SlideShare a Scribd company logo
1 of 11
Download to read offline
A single amino acid residue controls ROS production in the
respiratory Complex I from Escherichia coli
Juho Knuuti,1
Galina Belevich,1
Vivek Sharma,2
Dmitry A. Bloch1
* and Marina Verkhovskaya1
**
1
Helsinki Bioenergetics Group, Institute of
Biotechnology, University of Helsinki, PO Box 65
(Viikinkaari 1), FIN-00014, Helsinki, Finland.
2
Department of Physics, Tampere University of
Technology, P. O. Box 692, FIN-33101, Tampere,
Finland.
Summary
Reactive oxygen species (ROS) production by respira-
tory Complex I from Escherichia coli was studied in
bacterial membrane fragments and in the isolated and
purified enzyme, either solubilized or incorporated in
proteoliposomes. We found that the replacement of a
single amino acid residue in close proximity to the
nicotinamide adenine dinucleotide (NADH)-binding
catalytic site (E95 in the NuoF subunit) dramatically
increases the reactivity of Complex I towards dioxygen
(O2). In the E95Q variant short-chain ubiquinones
exhibit strong artificial one-electron reduction at the
catalytic site, also leading to a stronger increase in
ROS production. Two mechanisms can contribute to
the observed kinetic effects: (a) a change in the reac-
tivity of flavin mononucleotide (FMN) towards di-
oxygen at the catalytic site, and (b) a change in the
population of the ROS-generating state. We propose
the existence of two (closed and open) states of the
NAD+
-bound enzyme as one feature of the substrate-
binding site of Complex I. The analysis of the kinetic
model of ROS production allowed us to propose that
the population of Complex I with reduced FMN is
always low in the wild-type enzyme even at low
ambient redox potentials, minimizing the rate of reac-
tion with O2 in contrast to E95Q variant.
Introduction
Complex I [nicotinamide adenine dinucleotide (NADH):qui-
none oxidoreductase, type I] catalyses electron input to the
respiratory chain. It uses NADH as a two-electron donor
and transfers these electrons to quinone (most often ubi-
quinone) via flavin mono-nucleotide (FMN) and eight iron-
sulphur (FeS) clusters located in the hydrophilic domain of
the enzyme. Quinone reduction is coupled to translocation
of three to four protons across the inner mitochondrial
or bacterial cytoplasmic membrane (Wikström, 1984;
Hinkle et al., 1991; Galkin et al., 1999; 2006; Wikström and
Hummer, 2012). Complex I has also been suggested to be
one of the major sites of formation of superoxide by mito-
chondria (Turrens and Boveris, 1980; Galkin and Brandt,
2005; Kussmaul and Hirst, 2006), and has been therefore
linked to some neurodegenerative diseases and ageing
(Lin and Beal, 2006; Fukui and Moraes, 2008; Rhein et al.,
2009). Indeed, it was reported that Complex I is the major
site of superoxide formation in brain mitochondria under
normal conditions (Barja and Herrero, 1998), and that it
produces most of the reactive oxygen species (ROS) in
Parkinson’s disease and ageing (for a review see Turrens,
2003), although the relevance of superoxide formation
by Complex I under physiological conditions has been
questioned (Zoccarato et al., 2004; Grivennikova and
Vinogradov, 2006; Drechsel and Patel, 2010; Kareyeva
et al., 2012). The site of this ‘electron leak’ from Complex I
to O2 has been studied in different systems, ranging from
intact mitochondria (Treberg et al., 2011) to the isolated
enzyme (Kussmaul and Hirst, 2006; Ohnishi et al., 2010),
there are no definite conclusions as yet. Here, we have
studied ROS production by isolated and native membrane-
bound Complex I from Escherichia coli, with the aid of
site-specific mutations to alter vital electron transfer prop-
erties. Complex I contains non-covalently bound FMN,
which resides on the bottom of a cavity of the catalytic
subunit and is, therefore, accessible to oxygen from the
surroundings. FMN has a low midpoint redox potential
(Sled et al., 1993; Euro et al., 2008b), and has been pro-
posed to be the main site of O2
•−
generation in mitochon-
drial Complex I (Galkin and Brandt, 2005; Grivennikova
and Vinogradov, 2006; Kussmaul and Hirst, 2006).
However, the rates of ROS production by purified mito-
chondrial [16–22 nmol mg−1
min−1
(Galkin and Brandt,
2005; Kussmaul and Hirst, 2006; Kareyeva et al., 2012)],
and bacterial [25–50 nmol mg−1
min−1
(Esterhazy et al.,
2008; this article)] Complex I, are very slow compared
to the NADH oxidation rates. Furthermore, some other
NADH-dependent flavin-containing enzymes, such as
Accepted 3 October, 2013. For correspondence. *E-mail dmitry
.bloch@helsinki.fi; Tel. (+358) 9 191 59754; Fax (+358) 9 191 59920;
**E-mail marina.verkhovskaya@helsinki.fi; Tel. (+358) 9 191 59748;
Fax (+358) 9 191 59920.
Molecular Microbiology (2013) 90(6), 1190–1200 ■ doi:10.1111/mmi.12424
First published online 29 October 2013
© 2013 John Wiley & Sons Ltd
NADH/NADPH oxidases belonging to the Nox family (e.g.
MJ0649 from Methanocaldococcus jannaschii) produce
superoxide at much higher rates (60 μmol min−1
mg−1
),
which is c. 60% of the NADH oxidation rate (Case et al.,
2009). ROS production by mitochondria respiring on sub-
strates that reduce matrix NAD+
is also moderate (for a
review see Brand, 2010). This rate can be significantly
increased by blocking the electron output with a specific
Complex I inhibitor (e.g. Turrens and Boveris, 1980), but
such treatment raises the NADH/NAD+
ratio making it
possible that other NADH-dependent enzymes, such as
pyruvate or α-ketoglutarate dehydrogenase, could partici-
pate (Starkov et al., 2004). Under certain conditions mito-
chondrial dihydrolipoamide dehydrogenase was found to
be the major source of NADH-dependent ROS production
(Kareyeva et al., 2012).
In submitochondrial particles NADH-dependent O2
•−
generation was found to be only slightly increased
by rotenone (Vinogradov and Grivennikova, 2005;
Grivennikova and Vinogradov, 2006), and the isolated
enzyme showed no response (Galkin and Brandt, 2005;
Kussmaul and Hirst, 2006; Drose et al., 2009). The first
indication that reversed electron flow in mitochondria could
be a source of ROS was obtained by Hinkle et al. in 1967
(Hinkle et al., 1967). Later, it was reported that succinate or
glycerol-6-phosphate gives the highest rates of Complex
I-derived ROS production in mitochondria under condi-
tions of reversed electron transfer (see Brand, 2010 for
a review). This phenomenon has been explained by
the presence of two different O2
•−
-generating nucleotide
binding sites in Complex I, only one of which is active upon
forward electron transport (Vinogradov and Grivennikova,
2005), but this was not supported by the resolved structure
of Complex I (Sazanov and Hinchliffe, 2006). O2
•−
genera-
tion in the quinone binding site of Complex I was proposed
independently by Brand and Treberg (Lambert and Brand,
2004a,b; Treberg et al., 2011), but this suggestion was
disputed by Pryde and Hirst (Pryde and Hirst, 2011), who
concluded from experiments with submitochondrial parti-
cles that ROS production by Complex I is due to oxidation
of FMN by O2 during both forward and reversed electron
flow.
At any rate, it is well established that the rates of gen-
eration of O2
•−
by Complex I in mitochondria, submito-
chondrial particles, and purified Complex I are low relative
to the much faster rates of NADH oxidation by quinone. It
has been suggested that retaining NAD+
in the catalytic
site could diminish the accessibility of oxygen to FMN
(Kussmaul and Hirst, 2006), thus perhaps contributing to
the slow rates of ROS production. The resolved structure
of Complex I with the nucleotide bound [PBD entry 3IAS
(Berrisford and Sazanov, 2009)] indeed suggests that
NAD+
covers almost completely the part of the FMN
alloxazine ring that protrudes into the cavity. However, it is
difficult to predict to what extent such an effect might
prevent oxygen to react with reduced FMN.
The rate of ROS production by Complex I is expected to
be determined by the properties of the nucleotide binding
site and the bound FMN. The current study where a con-
served glutamate residue in the NuoF subunit (E95) is
mutated reveals a key element of the mechanism that
prevents massive electron leakage from the catalytic site
to dioxygen.
Results
Previously we have found that the replacement of the
invariant Glu95 with glutamine in the NADH- and FMN-
binding NuoF subunit results in moderate changes in
nucleotide affinity, but a significant upshift in the midpoint
redox potential of FMN (Euro et al., 2009a). The latter was
explained as an electrostatic effect on FMN due to replac-
ing the negatively charged carboxylate by the neutral
amide, since the side-chain of Glu95 resides in the nucleo-
tide binding cavity at a short distance from FMN. The
structure of the hydrophilic domain of Thermus thermophi-
lus Complex I (PDB entry 3IAM) reduced with NADH shows
that the latter is bound between E97, a counterpart of E95
in E. coli enzyme, and FMN. The distance between E97
and the nicotinamide ring of bound NADH is 3.2 Å. On the
other side E97 is located at the border of two subunits,
Nqo1 (NuoF) and Nqo2 (NuoE). The latter contains the
2Fe-2S cluster N1a, which lays 10 Å from E97 with Pro98
in between (Fig. 1).
An unexpected property of the E95Q variant of Complex
I is a strong increase of the rate of NADH-dependent
hydrogen peroxide production (Table 1). We compared
ROS production of wild-type enzyme, the E95Q variant,
and two other variants that lack quinone reductase activity.
In NuoCD R274A the electron transfer to ubiquinone was
blocked since the replacement of R274 by the neutral
alanine residue resulted in a strong decrease of the mid-
point redox potential of the last cluster in the intraprotein
FeS chain, N2 (Belevich et al., 2007), to such an extent that
it cannot be reduced by NADH. This cluster is the imme-
diate electron donor to ubiquinone. In the other variant,
NuoM E144A, the mutation was in the membrane subunit,
NuoM, quite distant from the quinone binding site (Euro
et al., 2008a). It is not yet known whether ubiquinone could
be reduced in this mutant, but the properties of the N2
cluster were unchanged. The rates of ROS production in
wild type and the NuoM E144A variant were rather similar,
and slightly lower in NuoCD R274A, in stark contrast to the
very high reactivity of the NuoF E95Q variant (Table 1). The
small decrease of ROS production rate in NuoCD R274A
can be explained by the lower content of Complex I in the
sample: normalization by hexaammineruthenium (III)
chloride (HAR)-reductase activity yields the same ratio of
ROS production by Complex I 1191
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
NADH-dependent HAR to oxygen reduction rate as for wild
type. This finding excludes any significant role of the N2
centre, or ubiquinone bound near that centre, in NADH-
dependent H2O2 generation by purified Complex I, since
both redox centres are reduced to a much smaller extent in
NuoCD R274A compared to wild type, yet with an insignifi-
cant change of ROS production velocity. No significant
difference in the rate of ROS production was observed
between solubilized Complex I and enzyme reconstituted
into liposomes (not shown).
ROS production by Complex I can be stimulated by
short-chain ubiquinones, insignificantly by decylubiqui-
none (DQ) and strongly by coenzyme Q1 (Q1) and mena-
dione (Table 2). This process is probably mediated by
formation of the respective semiquinone which interacts
with oxygen. It is known that short-chain ubiquinone ana-
logues, such as Q1, can be reduced by Complex I from an
artificial site resulting in O2
•−
generation (Cadenas et al.,
1977; Galkin and Brandt, 2005); the efficiency of this
process is increased with decrease of the quinone hydro-
phobicity (Cadenas et al., 1977). Most probably the site of
this artificial reaction is FMN (King et al., 2009). Menadi-
one is known to interact with flavin in many flavin-
containing enzymes that yields intensive ROS formation
(e.g. Matsuda et al., 2000). All in all, the data shown in
Table 2 indicates that FMN is the only significant site of
ROS production by Complex I under these conditions.
The dependence of the rate of ROS production on the
ambient redox potential calculated from the ratio [NADH]/
[NAD+
] is shown in Fig. 2. The experimental curve, which
is well reproducible and similar for wild type and the
NuoCD R274A and NuoM E144A variants, is less steep
and positively shifted (Fig. 2A) from a Nernstian n = 2
curve with Em = −350 mV, characteristic for FMN of E. coli
Complex I (Euro et al., 2008b). These data are in very
good agreement with that obtained on E. coli Complex I
by Esterhazy et al. (2008). In the mitochondrial enzyme
the rate of O2
•−
formation is half-maximal at −359 mV
(Kussmaul and Hirst, 2006). The titration of the rate of
ROS formation by the NuoF E95Q variant was steeper
Fig. 1. Disposition of E97 (counterpart of
E95 in E. coli Complex I) in the catalytic site
of Complex I from T. thermophilus (PDB
entry: 3IAM).
Table 1. ROS production by isolated Complex I does not correlate
with its NADH oxidase or ubiquinone reductase activities.
Variant
Complex I activity, μmol mg−1
min−1 Rate of NADH-
dependent
ROS production,
nmol mg−1
min−1
NADH:HAR
oxidoreductase
NADH:DQ
oxidoreductase
Wt 107.4 ± 7.0 26.9 ± 1.8 51.7 ± 3.5
NuoM E144A 109.7 ± 8.5 3.6 ± 0.5 55.6 ± 7.5
NuoCD R274A 72.4 ± 2.3 5.7 ± 0.3 44.1 ± 0.4
NuoF E95Q 24.4 ± 1.5 11.5 ± 1.5 749.4 ± 1.7
1192 J. Knuuti et al. ■
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
than that in the WT enzyme, and showed a clear positive
shift (Fig. 2B), which is in good agreement with the previ-
ously determined midpoint potentials of the two-electron
transition of FMN, −350 mV for wild type and −310 mV in
this mutant (Euro et al., 2009a). Assuming that both the
reduced and the half-reduced forms of FMN interact with
oxygen, it may be suggested that the deviation from a
two-electron Nernstian curve in the wild-type enzyme is
due to a somewhat higher semiflavin radical stability than
in the E95Q variant. Such an effect could explain the
change of shape of the titration curves, but not the strong
increase in the rate of ROS generation by the E95Q
variant.
The ubiquinone reductase activity of wild-type enzyme
is inhibited by rolliniastatin, the specific Complex I inhibitor
affecting the quinone binding site, by 90%, regardless of
whether DQ or Q1 are the acceptors, whereas the activity
of NuoF E95Q is inhibited by only 50% when the DQ is the
electron acceptor and almost not at all with Q1 (Fig. 3).
Since the NuoF E95 is quite distant (over 100 Å) from the
quinone binding site and can hardly affect its structure
directly, the observed relative insensitivity to rolliniastatin
indicates that one half of the DQ and almost all of the Q1
reductase activity are due to quinone reduction at another
artificial site, presumably FMN.
An increased capability of electron exchange in the
catalytic site of the NuoF E95Q variant is also revealed
by a very different reactivity to dithionite. The wild-type
enzyme is reduced slowly on a minutes timescale,
whereas the reduction of NuoF E95Q occurs much faster
(Fig. 4). The kinetics of optical changes at 450 nm upon
reduction, where the FMN contribution to the Complex I
spectrum is maximal, is shown in Fig. 4. The half times
are approximately 3 min and 10 s, for wild type and NuoF
E95Q respectively. Thus the acceleration of enzyme
reduction by the mutation is of the same order as the
acceleration of hydrogen peroxide production.
On the basis of our results we may conclude that the
catalytic site of Complex I becomes leaky with respect to
reactivity with O2 due to loss of the negative charge of
E95, although the structure of the cavity containing the
FMN and the nucleotide binding site was not significantly
perturbed, since the affinity for NADH was only slightly
changed (Euro et al., 2009a).
The fact that the rate of reduction of Complex I by
dithionite was considerably accelerated in the NuoF E95Q
Table 2. ROS production by isolated Complex I induced by short-chain ubiquinones.
Sample
ROS productiona
– DQ Q1 Menadione
Purified wt 1.00 ± 0.02 1.16 ± 0.01 9.28 ± 0.28 12.10 ± 0.31
Purified E95Q 15.98 ± 0.46 22.78 ± 0.04 95.46 ± 2.25 144.87 ± 1.17
Membranes wt 1.00 ± 0.14 1.07 ± 0.09 3.93 ± 0.02 5.55 ± 0.02
Membranes E95Qb
6.04 ± 0.90 5.98 ± 0.26 40.17 ± 1.47 59.07 ± 2.25
a. The initial rate of H2O2 production is normalized to that of wild type in the absence of added quinones, which are 51.7 ± 3.5 nmol mg−1
min−1
for purified enzyme and 2.49 ± 0.34 nmol mg−1
min−1
for native membrane-bound Complex I respectively.
b. Normalized to estimated 1.5 times lower content of mutated Complex I in native membranes. Estimation is based on the ratio of differences
between the specific quinone reductase activities of the membranes and purified proteins.
Fig. 2. Redox titration of the rate of H2O2
formation by purified Complex I. The
normalized rate is plotted against the
apparent ambient redox potential calculated
from the ratio [NADH]/[NAD+
].
A. Titration of wild-type Complex I, circles,
and mutants lacking ubiquinone reductase
activity, NuoCD R274A, squares, and NuoM
E144A, triangles.
B. Comparison of the redox titration of
wild-type Complex I, circles, and NuoF E95Q,
squares. Note that the absolute values for the
mutant are 15 times larger in comparison with
wild type.
ROS production by Complex I 1193
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
variant is consistent with the rate-limiting step being the
approach of the dithionite oxoanion to the active site. In
the wild-type enzyme this reaction is apparently limited by
the negative charge of the carboxylate group of E95.
However, this cannot explain the higher reactivity of the
E95Q variant with neutral short-chain ubiquinones.
Discussion
Here we report the ∼15-fold enhancement in the steady-
state rate of ROS production (vROS
) by the E95Q variant
of E. coli Complex I at low ambient redox potentials
(maintained by the NADH/NAD+
pair) as the result of the
increased electron leakage from the catalytic site. It comes
from the X-ray crystallography data that ROS formation is
very unlikely to occur at any of the FeS clusters due to their
shielding by the protein matrix; only FMN is partially
exposed to the solvent (PDB entry: 3IAS). We also assume
that the ubiquinone binding site does not participate in
ROS production, at least under the given experimental
conditions. Thus, our observations can only be accounted
for by changes in the properties of the active site caused by
the mutation.
In Complex I the flavin cofactor is partially buried at
the bottom of a deep cavity forming the catalytic site
(Sazanov and Hinchliffe, 2006). The side-chain of Glu95/
Glu97 (E. coli numbering, subunit NuoF, or T. thermophi-
lus numbering, subunit Nqo1) is located at the same
depth as flavin and around 6 Å (N5FMN-OE1E97) away from
the flavin π-system. The carboxylate and the flavin form a
binding pocket for the nucleotide, as shown in Fig. 1 (PDB
entry: 3IAM). It follows from the structure (see Supporting
Fig. 3. Relative insensitivity of ubiquinone
reductase activity of E95Q to rolliniastatin.
NADH oxidation was monitored in the
presence of DQ (A, B) and Q1 (C, D) by
purified wild-type (A, C) and E95Q (B, D)
Complexes I. NADH addition is indicated by
arrow. Rolliniastatin, 1 μM, was added at zero
time as indicated.
Fig. 4. Kinetics of wild-type and NuoF E95Q Complexes I reduction by dithionite. Set of redox spectra upon the reduction is shown for wild
type and E95Q variant. Kinetics of optical changes at 450 nm is characterized by τ1
2
≈ 3 min for wild type (circles) and τ1
2 ≈ 10 s for E95Q
(squares).
1194 J. Knuuti et al. ■
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
information, section 1) as well as from the effect of the
E95Q mutation on the Em of FMN (Euro et al., 2009a) that
the Glu95 side-chain is most probably in the anionic state.
In the nucleotide-bound state, the H-bonding between the
Glu95 and NAD+
is very unlikely; however, the electric
field of its charged carboxylate group is plausibly to affect
both redox potentials of the bound nucleotide and the
flavin, despite the polar nature of the solvent-exposed
nucleotide-binding cavity.
Here we do not consider oxygen interaction with the
FMN radical, despite the fact that the latter is known as a
powerful ROS generator in many chemical or enzymatic
systems (see, e.g. Massey, 1994; 1995; Massey and
Ghisla, 1974; Mattevi, 2006). NADH is a strict two-
electron donor so that any FMN radical would derive from
subsequent oxidation of the FMNH−
state. Thus the popu-
lation of the FMN radical state will be dependent on the
population of the FMNH−
state. Moreover, the participation
of the two-electron-reduced FMN in ROS production by
mitochondrial Complex I was shown experimentally
(Kussmaul and Hirst, 2006).
The mechanism of O2 interaction with the reduced flavin
is not known in Complex I. However, it is generally
accepted that the primary step in ROS generation in other
flavoproteins and chemical systems is the rate-limiting spin
conversion and charge transfer between FMNH−
and O2
resulting in the ‘caged’ radical pair formation followed by a
rebinding of O2
•−
at the C4a atom of the flavin with the
formation of a flavin hydroperoxide species, which can
further produce H2O2 (Massey, 1994; Mattevi, 2006). Pro-
duction of both O2
•−
and H2O2 was shown experimentally in
Complex I, although the ratio between these two species
varied and was also reported to be different in mitochon-
drial and bacterial enzymes (Esterhazy et al., 2008).
Hence, in this study, we do not discriminate between O2
•−
and H2O2 as both are considered ROS, the focus being
concentrated on the nature of the electron leakage from the
catalytic site.
Since ROS production is considered virtually irrevers-
ible, based on the very high quenching efficiency of the
detection system, its observed rate in Complex I at steady
state can be quantitatively expressed as
v kROS ROS
FMNH= × −
[ ] (1)
where kROS
is the rate constant of the reaction and
[FMNH−
] is the population of the reduced flavin. Given the
measured redox potentials of FMN in the nucleotide-free
WT and mutant enzymes [−350 mV and −310 mV respec-
tively (Euro et al., 2009a)], FMN should stay almost fully
reduced in both variants at Eh < −380 mV (ambient redox
potentials, typically provided by NADH). If it was indeed
the case, the E95Q mutant should have revealed much
higher reactivity of the reduced FMN towards dioxygen
(that is, a higher value of kROS
, e.g. due to the increased
accessibility of dioxygen to the reduced FMN) in order to
ensure the observed change in vROS
.
In Complex I direct interaction of O2 with FMN is blocked
at the flavin distal (opposite to the substrate-binding) side
due to the steric hindrance caused by the peptide back-
bone and side-chains of Gly183, Glu184 and Asn92
(T. thermophilus numbering, PDB entry: 3IAM); the flanks
of the flavin are also protected by the protein, except for its
redox-inactive dimethylbenzene ring and the N5 atom. The
latter was suggested as the site of O2
•−
formation in model
flavoproteins (McDonald et al., 2011). In Complex I it is
exposed to the cavity large enough to hold one or two H2O
molecules and possibly O2 or O2
•−
; however, the cavity itself
is isolated from the rest of the nucleotide-binding pocket by
Ser96, Glu97 and Tyr180 peptide backbone, Ser96 and
Asp94 side-chains (T. thermophilus numbering, PDB
entry: 3IAM), and by the C3 amide group of bound nicoti-
namide, preventing both O2 binding to, and O2
•−
escape
from the flavin N5 atom. Hence O2 can approach FMN only
from the proximal (NAD+
-binding) side, and it seems that
the presence of either NADH or NAD+
bound to FMN
prohibits the access of O2. Therefore, ROS production is
likely to occur either in the nucleotide-free enzyme (i.e.
after NAD+
dissociation) or in the situation, where NAD+
is
present in the binding pocket but does not form the π–π
stacked complex with FMNH−
(usually referred to as
charge–transfer complex, CTC, in many flavoproteins),
thereby allowing O2 to reach the proximal side of the flavin
plain. In the latter case, two conformational sub-states for
the NAD+
-bound enzyme could be proposed: a ‘closed’
state [(NAD+
-FR
)closed
] for the NAD+
-FR
CTC and an ‘open’
state [(NAD+
-FR
)open
], where such complex is not formed.
The latter possibility has indeed been discussed in the
literature for various flavoproteins (Gassner et al., 1996;
Palfey et al., 2001; Palfey and McDonald, 2010) and is
further supported by the observation that in the X-ray
crystal structure of Complex I (Berrisford and Sazanov,
2009) the planes of the flavin and nicotinamide rings are
not strictly parallel, indicating a possible mixture of several
sub-states. However, the E95Q replacement can hardly
affect the O2 accessibility to FMN in either case: in the
nucleotide-free catalytic site the distance between FMN
and Glu95 is large enough to accommodate an O2 mol-
ecule (> 5 Å); in (NAD+
-FR
)open
state the negatively charged
Glu95 can only attract NAD+
[in comparison with (NAD+
-
FR
)closed
state] providing more access of O2 to FMN and
more ROS production in WT.
The E95Q replacement could also affect the kROS
value
(Eq. 1) if NAD+
dissociation rate ( )koff
NAD+
became higher
for E95Q with respect to WT (in other words, NAD+
was
retained longer in the binding site of WT, attracted by
Glu95 and preventing the O2 access to FMN). The
decreased Km
NADH
, as well as the decreased Ki for NAD+
and ADP-ribose in the mutant with respect to wild type
ROS production by Complex I 1195
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
(Euro et al., 2009a) may argue in favour of this sugges-
tion. However, the changes in apparent affinities upon the
mutation [2 times for Km
NADH
, 10 μM in WT and 5 μM in
E95Q, and 2.4 times for Ki
NAD+
, 1 mM in WT and 0.39 mM
in E95Q respectively (Euro et al., 2009a)] are not large
enough to provide 15 times stronger binding of O2 in the
mutant.
Another possibility for the change in kROS
is the mutation-
specific change of the redox potential of the FMN-
bound dioxygen (see Supporting information, section 2).
However, the estimated effect ranging from 0 to −15 mV,
depending on the position of the bound O2, is too small to
account for the observed effect in full.
Therefore, we conclude that the change in reactivity of
the reduced flavin with respect to O2 can hardly explain
the observed mutation effects. Due to this we consider the
thermodynamic aspect of the reaction of Complex I with
oxygen. From the experimentally measured effect of the
E95Q mutation on the Em of FMN in the absence of
nucleotides [ΔEm = +40 mV (Euro et al., 2009a)], the esti-
mated value of the dielectric constant of the cavity (ε ≈ 53,
see Eq. S1), and the structural difference between the
nucleotide-free and nucleotide-bound enzymes, the inter-
action potential between Glu95 and FMNH−
was found,
δϕ1 = −35 meV (see Supporting information, section 2).
On the other hand, the interaction between the negatively
charged Glu95 and the positively charged, oxidized nico-
tinamide moiety of bound NAD+
is, by absolute value,
much stronger, δϕ2 = +59 meV. The sum δϕ1 + δϕ2 = +24
meV (+0.6 kcal mol−1
) indicates that in the WT enzyme the
carboxylate group of Glu95 stabilizes FMNH−
in the
nucleotide-bound state (NAD+
-FR
) relative to the same
state in the mutant. Similar conclusion was obtained from
the density functional theory (DFT) calculations using a
small model system comprising only of FMN and NAD+
or
NADH in the CTC configuration and the Glu95 residue
(see Supporting information, section 3, Tables S1 and S2,
and Fig. S1). Based on the electrostatic interaction ener-
gies calculated from the partial charges derived from DFT
calculations, the (NAD+
-FR
)closed
state is found more stable
in WT by 1.1 kcal mol−1
(at ε = 50). These results predict
the opposite effect of the mutation on ROS production.
Relative destabilization of the NAD+
-FR
state in the mutant
also means higher dissociation constant for NAD+
: a
change in 1.1 kcal mol−1
would raise KD
NAD+
in the mutant
for almost one order of magnitude. Keeping in mind the
∼ 40 meV stabilization of the nucleotide-free state FR
in
the mutant, the change in KD
NAD+
would be even larger,
which is very unlikely, given the Ki
NAD+
data (see above).
Therefore, the other explanation of the observed phenom-
enon should be suggested.
Finally, we analysed how the mutation affects the
steady-state probability, or life time, of transient enzyme
state(s) competent in ROS production ([FMNH−
] in Eq. 1).
The reaction scheme is shown in Fig. 5. After FMN reduc-
tion (step 3) there are two alternatives. In Cycle A, NAD+
dissociation (step 4) is followed by the O2 reaction with
state FR
(step 6). Note that Cycle A is realized only in the
situation when the FeS chain is reduced to a great extent,
so that the electron transfer from FMNH−
to the FeS chain
is no longer possible [step 5, kET2 0= , see for a review
(Verkhovskaya and Bloch, 2013)]. The other alternative is
Cycle B, where O2 reacts with the state NAD+
-FR
(step 7).
If NAD+
dissociation from FMNH−
is significantly slower
than ROS production ( )k koff
NAD ROS+
<< , then the contribu-
tion of Cycle A is negligible and Cycle B becomes pre-
dominant. Unlike Cycle A, Cycle B is independent of the
reduction level of the FeS chain. Although in various
NAD(P)H-dependent flavoenzymes NAD+
can leave the
binding site either before or after the oxidation of FMN to
Fig. 5. Kinetic scheme of the steady-state ROS production by Complex I. F and FR
stand for the nucleotide-free enzyme with oxidized and
two-electron-reduced FMN respectively; NAD+
-F and NAD+
-FR
are the enzyme with bound NAD+
; NADH-F is the enzyme with bound NADH.
Both ‘closed’ and ‘open’ states of NAD+
-FR
are shown. Redox state of the FeS cluster chain is not shown for simplicity. Highlighted are the
ROS-producing states (NAD+
-FR
)open
and FR
. Reaction steps are: NAD+
dissociation from oxidized enzyme (1), NADH binding to oxidized
enzyme (2), electron transfer from NADH to FMN (presumably through the hydride transfer within the charge–transfer complex, 3), NAD+
dissociation from the enzyme with reduced FMN (4), electron transfer from reduced FMN to the FeS cluster chain (5), interaction of reduced
FMN with dioxygen and ROS production (6, 7), respectively, as indicated by numbers. All reaction steps (1–5), except for the ROS production
(6, 7), are thermodynamically reversible; arrows show the preferred fluxes under typical conditions. Dotted line shows the pre-steady-state
reduction of the enzyme-bound FeS cluster chain by FR
. Dashed lines show the irreversible two-electron reduction of dioxygen.
1196 J. Knuuti et al. ■
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
a secondary substrate, there are examples in the litera-
ture where flavin remains reduced for a considerable time
(e.g. when the electron acceptors are not available) and
NAD+
stays attached to it as long as the flavin remains
reduced (Gassner et al., 1996), which should be due to
the electrostatic interactions between NAD+
and FMNH−
.
Seemingly contradictory with the latter proposal, the
experimentally measured, relatively high rate constant of
NAD+
dissociation [ koff
NAD
s
+
≈ −
103 1
(Verkhovskaya et al.,
2008)] was in fact determined under the conditions where
the FeS chain and FMN were not fully reduced; thus, two
different koff
NAD+
values should be considered, depending
on the reduction state of the flavin (Fig. 5, steps 1 and 4
respectively).
Noteworthy, Cycle B can be realized only when the
steady-state electron flux from NAD+
-FR
to O2 takes place.
In the absence of O2, the reaction will inevitably proceed
to the nucleotide-free, fully reduced state (FR
), despite its
accumulation is slow.
To quantitatively analyse the yield of the FMNH−
states,
the analytical solution was obtained for each cycle sepa-
rately (Cycle A, steps 2, 3, 4, and Cycle B, steps 1, 2, 3)
at equilibrium and without the electron leakage to O2
(kROS
= 0) (Fig. 6; see also: Supporting information,
section 4, Eqs S2–S4 and Figs S2 and S3). Cycle A does
not show any difference in the population of ROS-
production state (FR
) between WT and E95Q at low Eh
(< −380 mV). The equilibrium solution for Cycle B predicts
that the flavin can be reduced only to a certain, small
extent in the NADH-reduced enzyme even at low poten-
tials. Despite the enzyme is considered in redox equilib-
rium with NADH/NAD+
(given that the absolute ROS
production rate in Complex I, 0.5–1.0 s−1
in WT, is orders
of magnitude lower than the enzyme turnover rate,
kcat > 300 s−1
), in the absence of redox mediators the
sequence of the steps restricts the yield of FMNH−
: NADH
binding (Fig. 5, step 2) is followed by the electron redis-
tribution between NADH and FMN (step 3) according to
their Em; at the same time, the preceding NAD+
dissocia-
tion (step 1) does not contribute to repopulate FMNH−
.
When the electron transfer from NADH to FMN is uphill
( )/
ΔE E Em m
FMNH /FMN
m
NADH NAD
= − <
− +
0 , then [NAD+
-FR
] is low;
when it is downhill (ΔEm > 0), [NAD+
-FR
] is higher. Assum-
ing Em
NADH NAD
mV/ +
= −320 and the experimentally deter-
mined Em values for FMN, −350 (WT) and −310 (E95Q),
the former situation is realized in WT and the latter, in the
Fig. 6. Simulated Eh-dependences of the
equilibrium concentrations of intermediates in
Cycles A and B. The parameters used
are: Ntot
= [NADH] + [NAD+
] = 10−4
M,
KD
NADH
M= −
10 5
, KD
NAD
M
+
= −
10 5
,
Em
N
mV= −320 , Em
F
mV= −350 (WT),
Em
F
mV= −310 (E95Q). The ambient redox
potential is set by [NADH] and [NAD+
]:
E Eh m
N NADH
NAD
= − × +
30 lg
[ ]
[ ]
. All redox potentials
quoted, refer to the NHE scale. For the
respective kinetic schemes, see Fig. 5. For
further details, see Supporting information,
section 4. Note that for Cycle B, the
parameter set differs from that found by the
data fitting (see Fig. S5 and Table S3). The
parameter set used here is the one showing
the difference between the two cycles the
most (see Supporting information, section 4).
ROS production by Complex I 1197
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
mutant (Fig. 6, lower panel). By varying the KD values for
WT and E95Q, it was possible to obtain a good fit of the
vROS
data, given both experimental effects of the mutation:
the 15-fold increase in vROS
at low Eh and the ∼ 20 mV
upshift of the apparent midpoint potential (E 1
2) of the vROS
Eh-dependence (see Supporting information, sections 4
and 5, Eqs S5–S10, Figs S4 and S5, and Table S3).
As shown by the electrostatic and DFT calculations (see
above), the NAD+
-FR
state is relatively stable in WT as
compared with the mutant, which apparently contradicts
both the observations and the kinetic modelling. However,
the uncertainty in the position of nicotinamide in the ROS-
producing (NAD+
-FR
)open
state might affect the result of
such calculations. Since the nucleotide-binding cavity is
wide enough to accommodate NAD+
in a different position
as compared with the resolved structure (see Supporting
information, section 1), the electrostatic interaction
between Glu95, FMNH−
and NAD+
may vary strongly. The
lack of the expected mutation-caused change in the NAD+
affinity (Ki
NAD+
, see above) further supports this possibility.
We note that the equilibrium solution of Cycle B (Fig. 6,
lower panel) predicts a significant amount of the
nucleotide-bound enzyme state(s) at low Eh, which are
known to form CTC in many flavoproteins. However, CTC
can be difficult to detect due to strong variation in both
extinction coefficient and maximum wavelength; more-
over, both NADH-F and NAD+
-FR
CTC are often found to
be transient, short-lived states (see Massey and Ghisla,
1974; Gao and Ellis, 2007; Lin et al., 2012; and refer-
ences therein). No information about optical properties of
CTC in Complex I has been available yet. We were not
able to detect any signature of CTC in E. coli Complex I
under the ROS-producing conditions, either (data not
shown). Since in Complex I, the proposed ROS-producing
state (NAD+
-FR
)open
is not a CTC, it should not have an
absorption signature. On the other hand, the maximum
yield of [NAD+
-FR
] can be rather low in both WT and the
mutant in the NADH-non-saturating conditions and
depend on KD
NADH
(see Fig. S4 and Eq. S9).
An important implication of this analysis is the finding
that even low electron leakage from NAD+
-FR
results in
returning the enzyme to the non-ROS-producing state
(NADH-F), which accumulates. Therefore, the population
of the ROS producing states is always low in the WT
enzyme. We conclude that Complex I possesses an
inbuilt safety mechanism that prevents excessive reduc-
tion of FMN and consequently minimizes ROS production
at low ambient redox potentials regardless of the avail-
ability of the natural substrate, ubiquinone. This mecha-
nism is based on the stabilization of an intermediate state
(non-reactive with O2) of partially reduced enzyme, where
the FeS chain is either oxidized or reduced, but only to a
certain extent, and slow NAD+
release from the state with
reduced flavin.
Experimental procedures
Escherichia coli strains MWC215 (Δnhd) (Calhoun and
Gennis, 1993), GR70N (Green et al., 1984) (both have wild-
type Complex I) and variants NuoM E144A (Euro et al.,
2008a), NuoF E95Q (Euro et al., 2009a) and NuoCD R274A
(Belevich et al., 2007) were grown aerobically in LB medium
supplemented with appropriate antibiotics on demand at
37°C. Inverted cytoplasmic membrane vesicles were pre-
pared after cell wall removal by sonication as described pre-
viously (Belevich et al., 2007), and used the same day for
analysis. Enzyme extraction and further purification from iso-
lated membranes was done as previously described (Euro
et al., 2009b). The HAR and DQ reductase activities of the
membrane-bound and purified Complex I were measured as
in Belevich et al. (2011). Reconstitution of Complex I into
liposomes was performed as described in Verkhovskaya
et al. (2011). Protein concentration in the membrane prepa-
rations was determined by the BCA protein assay kit (Pierce)
using BSA as a standard. Absorbance at 280 nm was used to
quantify the purified protein (ε280 = 785160 M−1
cm−1
).
Complex I mediated NADH-dependent H2O2 production
was quantified by using 0.4 units ml−1
of horse-radish per-
oxidase (Sigma) catalysing oxidation of 10 μM Amplex
Red (Sigma) to resorufin (ε557–620 = 51.6 mM−1
cm−1
) at 30°C.
NADH was used for purified Complex I and nicotinamide
hypoxanthine dinucleotide (dNADH) for native membrane
vesicles of GR70N and its variants NuoF E95Q, NuoCD
R274A and NuoM E144A to avoid contribution of NADH:ubiq-
uinone reductase type II, which was completely removed
after purification. The MWC215 strain is a GR70N derivative
lacking the gene encoding this alternative NADH dehydroge-
nase; it was grown for high-yield wild-type enzyme purifica-
tion. H2O2 production was measured in 200 mM HEPES-KOH
pH 7.0 (RT) or 20 mM HEPES-KOH containing 360 mM man-
nitol pH 7.0 (RT). Initial ROS production rate was same in
both buffers, but the latter was chosen for the measurements
of Complex I proteoliposomes to preserve electrogenic prop-
erties of the enzyme. 0.005% n-dodecyl β-D-maltoside (DDM)
was included when solubilized enzyme was used, at con-
centrations of 0.6 μg ml-1
(1.0 nM) for E95Q variant and
6.1–7.8 μg ml−1
(11–15 nM) for other variants and WT respec-
tively. Media was supplemented with 1 μM rolliniastatin when
ROS production by membrane preparations was defined, at
total protein concentration of 45 μg ml−1
for E95Q variant and
90 μg ml-1
for other variants and WT respectively. H2O2 for-
mation by purified Complex I was titrated using the
NADH:NAD+
redox couple to set the ambient potential. The
nucleotides, NADH or dNADH and NAD+
were added to a
total concentration of 35 μM or (d)NADH concentration was
kept constant and NAD+
concentration was varied. The reac-
tion was initiated by Complex I addition and the rate, which
appeared non-linear, was quantified right after the enzyme
addition allowing short mixing and defining the maximal rate
as velocity without NAD+
addition. Background of H2O2 gen-
eration was determined in the presence of 525 units ml−1
catalase (Sigma).
Kinetic measurements were carried out using a high-
resolution CCD-array spectrometer (HR2000+, Ocean
Optics). The assay buffer comprised 50 mM HEPES-KOH,
pH 7.0, 15% glycerol and 0.005% DDM. Complex I was
added at a final concentration of 2.7–2.9 μM.
1198 J. Knuuti et al. ■
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
Numerical simulations and the data fit were performed
using Matlab R2010 (MathWorks, Natick, MA). The data fit
was accomplished using a non-linear, constrained solver
‘fmincon’. A full description of the kinetic models is given in
Supporting information. The details of DFT calculations are
also described in Supporting information (section 3).
Acknowledgements
This work was supported by grants from Biocentrum Helsinki,
the Sigrid Jusélius Foundation and the Academy of Finland.
V.S. acknowledges postdoctoral researcher funding from the
Academy of Finland. We thank Prof. Mårten Wikström for
extensive discussion and his interest to this work. We thank
Eija Haasanen for purification of Complex I. We are also
thankful to the Center for Scientific Computing (CSC),
Finland, for providing computing time.
References
Barja, G., and Herrero, A. (1998) Localization at complex I
and mechanism of the higher free radical production of
brain nonsynaptic mitochondria in the short-lived rat than in
the longevous pigeon. J Bioenerg Biomembr 30: 235–
243.
Belevich, G., Euro, L., Wikström, M., and Verkhovskaya, M.
(2007) Role of the conserved arginine 274 and histidine
224 and 228 residues in the NuoCD subunit of complex I
from Escherichia coli. Biochemistry 46: 526–533.
Belevich, G., Knuuti, J., Verkhovsky, M.I., Wikström, M., and
Verkhovskaya, M. (2011) Probing the mechanistic role of
the long alpha-helix in subunit L of respiratory Complex I
from Escherichia coli by site-directed mutagenesis. Mol
Microbiol 82: 1086–1095.
Berrisford, J.M., and Sazanov, L.A. (2009) Structural basis
for the mechanism of respiratory complex I. J Biol Chem
284: 29773–29783.
Brand, M.D. (2010) The sites and topology of mitochondrial
superoxide production. Exp Gerontol 45: 466–472.
Cadenas, E., Boveris, A., Ragan, C.I., and Stoppani, A.O.M.
(1977) Production of Superoxide radicals and hydrogen-
peroxide by NADH-ubiquinone reductase and ubiquinol-
cytochrome C reductase from beef-heart mitochondria.
Arch Biochem Biophys 180: 248–257.
Calhoun, M.W., and Gennis, R.B. (1993) Demonstration of
separate genetic loci encoding distinct membrane-bound
respiratory NADH dehydrogenases in Escherichia coli. J
Bacteriol 175: 3013–3019.
Case, C.L., Rodriguez, J.R., and Mukhopadhyay, B. (2009)
Characterization of an NADH oxidase of the flavin-
dependent disulfide reductase family from Methanocaldo-
coccus jannaschii. Microbiology 155: 69–79.
Drechsel, D.A., and Patel, M. (2010) Respiration-dependent
H2O2 removal in brain mitochondria via the thioredoxin/
peroxiredoxin system. J Biol Chem 285: 27850–27858.
Drose, S., Galkin, A., and Brandt, U. (2009) Measurement of
superoxide formation by mitochondrial complex I of Yar-
rowia lipolytica. Methods Enzymol 456: 475–490.
Esterhazy, D., King, M.S., Yakovlev, G., and Hirst, J. (2008)
Production of reactive oxygen species by complex I (NAD-
H:ubiquinone oxidoreductase) from Escherichia coli and
comparison to the enzyme from mitochondria. Biochemis-
try 47: 3964–3971.
Euro, L., Belevich, G., Verkhovsky, M.I., Wikström, M., and
Verkhovskaya, M. (2008a) Conserved lysine residues of
the membrane subunit NuoM are involved in energy con-
version by the proton-pumping NADH:ubiquinone oxidore-
ductase (Complex I). Biochim Biophys Acta 1777: 1166–
1172.
Euro, L., Bloch, D.A., Wikström, M., Verkhovsky, M.I., and
Verkhovskaya, M. (2008b) Electrostatic interactions
between FeS clusters in NADH:ubiquinone oxidoreductase
(Complex I) from Escherichia coli. Biochemistry 47: 3185–
3193.
Euro, L., Belevich, G., Bloch, D.A., Verkhovsky, M.I.,
Wikström, M., and Verkhovskaya, M. (2009a) The role of
the invariant glutamate 95 in the catalytic site of Complex I
from Escherichia coli. Biochim Biophys Acta 1787: 68–
73.
Euro, L., Belevich, G., Wikström, M., and Verkhovskaya, M.
(2009b) High affinity cation-binding sites in Complex I from
Escherichia coli. Biochim Biophys Acta 1787: 1024–1028.
Fukui, H., and Moraes, C.T. (2008) The mitochondrial impair-
ment, oxidative stress and neurodegeneration connection:
reality or just an attractive hypothesis? Trends Neurosci 31:
251–256.
Galkin, A., and Brandt, U. (2005) Superoxide radical forma-
tion by pure complex I (NADH:ubiquinone oxidoreductase)
from Yarrowia lipolytica. J Biol Chem 280: 30129–30135.
Galkin, A., Dröse, S., and Brandt, U. (2006) The proton
pumping stoichiometry of purified mitochondrial complex I
reconstituted into proteoliposomes. Biochim Biophys Acta
1757: 1575–1581.
Galkin, A.S., Grivennikova, V.G., and Vinogradov, A.D.
(1999) -->H+/2e- stoichiometry in NADH-quinone reduc-
tase reactions catalyzed by bovine heart submitochondrial
particles. FEBS Lett 451: 157–161.
Gao, B., and Ellis, H.R. (2007) Mechanism of flavin reduction
in the alkanesulfonate monooxygenase system. Biochim
Biophys Acta 1774: 359–367.
Gassner, G.T., Johnson, D.A., Liu, H.W., and Ballou, D.P.
(1996) Kinetics of the reductive half-reaction of the iron-
sulfur flavoenzyme CDP-6-deoxy-L-threo-D-glycero-4-
hexulose-3-dehydrase reductase. Biochemistry 35: 7752–
7761.
Green, G.N., Kranz, R.G., Lorence, R.M., and Gennis, R.B.
(1984) Identification of subunit I as the cytochrome b558
component of the cytochrome d terminal oxidase complex
of Escherichia coli. J Biol Chem 259: 7994–7997.
Grivennikova, V.G., and Vinogradov, A.D. (2006) Generation
of superoxide by the mitochondrial Complex I. Biochim
Biophys Acta 1757: 553–561.
Hinkle, P.C., Penefsky, H.S., and Racker, E. (1967) Partial
resolution of enzymes catalyzing oxidative phosphorylation.
12. H2
18
O-inorganic phosphate and H2
18
O-adenosine
triphosphate exchange reactions in submitochondrial parti-
cles from beef heart. J Biol Chem 242: 1788–1792.
Hinkle, P.C., Kumar, M.A., Resetar, A., and Harris, D.L.
(1991) Mechanistic stoichiometry of mitochondrial oxida-
tive phosphorylation. Biochemistry 30: 3576–3582.
Kareyeva, A.V., Grivennikova, V.G., and Vinogradov, A.D.
(2012) Mitochondrial hydrogen peroxide production as
ROS production by Complex I 1199
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
determined by the pyridine nucleotide pool and its redox
state. Biochim Biophys Acta 1817: 1879–1885.
King, M.S., Sharpley, M.S., and Hirst, J. (2009) Reduction of
hydrophilic ubiquinones by the flavin in mitochondrial NAD-
H:ubiquinone oxidoreductase (Complex I) and production
of reactive oxygen species. Biochemistry 48: 2053–2062.
Kussmaul, L., and Hirst, J. (2006) The mechanism of super-
oxide production by NADH:ubiquinone oxidoreductase
(complex I) from bovine heart mitochondria. Proc Natl Acad
Sci USA 103: 7607–7612.
Lambert, A.J., and Brand, M.D. (2004a) Inhibitors of the
quinone-binding site allow rapid superoxide production
from mitochondrial NADH:ubiquinone oxidoreductase
(complex I). J Biol Chem 279: 39414–39420.
Lambert, A.J., and Brand, M.D. (2004b) Superoxide produc-
tion by NADH:ubiquinone oxidoreductase (complex I)
depends on the pH gradient across the mitochondrial inner
membrane. Biochemistry 382: 511–517.
Lin, M.T., and Beal, M.F. (2006) Mitochondrial dysfunction
and oxidative stress in neurodegenerative diseases.
Nature 443: 787–795.
Lin, T.Y., Werther, T., Jeoung, J.H., and Dobbek, H. (2012)
Suppression of electron transfer to dioxygen by charge
transfer and electron transfer complexes in the FAD-
dependent reductase component of toluene dioxygenase.
J Biol Chem 287: 38338–38346.
McDonald, C.A., Fagan, R.L., Collard, F., Monnier, V.M., and
Palfey, B.A. (2011) Oxygen reactivity in flavoenzymes:
context matters. J Am Chem Soc 133: 16809–16811.
Massey, V. (1994) Activation of molecular oxygen by flavins
and flavoproteins. J Biol Chem 269: 22459–22462.
Massey, V. (1995) Introduction: flavoprotein structure and
mechanism. FASEB J 9: 473–475.
Massey, V., and Ghisla, S. (1974) Role of charge–transfer
interactions in flavoprotein catalysis. Ann N Y Acad Sci
227: 446–465.
Matsuda, H., Kimura, S., and Iyanagi, T. (2000) One-electron
reduction of quinones by the neuronal nitric-oxide synthase
reductase domain. Biochim Biophys Acta 1459: 106–116.
Mattevi, A. (2006) To be or not to be an oxidase: challenging
the oxygen reactivity of flavoenzymes. Trends Biochem Sci
31: 276–283.
Ohnishi, S.T., Shinzawa-Itoh, K., Ohta, K., Yoshikawa, S.,
and Ohnishi, T. (2010) New insights into the superoxide
generation sites in bovine heart NADH-ubiquinone oxidore-
ductase (Complex I): the significance of protein-associated
ubiquinone and the dynamic shifting of generation sites
between semiflavin and semiquinone radicals. Biochim
Biophys Acta 1797: 1901–1909.
Palfey, B.A., and McDonald, C.A. (2010) Control of catalysis
in flavin-dependent monooxygenases. Arch Biochem
Biophys 493: 26–36.
Palfey, B.A., Bjornberg, O., and Jensen, K.F. (2001) Insight
into the chemistry of flavin reduction and oxidation in
Escherichia coli dihydroorotate dehydrogenase obtained
by rapid reaction studies. Biochemistry 40: 4381–4390.
Pryde, K.R., and Hirst, J. (2011) Superoxide is produced by
the reduced flavin in mitochondrial complex I a single,
unified mechanism that applies during both forward and
reverse electron transfer. J Biol Chem 286: 18056–18065.
Rhein, V., Baysang, G., Rao, S., Meier, F., Bonert, A.,
Muller-Spahn, F., and Eckert, A. (2009) Amyloid-beta leads
to impaired cellular respiration, energy production and
mitochondrial electron chain complex activities in human
neuroblastoma cells. Cell Mol Neurobiol 29: 1063–1071.
Sazanov, L.A., and Hinchliffe, P. (2006) Structure of the
hydrophilic domain of respiratory complex I from Thermus
thermophilus. Science 311: 1430–1436.
Sled, V., Rudnitzky, N., Hatefi, Y., and Ohnishi, T. (1993)
Thermodynamic analysis of FMN in mitochondrial NADH
ubiquinone oxidoreductase (Complex-I). Biophys J 64:
A105–A105.
Starkov, A.A., Fiskum, G., Chinopoulos, C., Lorenzo, B.J.,
Browne, S.E., Patel, M.S., and Beal, M.F. (2004) Mitochon-
drial alpha-ketoglutarate dehydrogenase complex gener-
ates reactive oxygen species. J Neurosci 24: 7779–7788.
Treberg, J.R., Quinlan, C.L., and Brand, M.D. (2011) Evi-
dence for two sites of superoxide production by mitochon-
drial NADH-ubiquinone oxidoreductase (Complex I). J Biol
Chem 286: 27103–27110.
Turrens, J.F. (2003) Mitochondrial formation of reactive
oxygen species. J Physiol 552: 335–344.
Turrens, J.F., and Boveris, A. (1980) Generation of superox-
ide anion by the NADH dehydrogenase of bovine heart-
mitochondria. Biochem J 191: 421–427.
Verkhovskaya, M., and Bloch, D.A. (2013) Energy-converting
respiratory Complex I: on the way to the molecular mecha-
nism of the proton pump. Int J Biochem Cell Biol 45:
491–511.
Verkhovskaya, M., Knuuti, J., and Wikström, M. (2011) Role
of Ca2+
in structure and function of Complex I from Escheri-
chia coli. Biochim Biophys Acta 1807: 36–41.
Verkhovskaya, M.L., Belevich, N., Euro, L., Wikström, M.,
and Verkhovsky, M.I. (2008) Real-time electron transfer in
respiratory complex I. Proc Natl Acad Sci USA 105: 3763–
3767.
Vinogradov, A.D., and Grivennikova, V.G. (2005) Generation
of superoxide-radical by the NADH:ubiquinone oxidoreduc-
tase of heart mitochondria. Biochemistry (Mosc) 70: 120–
127.
Wikström, M. (1984) Two protons are pumped from the mito-
chondrial matrix per electron transferred between NADH
and ubiquinone. FEBS Lett 169: 300–304.
Wikström, M., and Hummer, G. (2012) Stoichiometry of proton
translocation by respiratory complex I and its mechanistic
implications. Proc Natl Acad Sci USA 109: 4431–4436.
Zoccarato, F., Cavallini, L., and Alexandre, A. (2004)
Respiration-dependent removal of exogenous H2O2 in
brain mitochondria – inhibition by Ca2+
. J Biol Chem 279:
4166–4174.
Supporting information
Additional supporting information may be found in the online
version of this article at the publisher’s web-site.
1200 J. Knuuti et al. ■
© 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200

More Related Content

What's hot

SHIHeterologousExpressionPLOSONE
SHIHeterologousExpressionPLOSONESHIHeterologousExpressionPLOSONE
SHIHeterologousExpressionPLOSONEPatrick McTernan
 
LGRAUX_Research summary
LGRAUX_Research summaryLGRAUX_Research summary
LGRAUX_Research summaryLionel Graux
 
Free radicals
Free radicalsFree radicals
Free radicalsakif12345
 
Postdoctoral Research Summary
Postdoctoral Research SummaryPostdoctoral Research Summary
Postdoctoral Research SummaryAdam Powell
 
Reactive Oxygen Species,Reactive Nitrogen Species and Redox Signaling
Reactive Oxygen Species,Reactive Nitrogen Species and Redox SignalingReactive Oxygen Species,Reactive Nitrogen Species and Redox Signaling
Reactive Oxygen Species,Reactive Nitrogen Species and Redox SignalingJeju National University
 
Copper complexes as anti cancer agent ppt
Copper complexes as anti cancer agent pptCopper complexes as anti cancer agent ppt
Copper complexes as anti cancer agent pptPARTHASARATHIMAHANTA
 
Pharmacology of free radicals
Pharmacology of free radicalsPharmacology of free radicals
Pharmacology of free radicalsShivam Diwaker
 
Oxidative stress and Alzheimer's disease
Oxidative stress and Alzheimer's diseaseOxidative stress and Alzheimer's disease
Oxidative stress and Alzheimer's diseaseMandeep Bhullar
 
Free radicals and antioxidants
Free radicals and antioxidantsFree radicals and antioxidants
Free radicals and antioxidantsDr Alok Bharti
 
Free radical induced injury
Free radical induced injuryFree radical induced injury
Free radical induced injuryPulak Agrawal
 
Final presantation mike
Final presantation mikeFinal presantation mike
Final presantation mikemrjanucik
 
Particle and Particle Systems Characterization_2013
Particle and Particle Systems Characterization_2013Particle and Particle Systems Characterization_2013
Particle and Particle Systems Characterization_2013Nidhi Basak
 
13. free radicals and antioxidants
13. free radicals and antioxidants13. free radicals and antioxidants
13. free radicals and antioxidantsMadhumita Sen
 
Tfa_N-methylation
Tfa_N-methylationTfa_N-methylation
Tfa_N-methylationJordan Gipe
 

What's hot (20)

SHIHeterologousExpressionPLOSONE
SHIHeterologousExpressionPLOSONESHIHeterologousExpressionPLOSONE
SHIHeterologousExpressionPLOSONE
 
Oxidative phosphorylation april batch 2014
Oxidative phosphorylation april batch 2014Oxidative phosphorylation april batch 2014
Oxidative phosphorylation april batch 2014
 
Mechanisim and generation of free radicals
Mechanisim and generation of free radicalsMechanisim and generation of free radicals
Mechanisim and generation of free radicals
 
Free radicals
Free radicalsFree radicals
Free radicals
 
LGRAUX_Research summary
LGRAUX_Research summaryLGRAUX_Research summary
LGRAUX_Research summary
 
Free radicals
Free radicalsFree radicals
Free radicals
 
J. Org. Chem. 2002 Grotjahn
J. Org. Chem. 2002 GrotjahnJ. Org. Chem. 2002 Grotjahn
J. Org. Chem. 2002 Grotjahn
 
2012Brookins_TeQuion
2012Brookins_TeQuion2012Brookins_TeQuion
2012Brookins_TeQuion
 
Postdoctoral Research Summary
Postdoctoral Research SummaryPostdoctoral Research Summary
Postdoctoral Research Summary
 
Copper in proteins and enzymes
Copper in proteins and enzymesCopper in proteins and enzymes
Copper in proteins and enzymes
 
Reactive Oxygen Species,Reactive Nitrogen Species and Redox Signaling
Reactive Oxygen Species,Reactive Nitrogen Species and Redox SignalingReactive Oxygen Species,Reactive Nitrogen Species and Redox Signaling
Reactive Oxygen Species,Reactive Nitrogen Species and Redox Signaling
 
Copper complexes as anti cancer agent ppt
Copper complexes as anti cancer agent pptCopper complexes as anti cancer agent ppt
Copper complexes as anti cancer agent ppt
 
Pharmacology of free radicals
Pharmacology of free radicalsPharmacology of free radicals
Pharmacology of free radicals
 
Oxidative stress and Alzheimer's disease
Oxidative stress and Alzheimer's diseaseOxidative stress and Alzheimer's disease
Oxidative stress and Alzheimer's disease
 
Free radicals and antioxidants
Free radicals and antioxidantsFree radicals and antioxidants
Free radicals and antioxidants
 
Free radical induced injury
Free radical induced injuryFree radical induced injury
Free radical induced injury
 
Final presantation mike
Final presantation mikeFinal presantation mike
Final presantation mike
 
Particle and Particle Systems Characterization_2013
Particle and Particle Systems Characterization_2013Particle and Particle Systems Characterization_2013
Particle and Particle Systems Characterization_2013
 
13. free radicals and antioxidants
13. free radicals and antioxidants13. free radicals and antioxidants
13. free radicals and antioxidants
 
Tfa_N-methylation
Tfa_N-methylationTfa_N-methylation
Tfa_N-methylation
 

Similar to ROS knuuti

Electron Transport Chain
Electron Transport ChainElectron Transport Chain
Electron Transport ChainAnujaParkhe
 
Oxidative phosphorylation1
Oxidative phosphorylation1Oxidative phosphorylation1
Oxidative phosphorylation1Navneet Kumar
 
ELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.ppt
ELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.pptELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.ppt
ELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.pptRamieMollid
 
BT631-26-Membrane_proteins
BT631-26-Membrane_proteinsBT631-26-Membrane_proteins
BT631-26-Membrane_proteinsRajesh G
 
reactive oxygen species in periodontal disease
reactive oxygen species in periodontal diseasereactive oxygen species in periodontal disease
reactive oxygen species in periodontal diseaseSuhani Goel
 
free radicals and antioxidants.pptx
free radicals and antioxidants.pptxfree radicals and antioxidants.pptx
free radicals and antioxidants.pptxDrManojAcharya1
 
Oxidative phosphorylation
Oxidative phosphorylationOxidative phosphorylation
Oxidative phosphorylationMohammedLuqman9
 
Chapter 7 Energy transduction in cells.pptx
Chapter 7 Energy transduction in cells.pptxChapter 7 Energy transduction in cells.pptx
Chapter 7 Energy transduction in cells.pptxAsmamawTesfaw1
 
Electron Transport Chain & Oxidative Phosphorylation
Electron Transport Chain & Oxidative PhosphorylationElectron Transport Chain & Oxidative Phosphorylation
Electron Transport Chain & Oxidative PhosphorylationSyed Muhammad Khan
 
Electron Transport and Oxidative Phosphorylation
Electron Transport and Oxidative PhosphorylationElectron Transport and Oxidative Phosphorylation
Electron Transport and Oxidative PhosphorylationHamid Ur-Rahman
 
Fielding_DoctoralResearch
Fielding_DoctoralResearchFielding_DoctoralResearch
Fielding_DoctoralResearchAndrew Fielding
 
Biological oxidation and etc
Biological oxidation and etcBiological oxidation and etc
Biological oxidation and etcNidhi Argade
 
Andover_poster_F_15
Andover_poster_F_15Andover_poster_F_15
Andover_poster_F_15Juho Knuuti
 
FREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdf
FREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdfFREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdf
FREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdfTARENIDAS
 
Electron Transport Chain lecture.ppt
Electron Transport Chain lecture.pptElectron Transport Chain lecture.ppt
Electron Transport Chain lecture.pptNatalya80
 

Similar to ROS knuuti (20)

Electron Transport Chain
Electron Transport ChainElectron Transport Chain
Electron Transport Chain
 
Oxidative phosphorylation1
Oxidative phosphorylation1Oxidative phosphorylation1
Oxidative phosphorylation1
 
ELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.ppt
ELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.pptELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.ppt
ELECTRON TRANSPORT CHAIN AND OXIDATIVE PHOSPHORYLATION.ppt
 
Biological oxidation
Biological oxidationBiological oxidation
Biological oxidation
 
Maryam Bachelor thesis
Maryam Bachelor thesisMaryam Bachelor thesis
Maryam Bachelor thesis
 
BT631-26-Membrane_proteins
BT631-26-Membrane_proteinsBT631-26-Membrane_proteins
BT631-26-Membrane_proteins
 
review catalysis heterogenius.pdf
review catalysis heterogenius.pdfreview catalysis heterogenius.pdf
review catalysis heterogenius.pdf
 
Biological oxidation
Biological oxidationBiological oxidation
Biological oxidation
 
Biological oxidation
Biological oxidationBiological oxidation
Biological oxidation
 
reactive oxygen species in periodontal disease
reactive oxygen species in periodontal diseasereactive oxygen species in periodontal disease
reactive oxygen species in periodontal disease
 
free radicals and antioxidants.pptx
free radicals and antioxidants.pptxfree radicals and antioxidants.pptx
free radicals and antioxidants.pptx
 
Oxidative phosphorylation
Oxidative phosphorylationOxidative phosphorylation
Oxidative phosphorylation
 
Chapter 7 Energy transduction in cells.pptx
Chapter 7 Energy transduction in cells.pptxChapter 7 Energy transduction in cells.pptx
Chapter 7 Energy transduction in cells.pptx
 
Electron Transport Chain & Oxidative Phosphorylation
Electron Transport Chain & Oxidative PhosphorylationElectron Transport Chain & Oxidative Phosphorylation
Electron Transport Chain & Oxidative Phosphorylation
 
Electron Transport and Oxidative Phosphorylation
Electron Transport and Oxidative PhosphorylationElectron Transport and Oxidative Phosphorylation
Electron Transport and Oxidative Phosphorylation
 
Fielding_DoctoralResearch
Fielding_DoctoralResearchFielding_DoctoralResearch
Fielding_DoctoralResearch
 
Biological oxidation and etc
Biological oxidation and etcBiological oxidation and etc
Biological oxidation and etc
 
Andover_poster_F_15
Andover_poster_F_15Andover_poster_F_15
Andover_poster_F_15
 
FREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdf
FREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdfFREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdf
FREE RADICAL CELL INJURY PPT BY Dr. Tareni Das.pdf
 
Electron Transport Chain lecture.ppt
Electron Transport Chain lecture.pptElectron Transport Chain lecture.ppt
Electron Transport Chain lecture.ppt
 

ROS knuuti

  • 1. A single amino acid residue controls ROS production in the respiratory Complex I from Escherichia coli Juho Knuuti,1 Galina Belevich,1 Vivek Sharma,2 Dmitry A. Bloch1 * and Marina Verkhovskaya1 ** 1 Helsinki Bioenergetics Group, Institute of Biotechnology, University of Helsinki, PO Box 65 (Viikinkaari 1), FIN-00014, Helsinki, Finland. 2 Department of Physics, Tampere University of Technology, P. O. Box 692, FIN-33101, Tampere, Finland. Summary Reactive oxygen species (ROS) production by respira- tory Complex I from Escherichia coli was studied in bacterial membrane fragments and in the isolated and purified enzyme, either solubilized or incorporated in proteoliposomes. We found that the replacement of a single amino acid residue in close proximity to the nicotinamide adenine dinucleotide (NADH)-binding catalytic site (E95 in the NuoF subunit) dramatically increases the reactivity of Complex I towards dioxygen (O2). In the E95Q variant short-chain ubiquinones exhibit strong artificial one-electron reduction at the catalytic site, also leading to a stronger increase in ROS production. Two mechanisms can contribute to the observed kinetic effects: (a) a change in the reac- tivity of flavin mononucleotide (FMN) towards di- oxygen at the catalytic site, and (b) a change in the population of the ROS-generating state. We propose the existence of two (closed and open) states of the NAD+ -bound enzyme as one feature of the substrate- binding site of Complex I. The analysis of the kinetic model of ROS production allowed us to propose that the population of Complex I with reduced FMN is always low in the wild-type enzyme even at low ambient redox potentials, minimizing the rate of reac- tion with O2 in contrast to E95Q variant. Introduction Complex I [nicotinamide adenine dinucleotide (NADH):qui- none oxidoreductase, type I] catalyses electron input to the respiratory chain. It uses NADH as a two-electron donor and transfers these electrons to quinone (most often ubi- quinone) via flavin mono-nucleotide (FMN) and eight iron- sulphur (FeS) clusters located in the hydrophilic domain of the enzyme. Quinone reduction is coupled to translocation of three to four protons across the inner mitochondrial or bacterial cytoplasmic membrane (Wikström, 1984; Hinkle et al., 1991; Galkin et al., 1999; 2006; Wikström and Hummer, 2012). Complex I has also been suggested to be one of the major sites of formation of superoxide by mito- chondria (Turrens and Boveris, 1980; Galkin and Brandt, 2005; Kussmaul and Hirst, 2006), and has been therefore linked to some neurodegenerative diseases and ageing (Lin and Beal, 2006; Fukui and Moraes, 2008; Rhein et al., 2009). Indeed, it was reported that Complex I is the major site of superoxide formation in brain mitochondria under normal conditions (Barja and Herrero, 1998), and that it produces most of the reactive oxygen species (ROS) in Parkinson’s disease and ageing (for a review see Turrens, 2003), although the relevance of superoxide formation by Complex I under physiological conditions has been questioned (Zoccarato et al., 2004; Grivennikova and Vinogradov, 2006; Drechsel and Patel, 2010; Kareyeva et al., 2012). The site of this ‘electron leak’ from Complex I to O2 has been studied in different systems, ranging from intact mitochondria (Treberg et al., 2011) to the isolated enzyme (Kussmaul and Hirst, 2006; Ohnishi et al., 2010), there are no definite conclusions as yet. Here, we have studied ROS production by isolated and native membrane- bound Complex I from Escherichia coli, with the aid of site-specific mutations to alter vital electron transfer prop- erties. Complex I contains non-covalently bound FMN, which resides on the bottom of a cavity of the catalytic subunit and is, therefore, accessible to oxygen from the surroundings. FMN has a low midpoint redox potential (Sled et al., 1993; Euro et al., 2008b), and has been pro- posed to be the main site of O2 •− generation in mitochon- drial Complex I (Galkin and Brandt, 2005; Grivennikova and Vinogradov, 2006; Kussmaul and Hirst, 2006). However, the rates of ROS production by purified mito- chondrial [16–22 nmol mg−1 min−1 (Galkin and Brandt, 2005; Kussmaul and Hirst, 2006; Kareyeva et al., 2012)], and bacterial [25–50 nmol mg−1 min−1 (Esterhazy et al., 2008; this article)] Complex I, are very slow compared to the NADH oxidation rates. Furthermore, some other NADH-dependent flavin-containing enzymes, such as Accepted 3 October, 2013. For correspondence. *E-mail dmitry .bloch@helsinki.fi; Tel. (+358) 9 191 59754; Fax (+358) 9 191 59920; **E-mail marina.verkhovskaya@helsinki.fi; Tel. (+358) 9 191 59748; Fax (+358) 9 191 59920. Molecular Microbiology (2013) 90(6), 1190–1200 ■ doi:10.1111/mmi.12424 First published online 29 October 2013 © 2013 John Wiley & Sons Ltd
  • 2. NADH/NADPH oxidases belonging to the Nox family (e.g. MJ0649 from Methanocaldococcus jannaschii) produce superoxide at much higher rates (60 μmol min−1 mg−1 ), which is c. 60% of the NADH oxidation rate (Case et al., 2009). ROS production by mitochondria respiring on sub- strates that reduce matrix NAD+ is also moderate (for a review see Brand, 2010). This rate can be significantly increased by blocking the electron output with a specific Complex I inhibitor (e.g. Turrens and Boveris, 1980), but such treatment raises the NADH/NAD+ ratio making it possible that other NADH-dependent enzymes, such as pyruvate or α-ketoglutarate dehydrogenase, could partici- pate (Starkov et al., 2004). Under certain conditions mito- chondrial dihydrolipoamide dehydrogenase was found to be the major source of NADH-dependent ROS production (Kareyeva et al., 2012). In submitochondrial particles NADH-dependent O2 •− generation was found to be only slightly increased by rotenone (Vinogradov and Grivennikova, 2005; Grivennikova and Vinogradov, 2006), and the isolated enzyme showed no response (Galkin and Brandt, 2005; Kussmaul and Hirst, 2006; Drose et al., 2009). The first indication that reversed electron flow in mitochondria could be a source of ROS was obtained by Hinkle et al. in 1967 (Hinkle et al., 1967). Later, it was reported that succinate or glycerol-6-phosphate gives the highest rates of Complex I-derived ROS production in mitochondria under condi- tions of reversed electron transfer (see Brand, 2010 for a review). This phenomenon has been explained by the presence of two different O2 •− -generating nucleotide binding sites in Complex I, only one of which is active upon forward electron transport (Vinogradov and Grivennikova, 2005), but this was not supported by the resolved structure of Complex I (Sazanov and Hinchliffe, 2006). O2 •− genera- tion in the quinone binding site of Complex I was proposed independently by Brand and Treberg (Lambert and Brand, 2004a,b; Treberg et al., 2011), but this suggestion was disputed by Pryde and Hirst (Pryde and Hirst, 2011), who concluded from experiments with submitochondrial parti- cles that ROS production by Complex I is due to oxidation of FMN by O2 during both forward and reversed electron flow. At any rate, it is well established that the rates of gen- eration of O2 •− by Complex I in mitochondria, submito- chondrial particles, and purified Complex I are low relative to the much faster rates of NADH oxidation by quinone. It has been suggested that retaining NAD+ in the catalytic site could diminish the accessibility of oxygen to FMN (Kussmaul and Hirst, 2006), thus perhaps contributing to the slow rates of ROS production. The resolved structure of Complex I with the nucleotide bound [PBD entry 3IAS (Berrisford and Sazanov, 2009)] indeed suggests that NAD+ covers almost completely the part of the FMN alloxazine ring that protrudes into the cavity. However, it is difficult to predict to what extent such an effect might prevent oxygen to react with reduced FMN. The rate of ROS production by Complex I is expected to be determined by the properties of the nucleotide binding site and the bound FMN. The current study where a con- served glutamate residue in the NuoF subunit (E95) is mutated reveals a key element of the mechanism that prevents massive electron leakage from the catalytic site to dioxygen. Results Previously we have found that the replacement of the invariant Glu95 with glutamine in the NADH- and FMN- binding NuoF subunit results in moderate changes in nucleotide affinity, but a significant upshift in the midpoint redox potential of FMN (Euro et al., 2009a). The latter was explained as an electrostatic effect on FMN due to replac- ing the negatively charged carboxylate by the neutral amide, since the side-chain of Glu95 resides in the nucleo- tide binding cavity at a short distance from FMN. The structure of the hydrophilic domain of Thermus thermophi- lus Complex I (PDB entry 3IAM) reduced with NADH shows that the latter is bound between E97, a counterpart of E95 in E. coli enzyme, and FMN. The distance between E97 and the nicotinamide ring of bound NADH is 3.2 Å. On the other side E97 is located at the border of two subunits, Nqo1 (NuoF) and Nqo2 (NuoE). The latter contains the 2Fe-2S cluster N1a, which lays 10 Å from E97 with Pro98 in between (Fig. 1). An unexpected property of the E95Q variant of Complex I is a strong increase of the rate of NADH-dependent hydrogen peroxide production (Table 1). We compared ROS production of wild-type enzyme, the E95Q variant, and two other variants that lack quinone reductase activity. In NuoCD R274A the electron transfer to ubiquinone was blocked since the replacement of R274 by the neutral alanine residue resulted in a strong decrease of the mid- point redox potential of the last cluster in the intraprotein FeS chain, N2 (Belevich et al., 2007), to such an extent that it cannot be reduced by NADH. This cluster is the imme- diate electron donor to ubiquinone. In the other variant, NuoM E144A, the mutation was in the membrane subunit, NuoM, quite distant from the quinone binding site (Euro et al., 2008a). It is not yet known whether ubiquinone could be reduced in this mutant, but the properties of the N2 cluster were unchanged. The rates of ROS production in wild type and the NuoM E144A variant were rather similar, and slightly lower in NuoCD R274A, in stark contrast to the very high reactivity of the NuoF E95Q variant (Table 1). The small decrease of ROS production rate in NuoCD R274A can be explained by the lower content of Complex I in the sample: normalization by hexaammineruthenium (III) chloride (HAR)-reductase activity yields the same ratio of ROS production by Complex I 1191 © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 3. NADH-dependent HAR to oxygen reduction rate as for wild type. This finding excludes any significant role of the N2 centre, or ubiquinone bound near that centre, in NADH- dependent H2O2 generation by purified Complex I, since both redox centres are reduced to a much smaller extent in NuoCD R274A compared to wild type, yet with an insignifi- cant change of ROS production velocity. No significant difference in the rate of ROS production was observed between solubilized Complex I and enzyme reconstituted into liposomes (not shown). ROS production by Complex I can be stimulated by short-chain ubiquinones, insignificantly by decylubiqui- none (DQ) and strongly by coenzyme Q1 (Q1) and mena- dione (Table 2). This process is probably mediated by formation of the respective semiquinone which interacts with oxygen. It is known that short-chain ubiquinone ana- logues, such as Q1, can be reduced by Complex I from an artificial site resulting in O2 •− generation (Cadenas et al., 1977; Galkin and Brandt, 2005); the efficiency of this process is increased with decrease of the quinone hydro- phobicity (Cadenas et al., 1977). Most probably the site of this artificial reaction is FMN (King et al., 2009). Menadi- one is known to interact with flavin in many flavin- containing enzymes that yields intensive ROS formation (e.g. Matsuda et al., 2000). All in all, the data shown in Table 2 indicates that FMN is the only significant site of ROS production by Complex I under these conditions. The dependence of the rate of ROS production on the ambient redox potential calculated from the ratio [NADH]/ [NAD+ ] is shown in Fig. 2. The experimental curve, which is well reproducible and similar for wild type and the NuoCD R274A and NuoM E144A variants, is less steep and positively shifted (Fig. 2A) from a Nernstian n = 2 curve with Em = −350 mV, characteristic for FMN of E. coli Complex I (Euro et al., 2008b). These data are in very good agreement with that obtained on E. coli Complex I by Esterhazy et al. (2008). In the mitochondrial enzyme the rate of O2 •− formation is half-maximal at −359 mV (Kussmaul and Hirst, 2006). The titration of the rate of ROS formation by the NuoF E95Q variant was steeper Fig. 1. Disposition of E97 (counterpart of E95 in E. coli Complex I) in the catalytic site of Complex I from T. thermophilus (PDB entry: 3IAM). Table 1. ROS production by isolated Complex I does not correlate with its NADH oxidase or ubiquinone reductase activities. Variant Complex I activity, μmol mg−1 min−1 Rate of NADH- dependent ROS production, nmol mg−1 min−1 NADH:HAR oxidoreductase NADH:DQ oxidoreductase Wt 107.4 ± 7.0 26.9 ± 1.8 51.7 ± 3.5 NuoM E144A 109.7 ± 8.5 3.6 ± 0.5 55.6 ± 7.5 NuoCD R274A 72.4 ± 2.3 5.7 ± 0.3 44.1 ± 0.4 NuoF E95Q 24.4 ± 1.5 11.5 ± 1.5 749.4 ± 1.7 1192 J. Knuuti et al. ■ © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 4. than that in the WT enzyme, and showed a clear positive shift (Fig. 2B), which is in good agreement with the previ- ously determined midpoint potentials of the two-electron transition of FMN, −350 mV for wild type and −310 mV in this mutant (Euro et al., 2009a). Assuming that both the reduced and the half-reduced forms of FMN interact with oxygen, it may be suggested that the deviation from a two-electron Nernstian curve in the wild-type enzyme is due to a somewhat higher semiflavin radical stability than in the E95Q variant. Such an effect could explain the change of shape of the titration curves, but not the strong increase in the rate of ROS generation by the E95Q variant. The ubiquinone reductase activity of wild-type enzyme is inhibited by rolliniastatin, the specific Complex I inhibitor affecting the quinone binding site, by 90%, regardless of whether DQ or Q1 are the acceptors, whereas the activity of NuoF E95Q is inhibited by only 50% when the DQ is the electron acceptor and almost not at all with Q1 (Fig. 3). Since the NuoF E95 is quite distant (over 100 Å) from the quinone binding site and can hardly affect its structure directly, the observed relative insensitivity to rolliniastatin indicates that one half of the DQ and almost all of the Q1 reductase activity are due to quinone reduction at another artificial site, presumably FMN. An increased capability of electron exchange in the catalytic site of the NuoF E95Q variant is also revealed by a very different reactivity to dithionite. The wild-type enzyme is reduced slowly on a minutes timescale, whereas the reduction of NuoF E95Q occurs much faster (Fig. 4). The kinetics of optical changes at 450 nm upon reduction, where the FMN contribution to the Complex I spectrum is maximal, is shown in Fig. 4. The half times are approximately 3 min and 10 s, for wild type and NuoF E95Q respectively. Thus the acceleration of enzyme reduction by the mutation is of the same order as the acceleration of hydrogen peroxide production. On the basis of our results we may conclude that the catalytic site of Complex I becomes leaky with respect to reactivity with O2 due to loss of the negative charge of E95, although the structure of the cavity containing the FMN and the nucleotide binding site was not significantly perturbed, since the affinity for NADH was only slightly changed (Euro et al., 2009a). The fact that the rate of reduction of Complex I by dithionite was considerably accelerated in the NuoF E95Q Table 2. ROS production by isolated Complex I induced by short-chain ubiquinones. Sample ROS productiona – DQ Q1 Menadione Purified wt 1.00 ± 0.02 1.16 ± 0.01 9.28 ± 0.28 12.10 ± 0.31 Purified E95Q 15.98 ± 0.46 22.78 ± 0.04 95.46 ± 2.25 144.87 ± 1.17 Membranes wt 1.00 ± 0.14 1.07 ± 0.09 3.93 ± 0.02 5.55 ± 0.02 Membranes E95Qb 6.04 ± 0.90 5.98 ± 0.26 40.17 ± 1.47 59.07 ± 2.25 a. The initial rate of H2O2 production is normalized to that of wild type in the absence of added quinones, which are 51.7 ± 3.5 nmol mg−1 min−1 for purified enzyme and 2.49 ± 0.34 nmol mg−1 min−1 for native membrane-bound Complex I respectively. b. Normalized to estimated 1.5 times lower content of mutated Complex I in native membranes. Estimation is based on the ratio of differences between the specific quinone reductase activities of the membranes and purified proteins. Fig. 2. Redox titration of the rate of H2O2 formation by purified Complex I. The normalized rate is plotted against the apparent ambient redox potential calculated from the ratio [NADH]/[NAD+ ]. A. Titration of wild-type Complex I, circles, and mutants lacking ubiquinone reductase activity, NuoCD R274A, squares, and NuoM E144A, triangles. B. Comparison of the redox titration of wild-type Complex I, circles, and NuoF E95Q, squares. Note that the absolute values for the mutant are 15 times larger in comparison with wild type. ROS production by Complex I 1193 © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 5. variant is consistent with the rate-limiting step being the approach of the dithionite oxoanion to the active site. In the wild-type enzyme this reaction is apparently limited by the negative charge of the carboxylate group of E95. However, this cannot explain the higher reactivity of the E95Q variant with neutral short-chain ubiquinones. Discussion Here we report the ∼15-fold enhancement in the steady- state rate of ROS production (vROS ) by the E95Q variant of E. coli Complex I at low ambient redox potentials (maintained by the NADH/NAD+ pair) as the result of the increased electron leakage from the catalytic site. It comes from the X-ray crystallography data that ROS formation is very unlikely to occur at any of the FeS clusters due to their shielding by the protein matrix; only FMN is partially exposed to the solvent (PDB entry: 3IAS). We also assume that the ubiquinone binding site does not participate in ROS production, at least under the given experimental conditions. Thus, our observations can only be accounted for by changes in the properties of the active site caused by the mutation. In Complex I the flavin cofactor is partially buried at the bottom of a deep cavity forming the catalytic site (Sazanov and Hinchliffe, 2006). The side-chain of Glu95/ Glu97 (E. coli numbering, subunit NuoF, or T. thermophi- lus numbering, subunit Nqo1) is located at the same depth as flavin and around 6 Å (N5FMN-OE1E97) away from the flavin π-system. The carboxylate and the flavin form a binding pocket for the nucleotide, as shown in Fig. 1 (PDB entry: 3IAM). It follows from the structure (see Supporting Fig. 3. Relative insensitivity of ubiquinone reductase activity of E95Q to rolliniastatin. NADH oxidation was monitored in the presence of DQ (A, B) and Q1 (C, D) by purified wild-type (A, C) and E95Q (B, D) Complexes I. NADH addition is indicated by arrow. Rolliniastatin, 1 μM, was added at zero time as indicated. Fig. 4. Kinetics of wild-type and NuoF E95Q Complexes I reduction by dithionite. Set of redox spectra upon the reduction is shown for wild type and E95Q variant. Kinetics of optical changes at 450 nm is characterized by τ1 2 ≈ 3 min for wild type (circles) and τ1 2 ≈ 10 s for E95Q (squares). 1194 J. Knuuti et al. ■ © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 6. information, section 1) as well as from the effect of the E95Q mutation on the Em of FMN (Euro et al., 2009a) that the Glu95 side-chain is most probably in the anionic state. In the nucleotide-bound state, the H-bonding between the Glu95 and NAD+ is very unlikely; however, the electric field of its charged carboxylate group is plausibly to affect both redox potentials of the bound nucleotide and the flavin, despite the polar nature of the solvent-exposed nucleotide-binding cavity. Here we do not consider oxygen interaction with the FMN radical, despite the fact that the latter is known as a powerful ROS generator in many chemical or enzymatic systems (see, e.g. Massey, 1994; 1995; Massey and Ghisla, 1974; Mattevi, 2006). NADH is a strict two- electron donor so that any FMN radical would derive from subsequent oxidation of the FMNH− state. Thus the popu- lation of the FMN radical state will be dependent on the population of the FMNH− state. Moreover, the participation of the two-electron-reduced FMN in ROS production by mitochondrial Complex I was shown experimentally (Kussmaul and Hirst, 2006). The mechanism of O2 interaction with the reduced flavin is not known in Complex I. However, it is generally accepted that the primary step in ROS generation in other flavoproteins and chemical systems is the rate-limiting spin conversion and charge transfer between FMNH− and O2 resulting in the ‘caged’ radical pair formation followed by a rebinding of O2 •− at the C4a atom of the flavin with the formation of a flavin hydroperoxide species, which can further produce H2O2 (Massey, 1994; Mattevi, 2006). Pro- duction of both O2 •− and H2O2 was shown experimentally in Complex I, although the ratio between these two species varied and was also reported to be different in mitochon- drial and bacterial enzymes (Esterhazy et al., 2008). Hence, in this study, we do not discriminate between O2 •− and H2O2 as both are considered ROS, the focus being concentrated on the nature of the electron leakage from the catalytic site. Since ROS production is considered virtually irrevers- ible, based on the very high quenching efficiency of the detection system, its observed rate in Complex I at steady state can be quantitatively expressed as v kROS ROS FMNH= × − [ ] (1) where kROS is the rate constant of the reaction and [FMNH− ] is the population of the reduced flavin. Given the measured redox potentials of FMN in the nucleotide-free WT and mutant enzymes [−350 mV and −310 mV respec- tively (Euro et al., 2009a)], FMN should stay almost fully reduced in both variants at Eh < −380 mV (ambient redox potentials, typically provided by NADH). If it was indeed the case, the E95Q mutant should have revealed much higher reactivity of the reduced FMN towards dioxygen (that is, a higher value of kROS , e.g. due to the increased accessibility of dioxygen to the reduced FMN) in order to ensure the observed change in vROS . In Complex I direct interaction of O2 with FMN is blocked at the flavin distal (opposite to the substrate-binding) side due to the steric hindrance caused by the peptide back- bone and side-chains of Gly183, Glu184 and Asn92 (T. thermophilus numbering, PDB entry: 3IAM); the flanks of the flavin are also protected by the protein, except for its redox-inactive dimethylbenzene ring and the N5 atom. The latter was suggested as the site of O2 •− formation in model flavoproteins (McDonald et al., 2011). In Complex I it is exposed to the cavity large enough to hold one or two H2O molecules and possibly O2 or O2 •− ; however, the cavity itself is isolated from the rest of the nucleotide-binding pocket by Ser96, Glu97 and Tyr180 peptide backbone, Ser96 and Asp94 side-chains (T. thermophilus numbering, PDB entry: 3IAM), and by the C3 amide group of bound nicoti- namide, preventing both O2 binding to, and O2 •− escape from the flavin N5 atom. Hence O2 can approach FMN only from the proximal (NAD+ -binding) side, and it seems that the presence of either NADH or NAD+ bound to FMN prohibits the access of O2. Therefore, ROS production is likely to occur either in the nucleotide-free enzyme (i.e. after NAD+ dissociation) or in the situation, where NAD+ is present in the binding pocket but does not form the π–π stacked complex with FMNH− (usually referred to as charge–transfer complex, CTC, in many flavoproteins), thereby allowing O2 to reach the proximal side of the flavin plain. In the latter case, two conformational sub-states for the NAD+ -bound enzyme could be proposed: a ‘closed’ state [(NAD+ -FR )closed ] for the NAD+ -FR CTC and an ‘open’ state [(NAD+ -FR )open ], where such complex is not formed. The latter possibility has indeed been discussed in the literature for various flavoproteins (Gassner et al., 1996; Palfey et al., 2001; Palfey and McDonald, 2010) and is further supported by the observation that in the X-ray crystal structure of Complex I (Berrisford and Sazanov, 2009) the planes of the flavin and nicotinamide rings are not strictly parallel, indicating a possible mixture of several sub-states. However, the E95Q replacement can hardly affect the O2 accessibility to FMN in either case: in the nucleotide-free catalytic site the distance between FMN and Glu95 is large enough to accommodate an O2 mol- ecule (> 5 Å); in (NAD+ -FR )open state the negatively charged Glu95 can only attract NAD+ [in comparison with (NAD+ - FR )closed state] providing more access of O2 to FMN and more ROS production in WT. The E95Q replacement could also affect the kROS value (Eq. 1) if NAD+ dissociation rate ( )koff NAD+ became higher for E95Q with respect to WT (in other words, NAD+ was retained longer in the binding site of WT, attracted by Glu95 and preventing the O2 access to FMN). The decreased Km NADH , as well as the decreased Ki for NAD+ and ADP-ribose in the mutant with respect to wild type ROS production by Complex I 1195 © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 7. (Euro et al., 2009a) may argue in favour of this sugges- tion. However, the changes in apparent affinities upon the mutation [2 times for Km NADH , 10 μM in WT and 5 μM in E95Q, and 2.4 times for Ki NAD+ , 1 mM in WT and 0.39 mM in E95Q respectively (Euro et al., 2009a)] are not large enough to provide 15 times stronger binding of O2 in the mutant. Another possibility for the change in kROS is the mutation- specific change of the redox potential of the FMN- bound dioxygen (see Supporting information, section 2). However, the estimated effect ranging from 0 to −15 mV, depending on the position of the bound O2, is too small to account for the observed effect in full. Therefore, we conclude that the change in reactivity of the reduced flavin with respect to O2 can hardly explain the observed mutation effects. Due to this we consider the thermodynamic aspect of the reaction of Complex I with oxygen. From the experimentally measured effect of the E95Q mutation on the Em of FMN in the absence of nucleotides [ΔEm = +40 mV (Euro et al., 2009a)], the esti- mated value of the dielectric constant of the cavity (ε ≈ 53, see Eq. S1), and the structural difference between the nucleotide-free and nucleotide-bound enzymes, the inter- action potential between Glu95 and FMNH− was found, δϕ1 = −35 meV (see Supporting information, section 2). On the other hand, the interaction between the negatively charged Glu95 and the positively charged, oxidized nico- tinamide moiety of bound NAD+ is, by absolute value, much stronger, δϕ2 = +59 meV. The sum δϕ1 + δϕ2 = +24 meV (+0.6 kcal mol−1 ) indicates that in the WT enzyme the carboxylate group of Glu95 stabilizes FMNH− in the nucleotide-bound state (NAD+ -FR ) relative to the same state in the mutant. Similar conclusion was obtained from the density functional theory (DFT) calculations using a small model system comprising only of FMN and NAD+ or NADH in the CTC configuration and the Glu95 residue (see Supporting information, section 3, Tables S1 and S2, and Fig. S1). Based on the electrostatic interaction ener- gies calculated from the partial charges derived from DFT calculations, the (NAD+ -FR )closed state is found more stable in WT by 1.1 kcal mol−1 (at ε = 50). These results predict the opposite effect of the mutation on ROS production. Relative destabilization of the NAD+ -FR state in the mutant also means higher dissociation constant for NAD+ : a change in 1.1 kcal mol−1 would raise KD NAD+ in the mutant for almost one order of magnitude. Keeping in mind the ∼ 40 meV stabilization of the nucleotide-free state FR in the mutant, the change in KD NAD+ would be even larger, which is very unlikely, given the Ki NAD+ data (see above). Therefore, the other explanation of the observed phenom- enon should be suggested. Finally, we analysed how the mutation affects the steady-state probability, or life time, of transient enzyme state(s) competent in ROS production ([FMNH− ] in Eq. 1). The reaction scheme is shown in Fig. 5. After FMN reduc- tion (step 3) there are two alternatives. In Cycle A, NAD+ dissociation (step 4) is followed by the O2 reaction with state FR (step 6). Note that Cycle A is realized only in the situation when the FeS chain is reduced to a great extent, so that the electron transfer from FMNH− to the FeS chain is no longer possible [step 5, kET2 0= , see for a review (Verkhovskaya and Bloch, 2013)]. The other alternative is Cycle B, where O2 reacts with the state NAD+ -FR (step 7). If NAD+ dissociation from FMNH− is significantly slower than ROS production ( )k koff NAD ROS+ << , then the contribu- tion of Cycle A is negligible and Cycle B becomes pre- dominant. Unlike Cycle A, Cycle B is independent of the reduction level of the FeS chain. Although in various NAD(P)H-dependent flavoenzymes NAD+ can leave the binding site either before or after the oxidation of FMN to Fig. 5. Kinetic scheme of the steady-state ROS production by Complex I. F and FR stand for the nucleotide-free enzyme with oxidized and two-electron-reduced FMN respectively; NAD+ -F and NAD+ -FR are the enzyme with bound NAD+ ; NADH-F is the enzyme with bound NADH. Both ‘closed’ and ‘open’ states of NAD+ -FR are shown. Redox state of the FeS cluster chain is not shown for simplicity. Highlighted are the ROS-producing states (NAD+ -FR )open and FR . Reaction steps are: NAD+ dissociation from oxidized enzyme (1), NADH binding to oxidized enzyme (2), electron transfer from NADH to FMN (presumably through the hydride transfer within the charge–transfer complex, 3), NAD+ dissociation from the enzyme with reduced FMN (4), electron transfer from reduced FMN to the FeS cluster chain (5), interaction of reduced FMN with dioxygen and ROS production (6, 7), respectively, as indicated by numbers. All reaction steps (1–5), except for the ROS production (6, 7), are thermodynamically reversible; arrows show the preferred fluxes under typical conditions. Dotted line shows the pre-steady-state reduction of the enzyme-bound FeS cluster chain by FR . Dashed lines show the irreversible two-electron reduction of dioxygen. 1196 J. Knuuti et al. ■ © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 8. a secondary substrate, there are examples in the litera- ture where flavin remains reduced for a considerable time (e.g. when the electron acceptors are not available) and NAD+ stays attached to it as long as the flavin remains reduced (Gassner et al., 1996), which should be due to the electrostatic interactions between NAD+ and FMNH− . Seemingly contradictory with the latter proposal, the experimentally measured, relatively high rate constant of NAD+ dissociation [ koff NAD s + ≈ − 103 1 (Verkhovskaya et al., 2008)] was in fact determined under the conditions where the FeS chain and FMN were not fully reduced; thus, two different koff NAD+ values should be considered, depending on the reduction state of the flavin (Fig. 5, steps 1 and 4 respectively). Noteworthy, Cycle B can be realized only when the steady-state electron flux from NAD+ -FR to O2 takes place. In the absence of O2, the reaction will inevitably proceed to the nucleotide-free, fully reduced state (FR ), despite its accumulation is slow. To quantitatively analyse the yield of the FMNH− states, the analytical solution was obtained for each cycle sepa- rately (Cycle A, steps 2, 3, 4, and Cycle B, steps 1, 2, 3) at equilibrium and without the electron leakage to O2 (kROS = 0) (Fig. 6; see also: Supporting information, section 4, Eqs S2–S4 and Figs S2 and S3). Cycle A does not show any difference in the population of ROS- production state (FR ) between WT and E95Q at low Eh (< −380 mV). The equilibrium solution for Cycle B predicts that the flavin can be reduced only to a certain, small extent in the NADH-reduced enzyme even at low poten- tials. Despite the enzyme is considered in redox equilib- rium with NADH/NAD+ (given that the absolute ROS production rate in Complex I, 0.5–1.0 s−1 in WT, is orders of magnitude lower than the enzyme turnover rate, kcat > 300 s−1 ), in the absence of redox mediators the sequence of the steps restricts the yield of FMNH− : NADH binding (Fig. 5, step 2) is followed by the electron redis- tribution between NADH and FMN (step 3) according to their Em; at the same time, the preceding NAD+ dissocia- tion (step 1) does not contribute to repopulate FMNH− . When the electron transfer from NADH to FMN is uphill ( )/ ΔE E Em m FMNH /FMN m NADH NAD = − < − + 0 , then [NAD+ -FR ] is low; when it is downhill (ΔEm > 0), [NAD+ -FR ] is higher. Assum- ing Em NADH NAD mV/ + = −320 and the experimentally deter- mined Em values for FMN, −350 (WT) and −310 (E95Q), the former situation is realized in WT and the latter, in the Fig. 6. Simulated Eh-dependences of the equilibrium concentrations of intermediates in Cycles A and B. The parameters used are: Ntot = [NADH] + [NAD+ ] = 10−4 M, KD NADH M= − 10 5 , KD NAD M + = − 10 5 , Em N mV= −320 , Em F mV= −350 (WT), Em F mV= −310 (E95Q). The ambient redox potential is set by [NADH] and [NAD+ ]: E Eh m N NADH NAD = − × + 30 lg [ ] [ ] . All redox potentials quoted, refer to the NHE scale. For the respective kinetic schemes, see Fig. 5. For further details, see Supporting information, section 4. Note that for Cycle B, the parameter set differs from that found by the data fitting (see Fig. S5 and Table S3). The parameter set used here is the one showing the difference between the two cycles the most (see Supporting information, section 4). ROS production by Complex I 1197 © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 9. mutant (Fig. 6, lower panel). By varying the KD values for WT and E95Q, it was possible to obtain a good fit of the vROS data, given both experimental effects of the mutation: the 15-fold increase in vROS at low Eh and the ∼ 20 mV upshift of the apparent midpoint potential (E 1 2) of the vROS Eh-dependence (see Supporting information, sections 4 and 5, Eqs S5–S10, Figs S4 and S5, and Table S3). As shown by the electrostatic and DFT calculations (see above), the NAD+ -FR state is relatively stable in WT as compared with the mutant, which apparently contradicts both the observations and the kinetic modelling. However, the uncertainty in the position of nicotinamide in the ROS- producing (NAD+ -FR )open state might affect the result of such calculations. Since the nucleotide-binding cavity is wide enough to accommodate NAD+ in a different position as compared with the resolved structure (see Supporting information, section 1), the electrostatic interaction between Glu95, FMNH− and NAD+ may vary strongly. The lack of the expected mutation-caused change in the NAD+ affinity (Ki NAD+ , see above) further supports this possibility. We note that the equilibrium solution of Cycle B (Fig. 6, lower panel) predicts a significant amount of the nucleotide-bound enzyme state(s) at low Eh, which are known to form CTC in many flavoproteins. However, CTC can be difficult to detect due to strong variation in both extinction coefficient and maximum wavelength; more- over, both NADH-F and NAD+ -FR CTC are often found to be transient, short-lived states (see Massey and Ghisla, 1974; Gao and Ellis, 2007; Lin et al., 2012; and refer- ences therein). No information about optical properties of CTC in Complex I has been available yet. We were not able to detect any signature of CTC in E. coli Complex I under the ROS-producing conditions, either (data not shown). Since in Complex I, the proposed ROS-producing state (NAD+ -FR )open is not a CTC, it should not have an absorption signature. On the other hand, the maximum yield of [NAD+ -FR ] can be rather low in both WT and the mutant in the NADH-non-saturating conditions and depend on KD NADH (see Fig. S4 and Eq. S9). An important implication of this analysis is the finding that even low electron leakage from NAD+ -FR results in returning the enzyme to the non-ROS-producing state (NADH-F), which accumulates. Therefore, the population of the ROS producing states is always low in the WT enzyme. We conclude that Complex I possesses an inbuilt safety mechanism that prevents excessive reduc- tion of FMN and consequently minimizes ROS production at low ambient redox potentials regardless of the avail- ability of the natural substrate, ubiquinone. This mecha- nism is based on the stabilization of an intermediate state (non-reactive with O2) of partially reduced enzyme, where the FeS chain is either oxidized or reduced, but only to a certain extent, and slow NAD+ release from the state with reduced flavin. Experimental procedures Escherichia coli strains MWC215 (Δnhd) (Calhoun and Gennis, 1993), GR70N (Green et al., 1984) (both have wild- type Complex I) and variants NuoM E144A (Euro et al., 2008a), NuoF E95Q (Euro et al., 2009a) and NuoCD R274A (Belevich et al., 2007) were grown aerobically in LB medium supplemented with appropriate antibiotics on demand at 37°C. Inverted cytoplasmic membrane vesicles were pre- pared after cell wall removal by sonication as described pre- viously (Belevich et al., 2007), and used the same day for analysis. Enzyme extraction and further purification from iso- lated membranes was done as previously described (Euro et al., 2009b). The HAR and DQ reductase activities of the membrane-bound and purified Complex I were measured as in Belevich et al. (2011). Reconstitution of Complex I into liposomes was performed as described in Verkhovskaya et al. (2011). Protein concentration in the membrane prepa- rations was determined by the BCA protein assay kit (Pierce) using BSA as a standard. Absorbance at 280 nm was used to quantify the purified protein (ε280 = 785160 M−1 cm−1 ). Complex I mediated NADH-dependent H2O2 production was quantified by using 0.4 units ml−1 of horse-radish per- oxidase (Sigma) catalysing oxidation of 10 μM Amplex Red (Sigma) to resorufin (ε557–620 = 51.6 mM−1 cm−1 ) at 30°C. NADH was used for purified Complex I and nicotinamide hypoxanthine dinucleotide (dNADH) for native membrane vesicles of GR70N and its variants NuoF E95Q, NuoCD R274A and NuoM E144A to avoid contribution of NADH:ubiq- uinone reductase type II, which was completely removed after purification. The MWC215 strain is a GR70N derivative lacking the gene encoding this alternative NADH dehydroge- nase; it was grown for high-yield wild-type enzyme purifica- tion. H2O2 production was measured in 200 mM HEPES-KOH pH 7.0 (RT) or 20 mM HEPES-KOH containing 360 mM man- nitol pH 7.0 (RT). Initial ROS production rate was same in both buffers, but the latter was chosen for the measurements of Complex I proteoliposomes to preserve electrogenic prop- erties of the enzyme. 0.005% n-dodecyl β-D-maltoside (DDM) was included when solubilized enzyme was used, at con- centrations of 0.6 μg ml-1 (1.0 nM) for E95Q variant and 6.1–7.8 μg ml−1 (11–15 nM) for other variants and WT respec- tively. Media was supplemented with 1 μM rolliniastatin when ROS production by membrane preparations was defined, at total protein concentration of 45 μg ml−1 for E95Q variant and 90 μg ml-1 for other variants and WT respectively. H2O2 for- mation by purified Complex I was titrated using the NADH:NAD+ redox couple to set the ambient potential. The nucleotides, NADH or dNADH and NAD+ were added to a total concentration of 35 μM or (d)NADH concentration was kept constant and NAD+ concentration was varied. The reac- tion was initiated by Complex I addition and the rate, which appeared non-linear, was quantified right after the enzyme addition allowing short mixing and defining the maximal rate as velocity without NAD+ addition. Background of H2O2 gen- eration was determined in the presence of 525 units ml−1 catalase (Sigma). Kinetic measurements were carried out using a high- resolution CCD-array spectrometer (HR2000+, Ocean Optics). The assay buffer comprised 50 mM HEPES-KOH, pH 7.0, 15% glycerol and 0.005% DDM. Complex I was added at a final concentration of 2.7–2.9 μM. 1198 J. Knuuti et al. ■ © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 10. Numerical simulations and the data fit were performed using Matlab R2010 (MathWorks, Natick, MA). The data fit was accomplished using a non-linear, constrained solver ‘fmincon’. A full description of the kinetic models is given in Supporting information. The details of DFT calculations are also described in Supporting information (section 3). Acknowledgements This work was supported by grants from Biocentrum Helsinki, the Sigrid Jusélius Foundation and the Academy of Finland. V.S. acknowledges postdoctoral researcher funding from the Academy of Finland. We thank Prof. Mårten Wikström for extensive discussion and his interest to this work. We thank Eija Haasanen for purification of Complex I. We are also thankful to the Center for Scientific Computing (CSC), Finland, for providing computing time. References Barja, G., and Herrero, A. (1998) Localization at complex I and mechanism of the higher free radical production of brain nonsynaptic mitochondria in the short-lived rat than in the longevous pigeon. J Bioenerg Biomembr 30: 235– 243. Belevich, G., Euro, L., Wikström, M., and Verkhovskaya, M. (2007) Role of the conserved arginine 274 and histidine 224 and 228 residues in the NuoCD subunit of complex I from Escherichia coli. Biochemistry 46: 526–533. Belevich, G., Knuuti, J., Verkhovsky, M.I., Wikström, M., and Verkhovskaya, M. (2011) Probing the mechanistic role of the long alpha-helix in subunit L of respiratory Complex I from Escherichia coli by site-directed mutagenesis. Mol Microbiol 82: 1086–1095. Berrisford, J.M., and Sazanov, L.A. (2009) Structural basis for the mechanism of respiratory complex I. J Biol Chem 284: 29773–29783. Brand, M.D. (2010) The sites and topology of mitochondrial superoxide production. Exp Gerontol 45: 466–472. Cadenas, E., Boveris, A., Ragan, C.I., and Stoppani, A.O.M. (1977) Production of Superoxide radicals and hydrogen- peroxide by NADH-ubiquinone reductase and ubiquinol- cytochrome C reductase from beef-heart mitochondria. Arch Biochem Biophys 180: 248–257. Calhoun, M.W., and Gennis, R.B. (1993) Demonstration of separate genetic loci encoding distinct membrane-bound respiratory NADH dehydrogenases in Escherichia coli. J Bacteriol 175: 3013–3019. Case, C.L., Rodriguez, J.R., and Mukhopadhyay, B. (2009) Characterization of an NADH oxidase of the flavin- dependent disulfide reductase family from Methanocaldo- coccus jannaschii. Microbiology 155: 69–79. Drechsel, D.A., and Patel, M. (2010) Respiration-dependent H2O2 removal in brain mitochondria via the thioredoxin/ peroxiredoxin system. J Biol Chem 285: 27850–27858. Drose, S., Galkin, A., and Brandt, U. (2009) Measurement of superoxide formation by mitochondrial complex I of Yar- rowia lipolytica. Methods Enzymol 456: 475–490. Esterhazy, D., King, M.S., Yakovlev, G., and Hirst, J. (2008) Production of reactive oxygen species by complex I (NAD- H:ubiquinone oxidoreductase) from Escherichia coli and comparison to the enzyme from mitochondria. Biochemis- try 47: 3964–3971. Euro, L., Belevich, G., Verkhovsky, M.I., Wikström, M., and Verkhovskaya, M. (2008a) Conserved lysine residues of the membrane subunit NuoM are involved in energy con- version by the proton-pumping NADH:ubiquinone oxidore- ductase (Complex I). Biochim Biophys Acta 1777: 1166– 1172. Euro, L., Bloch, D.A., Wikström, M., Verkhovsky, M.I., and Verkhovskaya, M. (2008b) Electrostatic interactions between FeS clusters in NADH:ubiquinone oxidoreductase (Complex I) from Escherichia coli. Biochemistry 47: 3185– 3193. Euro, L., Belevich, G., Bloch, D.A., Verkhovsky, M.I., Wikström, M., and Verkhovskaya, M. (2009a) The role of the invariant glutamate 95 in the catalytic site of Complex I from Escherichia coli. Biochim Biophys Acta 1787: 68– 73. Euro, L., Belevich, G., Wikström, M., and Verkhovskaya, M. (2009b) High affinity cation-binding sites in Complex I from Escherichia coli. Biochim Biophys Acta 1787: 1024–1028. Fukui, H., and Moraes, C.T. (2008) The mitochondrial impair- ment, oxidative stress and neurodegeneration connection: reality or just an attractive hypothesis? Trends Neurosci 31: 251–256. Galkin, A., and Brandt, U. (2005) Superoxide radical forma- tion by pure complex I (NADH:ubiquinone oxidoreductase) from Yarrowia lipolytica. J Biol Chem 280: 30129–30135. Galkin, A., Dröse, S., and Brandt, U. (2006) The proton pumping stoichiometry of purified mitochondrial complex I reconstituted into proteoliposomes. Biochim Biophys Acta 1757: 1575–1581. Galkin, A.S., Grivennikova, V.G., and Vinogradov, A.D. (1999) -->H+/2e- stoichiometry in NADH-quinone reduc- tase reactions catalyzed by bovine heart submitochondrial particles. FEBS Lett 451: 157–161. Gao, B., and Ellis, H.R. (2007) Mechanism of flavin reduction in the alkanesulfonate monooxygenase system. Biochim Biophys Acta 1774: 359–367. Gassner, G.T., Johnson, D.A., Liu, H.W., and Ballou, D.P. (1996) Kinetics of the reductive half-reaction of the iron- sulfur flavoenzyme CDP-6-deoxy-L-threo-D-glycero-4- hexulose-3-dehydrase reductase. Biochemistry 35: 7752– 7761. Green, G.N., Kranz, R.G., Lorence, R.M., and Gennis, R.B. (1984) Identification of subunit I as the cytochrome b558 component of the cytochrome d terminal oxidase complex of Escherichia coli. J Biol Chem 259: 7994–7997. Grivennikova, V.G., and Vinogradov, A.D. (2006) Generation of superoxide by the mitochondrial Complex I. Biochim Biophys Acta 1757: 553–561. Hinkle, P.C., Penefsky, H.S., and Racker, E. (1967) Partial resolution of enzymes catalyzing oxidative phosphorylation. 12. H2 18 O-inorganic phosphate and H2 18 O-adenosine triphosphate exchange reactions in submitochondrial parti- cles from beef heart. J Biol Chem 242: 1788–1792. Hinkle, P.C., Kumar, M.A., Resetar, A., and Harris, D.L. (1991) Mechanistic stoichiometry of mitochondrial oxida- tive phosphorylation. Biochemistry 30: 3576–3582. Kareyeva, A.V., Grivennikova, V.G., and Vinogradov, A.D. (2012) Mitochondrial hydrogen peroxide production as ROS production by Complex I 1199 © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200
  • 11. determined by the pyridine nucleotide pool and its redox state. Biochim Biophys Acta 1817: 1879–1885. King, M.S., Sharpley, M.S., and Hirst, J. (2009) Reduction of hydrophilic ubiquinones by the flavin in mitochondrial NAD- H:ubiquinone oxidoreductase (Complex I) and production of reactive oxygen species. Biochemistry 48: 2053–2062. Kussmaul, L., and Hirst, J. (2006) The mechanism of super- oxide production by NADH:ubiquinone oxidoreductase (complex I) from bovine heart mitochondria. Proc Natl Acad Sci USA 103: 7607–7612. Lambert, A.J., and Brand, M.D. (2004a) Inhibitors of the quinone-binding site allow rapid superoxide production from mitochondrial NADH:ubiquinone oxidoreductase (complex I). J Biol Chem 279: 39414–39420. Lambert, A.J., and Brand, M.D. (2004b) Superoxide produc- tion by NADH:ubiquinone oxidoreductase (complex I) depends on the pH gradient across the mitochondrial inner membrane. Biochemistry 382: 511–517. Lin, M.T., and Beal, M.F. (2006) Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 443: 787–795. Lin, T.Y., Werther, T., Jeoung, J.H., and Dobbek, H. (2012) Suppression of electron transfer to dioxygen by charge transfer and electron transfer complexes in the FAD- dependent reductase component of toluene dioxygenase. J Biol Chem 287: 38338–38346. McDonald, C.A., Fagan, R.L., Collard, F., Monnier, V.M., and Palfey, B.A. (2011) Oxygen reactivity in flavoenzymes: context matters. J Am Chem Soc 133: 16809–16811. Massey, V. (1994) Activation of molecular oxygen by flavins and flavoproteins. J Biol Chem 269: 22459–22462. Massey, V. (1995) Introduction: flavoprotein structure and mechanism. FASEB J 9: 473–475. Massey, V., and Ghisla, S. (1974) Role of charge–transfer interactions in flavoprotein catalysis. Ann N Y Acad Sci 227: 446–465. Matsuda, H., Kimura, S., and Iyanagi, T. (2000) One-electron reduction of quinones by the neuronal nitric-oxide synthase reductase domain. Biochim Biophys Acta 1459: 106–116. Mattevi, A. (2006) To be or not to be an oxidase: challenging the oxygen reactivity of flavoenzymes. Trends Biochem Sci 31: 276–283. Ohnishi, S.T., Shinzawa-Itoh, K., Ohta, K., Yoshikawa, S., and Ohnishi, T. (2010) New insights into the superoxide generation sites in bovine heart NADH-ubiquinone oxidore- ductase (Complex I): the significance of protein-associated ubiquinone and the dynamic shifting of generation sites between semiflavin and semiquinone radicals. Biochim Biophys Acta 1797: 1901–1909. Palfey, B.A., and McDonald, C.A. (2010) Control of catalysis in flavin-dependent monooxygenases. Arch Biochem Biophys 493: 26–36. Palfey, B.A., Bjornberg, O., and Jensen, K.F. (2001) Insight into the chemistry of flavin reduction and oxidation in Escherichia coli dihydroorotate dehydrogenase obtained by rapid reaction studies. Biochemistry 40: 4381–4390. Pryde, K.R., and Hirst, J. (2011) Superoxide is produced by the reduced flavin in mitochondrial complex I a single, unified mechanism that applies during both forward and reverse electron transfer. J Biol Chem 286: 18056–18065. Rhein, V., Baysang, G., Rao, S., Meier, F., Bonert, A., Muller-Spahn, F., and Eckert, A. (2009) Amyloid-beta leads to impaired cellular respiration, energy production and mitochondrial electron chain complex activities in human neuroblastoma cells. Cell Mol Neurobiol 29: 1063–1071. Sazanov, L.A., and Hinchliffe, P. (2006) Structure of the hydrophilic domain of respiratory complex I from Thermus thermophilus. Science 311: 1430–1436. Sled, V., Rudnitzky, N., Hatefi, Y., and Ohnishi, T. (1993) Thermodynamic analysis of FMN in mitochondrial NADH ubiquinone oxidoreductase (Complex-I). Biophys J 64: A105–A105. Starkov, A.A., Fiskum, G., Chinopoulos, C., Lorenzo, B.J., Browne, S.E., Patel, M.S., and Beal, M.F. (2004) Mitochon- drial alpha-ketoglutarate dehydrogenase complex gener- ates reactive oxygen species. J Neurosci 24: 7779–7788. Treberg, J.R., Quinlan, C.L., and Brand, M.D. (2011) Evi- dence for two sites of superoxide production by mitochon- drial NADH-ubiquinone oxidoreductase (Complex I). J Biol Chem 286: 27103–27110. Turrens, J.F. (2003) Mitochondrial formation of reactive oxygen species. J Physiol 552: 335–344. Turrens, J.F., and Boveris, A. (1980) Generation of superox- ide anion by the NADH dehydrogenase of bovine heart- mitochondria. Biochem J 191: 421–427. Verkhovskaya, M., and Bloch, D.A. (2013) Energy-converting respiratory Complex I: on the way to the molecular mecha- nism of the proton pump. Int J Biochem Cell Biol 45: 491–511. Verkhovskaya, M., Knuuti, J., and Wikström, M. (2011) Role of Ca2+ in structure and function of Complex I from Escheri- chia coli. Biochim Biophys Acta 1807: 36–41. Verkhovskaya, M.L., Belevich, N., Euro, L., Wikström, M., and Verkhovsky, M.I. (2008) Real-time electron transfer in respiratory complex I. Proc Natl Acad Sci USA 105: 3763– 3767. Vinogradov, A.D., and Grivennikova, V.G. (2005) Generation of superoxide-radical by the NADH:ubiquinone oxidoreduc- tase of heart mitochondria. Biochemistry (Mosc) 70: 120– 127. Wikström, M. (1984) Two protons are pumped from the mito- chondrial matrix per electron transferred between NADH and ubiquinone. FEBS Lett 169: 300–304. Wikström, M., and Hummer, G. (2012) Stoichiometry of proton translocation by respiratory complex I and its mechanistic implications. Proc Natl Acad Sci USA 109: 4431–4436. Zoccarato, F., Cavallini, L., and Alexandre, A. (2004) Respiration-dependent removal of exogenous H2O2 in brain mitochondria – inhibition by Ca2+ . J Biol Chem 279: 4166–4174. Supporting information Additional supporting information may be found in the online version of this article at the publisher’s web-site. 1200 J. Knuuti et al. ■ © 2013 John Wiley & Sons Ltd, Molecular Microbiology, 90, 1190–1200