You are on page 1of 9

Biochimica et Biophysica Acta 1857 (2016) 863–871

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

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

Review

Oxidation of NADH and ROS production by respiratory complex I☆


Andrei D. Vinogradov ⁎, Vera G. Grivennikova
Department of Biochemistry, School of Biology, Moscow State University, Moscow 119991

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

Article history: Kinetic characteristics of the proton-pumping NADH:quinone reductases (respiratory complexes I) are reviewed.
Received 29 September 2015 Unsolved problems of the redox-linked proton translocation activities are outlined. The parameters of complex
Received in revised form 2 November 2015 I-mediated superoxide/hydrogen peroxide generation are summarized, and the physiological significance
Accepted 7 November 2015
of mitochondrial ROS production is discussed. This article is part of a Special Issue entitled Respiratory complex
Available online 10 November 2015
I, edited by Volker Zickermann and Ulrich Brandt.
Keywords:
© 2015 Elsevier B.V. All rights reserved.
NADH:quinone oxidoreductase
Proton pumping
Hydrogen peroxide
Superoxide
Bacterial plasma membranes
Mitochondria

1. Introduction bound to a membrane freely permeable for protons catalyze reaction


(1) irreversibly due to a significant thermodynamic gap between the
Respiratory complex I (proton-translocating NADH:quinone oxido- standard redox potentials of NADH/NAD+ (− 320 mV) and QH2/Q
reductases, mitochondrial complex I, bacterial NDH-1)1 catalyzes a (+60 mV).
reaction of great metabolic significance: The membrane-bound and all solubilized or detergent-dispersed
preparations of the enzyme catalyze the irreversible reaction:

ð1Þ NADH þ 2Aox →NADþ þ 2Ared þ 2Hþ ð2Þ


where Aox and Ared stand for artificial oxidized and reduced forms
where Q stands for the natural membrane-bound quinone (ubiquinone of the A red /A ox couple. Ferricyanide 3 − anion (Ferri) and/or
with various number of isoprenoid units at position 6 of benzoquinone hexaammineruthenium3 + cation (HAR) are routinely used in assays
ring) or menaquinone, H+ +
in and Hout stand for intramitochondrial of the enzyme. Reaction (2) is not coupled to proton translocating activ-
(cytoplasmic) and extruded protons, respectively, and p.m.f. (proton- ity. Membrane-bound and various solubilized preparations of complex I
motive force) is the transmembrane gradient of electrochemical activity as well as their FMN-containing fragments catalyze the reversible
of H+ (Δμ~ Hþ ). NADH:APAD+ transhydrogenase reaction. Mitochondrial and bacterial
Reaction (1) is only reversible when the oxidoreduction is tightly complex I also catalyzes NADH oxidation by oxygen:
coupled to proton pumping activity. Thus, this reversible reaction only
proceeds when the enzyme is bound to a membrane that is poorly NADH þ О2 þ Нþ →NADþ þ Н2 О2 ð3Þ
permeable to protons. The enzyme detached from the membrane or
and/or
+
Abbreviations: APAD , 3-acetylpyridine adenine dinucleotide; Ferri, ferricyanide;
FMN, flavin mononucleotide; HAR, hexaammineruthenium(III); Qn, 2,3-dimethoxy,5- NADH þ 2О2 →NADþ þ 2O•2 þ Hþ ð4Þ
methyl,6-(n)isoprenyl,1,4-benzoquinone; ROS, reactive oxygen species; SMP, submito-
chondrial particles; O•2 , superoxide anion; SOD, superoxide dismutase; p.m.f., proton-
motive force (transmembrane gradient of electrochemical activity of H+, Δμ~ Ηþ ). the formation of so-called ROS (reactive oxygen species – hydrogen
☆ This article is part of a Special Issue entitled Respiratory complex I, edited by Volker peroxide and superoxide) at the highest rate of about 0.2–0.3% of
Zickermann and Ulrich Brandt. those depicted by Eqs. (1) and (2).
⁎ Corresponding author. Spectacular progress in visualization of atomic structures of the frag-
E-mail address: adv@biochem.bio.msu.su (A.D. Vinogradov).
1
Bacterial plasma membrane proton pumping NADH:quinone oxidoreductases (NDH-1)
ments [1,2] and the entire complex I [3–5] isolated from various sources
and corresponding mitochondrial enzymes are designated as complex I throughout the text have been achieved during the last decade. However, as important as
for the sake of simplicity. the information provided by the molecular architecture of the enzyme

http://dx.doi.org/10.1016/j.bbabio.2015.11.004
0005-2728/© 2015 Elsevier B.V. All rights reserved.
864 A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871

is, it serves only as excellent groundwork for hypotheses on possible of the enzyme, and how it can be regulated in situ? Natural ubiquinones
catalytic mechanisms. We believe that studies on the catalytic activities are practically insoluble in aqueous media, and their analogs or homo-
still are and will be the powerful tool for understanding the mechanistic logs at “saturating” concentrations of about 100 μM are routinely used
and, perhaps more physiologically important, regulatory properties of for assays of the membrane-bound or detergent-solubilized enzyme.
complex I. The content of ubiquinone in heart inner mitochondrial membrane is
This short review discusses available information on some features about 5 nmol per mg of protein [37]. The bulk quinone is located in
of eukaryotic and bacterial complex I as they appear from studies of the lipid phase of the membrane. The content of phospholipids in the
the reactions (1 and 2). We will also discuss reactions (3 and 4) and inner mitochondrial membrane (bovine heart) is about 0.4 mg per mg
their possible physiological significance. Because little information is of protein [37]. Simple approximate calculations show that the concen-
available on the catalytic properties of plant complex I, they are not tration of ubiquinone in the lipid phase is as high as about 10 mM. Thus,
discussed here; an interested reader is addressed to Refs. [6,7]. in situ quinone reduction by complex I proceeds at concentrations of the
substrate of about two orders of magnitude higher than that used under
2. Intramolecular electron transfer standard assay systems with added artificial quinone acceptors. The
true concentration of “water-soluble” quinones in the membranes and
Complex I contains up to ten active redox components (FMN, 7–8 their respective reactivity certainly depend on the lipid/water partition
iron-sulfur clusters, and bound quinone(s)). The structures of complex coefficient [38]. The sequence of events during the steady-state
I and fragments derived therefrom [1–4] unambiguously suggest the NADH:externally added quinone-acceptor reductase reaction catalyzed
following sequence of intramolecular electron transfer on the way by particulate complex I is not clear. External quinone might either
from NADH to the final quinone acceptor: directly bind to the ubiquinone-specific site or oxidize tightly bound
Q, or oxidize bulk natural reduced quinone swimming in the lipid phase.
NADH→FMN→N3ðN1aÞ→N1b→N4→N5→N6a→N6b→N2→Q ð5Þ In studies on Q-pool behavior in the respiratory chain [39] recently
During steady-state coupled or uncoupled NADH oxidation by bo- confirmed by others [40], linear dependence of NADH oxidase activity
vine heart submitochondrial particles (SMP), all EPR-detectable iron- of bovine heart SMP on the content (concentration) of oxidized ubiqui-
sulfur centers of complex I are almost completely reduced [8,9]. Thus, none have been demonstrated. If extrapolated to the state where
the rate-limiting step of the overall NADH oxidase is reoxidation of ubiquinone is fully oxidized, the data would correspond to turnover
the terminal electron-transferring component (N2, or specifically and number of complex I of about 500 s−1 (pH 7.4, 25 °C), a value that is
tightly bound quinone) by bulk quinone. Only a very short description substantially higher than those (averaged) measured at saturating con-
of the intramolecular electron transfer is needed for further discussion centrations of water-soluble quinones or in steady-state NADH oxidase
of the steady-state catalytic activity. Recently, the kinetics of intramo- assays (~250 s−1, see above). This is not surprising because during the
lecular electron transfer in purified Escherichia coli complex I using steady state, fully uncoupled NADH oxidase operates at Qred/Qox ratio
ultrafast freeze-quenching techniques followed by EPR and optical of about 1. Linear, not hyperbolic, dependence of the NADH oxidase
spectroscopic analysis have been reported [10–12]. The results obtained rate on molar fraction of oxidized ubiquinone (Qox/(Qox + Qred)) has
by these groups and their interpretation significantly differ. Biphasic been demonstrated [39]. Such dependence is expected either if the
reduction of iron-sulfur clusters N1a and N2 (~ 50 μsec) followed total concentration of quinone is much lower than the apparent Km for
by much slower reduction of N1b and N6b was interpreted by Qox or if Qred competes with Qox for binding at the reactive site with
Verkhovskaya et al. so as to suggest that NAD+ dissociation from the similar affinity. The first possibility seems to be unlikely because of
nucleotide-binding site is the rate-limiting step in the overall reaction high concentration of ubiquinone in the membrane (see above). Kinetic
(rapid oxidation of the first NADH molecule results in reduction of competition between oxidized and reduced forms of ubiquinone (with
N1a and N2, and further reduction of the enzyme proceeds significantly similar affinity) seems the more plausible explanation and corroborates
more slowly because the next NADH molecule can bind to the empty with well-established reversibility of reaction (1). Weak (competitive
active site only after dissociation of NAD+) [10,11]. De Vries et al. with oxidized Q1) inhibition of NADH:Q1 reductase activity catalyzed
interpreted their result as to suggest that rapid reduction of N2 by highly purified bovine heart complex I by the product (Q1H2) have
results in six-fold decrease in electron tunneling rate between other been reported [41]. Also, a competitive relation between oxidized
iron-sulfur centers, thus synchronizing electron transfer with the and reduced quinone acceptor (Q2) was documented for respiratory
proton pumping activity [12]. Whatever interpretation is correct, the complex II [42].
enzyme turnover when it operates as a member of the respiratory The linear dependence of NADH oxidase activity on the molar frac-
chain (~hundreds per sec) is substantially slower than any intramolec- tion of oxidized ubiquinone might shed some light on the mechanism
ular redox reaction. of the respiratory control phenomenon. NADH oxidation in controlled
state 4 can equally be slow either because p.m.f. inhibits proton
translocating activity (see Eq. (1)) or because an increase in reduced
3. Steady-state activity
ubiquinone concentration due to a decrease in Q-pool oxidation
by further components of the respiratory chain (complex III and
The accumulated data on NADH-dehydrogenating activities of
cytochrome c reductase), which are under the control of p.m.f. The rela-
membrane-bound complex I scattered in the literature are given in
tive contributions of “thermodynamic” (p.m.f.) and “kinetic” (Qred/Qox
Table 1. Averaged NADH oxidase activity of fully activated2 uncoupled
ratio) factors to the respiratory control phenomenon at the level of
bovine heart SMP is about 1.5 μmol of NADH oxidized per min per mg
complex I remain to be established.
of protein (Table 1), a value that is close to that seen at “saturating”
concentrations of artificial quinone acceptors and other analogs or
4. Artificial electron acceptors
homologs of ubiquinone (Q1). Assuming a molar content of complex I
in heart mitochondrial membranes of about 0.1 nmol per mg of protein
Ferricyanide (Ferri) [43,44] and hexaammineruthenium III (HAR)
[37], these numbers correspond to maximal steady-state enzyme
[21] are the most frequently used artificial electron acceptors for quan-
turnover of ~ 250 s−1. Does this turnover reflect the maximal capacity
titation of complex I catalytic activity of preparations of different degree
2
of resolution. The substrate (NADH) concentration–activity profile
The mitochondrial enzyme exhibits complex kinetics in the quinone reductase activity
due to slow transition between catalytically inert D- and active A-forms (operationally
for Ferri reductase catalyzed by the membrane-bound or soluble, so-
called the A/D-transition). This phenomenon reviewed in Refs [35,36] is beyond the scope called high molecular mass preparations of complex I appears as a
of the present discussion. bell-shaped curve with a relatively sharp maximum. For any
A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871 865

Table 1
Specific NADH-oxidizing activities of mitochondrial or plasma membrane respiratory complex I (representative values) a

Activity (μmol/min per mg)

NADH oxidase NADH:quinone reductase NADH:ferricyanide reductase NADH:HAR reductase NADH:APAD+ reductase
(Ferri 0.5 mM) (HAR 2 mM)

Eukaryotes
Bos taurus 0.7–2.0 [13–17] 0.3–1.6 [13,18,19] 2.0–8.4 [20,21] 18.0b [22] 0.8 [24]
2.7–8.0 [20,21,23]
Y. lipolytica 0.2 [25] 0.4 [25] 0.9–1.2 [25,26]
N. crassa 0.2–0.3 [27] 0.1–0.4 [27,28] 2.5 [28]
Prokaryotes
E. coli 0.4 [29] 0.4–3.1 [29,30] 1.4–4.6 [29–31] 2.2d [32]
P. denitrificans 0.8 [14,15,33,34] 0.8 [34] 0.7 [20] 2.0 [20]
1.6–2.1c [33]
R. capsulatus 0.5 [34] 0.5 [34]
a
Specific activities reported in the literature even for the same type of preparations are inherently variable due to somewhat different assay conditions. This is particularly true for data
where artificial electron acceptors were used. To the best of our knowledge, no study was reported where oxidase and acceptor reductase activities extrapolated to infinity concentrations
of the latter have been compared.
b
Extrapolated to infinity HAR concentration.
c
1 mM Ferri in the reaction mixture.
d
0.35 mM HAR in the reaction mixture.

oxidoreductase reaction, the enzyme turnover number or Vmax is mean- proposed by Birrell et al. [45]. Whether an allosteric ATP-binding site
ingful only if the activity determined at various concentration of an elec- presumably located at some accessory subunit [49] of eukaryotic com-
tron acceptor is extrapolated to its infinite concentration. Because of the plexes has any physiological significance is an open question. Only the
evidently non-hyperbolic NADH concentration–activity profile, such HAR reductase activity of eukaryotic complex I is affected by ATP.
extrapolation is somewhat ambiguous, and great precautions should Recently, a striking observation relevant to HAR reductase activity
be taken if complex I catalytic activity is to be quantitated using Ferri was reported by Varghese et al. [50]. A point mutation (A341V) in the
as the electron acceptor. Y. lipolytica FMN-containing subunit eliminated NADH:HAR reductase
The simplest mechanistic explanation for the bell-shaped depen- activity, leaving all other activities including paraquat reductase unaf-
dence is that the reaction proceeds by a ping-pong mechanism, and fected [50]. This observation suggests a different reaction mechanism
NADH (at high concentrations) binds to the active site of the reduced for paraquat and HAR reduction and demonstrates that our current
enzyme, thus sterically protecting accessibility of the Ferri anion to knowledge on the reduction of the artificial electron acceptors is far
FMN buried in a deep cleft. This interpretation hardly agrees with the ki- from complete.
netics of NADH-HAR reductase: it proceeds according to a ternary com-
plex mechanism and shows no inhibition at high NADH concentration 5. Proton pumping activity
[20,21,45]. Thus, the kinetic behavior makes HAR a convenient acceptor
for studies directed to quantitation of the enzyme turnover or its con- At present, the molecular mechanism of the redox-linked proton
tent (that is assumed to be proportional to Vmax) in various membrane translocation at coupling Site 1 remains a black box. Numerous earlier
or solubilized preparations. The specific NADH dehydrogenating activity Mitchellian direct transmembrane redox loop models should be
of membrane-bound complex I in bovine heart SMP as extrapolated to discarded in light of the structural arrangement of multiple redox com-
Vmax is at least by an order of magnitude higher than their NADH ponents within the enzyme structure. The ubiquinone reactive site
oxidase corresponding to turnover of about 2500 s−1. This number de- (iron-sulfur cluster N2) is located approximately 30 Å from the coupling
termines the lower limit for the rate of product release from the enzyme membrane plane [51,52]. The hydrophobic domain is composed of a
active site: NAD+ dissociation is an inevitable step independent of the number of subunits including three transmembrane homologs of bacte-
particular site where any electron acceptor interacts with the enzyme. rial Na+/H+ antiporters [53] (subunits L, M, and N according to E. coli
Thus, at least for the mammalian enzyme, the half time of NAD+ release nomenclature) [2]. A simple and attractive possibility is that these
should be less than 0.3 msec. This value is substantially lower than that subunits operate as the proton-conductive channels required for the
for the slow phase of N1b and N6b reduction (about 2 msec) in purified pumping mechanism. For any model of coupling, the stoichiometric
E. coli complex I interpreted as a characteristic time for NAD+ dissocia- coefficient, n in Eq. (1), should be defined. A consensus has been
tion [10,11]. reached that 4 protons are translocated per molecule of NADH oxidized
The NADH:HAR reductase activity of SMP and purified bovine and ubiquinone reduced (2ē) [54,55,and references cited therein]. Here
heart complex I at low concentrations of the acceptor is strongly we will briefly discuss several points that might cast a new perspective
(up to 10-fold) stimulated by ATP and other purine tri- and on studies of the coupling mechanism.
diphosphonucleotides [45,46]. This effect was interpreted as evidence
for an allosteric nucleotide-binding site in the mammalian enzyme (i) The basic structure of Thermus thermophilus complex I is remark-
[46] or, alternatively, as a change in the reaction pathways of the reac- ably similar to that of the E. coli and eukaryotic enzymes.
tion [45] in the presence of ATP, ADP, and ADP-ribose, competitive Menaquinone (E7m ~ −70 mV) serves as the electron acceptor in
inhibitors of the NADH/NAD+-binding site [47]. Kinetically, ATP acts T. thermophilus membranes [56], whereas ubiquinones
as a competitive (with NADH) activator, thus decreasing the apparent (E7m ~ +60 mV) are the acceptors in eukaryotic membranes. The
Km for NADH [48]. Grivennikova et al. suggest that ATP bound at a puta- thermodynamic restrictions hardly permit the T. thermophilus
tive allosteric site increases the midpoint potential of an enzyme com- enzyme and ubiquinone-specific complexes I to operate by the
ponent that donates electrons to HAR [46]. Remarkably, no activating same proton pumping mechanism. To the best of our knowledge,
effect of ATP was observed for prokaryotic Paracoccus denitrificans the value of n in Eq. (1) as applied to T. thermophilus is not known.
membranes, whereas it was seen for eukaryotic Yarrowia lipolytica In fact, among prokaryotes the n value was experimentally evalu-
SMP. This species-dependent activation by ATP hardly corroborates ated only for E. coli [57] and P. denitrificans [58,59]. Presumably,
with reaction pathway change induced by nucleotides, the mechanism the same number of proton-conducting subunits is present in
866 A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871

prokaryotic and eukaryotic enzymes. The possibility cannot be rotenone (and other multiple rotenone-like inhibitors) or by
excluded that variable stoichiometry exists in the reaction transformation of the enzyme to its inactive D-form is not
(Eq. 1), being controlled by some special regulatory mechanism known. It could well be the initial one-electron reduction of the
that turns “on” one, two, or three proton-translocating subunits specifically tightly bound Q, or accepting the second electron to
depending on particular physiological conditions (steady-state form reduced bound quinol, or dismutation of two semiquinones
p.m.f.). It does not seem unlikely that if a channel-forming subunit to form bound QH2, or electron exchange between bound QH2
is specialized for redox-linked proton translocation its functional and Q arriving from the bulk quinone pool. The key question of
state (conformation) would be voltage-dependent, as it is well what the particular step(s) is(are) coupled with proton pumping
established for a number of cation-specific transporters [60–62]. remains to be answered.
The structures of putative H+-translocating subunits N, L, and M
are different, thus their “voltage sensitivity” (if it exists) might 6. Purified complex I
also be different, resulting in variable stoichiometry of the overall
reaction (1) at different p.m.f. values. It should be emphasized that Only limited data on catalytic activities of purified complex I are
an attractive possibility of the direct participation of N, L, and M available, although several properties of the enzymes isolated from
subunits in redox-linked energy coupling remains just a proposal mammalian [65–67] and yeast [26,68] mitochondria, or plasma mem-
having no strong experimental evidence, although “partial” branes of E. coli [30–32], P. denitrificans [33], and thermophilic bacteria
uncoupling has been observed for Yarrowia complex I: variants (T. thermophilus [69], Aquifex aeolicus [70], Rhodothermus marinus [71])
in Yarrowia complex I deleted by two of the three subunits that have been described. A requirement for added phospholipids for the
are assumed to be candidates for proton pumping translocate rotenone-sensitive activity of purified bovine heart complex I was dem-
protons with half of the stoichiometry observed for the intact onstrated many years ago [72]. More recently, the same phenomenon
parental enzyme [63]. was demonstrated for the detergent solubilized purified enzymes
(ii) Counter cation translocation is required to measure proton from Bos taurus [66], Y. lipolytica [73], and E. coli [32]. The turnover
movement across the coupling membrane as a pH-change be- numbers of purified complex I are significantly lower than for complex
cause the electrical component of the p.m.f. should be dissipated. I-catalyzed NADH oxidase activity of the parent membranes. This is
Unexpectedly, valinomycin did not affect the initial rate or the explained either by an inadequate assay using externally added
maximal pH change if the proton translocation was initiated by “water-soluble” quinones (see above) or by the presence of some, yet
NADH-external quinone oxidoreduction in coupled SMP [54]. unidentified specific factor(s) required for full catalytic activity in the
Strong stimulation of proton translocation by valinomycin, as ex- natural membranes.
pected, was seen under the same experimental conditions if the
overall NADH oxidase (all three coupling sites) was activated. 7. Complex I-mediated ROS production
Formally, these data can be interpreted as to suggest that, in con-
trast to complex III and cytochrome oxidase, complex I operates Membrane-bound or purified complex I from all species studied so
solely as the ΔpH generator. Work aimed to confirm or discard far catalyze reactions (3) and/or (4). Quantitative and even qualitative
such a possibility is in progress in our laboratory. appraisals of the contribution of complex I to overall ROS production
(iii) The NADH oxidase activity of bovine heart SMP is completely by intact mitochondria are extremely difficult due to the presence of
(more than 90%) inhibited by specific N2-ubiquinone junction multiple intramitochondrial enzymes involved in the formation (for
site-directed inhibitors – rotenone and piericidin. When com- example, dihydrolipoyl dehydrogenase) and utilization of superoxide
plex I activity is assayed with externally added ubiquinone (superoxide dismutases) and hydrogen peroxide (thioredoxins, gluta-
homologs, a substantial rotenone-insensitive reaction is observed. thione peroxidase, catalase). Here we discuss only the data obtained
The trivial explanation of this phenomenon is that water-soluble for systems where ROS production can be unambiguously attributed
quinones accept electrons from some other than the natural to the membrane-bound or purified complex I. NADH or succinate
ubiquinone reactive site, just as other artificial acceptors do oxidation by tightly coupled bovine heart SMP is accompanied by super-
(Ferri, HAR). However, studies on the proton pumping activity of oxide production at the rate of about 1 nmol/min per mg of protein
bovine heart SMP do not agree with this simple interpretation corresponding to about 0.2–0.3% of the total oxygen consumption. The
[54]. Rotenone-inhibited particles catalyze the NADH:Q1 reductase contribution of particular respiratory complexes to the total superoxide
reaction coupled with proton translocation with the same stoi- production can be quantitatively evaluated as shown in Table 2. The
chiometry of 4 H+/2ē [54]. Moreover, the irreversibly stabilized highest production, which is inhibited by rotenone and uncouplers, is
D-form of complex I, which is phenomenologically equivalent seen during coupled succinate oxidation, thus suggesting that about
to the rotenone-inhibited enzyme, also shows proton 70% of one-electron oxygen reduction proceeds via energy-dependent
translocating activity with the same stoichiometry [64]. These reverse electron transfer (reversal of reaction (1)). The residual
unexpected observations are certainly relevant to the mecha- rotenone-insensitive generation corresponds to the combined activities
nism of redox coupled proton translocation. The particular step of complexes II and III. It was somehow surprising that the rate of gen-
of the electron transfer from N2 to bulk ubiquinone blocked by eration during more rapid coupled NADH oxidation is about three-fold

Table 2
Oxidase activities and superoxide generation by bovine heart coupled SMP (pH 8.0, 30 °C)a

Substrate-donor Oxidase activity, μmol/min per mg Generation of O•2 , nmol/min per mg

– uncoupler (state 4) + uncoupler (state 3) – uncoupler (state 4) + uncoupler (state 3)

1. Succinate (10 mM) 0.3 0.9 1.0 0.1


+ rotenone 0.3 – 0.3 –
2. Succinate (10 mM) + NADH (1 mM) 0.7 – 0.3 –
3. NADH (1 mM) 0.6 1.9 0.4 0.1
4. NADH (50 μM) 0.6 1.8 1.0 0.9
+ rotenone b0.002 b0.002 1.4 –
a
Adapted from Ref. [17,74].
A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871 867

lower than the succinate-supported reaction. Moreover, the addition of of the wild and mutated strains. Normalization of their data to the
NADH (1 mM) to a sample where complex I produces superoxide via “natural” activities shows 60% quinone reductase. This agrees with
succinate-supported reverse electron transfer inhibits the reaction. data reported previously by the same group [79] and originally
The superoxide production increases and reaches the level of the interpreted as to suggest that the mutants lacking N2 were still capable
succinate-supported rate at low concentration of NADH. The depen- of ubiquinone reduction at near normal rates, a possibility that seems
dences of superoxide production and ferricyanide reductase activity of hardly probable. Rather, the R141M mutation changes the spectral
complex I on NADH concentration are almost the same: both appear and redox properties of N2, as later demonstrated by Zwicker et al. for
as bell-shaped curves with a maximum at about 50 μM [17,75]. The the H226M mutant, showing 80 mV negative shift of N2 midpoint
simplest interpretation of bell-shaped dependence is that oxygen potential, absence of its pH-dependence, and still being capable of
reduction proceeds via a ping-pong mechanism, and the accessibility redox-linked proton translocation with unaltered stoichiometry [80].
of a one-electron donating site for oxygen is restricted in the reduced We believe that exclusion of iron-sulfur cluster N2 as a possible site of
enzyme–NADH complexes. Other possible interpretations, such as the superoxide production should wait for more experimental verification.
presence of two NADH/NAD+-binding sites in mammalian complex I, Hydrogen peroxide formation depends on NADH concentration quite
have been discussed [17]. Superoxide production at low (optimal) differently. At low substrate concentrations (up to 3 μM), no production
concentration of NADH is activated by rotenone. At very low NADH is seen (Fig. 1), and the rate reaches a constant value at higher (up to
concentrations, purified bovine heart complex I produces ROS mostly millimolar) range of NADH concentration, where its relative contribu-
(90%) as superoxide [18], whereas purified E. coli enzyme generates tion to the overall ROS production is about 60% [75].
ROS as hydrogen peroxide [76]. Recently, membrane-bound complex I Complex I-catalyzed ROS production is inhibited by μmolar NAD+
(SMP) as well as the purified enzyme were shown to generate both concentrations [17]. Because of the very low activity of complex I in
hydrogen peroxide and superoxide [75,77]. The partitioning between ROS generation as compared with major oxidase or NADH:artificial
the products depends on NADH concentration as shown in Fig. 1. The acceptor reductase reactions, it is safe to assume that all redox compo-
superoxide production reaches a maximum at 10–50 μM NADH and nents of the enzyme are in equilibrium with the NAD+/NADH couple
gradually decreases in the millimolar range of the substrate [17,75]. during the steady-state reaction (the contribution of the kinetic term
The apparent KNADH
m as determined from the linear double-reciprocal to the apparent Km is negligible). Thus, the dependence of ROS produc-
plot for the ascending part of the superoxide production titration tion on NAD+/NADH ratio is indicative of the midpoint redox potential
curve is as low as about 0.5 μM [75]. As noted above (see 4. Artificial of the component(s) reacting with oxygen. The NAD+/NADH ratios
electron acceptors section), the apparent Km value for the substrate- reported in the literature for half-maximal ROS production by complex
donor in the reaction catalyzed by any oxidoreductase depends on the I are greatly variable (from 0.01 up to 7.0) [8,18,81–83]. Titration of
redox potential gap between the primary electron acceptor (FMN for complex I by the NAD+/NADH couple is not a true redox titration
complex I) and the component that reacts with an acceptor. The very because the relative binding affinities of the reduced and oxidized
low KNADH
m for superoxide production [18,75] indicates that a compo- enzyme to NAD+ and NADH [84] can significantly contribute to the
nent immediately reacting with oxygen has a substantially higher apparent (NAD+/NADH)0,5 ratio required for ROS production. This
redox potential than FMN. The iron-sulfur cluster N2 might serve as a ratio is dependent on the total pool nucleotide concentration. For exam-
one-electron donor for oxygen reduction as it has been proposed by ple, it decreases from 0.2 to 0.05 for total ROS production by SMP when
Genova et al. [78]. This cluster is located close to the ubiquinone- total NAD+ plus NADH concentration increases from 50 to 500 μM [83].
binding site in a funnel-like area at the distance of 25–30 Å from the At optimal nucleotide concentration (50 μM), the rate of Н2О2
membrane plane [51,52]. All other iron-sulfur centers are well insulated production fits the Nernst equation for a two-electron reaction with
by the protein environment. A direct approach to exclude cluster N2 as midpoint redox potential of − 350 mV ((NAD+/NADH)0.5 = 0.13)
the site of one-electron oxygen reduction was used by Galkin and [75], a value close to the midpoint potential of FMN in complex I
Brandt [25]. They found that a variant form of Y. lipolytica isolated (− 370 mV at рН 8.0) [85]. The same dependence for superoxide
complex I (R141M) that showed no EPR-detectable center N2 produced production does not fit the Nernst equation, and half-maximal activity
superoxide with the same rate (96%) as did the enzyme from the wild is observed at a significantly higher NAD+/NADH ratio (0.33) [75].
strain. They normalized their data to NADH:HAR reductase activities Superoxide production does proceed even when the nucleotide pool is

Fig. 1. Partitioning between superoxide and hydrogen peroxide as the products of complex I-mediated ROS production. Red and blue bars are superoxide and hydrogen peroxide,
respectively. The lines on left diagram are drawn to emphasize complex kinetics of the NADH concentration dependence. Adapted from Ref. [75].
868 A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871

90% oxidized, i.e. at much more positive potential than that of the was approximated as about 8 [101]. In isolated cardiomyocytes respir-
FMNH−/FMN couple. The dependence of superoxide production by ing on 10 mM glucose [102] or in heart perfused by 10 mM glucose,
purified complex I on the NAD+/NADH ratio measured by Kussmaul this ratio is closed to 1 [103]. Thus, the apparent redox potential in the
and Hirst fits the theoretical curve of the two-electron titration of the matrix is significantly higher than that conventionally used in the
FMNH−/FMN couple with midpoint potential of −360 mV [18]. model experiments on ROS production. It appears that in vivo complex
A single amino acid replacement (E95, a glutamate residue located I-mediated ROS production is many-fold lower than measured under
near FMN) by glutamine in E. coli complex I results in a 15-fold increase experimentally “optimal” conditions.
in the rate of NADH-dependent hydrogen peroxide production [86]. The In recent years, the term “p.m.f.(or energy)-dependent ROS-
authors proposed the existence of closed and open states of the production” is frequently used in the literature, thus implicitly assuming
nucleotide-binding site in the FMN-containing subunit with different that the membrane energization and deenergization increases and de-
reactivity (accessibility) of flavin to oxygen, a model where the E95 creases ROS production, respectively. We believe that this terminology
residue keeps the site in the closed state. This observation seems to be can be misinterpreted by a reader who is only superficially familiar
relevant to the well-known stimulatory effects of guanidine on the with mitochondrial bioenergetics. The production of ROS depends on
ferricyanide reductase [87] and superoxide producing activities [75] of the amount (concentration) of reactive sites accessible to oxygen. The
mammalian complex I. The electrostatic interaction between the large steady-state NADH/NAD+ ratio, indeed, is decreased when respiration
guanidinium cation and the glutamate anion might stabilize the open is activated by ADP (small drop of р.m.f.) or by uncoupler (complete
state. Interestingly, this negatively charged/neutral residue replacement dissipation of р.m.f.). However, it seems important to emphasize
was accompanied by about 5-fold decrease in HAR reductase activity that р.m.f. itself does not affect the generation rate; it influences redox
[86]. state of the oxygen reactive sites. No correlation between membrane
The well-known “burst” of ROS upon anaerobic–aerobic state transi- energization and ROS production exists; for example, rotenone stimu-
tion, such as organ reperfusion after ischemia, is a phenomenon that lates mitochondrial ROS production, whereas it completely deenergizes
might be relevant to the unusual hysteretic kinetics of complex I. If no membranes. It might appear that this note is a matter of semantics;
oxidized ubiquinone is available, the enzyme is transformed to the so- however, the expression “energy-dependent ROS generation” (by
called deactivated state where electron transfer from N2 to ubiquinone complex I) is misleading.
is blocked (see Refs. [35,88] for reviews). The active (A) state-to-
deactivated state (D) transition has been detected for isolated complex 8. Notes on physiological and pathophysiological significance of
I [89], SMP [13], intact heart mitochondria [90], and in ex vivo studies of complex I and other mitochondrial enzyme-mediated
perfused hearts [91]. In terms of catalytic activities, deactivation of ROS production
complex I is equivalent to its inhibition by rotenone, which is known
to increase ROS production. The back transformation of the D- to A- Mitochondria contain a number of enzymes capable of ROS genera-
form is a slow process that is inhibited by free fatty acids and divalent tion. Their relative contributions along with complex I to the level of
metal cations [92–94]. Taken together, these data suggest the following intracellular hydrogen peroxide are expected to be species- and
scenario of the normal state → ischemia → reperfusion transition. tissue-dependent. To the best of our knowledge, only one study
Negligible complex I-mediated ROS production under the initial normal published more than 40 years ago was designed for very approximate
state occurs. It stops when no oxygen is available, and complex I become quantitation of the relative contribution of various enzyme systems to
deactivated. A sudden increase in oxygen to the normal level would H2O2 production in rat liver [104]. According to that report, about 15%
result in a burst of ROS because the deactivated enzyme will be directly of intracellular hydrogen peroxide is produced by mitochondria. The
oxidized by oxygen, not by ubiquinone. Increased ROS production is widespread opinion that mitochondria are the major source of ROS,
expected for the time needed for the slow D-to-A transformation and apparently because mitochondria are indeed the major consumers of
restoration of the normal ubiquinone reductase activity. oxygen, is exemplified by citing D. Harman who wrote, “Mitochondria
When discussing forward and reverse electron transfer in complex would be expected to be particularly subject to free radical induced
I-mediated ROS production, a note should be made concerning the use change as over 90% of oxygen utilized by mammals takes place in
of succinate as the respiratory substrate. The respiratory chain of them” [105]. This and other similar statements circulating in the litera-
coupled mitochondria or SMP oxidizing externally added succinate ture are somehow misleading. Mitochondrial respiration inevitably
cannot be considered as a model of any physiologically conceivable decreases the local concentration of oxygen in the vicinity of the centers
situation. Succinate, an intermediate of the Krebs cycle, is produced potentially capable of ROS production, which results in a decrease in the
and utilized in the mitochondrial matrix, providing one fifth of the generation rate.
reducing equivalents during complete oxidation of pyruvate. No other The increase in atmospheric oxygen in the history of the Earth has
quantitatively significant cytoplasmic sources of succinate exists in led to great improvement in the energetics of life. On the other hand,
aerobic metabolic pathways (α-oxoglutarate-dependent proline it also necessitates protection against undesired deteriorating oxidative
hydroxylation [95] and succinic semialdehyde transformation [96] are reactions. One strategy to solve the problem is to protect potentially
minor contributors to the total respiratory activity of mammalian reactive centers (flavins and/or iron-sulfur clusters) by the specific
tissues). This by no means excludes reverse electron transfer as a arrangement of the protein structure, including attachment of addition-
pathway for ROS production under some pathophysiological condi- al subunits that have no other function than to avoid their oxygen
tions, such as anoxia, where a significant amount of intramitochondrial reactivity. Numerous experimental data show that almost all flavo-
succinate is accumulated [97]. Also, ubiquinol, the actual substrate for and iron-sulfur proteins including those knowingly not operating
the reversal, is produced in several mitochondrial metabolic pathways functionally as oxidases do react with oxygen, thus producing either su-
such as fatty acid β-oxidation or oxidation of α-glycerophosphate. peroxide or hydrogen peroxide. In other words, their specific protection
As discussed above, ROS-producing activity of complex I depends on is not perfect. This is compensated by the widespread presence of SODs,
the matrix redox potential, i.e. NAD+/NADH ratio [83]. The total amount peroxidases, and catalase. Does mitochondrial ROS production have any
of pyridine nucleotides in heart mitochondria can be approximated as physiological function? Numerous reports have recently appeared in
4–7 nmol per mg of protein [98–100], these values corresponding the literature where signaling cascade reactions with the participation
to 4–7 millimolar concentration. It can thus be assumed that the of hydrogen peroxide are suggested and discussed (see for example re-
nucleotide-binding sites of complex I are always saturated. However, view [106]). Note should be made that signaling molecules are defined,
the concentrations of free NAD+/NADH nucleotides are not known. in contrast to metabolites, as those that bind to a receptor and induce its
The physiologically relevant NAD+/NADH ratio in liver mitochondria structural change without their chemical transformation. The structural
A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871 869

change of a receptor then leads to activation (or inhibition) of some [4] K.R. Vinothkumar, J. Zhu, J. Hirst, Architecture of mammalian respiratory complex I,
Nature 515 (2014) 80–84.
metabolic pathways. Many hormones, second messengers (c-AMP, [5] V. Zickermann, C. Wirth, H. Nasiri, K. Siegmund, H. Schwalbe, C. Hunte, U. Brandt,
Са2 +), and protein factors are true signaling molecules. If expanded Structural biology. Mechanistic insight from the crystal structure of mitochondrial
(unjustified in our opinion), the meaning of “signaling molecule” complex I, Science 347 (2015) 44–49.
[6] J.L. Heazlewood, K.A. Howell, A.H. Millar, Mitochondrial complex I from
could be applied to any metabolite. In most schemes, the mechanism Arabidopsis and rice: orthologs of mammalian and fungal components coupled
of hydrogen peroxide “signaling” is postulated as a trivial nonenzymatic with plant-specific subunits, Biochim. Biophys. Acta 1604 (2003) 159–169.
oxidation of deprotonated sulfhydryl groups of the “target” that results [7] E.H. Meyer, Proteomic investigations of complex I composition: how to define a
subunit? Front. Plant Sci. 3 (2012) 106.
in their catalytic (or further signaling) activities. It appears that if a reg- [8] A.I.G. Krishnamoorthy, P. Hinkle, Studies on the electron transfer pathway, topog-
ulatory signaling cascade does exist, all its steps must be enzymatically raphy of iron-sulfur centers, and site of coupling in NADH-Q oxidoreductase, J. Biol.
catalyzed, as is evident for the classic cascade regulation of glycogen Chem. 263 (1988) 17566–17575.
[9] D.S. Burbaev, I.A. Moroz, A.B. Kotlyar, V.D. Sled, A.D. Vinogradov, Ubisemiquinone
metabolism.
in the NADH-ubiquinone reductase region of the mitochondrial respiratory
Another aspect worth brief discussion relevant to the physiological chain, FEBS Lett. 254 (1989) 47–51.
significance of mitochondrial ROS production is its dependence on [10] M.L. Verkhovskaya, N. Belevich, L. Euro, M. Wikström, M.I. Verkhovsky, Real-time
oxygen concentration. In contrast to cytochrome oxidase, a reaction electron transfer in respiratory complex I, Proc. Natl. Acad. Sci. U. S. A. 105
(2008) 3763–3767.
which has an apparent Km for oxygen significantly lower than the [11] N. Belevich, G. Belevich, M. Verkhovskaya, Real-time optical studies of respiratory
physiologically conceivable concentration (about 5-fold less than that complex I turnover, Biochim. Biophys. Acta 1837 (2014) 1973–1980.
in air-saturated aqueous solutions at normal pressure [107]), the specif- [12] S. de Vries, K. Dörner, M.J. Strampraad, T. Friedrich, Electron tunneling rates in
respiratory complex I are tuned for efficient energy conversion, Angew. Chem.
ic activity of “major” respiratory chain-linked ROS generator, complex I Int. Ed. Engl. 54 (2015) 2844–2848.
[74,108,109], is simply proportional to oxygen concentration. Hyperbol- [13] A.B. Kotlyar, A.D. Vinogradov, Slow active/inactive transition of the mitochondrial
ic or other complex kinetics of complex I-mediated ROS production NADH-ubiquinone reductase, Biochim. Biophys. Acta 1019 (1990) 151–158.
[14] A.B. Kotlyar, S.P. Albracht, R.J. van Spanning, Comparison of energization of
might be expected if it does have physiological function. On the other complex I in membrane particles from Paracoccus denitrificans and bovine heart
hand, simple proportionality of ROS formation to oxygen concentration mitochondria, Biochim. Biophys. Acta 1365 (1998) 53–59.
might serve as an ideal oxygen sensing mechanism. [15] A.B. Kotlyar, N. Borovok, NADH oxidation and NAD+ reduction catalysed by tightly
coupled inside-out vesicles from Paracoccus denitrificans, Eur. J. Biochem. 269
The enzymes utilizing superoxide and hydrogen peroxide are (2002) 4020–4024.
frequently considered as an “antioxidant defense system”, whereas [16] K.R. Pryde, J. Hirst, Superoxide is produced by the reduced flavin in mitochondrial
generation of ROS is considered as an evolutionarily unfavorable complex I: a single, unified mechanism that applies during both forward and
reverse electron transfer, J. Biol. Chem. 286 (2011) 18056–18065.
leakage reaction (except for NAD(P)H oxidases, which perform a clearly
[17] V.G. Grivennikova, A.D. Vinogradov, Generation of superoxide by the mitochondri-
defined function). An alternative view is that hydrogen peroxide is a al complex I, Biochim. Biophys. Acta 1757 (2006) 553–561.
normal metabolic intermediate that fulfills some important not yet [18] L. Kussmaul, J. Hirst, The mechanism of superoxide production by NADH:ubiqui-
clear function(s). If a “leakage” hypothesis and “antioxidant defense none oxidoreductase (complex I) from bovine heart mitochondria, Proc. Natl.
Acad. Sci. U. S. A. 103 (2006) 7607–7612.
system” are to be accepted, many more or less experimentally justified [19] R. Fato, E. Estornell, S. Di Bernardo, F. Pallotti, G. Parenti Castelli, G. Lenaz, Steady-
efforts to improve the defense (antioxidant drugs, particularly those state kinetics of the reduction of coenzyme Q analogs by complex I (NADH:ubiqui-
mitochondrially targeted [110,111]) are certainly warranted. On the none oxidoreductase) in bovine heart mitochondria and submitochondrial
particles, Biochemistry 35 (1996) 2705–2716.
other hand, if the alternative view is correct, the use of so-called antiox- [20] V. Zickermann, S. Kurki, M. Kervinen, I. Hassinen, M. Finel, The NADH oxidation
idant drugs could result in unfavorable consequences. When discussing domain of complex I: do bacterial and mitochondrial enzymes catalyze ferricya-
physiological and pathophysiological “oxidative stress”, one should nide reduction similarly? Biochim. Biophys. Acta 1459 (2000) 61–68.
[21] V.D. Sled, A.D. Vinogradov, Kinetics of the mitochondrial NADH-ubiquinone
keep in mind metabolic effects of the “reductive stress” that can be oxidoreductase interaction with hexammineruthenium(III), Biochim. Biophys.
induced by the antioxidants [112]. Obvious expected consequences of Acta 1141 (1993) 262–268.
“overreduction” are: 1) abnormal anabolic activity including malignant [22] E.V. Gavrikova, V.G. Grivennikova, V.D. Sled, T. Ohnishi, A.D. Vinogradov, Kinetics of
the mitochondrial three-subunit NADH dehydrogenase interaction with
growth; 2) an increase in fatty acid and fat biosynthesis; 3) disorder hexammineruthenium(III), Biochim. Biophys. Acta 1230 (1995) 23–30.
of amino acid transport systems (participation of glutathione in the [23] J.G. Okun, V. Zickermann, K. Zwicker, H. Schägger, U. Brandt, Binding of detergents
γ-glutamate cycle [113]); 4) a decrease of normal insulin level due and inhibitors to bovine complex I - a novel purification procedure for bovine
complex I retaining full inhibitor sensitivity, Biochim. Biophys. Acta 1459 (2000)
to an increase in the insulin:glutathione transhydrogenase reaction
77–87.
[114,115]. Last, not the least consequence of reductive stress is a possi- [24] N.V. Zakharova, T.V. Zharova, A.D. Vinogradov, Kinetics of transhydrogenase
ble paradoxical activation of ROS production. reaction catalyzed by the mitochondrial NADH-ubiquinone oxidoreductase
(complex I) imply more than one catalytic nucleotide-binding sites, FEBS Lett.
444 (1999) 211–216.
Conflict of Interest [25] A. Galkin, U. Brandt, Superoxide radical formation by pure complex I (NADH:
ubiquinone oxidoreductase) from Yarrowia lipolytica, J. Biol. Chem. 280 (2005)
We declare no conflict of interest. 30129–30135.
[26] N. Kashani-Poor, S. Kerscher, V. Zickermann, U. Brandt, Efficient large scale
purification of his-tagged proton translocating NADH:ubiquinone oxidoreductase
Acknowledgements (complex I) from the strictly aerobic yeast Yarrowia lipolytica, Biochim. Biophys.
Acta 1504 (2001) 363–370.
[27] V.G. Grivennikova, D.V. Serebryanaya, E.P. Isakova, T.A. Belozerskaya, A.D.
We are indebted to Dr. J. Hirst for providing us data (Ref. [50]) before Vinogradov, The transition between active and de-activated forms of NADH:
publication. We are grateful to anonymous Reviewer for his (her) kind ubiquinone oxidoreductase (complex I) in the mitochondrial membrane of
linguistic corrections of our manuscript. The experimental works done Neurospora crassa, Biochem. J. 369 (2003) 619–626.
[28] A.V. Ushakova, M. Duarte, A.D. Vinogradov, A. Videira, The 29.9 kDa subunit of
in our laboratory were supported by Russian Foundation for Fundamen- mitochondrial complex I is involved in the enzyme active/de-active transitions, J.
tal Research Grant 14-04-00279 to A.D.V. and Grant 15-04-02957 to Mol. Biol. 351 (2005) 327–333.
V.G.G. [29] K. Matsushita, T. Ohnishi, H.R. Kaback, NADH-ubiquinone oxidoreductases of the
Escherichia coli aerobic respiratory chain, Biochemistry 26 (1987) 7732–7777.
[30] T. Pohl, M. Uhlmann, M. Kaufenstein, T. Friedrich, Lambda Red-mediated mutagen-
References esis and efficient large scale affinity purification of the Escherichia coli NADH:
ubiquinone oxidoreductase (complex I), Biochemistry 46 (2007) 10694–10702.
[1] L.A. Sazanov, P. Hinchliffe, Structure of the hydrophilic domain of respiratory [31] L.A. Sazanov, J. Carroll, P. Holt, L. Toime, I.M. Fearnley, A role for native lipids in
complex I from Thermus thermophilus, Science 311 (2006) 1430–1436. the stabilization and two-dimensional crystallization of the Escherichia coli
[2] R.G. Efremov, R. Baradaran, L.A. Sazanov, The architecture of respiratory complex I, NADH-ubiquinone oxidoreductase (complex I), J. Biol. Chem. 278 (2003)
Nature 465 (2010) 441–445. 19483–19491.
[3] C. Hunte, V. Zickermann, U. Brandt, Functional modules and structural basis of con- [32] L. Sinegina, M. Wikström, M.I. Verkhovsky, M.L. Verkhovskaya, Activation of
formational coupling in mitochondrial complex I, Science 329 (2010) 448–451. isolated NADH:ubiquinone reductase I (complex I) from Escherichia coli by
870 A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871

detergent and phospholipids. Recovery of ubiquinone reductase activity and [63] S. Dröse, S. Krack, L. Sokolova, K. Zwicker, H.D. Barth, N. Morgner, H. Heide, et al.,
changes in EPR signals of iron-sulfur clusters, Biochemistry 44 (2005) Functional dissection of the proton pumping modules of mitochondrial complex
8500–8506. I, PLoS Biol. 9 (2011), e1001128.
[33] C.Y. Yip, M.E. Harbour, K. Jayawardena, I.M. Fearnley, L.A. Sazanov, Evolution of [64] A.S. Galkin, V.G. Grivennikova, A.D. Vinogradov, →H+/2ē stoichiometry of the
respiratory complex I: "supernumerary" subunits are present in the alpha- NADH:ubiquinone reductase reaction catalyzed by submitochondrial particles,
proteobacterial enzyme, J. Biol. Chem. 286 (2011) 5023–5033. Biochemistry (Mosc) 66 (2001) 435–443.
[34] V.G. Grivennikova, R. Roth, N.V. Zakharova, C. Hägerhäll, A.D. Vinogradov, The [65] Y. Hatefi, A.G. Haavik, D.E. Griffiths, Studies on the electron transfer system. XL.
mitochondrial and prokaryotic proton-translocating NADH:ubiquinone oxidore- Preparation and properties of mitochondrial DPNH-coenzyme Q reductase, J.
ductases: similarities and dissimilarities of the quinone-junction sites, Biochim. Biol. Chem. 237 (1962) 1676–1680.
Biophys. Acta 1607 (2003) 79–90. [66] M.S. Sharpley, R.J. Shannon, F. Draghi, J. Hirst, Interactions between phospholipids
[35] A.D. Vinogradov, V.G. Grivennikova, The mitochondrial complex I: progress in and NADH:ubiquinone oxidoreductase (complex I) from bovine mitochondria,
understanding of catalytic properties, IUBMB Life 52 (2001) 129–134. Biochemistry 45 (2006) 241–248.
[36] M. Babot, A. Birch, P. Labarbuta, A. Galkin, Characterisation of the active/de-active [67] K. Shinzawa-Itoh, J. Seiyama, H. Terada, R. Nakatsubo, K. Naoki, Y. Nakashima, S.
transition of mitochondrial complex I, Biochim. Biophys. Acta 1837 (2014) Yoshikawa, Bovine heart NADH-ubiquinone oxidoreductase contains one molecule
1083–1092. of ubiquinone with ten isoprene units as one of the cofactors, Biochemistry 49
[37] A.D. Vinogradov, T.E. King, The Keilin-Hartree heart muscle preparation, Methods (2010) 487–492.
Enzymol. 55 (1979) 118–127. [68] H.R. Bridges, L. Grgic, M.E. Harbour, J. Hirst, The respiratory complexes I from the
[38] R. Fato, M. Battino, M. Degli Esposti, G. Parenti Castelli, G. Lenaz, Determination of mitochondria of two Pichia species, Biochem. J. 422 (2009) 151–159.
partition and lateral diffusion coefficients of ubiquinones by fluorescence [69] P. Hinchliffe, J. Carroll, L.A. Sazanov, Identification of a novel subunit of respiratory
quenching of n-(9-anthroyloxy)stearic acids in phospholipid vesicles and complex I from Thermus thermophilus, Biochemistry 45 (2006) 4413–4420.
mitochondrial membranes, Biochemistry 25 (1986) 3378–3390. [70] G. Peng, G. Fritzsch, V. Zickermann, H. Schägger, R. Mentele, F. Lottspeich, M.
[39] A. Kröger, M. Klingenberg, The kinetics of the redox reactions of ubiquinone related Bostina, et al., Isolation, characterization and electron microscopic single particle
to the electron-transport activity in the respiratory chain, Eur. J. Biochem. 34 analysis of the NADH:ubiquinone oxidoreductase (complex I) from the hyperther-
(1973) 358–368. mophilic eubacterium Aquifex aeolicus, Biochemistry 42 (2003) 3032–3039.
[40] J.N. Blaza, R. Serreli, A.J. Jones, K. Mohammed, J. Hirst, Kinetic evidence against [71] A.S. Fernandes, M.M. Pereira, M. Teixeira, Purification and characterization of the
partitioning of the ubiquinone pool and the catalytic relevance of respiratory- complex I from the respiratory chain of Rhodothermus marinus, J. Bioenerg.
chain supercomplexes, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) 15735–15740. Biomembr. 34 (2002) 413–421.
[41] Y. Nakashima, K. Shinzawa-Itoh, K. Watanabe, K. Naoki, N. Hano, S. Yoshikawa, [72] C.I. Ragan, The role of phospholipids in the reduction of ubiquinone analogues
Steady-state kinetics of NADH:coenzyme Q oxidoreductase isolated from bovine by the mitochondrial reduced nicotinamide-adenine dinucleotide-ubiquinone
heart mitochondria, J. Bioenerg. Biomembr. 34 (2002) 11–19. oxidoreductase complex, Biochem. J. 172 (1978) 539–547.
[42] V.G. Grivennikova, A.D. Vinogradov, Kinetics of ubiquinone reduction by the [73] S. Dröse, K. Zwicker, U. Brandt, Full recovery of the NADH:ubiquinone activity of
resolved succinate: ubiquinone reductase, Biochim. Biophys. Acta 682 (1982) complex I (NADH:ubiquinone oxidoreductase) from Yarrowia lipolytica by the
491–495. addition of phospholipids, Biochim. Biophys. Acta 1556 (2002) 65–72.
[43] S. Minakami, R.L. Ringler, T.P. Singer, Studies on the respiratory chain-linked [74] A.D. Vinogradov, V.G. Grivennikova, Generation of superoxide-radical by the
dihydrodiphosphopyridine nucleotide dehydrogenase. I. Assay of the enzyme in NADH:ubiquinone oxidoreductase of heart mitochondria, Biochemistry (Mosc)
particulate and in soluble preparations, J. Biol. Chem. 237 (1962) 569–576. 70 (2005) 120–127.
[44] G. Dooijewaard, E.C. Slater, Steady-state kinetics of high molecular weight (type-I) [75] V.G. Grivennikova, A.D. Vinogradov, Partitioning of superoxide and hydrogen
NADH dehydrogenase, Biochim. Biophys. Acta 440 (1976) 1–15. peroxide production by mitochondrial respiratory complex I, Biochim. Biophys.
[45] J.A. Birrell, M.S. King, J. Hirst, A ternary mechanism for NADH oxidation by positive- Acta 1827 (2013) 446–454.
ly charged electron acceptors, catalyzed at the flavin site in respiratory complex I, [76] D. Esterházy, M.S. King, G. Yakovlev, J. Hirst, Production of reactive oxygen species
FEBS Lett. 585 (2011) 2318–2322. by complex I (NADH:ubiquinone oxidoreductase) from Escherichia coli and com-
[46] V.G. Grivennikova, G.V. Gladyshev, A.D. Vinogradov, Allosteric nucleotide-binding parison to the enzyme from mitochondria, Biochemistry 47 (2008) 3964–3971.
site in the mitochondrial NADH:ubiquinone oxidoreductase (respiratory complex [77] T.Y. Pepelina, R.V. Chertkova, T.V. Ostroverkhova, D.A. Dolgikh, M.P. Kirpichnikov,
I), FEBS Lett. 585 (2011) 2212–2216. V.G. Grivennikova, A.D. Vinogradov, Site-directed mutagenesis of cytochrome c:
[47] T.V. Zharova, A.D. Vinogradov, A competitive inhibition of the mitochondrial reactions with respiratory chain components and superoxide radical, Biochemistry
NADH-ubiquinone oxidoreductase (complex I) by ADP-ribose, Biochim. Biophys. (Mosc) 74 (2009) 625–632.
Acta 1320 (1997) 256–264. [78] M.L. Genova, B. Ventura, G. Giuliano, C. Bovina, G. Formiggini, G. Parenti Castelli, G.
[48] A.D. Vinogradov, NADH/NAD+ interaction with NADH: ubiquinone oxidoreductase Lenaz, The site of production of superoxide radical in mitochondrial complex I is
(complex I), Biochim. Biophys. Acta 1777 (2008) 729–734. not a bound ubisemiquinone but presumably iron-sulfur cluster N2, FEBS Lett.
[49] K. Kmita, V. Zickermann, Accessory subunits of mitochondrial complex I, Biochem. 505 (2001) 364–368.
Soc. Trans. 41 (2013) 1272–1279. [79] L. Grgic, K. Zwicker, N. Kashani-Poor, S. Kerscher, U. Brandt, Functional significance
[50] F. Varghese, E. Atcheson, H.R. Bridges, J. Hirst, Characterization of clinically identi- of conserved histidines and arginines in the 49-kDa subunit of mitochondrial
fied mutations in NDUFV1, the flavin-binding subunit of respiratory complex I, complex I, J. Biol. Chem. 279 (2004) 21193–21199.
using a yeast model system, Hum. Mol. Genet. 1-11 (2015). [80] K. Zwicker, A. Galkin, S. Dröse, L. Grgic, S. Kerscher, U. Brandt, The redox-Bohr
[51] V. Zickermann, M. Bostina, C. Hunte, T. Ruiz, M. Radermacher, U. Brandt, Functional group associated with iron-sulfur cluster N2 of complex I, J. Biol. Chem. 281
implications from an unexpected position of the 49-kDa subunit of NADH: (2006) 23013–23017.
ubiquinone oxidoreductase, J. Biol. Chem. 278 (2003) 29072–29078. [81] Y. Kushnareva, A.N. Murphy, A. Andreyev, Complex I-mediated reactive oxygen
[52] R. Baradaran, J.M. Berrisford, G.S. Minhas, L.A. Sazanov, Crystal structure of the species generation: modulation by cytochrome c and NAD(P)+ oxidation-
entire respiratory complex I, Nature 494 (2013) 443–448. reduction state, Biochem. J. 368 (2002) 545–553.
[53] C. Mathiesen, C. Hägerhäll, Transmembrane topology of the NuoL, M and N [82] A.P. Kudin, N.Y. Bimpong-Buta, S. Vielhaber, C.E. Elger, W.S. Kunz, Characterization
subunits of NADH:quinone oxidoreductase and their homologues among of superoxide-producing sites in isolated brain mitochondria, J. Biol. Chem. 279
membrane-bound hydrogenases and bona fide antiporters, Biochim. Biophys. (2004) 4127–4135.
Acta 1556 (2002) 121–132. [83] A.V. Kareyeva, V.G. Grivennikova, A.D. Vinogradov, Mitochondrial hydrogen perox-
[54] A.S. Galkin, V.G. Grivennikova, A.D. Vinogradov, →H+/2ē stoichiometry in NADH- ide production as determined by the pyridine nucleotide pool and its redox state,
quinone reductase reactions catalyzed by bovine heart submitochondrial particles, Biochim. Biophys. Acta 1817 (2012) 1879–1885.
FEBS Lett. 451 (1999) 157–161. [84] V.G. Grivennikova, A.B. Kotlyar, J.S. Karliner, G. Cecchini, A.D. Vinogradov, Redox-
[55] M. Wikström, G. Hummer, Stoichiometry of proton translocation by respiratory dependent change of nucleotide affinity to the active site of the mammalian
complex I and its mechanistic implications, Proc. Natl. Acad. Sci. U. S. A. 109 complex I, Biochemistry 46 (2007) 10971–10978.
(2012) 4431–4436. [85] V.D. Sled, N.I. Rudnitzky, Y. Hatefi, T. Ohnishi, Thermodynamic analysis of flavin in
[56] S.W. Meinhardt, D.C. Wang, K. Hon-nami, T. Yagi, T. Oshima, T. Ohnishi, Studies on mitochondrial NADH:ubiquinone oxidoreductase (complex I), Biochemistry 33
the NADH-menaquinone oxidoreductase segment of the respiratory chain in (1994) 10069–10075.
Thermus thermophilus HB-8, J. Biol. Chem. 265 (1990) 1360–1368. [86] J. Knuuti, G. Belevich, V. Sharma, D.A. Bloch, M. Verkhovskaya, A single amino acid
[57] A.V. Bogachev, R.A. Murtazina, V.P. Skulachev, H+/ē stoichiometry for NADH residue controls ROS production in the respiratory complex I from Escherichia coli,
dehydrogenase I and dimethyl sulfoxide reductase in anaerobically grown Mol. Microbiol. 90 (2013) 1190–1200.
Escherichia coli cells, J. Bacteriol. 178 (1996) 6233–6237. [87] Y. Hatefi, K.E. Stempel, W.G. Hanstein, Inhibitors and activators of the mitochondri-
[58] V.G. Grivennikova, A.V. Ushakova, C. Hagerhall, A.D. Vinogradov, Proton translocation al reduced diphosphopyridine nucleotide dehydrogenase, J. Biol. Chem. 244 (1969)
catalyzed by Paracoccus denitrificans NADH:quinone oxidoreductase (NDH-1), 2358–2365.
Biochim. Biophys. Acta EBEC Short Reports 12 (2002) 208. [88] A.D. Vinogradov, Catalytic properties of the mitochondrial NADH-ubiquinone
[59] P. Scholes, P. Mitchell, Respiration-driven proton translocation in Micrococcus oxidoreductase (complex I) and the pseudo-reversible active/inactive enzyme
denitrificans, J. Bioenerg. 1 (1971) 309–323. transition, Biochim. Biophys. Acta 1364 (1998) 169–185.
[60] S.B. Long, E.B. Campbell, R. Mackinnon, Crystal structure of a mammalian voltage- [89] E.O. Maklashina, V.D. Sled, A.D. Vinogradov, Hysteresis behavior of complex I
dependent Shaker family K+ channel, Science 309 (2005) 897–903. from bovine heart mitochondria: kinetic and thermodynamic parameters of retarded
[61] R. Mackinnon, Potassium channels and the atomic basis of selective ion conduction reverse transition from the inactive to active state, Biokhimiia 59 (1994) 707–714.
(Nobel lecture), Angew. Chem. Int. Ed. 43 (2004) 4265–4277. [90] V.G. Grivennikova, A.N. Kapustin, A.D. Vinogradov, Catalytic activity of NADH-
[62] W.A. Catterall, From ionic currents to molecular mechanisms: the structure and ubiquinone oxidoreductase (complex I) in intact mitochondria. Evidence for the
function of voltage-gated sodium channels, Neuron 26 (2000) 13–25. slow active/inactive transition, J. Biol. Chem. 276 (2001) 9038–9044.
A.D. Vinogradov, V.G. Grivennikova / Biochimica et Biophysica Acta 1857 (2016) 863–871 871

[91] E.O. Maklashina, Y. Sher, H.-Z. Zhou, M.O. Gray, J.S. Karliner, G. Cecchini, Effect of [102] J. Eng, R.M. Lynch, R.S. Balaban, Nicotinamide adenine dinucleotide fluorescence
anoxia/reperfusion on the reversible active/de-active transition of NADH- spectroscopy and imaging of isolated cardiac myocytes, Biophys. J. 55 (1989)
ubiquinone oxidoreductase (complex I) in rat heart, Biochim. Biophys. Acta 1556 621–630.
(2002) 6–12. [103] I.E. Hassinen, Mitochondrial respiratory control in the myocardium, Biochim.
[92] M.V. Loskovich, V.G. Grivennikova, G. Cecchini, A.D. Vinogradov, Inhibitory effect of Biophys. Acta 853 (1986) 135–151.
palmitate on the mitochondrial NADH:ubiquinone oxidoreductase (complex I) as [104] A. Boveris, N. Oshino, B. Chance, The cellular production of hydrogen peroxide,
related to the active-de-active enzyme transition, Biochem. J. 387 (2005) 677–683. Biochem. J. 128 (1972) 617–630.
[93] A.B. Kotlyar, V.D. Sled, A.D. Vinogradov, Effect of Ca2+ ions on the slow active/inac- [105] D. Harman, The biologic clock: the mitochondria? J. Am. Geriatr. Soc. 20 (1972)
tive transition of the mitochondrial NADH-ubiquinone reductase, Biochim. 145–147.
Biophys. Acta 1098 (1992) 144–150. [106] E.A. Veal, A.M. Day, B.A. Morgan, Hydrogen peroxide sensing and signaling, Mol.
[94] D.S. Kalashnikov, V.G. Grivennikova, A.D. Vinogradov, Synergetic inhibition of the Cell 26 (2007) 1–14.
brain mitochondrial NADH:ubiquinone oxidoreductase (complex I) by fatty acids [107] D.F. Wilson, Quantifying the role of oxygen pressure in tissue function, Am. J.
and Ca2+, Biochemistry (Mosc) 76 (2011) 968–975. Physiol. Heart Circ. Physiol. 294 (2008) H11–H13.
[95] C.J. Schofield, Z. Zhang, Structural and mechanistic studies on 2-oxoglutarate- [108] A. Boveris, B. Chance, The mitochondrial generation of hydrogen peroxide. General
dependent oxygenases and related enzymes, Curr. Opin. Struct. Biol. 9 (1999) properties and effect of hyperbaric oxygen, Biochem. J. 134 (1973) 707–716.
722–731. [109] J.F. Turrens, B.A. Freeman, J.G. Levitt, J.D. Crapo, The effect of hyperoxia on superox-
[96] E. Nguyen, M.J.S. Picklo, Inhibition of succinic semialdehyde dehydrogenase ide production by lung submitochondrial particles, Arch. Biochem. Biophys. 217
activity by alkenal products of lipid peroxidation, Biochim. Biophys. Acta 1637 (1982) 401–410.
(2003) 107–112. [110] V.P. Skulachev, SkQ1 treatment and food restriction – two ways to retard an aging
[97] E.T. Chouchani, V.R. Pell, E. Gaude, D. Aksentijevic, S.Y. Sundier, E.L. Robb, A.J. Dare, program of organisms, Aging 3 (2011) 1045–1050.
et al., Ischaemic accumulation of succinate controls reperfusion injury through [111] M.P. Murphy, Antioxidants as therapies: can we improve on nature? Free Radic.
mitochondrial ROS, Nature 515 (2014) 431–435. Biol. Med. 66 (2014) 20–23.
[98] K.B. Jacobson, N.O. Kaplan, Pyridine coenzymes of subcellular tissue fractions, J. [112] H. Zhang, P. Limphong, J. Pieper, Q. Liu, C.K. Rodesch, E. Christians, I.J. Benjamin,
Biol. Chem. 226 (1957) 603–613. Glutathione-dependent reductive stress triggers mitochondrial oxidation and
[99] R.L. Lester, Y. Hatefi, Studies on the mechanism of oxidative phosphorylation. IV. cytotoxicity, FASEB J. 26 (2012) 1442–1451.
Pyridine nucleotide binding and its relation to activity in heart mitochondria, [113] O.W. Griffith, R.J. Bridges, A. Meister, Transport of γ-glutamyl amino acids: role of
Biochim. Biophys. Acta 29 (1958) 103–112. glutathione and γ-glutamyl transpeptidase, Proc. Natl. Acad. Sci. U. S. A. 76 (1979)
[100] M. Klingenberg, W. Slenczka, E. Ritt, Vergleichende biochemie der 6319–6322.
pyridinnucleotid-systeme in mitochondrien verschiedener organe, Biochem. Z. [114] M.L. Chandler, P.T. Varandani, Kinetic analysis of the mechanism of insulin
332 (1959) 47–66. degradation by glutathione-insulin transhydrogenase (thiol: protein-disulfide
[101] D.H. Williamson, P. Lund, H.A. Krebs, The redox state of free nicotinamide-adenine oxidoreductase), Biochemistry 14 (1975) 2107–2115.
dinucleotide in the cytoplasm and mitochondria of rat liver, Biochem. J. 103 (1967) [115] C.M. Cordes, R.G. Bennett, G.L. Siford, F.G. Hamel, Redox regulation of insulin
514–527. degradation by insulin-degrading enzyme, PLoS One 6 (2011), e18138.

You might also like