Direct catalytic electrochemistry of sulfite dehydrogenase: Mechanistic insights and contrasts with...

7
Direct catalytic electrochemistry of sulte dehydrogenase: Mechanistic insights and contrasts with related Mo enzymes Trevor D. Rapson, Ulrike Kappler, Paul V. Bernhardt Centre for Metals in Biology, School of Molecular and Microbial Sciences, University of Queensland, Brisbane, 4072, Australia abstract article info Article history: Received 16 April 2008 Received in revised form 6 June 2008 Accepted 6 June 2008 Available online 13 June 2008 Keywords: Molybdenum Enzyme Electrochemistry Under hydrodynamic electrochemical conditions with slow cyclic voltammetry sweep rates we have been able to probe catalytic events at the molybdenum active site of sulte dehydrogenase (SDH) from Starkeya novella adsorbed on an edge plane graphite electrode within a polylysine lm. The electrochemically driven catalytic behaviour of SDH mirrors that seen in solution assays suggesting that the adsorbed enzyme retains its native activity. However, at high sulte concentrations, the voltammetric waveform transforms from the expected sigmoidal prole to a peak-shaped response, similar to that reported for the molybdenum enzymes DMSO reductase and nitrate reductase (NarGHI and NapAB) where a redox reaction at the active site has been associated with a switch to lower activity at high overpotentials. This is the rst time a similar phenomenon has been observed in a Mo-containing oxidase/dehydrogenase, which raises a number of interesting mechanistic problems. The potential at which the activity of SDH becomes attenuated only emerges at saturating substrate conditions and occurs at a potential (ca. + 320mV vs NHE) well removed from any known redox couple in the enzyme. These results cannot be explained by the same mechanism adopted for DMSO reductase and nitrate reductase catalysis. © 2008 Elsevier B.V. All rights reserved. 1. Introduction Adsorption of oxidoreductase enzymes on a chemically modied electrode, a technique known as protein lm voltammetry, allows direct electron transfer between the enzyme and the electrode, enabling electrochemically driven catalysis [1]. The electrode replaces the natural electron exchange partner of the enzyme, whilst connement of the enzyme to the electrode surface overcomes the disadvantage of slow enzyme diffusion to the electrode surface. Ideally the enzyme retains its full native activity whilst conned to the electrode [2,3]. A powerful feature of the method is that the catalytic activity of the enzyme can be investigated as a function of electrochemical potential. The operat- ing potential of an enzyme relates to the minimal electrochemical driving force which must be provided for catalysis and is typically in the vicinity of the formal potential of the enzyme's physiological redox partner [4]. A class of enzymes that we [59] and others [1016] have studied in some detail is the mononuclear Mo enzymes, which catalyze 2- electron, 2-proton, O-atom transfer reactions on a remarkable array of organic and inorganic substrates. Hille has classied these enzymes into three families according to the coordination environment of the Mo active site; the xanthine oxidase, sulte oxidase and DMSO reductase families [17]. The Mo active sites all cycle between their Mo VI and Mo IV oxidation states and an oxo or hydroxo ligand is exchanged between the substrate and Mo during turnover. Although the Mo VI form is the active oxidation state for the oxidase/dehydrogenase enzymes and the Mo IV form is active for the reductases, the Mo V oxidation state is often a stable intermediate. The sulte oxidase family of Mo enzymes share an active site comprising a ve-coordinate Mo ion bearing a single molybdopterin chelating ligand, a Cys residue and two oxo ligands in its Mo VI form (Fig. 1). The most intensively studied members of this family [18] comprise the human and chicken sulte oxidases (HSO and CSO), plant sulte oxidase (PSO) and a sulte dehydrogenase (SDH) isolated from the soil bacterium Starkeya novella [19]; the subject of this investigation. In each case, the oxidation of sulte to sulfate involves transfer of the equatorially coordinated oxo ligand (Fig. 1) to the substrate and this is summarised in Eq. (1). The sulfate ligand is then Biochimica et Biophysica Acta 1777 (2008) 13191325 Abbreviations: SDH, sulte dehydrogenase; DMSO, dimethyl sulfoxide; Cys, cysteine; HSO, human sulte oxidase; CSO, chicken sulte oxidase; PSO, plant sulte oxidase; EDTA, ethylenediamine-tetraacetete; tris, tris(hydroxymethyl)-methylammo- nium; NHE, normal hydrogen electrode; RDV, rotating disk voltammetry; EPG, edge plane pyrolytic graphite; DDAB, didodecyldimethylammonium bromide; SDH, sulte dehydrogenase; DMSO, dimethyl sulfoxide; Cys, cysteine; HSO, human sulte oxidase; CSO, chicken sulte oxidase; PSO, plant sulte oxidase; EDTA, ethylenediamine- tetraacetete; tris, tris(hydroxymethyl)-methylammonium; NHE, normal hydrogen electrode; RDV, rotating disk voltammetry; EPG, edge plane pyrolytic graphite; DDAB, didodecyldimethylammonium bromide; Tyr, tyrosine; Phe, phenylalanine Corresponding author. E-mail address: [email protected] (P.V. Bernhardt). 0005-2728/$ see front matter © 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.bbabio.2008.06.005 Contents lists available at ScienceDirect Biochimica et Biophysica Acta journal homepage: www.elsevier.com/locate/bbabio

Transcript of Direct catalytic electrochemistry of sulfite dehydrogenase: Mechanistic insights and contrasts with...

Biochimica et Biophysica Acta 1777 (2008) 1319–1325

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

j ourna l homepage: www.e lsev ie r.com/ locate /bbab io

Direct catalytic electrochemistry of sulfite dehydrogenase: Mechanistic insights andcontrasts with related Mo enzymes

Trevor D. Rapson, Ulrike Kappler, Paul V. Bernhardt ⁎Centre for Metals in Biology, School of Molecular and Microbial Sciences, University of Queensland, Brisbane, 4072, Australia

Abbreviations: SDH, sulfite dehydrogenase; DMScysteine; HSO, human sulfite oxidase; CSO, chicken suloxidase; EDTA, ethylenediamine-tetraacetete; tris, tris(hnium; NHE, normal hydrogen electrode; RDV, rotatingplane pyrolytic graphite; DDAB, didodecyldimethylammdehydrogenase; DMSO, dimethyl sulfoxide; Cys, cysteinCSO, chicken sulfite oxidase; PSO, plant sulfite oxidtetraacetete; tris, tris(hydroxymethyl)-methylammonelectrode; RDV, rotating disk voltammetry; EPG, edge pdidodecyldimethylammonium bromide; Tyr, tyrosine; P⁎ Corresponding author.

E-mail address: [email protected] (P.V. Bernha

0005-2728/$ – see front matter © 2008 Elsevier B.V. Aldoi:10.1016/j.bbabio.2008.06.005

a b s t r a c t

a r t i c l e i n f o

Article history:

Under hydrodynamic electr Received 16 April 2008Received in revised form 6 June 2008Accepted 6 June 2008Available online 13 June 2008

Keywords:MolybdenumEnzymeElectrochemistry

ochemical conditions with slow cyclic voltammetry sweep rates we have beenable to probe catalytic events at the molybdenum active site of sulfite dehydrogenase (SDH) from Starkeyanovella adsorbed on an edge plane graphite electrode within a polylysine film. The electrochemically drivencatalytic behaviour of SDH mirrors that seen in solution assays suggesting that the adsorbed enzyme retainsits native activity. However, at high sulfite concentrations, the voltammetric waveform transforms from theexpected sigmoidal profile to a peak-shaped response, similar to that reported for the molybdenum enzymesDMSO reductase and nitrate reductase (NarGHI and NapAB) where a redox reaction at the active site hasbeen associated with a switch to lower activity at high overpotentials. This is the first time a similarphenomenon has been observed in a Mo-containing oxidase/dehydrogenase, which raises a number ofinteresting mechanistic problems. The potential at which the activity of SDH becomes attenuated onlyemerges at saturating substrate conditions and occurs at a potential (ca. + 320mV vs NHE) well removed fromany known redox couple in the enzyme. These results cannot be explained by the same mechanism adoptedfor DMSO reductase and nitrate reductase catalysis.

© 2008 Elsevier B.V. All rights reserved.

1. Introduction

Adsorption of oxidoreductase enzymes on a chemically modifiedelectrode, a technique known as protein film voltammetry, allowsdirect electron transfer between the enzyme and the electrode,enabling electrochemically driven catalysis [1]. The electrodereplaces the natural electron exchange partner of the enzyme,whilst confinement of the enzyme to the electrode surfaceovercomes the disadvantage of slow enzyme diffusion to theelectrode surface. Ideally the enzyme retains its full native activitywhilst confined to the electrode [2,3]. A powerful feature of themethod is that the catalytic activity of the enzyme can beinvestigated as a function of electrochemical potential. The operat-

O, dimethyl sulfoxide; Cys,fite oxidase; PSO, plant sulfiteydroxymethyl)-methylammo-disk voltammetry; EPG, edgeonium bromide; SDH, sulfite

e; HSO, human sulfite oxidase;ase; EDTA, ethylenediamine-ium; NHE, normal hydrogenlane pyrolytic graphite; DDAB,he, phenylalanine

rdt).

l rights reserved.

ing potential of an enzyme relates to the minimal electrochemicaldriving force which must be provided for catalysis and is typically inthe vicinity of the formal potential of the enzyme's physiologicalredox partner [4].

A class of enzymes thatwe [5–9] and others [10–16] have studied insome detail is the mononuclear Mo enzymes, which catalyze 2-electron, 2-proton, O-atom transfer reactions on a remarkable array oforganic and inorganic substrates. Hille has classified these enzymesinto three families according to the coordination environment of theMo active site; the xanthine oxidase, sulfite oxidase and DMSOreductase families [17]. TheMo active sites all cycle between theirMoVI

and MoIV oxidation states and an oxo or hydroxo ligand is exchangedbetween the substrate and Mo during turnover. Although the MoVI

form is the active oxidation state for the oxidase/dehydrogenaseenzymes and the MoIV form is active for the reductases, the MoV

oxidation state is often a stable intermediate.The sulfite oxidase family of Mo enzymes share an active site

comprising a five-coordinate Mo ion bearing a single molybdopterinchelating ligand, a Cys residue and two oxo ligands in its MoVI form(Fig. 1). The most intensively studied members of this family [18]comprise the human and chicken sulfite oxidases (HSO and CSO),plant sulfite oxidase (PSO) and a sulfite dehydrogenase (SDH) isolatedfrom the soil bacterium Starkeya novella [19]; the subject of thisinvestigation. In each case, the oxidation of sulfite to sulfate involvestransfer of the equatorially coordinated oxo ligand (Fig. 1) to thesubstrate and this is summarised in Eq. (1). The sulfate ligand is then

Fig. 1. Active site of enzymes from the sulfite oxidase family.

1320 T.D. Rapson et al. / Biochimica et Biophysica Acta 1777 (2008) 1319–1325

substituted by an aqua ligand and the Mo ion must be reoxidized to itsMoVI state for catalysis to continue.

O ¼ MoVI þ SO2–3 ⇌MoIV–OSO3 ð1Þ

SDH is a heterodimer comprising molybdenum and heme cbinding subunits [19] and these separate subunits occupy fixedpositions relative to one another during catalysis [20,21]. This is incontrast to CSO [22] and HSO where a heme b domain is connected tothe molybdenum binding domain by a flexible hinge. The hemecofactor is the site at which electrons are passed to the co-substrate(dioxygen) or electron transfer partner (cytochrome c). No hemecofactor has been identified in PSO [23] which highlights importantmechanistic diversity within this family.

In an earlier paper [6] we reported both non-turnover and theelectrocatalytic voltammetry of SDH. However, some interestingfeatures of the voltammetry from SDH were apparent including anon-ideal peak-shaped catalytic waveform rather than the expectedsigmoidal profile characteristic of a classical steady state voltammo-gram [2,24]. Similar features have continued to appear in the Moenzyme electrochemistry literature over the recent years with nitratereductases NarGH and NapAB [10,11,15,16] and DMSO reductase [12](all members of the DMSO reductase family). There has been progresstoward a unifying mechanism that explains this non-ideal behaviourand this has involved MoV participating as an important intermediatein the catalytic mechanism [25]. This is an interesting proposition andone that we have considered here with SDH; notably an oxidizingenzyme as opposed to all other Mo (reductase) enzymes that haveshown this feature. The mechanistic consequences of this reversal inreaction direction are very important as we shall illustrate.

Fig. 2. Catalytic cyclic voltammogram of SDH adsorbed on a kanamycin-modified EPGelectrode in the presence of 90 μM sulfite (solid curve) and without sulfite (brokencurve). Experimental conditions: pH 8.0, sweep rate 5mV s−sec−1, electrode rotation rate500 rpm, 298K. The arrow indicates the sweepdirection and the catalytic current at highpotential ilim is shown.

2. Materials and methods

2.1. Materials

SorAB (SDH) was purified from a heterologous expression systemin Rhodobacter capsulatus described previously [26]. All reagents usedwere of analytical grade purity and used without any further pre-treatment. All solutions were prepared in purified water (Millipore,18.2 MΩ.cm). Tris acetate (20 mM) was used for experimentsconducted in the pH range of 7.5 to 8.5. When a wider pH range (6–10) was investigated, a buffer mixture containing both bis–trispropane (10 mM) and 2-amino-2-methylpropan-1-ol (10 mM) wasused, titrated with acetic acid to give the desired pH. Sulfite andinhibitor anions investigatedwere added from a stock solution, freshlyprepared in a solution of tris acetate (50mM), pH 8.8with 5mMEDTA.

2.2. Electrochemical measurements and electrode preparation

Cyclic voltammetry and chronoamperometry were carried outwith a BAS100B/W electrochemical workstation using a threeelectrode system consisting of an edge plane pyrolytic graphite(EPG) working electrode, a platinumwire counter electrode and a Ag/AgCl reference electrode (+ 196mV vs NHE). All potentials have been

corrected relative to NHE. Experiments were carried out in Ar purgedsolutions on the bench or within a Belle Technology anaerobic boxunder an atmosphere of N2 (O2b10 ppm) at 25°C. The workingelectrode was attached to a BAS RDE-3 rotating disk cell stand androtated typically at 500rpm.

The working electrode (surface area ~0.1cm2) was prepared bycleaving a 1μm layer from the face of the electrode with a microtomeand then cleaned by sonication in MilliQ water. No abrasives wereused. The enzyme (2μL, 25 μM) was co-adsorbed onto the dry exposedgraphite electrode surface with 2μL solutions of either polylysine(1mg/mL), chitosan (10mg/mL), polyethylene imine (5% v/w) orkanamycin (10mg/mL) then air dried at 4 °C for ca. 3h.

In a typical cyclic voltammetry experiment, the potential wascycled between −100 mV and +400 mV vs NHE with an electroderotation rate of 500rpm and a scan rate of 5 mV s−1. Enzyme filmdegradation over time (as judged by periodic monitoring of thecurrent obtained from standard solutions) was negligible. In multi-sweep experiments 5cycles were carried out consecutively. Chron-oamperometric determinations of catalytic current were carried outby poising the rotating electrode at the potential of the maximumcatalytic current (determined from cyclic voltammograms) in thepresence of various concentrations of substrate and inhibitors.

2.3. Analysis of voltammogram shape

Voltammograms were analyzed using approaches describedpreviously [24]. The magnitude of the catalytic current wasdetermined by subtracting the voltammogram obtained from anexperiment carried out in the absence of substrate using theBAS100W software (vers. 2.3). Plotting the steady state catalyticcurrent (ilim, Fig. 2) as a function of sulfite concentration allowed thedetermination of the electrochemical Michaelis constant (KM,sulfite)according to Eq. (2), where imax is the catalytic current at saturatingsubstrate concentrations.

ilim =imax½SO2−

3 �KM;sulfite + ½SO2−

3 �ð2Þ

The catalytic potential and the steepness of the voltammetricsigmoidal waveform were determined from the first derivative

Fig. 3. (A) The pH dependence of SDH catalytic voltammetry (anodic sweeps shown forclarity). Conditions: 300 μM sulfite, sweep rate 5 mV/s, rotation rate 500 rpm. Note the−59 mV/pH unit shift of the catalytic wave as well as the decline in current above andbelow the pH optimum of 8; (B) the pH dependence of the catalytic potential (Ecat, blackdiamonds) reported here relative to the MoVI/V (grey circles), MoV/IV (grey triangles) andheme (FeIII/II, grey squares) redox potentials of SDH determined previously in Ref. [6].

1321T.D. Rapson et al. / Biochimica et Biophysica Acta 1777 (2008) 1319–1325

catalytic voltammogram (di/dE) as described [24]. The catalyticpotential is obtained from the maximum in the first derivativevoltammogram. The steepness of the wave, determined from thewidth of the first derivative at half height (δ), allows the cooperativitycoefficient (apparent number of electrons, napp) transferred in the ratedefining step of catalysis to be calculated from Eq. (3):

δ =3:53RTnappF

=90:6napp

mV at 25 ˚ C ð3Þ

3. Results

3.1. Optimisation of experimental conditions

For the theoretical models developed by Heering et al. [24] to beapplied in a useful way it is important to ensure that the overallcatalytic reaction is not impaired by slow substrate diffusion orsluggish interfacial electron transfer. Substrate depletion at the activesite occurs when the rate of the enzyme–substrate reaction exceedsthe rate at which substrate can be replenished by diffusion alone. Thisproblem may be overcome through rotating disk voltammetry (RDV)to provide a constant and rapid delivery of substrate to the active site[24].

The correct choice of working electrode surface and co-adsorbateis critical in ensuring effective interfacial electron transfer and also instabilising the adsorbed enzyme film. Enzyme desorption can beproblematic under the hydrodynamic demands of RDV. On the basis ofreproducibility and magnitude of the response, kanamycin wasinitially chosen as the preferred electro-inactive co-adsorbate (orpromoter) to stabilise the SDH film whilst adsorbed on an edge planepyrolytic graphite (EPG) rotating disk working electrode. However, thekanamycin-SDH filmwas not stable above pH 8.5. The aminoglycosidekanamycin bears four primary amino groups with pKa values in therange 6.2–9.0 [27]. Therefore, above pH 8.5 most of the positive chargeof kanamycin that is essential in forming a stable ternary systemwith(negatively charged EPG and SDH) is lost and the film desorbs from theelectrode. By contrast, the (polymeric) promoter polylysine formed amore stable (though somewhat less active) enzyme film over theentire pH range investigated and this was used for full range pH-dependent experiments. Under the hydrodynamic electrochemicalconditions employed with efficient electron transfer between theelectrode and the enzyme, the rate limiting event in the catalytic cyclewas the enzyme–sulfite reaction.

The catalytic cyclic voltammetry of SDH under hydrodynamicconditions comprised a sigmoidal waveform in the presence of sulfitecommencing ca. + 100 mV vs NHE (Fig. 2, solid curve) and reaching anapproximate plateau current at high potential (ilim) where the activeMoVI form of the enzyme was continually and rapidly regenerated. Inthe absence of sulfite and under the specific conditions employedhere, no responses from the Mo or heme cofactors were obtained (Fig.2, broken curve) and this voltammogram served as a blank whichcould be subtracted from all other catalytic voltammograms. Ourprevious study using other promoters including a surfactant DDABenabled the resolution of non-turnover signals from the heme andMocofactors [7]. In control experiments, when no enzyme was present,only direct sulfite oxidation was observed at a significantly higherpotential (N +400 mV, Supplementary Fig. S1). These results indicatethat the observed current is due to an electrochemically drivenenzymatic oxidation of sulfite.

In the absence of electrode rotation the SDH catalytic voltammo-gram is distinctly asymmetric (Supplementary Fig. S2, broken line).This is consistent with substrate depletion at the surface of theelectrode [28]. In other words, during the timescale of the sweep, thesubstrate concentration gradient at the electrode decreases due torapid consumption of sulfite by the enzyme and this cannot be com-

pensated bydiffusion alone.When the electrode is rotated at 1000rpm,substrate diffusion is accelerated and a steady state is established. Inthis case the forward and reverse voltammetric sweeps are identical(Fig. 2) when the capacitive component of the current is removed (seealso Supplementary Fig. S2, solid curve). At electrode rotation rateshigher than 500rpm there was no further increase in themagnitude ofthe catalytic current, indicating that sulfite delivery to the enzyme issufficiently rapid that the current is only limited by the potentialdependent enzyme–sulfite reaction.

To ensure that interfacial electron transfer between the enzymeand the electrode was fast, the effect of the voltammetric sweep rateunder electrode rotationwas investigated. The voltammetric responsewas independent of sweep rates between 1 and 20 mV s−1 indicatingthat electron transfer between the electrode and enzyme is facile andthe voltammetric response is not limited by heterogeneous electronexchange between the enzyme and the electrode at these sweep rates.A slight change in the symmetry of the forward and reversevoltammetric sweeps was noted at higher scan rates. Optimalexperimental conditions of 500rpm electrode rotation and 5 mV s−1

Fig. 4. KM,sulfite values determined amperometrically (filled circles) with SDHimmobilised in a polylysine film on an EPG electrode with electrode rotation rate of500 rpm using a buffer mixture of 10 mM bis–tris propane and 10 mM 2-amino-2-methyl-1-propanol. Comparative data from solution assays (open triangles) werecarried out using horse heart cytochrome c as an electron acceptor [29].

1322 T.D. Rapson et al. / Biochimica et Biophysica Acta 1777 (2008) 1319–1325

sweep rate provided conditions whereby the voltammetric responsetruly reflected the characteristics of the enzyme–substrate reaction.

3.2. Catalytic properties: pH and substrate dependence

The effect of pH on the voltammetric response of SDH wasinvestigated whilst keeping the sulfite concentration constant andhigh (Fig. 3A). A pH dependence of −59 mV per pH unit was observedfor the catalytic half-wave potential Ecat that tracks the potential of theMoVI/V couple reported earlier (Fig. 3B) [7,29]. The pH dependence isconsistent with a single-electron, single-proton-coupled redox reac-tion fromHO–MoV to O=MoVI. A similar pH dependencewas seen at allsulfite concentrations investigated (300 μM to 3 mM).

At each pH investigated, the catalytic current at high potentialincreasedwith sequential additions of sulfite and then reached a plateauat high substrate concentrations. The substrate dependence of ilimfollowed Eq. (2) and electrochemical Michaelis constants were obtainedacross the pH range 6–10. The results are presented in Fig. 4 and Table 1in comparisonwith the data determined from solution chemical assayswith horse heart cytochrome c as the electron acceptor [30].

The electrochemical KM,sulfite values for electrode-immobilised SDHwere very close to those seen in solution assays using horse heartcytochrome c as an electron acceptor. In both cases, KM,sulfite increasessteeply above pH 8.5. This characteristic change in substrate affinity athigher pH has been reported in all sulfite oxidizing enzymes to date[31,32]. Deprotonation of a universally conserved Tyr residue (Tyr236)close to the sulfite binding site on the Mo ion has been often associatedwith this rise in KM,sulfite at high pH. However, our recent report of thecrystal structure and enzymology of a Tyr236Phe variant of SDH, whichalso shows a rise in KM,sulfite at high pH despite the absence of anionisable phenol group at the active site, suggests that there may bemore than one deprotonation involved in this change in activity [29].

3.3. Voltammetric waveform

The cyclic voltammetric waveform was analyzed with methodspreviously described by Heering et al. using three parameters; the

Table 1The effect of pH on KM,sulfite (μM) and on the degree of distortion of the ideally sigmoidalwaveform (expressed as the ratio of limiting (high potential) to peak (maximum)currents for SDH within the pH range 7.7 to 8.5

pH KM (μM, at peak potential) KM (μM, at high potential) ilim/ipeak

7.7 12±0.59 (294 mV) 12.5±0.64 (360 mV) 0.9±0.058.0 63±3 (264 mV) 54±3 (345 mV) 0.85±0.058.5 283±26 (244 mV) 269±37 (345 mV) 0.77±0.05

high potential current magnitude (ilim), the number of electronstransferred in the rate defining step (napp), and the catalytic operatingpotential (Ecat) [24]. The voltammetric responses of SDH at 60 and300µM sulfite are shown in Fig. 5A. From Eq. (3), the number ofelectrons transferred in the rate defining step can be calculated fromthe peak width at half height (δ) of the first derivative voltammogram(Fig. 5B). A value of δ ~100 mV (napp=0.9≈1) was found, which did notvary significantly as a function of either substrate concentration or pH.The most important point is that a cooperative 2-electron transfer(napp=2) at the active site can be ruled out. Note that napp is distinctfrom the obligate 2-electron stoichiometry of the sulfite oxidationreaction.

Our previous studies [7,29] have determined that the MoVI/V andMoV/IV redox potentials are well separated and that the MoV form isdominant in the potential range −100 to +150 mV vs NHE. Within thepotential range of the voltammograms presented in Fig. 1, MoIV isnever present (except transiently following turnover). In other wordsonly the components of the FeIII/II couple of the heme and the MoVI/V

couple are addressed within the potential ranges shown. The pHdependence of the catalytic potential and the approximately oneelectron (napp = 0.9) stoichiometry of the rate defining redox reactionsuggests that the MoVI/V couple is being addressed directly. The hemeredox potential is pH independent [7] and if this was the site thatexchanged electrons wewould not expect themarked pH dependenceof Ecat apparent in Fig. 3B.

Fig. 5. (A) The effect of sulfite concentration (60 and 300 μM) on the voltammetricwaveform of SDH immobilised on a kanamycin-modified EPG electrode (pH 8.0,500 rpm, 5 mV/s); (B) the first derivative of the catalytic current (di/dE) with respect toapplied potential at increasing sulfite concentrations. The switch potential (Esw) isindicated.

Table 2Percentage inhibition of sulfite oxidase activity determined amperometrically at 270mVvs NHE with SDH immobilised in a polylysine film (pH 8.0, 300 μM sulfite) andcomparative solution assay data

Inhibitor Electrochemical activity at 100 mMinhibitor concentration

Inhibitor conc. for 50 % inhibitionin solution assays (ref. [19])

SO42− 19% 15 mM

Cl− 30% 50 mMHPO4

2− 18% 20 mMNO3

− 37% 3 mM

1323T.D. Rapson et al. / Biochimica et Biophysica Acta 1777 (2008) 1319–1325

By far the most pronounced and interesting effect apparent withincreasing substrate concentration was the change in voltammogramshape (Fig. 5A). At low sulfite concentrations (b 90 μM) the expectedsigmoidal voltammogram shape was obtained. However, as thesubstrate concentration increased towards saturating levels, thevoltammogram became distinctly peak-shaped i.e. the currentdecreased as the catalytic potential was traversed. This change ismore clearly illustrated by the accompanying first derivative voltam-mogram (Fig. 5B)where a pronounced trough appears at high potential(labelled Esw). Ideally the catalytic current should reach a plateau andremain potential independent thenceforth under steady state condi-tions. Attenuation of enzymatic activity at high potential (at saturatingconcentrations of sulfite) was seen in all cases regardless of thepromoter or buffer that was employed.

From inspection of the first derivative voltammogram (Fig. 5B), thecatalytic potential (Ecat) appears to shift slightly to higher potentialwith increasing sulfite concentrations. However, as the distortion froman ideal sigmoidal wave becomes greater (with increasing sulfiteconcentration), the trough in the high potential region of the firstderivative voltammogram grows in intensity. This distortion of theideally symmetrical peak-shaped first derivative voltammogram leadsto an apparent shift of the maximum to higher potential. In otherwords, the maximum in the first derivative voltammogram, whichwould coincide with Ecat in an ideally sigmoidal voltammogram, doesnot accurately reflect the catalytic potential. Therefore, we have noevidence that Ecat is dependent on sulfite concentration under theseconditions.

Peak-shaped voltammetry under ideally steady state conditionshas been modelled by other groups as a convolution of two ideallysigmoidal voltammograms which are related by a redox switch atpotential Esw [11,12,15,16,33]. In all cases the activity at high over-potential (driving force) was assumed to be lower than at moderateoverpotentials thus leading to amaximumpeak current that decreasedto a plateau as the driving force increased. The potential at which theenzyme switches from high to low activity of Esw is most easilyresolved in the first derivative voltammogram [11,12] (which corre-sponds with the inflection point after the peak current is traversed inthe normal voltammogram).

We have analysed our data in the same manner (Fig. 5B). It isapparent that there is very little change in Esw as sulfite concentrationis raised. However, the sigmoidal waveform becomes more (not less)distorted at higher sulfite concentrations; the exact opposite of thatseen in voltammetry studies of other Mo enzymes [11,16].

The degree of current attenuationwas quantified as the ratio of thecurrent at high potential (ilim) over themaximum(peak) current (ipeak).This ratio ilim/ipeak reached a minimum of 0.77 ± 0.05 at pH 8 (Table 1).The pH dependence of the waveform was also investigated, but thiscould only be done over a relatively narrow range. Below pH 7.7 thecatalytic waveform and direct (non-specific) sulfite oxidation waveoverlap thereby masking any attenuation that might be occurring.Indeed below pH 6 no catalytic current could be seen. This is due to (i)the anodic shift of the catalytic wave close to the non-enzymatic sulfiteoxidation response (Supplementary Fig. S1) and (ii) the inherentlylower activity of the enzyme at pH 6 [7,19], both apparent in Fig. 3A.Above pH 8.5 much larger KM values demanded high sulfiteconcentrations to saturate the enzyme, which also increased inter-ference from the non-enzymatic sulfite oxidation current.

With kanamycin as a promoter for SDH electrochemistry andwithin the range 7.7bpHb8.5, interference from direct sulfiteoxidation on the catalytic waveform was negligible. The KM,sulfite

values in this pH range were compared for the potential regions ofmaximal and attenuated catalytic activity (Table 1). It is apparent thatthere is very little difference between the KM,sulfite values at the peak(optimum) potential or at high (attenuated) potential within the pHrange investigated and so we can only conclude that the origin of thepeak-shaped voltammetry is not due to proton transfer effects at the

active site at least within the pH range that we have been able toprobe. This is distinct from the work of Heffron et al. where significantpH dependence of the catalytic voltammetric waveform was indica-tive of rate limiting protonation events at the active site in E. coliDMSO reductase [12].

3.4. Inhibition effects

The use of inhibitors has helped determine the nature of catalyticattenuation especially in nitrate reductases where azide and thiocya-nate are both competitive inhibitors [15]. Sulfite oxidizing enzymeshave a number of known inhibitors such as phosphate, chloride,nitrate and the enzyme product sulfate [19,34]. The relative effects ofthese anions on SDH activity were investigated (Table 2). Interestingly,all of the anions investigated here were much weaker inhibitors of theelectrode-confined enzyme than has been observed before in solutionassays [19]. Indeed, 50% enzyme inhibition was unattainable atexperimentally practical anion concentrations that did not perturbthe solution ionic strength. Therefore, the percentage inhibition at100 mM concentrations of each inhibitor was determined to enable acomparison to be made. Although the inhibitor anions attenuated thecatalytic current to varying degrees, they had no effect whatsoever onthe voltammetric waveforms and the peak-shaped profiles remained.

4. Discussion

Introducing the potential domain to enzyme catalysis throughprotein film voltammetry has led to a number of interesting andunexpected insights into the analysis of enzyme mechanism [1,2]. Anincreasing number of enzymes do not exhibit the expected idealsigmodial voltammetric response predicted by the steady statedmodels developed by Heering et al. [24]. Of particular interest here areMo enzymes that show an apparent potential-optimised activity,namely the nitrate reductases (NapAB and NarGH) and E. coli DMSOreductase [11,12,16]. In an attempt to explain the origin of theseunexpected voltammetric responses it has been proposed thatsubstrate binding can occur to intermediate redox states of the activesite which themselves are not catalytically competent [4,25].Specifically it has been suggested that MoV may play an importantmechanistic role whereby either substrate or proton binding occurs toMoV in a kinetically favoured pathway [12,16]. A number of modelshave been developed to explain such phenomena, which combinehigh and low activity steady state (sigmoidal) functions linked by aredox event at potential Esw [11,12,15]. Recently Leger and co-workershave extended this model to include other events during the catalyticcycle in an effort tomodel similar complex wave shapes. However, thismodel only applies to low substrate concentrations [4,16]. In thepresent study we have observed the exact opposite phenomenon i.e.distortion from an ideal sigmoidal waveform only emerges at highsubstrate concentrations.

There are a number of reasons why we believe that the results wepresent here for an oxidising Mo enzyme cannot be explained by themodels developed for corresponding reducing Mo enzymes. Theobserved attenuation in enzymatic activity of SDH occurs at apotential (Esw ca. +320 mV vs NHE) much higher than any other

Fig. 6. Simplified catalytic schemes for (A) a typicalMo reductase (nitrate reductase) and(B) Mo oxidase/dehydrogenase (sulfite dehydrogenase). In (A) the two competingpathways for substrate binding with rate constants kV,on (operative close to catalyticpotential) and kIV,on (operative at large overpotential)) are proposed [15] to besufficiently different to lead to the potential dependence of catalysis. In (B) there is nocorresponding rate discrimination as sulfite can only interact with the MoVI form.

1324 T.D. Rapson et al. / Biochimica et Biophysica Acta 1777 (2008) 1319–1325

known redox couple in SDH (e.g. FeIII/II + 177mV vs NHE and MoVI/V +166mV vs NHE at pH 8) [7,29]. This result is in contrast to otherreported cases of peak-shaped Mo enzyme voltammetry where theoptimal potential window for enzymatic activity coincided with theregion in which the MoV active site underwent a change in oxidationstate [12].

All Mo enzymes reported to date that have shown an optimalpotential window of activity are (nitrate or DMSO) reductases. In thiswork we report a similar phenomenon in an oxidising Mo enzyme.This requires a reversal in the order of events at the active site duringcatalysis, as illustrated in Fig. 6. In the reductases, the substratecoordinates to Mo ion prior to turnover and this may involve eitherMoV, with rate kV,on, or MoIV, with rate kIV,on (Fig. 6A). Three differentgroups have each proposed models [11,12,16] of similar origin toexplain the phenomenon of potential optima in Mo enzymevoltammetry. The current consensus view is that substrate bindingto MoV is faster than to MoIV and that at large overpotentials, theslower pathway becomes dominant (MoV is bypassed electrochemi-cally). In SDH (and other sulfite oxidising enzymes) the oxo ligand istransferred to the substrate in what is essentially a concerted 2-electron, O-atom transfer (Eq. (1)). In the present case there is onlyone pathway to the Michaelis complex (Fig. 6B) thus we cannotidentify separate potential dependent, rate limiting events that wouldexplain the attenuation of activity at higher potential. Furthermore theemergence of this feature at high sulfite concentrations is at odds withprevious results of reductase enzymes.

We note that EPG itself exhibits voltammetric responses due toproton coupled oxidation/reduction reactions of phenolic residues onthe surface of the electrode i.e. quinone/hydroquinone couples [35],and these processes will alter the electrostatic and H-bondingproperties of the surface as a function of potential. It is conceivablethat this could affect the orientation of the enzyme on the surfacewhich in turnmayaffect substrate access to the active site or interfacialelectron transfer. At pH 7 the EPG surface redox process appears at ca.+250 mV vs NHE [35] and exhibits a pH dependence of −59 mV/pHunit. Although this potential is in the vicinity of Esw (Fig. 5B), the lack ofa pH dependence of Esw in the present case and the emergence of thepeak-shaped waveform only at high sulfite concentrations is incon-sistent with the observed attenuation in activity.

It is interesting to note that SDH co-adsorbed with eitherkanamycin or polylysine is much less susceptible to inhibition byanions than has been observed for the freely diffusing enzyme insolution assays [19]. It is possible that these inhibitor anions are notbinding to theMo at the active site but insteadmay be interferingwiththe enzyme's interaction with cytochrome c in the solution. Thisproposal supports the observations from EPR spectroscopy thatrevealed the MoV signal is unaffected by anions such as chloride andphosphate [19], unlike other sulfite oxidising enzymes studied.

In conclusion, the bacterial sulfite dehydrogenase we have studiedhere is an excellentmodel enzyme for probing electron transfer eventsoccurring at the active site of sulfite oxidizing enzymes. The catalyticbehaviour of SDH immobilised on an EPG electrode within apolylysine film mirrors that seen in solution assays suggesting thatthe enzyme retains its native activity. The voltammetric response isnot affected by slow intramolecular electron transfer and under theoptimised experimental conditions substrate diffusion limitationshave been overcome. We have not been able to determine the cause ofthe unusual peak-shaped voltammetry at high concentrations ofsulfite. However, we have been able to conclude that MoV does notplay a rate defining role in the catalytic mechanism of SDH beforeturnover. This result is important given the growing number of Moenzymes which have been suggested to follow a mechanism in whichMoV intermediates have been implicated. Our voltammetric study hasrevealed some interesting inhibition differences between the immo-bilised enzyme and the enzyme in solution. To try to determine theorigin of the potential optimal behaviour of SDH we are investigatingthe direct electrochemistry of SDH variants with changes in key aminoacid residues required for substrate binding and intramolecularelectron transfer.

Acknowledgements

We wish to thank the Australian Research Council for grantsupport (PVB: DP0343405 and DP0880288 and UK: DP0878525) andthe University of Queensland for financial support of this work. TDRacknowledges the award of an Endeavour International PostgraduateResearch Scholarship.

Appendix A. Supplementary data

Supplementary data associated with this article can be found, inthe online version, at doi:10.1016/j.bbabio.2008.06.005.

References

[1] C. Leger, S.J. Elliott, K.R. Hoke, L.J.C. Jeuken, A.K. Jones, F.A. Armstrong, Enzymeelectrokinetics: using protein film voltammetry to investigate redox enzymes andtheir mechanisms, Biochemistry 42 (2003) 8653–8662.

[2] P.V. Bernhardt, Enzyme electrochemistry — biocatalysis on an electrode, Aust. J.Chem. 59 (2006) 233–256.

[3] F.A. Armstrong, G.S. Wilson, Recent developments in faradaic bioelectrochemistry,Electrochim. Acta 45 (2000) 2623–2645.

[4] P. Bertrand, B. Frangioni, S. Dementin, M. Sabaty, P. Arnoux, B. Guigliarelli, D.Pignol, C. Leger, Effects of slow substrate binding and release in redox enzymes:theory and application to periplasmic nitrate reductase, J. Phys. Chem. B 111(2007) 10300–10311.

[5] K.-F. Aguey-Zinsou, P.V. Bernhardt, A.G. McEwan, J.P. Ridge, The first non-turnovervoltammetric response from a molybdenum enzyme: direct electrochemistry ofdimethylsulfoxide reductase from Rhodobacter capsulatus, J. Biol. Inorg. Chem. 7(2002) 879–883.

[6] K.F. Aguey-Zinsou, P.V. Bernhardt, S. Leimkuehler, Protein film voltammetry ofRhodobacter Capsulatus xanthine dehydrogenase, J. Am. Chem. Soc. 125 (2003)15352–15358.

[7] K.-F. Aguey-Zinsou, P.V. Bernhardt, U. Kappler, A.G. McEwan, Direct electrochem-istry of a bacterial sulfite dehydrogenase, J. Am. Chem. Soc. 125 (2003) 530–535.

[8] P.V. Bernhardt, J.M. Santini, Protein film voltammetry of arsenite oxidase from thechemolithoautotrophic arsenite-oxidizing bacterium NT-26, Biochemistry 45(2006) 2804–2809.

[9] P.V. Bernhardt, M.J. Honeychurch, A.G. McEwan, Direct electrochemically drivencatalysis of bovine milk xanthine oxidase, Electrochem. Commun. 8 (2006)257–261.

1325T.D. Rapson et al. / Biochimica et Biophysica Acta 1777 (2008) 1319–1325

[10] L.J. Anderson, D.J. Richardson, J.N. Butt, Using direct electrochemistry to probe ratelimiting events during nitrate reductase turnover, Faraday Discuss. 116 (2000)155–169.

[11] L.J. Anderson, D.J. Richardson, J.N. Butt, Catalytic protein film voltammetry from arespiratory nitrate reductase provides evidence for complex electrochemicalmodulation of enzyme activity, Biochemistry 40 (2001) 11294–11307.

[12] K. Heffron, C. Leger, R.A. Rothery, J.H. Weiner, F.A. Armstrong, Determination of anoptimal potential window for catalysis by E. coli dimethyl sulfoxide reductase andhypothesis on the role of Mo(V) in the reaction pathway, Biochemistry 40 (2001)3117–3126.

[13] S.J. Elliott, A.E. McElhaney, C. Feng, J.H. Enemark, F.A. Armstrong, A voltammetricstudy of interdomain electron transfer within sulfite oxidase, J. Am. Chem. Soc. 124(2002) 11612–11613.

[14] E.E. Ferapontova, T. Ruzgas, L. Gorton, Direct electron transfer of heme- andmolybdopterin cofactor-containing chicken liver sulfite oxidase on alkanethiol-modified gold electrodes, Anal. Chem. 75 (2003) 4841–4850.

[15] S.J. Elliott, K.R. Hoke, K. Heffron, M. Palak, R.A. Rothery, J.H. Weiner, F.A.Armstrong, Voltammetric studies of the catalytic mechanism of the respiratorynitrate reductase from Escherichia coli: how nitrate reduction and inhibitiondepend on the oxidation state of the active site, Biochemistry 43 (2004)799–807.

[16] B. Frangioni, P. Arnoux, M. Sabaty, D. Pignol, P. Bertrand, B. Guigliarelli, C. Leger, InRhodobacter sphaeroides respiratory nitrate reductase, the kinetics of substratebinding favors intramolecular electron transfer, J. Am. Chem. Soc 126 (2004)1328–1329.

[17] R. Hille, The mononuclear molybdenum enzymes, Chem. Rev 96 (1996)2757–2816.

[18] C. Feng, G. Tollin, J.H. Enemark, Sulfite oxidizing enzymes, Biochim. Biophys. Acta,Proteins and Proteomics 1774 (2007) 527–539.

[19] U. Kappler, B. Bennett, J. Rethmeier, G. Schwarz, R. Deutzmann, A.G. McEwan, C.Dahl, Sulfite: Cytochrome c oxidoreductase from Thiobacillus novellus. Purification,characterization, and molecular biology of a heterodimeric member of the sulfiteoxidase family, J. Biol. Chem 275 (2000) 13202–13212.

[20] U. Kappler, S. Bailey, Molecular basis of intramolecular electron transfer in sulfite-oxidizing enzymes is revealed by high resolution structure of a heterodimericcomplex of the catalytic molybdopterin subunit and a c-type cytochrome subunit,J. Biol. Chem 280 (2005) 24999–25007.

[21] U. Kappler, S. Bailey, Crystallization and preliminary X-ray analysis of sulfitedehydrogenase from Starkeya novella, Acta Crystallographica, Section D: BiologicalCrystallography D60 (2004) 2070–2072.

[22] C. Kisker, H. Schindelin, A. Pacheco, W.A. Wehbi, R.M. Garrett, K.V. Rajagopalan, J.H. Enemark, D.C. Rees, Molecular basis of sulfite oxidase deficiency from thestructure of sulfite oxidase, Cell 91 (1997) 973–983.

[23] N. Schrader, K. Fischer, K. Theis, R.R. Mendel, G. Schwarz, C. Kisker, The crystalstructure of plant sulfite oxidase provides insights into sulfite oxidation in plantsand animals, Structure 11 (2003) 1251–1263.

[24] H.A. Heering, J. Hirst, F.A. Armstrong, Interpreting the catalytic voltammetry ofelectroactive enzymes adsorbed on electrodes, J. Phys. Chem. B 102 (1998)6889–6902.

[25] F.A. Armstrong, Recent developments in dynamic electrochemical studies ofadsorbed enzymes and their active sites, Curr. Opin. Chem. Biol. 9 (2005) 110–117.

[26] U. Kappler, A.G. McEwan, A system for the heterologous expression of complexredox proteins in Rhodobacter capsulatus: characterisation of recombinantsulphite:cytochrome c oxidoreductase from Starkeya novella, FEBS Lett. 529(2002) 208–214.

[27] W. Szczepanik, P. Kaczmarek, J. Sobczak, W. Bal, K. Gatner, M. Jezowska-Bojczuk,Copper(II) binding by kanamycin A and hydrogen peroxide activation by resultingcomplexes, New J. Chem. 26 (2002) 1507–1514.

[28] A.J. Bard, L.R. Faulkner, Electrochemical methods: fundamentals and applications1982.

[29] U. Kappler, S. Bailey, C. Feng, M.J. Honeychurch, G.R. Hanson, P.V. Bernhardt, G.Tollin, J.H. Enemark, Kinetic and structural evidence for the importance of Tyr236for the integrity of the Mo active site in a bacterial sulfite dehydrogenase,Biochemistry 45 (2006) 9696–9705.

[30] S. Bailey, T.D. Rapson, K. Johnson-Winters, J.H. Enemark, U. Kappler, MolecularBasis For Enzymatic sulfite oxidation — how three conserved active site residuesshape enzyme activity, J. Biol. Chem. (2008) accepted.

[31] M.S. Brody, R. Hille, The kinetic behavior of chicken liver sulfite oxidase,Biochemistry 38 (1999) 6668–6677.

[32] H.L. Wilson, K.V. Rajagopalan, The role of tyrosine 343 in substrate binding andcatalysis by human sulfite oxidase, J. Biol. Chem. 279 (2004) 15105–15113.

[33] S.J. Elliott, C. Leger, H.R. Pershad, J. Hirst, K. Heffron, N. Ginet, F. Blasco, R.A.Rothery, J.H. Weiner, F.A. Armstrong, Detection and interpretation of redoxpotential optima in the catalytic activity of enzymes, Biochim. Biophys. Acta,Bioenergetics 1555 (2002) 54–59.

[34] D.L. Kessler, K.V. Rajagopalan, Hepatic sulfite oxidase. Effect of anions oninteraction with cytochrome, Biochim. Biophys. Acta, Enzymology 370 (1974)389–398.

[35] J.F. Evans, T. Kuwana, Radiofrequency oxygen plasma treatment of pyrolyticgraphite electrode surfaces, Anal. Chem. 49 (1977) 1632–1635.