Rubpy. This approach al-
lowed to study the kinetics of membrane potential gen-
eration associated with the single-electron reduction of
the metastable oxidized state O
H
of cytochrome oxidase.
It was shown with time resolution [85,86] that rapid oxi-
dation of completely reduced COX was immediately fol-
lowed by the short-lived oxidized state O
H
characterized
by a much more positive redox potential of BNC than the
O state. In contrast to the O state, electron injection into
COX in the O
H
state was accompanied by rapid electron
transfer to the BNC and associated stages of electrogen-
ic proton translocation [65, 85, 86, 91, 92]. The charac-
teristics of the electrogenic phases in the single-electron
O
H
→ E
H
transition [85, 86] generally resembled those
for the transitions in the oxidative phase and are consis-
tent with the transmembrane transfer of a single proton.
Proton pumping during the O
H
→ E
H
transition was also
confirmed in the case of natural (and not artificial, like
Rubpy) electron donor in the study of family A (subfam-
ily A2) heme-copper oxidase caa
3
from T. thermophilus,
which has an additional redox center (cytochrome c)
and, accordingly, five electrons in a completely reduced
state [11]. When the fully reduced caa
3
oxidase from
T. thermophilus was oxidized by an oxygen molecule in
the single-turnover mode, the final transition was the
O
H
→ E
H
transition [92]. The final electron acceptor in
the O
H
→ E
H
transition of this enzyme, as in the case of
the artificial electron donor Rubpy, was Cu
B
.
The E
H
R transition is presumably associated with
the pumping of one proton. However, the stage of the
second electron transfer in the COX catalytic cycle still
remains the least studied process, since it was not pos-
sible to obtain a homogeneous population of enzyme
molecules in the single-electron state (E
H
). This is pri-
marily due to the difficulty of fixing the enzyme in the
E
H
state and existence of multiple possible states of the
single-electron (E and E
H
) and oxidized (O and O
H
) en-
zyme differing in their functional properties. Treating
COX in the Fstate with carbon monoxide under steady-
state conditions converted it to the E state, which has
one electron equivalent more than the oxidized stateO.
Electron injection from Rubpy into this state led to a
photoelectric response similar in its characteristics to
the F O transition [84]. However, electron distribution
between the redox centers in the resulting E state dif-
fered significantly from the distribution observed for the
E
H
state formed by the electron injection from Rubpy
into COX in the O
H
state [65, 86, 92].
FUNCTIONAL ROLE
OF PROTON-CONDUCTING PATHWAYS
IN THE CATALYTIC CYCLE
OF FAMILYA COX ENZYMES
The studies on the effect of mutations of proton-ex-
changing groups on the functional characteristics of the
coupled proton transfer demonstrated that electrogenic
translocation of protons by familyA COX enzymes oc-
curs through the proton-conducting structures (“chan-
nels” D and K) containing critical conserved proton-ex-
change residues [30, 31, 45, 81]. However, participation
of these channels in the conduction of different protons
at the individual stages of the catalytic cycle is organized
in a non-trivial way. Based on the effect of mutations of
particular amino acid residues [93, 94] and the three-
dimensional structure of COX [32, 36], it was assumed
that these two channels are specialized for conducting
substrate protons (K channel) and protons pumped
through the membrane (D channel).
If the D and K channels were responsible for the
transfer of protons of different types, then mutations
blocking their activity should have inhibited the F O
transition. However, while amino acid substitutions in
the D channel suppressed the millisecond components
of the potential generation and, therefore, electrogenic
proton transfer at the F O stage, the blockade of the
K channel did not inhibit electrogenic proton transfer
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1519
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
at this stage [30, 31, 45, 81]. Moreover, in the presence
of excess respiration substrates, i.e., when the influx of
electrons was not the reaction rate-limiting stage, reduc-
tion of the BNC (O R transition) in the K-channel
mutants was hindered due to the blockade of proton cap-
ture from the matrix into the BNC during the reducing
phase of the catalytic cycle. It was concluded that the
D and K channels differ not in the nature of conduct-
ed protons (substrate or pumped), but in the type of the
catalytic cycle half-reaction they serve [45, 81, 85].
Later experimental results indicated that the K
channel does not appear to be involved in the transfer of
pumped protons at all, while the D channel provides the
transfer of substrate protons in the oxidative part of the
cycle and all pumped protons in the catalytic cycle (as
in oxidative phase and in the reducing phase). Indeed,
if the transfer of pumped protons occurred through the
K channel, then the N139D mutation in the D chan-
nel (which preserves the oxygen reductase function, but
completely inhibits proton pumping through the mem-
brane [30,87]) should not affect the transfer of pumped
protons in the reducing phase of the catalytic cycle.
In accordance with this, no mutations in the K channel
have been identified that specifically inhibited the pump-
ing of protons across the membrane. It is currently ac-
cepted that the K channel transfers one or both substrate
protons during the reducing part of the cycle (Fig. 2) [12,
85, 95]. It was found that the hydrogen bond between
the hydroxyl group of the heme a
3
farnesyl substituent
and redox-active tyrosine residue (Y288), which is pres-
ent in the oxidized COX, was absent in the crystals of
reduced COX form [40] and was replaced by water mole-
cules (i.e., this bond can act as a “latch” in the upper part
of the K channel).
The signal to turn off the K channel can be forma-
tion of the oxoferryl state of heme a
3
during generation
of intermediate P in the A P transition. The resulting
oxene state of the oxygen atom (as a strong axial ligand
of heme iron) can cause changes in the enzyme con-
formation, similar to the R T transition in hemoglo-
bin [45]. A slight shift of charges inside the K channel,
which is closed in its upper part, can also occur during
the oxidative half-reaction of the catalytic cycle [31, 82].
The role of such charge shift inside the channel is not
clear; for example, it may lead to a decrease in the en-
ergy barrier for the electron transfer reactions inside the
COX hydrophobic core.
Beside the D and K channels, mitochondrial COX
might have the third proton pathway, the H channel,
whose role in the conduction of protons remains a mat-
ter of debates [34, 96]. According to the authors of the
study on the X-ray structural analysis of mammalian
COXs [34, 35, 96], who identified this hypothetical pro-
ton pathway, the H channel conducts pumped protons
into COX from the N side of the membrane through
the vicinity of heme a to the D51 residue located near
the P side of the membrane. The function of this chan-
nel and regulation of proton translocation through it, in-
cluding physiological aspects of this regulation, have been
poorly studied [14]. It was suggested that the H channel
is involved in the conduction of pumped protons across
the entire membrane [97]; according to other hypoth-
esis, only the upper part of this pathway is used [98].
Finally, the third hypothesis states that the H channel is
not involved in the proton pumping, but represents the
so-called dielectric “well”, whose role may be modula-
tion of the redox potential of the nearby electron trans-
fer, hemea [99, 100].
There are indications that mutations in the H chan-
nel inhibit proton translocation by the mitochondrial
COX [34, 96]. However, targeted mutagenesis of enzymes
from higher eukaryotes is technically difficult, which
complicates interpretation of the obtained results [101].
At the same time, the exit portion of the proton H chan-
nel, whose conformational changes are the key element
of the alternative hypothetical mechanism of proton
pumping by the mitochondrial COX, is not conserved
in homologous oxidases from bacteria [102, 103]. More-
over, mutations of homologous residues in the H chan-
nels of bacterial aa
3
-type COXs do not inhibit these
enzymes [103, 104]. The studies of COX proteins from
lower eukaryotes (yeast) do not support involvement of
the H channel in proton pumping [16].
MECHANISM OF MEMBRANE
POTENTIAL GENERATION DURING
THE SINGLE-ELECTRON TRANSITION F → O
Studying the effects of isotope substitution and zinc
ions has been instrumental in the identification of elec-
trogenic processes underlying the “medium” and “slow”
phases of the membrane potential generation kinetics
during the F O transition. The effects of isotope sub-
stitution differ significantly between the electrogenic
phases, which allowed to identify analogs of these phases
in COX mutants and to correlate them with the transfer
of different types of protons [30]. The “medium” phase
was absent in the uncoupled mutant with the N139D
substitution in the D channel, which retained the oxy-
gen reductase activity but was unable to pump protons
through the membrane. This suggests the association of
the “medium” phase with the transfer of pumped pro-
ton, as well as that the pumped proton is translocated
before the substrate one. Accordingly, the “slow” phase,
which was retained in the mutant enzyme, is interpreted
as the substrate proton transfer to the BNC.
In the presence of zinc ions (proton transport in-
hibitors) added externally to proteoliposomes with the
mitochondrial COX or wild-type cytochrome aa
3
oxi-
dase from R. sphaeroides [15, 105], the “slow” electro-
genic phase decelerated. No such effect was observed
SILETSKY1520
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
for the “medium” electrogenic phase. Zinc ions also
had no impact on the protonic electrogenic phase in the
uncoupled N139D mutant. Taken together, these data
indicate that the “slow” electrogenic phase in the wild-
type enzyme includes a stage of the pumped proton re-
lease to the membrane outer side and that this process
is likely the rate-limiting reaction for the entire “slow”
electrogenic phase [15]. In contrast to the proton entry
pathways, the trajectory of proton exit to the outer mem-
brane has been poorly studied. There are indications that
the exit of water molecules is organized in a form of dis-
crete trajectory through two channel-like structures that
can also be involved in conducting the pumped proton
to the outer side of the membrane [52, 106]. Studying
the effects of different zinc concentrations on the ki-
netics of membrane potential generation in proteolipo-
somes with membrane-incorporated cytochrome aa
3
oxidase from R.sphaeroides indicated the presence of at
least two separate effective binding sites for Zn
2+
ions on
the P side of the membrane. It was also shown that the
release of the pumped proton from the proton loading
site (PLS) of COX into the external aqueous phase can
occur along several trajectories [15].
Analysis of the “medium” and “slow” phases us-
ing sequential reaction model revealed that the values of
their relative amplitudes were close [107, 108]. Thus, the
relative amplitudes of the “medium” and “slow” elec-
trogenic protonic phases for the F O transition in the
mitochondrial enzyme were ~1.9-2.1 of the value for the
“fast” electrogenic phase, i.e., which together corre-
sponded to the transfer of two protons through most of
the membrane dielectric [108]. As mentioned above, the
D channel ensures the transfer of both pumped and sub-
strate protons during transitions of the oxidative phase
of the COX catalytic cycle [45, 81]. Therefore, the “me-
dium” and “slow” electrogenic protonic phases can be
interpreted as two sequential identical processes of re-
protonation of some key residue located in the vicinity
of the BNC in the D channel, with a minor contribu-
tion from other coupled charge transitions in the protein
[107, 109, 110].
The conserved E286 residue located in the upper
part of the D channel serves as an intermediate proton
donor that ensures conduction of the pumped and sub-
strate protons (Fig.3). E286 is a bifurcation point in the
translocation of protons through the D channel that can
occur either along the substrate proton pathway to the
heme a
3
/Cu
B
center or to the outer side of the mem-
brane. Replacement of E286 with a non-protonatable
analog (E286Q) inhibited the transfer of both pumped
and substrate protons in the single-electron F → O tran-
sition [45, 81]. Experiments with the enzyme carrying
the N139L mutation, which blocks the entrance to the
D channel, allowed to identify the electrogenic stage of
proton transfer from the putative primary proton donor
(residue E286) to the BNC and to establish the electro-
genic distance between them (~0.15 of thickness of the
membrane dielectric) [31]. This is in good agreement
with the structural data, according to which the distance
from the internal aqueous phase to E286 is ~0.56 of the
geometric thickness of the membrane [107]. In the case
of proton deficiency in the D channel (deprotonated
E286), the role of the proton donor for the BNC can be
performed by the Y35 residue in the middle part of the
channel [27].
Single-electron transition F O in family A heme-
copper oxidases upon electron injection from Rubpy
begins with the reduction of Cu
A
and electrogenic trans-
fer of electron from Cu
A
to heme a (“fast” component
of the ΔΨ generation kinetics) [17] (Fig. 3). A possible
shift of the positive charge within the gated K channel
in response to the heme a reduction may facilitate this
process (step 2′ in Fig. 3) [31]. Electron transfer from
hemea to heme a
3
4+
=O
2–
occurs along the membrane
plane in almost electrically neutral manner. Proton
translocation begins with the transfer of the pumped
proton from E286 to the proton “trap” (PLS) located
upstream of the BNC. In different family A heme-cop-
per oxidases, heme a
3
propionate A and/or one of the
histidine ligands of Cu
B
can be considered as the PLS
[111, 112]. There are indications that the role of PLS can
belong to a cluster of proton-exchange groups, includ-
ing propionates A and D of heme a
3
and nearby residues
(D52 and K171 in mitochondrial COX) [113], or by the
hydrophilic domain located above hemea [98].
In the F → O transition, the substrate proton is also
transferred from E286 directly to the BNC, and E286
is reprotonated twice through the D channel during the
entire transition. The “medium” electrogenic phase in-
cludes electrogenic proton transfer from E286 to the
primary proton acceptor (PLS), as well as one of the
successive electrogenic reprotonations of E286 from
the internal aqueous N phase through the D channel
(Fig. 3). The “slow” electrogenic phase involves the sec-
ond reprotonation of E286 following the transfer from
E286–COOH to the BNC of the proton involved in the
oxygen chemical conversion.
There are indications that the proton transfer
from E286 to the BNC (stage 5) is a part of the “slow”
(in R. sphaeroides COX) or “medium” (in mitochon-
drial COX) electrogenic phases [28]. The transfer of the
substrate proton to the BNC leads to the neutralization
of transferred electron, the negative charge of which
could stabilize the pumped proton located in the PLS.
The transfer of a new proton deep into the membrane
dielectric upon E286 reprotonation should lead to the
electrostatic expulsion of the pumped proton from
thePLS [112, 114, 115]. Accordingly, the “slow” electro-
genic phase also involves proton release from the PLS
to the outer side of the membrane, which can be slowed
down by the addition of zinc ions from the P side of the
membrane (Fig.3).
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1521
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 3. The figure shows the electrogenic stages resolved by single-electron photoinjection using Rubpy in the F O transition (transfer of the
fourth electron in the COX catalytic cycle). Before the reaction begins, hemea
3
is in the oxoferryl state (not shown). Red arrows indicate direc-
tions of electron transfer; blue arrows indicate the stages of proton translocation. Two proton pathways (D and K channels) and their conserved
amino acid residues are shown. Steps1 and3 reflect electron transfer from Cu
A
to hemea and from hemea to the BNC, respectively. During the
F → O transition, both protons (substrate and pumped) are transferred through the D channel. It is assumed that isomerization of E286 (transition
between the E286
1
and E286
2
conformations) precedes proton transfer from E286 to the PLS. Individual stages of proton transfer are components
of the “medium” (2 and 4) or the “slow” (6 and 7) electrogenic phases. An additional pathway for the pumped proton release from the PLS
(step8) was revealed in the presence of zinc ions [15]. Proton transfer from E286 to the BNC (step5) is presumably a component of the “medium”
(in mitochondrial COX) or “slow” (in R.sphaeroides COX) electrogenic phases. Stage 2′ is a possible charge shift in the K channel in response to
the hemea reduction [31].
THE USE OF DIRECT ELECTROMETRIC
METHOD IN THE STUDIES OF FAMILYB
HEME-COPPER OXIDASES
To understand the mechanism of coupled proton
pumping by COX, it is important to study the diversity of
heme-copper terminal oxidases [8], including enzymes
of much less investigated and evolutionarily distant
families B and C [19, 116, 117]. Although the studies of
family B and C oxidases have started relatively recently,
they have attracted a lot interest [29, 95, 118, 119]. Toa
great extent this is due to the common occurrence of
these enzymes (along with copper-free bd-type oxidases)
in pathogenic microorganisms and, therefore, their bio-
medical significance [120].
The use of direct electrometric method allowed
to discover an interesting feature of the BNC from the
ba
3
-type COX from T. thermophilus, a typical repre-
sentative of family B [121]. In the initial oxidized state,
this enzyme does not react with exogenous ligands and,
therefore, cannot be converted and stabilized in the P or F
states by the treatment with hydrogen peroxide or carbon
monoxide [121], as it was done for typical aa
3
-type oxi-
dases [63]. However, when an electron from Rubpy was
injected into the fully oxidized (state O) COX ba
3
from
T. thermophilus in the presence of hydrogen peroxide, an
additional electrogenic reaction was observed, the rate
of which was directly proportional to the concentration
of the added peroxide. In other words, the BNC of this
enzyme, which was inactive in the oxidized state toward
external ligands, acquired the ability to bind ligands (and
transit to the active/open state) in response to the sin-
gle-electron injection and reduction of hemeb [121].
An important aspect in the context of uniqueness
and usefulness of direct electrometric method in the
studies of the ba
3
-type COX from T. thermophilus is that
the average parameters of the catalytic cycle of this en-
zyme assessed by the degree of acidification of the outer
compartment of proteoliposomes in the stationary oxy-
gen reductase reaction, indicated the variability in the
stoichiometry of proton pumping. Unlike family A oxi-
dases, the stoichiometry of proton pumping by family B
and C oxidases in the stationary measurements can vary
within a significant range: from 0.5 to ~0.85 protons per
electron (H
+
/e
) entering the BNC [66, 118, 121-124].
Application of direct electrometric method makes it
possible to obtain and compare electrogenic parameters
of individual single-electron transitions and to obtain
SILETSKY1522
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
unique information about the features of the coupling
mechanism.
In the ba
3
-type COX from T. thermophilus (a typical
representative of family B proteins), the stages of elec-
trogenic proton transfer are resolved during the oxida-
tive phase [122] and the first transition of the reducing
phase of the catalytic cycle [66]. The stages during which
electron transfer is not associated with the proton pump-
ing across the membrane have been identified, which
explained the decrease in the effective stoichiometry of
proton pumping and its significant variability in family B
oxidases [66,122]. In particular, it was shown that in con-
trast to family A enzymes, one proton is pumped through
the membrane instead of two in the oxidative phase of
the catalytic cycle of ba
3
-type COX from T. thermophilus
[122]. At the same time, the transfer of the 1st electron in
the reducing phase (transition O
H
→ E
H
) is also not asso-
ciated with the proton pumping through the membrane,
which can be caused by the membrane potential forma-
tion during the oxidative phase [66].
Apparently, family B enzymes have preserved only
one functional proton entry channel, which is homolo-
gous to the K channel of familyA oxidases [5, 29]. Using
direct electrometric method, it was shown that in con-
trast to familyA enzymes, familyB oxidases use the K
channel for the transfer of both substrate and pumped
protons in the oxidative phase. The T315V mutation in
the K-channel of ba
3
-type COX from T. thermophilus
causes a slowdown of the F O
H
transition due to the
deceleration of the substrate proton transfer to the BNC,
as well as uncoupling, i.e., complete absence of proton
pumping in the oxidative phase of the catalytic cycle [29].
Funding. The study was supported by the Russian
Science Foundation (project no.23-24-00143).
Acknowledgments. The author would like to ex-
press his deepest gratitude to the untimely deceased
A. A. Konstantinov, A. D. Kaulen, M. I. Verkhovsky,
L. A. Drachev, and V. P. Skulachev. Collaboration with
these outstanding scientists at different times has made
an indelible impression on the author. The author is
grateful to D. L. Zaslavsky, I. A. Smirnova, I. N. Be-
levich, Prof. M. Wikström and Prof. R. Gennis for col-
laboration in the studies of heme-copper oxidases by
direct electrometric method.
Ethics declarations. The author declares no conflict
of interest in financial or any other area. This article
does not contain description of studies with human par-
ticipants or animals performed by the author.
REFERENCES
1. Anraku, Y. (1988) Bacterial electron transport chains,
Ann. Rev. Biochem., 57, 101-132, doi: 10.1146/annurev.
bi.57.070188.000533.
2. Garcia-Horsman, J. A., Barquera, B., Rumbley, J., Ma, J.,
and Gennis, R.B. (1994) The superfamily of heme-cop-
per respiratory oxidases, J. Bacteriol., 176, 5587-5600,
doi:10.1128/jb.176.18.5587-5600.1994.
3. Babcock, G. T., and Wikström, M. (1992) Oxygen Acti-
vation and the conservation of energy in cell respiration,
Nature, 356, 301-309, doi:10.1038/356301a0.
4. Ferguson-Miller, S., and Babcock, G. T. (1996) Heme/
copper terminal oxidases, Chem. Rev., 7, 2889-2907,
doi:10.1021/cr950051s.
5. Hemp, J., and Gennis, R. B. (2008) Diversity of the
heme-copper superfamily in archaea: insights from ge-
nomics and structural modeling, Results Probl. Cell Differ.,
45, 1-31, doi:10.1007/400_2007_046.
6. Rich, P. R. (2017) Mitochondrial cytochrome c oxi-
dase: catalysis, coupling and controversies, Biochem. Soc.
Trans., 45, 813-829, doi: 10.1042/BST20160139.
7. Siletsky, S. A., and Borisov, V. B. (2021) Proton pump-
ing and non-pumping terminal respiratory oxidases: ac-
tive sites intermediates of these molecular machines and
their derivatives, Int.J. Mol. Sci., 22, 10852, doi:10.3390/
ijms221910852.
8. Pereira, M. M., Santana, M., and Teixeira, M. (2001)
A novel scenario for the evolution of haem-copper oxy-
gen reductases, Biochim. Biophys. Acta, 1505, 185-208,
doi:10.1016/S0005-2728(01)00169-4.
9. Yoshikawa, S., and Shimada, A. (2015) Reaction mech-
anism of cytochrome c oxidase, Chem. Rev., 115, 1936-
1989, doi:10.1021/cr500266a.
10. Fee, J. A., Yoshida, T., Surerus, K. K., and Mather, M.W.
(1993) Cytochrome caa
3
from the thermophilic bacterium
Thermus thermophilus: a member of the heme-copper
oxidase superfamily, J. Bioenerg. Biomembr., 25, 103-114,
doi: 10.1007/BF00762852.
11. Siletsky, S. A., Belevich, I., Soulimane, T., Verkhovsky,
M.I., and Wikström, M. (2013) Thefifth electron in the
fully reduced caa
3
from Thermus thermophilus is compe-
tent in proton pumping, Biochim. Biophys. Acta, 1827,
1-9, doi:10.1016/j.bbabio.2012.09.013.
12. Wikström, M., Krab, K., and Sharma, V. (2018) Oxy-
gen activation and energy conservation by cytochrome c
oxidase, Chem. Rev., 118, 2469-2490, doi: 10.1021/acs.
chemrev.7b00664.
13. Forte, E., Borisov, V. B., Siletsky, S. A., Petrosino, M.,
and Giuffre, A. (2019) In the respiratory chain of Esch-
erichia coli cytochromes bd-I and bd-II are more sen-
sitive to carbon monoxide inhibition than cytochrome
bo
3
, Biochim. Biophys. Acta Bioenergetics, 1860, 148088,
doi:10.1016/j.bbabio.2019.148088.
14. Bjorck, M. L., Vilhjalmsdottir, J., Hartley, A. M., Meu-
nier, B., Nasvik Ojemyr, L., Marechal, A., and Brzezin-
ski,P. (2019) Proton-transfer pathways in the mitochon-
drial S. cerevisiae cytochrome c oxidase, Sci. Rep., 9,
20207, doi:10.1038/s41598-019-56648-9.
15. Siletsky, S. A., and Gennis, R. B. (2021) Time-resolved
electrometric study of the F O transition in cytochrome c
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1523
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
oxidase. The effect of Zn
2+
ions on the positive side of
the membrane, Biochemistry (Moscow), 86, 105-122,
doi:10.1134/S0006297921010107.
16. Marechal, A., Xu, J. Y., Genko, N., Hartley, A. M.,
Haraux, F., Meunier, B., and Rich, P.R. (2020) Acom-
mon coupling mechanism for A-type heme-copper oxi-
dases from bacteria to mitochondria, Proc. Natl. Acad. Sci.
USA, 117, 9349-9355, doi:10.1073/pnas.2001572117.
17. Siletsky, S. A. (2013) Steps of the coupled charge translo-
cation in the catalytic cycle of cytochromec oxidase, Front.
Biosci. (Landmark Ed), 18, 36-57, doi:10.2741/4086.
18. Fee, J. A., Sanders, D., Slutter, C. E., Doan, P. E.,
Aasa, R., Karpefors, M., and Vänngård, T. (1995)
Multi-frequency epr evidence for a binuclear Cu
A
center
in cytochrome c oxidase: studies with a
63
Cu- and
65
Cu-
enriched, soluble domain of the cytochrome ba
3
, sub-
unitII from Thermus Thermophilus, Biochem. Biophys. Res.
Commun., 212, 77-83, doi:10.1006/bbrc.1995.1938.
19. Soulimane, T., Buse, G., Bourenkov, G. B., Bartunik,
H. D., Huber, R., and Than, M.E. (2000) Structure and
mechanism of the aberrant ba
3
-cytochrome c oxidase
from Thermus thermophilus, EMBO J., 19, 1766-1776,
doi:10.1093/emboj/19.8.1766.
20. Pitcher, R. S., and Watmough, N. J. (2004) Thebacterial
cytochrome cbb
3
oxidases, Biochim. Biophys. Acta, 1655,
388-399, doi:10.1016/j.bbabio.2003.09.017.
21. Buschmann, S., Warkentin, E., Xie, H., Langer, J. D.,
Ermler, U., and Michel, H. (2010) Thestructure of cbb
3
cytochrome oxidase provides insights into proton pump-
ing, Science, 329, 327-330, doi:10.1126/science.1187303.
22. Mitchell, P. (1968) Chemiosmotic coupling and energy
transduction, Glynn Research Ltd., Bodmin.
23. Wikström, M. (1977) Proton pump coupled to cyto-
chromec oxidase in mitochondria, Nature, 266, 271-273,
doi:10.1038/266271a0.
24. Wikström, M. (2004) Cytochrome c oxidase: 25years of
the elusive proton pump, Biochim. Biophys. Acta, 1655,
241-247, doi:10.1016/j.bbabio.2003.07.013.
25. Drachev, L. A., Jasaitis, A. A., Kaulen, A. D., Kondrashin,
A. A., Liberman, E. A., Nemecek, I. B., Ostroumov, S. A.,
Semenov, A., and Skulachev, V. P. (1974) Direct measure-
ment of electric current generation by cytochrome oxidase,
H
+
-ATPase and bacteriorhodopsin, Nature, 249, 321-324,
doi:10.1038/249321a0.
26. Drachev, L. A., Kaulen, A. D., Khitrina, L. V., and
Skulachev, V. P. (1981) Fast stages of photoelectric pro-
cesses in biological membranes. I. Bacteriorhodopsin,
Eur. J. Biochem., 117, 461-470, doi:10.1111/j.1432-1033.
1981.tb06361.x.
27. Belevich, I., Gorbikova, E., Belevich, N. P., Rauhama-
ki, V., Wikström, M., and Verkhovsky, M.I. (2010) Ini-
tiation of the proton pump of cytochromec oxidase, Proc.
Natl. Acad. Sci. USA, 107, 18469-18474, doi: 10.1073/
pnas.1010974107.
28. Siletsky, S. A., and Konstantinov, A. A. (2012) Cytochrome
c oxidase: charge translocation coupled to single-electron
partial steps of the catalytic cycle, Biochim. Biophys. Acta,
1817, 476-488, doi:10.1016/j.bbabio.2011.08.003.
29. Siletsky, S. A., Soulimane, T., Belevich, I., Gennis, R.B.,
and Wikström, M. (2021) Specific inhibition of proton
pumping by the T315V mutation in the K channel of cy-
tochrome ba
3
from Thermus thermophilus, Biochim. Bio-
phys. Acta Bioenergetics, 1862, 148450, doi: 10.1016/
j.bbabio.2021.148450.
30. Siletsky, S. A., Pawate, A. S., Weiss, K., Gennis, R. B.,
and Konstantinov, A. A. (2004) Transmembrane charge
separation during the ferryl-oxo oxidized transition in
a non-pumping mutant of cytochromec oxidase, J.Biol.
Chem., 279, 52558-52565, doi:10.1074/jbc.M407549200.
31. Siletsky, S. A., Zhu, J., Gennis, R. B., and Konstan-
tinov, A. A. (2010) Partial steps of charge transloca-
tion in the nonpumping N139L mutant of Rhodo-
bacter sphaeroides cytochrome c oxidase with a blocked
D-channel, Biochemistry, 49, 3060-3073, doi: 10.1021/
bi901719e.
32. Tsukihara, T., Aoyama, H., Yamashita, E., Takashi, T.,
Yamaguichi, H., Shinzawa-Itoh, K., Nakashima, R.,
Yaono, R., and Yoshikawa, S. (1996) The whole struc-
ture of the 13-subunit oxidized cytochrome c oxidase
at 2.8 Å, Science, 272, 1136-1144, doi: 10.1126/science.
272.5265.1136.
33. Yoshikawa, S., Shinzawa-Itoh, K., Nakashima, R., Ya-
ono, R., Inoue, N., Yao, M., Fei, M. J., Libeu, C. P.,
Mizushima, T., Yamaguchi, H., Tomizaki, T., and Tsuki-
hara, T. (1998) Redox-coupled crystal structural chang-
es in bovine heart cytochrome c oxidase, Science, 280,
1723-1729, doi:10.1126/science.280.5370.1723.
34. Tsukihara, T., Shimokata, K., Katayama, Y., Shimada,H.,
Muramoto, K., Aoyama, H., Mochizuki, M., Shinzawa-
Itoh, K., Yamashita, E., Yao, M., Ishimura, Y., and
Yoshikawa, S. (2003) The low-spin heme of cytochromec
oxidase as the driving element of the proton-pumping
process, Proc. Natl. Acad. Sci. USA, 100, 15304-15309,
doi:10.1073/pnas.2635097100.
35. Muramoto, K., Hirata, K., Shinzawa-Itoh, K., Yoko-o,S.,
Yamashita, E., Aoyama, H., Tsukihara, T., and Yoshika-
wa, S. (2007) Ahistidine residue acting as a controlling
site for dioxygen reduction and proton pumping by cyto-
chromec oxidase, Proc. Natl. Acad. Sci. USA, 104, 7881-
7886, doi:10.1073/pnas.0610031104.
36. Iwata, S., Ostermeier, C., Ludwig, B., and Michel, H.
(1995) Structure at 2.8Å resolution of cytochromec oxi-
dase from Paracoccus denitrificans, Nature, 376, 660-669,
doi:10.1038/376660a0.
37. Ostermeier, C., Iwata, S., Ludwig, B., and Michel, H.
(1995) F
V
fragment-mediated crystallization of the mem-
brane protein bacterial cytochromec oxidase, Nat. Struct.
Biol., 2, 842, doi:10.1038/nsb1095-842.
38. Koepke, J., Olkhova, E., Angerer, H., Muller, H.,
Peng, G., and Michel, H. (2009) High resolution crys-
tal structure of Paracoccus denitrificans cytochrome c
oxidase: New insights into the active site and the proton
SILETSKY1524
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
transfer pathways, Biochim. Biophys. Acta, 1787, 635-645,
doi:10.1016/j.bbabio.2009.04.003.
39. Svensson-Ek, M., Abramson, J., Larsson, G., Torn-
roth,S., Brzezinski, P., and Iwata, S. (2002) The X-ray
crystal structures of wild-type and EQ(I-286) mutant
cytochrome c oxidases from Rhodobacter sphaeroides,
J. Mol. Biol., 321, 329-339, doi: 10.1016/S0022-2836
(02)00619-8.
40. Qin, L., Liu, J., Mills, D. A., Proshlyakov, D. A.,
Hiser,C., and Ferguson-Miller, S. (2009) Redox depen-
dent conformational changes in cytochromec oxidase sug-
gest a gating mechanism for proton uptake, Biochemistry,
48, 5121-5130, doi:10.1021/bi9001387.
41. Abramson, J., Riistama, S., Larsson, G., Jasaitis, A., Svens-
son-Ek, M., Laakkonen, L., Puuustinen, A., Iwata, S.,
and Wikström, M. (2000) The structure of the ubiquinol
oxidase from Escherichia coli and its ubiquinone binding
site, Nat. Struct. Biol., 7, 910-917, doi:10.1038/82824.
42. Noor, M. R., and Soulimane, T. (2013) Structure of caa
(3)
cytochrome c oxidase – a nature-made enzyme-sub-
strate complex, Biol. Chem., 394, 579-591, doi: 10.1515/
hsz-2012-0343.
43. Luna, V. M., Chen, Y., Fee, J. A., and Stout, C. D. (2008)
Crystallographic studies of Xe and Kr binding within the
large internal cavity of cytochrome ba
3
from Thermus
thermophilus: structural analysis and role of oxygen trans-
port channels in the heme-Cu oxidases, Biochemistry,
47, 4657-4665, doi:10.1021/bi800045y.
44. Fetter, J. R., Qian, J., Shapleigh, J., Thomas, J. W.,
Garcia- Horsman, A., Schmidt, E., Hosler, J., Babcock,
G. T., Gennis, R. B., and Ferguson-Miller, S. (1995)
Possible proton relay pathways in cytochrome c oxidase,
Proc. Natl. Acad. Sci. USA, 92, 1604-1608, doi:10.1073/
pnas.92.5.1604.
45. Konstantinov, A. A., Siletsky, S., Mitchell, D., Kaulen,A.,
and Gennis, R. B. (1997) Theroles of the two proton in-
put channels in cytochrome c oxidase from Rhodobacter
sphaeroides probed by the effects of site-directed mutations
on time resolved electrogenic intraprotein proton transfer,
Proc. Natl. Acad. Sci. USA, 94, 9085-9090, doi:10.1073/
pnas.94.17.9085.
46. Gennis, R. B. (1998) Multiple proton-conducting path-
ways in cytochrome oxidase and a proposed role for
the active-site tyrosine, Biochim. Biophys. Acta, 1365,
241-248, doi:10.1016/S0005-2728(98)00075-9.
47. Branden, M., Tomson, F., Gennis, R. B., and Brzezin-
ski, P. (2002) The entry point of the K-proton-trans-
fer pathway in cytochrome c oxidase, Biochemistry, 41,
10794-10798, doi:10.1021/bi026093+.
48. Hofacker, I., and Schulten, K. (1998) Oxygen and pro-
ton pathways in cytochrome c oxidase, Proteins: Struct.
Funct., 30, 100-107, doi: 10.1002/(SICI)1097-0134
(199801)30:1<100::AID-PROT9>3.0.CO;2-S.
49. Popovic, D. M., and Stuchebrukhov, A. A. (2005) Pro-
ton exit channels in bovine cytochromec oxidase, J.Phys.
Chem.B, 109, 1999-2006, doi:10.1021/jp0464371.
50. Riistama, S., Puustinen, A., Verkhovsky, M. I., Morgan,
J. E., and Wikström, M. (2000) Binding of O
2
and its re-
duction are both retarded by replacement of valine 279
by isoleucine in cytochrome c oxidase from Paracoccus
denitrificans, Biochemistry, 39, 6365-6372, doi: 10.1021/
bi000123w.
51. Salomonsson, L., Lee, A., Gennis, R. B., and Brzezin-
ski, P. (2004) Asingle-amino-acid lid renders a gas-tight
compartment within a membrane-bound transporter,
Proc. Natl. Acad. Sci. USA, 101, 11617-11621, doi:10.1073/
pnas.0402242101.
52. Schmidt, B., McCracken, J., and Ferguson-Miller, S.
(2003) A discrete water exit pathway in the membrane
protein cytochromec oxidase, Proc. Natl. Acad. Sci. USA,
100, 15539-15542, doi:10.1073/pnas.2633243100.
53. Verkhovsky, M. I., Jasaitis, A., Verkhovskaya, M. L.,
Morgan, L., and Wikström, M. (1999) Proton translo-
cation by cytochrome c oxidase, Nature, 400, 480-483,
doi:10.1038/22813.
54. Chance, B., Saronio, C., and Leigh, J. S., Jr. (1975) Func-
tional intermediates in the reaction of membrane-bound
cytochrome oxidase with oxygen, J. Biol. Chem., 250,
9226-9237, doi:10.1016/S0021-9258(19)40634-0.
55. Hill, B. C., and Greenwood, C. (1983) Spectroscopic ev-
idence for the participation of compound A (Fea
3
2+
-O
2
)
in the reaction of mixed-valence cytochrome c oxidase
with oxygen at room temperature, Biochem.J., 215, 659-
667, doi:10.1042/bj2150659.
56. Muramoto, K., Ohta, K., Shinzawa-Itoh, K., Kanda,K.,
Taniguchi, M., Nabekura, H., Yamashita, E., Tsukiha-
ra, T., and Yoshikawa, S. (2010) Bovine cytochrome c
oxidase structures enable O
2
reduction with minimization
of reactive oxygens and provide a proton-pumping gate,
Proc. Natl. Acad. Sci. USA, 107, 7740-7745, doi:10.1073/
pnas.0910410107.
57. Kitagawa, T., and Ogura, T. (1997) Oxygen activation
mechanism at the binuclear site of heme-copper oxi-
dase superfamily as revealed by time-resolved resonance
Raman spectroscopy, in Progress in Inorganic Chem-
istry (Karlin, K. D., ed) Wiley & Sons, pp. 431-479,
doi:10.1002/9780470166468.ch6.
58. Proshlyakov, D. A., Pressler, M. A., and Babcock,
G. T. (1998) Dioxygen activation and bond cleavage by
mixed-valence cytochromec oxidase, Proc. Natl. Acad.Sci.
USA, 95, 8020-8025, doi:10.1073/pnas.95.14.8020.
59. Proshlyakov, D. A., Pressler, M. A., DeMaso, C., Leykam,
J. F., DeWitt, D. L., and Babcock, G. T. (2000) Oxy-
gen activation and reduction in respiration: involvement
of redox-active tyrosine 244, Science, 290, 1588-1591,
doi:10.1126/science.290.5496.1588.
60. Buse, G., Soulimane, T., Dewor, M., Meyer, H. E., and
Bloggel, M. (1999) Evidence for a copper coordinated his-
tidine-tyrosine crosslink in the active site of cytochrome
oxidase, Protein Sci., 8, 985-990, doi:10.1110/ps.8.5.985.
61. Rauhamaki, V., Baumann, M., Soliymani, R., Puusti-
nen, A., and Wikström, M. (2006) Identification
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1525
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
of histidin-tyrosin cross-link in the active site of the
cbb
3
-type cytochrome c oxidase from Rhodobacter
sphaeroides, Proc. Natl. Acad. Sci. USA, 103, 16135-16140,
doi:10.1073/pnas.0606254103.
62. Babcock, G. T. (1999) How oxygen is activated and re-
duced in respiration, Proc. Natl. Acad. Sci. USA, 96,
12971-12973, doi:10.1073/pnas.96.23.12971.
63. Siletsky, S., Kaulen, A. D., and Konstantinov, A. A.
(1999) Resolution of electrogenic steps coupled to conver-
sion of cytochromec oxidase from the peroxy to the fer-
ryl-oxo state, Biochemistry, 38, 4853-4861, doi: 10.1021/
bi982614a.
64. Sharma, V., Karlin, K. D., and Wikstrom, M. (2013) Com-
putational study of the activated O
H
state in the catalytic
mechanism of cytochrome c oxidase, Proc. Natl. Acad. Sci.
USA, 110, 16844-16849, doi: 10.1073/pnas.1220379110.
65. Siletsky, S. A., Belevich, I., Wikström, M., Soulimane, T.,
and Verkhovsky, M. I. (2009) Time-resolved O
H
→ E
H
tran-
sition of the aberrant ba
3
oxidase from Thermus thermoph-
ilus, Biochim. Biophys. Acta, 1787, 201-205, doi:10.1016/
j.bbabio.2008.12.020.
66. Siletsky, S. A., Belevich, I., Belevich, N. P., Souli-
mane,T., and Wikström, M. (2017) Time-resolved gener-
ation of membrane potential by ba
3
cytochromec oxidase
from Thermus thermophilus coupled to single electron in-
jection into the O and OH states, Biochim. Biophys. Acta,
1858, 915-926, doi:10.1016/j.bbabio.2017.08.007.
67. Siletsky, S. A., Mamedov, M. D., Lukashev, E. P., Bal-
ashov, S. P., Dolgikh, D.A., Rubin, A.B., Kirpichnikov,
M. P., and Petrovskaya, L. E. (2016) Electrogenic steps
of light-driven proton transport in ESR, a retinal protein
from Exiguobacterium sibiricum, Biochim. Biophys. Acta,
1857, 1741-1750, doi:10.1016/j.bbabio.2016.08.004.
68. Kaulen, A. D. (2000) Electrogenic processes and protein
conformational changes accompanying the bacteriorho-
dopsin photocycle, Biochim. Biophys. Acta, 1460, 204-219,
doi:10.1016/S0005-2728(00)00140-7.
69. Gibson, Q. H., and Greenwood, C. (1963) Reactions of
cytochrome oxidase with oxygen and carbon monoxide,
Biochem.J., 86, 541-554, doi:10.1042/bj0860541.
70. Nilsson, T. (1992) Photoinduced electron transfer from
tris(2,2′-bipyridyl)ruthenium to cytochrome c oxidase,
Proc. Natl. Acad. Sci. USA, 89, 6497-6501, doi:10.1073/
pnas.89.14.6497.
71. Belevich, I., and Verkhovsky, M. I. (2008) Molecular
mechanism of proton translocation by cytochromec ox-
idase, Antioxid. Redox Signal., 10, 1-29, doi: 10.1089/
ars.2007.1705.
72. Kaila, V. R., Verkhovsky, M. I., and Wikström, M. (2010)
Proton-coupled electron transfer in cytochrome oxidase,
Chem. Rev., 110, 7062-7081, doi:10.1021/cr1002003.
73. Zaslavsky, D., Kaulen, A., Smirnova, I. A., Vygodina,
T. V., and Konstantinov, A. A. (1993) Flash-induced
membrane potential generation by cytochrome c oxi-
dase, FEBS Lett., 336, 389-393, doi: 10.1016/0014-5793
(93)80843-J.
74. Zaslavsky, D. L., Smirnova, I. A., Siletsky, S. A., Kaulen,
A.D., Millett, F., and Konstantinov, A. A. (1995) Rap-
id kinetics of membrane potential generation by cyto-
chromec oxidase with the photoactive Ru(II)-tris-bipyr-
idyl derivative of cytochrome c as electron donor, FEBS
Lett., 359, 27-30, doi:10.1016/0014-5793(94)01443-5.
75. Siletsky, S. A., Kaulen, A. D., and Konstantinov, A. A.
(1997) Electrogenic events associated with peroxy- to fer-
ryl-oxo state transition in cytochrome c oxidase, Eur. J.
Biophys., 26, 98.
76. Verkhovsky, M. I., Tuukkanen, A., Backgren, C., Puusti-
nen, A., and Wikström, M. (2001) Charge translocation
coupled to electron injection into oxidized cytochromec
oxidase from Paracoccus denitrificans
, Biochemistry, 40,
7077-7083, doi:10.1021/bi010030u.
77. Verkhovsky, M. I., Morgan, J. E., Verkhovskaya, M.,
and Wikström, M. (1997) Translocation of electrical
charge during a single turnover of cytochrome c oxi-
dase, Biochim. Biophys. Acta, 1318, 6-10, doi: 10.1016/
S0005-2728(96)00147-8.
78. Hinkle, P., and Mitchell, P. (1970) Effect of membrane
potential on the redox poise between cytochrome a and
cytochrome c in rat liver mitochondria, J. Bioenerg.,
1, 45-60, doi:10.1007/BF01516088.
79. Wikström, M. (1989) Identification of the electron transfers
in cytochrome oxidase that are coupled to proton-pump-
ing, Nature, 338, 776-778, doi:10.1038/338776a0.
80. Wikström, M., Krab, K., and Saraste, M. (1981) Cyto-
chrome Oxidase– ASynthesis, Academic Press, New York.
81. Siletsky, S. A., Kaulen, A. D., Mitchell, D., Gennis, R.B.,
and Konstantinov, A.A. (1996) Resolution of two proton
conduction pathways in cytochrome c oxidase, EBEC
Short Rep., 9, 90.
82. Lepp, H., Svahn, E., Faxen, K., and Brzezinski, P.
(2008) Charge transfer in the K proton pathway linked to
electron transfer to the catalytic site in cytochromec oxi-
dase, Biochemistry, 47, 4929-4935, doi:10.1021/bi7024707.
83. Ruitenberg, M., Kannt, A., Bamberg, E., Ludwig, B., Mi-
chel, H., and Fendler, K. (2000) Single-electron reduc-
tion of the oxidized state is coupled to proton uptake via
the K pathway in Paracoccus denitrificans cytochrome c
oxidase, Proc. Natl. Acad. Sci. USA, 97, 4632-4636,
doi:10.1073/pnas.080079097.
84. Ruitenberg, M., Kannt, A., Bamberg, E., Fendler, K., and
Michel, H. (2002) Reduction of cytochromec oxidase by
a second electron leads to proton translocation, Nature,
417, 99-102, doi:10.1038/417099a.
85. Bloch, D., Belevich, I., Jasaitis, A., Ribacka, C., Puusti-
nen, A., Verkhovsky, M. I., and Wikström, M. (2004)
The catalytic cycle of cytochromec oxidase is not the sum
of its two halves, Proc. Natl. Acad. Sci. USA, 101, 529-533,
doi:10.1073/pnas.0306036101.
86. Belevich, I., Bloch, D. A., Wikström, M., and Verkhovsky,
M. I. (2007) Exploring the proton pump mechanism of
cytochromec oxidase in real time, Proc. Natl. Acad. Sci.
USA, 104, 2685-2690, doi:10.1073/pnas.0608794104.
SILETSKY1526
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
87. Pawate, A. S., Morgan, J., Namslauer, A., Mills, D.,
Brzezinski, P., Ferguson-Miller, S., and Gennis, R. B.
(2002) Amutation in subunitI of cytochrome oxidase from
Rhodobacter sphaeroides results in an increase in steady-
state activity but completely eliminates proton pumping,
Biochemistry, 41, 13417-13423, doi:10.1021/bi026582+.
88. Pfitzner, U., Hoffmeier, K., Harrenga, A., Kannt, A.,
Michel, H., Bamberg, E., Richter, O.-M. H., and Lud-
wig, B. (2000) Tracing the D-pathway in reconstitut-
ed site-directed mutants of cytochrome c oxidase from
Paracoccus denitrificans, Biochemistry, 39, 6756-6762,
doi:10.1021/bi992235x.
89. Ribacka, C., Verkhovsky, M. I., Belevich, I., Bloch, D. A.,
Puustinen, A., and Wikström, M. (2005) An elementary
reaction step of the proton pump is revealed by mutation of
tryptophan-164 to phenylalanine in cytochrome c oxidase
from Paracoccus denitrificans, Biochemistry, 44, 16502-
16512, doi:10.1021/bi0511336.
90. Faxen, K., Gilderson, G., Adelroth, P., and Brzezinski,P.
(2005) Amechanistic principle for proton pumping by cy-
tochrome c oxidase, Nature, 437, 286-289, doi:10.1038/
nature03921.
91. Brand, S. E., Rajagukguk, S., Ganesan, K., Geren, L.,
Fabian, M., Han, D., Gennis, R. B., Durham, B., and
Millett, F. (2007) A new ruthenium complex to study
single-electron reduction of the pulsed O
H
state of deter-
gent-solubilized cytochrome oxidase, Biochemistry, 46,
14610-14618, doi:10.1021/bi701424d.
92. Siletsky, S. A., Belevich, I., Belevich, N. P., Souli-
mane, T., and Verkhovsky, M. I. (2011) Time-resolved
single-turnover of caa
3
oxidase from Thermus thermophi-
lus. Fifth electron of the fully reduced enzyme converts
O
H
into E
H
state, Biochim. Biophys. Acta, 1807, 1162-1169,
doi:10.1016/j.bbabio.2011.05.006.
93. Thomas, J. W., Puustinen, A., Alben, J. O., Gennis, R.B.,
and Wikström, M. (1993) Substitution of asparagine for
aspartate-135 in subunit I of the cytochrome bo ubiqui-
nol oxidase of Escherichia coli eliminates proton-pump-
ing activity, Biochemistry, 32, 10923-10928, doi:10.1021/
bi00091a048.
94. Garcia-Horsman, J. A., Puustinen, A., Gennis, R.B., and
Wikström, M. (1995) Proton transfer in cytochrome bo
3
ubiquinol oxidase of Escherichia coli: second site muta-
tions in subunitI that restore proton pumping in the de-
coupled mutant Asp135Asn, Biochemistry, 34, 4428-4433,
doi:10.1021/bi00013a035.
95. Siletsky, S. A., Borisov, V. B., and Mamedov, M. D. (2017)
PhotosystemII and terminal respiratory oxidases: molecu-
lar machines operating in opposite directions, Front. Bios-
ci. (Landmark Ed), 22, 1379-1426, doi:10.2741/4550.
96. Shimokata, K., Katayama, Y., Murayama, H., Sue-
matsu, M., Tsukihara, T., Muramoto, K., Aoyama, H.,
Yoshikawa, S., and Shimada, H. (2007) The proton
pumping pathway of bovine heart cytochrome c oxidase,
Proc. Natl. Acad. Sci. USA, 104, 4200-4205, doi:10.1073/
pnas.0611627104.
97. Yoshikawa, S., Shimada, A., and Shinzawa-Itoh, K.
(2015) Respiratory conservation of energy with dioxygen:
cytochrome c oxidase, Met. Ions Life Sci., 15, 89-130,
doi:10.1007/978-3-319-12415-5_4.
98. Capitanio, N., Palese, L. L., Capitanio, G., Martino, P.L.,
Richter, O. M., Ludwig, B., and Papa, S. (2012) Allosteric
interactions and proton conducting pathways in proton
pumping aa
(3)
oxidases: heme a as a key coupling element,
Biochim. Biophys. Acta, 1817, 558-566, doi: 10.1016/
j.bbabio.2011.11.003.
99. Rich, P. R., and Marechal, A. (2013) Functions of the hydro-
philic channels in protonmotive cytochromec oxidase, J.R.
Soc. Interface, 10, 20130183, doi:10.1098/rsif.2013.0183.
100. Sharma, V., Jambrina, P. G., Kaukonen, M., Rosta,E.,
and Rich, P. R. (2017) Insights into functions of the
H channel of cytochrome c oxidase from atomistic mo-
lecular dynamics simulations, Proc. Natl. Acad. Sci. USA,
114, E10339-E10348, doi:10.1073/pnas.1708628114.
101. Wikström, M., and Verkhovsky, M. I. (2007) Mechanism
and energetics of proton translocation by the respirato-
ry heme-copper oxidases, Biochim. Biophys. Acta, 1767,
1200-1214, doi:10.1016/j.bbabio.2007.06.008.
102. Salje, J., Ludwig, B., and Richter, O.-M. H. (2005) Is a
third proton-conducting pathway operative in bacterial
cytochrome c oxidase? Biochem. Soc. Trans., 33, 829-831,
doi:10.1042/BST0330829.
103. Lee, H.-m., Das, T. K., Rousseau, D. L., Mills, D., Fer-
gusson-Miller, S., and Gennis, R. (2000) Mutations in
the putative H-channel in the cytochromec oxidase from
Rhodobacter sphaeroides show that this channel is not im-
portant for proton conduction but reveals modulation of
the properties of heme a, Biochemistry, 39, 2989-2996,
doi:10.1021/bi9924821.
104. Pfitzner, U., Odenwald, A., Ostermann, T., Weingard,L.,
Ludwig, B., and Richter, O. M. H. (1998) Cyto-
chrome c oxidase (heme aa
3
) from Paracoccus denitri-
ficans: analysis of mutations in putative proton chan-
nels of subunit I, J. Bioenerg. Biomembr., 30, 89-97,
doi:10.1023/A:1020515713103.
105. Kuznetsova, S. S., Azarkina, N. V., Vygodina, T. V., Siletsky,
S.A., and Konstantinov, A.A. (2005) Zink ions as cyto-
chromec oxidase inhibitors: two sites of action, Biochemistry
(Moscow), 70, 128-136, doi:10.1007/s10541-005-0091-6.
106. Sugitani, R., and Stuchebrukhov, A. A. (2009) Molec-
ular dynamics simulation of water in cytochrome c oxi-
dase reveals two water exit pathways and the mechanism
of transport, Biochim. Biophys. Acta, 1787, 1140-1150,
doi:10.1016/j.bbabio.2009.04.004.
107. Medvedev, D. M., Medvedev, E. S., Kotelnikov, A. I., and
Stuchebrukhov, A.A. (2005) Analysis of the kinetics of the
membrane potential generated by cytochrome c oxidase
upon single electron injection, Biochim. Biophys. Acta,
1710, 47-56, doi:10.1016/j.bbabio.2005.08.008.
108. Siletsky, S., and Konstantinov, A. A. (2006) Electrogenic
mechanism of cytochromec oxidase, The 8th European Bi-
ological Chemistry Conference, p.80.
MECHANISM OF MEMBRANE POTENTIAL GENERATION 1527
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
109. Siletsky, S. A., Han, D., Brand, S., Morgan, J. E., Fabi-
an, M., Geren, L., Millett, F., Durham, B., Konstantinov,
A.A., and Gennis, R.B. (2006) Single-electron photore-
duction of the P
M
intermediate of cytochromec oxidase,
Biochim. Biophys. Acta, 1757, 1122-1132, doi: 10.1016/
j.bbabio.2006.07.003.
110. Sugitani, R., Medvedev, E. S., and Stuchebrukhov, A.A.
(2008) Theoretical and computational analysis of the
membrane potential generated by cytochrome c oxidase
upon single electron injection into the enzyme, Biochim.
Biophys. Acta, 1777, 1129-1139, doi: 10.1016/j.bbabio.
2008.05.006.
111. Kaila, V. R., Sharma, V., and Wikström, M. (2011)
The identity of the transient proton loading site of the
proton-pumping mechanism of cytochrome c oxi-
dase, Biochim. Biophys. Acta, 1807, 80-84, doi: 10.1016/
j.bbabio.2010.08.014.
112. Popovic, D. M., and Stuchebryukhov, A. A. (2004) Proton
pumping mechanism and catalytic cycle of cytochromec
oxidase: Coulomb pump model with kinetic gating, FEBS
Lett., 566, 126-130, doi:10.1016/j.febslet.2004.04.016.
113. Lu, J., and Gunner, M. R. (2014) Characterizing the
proton loading site in cytochromec oxidase, Proc. Natl.
Acad. Sci. USA, 111, 12414-12419, doi: 10.1073/pnas.
1407187111.
114. Rich, P. R. (1995) Towards an understanding of the chem-
istry of oxygen reduction and proton translocation in the
iron-copper respiratory oxidases, Aust. J. Plant Physiol.,
22, 479-486, doi:10.1071/PP9950479.
115. Wikström, M., and Sharma, V. (2018) Proton pumping by
cytochrome c oxidase – a 40 year anniversary, Biochim.
Biophys. Acta Bioenergetics, 1859, 692-698, doi: 10.1016/
j.bbabio.2018.03.009.
116. Zimmermann, B. H., Nitsche, C. I., Fee, J. A., Rusnak,F.,
and Munck, E. (1988) Properties of a copper-contain-
ing cytochrome ba
3
: a second terminal oxidase from
the extreme thermophile Thermus Thermophilus, Proc.
Natl. Acad. Sci. USA, 85, 5779-5783, doi: 10.1073/pnas.
85.16.5779.
117. Rauhamaki, V., Bloch, D. A., Verkhovsky, M. I., and Wik-
ström, M. (2009) Active site of cytochromecbb
3
, J.Biol.
Chem., 284, 11301-11308, doi:10.1074/jbc.M808839200.
118. Rauhamaki, V., Bloch, D. A., and Wikström, M. (2012)
Mechanistic stoichiometry of proton translocation by cy-
tochromecbb
3
, Proc. Natl. Acad. Sci. USA, 109, 7286-7291,
doi:10.1073/pnas.1202151109.
119. Steimle, S., van Eeuwen, T., Ozturk, Y., Kim, H. J., Brait-
bard, M., Selamoglu, N., Garcia, B. A., Schneidman-
Duhovny, D., Murakami, K., and Daldal, F. (2021)
Cryo-EM structures of engineered active bc
1
-cbb
3
type
CIII
2
CIV super-complexes and electronic communi-
cation between the complexes, Nat. Commun., 12, 929,
doi: 10.1038/s41467-021-21051-4.
120. Borisov, V. B., Siletsky, S. A., Nastasi, M. R., and Forte,E.
(2021) ROS defense systems and terminal oxidases in
bacteria, Antioxidants (Basel), 10, 839, doi: 10.3390/an-
tiox10060839.
121. Siletskiy, S., Soulimane, T., Azarkina, N., Vygodina, T.V.,
Buse, G., Kaulen, A., and Konstantinov, A. (1999) Time-
resolved generation of membrane potential by ba
3
cyto-
chrome c oxidase from Thermus thermophilus. Evidence for
reduction-induced opening of the binuclear centre, FEBS
Lett., 457, 98-102, doi:10.1016/S0014-5793(99)01019-4.
122. Siletsky, S. A., Belevich, I., Jasaitis, A., Konstantinov,
A. A., Wikström, M., Soulimane, T., and Verkhovsky,
M.I. (2007) Time-resolved single-turnover of ba
3
oxidase
from Thermus thermophilus, Biochim. Biophys. Acta, 1767,
1383-1392, doi:10.1016/j.bbabio.2007.09.010.
123. Kannt, A., Lancaster, C. R., and Michel, H. (1998)
Thecoupling of electron transfer and proton translocation:
electrostatic calculations on Paracocus denitrificans cyto-
chromec oxidase, Biophys.J., 74, 708-721, doi:10.1016/
S0006-3495(98)73996-7.
124. Rauhamaki, V., and Wikström, M. (2014) Thecauses of
reduced proton-pumping efficiency in type B and C respi-
ratory heme-copper oxidases, and in some mutated vari-
ants of type A, Biochim. Biophys. Acta, 1837, 999-1003,
doi:10.1016/j.bbabio.2014.02.020.