Vous êtes sur la page 1sur 7

Biochimica et Biophysica Acta 1857 (2016) 380–386

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

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

Electron transfer between the QmoABC membrane complex and


adenosine 5′-phosphosulfate reductase
Américo G. Duarte ⁎, André A. Santos, Inês A.C. Pereira ⁎
Instituto de Tecnologia Química e Biológica António Xavier, Universidade Nova de Lisboa, Av. da República, Estação Agronómica Nacional, 2780-157, Oeiras, Portugal

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

Article history: The dissimilatory adenosine 5′-phosphosulfate reductase (AprAB) is a key enzyme in the sulfate reduction
Received 16 September 2015 pathway that catalyzes the reversible two electron reduction of adenosine 5′-phosphosulfate (APS) to sulfite
Received in revised form 30 December 2015 and adenosine monophosphate (AMP). The physiological electron donor for AprAB is proposed to be the
Accepted 4 January 2016
QmoABC membrane complex, coupling the quinone-pool to sulfate reduction. However, direct electron transfer
Available online 6 January 2016
between these two proteins has never been observed. In this work we demonstrate for the first time direct
Keywords:
electron transfer between the Desulfovibrio desulfuricans ATCC 27774 QmoABC complex and AprAB. Cyclic
Dissimilatory sulfur metabolism voltammetry conducted with the modified Qmo electrode and AprAB in the electrolyte solution presented the
Respiratory membrane complex Qmo electrochemical signature with two additional well-defined one electron redox processes, attributed to
Electron-transfer the AprAB FAD redox behavior. Moreover, experiments performed under catalytic conditions using the QmoABC
Electrochemistry modified electrode, with AprAB and APS in solution, show a catalytic current peak develop in the cathodic wave,
Quinone-pool attributed to substrate reduction, and which is not observed in the absence of QmoABC. Substrate dependence
conducted with different electrode preparations (with and without immobilized Qmo) demonstrated that the
QmoABC complex is essential for efficient electron delivery to AprAB, in order to sustain catalysis. These results
confirm the role of Qmo in electron transfer to AprAB.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction Dissimilatory AprAB has been purified from different microorgan-


isms [4–9]. It is a flavoenzyme, characterized as an αβ heterodimer,
Sulfate reducing prokaryotes (SRP) are widespread in nature [1] where the α-subunit (͠75 kDa) comprises the catalytic site, a non-
particularly in anaerobic sulfate-rich environments, such as marine covalently bound flavin adenosine dinucleotide (FAD) cofactor [9]. The
sediments, given their respiratory metabolism that converts sulfate to small β-subunit with approximately 18 kDa harbors two [4Fe4S] clus-
sulfide. Although, many decades have been passed since sulfate reduc- ters, which function in electron transfer [10]. Crystal structures showed
tion was first associated with oxidative phosphorylation [2], the that the catalytic FAD cofactor is close to one of the [4Fe4S] centers
complete electron transfer pathways and mechanisms for energy (center I), which is buried inside the β-subunit, whereas the other FeS
conservation are not fully elucidated. The proteins required for dissim- cluster (center II), is more solvent exposed and responsible for
ilatory sulfate reduction are highly conserved among SRP and some of accepting electrons from the enzyme physiological electron donor
them are also present in sulfur-oxidizing organisms [3]. Sulfate reduc- [7,10,11]. The midpoint redox potentials of the Fe–S clusters were deter-
tion is an intracellular pathway that requires the transport of sulfate mined by EPR spectroscopy [5] as −60 mV and −540 mV (vs NHE), for
to the cytoplasm and its subsequent activation, a step catalyzed by the centers I and II, respectively. The discrepancy between the two redox
ATP sulfurylase, producing adenosine 5′-phosphosulfate (APS). APS is potentials was attributed to the different environment surrounding
then reduced to sulfite by the adenosine 5′-phosphosulfate reductase the Fe–S clusters, where the number of dipoles surrounding the centers
(AprAB), and sulfite is converted to sulfide by the dissimilatory sulfite is modulating their redox potential [11] (and references therein). It was
reductase (DsrAB) with the involvement of DsrC (Fig. 1A). proposed that electrons enter the enzyme by center II and are trans-
ferred to center I, towards the catalytic FAD moiety [7,11]. Substrate dif-
Abbreviations: AMP, adenoside monophosphate; APS, adenosine 5′-phosphosulfate; fusion to the buried FAD moiety is assured by a long and wide channel
AprAB, adenosine 5′-phosphosulfate reductase; BN-PAGE, blue native polyacrylamide (approximately 1.7 nm long and 1.0 nm wide), towards the catalytic
gel electrophoresis; CV, cyclic voltammogram; DDM, dodecyl-β-D-maltoside; FAD, flavin site that is partially solvent exposed [7,10]. The physiological electron
adenosine dinucleotide; KPB, potassium phosphate buffer; QmoABC, quinone- donor for AprAB is believed to be the quinone-interacting respiratory
interacting membrane-bound oxidoreductase; SRP, sulfate reducing prokaryotes.
⁎ Corresponding authors.
complex, QmoABC, whose genes are often found next to aprAB
E-mail addresses: americoduarte@itqb.unl.pt (A.G. Duarte), ipereira@itqb.unl.pt (Fig. 1B) [12]. A Desulfovibrio vulgaris qmoABC deletion mutant was un-
(I.A.C. Pereira). able to grow in lactate-sulfate, but remained capable of growing in

http://dx.doi.org/10.1016/j.bbabio.2016.01.001
0005-2728/© 2016 Elsevier B.V. All rights reserved.
A.G. Duarte et al. / Biochimica et Biophysica Acta 1857 (2016) 380–386 381

ε392 nm = 50 mM−1·cm−1 [8]. Pure protein samples were stored in


100 mM potassium phosphate buffer (KPB) 10% glycerol (and 0.1%
w/v DDM, for QmoABC) at − 80 °C prior to use. The purified enzyme
had a Vmax of 5.8 μmol APS·min−1·mg−1, which is similar to previously
reported [5].

2.2. Electrochemical methods

2.2.1. Protein immobilization to the graphite electrode surface


First, the pyrolytic graphite electrode (∅ = 0.3 cm) was immersed in
a diluted nitric acid solution, rinsed in deionized water, hand polished in
1 μm alumina (for at least 30 min), briefly sonicated and finally rinsed
with deionized water. Drops of protein solution (QmoABC 69 μM and
AprAB 480 μM) from 4 to 8 μl were added to the polished pyrolytic
graphite electrode surface in order to form a protein film using the
solvent casting technique. Electrochemical measurements with AprAB
in solution were made by additions of the protein stock solution to the
electrolyte solution.

2.2.2. Direct electrochemical studies


All electrochemical measurements were conducted inside a Coy an-
aerobic chamber (98% Argon, 2% H2, atmosphere), at room temperature
Fig. 1. Schematic representation of the sulfate reduction pathway. A) The complete sulfate (23 °C) using an electrochemical cell in a three electrode system: pyro-
reduction pathway. B) The proposed electron transfer between the quinol pool and the lytic graphite electrode as the working electrode, Ag/AgCl electrode as
APS reduction step. Blue arrows indicate electron movement.
the reference electrode and a platinum wire as the counter electrode.
Electrodes were connected to a CH potentiostat (Electrochemical ana-
lactate-sulfite, demonstrating the essential role of the QmoABC complex lyzer) with data acquired by the manufacturer's software (CHI1201a).
for sulfate, but not sulfite, reduction [13]. Several interaction studies, The electrolyte buffer solution used in all experiments was 250 mM
including co-immunoprecipitation, surface plasmon resonance, cross- KPB pH 6.8.
linking Far-Western blot and tag-affinity purification demonstrated a Cyclic voltammetry (CV) was evaluated in different scan rates,
physical interaction between QmoABC and AprAB [14]. However, direct ranging from 5 to 5000 mV s−1 and control experiments were conduct-
electron transfer between these two proteins was never reported. ed for all assays with the pyrolytic graphite bare electrode. Subtractions
The QmoABC complex, is composed by two cytoplasmic subunits and voltammogram treatment were conducted with the Origin soft-
(QmoAB) and one membrane subunit (QmoC). Its biochemical charac- ware. In all experiments, the first voltammograms were discarded. All
terization revealed the presence of multiple redox cofactors: two potentials mentioned in this work were corrected to the Standard
b-type hemes, two FAD and several FeS centers [12,14]. The QmoC sub- Hydrogen Electrode (SHE), with a 197 mV correction factor.
unit belongs to the family of membrane cytochromes b where its hydro- For non-control diffusion redox processes, the number of electrons
phobic C-terminal domain includes six transmembrane helices binding involved was estimated by the signal peak width at half current height
two heme b groups [12,15]. UV–visible titration revealed the midpoint (ΔEp1/2) [17] and the heterogeneous rate constant (ks) was estimated
redox potentials of the b-type hemes as +75 and −20 mV. Moreover, interpolating the anodic to cathodic peak separation (ΔEp) according
using menadiol (a menaquinol analog) as electron donor, it was possi- Laviron's formulation [18] (Supporting information).
ble to completely reduce the hemes, supporting quinol interaction of
the QmoABC complex [12]. 2.2.3. Electrochemical measurements under turnover conditions
In this work we demonstrate for the first time, direct electron The electrochemical response under turnover conditions was evalu-
transfer between the Desulfovibrio desulfuricans ATCC 27774 QmoABC ated using the modified QmoABC graphite electrode in the presence of
complex and the soluble AprAB. The QmoABC complex was immobilized AprAB in solution. In these experiments increasing amounts of substrate
on a pyrolytic graphite electrode, establishing direct electron transfer (APS) were added to the electrolyte buffer from anaerobic stock
with AprAB present in the electrolyte solution, in turnover and non- solutions. Control experiments were conducted with the bare electrode
turnover conditions. These experiments demonstrate direct electron with and without AprAB in solution, or with the modified Qmo
transfer between the two proteins, confirming the role of the QmoABC electrode without AprAB in solution.
complex as electron donor for AprAB.
3. Results and discussion
2. Materials and methods
3.1. QmoABC electrochemical characterization
2.1. Cell growth and protein purification
A direct electrochemical response of the immobilized QmoABC com-
D. desulfuricans ATCC 27774 was grown as previously reported [14]. plex was observed by the presence of a quasi-reversible redox process
Both QmoABC complex and AprAB purification were carried in anaero- (Fig. 2). This broad signal starts to develop between −160 and 10 mV
bic conditions inside a Coy anaerobic chamber (98% Argon, 2% H2 atmo- in the cathodic wave and between − 60 and 150 mV in the anodic
sphere) as previously described by Ramos et al. [14]. The purification wave. Subtraction of control voltammograms without protein clearly
steps were monitored by UV–visible spectroscopy and polyacrylamide evidenced the anodic and cathodic peaks (Fig. 2B) centered at − 9 ±
gel electrophoresis (SDS-PAGE for AprAB; and BN-PAGE, for the 7 mV. This well–defined redox process presents a current peak intensity
QmoABC complex). Protein concentration was estimated by colorimet- ratio close to one that is dependent with the scan rate increase, as
ric assays [16] and a molar extinction coefficient of ε408 nm = shown in Fig. 2C, characteristic of a surface confined redox process
91.4 mM− 1·cm− 1 was estimated for the Qmo complex. The AprAB [17]. This is the first electrochemical study performed with the QmoABC
concentration was determined by the molar extinction coefficient of complex adsorbed to an electrode surface, showing that it can establish
382 A.G. Duarte et al. / Biochimica et Biophysica Acta 1857 (2016) 380–386

Fig. 2. Cyclic voltammograms of immobilized QmoABC complex (100 mV s−1). A) CV of immobilized QmoABC complex (black line) and control experiment without protein (dashed line);
B) subtraction of the voltammograms in A; C) scan rate dependence of the anodic and cathodic current peak intensities of the QmoABC redox process.

direct electron transfer. Since the QmoABC complex has multiple redox two electron redox processes arising from the QmoABC cofactors, as
cofactors (hemes, Fe-S clusters and FAD [12]) with unclear stoichiome- described in the literature for other multi-cofactor proteins [19–21].
try and no detailed midpoint redox potential characterization, it is im- The scan rate dependence showed an increase on ΔEp and on the
possible to conduct a full deconvolution of the QmoABC redox signal. anodic and cathodic peaks area, more evident at high scan rates
The broad signal peak width at half current height (ΔEp1/2) would (Table S1 — Supporting information). This may be expected for multi
suggest one electron is involved in the redox process. However, the redox center proteins such as Qmo, since the increase in scan rate will
redox process observed should account for the sum of several one or favor particularly the electron transfer of those redox centers that are

Fig. 3. Direct electrochemical experiments of immobilized and soluble AprAB. All figures show on top the experiment (black) and control (dashed) voltammograms, and below the
respective subtraction. A) CV of the immobilized AprAB (100 mV s−1); B) CV of the immobilized free FAD cofactor (100 mV s−1), Γ = 51 ± 6 pmol·cm−2; C) CV of AprAB in solution
(39 μM, 100 mV s−1).
A.G. Duarte et al. / Biochimica et Biophysica Acta 1857 (2016) 380–386 383

directly in contact with the electrode surface. Nevertheless, it should be


noted that the potentials where the redox signal starts to develop are in
agreement with the midpoint redox potentials of the QmoABC b-type
hemes (+75 and −20 mV) [12]). This suggests that these redox centers
may be involved in direct electron transfer towards the electrode.

3.2. AprAB electrochemical characterization

The AprAB electrochemical response was evaluated with immobili-


zation to the pyrolytic graphite electrode or in solution (Fig. 3). The
immobilized AprAB electrochemical response showed two close redox
processes, named here as process 1 (at −282 ± 5) and 2 (at −197 ±
4 mV) (Fig. 3A). Process 1 was attributed to the electrochemical
response from the native AprAB, while the much less intense process
2 may arise from an alternative AprAB conformation immobilized at
the electrode surface, such as a minor protein fraction with an exposed
flavin cofactor, since this is not covalently bond to the enzyme. Both the
CV shape and the measured redox potentials are in agreement with
those observed for other flavin-containing proteins [20,22]. The current
peak intensity from both anodic and cathodic processes was propor-
tional to the scan rate increase (supporting information, Fig S1) and
its ratio close to one. Anodic and cathodic peak areas are proportional
and ΔEp remains small (≤67 mV) with scan rate dependence, which is
characteristic of strongly adsorbed redox species at the electrode
surface (Table S1, Supporting information). The measured ΔEp1/2 from
process 1 suggests a one electron redox process is present. However,
the FAD accounts for a two electron redox process (Fig. 3B), so the Fig. 4. Direct electron transfer between immobilized QmoABC complex and AprAB in
broad peak must correspond to the overlap of two one-electron redox solution. Top — CV with the modified QmoABC electrode with 20 μM AprAB in solution
(50 mV s−1, black line), and control experiment (dashed line); bottom — voltammogram
processes. The electrochemical parameters are similar to those observed
subtraction.
for the free FAD cofactor immobilized to the pyrolytic graphite electrode
surface (−317 ± 4 and −220 ± 7 mV, Fig. 3B), suggesting that the two
redox processes observed for the immobilized AprAB are from the direct remains evident from − 110 to 50 mV in the difference voltammo-
electrochemical response of the catalytic FAD. Using process 1 ΔEp and grams, besides the redox processes 1 and 2 of AprAB. All these signals
applying Laviron's formulation [18] (supporting information), the are observed at different scan-rates as shown in supporting information
heterogeneous rate constant for electron transfer (ks) was determined Fig. S2, where the two well-defined AprAB redox processes become
to be in the order of 4 s−1. The surface coverage was determined to be more evident with scan rate increase (between 50 and 200 mV·s−1).
51 ± 6 pmol/cm2, which according to the AprAB crystal structures According to the measured ΔEp1/2, both redox processes of AprAB
[7,11], corresponds to a multilayer coating. account for one electron reduction and the current peak intensity ratio
When the enzyme is added to the electrolyte solution, the recorded from the anodic and cathodic peaks is close to one,. However, the anodic
CVs present only one redox process (Fig. 3C). When AprAB (or free FAD) and cathodic current peak intensity is now proportional to the scan rate
is immobilized, the flavin cofactor is orientated towards the electrode, increase, as characteristic of surface confined systems whereas for iso-
favoring electron transfer, which is less evident when the enzyme is lated AprAB in solution the current peak intensity is proportional to
maintained in solution. A close examination of the baseline subtracted the scan rate square root (supporting information Fig. S1 and
voltammograms shows a broad redox peak, developing a shoulder Table S1). Additionally, increasing protein concentration of AprAB did
only in the anodic curve, at higher potential values, while the cathodic not alter significantly the current peak intensity for both anodic and
wave presents a broader signal. Therefore, according to what was cathodic peaks (Supporting information S3). This is consistent with a
observed for the immobilized enzyme, this signal was assigned to the physical interaction between the two proteins, producing a QmoABC–
interaction of the AprAB catalytic FAD with the electrode. The midpoint AprAB electron transfer complex, as previously reported [14]. Therefore,
redox potential determined was −309 ± 5 mV. The determined ΔEp is considering the electrode-bound QmoABC as the electron donor unit for
now higher than that observed when the enzyme was immobilized to AprAB, it is possible to estimate the heterogeneous rate constant
the electrode surface. The anodic and cathodic current peak intensity between the modified QmoABC electrode and the soluble AprAB. The
ratio was close to one and proportional to the scan rate square-root, determined values were in the order of 21 and 68 ± s− 1, for AprAB
which is expected for diffusion controlled systems, since the enzyme processes 1 and 2, respectively. These values are higher than the ones
was added to the electrolyte solution (supporting information Fig. S1 determined for immobilized AprAB alone and indicate a better interac-
and Table S1). tion and faster electron transfer of AprAB with the modified QmoABC
electrode than with the bare electrode. In order to assure that the ob-
3.3. Direct electron transfer from immobilized QmoABC to AprAB served redox signal is from AprAB interacting with the immobilized
QmoABC complex and not from an unspecific flavin in solution
In order to evaluate direct electron transfer between QmoABC and (e.g. FAD released from denatured enzyme), some free FAD was added
AprAB, electrochemical experiments were performed with the modified to the electrolyte solution with the QmoABC complex adsorbed to the
QmoABC graphite electrode adding AprAB to the electrolyte solution. working electrode. The obtained CVs (Supporting information, Fig. S4)
Baseline subtracted voltammograms show two distinct well-defined show two redox processes quite similar to the results observed for the
redox processes at − 300 ± 7 mV and − 202 ± 6 mV (Fig. 4). immobilized free FAD (Fig. 3B), but dissimilar from those observed
These redox processes are similar to those observed for the direct elec- with AprAB.
trochemical characterization of AprAB either immobilized or in solution Electrochemical measurements have been reported as useful tools
(Section 3.2). The signal resulting from QmoABC reduction (Section 3.1) to study protein–protein electron transfer, under turnover and non-
384 A.G. Duarte et al. / Biochimica et Biophysica Acta 1857 (2016) 380–386

turnover conditions [23–25]. Our data analysis demonstrate for the first to the AprAB catalytic activity of APS reduction. Importantly, absence of
time direct electron transfer between the immobilized QmoABC QmoABC from the electrode prevented the production of a catalytic cur-
complex and AprAB. When using the bare electrode to transfer elec- rent (Fig. 5A2 and B2). The measured current amplitude of the cathodic
trons to AprAB (immobilized or in solution) the two redox processes current peak shows a substrate concentration dependence (Fig. 5C). The
of FAD bound to AprAB appear less resolved than in the presence of hyperbolic shape observed with the modified QmoABC electrode is typi-
immobilized QmoABC. When QmoABC is adsorbed to the electrode, cal of an enzyme kinetic behavior, showing a maximum current (Imax)
electrons are transferred to the AprAB catalytic site through the Qmo of 0.78 μA and an observed affinity constant (kobs) of 61 μM, which is in
and AprAB FeS centers. The detection of two well-defined one electron the same order of the KM reported for this enzyme [5]. In contrast, the cat-
redox processes, suggest a stable flavosemiquinone state in the alytic activity of AprAB with the bare electrode, without immobilized
presence of QmoABC, a feature observed in AprAB enzymes isolated QmoABC, showed no significant cathodic current variations (Fig. 5C).
from sulfate reducing bacteria [5]. Other control experiments conducted with the bare electrode or the
modified QmoABC electrode without AprAB, using different APS concen-
3.4. AprAB catalytic activity with modified QmoABC graphite electrode trations, did not exhibit substantial catalytic current variations (see
Supporting information, Fig. S5).
Electron transfer between the modified QmoABC electrode and
soluble AprAB was also studied under turnover conditions. For this 3.5. Concluding remarks
purpose, the AprAB catalytic behavior was monitored in the presence
of the APS substrate, with the modified QmoABC graphite electrode as CVs obtained with the modified QmoABC electrode and AprAB in the
electron donor. In these experiments a catalytic current was observed, electrolyte buffer, clearly show an electrochemical signal from AprAB
starting to develop at approximately −150 mV on the cathodic wave catalytic site, meaning that when AprAB interacts with the adsorbed
(Fig. 5A1). Baseline voltammogram subtraction produced a defined ca- Qmo, electrons enter the enzyme and the flavin semi-reduced state is
thodic catalytic peak, which increased in intensity with higher APS con- stabilized. This is supported by the two well-defined redox processes
centrations (Fig. 5B1). Since the kinetic studies were conducted with a observed, only when Qmo was immobilized to the electrode (Fig. 4),
stationary pyrolitic graphite electrode, the catalytic current measured de- which are merged in the free enzyme in solution (Fig. 3). Electrochem-
pends on substrate diffusion. The described catalytic peak was attributed ical measurements under turnover conditions using the QmoABC

Fig. 5. Direct electron transfer between immobilized QmoABC complex and AprAB, under turnover conditions. A1) Modified QmoABC electrode with AprAB in solution with increasing APS
concentrations; A2) bare pyrolytic graphite electrode with AprAB in solution with increasing APS concentrations; B1) baseline subtracted cathodic wave from modified QmoABC electrode
with AprAB in solution; with increasing APS concentrations; B2) baseline subtracted cathodic wave from bare pyrolytic graphite electrode with AprAB in solution with increasing APS
concentrations; the small black arrows point the minimum and maximum points where current peak intensity was measured. C) Catalytic current of APS reduction measured with
AprAB present in the electrolyte solution and: modified QmoABC electrode (black squares), bare graphite electrode (open squares). The black line corresponds to an enzyme kinetic
hyperbolic plot; Imax = 0.78 μA, kobs = 61 μM. In all experiments: [AprAB] = 5 μM; scan rate — 50 mV s−1.
A.G. Duarte et al. / Biochimica et Biophysica Acta 1857 (2016) 380–386 385

modified electrode also demonstrate an effective electron transfer Qmo electrode, were the AprAB catalytic FAD redox signal presents
between the membrane complex and the enzyme, leading to APS reduc- two consecutive one electron redox processes (Fig. 4). Physiological
tion (Fig. 5). Electrons are transferred from the graphite electrode to the electron transfer between QmoABC and AprAB has been proposed to re-
Qmo complex, from where they are correctly orientated to AprAB. Since quire the presence of a third redox partner, in an electron confurcation
QmoA is the subunit proposed to interact with AprAB [14], it is likely to process, as direct electron transfer between a menaquinol analog and
dock with the AprB subunit since this has the solvent exposed [4Fe4S] AprAB through the Qmo complex has never been observed [14]. By
electron transfer center. The midpoint redox potentials of the AprAB using this electrochemical strategy, the modified Qmo electrode is capa-
FeS centers and their proximity drives electron transfer towards the cat- ble of transferring electrons to the soluble AprAB, sustaining catalysis,
alytic site [7] for APS reduction (Fig. 6A). Additionally, since QmoABC which was never observed in solution. This might be explained by
was characterized as a quinol-oxidizing complex [12], it is proposed the negative overpotential values imposed by the electrode (below
that electrons are transferred from the quinol-pool, through Qmo, to − 100 mV), that probably drive the reaction, whereas in vivo the
the AprAB catalytic site. menaquinol pool does not have a low enough redox potential to achieve
APS reduction was not detected when the QmoABC complex was ab- reduction of APS, requiring the involvement of a third partner This is the
sent from the electrode surface, meaning the electron exchange with first electrochemical study conducted with enzymes and respiratory
the enzyme catalytic site, observed for isolated AprAB (Fig. 3), does complexes involved in sulfate reduction. In this work we demonstrate
not sustain catalysis (Fig. 6B). Most likely, when immobilized, AprAB in- that there is direct electron transfer between the soluble dissimilatory
teracts with the electrode in an orientation that favors FAD direct elec- AprAB and the membrane QmoABC complex. The proposal of electron
tron transfer but blocks the enzyme substrate channel [7,10], transfer between QmoABC and AprAB was suggested more than ten
preventing substrate diffusion towards the catalytic site. Therefore, years ago but was never demonstrated. By means of electrochemical
the AprAB catalytic mechanism require that electrons reach the catalytic measurements it was possible to detect protein–protein electron trans-
FAD site from the adjacent [4Fe4S] center I, and for this electrons must fer under turnover and non-turnover conditions. The experiments
enter the enzyme via center II. In the present experiments, this electron conducted under non-turnover conditions show clearly electron
transfer mechanism was only accomplished when using the modified transfer between QmoABC membrane complex and AprAB. Further-
more, this electron transfer supports electrochemical catalytic activity
of AprAB. The electrochemical and kinetic data confirm the role of the
QmoABC complex as electron donor for AprAB.

Transparency document

The Transparency document associated with this article can be


found, in online version.

Acknowledgments

The authors acknowledge to Ricardo O. Louro and Catarina M.


Paquete for the use of the electrochemical apparatus and anaerobic
chamber, and Antonio L. de Lacey for scientific discussions. This work
was supported by the Fundação para a Ciência e Tecnologia (FCT/
MTCES) through grants PTDC/BBB-BQB/0684/2012 and UID/Multi/
04551/2013, and fellowships SFRH/BPD/84607/2012 (AGD), and SFRH/
BD/77940/2011 (AAS).

Appendix A. Supplementary data

Supplementary data to this article can be found online at http://dx.


doi.org/10.1016/j.bbabio.2016.01.001.

References

[1] G. Muyzer, A.J. Stams, The ecology and biotechnology of sulphate-reducing bacteria,
Nat. Rev. Microbiol. 6 (2008) 441–454.
[2] H.D. Peck Jr., Evidence for oxidative phosphorylation during the reduction of sulfate
with hydrogen by Desulfovibrio desulfuricans, J. Biol. Chem. 235 (1960) 2734–2738.
[3] I.A. Pereira, A.R. Ramos, F. Grein, M.C. Marques, S.M. da Silva, S.S. Venceslau, A
comparative genomic analysis of energy metabolism in sulfate reducing bacteria
and archaea, Front. Microbiol. 2 (2011) 69.
[4] G. Fritz, T. Buchert, H. Huber, K.O. Stetter, P.M. Kroneck, Adenylylsulfate reductases
from archaea and bacteria are 1:1 alphabeta-heterodimeric iron–sulfur
flavoenzymes — high similarity of molecular properties emphasizes their central
role in sulfur metabolism, FEBS Lett. 473 (2000) 63–66.
[5] G. Fritz, T. Buchert, P.M. Kroneck, The function of the [4Fe–4S] clusters and FAD in
bacterial and archaeal adenylylsulfate reductases. Evidence for flavin-catalyzed re-
duction of adenosine 5′-phosphosulfate, J. Biol. Chem. 277 (2002) 26066–26073.
[6] H. Ogata, A.G. Agrawal, A.P. Kaur, R. Goddard, W. Gartner, W. Lubitz, Purification,
crystallization and preliminary X-ray analysis of adenylylsulfate reductase from
Desulfovibrio vulgaris Miyazaki F, Acta Crystallogr. Sect. F: Struct. Biol. Cryst.
Commun. 64 (2008) 1010–1012.
Fig. 6. Electron transfer between the QmoABC complex and dissimilatory AprAB. [7] Y.L. Chiang, Y.C. Hsieh, J.Y. Fang, E.H. Liu, Y.C. Huang, P. Chuankhayan, J. Jeyakanthan,
A) QmoABC complex immobilized to a pyrolytic graphite surface, transferring electrons M.Y. Liu, S.I. Chan, C.J. Chen, Crystal structure of adenylylsulfate reductase from
to AprAB for APS reduction (blue arrows). B) Electron transfer between the bare Desulfovibrio gigas suggests a potential self-regulation mechanism involving the C
pyrolytic graphite electrode and AprAB for APS reduction. terminus of the beta-subunit, J. Bacteriol. 191 (2009) 7597–7608.
386 A.G. Duarte et al. / Biochimica et Biophysica Acta 1857 (2016) 380–386

[8] J. Lampreia, I. Moura, M. Teixeira, H.D. Peck, J. Legall, B.H. Huynh, J.J.G. Moura, The [18] E. Laviron, General expression of the linear potential sweep voltammogram in the
active-centers of adenylylsulfate reductase from desulfovibrio-gigas — characterization case of diffusionless electrochemical systems, J. Electroanal. Chem. Interfacial
and spectroscopic studies, Eur. J. Biochem. 188 (1990) 653–664. Electrochem. 101 (1979) 19–28.
[9] J. Lampreia, A.S. Pereira, J.J.G. Moura, Adenylylsulfate reductases from sulfate-reducing [19] H.A. Heering, J.H. Weiner, F.A. Armstrong, Direct detection and measurement of
bacteria, Inorg. Microb. Sulfur Metab. 243 (1994) 241–260. electron relays in a multicentered enzyme: voltammetry of electrode-surface films
[10] G. Fritz, A. Schiffer, A. Behrens, T. Büchert, U. Ermler, P.H. Kroneck, Living on sulfate: of E. coli fumarate reductase, an iron–sulfur flavoprotein, J. Am. Chem. Soc. 119
three-dimensional structure and spectroscopy of adenosine 5′-phosphosulfate re- (1997) 11628–11838.
ductase and dissimilatory sulfite reductase, in: C. Dahl, C. Friedrich (Eds.), Microbial [20] M.J. Hamill, S.E. Chobot, H.H. Hernandez, C.L. Drennan, S.J. Elliott, Direct electro-
Sulfur Metabolism, Springer, Berlin Heidelberg 2008, pp. 13–23. chemical analyses of a thermophilic thioredoxin reductase: interplay between
[11] G. Fritz, A. Roth, A. Schiffer, T. Buchert, G. Bourenkov, H.D. Bartunik, H. Huber, K.O. conformational change and redox chemistry, Biochemistry 47 (2008) 9738–9746.
Stetter, P.M. Kroneck, U. Ermler, Structure of adenylylsulfate reductase from the [21] R.M.A.S. Doyle, D.J. Richardson, T.A. Clarke, J.N. Butt, Freely diffusing versus adsorbed
hyperthermophilic Archaeoglobus fulgidus at 1.6-A resolution, Proc. Natl. Acad. Sci. protein: which better mimics the cellular state of a redox protein? Electrochim. Acta
U. S. A. 99 (2002) 1836–1841. 110 (2013) 73–78.
[12] R.H. Pires, A.I. Lourenco, F. Morais, M. Teixeira, A.V. Xavier, L.M. Saraiva, I.A. Pereira, [22] S. Mayhew, Potentiometric measurement of oxidation–reduction potentials, in: S.
A novel membrane-bound respiratory complex from Desulfovibrio desulfuricans Chapman, G. Reid (Eds.), Flavoprotein Protocols, Humana Press 1999, pp. 49–59.
ATCC 27774, Biochim. Biophys. Acta 1605 (2003) 67–82. [23] H.A. Heering, F.G. Wiertz, C. Dekker, S. de Vries, Direct immobilization of native
[13] G.M. Zane, H.C. Yen, J.D. Wall, Effect of the deletion of qmoABC and the promoter- yeast iso-1 cytochrome C on bare gold: fast electron relay to redox enzymes and
distal gene encoding a hypothetical protein on sulfate reduction in Desulfovibrio zeptomole protein-film voltammetry, J. Am. Chem. Soc. 126 (2004) 11103–11112.
vulgaris Hildenborough, Appl. Environ. Microbiol. 76 (2010) 5500–5509. [24] M. Guiral, G. Leroy, P. Bianco, P. Gallice, B. Guigliarelli, M. Bruschi, W. Nitschke, M.T.
[14] A.R. Ramos, K.L. Keller, J.D. Wall, I.A. Pereira, The membrane QmoABC complex Giudici-Orticoni, Interaction and electron transfer between the high molecular weight
interacts directly with the dissimilatory adenosine 5′-phosphosulfate reductase in cytochrome and cytochrome c(3) from Desulfovibrio vulgaris Hildenborough: kinetic,
sulfate reducing bacteria, Front. Microbiol. 3 (2012) 137. microcalorimetric, EPR and electrochemical studies, Biochim. Biophys. Acta 1723
[15] I.A.C. Pereira, Membrane complexes in Desulfovibrio, in: C. Friedrich, C. Dahl (Eds.), (2005) 45–54.
Microbial Sulfur Metabolism, Springer-Verlag 2008, pp. 24–35. [25] D.G.G. McMillan, S.J. Marritt, M.A. Firer-Sherwood, L. Shi, D.J. Richardson, S.D. Evans,
[16] O.H. Lowry, N.J. Rosebrough, A.L. Farr, R.J. Randall, Protein measurement with the S.J. Elliott, J.N. Butt, L.J.C. Jeuken, Protein–protein interaction regulates the direction
folin phenol reagent, J. Biol. Chem. 193 (1951) 265–275. of catalysis and electron transfer in a redox enzyme complex, J. Am. Chem. Soc. 135
[17] A.J. Bard, L.R. Faulkener, Electrochemical Methods, John Wiley & Sons, Inc., USA, (2013) 10550–10556.
2001.

Vous aimerez peut-être aussi