electrochemicalmeasurementofelectrontransferkinetics by ... · when peak heights had stabilized...

10
Electrochemical Measurement of Electron Transfer Kinetics by Shewanella oneidensis MR-1 * Received for publication, July 10, 2009, and in revised form, August 4, 2009 Published, JBC Papers in Press, August 6, 2009, DOI 10.1074/jbc.M109.043455 Daniel Baron , Edward LaBelle , Dan Coursolle § , Jeffrey A. Gralnick ‡¶ , and Daniel R. Bond ‡¶1 From the BioTechnology Institute, the § Department of Biochemistry, Biophysics, and Molecular Biology, and the Department of Microbiology, University of Minnesota, St. Paul, Minnesota 55108 Shewanella oneidensis strain MR-1 can respire using carbon electrodes and metal oxyhydroxides as electron acceptors, requiring mechanisms for transferring electrons from the cell interior to surfaces located beyond the cell. Although purified outer membrane cytochromes will reduce both electrodes and metals, S. oneidensis also secretes flavins, which accelerate elec- tron transfer to metals and electrodes. We developed tech- niques for detecting direct electron transfer by intact cells, using turnover and single turnover voltammetry. Metabolically active cells attached to graphite electrodes produced thin (submono- layer) films that demonstrated both catalytic and reversible electron transfer in the presence and absence of flavins. In the absence of soluble flavins, electron transfer occurred in a broad potential window centered at 0V(versus standard hydrogen electrode), and was altered in single (omcA, mtrC) and dou- ble deletion (omcA/mtrC) mutants of outer membrane cyto- chromes. The addition of soluble flavins at physiological concen- trations significantly accelerated electron transfer and allowed catalytic electron transfer to occur at lower applied potentials (0.2 V). Scan rate analysis indicated that rate constants for direct electron transfer were slower than those reported for pure cyto- chromes (1s 1 ). These observations indicated that anodic cur- rent in the higher (>0 V) window is due to activation of a direct transfer mechanism, whereas electron transfer at lower potentials is enabled by flavins. The electrochemical dissection of these activ- ities in living cells into two systems with characteristic midpoint potentials and kinetic behaviors explains prior observations and demonstrates the complementary nature of S. oneidensis electron transfer strategies. Respiratory electron flow typically occurs at the inner mem- brane, where oxidation and reduction can be easily linked to intracellular electron carriers and used to generate a membrane potential. However, when the electron acceptor or donor is insoluble, bacteria must possess a mechanism for transferring electrons beyond their inner membrane (1). This is especially true for Proteobcteria, which have an outer membrane that fur- ther insulates cytoplasmic and inner membrane processes from insoluble substrates. Metal oxides (such as Fe(III) and Mn(IV) oxyhydroxides) are well recognized naturally occurring elec- tron acceptors that demand such an electron transfer strategy (2– 4). Shewanella oneidensis MR-1, a metabolically versatile mem- ber of the gammaproteobacteria (5), is capable of reducing insoluble metals, and this phenotype has been linked to a col- lection of interacting multiheme cytochromes spanning the inner membrane, periplasmic space, and outer membrane (6 –12). There is also evidence that some of these cytochromes decorate the surface of pili-like structures extending from the cell surface (13, 14). Regardless of the ultimate location of the cytochromes, in all models of electron transfer, electrons must hop from these proteins to a solid surface or be transferred to a soluble mediator that can diffuse to a final destination (15, 16). Although chelation of a metal oxide is a third option (17, 18), the fact that Shewanella is able to transfer electrons to solid graphite electrodes (19 –23) underscores the need for a direct or diffusion-based electron transfer mechanism to link cellular proteins and surfaces. Recent work has shown that Shewanella species secrete sol- uble flavins (FMN and riboflavin) that facilitate electron trans- fer to both metals and electrodes (23, 24). For example, removal of accumulated soluble flavins decreases the rate of electron transfer by Shewanella biofilms to electrodes over 80%. Con- sistent with this observation, kinetic measurements with pure MtrC and OmcA (25) showed that direct reduction of solid metal oxides by these cytochromes was too slow to explain physiological rates of electron transfer, whereas turnover rates of these enzymes with soluble flavins were orders of magnitude larger. These studies suggest that the kinetics of electron trans- fer from cytochromes on the outer surface of Shewanella to electrodes will be significantly altered in the absence of diffus- ible mediators (9 –11, 26 –34). Voltammetry has proven a useful technique for the analysis of electron transfer rates and pathways using purified proteins (35–39) and has recently been extended to the study of intact bacteria (23, 40 – 42). In slow scan rate cyclic voltammetry, the rate of electron transfer from respiring Shewanella biofilms to electrodes rises sharply at the E° of riboflavin and FMN (0.2 V versus SHE) 2 (23). Such measurements relating thermody- namic driving force to turnover kinetics would be difficult with whole cell:Fe(III) oxide incubations, which do not allow fine control over the electron acceptor redox potential or real time recording of electron transfer rates. In addition, voltammetry provides tools to observe electron movement under single- * This work was supported, in whole or in part, by National Institutes of Health Grant 2T32-GM008347-16 (to D. C.). This work was also supported by Office of Naval Research Grants N000140810166 (to J. A. G.) and N000140810162 (to D. R. B.). 1 To whom correspondence should be addressed: BioTechnology Institute, University of Minnesota, 140 Gortner, 1479 Gortner Ave, St. Paul, MN 55108. Tel.: 612-624-8619; Fax: 612-625-1700; E-mail:: [email protected]. 2 The abbreviation used is: SHE, standard hydrogen electrode. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 284, NO. 42, pp. 28865–28873, October 16, 2009 © 2009 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A. OCTOBER 16, 2009 • VOLUME 284 • NUMBER 42 JOURNAL OF BIOLOGICAL CHEMISTRY 28865 by guest on January 13, 2021 http://www.jbc.org/ Downloaded from

Upload: others

Post on 21-Sep-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

Electrochemical Measurement of Electron Transfer Kineticsby Shewanella oneidensis MR-1*

Received for publication, July 10, 2009, and in revised form, August 4, 2009 Published, JBC Papers in Press, August 6, 2009, DOI 10.1074/jbc.M109.043455

Daniel Baron‡, Edward LaBelle‡, Dan Coursolle§, Jeffrey A. Gralnick‡¶, and Daniel R. Bond‡¶1

From the ‡BioTechnology Institute, the §Department of Biochemistry, Biophysics, and Molecular Biology, and the ¶Department ofMicrobiology, University of Minnesota, St. Paul, Minnesota 55108

Shewanella oneidensis strain MR-1 can respire using carbonelectrodes and metal oxyhydroxides as electron acceptors,requiring mechanisms for transferring electrons from the cellinterior to surfaces located beyond the cell. Although purifiedouter membrane cytochromes will reduce both electrodes andmetals, S. oneidensis also secretes flavins, which accelerate elec-tron transfer to metals and electrodes. We developed tech-niques for detecting direct electron transfer by intact cells, usingturnover and single turnover voltammetry. Metabolically activecells attached to graphite electrodes produced thin (submono-layer) films that demonstrated both catalytic and reversibleelectron transfer in the presence and absence of flavins. In theabsence of soluble flavins, electron transfer occurred in a broadpotential window centered at �0 V (versus standard hydrogenelectrode), and was altered in single (�omcA, �mtrC) and dou-ble deletion (�omcA/�mtrC) mutants of outer membrane cyto-chromes. The addition of soluble flavins at physiological concen-trations significantly accelerated electron transfer and allowedcatalytic electron transfer to occur at lower applied potentials(�0.2V). Scan rate analysis indicated that rate constants for directelectron transfer were slower than those reported for pure cyto-chromes (�1 s�1). These observations indicated that anodic cur-rent in the higher (>0 V) window is due to activation of a directtransfer mechanism, whereas electron transfer at lower potentialsis enabled by flavins. The electrochemical dissectionof these activ-ities in living cells into two systems with characteristic midpointpotentials and kinetic behaviors explains prior observations anddemonstrates the complementary nature of S. oneidensis electrontransfer strategies.

Respiratory electron flow typically occurs at the inner mem-brane, where oxidation and reduction can be easily linked tointracellular electron carriers and used to generate amembranepotential. However, when the electron acceptor or donor isinsoluble, bacteria must possess a mechanism for transferringelectrons beyond their inner membrane (1). This is especiallytrue for Proteobcteria, which have an outer membrane that fur-ther insulates cytoplasmic and innermembrane processes frominsoluble substrates. Metal oxides (such as Fe(III) and Mn(IV)

oxyhydroxides) are well recognized naturally occurring elec-tron acceptors that demand such an electron transfer strategy(2–4).Shewanella oneidensisMR-1, a metabolically versatile mem-

ber of the gammaproteobacteria (5), is capable of reducinginsoluble metals, and this phenotype has been linked to a col-lection of interacting multiheme cytochromes spanning theinner membrane, periplasmic space, and outer membrane(6–12). There is also evidence that some of these cytochromesdecorate the surface of pili-like structures extending from thecell surface (13, 14). Regardless of the ultimate location of thecytochromes, in all models of electron transfer, electrons musthop from these proteins to a solid surface or be transferred to asoluble mediator that can diffuse to a final destination (15, 16).Although chelation of a metal oxide is a third option (17, 18),the fact that Shewanella is able to transfer electrons to solidgraphite electrodes (19–23) underscores the need for a director diffusion-based electron transfer mechanism to link cellularproteins and surfaces.Recent work has shown that Shewanella species secrete sol-

uble flavins (FMN and riboflavin) that facilitate electron trans-fer to bothmetals and electrodes (23, 24). For example, removalof accumulated soluble flavins decreases the rate of electrontransfer by Shewanella biofilms to electrodes over 80%. Con-sistent with this observation, kinetic measurements with pureMtrC and OmcA (25) showed that direct reduction of solidmetal oxides by these cytochromes was too slow to explainphysiological rates of electron transfer, whereas turnover ratesof these enzymes with soluble flavins were orders of magnitudelarger. These studies suggest that the kinetics of electron trans-fer from cytochromes on the outer surface of Shewanella toelectrodes will be significantly altered in the absence of diffus-ible mediators (9–11, 26–34).Voltammetry has proven a useful technique for the analysis

of electron transfer rates and pathways using purified proteins(35–39) and has recently been extended to the study of intactbacteria (23, 40–42). In slow scan rate cyclic voltammetry, therate of electron transfer from respiring Shewanella biofilms toelectrodes rises sharply at the E°� of riboflavin and FMN (�0.2V versus SHE)2 (23). Such measurements relating thermody-namic driving force to turnover kinetics would be difficult withwhole cell:Fe(III) oxide incubations, which do not allow finecontrol over the electron acceptor redox potential or real timerecording of electron transfer rates. In addition, voltammetryprovides tools to observe electron movement under single-

* This work was supported, in whole or in part, by National Institutes of HealthGrant 2T32-GM008347-16 (to D. C.). This work was also supported by Officeof Naval Research Grants N000140810166 (to J. A. G.) and N000140810162(to D. R. B.).

1 To whom correspondence should be addressed: BioTechnology Institute,University of Minnesota, 140 Gortner, 1479 Gortner Ave, St. Paul, MN55108. Tel.: 612-624-8619; Fax: 612-625-1700; E-mail:: [email protected]. 2 The abbreviation used is: SHE, standard hydrogen electrode.

THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 284, NO. 42, pp. 28865–28873, October 16, 2009© 2009 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.

OCTOBER 16, 2009 • VOLUME 284 • NUMBER 42 JOURNAL OF BIOLOGICAL CHEMISTRY 28865

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 2: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

turnover conditions (in the absence of electron donor), allow-ing reversible oxidation and reduction of proteins accessible tothe electrode and study of kinetic behavior (43, 44).In this work, techniques of turnover (sustained electron

transfer from cells to electrode in the presence of electrondonor) and single turnover (reversible oxidation and reductionin the absence of electron donor) voltammetry were harnessedto investigate the role of outer membrane proteins in electrontransfer from Shewanella to electrodes. In all of these studies,intact metabolically active cells were used, along with electrodesurfaces known to act as acceptors for Shewanella. The resultsin the absence of soluble mediators provide evidence that elec-tron transfer betweenMtrC and OmcA and surfaces requires ahigher potential compared with when flavins are present toshuttle electrons to the surface. Mutant analysis also demon-strates that cells possessing different outer membrane cyto-chromes have differing abilities for direct and mediator-en-abled electron transfer.

EXPERIMENTAL PROCEDURES

Strains and Growth Conditions—For each trial, S. oneidensisMR-1 (ATCC 70050) or mutant cultures were grown from afrozen stock, first aerobically in Shewanella basal medium(described below) with 30 mM lactate acting as the electrondonor. After 12 h, the culture was transferred into 0.4 liters ofanaerobic Shewanella basal medium, containing (per liter)0.46 g of NH4Cl, 0.225 g of K2HPO4, 0.225 g of KH2PO4, 0.117g ofMgSO4�7H2O, and 0.225 g of (NH4)2SO4. Prior to autoclav-ing, 5 ml of a mineral mix (containing per liter 1.5 g of nitrilo-triacetic acid, 0.1 g ofMnCl2�4H2O, 0.3 g of FeSO4�7H2O, 0.17 gofCoCl2�6H2O, 0.1 g of ZnCl2, 0.04 g ofCuSO4�5H2O, 0.005 g ofAlK(SO4)2�12H2O, 0.005 g ofH3BO3, 0.09 g ofNa2MoO4, 0.12 gofNiCl2, 0.02 g ofNaWO4�2H2O, and 0.10 g ofNa2SeO4) aswellas 5 ml of a vitamin mix (containing per liter 0.002 g of biotin,0.002 g of folic acid, 0.02 g of pyridoxineHCl, 0.005 of thiamine,0.005 g of nicotinic acid, 0.005 g of pantothenic acid, 0.0001 g ofB-12, 0.005 g of p-aminobenzoic acid, and 0.005 g of thiocticacid)were added.Tobuffer against pHchanges, 100mMHEPESwas added, and 15 mM lactate and 40 mM fumarate were pro-vided as the electron donor and electron acceptor, respectively.Medium was sparged with N2 gas that had been passed over aheated copper column to remove oxygen, adjusted to pH 7.2,and autoclaved. After autoclaving, 0.05% (w/v) filter sterilizedcasamino acids were also added.To prepare thin films of attached cells, bacteria were grown

to stationary phase (A600 between 0.5 and 0.6), which repre-sented depletion of the electron donor. The anaerobic culturewas then transferred to centrifuge tubes in an anaerobic cham-ber (CO2:N2:H2 20:75:5) and centrifuged at 5000 rpm for 20min. In an anaerobic chamber, the pellet was gently resus-pended in 20 ml of Shewanella basal medium that did not con-tain electron donor, electron acceptor, or casamino acids. Thecells were centrifuged and resuspended again and used to inoc-ulate sterile, anaerobic electrochemical reactors in an anaero-bic glove bag. The electrochemical cells were incubated at 30 °Cunder a stream of nitrogen gas. Base-line voltammetry experi-ments were performed, and electrodes were poised at �0.24 Vversus SHE for at least 6 h to facilitate attachment to electrodes

and further depletion of intracellular donors (see “Electro-chemistry of Attached Films” below for details on analysis offilms). When cells were allowed to colonize electrodes over thecourse of several days, similarly grown cultures were directlyinoculated into the same electrochemical chambers withoutanywashing steps, and electrodeswere poised at�0.24V versusSHE continuously.Deletion Mutants—S. oneidensis MR-1 mutant strains con-

tainingmarkerless deletion constructs formtrC and omcAwerecreated as described inRef. 45. Briefly, upstreamanddownstreamregions of a targeted gene were ligated in a pSMV-3 plasmid con-taining counter-selectable markers, which was subsequentlytransformed into an Escherichia coliWM3064mating strain. Fol-lowing conjugation between the mating strain and S. oneidensisMR-1 wild type cells, recombination replaced the targeted genewith the deletion construct. Incubation under conditions favoringloss of the insert produced strains with deletions in the desiredregion. Mutations were verified using PCR with oligonucleotideprimers that spanned the deleted region, and proper expression ofnearby genes was verified by quantitative PCR.Electrochemical Techniques—5X-AQcarbon (PocoGraphite

Company, Decatur, TX) was used as the working electrode.Electrodes machined to 0.5 cm � 2 cm � 1 mmwere preparedby polishingwith 400 grit paper, rinsed, and sonicated in deion-ized water. The electrodes were cleaned for 24 h in 1 MHCl andstored in deionized water. 0.22-mm-thick platinum wire wasattached to the electrodes via a 3.5-mm-long nylon screw(Small Parts Inc., Miramar, FL) and a nylon nut. The platinumwire was soldered to a length of insulated copper wire within a3-mm-diameter glass capillary tube to provide a sealed outletextending from the electrochemical cell. The resistancebetween the carbon electrode and the copper wire was alwaysless than 0.5 �.The electrochemical reactors used in this study were 25-ml

glass cones (Bioanalytical Systems, West Lafayette, IN). A Tef-lon top was altered to contain a butyl rubber O-ring (limitingoxygen entry) and a series of holes to allow working, counter,reference electrodes, and gas lines to be introduced. A glassbody saturated calomel electrode (Fisher) was used as the ref-erence electrode. A 5-mm-diameter glass capillary tube wascapped with a nanoporous vycor frit (Bioanalytical Systems,West Lafayette, IN) connected to a larger diameter tube filledwith 1% agar in 0.1 M sodium sulfate solution as a salt bridge.Electrochemistry of Attached Films—After placing Shewanella

suspensions in anaerobic reactors with electrodes, the workingelectrode was poised at �0.24 V versus SHE, using a Bio-LogicVMP� Multichannel Potentiostat. During growth and duringcatalytic voltammetry, reactors were stirred with a magneticstir bar (Ax-Man, St. Paul, MN). Chronoamperometry wasmonitored until current levels fell to near background (�1�A/cm2) to ensure that cells attached and were depleted ofelectron donors. At various time points, chronoamperometrywas halted, and cyclic voltammetry (CV) was performed, untilpeaks in CV analysis became consistent. CV sweeps were per-formed from �0.56 to 0.45 V (versus SHE). The scan directionwas then reversed and swept to the original value of �0.56 V.CV was performed at a scan rate of 1.0 mV/s.

Direct Electron Transfer by Shewanella

28866 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284 • NUMBER 42 • OCTOBER 16, 2009

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 3: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

When peak heights had stabilized (typically within 6 h),working electrodes were washed to remove loosely attachedcells by replacing the medium in the reactor. The medium wasremoved with a sterile nitrogen-flushed syringe and replacedwith 5ml of fresh anaerobicmediumwithout electron donor oracceptor. After a brief period of stirring, the replacement proc-ess with fresh medium was performed again. Preliminary CVanalysis after eachwash showed that no changes in peak heightswere observed after a total of five washes, and flavin peaks werealso undetectable at this point. Three complete cycles of singleturnover slow scan rate CV data were collected at 1 mV/s foreach film. Scan rate analysis was then performed on cell filmswith a Gamry PCI4 Femtostat (Gamry Instruments, Warmin-ster, PA). All of the sweeps were performed from�0.56 to 0.8 V(versus SHE) with a scan rate of 1–1000mV/s with two sweeps/scan rate. Following scan rate analysis, the electrodes were dis-carded. The data were analyzed (e.g. base-line removal, deriva-tive analysis) using SOAS software (recently described in Ref.46). Rate constants were also estimated using the program Jel-lyfit (courtesy of L. Jeuken).Microscopy—Electrodes at various stages were placed in

amber 1.5-ml centrifuge tubes containing 1 ml of mediumwith60 �M propidium iodide and 10 �M Syto 9 from a Live/DeadBacLight bacterial viability kit (Invitrogen). The film wasstained for 20 min, rinsed, and placed in the center of a micro-scope slide surrounded by a gasket slightly thicker than theelectrode. The basinwas filledwith 50�l ofmedium, and a largecoverslip was sealed in place (ColorStay, RevlonCosmetics, LosAngeles, CA). The stained filmwas viewedwith aNikon EclipseC1 confocal microscope using 488- and 561-nm filters. Themorphology and coverage of the biofilm were evaluated at sev-eral locations, and representative sites were chosen for the col-lection of three-dimensional images using 0.5-�m intervals,beginning 20 �m above the top of the film and continuing until20 �m beyond the electrode/biofilm interface. The resultingsliceswere rendered using EZ-C1 3.20 FreeViewer software andImageJ 1.34s.A S3500N scanning electron microscope (Hitachi, Japan)

was also used to image freshly polished electrodes and elec-trodes with attached cells. The electrodes were fixed in 2.5%glutaraldehyde and dehydrated for 5 min in 25, 50, 75, 95, and100% ethanol. The electrodes then were dried in a DentonDV-502 vacuum evaporator (Moorestown, New Jersey), sput-ter-coated (Fullam EffaCoater, Clifton Park, New York), andfixed through graphite tape to a sample stage. The sampleswereplaced directly into the scanning electron microscope vacuumcompartment and examined at accelerating voltage of 5 kV.

RESULTS

Evidence for Two Potential Ranges for Electron Transfer inGrowing Shewanella Biofilms—It was recently discovered thatShewanella strains secrete redox-active flavins (24), which canaccelerate the rate that electrons are transferred from cells tometals and electrodes (23, 24). However, other observationsshowing interactions between purified cytochromes and solidelectron acceptors suggested that a mechanism independent offlavins could also be functional. To characterize electron trans-fer across a range of applied potentials by S. oneidensis MR-1

under growth conditions, cultures were first grown on 2-cm2

planar electrodes held at oxidizing potentials (�0.2 V versusSHE) and analyzed by cyclic voltammetry (Fig. 1).The concentration of soluble flavins in a S. oneidensisMR-1

culture can vary from 0.2 to 1�M, because of factors such as celldensity and the length of incubation (23, 24). To initiate growthin the presence of varying flavin concentrations, biofilms wereamended with increasing levels of flavins. For example, whenlate exponential phase cultures of S. oneidensis MR-1 wereused, flavin concentrations in the bioreactor were 0.1–0.2 �M

(FMN plus riboflavin). Over the next 96 h, anodic current rose(doubling nearly every 24 h), and leveled off at 35.6 �A � 6.1(n 3). When flavins were supplemented at inoculation, therate of increase was faster (doubling time approaching 12 h for1 �M added flavins), and current stabilized at levels over 3-foldhigher (110 �A � 5.1, n 3).

Representative cyclic voltammetry data for biofilms grown inthe presence of increasing levels of flavins are shown in Fig. 1A,with the derivative of each voltammogram shown in Fig. 1B, toallow visualization of midpoint potentials and wave symmetryin catalytic data. Voltammetry of all biofilms was dominated bya catalytic wave that initiated by �0.25 V and rose steeply at amidpoint potential of �0.2 V. The magnitude of this responseincreased with flavin level (23).Interfacial electron transfer increases exponentially with the

potential difference between donors and acceptors. Thus, smallchanges near the midpoint potential have a large effect on elec-tron transfer rate, but as potential rises, reactions responsiblefor bringing electrons to the interface approach a limit. Thesharp rise in the rate of electron transfer, centered at the mid-

FIGURE 1. A, cyclic voltammetry (1 mV/s) of S. oneidensis MR-1 after 96 h ofgrowth on planar 2-cm2 electrodes at �0. 24 V versus SHE. After the additionof cells, the initial levels of flavins were 0.2 �M, unless riboflavin was providedat inoculation as shown. B, first derivative plots of data in A, showing midpointpotentials of initial potential increase and similar slopes in higher potentialregions.

Direct Electron Transfer by Shewanella

OCTOBER 16, 2009 • VOLUME 284 • NUMBER 42 JOURNAL OF BIOLOGICAL CHEMISTRY 28867

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 4: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

point potential of flavins, provided an example of such behaviorcatalyzed by Shewanella.

However, voltammograms always demonstrated a secondaryresponse to applied potential, as shown by the slow, nearly lin-ear rise in anodic current at potentials above �0.1 V in allexperiments (Fig. 1A). Themagnitude of this boost was similar,because current rose 10–15�Aas potential was increased from�0.1 to �0.25 V. Although flavins altered electron transfer atthe lower potential window (near �0.2 V), the higher potentialincrease always remained and was not sensitive to flavin con-centrations. This suggested a second mechanism that wasactive at higher potentials, was not dependent upon flavins, andwas broader in its response to potential.To further characterize the effect flavins exerted on reaction

rates inwhole cells, Fig. 2 shows the relationship between addedriboflavin and specific rates of electron transfer per unit ofattached biomass (as �A/�g protein, 1 �A nearly 10 pmol ofelectrons/s). Soluble mediators increased the amount of cellbiomass that was able to colonize electrodes but also increasedthe specific rate each unit of biomass could transfer electrons tothe surface. This further demonstrated that Shewanella pos-sessed an excess of lactate oxidation activity and that higherspecific rates of electron transfer per cell were achieved whenriboflavin was also present to shuttle electrons to the surface.Adsorption of Cells to Electrode Surfaces for the Study of

Direct Transfer—The remainder of the experiments wereaimed at explaining the origin of the higher potential electrontransfer behavior, which was not affected by the level of solubleflavins. Tomore precisely study wild type andmutant strains, amethod was developed to obtain repeatable electrodes inter-faced with cell films. Identical electrodes were exposed to S.oneidensis that had been grown under anaerobic conditions.When an electrode (poised at�0.24 versus SHE) was immersedin a cell suspension of S. oneidensis MR-1 lacking electronacceptors, cyclic voltammetry detected reversible peaks thatstabilized in terms of height and location within 6 h. Whensimilar experiments were performed with chloramphenicol-treated cells, identical peak development over time wasobserved (data not shown), ruling out synthesis of new proteinsas a significant cause of the new voltammetry features.

Confocal microscopy of electrodes revealed a dense, 1–3-�m-thick layer of cells initially attached to surfaces (Fig. 3).However, when electrodes were washed five times (via fivechanges of anaerobic cell-free electrolyte), voltammetry fea-tures became lower in magnitude but more defined. Confocalmicroscopy of these washed electrodes revealed a thinning ofthe adsorbed cell layer, with individual cells and microcolonies(Fig. 3). Scanning electron microscopy also verified that elec-trodes retained less than a monolayer of intact cells, whichoften displayed thin cellular structures (Fig. 3). Thus, this pro-tocol removed all but themost firmly attached cells, limited theheterogeneity of the cell-electrode interface, and maximizeddiffusion to the electrode surface.A series of control experiments were also conducted, includ-

ing exposure of electrodes to sonicated cells, sonicated cellextracts (supernatants or pellet fractions), and supernatantsfrom cell suspensions that had preincubated for 6 h at 30 °C toallow lysis or secretion of redox-active components to occur.These treatments tested the hypothesis that observed peakswere due to secreted proteins, compounds shed from cells, orproducts of cell lysis. None of these extracts from lysed, soni-cated, or aged cultures were able to reproduce peaks or redoxbehavior observed with whole cells. These observations indi-cated that intact cells attached to electrodes were required toproduce the electrochemical activity.Metabolic Activity of Films on Electrodes—To confirm that

these submonolayers of cells were metabolically active, lactatewas added, whereas electrodes were held at an oxidizing poten-tial (�0.24 V versus SHE). The addition of 20mM lactate imme-diately produced an anodic current, which stabilized within0.5 h. At this point, cyclic voltammetry of electrodes revealed acatalytic response (Fig. 4,A), unlike those typically observed forShewanella (such as those shown in Fig. 1A). Instead of a rapidonset centered at �0.2 V, current rose across a broad higherpotential window.When 1 �M riboflavin was added to these lactate-oxidizing

cell films, anodic current immediately increased. In the pres-ence of riboflavin, cyclic voltammetry returned to the familiarpattern of low onset potential, with a steep increase in currentcentered at �0.2 V. However, as observed in natural biofilms,the broad linear rise in anodic current at higher potentials stillremained. Thus, the behavior seen in naturally grown biofilmscould be recreated in thin films, but the films produced a morehomogeneous population of independent cells, withoutdemanding growth of the cells (allowing study of mutants), allof which could also be washed to remove electron donors orflavins.Cells lacking the outer membrane decaheme cytochromes

OmcA and/orMtrC could attach to electrodes in a similar fash-ion and be tested for their ability to transfer electrons to elec-trodes (Fig. 4, B andC). Strains containing deletions in either ofthese cytochromes were severely impaired in rates of sustainedelectron transfer, most dramatically when MtrC was deleted.Because similar levels of attached cells were detected on allelectrodes (14 �g of cell protein/electrode of wild type andMtrC, 11�g of cell protein/electrode forOmcA), a failure toadhere could not explain these large differences in rates of elec-tron flux.

FIGURE 2. Attached S. oneidensis MR-1 protein after 96 h of growth onplanar electrodes at �0.24 V versus SHE. The ratio of attached biomass tolimiting current produced in cyclic voltammetry is shown on the right axis.The error bars show standard deviations from three independent reactors.

Direct Electron Transfer by Shewanella

28868 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284 • NUMBER 42 • OCTOBER 16, 2009

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 5: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

Single Turnover Voltammetry of Wild Type MR-1—Becausefilms of live cells and mutants could be attached to electrodes,in the absence of both electron donors and electron shuttles,the reversible exchange of electrons between electrodes andproteins on the outer surface of cells could be probed for thefirst time. All of the following experiments were performedwith at least three freshly attached thin films of S. oneidensisMR-1, in which the medium was exchanged repeatedly toremove loosely bound cells and soluble electron donors. Inaddition, all cells were incubated with electrodes poised at oxi-dizing potentials prior to analysis to deplete internal electrondonor pools and facilitate attachment. Thin films were usedonce (e.g. for scan rate analysis) and discarded; all replicateswere performed with freshly polished electrodes incubatedwith newly adsorbed cells.Films were used to make three measurements: the estimated

total number of accessible redox-active centers, the potentialrange of reversible electron transfer, and interfacial rate con-stants. The moles of electroactive redox centers/cm2 dis-charged during a single sweep (coverage) was estimated byEquation 1,

� �Q

nFA(Eq. 1)

where Q is the total charge, in Cou-lombs, obtained by integration ofpeaks (dQ/dt, to adjust for scanrate), and n 1 (38).At slow scan rates (1 mV/s) the

total sum of discharged electronsduring a single sweep provided val-ues as high as 3,000 pmol of elec-trons/cm2 (for wild type). However,as scan rates exceeded 10 mV/s,total discharge of the film con-verged on a much lower value,approaching 250 pmol/cm2. As theelectrode-interacting cytochromeswere embedded in live cells, whichstill produced a small backgroundlevel of anodic current, this largerpeak area at slow scan rates could bedue to additional electrons flowingin and out of periplasmic or quinonepools. However, the rapid diminish-ing of this “reservoir” of electrons atmoderate scan rates (as shown bythe stable integrated peak areas inFig. 6A) prevented this pool fromcontributing to the higher scan ratedata used to infer kinetics. Based oncalculations of cytochromes pack-ing onto a surface (such as the85-kDa MtrC) and the rougher pol-ishing treatment used in our exper-iments, coverage levels on the orderof 50 pmol of electrons/cm2

(accounting for the decahemenature of MtrC) would be expected

if the electrode were only in communication with proteins ableto physically bind the electrode surface (11, 35, 47). The factthat our values at lower scan rates exceeded this range furthersuggested that proteins on the cell surface could rapidlyexchange electrons from sites inside the periplasmor on the cellmembrane, independent of the time scale of the experiment.At slow scan rates (1 mV/s), reversible peaks were separated

by a value of�150mV (Fig. 5). Such peak separation is commonfor adsorbed proteins and cytochromes (48). For films of wildtypeMR-1, discharge of the entire film spanned a window of atleast 400 mV, which is similar to what has been observed inpurified multiheme outer membrane cytochromes fromShewanella (11, 29). However, the potential at which electrontransfer from the cell surface occurredwas centered nearly 0.1–0.2 V higher than what has been reported for most purifiedShewanella cytochromes (11, 29), suggesting further heteroge-neity or differences in protein orientationwhen incorporated inthe cell membrane.An important observation throughout all of these experi-

ments was that the height of base line-corrected peaks was pro-portional to scan rate (Fig. 6B). This relationship is the classicaldemonstration of thin-film behavior (49, 50), as described byEquation 2.

FIGURE 3. A and B, scanning electron micrographs of electrodes before (A) and after (B) preparation of S.oneidensis MR-1 films. C, confocal microscopy (maximum projections, top view) of similar electrodes afterexposure to cells and (D) after washing five times to remove loosely attached cells.

Direct Electron Transfer by Shewanella

OCTOBER 16, 2009 • VOLUME 284 • NUMBER 42 JOURNAL OF BIOLOGICAL CHEMISTRY 28869

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 6: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

ip �n2F2Av�

4RT(Eq. 2)

Stated differently, the entire pool of redox-active sites, �,discharged completely, causing ip to vary as a function of �. Thiswas consistent with the rate of any electron delivery to the cell-electrode interface (e.g. from the cell interior to outer mem-brane cytochromes) being faster than the rate of interfacialelectron transfer itself (the final hop to electrodes). With nodiffusional limitations confounding the analyses, analysis athigher scan rates could be used to explore interfacial electrontransfer behavior (49). Although this analysis has been appliedto adsorbed compounds (49) and isolated proteins (35, 51),

these films of cells provided an opportunity to infer the sameinformation from intact bacteria for the first time.At scan rates as slow as 50 mV/s, peak separation became

evident, as shown in Fig. 7. This peak separation was used toestimate a range for interfacial electron transfer rate constantsbetween electrodes and surface-attached proteins, using boththe Laviron approach, and the fitting program Jellyfit (29, 49).As noted by others (29), because the complex multiheme pro-teins responsible for this electron transfer likely demonstratedispersion and may represent averages from a population of

FIGURE 4. A, cyclic voltammetry (1 mV/s) of electrodes with S. oneidensis MR-1cell films under starvation conditions (red trace), 30 min after the addition of20 mM lactate as an electron donor (blue trace), and 30 min after the additionof 1 �M riboflavin to lactate-amended cells (black trace). B and C, identicalexperiments for thin films of mutants containing deletions in omcA and mtrC,respectively. Two sweeps were performed for each CV, and the second isshown. WT, wild type.

FIGURE 5. Base line-subtracted cyclic voltammetry data (1 mV/s) showingreversible electron transfer by adsorbed S. oneidensis MR-1 cell films, inblack. The data from mtrC is in red, that from omcA is in blue, and that frommtrC/omcA is in green. The current vales are normalized to represent allelectrodes containing 1 �g of protein/cm2. The two similarly colored tracesshow data from two independent replicates, showing variability betweenfilms. Two sweeps were performed for each CV, and the second is shown. WT,wild type.

FIGURE 6. A, integrated peak area for cathodic (open symbols) and anodic(closed symbols) for wild type and mutant cell films as a function of scan rate.B, relationship between peak height and scan rate for representative S. onei-densis wild type and mutant cell films for both anodic and cathodic peaks. WT,wild type.

Direct Electron Transfer by Shewanella

28870 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284 • NUMBER 42 • OCTOBER 16, 2009

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 7: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

redox centers, the values should be interpreted as ranges ratherthan as precise constants.Significant peak separation always occurred at scan rates

lower than those observed for pure proteins (Fig. 7). This indi-cated that the electron transfer pathway of whole cells behavedas a system with a k0 more on the order of 1 s�1. This peakseparation behavior and estimated k0 window (estimates rang-ing from 0.5 to 3 s�1 were obtained) were at least 2 orders ofmagnitude slower than what has been observed for isolatedShewanella cytochromes on electrodes but were similar toturnover rates reported for cytochromes in membrane prepa-rations (as well as in whole cells) interacting with goethite (25).Electron Transfer Behavior of MR-1 Mutants—As shown in

Fig. 4, cultures containing deletions in omcA ormtrCwere ableto attach to electrodes but had different abilities to sustain elec-tron transfer from cells. Because these mutants could notrespire sufficiently to external acceptors to establish biofilms inlong term electrode growth experiments (such as those shownin Fig. 1), the thin film procedure provided a new means tocompare mutant constructs.For example, when strains containing a deletion in the gene

encoding the outer membrane decaheme cytochrome MtrCwere adsorbed to electrodes, reversible peaks could still bedetected, although the absolute height and integrated areaunder them was lower. Because the data in Fig. 5 has been nor-malized to account for slight differences in cell attachment, thisindicated less accessible redox centers per cell in the mutant.The shape of the anodic peak was altered slightly and was cen-tered at a potential 50 mV higher than the wild type. Similarly,when a mutant lacking the outer membrane decaheme cyto-chrome OmcA was adsorbed to electrodes, the total amount ofredox activity decreased further, and the potential windowshifted again (Fig. 5). Little difference was observed in peakseparation data at higher scan rates, indicating that the k0trends for each cytochrome mutant were similar.A double mutant lacking both outer membrane decaheme

cytochromes (MtrC and OmcA) demonstrated much lowerpeak heights (�10-fold lower than wild type or single outermembrane cytochrome deletion mutant, when adjusted forattached protein). The fact that this double mutant was still

capable of a small amount of reversible electron transfer sug-gested that additional outer membrane redox active proteinsremain to be identified (Fig. 5). Because of the small peakheight, kinetic measurements were not made for this mutant.

DISCUSSION

Biochemical, genetic, and genomic analyses have shown thatShewanella strains share a conserved suite of cytochromesresponsible for transfer of electrons from the inner to the outermembrane, where they are accessible to metals, mediators, orelectrodes (7, 9, 12, 26–29, 31, 52). Incubations of outer mem-brane cytochromes isolated from Shewanella have demon-strated that purified versions of these proteins will reduce met-als (9–11, 26–34). The recent finding that Shewanella alsosecretes soluble flavins that accelerate metal or surface reduc-tion showed that a soluble agent could help solve the problemofthe last “hop” fromproteins to surfaces (23, 24). In thiswork, wehave provided observations of electron transfer from intactcells without the apparent aid of flavins, shown that this elec-tron transfer occurs at specific potentials, and provided esti-mates of rate constants for the process in living cells. Thus, bothmechanisms of electron transfer may be operational, with therelative contributions determined by flavin concentrations, dif-fusional distances (formediated transfer), and driving force (fordirect transfer).When enzymes (35, 36, 53) or bacteria (40, 41, 54, 55) are

linked to electrodes via pathways with rapid interfacial rates ofelectron transfer, an increase in applied potential causes anidentical increase in the potential of attached enzyme redoxcenter(s). This creates the classic sigmoidal i-V wave, whichrises at a single redox potential and plateaus at a limiting cur-rent reflecting a limiting enzymatic turnover rate (42). In con-trast, the more linear dependence of electron transfer rate withvoltage observed in higher potential regions with Shewanella(Figs. 1 and 4) suggested an interaction with slower or mixedinterfacial kinetics. The response to soluble flavins in both nat-urally growing cells and cell films further illustrated that therewas an excess of enzyme activity at the electrode but a bottle-neck in the electron transfer pathway that could be alleviated bythe presence of a soluble redox mediator.The results obtained with intact cells of Shewanella allow

some comparison with those obtained in vitro with recombi-nant proteins in two key parameters; apparentmidpoint poten-tials, and estimated electron transfer rates. It is important tonote that recombinant proteins are typically purifiedwith affin-ity tags, expressed without lipid-binding domains, incubated inthe presence of detergents, studied with soluble iron ligands,and/or are bound to electrodes with the aid of agents such aspolymixin (10, 27, 29, 30, 56). Such treatments could alter theexposed surface of the protein, increase flexibility, allow closercontact, or facilitate more rapid turnover.For example,MtrChas been shown in two separate studies to

be reversibly oxidized at basal plane graphite electrodes acrossa wide potential range, similar to what was observed in ourstudy with whole cells (of over 400mV). Films of purifiedMtrCproduced broad redox peaks centered at �138 mV (29) and�100 mV (11). Using cytochromes attached to scanning tun-neling microscope tips, an equally wide potential window cen-

FIGURE 7. Representative data showing the peak potential separationversus log (scan rate) for S. oneidensis MR-1 cell films. The solid line showsan example of a fit for a wild type film, where the data were fit to a k0 of 2 s�1.The dotted lines show peak separation ranges consistent with 0.1, 1, and 10s�1, respectively.

Direct Electron Transfer by Shewanella

OCTOBER 16, 2009 • VOLUME 284 • NUMBER 42 JOURNAL OF BIOLOGICAL CHEMISTRY 28871

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 8: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

tered near �200 mV was observed for MtrC (30). Films ofOmcA have also been reported to have potentials in the �150to �100 mV range, using metal oxide or graphite electrodes(29, 56).In contrast, direct electron transfer from intact Shewanella

films to graphite electrodes described here required potentialsnear 0 V, or at least 100–200mVmore positive than previouslyreported. Data from Shewanella sp. strain MR-4 biofilms alsodetected reversible electron transfer reactions in the same 0 Vrange (23). In this work, deletion of specific cytochromesaltered the redox behavior of whole cells, which supported thehypothesis that we were observing electron transfer as cata-lyzed by these well studied outer membrane (or pili-attached)proteins.A second comparison is related to rates of electron transfer.

Estimates for interfacial rate constants in excess of 100 s�1 havebeen reported for purified MtrC (11, 29). Such rates (in the100–1000 s�1 range) are highly typical of adsorbed enzymesstudied by film voltammetry (38, 57). Rates for cytochromesinteracting with solid electron acceptors are more variable andare often based on measurements during electron transfer tosolid substrates (expressed as kcat or rates observed undermod-erate driving force).Using purified OmcA, both Eggleston et al. (56) and Xiong et

al. (58) observed electron transfer to Fe(III) nanoparticles oroxide-coated electrodes on the order of 10�13mol e� cm�2 s�1,equivalent to �0.1 s�1/cytochrome. Incubations with wholeShewanella putrefaciens and Fe(III) nanoparticles producedestimates of 0.001 s�1 (59), although only a percentage of cyto-chromeswere likely reactingwith Fe(III), causing this latter rateto represent a lower boundary. Similarly low rates of single-protein turnover to goethite of 0.5–5 s�1 were recently pro-vided by Ross et al. (25). Because an overpotential of at least 100mV is present to estimate kcat, ko values for intact cells could beexpected to be at least this low.Our scan rate analysis data from whole cells provided inde-

pendent estimates of interfacial electron transfer to externalacceptors. Although these values represent averages of all elec-tron transfer reactions involved in direct transfer for an intactShewanella cell, they also represent a glimpse into the realworld conditions experienced by a cell attempting utilize itsintact electron transfer system to reduce this surface, with pro-teins confined to complexes within the membrane. All of ourmeasurements were consistent with rates on the order of 1 s�1

or slower, which were 10–100-fold lower than k0 valuesobserved for pure proteins adsorbed to electrodes (11, 29).When the maximum number of redox centers able to form a

cell-electrode connection and rate constants fromour scan rateanalysis are combined, a prediction of the exchange currentdensity (j0) for the nonmediated pathway can be made. Assum-ing that the final step of electron transfer occurs through cyto-chromes with a size similar to OmcA or MtrC, the most elec-tron transfer sites (regardless of whether cytochromes are onthe cell surface or on cell pili) that can possibly be in physicalcontactwith any electrode is on the order of 10�12mol/cm2 (11,60). Based on estimated k0 values for this pathway (�1 s�1), thisis equivalent to an exchange current between cells and elec-trodes of only �1 � 10�7 A/cm2.

Values in this low j0 range require significant overpotentials(of at least �100 mV) simply to achieve a forward anodic cur-rent of the magnitude attributed to the non-flavin-mediatedpathway (�10 �A/cm2) (50). This requirement for overpoten-tial is compounded by the broad potential window of the cyto-chromes, meaning that at a potential �0 V, only a fraction ofredox sites are experiencing conditions that enable forwardelectron transfer. These calculations are also consistent withestimates by others (11, 58), where films of pure OmcA sub-jected to moderate driving force supported electron transferrates equivalent to only �1.6 �A/cm2.

Taken together, the boost in electron transfer in naturallygrown biofilms above 0 V (Fig. 1) and the electron transferobserved in the same potential rangewith adsorbed Shewanellain the absence of flavins (Fig. 4) are likely due to cytochrome-electrode interactions. The relationship between appliedpotential and electron transfer at higher potentials is consistentwith a step in the direct pathway being relatively slow, com-paredwith reactions that supply electrons to this interface. Thisis also supported by the thin-film behavior relationship, whichwould be expected if reaction(s) feeding electrons to the inter-face were faster than the interface reaction itself.In contrast, even small amounts of flavins (when present) can

facilitate the transfer of electrons to electrodes at lower poten-tials and cycle back to Shewanella cell surface cytochromes.Diffusion limits the contribution of electron transfer by thismechanism, as shown by the strong relationship between solu-ble flavin concentration and anodic current (Figs. 1 and 4) andcurrent/protein ratios (Fig. 2).Implications for Shewanella—These results show that redox-

active proteins can be present on the outer surface of cells butnot interfaced properly with the cell interior (e.g. as seen inMtrC mutants that produce peaks in Fig. 5 but no catalyticactivity in Fig. 4). In addition, the cytochromes present on theouter surface appear to experience a bottleneck in electrontransfer to solid phase electron acceptors, a conclusion alsoreached in recent experiments with purified OmcA and MtrCby Ross et al. (25).Molecular computations of multiheme cytochromes (33)

have shown that subtle variations in the orientation of hemesrelative to a surface can slow rates of electron transfer by thesame protein from nearly 100 to 0.01 s�1. Because changes inthe form of chelated Fe(III) (27), the nature of oxide surfaces(56), and even alterations to the orientation of hemes at a sur-face (33) can dramatically alter electron transfer rates, surfacesother than carbon electrodes may interact differently with thedirect transfer pathway of Shewanella.

In the environment, the diversity of reducible oxide surfacesmay select for flavin secretion as ameans for cells to gain accessto wider range of acceptors, without requiring the organism toexpress many different cytochromes tuned to a variety ofunpredictable surfaces. Thus, flavins could act as a kind of “uni-versal translator,” allowing Shewanella to gain access to adiverse array of redox-active substrates, using only a singlecytochrome-based pathway.These observations also argue that, under conditions such as

microbial fuel cells where anodes stabilize at low (�0.1 V)potentials, Shewanellawould not compete with organisms that

Direct Electron Transfer by Shewanella

28872 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 284 • NUMBER 42 • OCTOBER 16, 2009

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 9: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

are capable of higher rates of respiration in the �0.1 V range(40, 41, 61) unless mediators accumulated to significantly facil-itate electron transfer. Alternatively, if an anode was main-tained at a higher potential (via use of a lower external resist-ance or potentiostat control), these observations provide athermodynamic basis for enrichment of different populationsof bacteria, based on the differing responses to potential in thisrelatively narrow range.

REFERENCES1. Richardson, D. J. (2000)Microbiology 146, 551–5712. Nealson, K. H., and Saffarini, D. (1994)Annu. Rev. Microbiol. 48, 311–3433. Lovley, D. R., Holmes, D. E., and Nevin, K. P. (2004) Adv. Microb. Physiol.

49, 219–2864. Gralnick, J. A., and Newman, D. K. (2007)Mol. Microbiol. 65, 1–115. Nealson, K. H., and Scott, J. (2005) in The Prokaryotes, pp. 1133–1151,

Springer-Verlag New York Inc., New York6. Myers, J. M., andMyers, C. R. (1998) BBA-Biomembranes 1373, 237–2517. Myers, J. M., and Myers, C. R. (2000) J. Bacteriol. 182, 67–758. Myers, J. M., andMyers, C. R. (2001)Appl. Environ. Microbiol. 67, 260–2699. Ross, D. E., Ruebush, S. S., Brantley, S. L., Hartshorne, R. S., Clarke, T. A.,

Richardson, D. J., and Tien, M. (2007) Appl. Environ. Microbiol. 73,5797–5808

10. Shi, L., Chen, B., Wang, Z., Elias, D. A., Mayer, M. U., Gorby, Y. A., Ni, S.,Lower, B. H., Kennedy, D. W., Wunschel, D. S., Mottaz, H. M., Marshall,M. J., Hill, E. A., Beliaev, A. S., Zachara, J.M., Fredrickson, J. K., and Squier,T. C. (2006) J. Bacteriol. 188, 4705–4714

11. Hartshorne, R. S., Jepson, B. N., Clarke, T. A., Field, S. J., Fredrickson, J.,Zachara, J., Shi, L., Butt, J. N., and Richardson, D. J. (2007) J. Biol. Inorg.Chem. 12, 1083–1094

12. Ruebush, S. S., Brantley, S. L., and Tien, M. (2006) Appl. Environ. Micro-biol. 72, 2925–2935

13. El-Naggar,M. Y., Gorby, Y. A., Xia,W., andNealson, K. H. (2008)Biophys.J. 95, L10–L12

14. Gorby, Y. A., Yanina, S., McLean, J. S., Rosso, K. M., Moyles, D.,Dohnalkova, A., Beveridge, T. J., Chang, I. S., Kim, B. H., Kim, K. S., Culley,D. E., Reed, S. B., Romine, M. F., Saffarini, D. A., Hill, E. A., Shi, L., Elias,D. A., Kennedy, D. W., Pinchuk, G., Watanabe, K., Ishii, S., Logan, B.,Nealson, K. H., and Fredrickson, J. K. (2006) Proc. Natl. Acad. Sci. U.S.A.103, 11358–11363

15. Newman, D. K., and Kolter, R. (2000) Nature 405, 94–9716. Lies, D. P., Hernandez,M. E., Kappler, A.,Mielke, R. E., Gralnick, J. A., and

Newman, D. K. (2005) Appl. Environ. Microbiol. 71, 4414–442617. Haas, J. R., andDiChristina, T. J. (2002) Environ. Sci. Technol. 36, 373–38018. Taillefert, M., Beckler, J. S., Carey, E., Burns, J. L., Fennessey, C. M., and

DiChristina, T. J. (2007) J. Inorg. Biochem. 101, 1760–176719. Kim, B. H., Ikeda, T., Park, H. S., Kim, H. J., Hyun, M. S., Kano, K., Takagi,

K., and Tatsumi, H. (1999) Biotechnol. Tech. 13, 475–47820. Kim, B. H., Kim, H. J., Hyun, M. S., and Park, D. H. (1999) J. Microbiol.

Biotechnol. 9, 127–13121. Kim, H. J., Hyun, M. S., Chang, I. S., and Kim, B. H. (1999) J. Microbiol.

Biotechnol. 9, 365–36722. Kim, H. J., Park, H. S., Hyun, M. S., Chang, I. S., Kim, M., and Kim, B. H.

(2002) Enzyme Microb. Technol. 30, 145–15223. Marsili, E., Baron, D. B., Shikhare, I. D., Coursolle, D., Gralnick, J. A., and

Bond, D. R. (2008) Proc. Natl. Acad. Sci. U.S.A. 105, 3968–397324. von Canstein, H., Ogawa, J., Shimizu, S., and Lloyd, J. R. (2008) Appl.

Environ. Microbiol. 74, 615–62325. Ross, D. E., Brantley, S. L., and Tien, M. (2009) Appl. Environ. Microbiol.

75, 5218–522626. Zhang, H., Tang, X., Munske, G. R., Zakharova, N., Yang, L., Zheng, C.,

Wolff, M. A., Tolic, N., Anderson, G. A., Shi, L., Marshall, M. J., Fredrick-son, J. K., and Bruce, J. E. (2008) J. Proteome. Res. 7, 1712–1720

27. Wang, Z., Liu, C., Wang, X., Marshall, M. J., Zachara, J. M., Rosso, K. M.,Dupuis, M., Fredrickson, J. K., Heald, S., and Shi, L. (2008) Appl. Environ.Microbiol. 74, 6746–6755

28. Shi, L.,Deng, S.,Marshall,M. J.,Wang,Z.,Kennedy,D.W.,Dohnalkova,A.C.,Mottaz, H. M., Hill, E. A., Gorby, Y. A., Beliaev, A. S., Richardson, D. J., Za-chara, J. M., and Fredrickson, J. K. (2008) J. Bacteriol. 190, 5512–5516

29. Firer-Sherwood, M., Pulcu, G. S., and Elliott, S. J. (2008) J. Biol. Inorg.Chem. 13, 849–854

30. Wigginton, N. S., Rosso, K. M., and Jr., M. F. (2007) J. Phys. Chem. B 111,12857–12864

31. Shi, L., Squier, T. C., Zachara, J. M., and Fredrickson, J. K. (2007) Mol.Microbiol. 65, 12–20

32. Lower, B. H., Shi, L., Yongsunthon, R., Droubay, T. C., McCready, D. E.,and Lower, S. K. (2007) J. Bacteriol. 189, 4944–4952

33. Kerisit, S., Rosso, K.M., Dupuis,M., andValiev,M. (2007) J. Phys. Chem. C111, 11363–11375

34. Marshall, M. J., Beliaev, A. S., Dohnalkova, A. C., Kennedy, D. W., Shi, L.,Wang, Z., Boyanov, M. I., Lai, B., Kemner, K.M., McLean, J. S., Reed, S. B.,Culley, D. E., Bailey, V. L., Simonson, C. J., Saffarini, D. A., Romine, M. F.,Zachara, J. M., and Fredrickson, J. K. (2006) PLoS Biol. 4, e268

35. Heering, H. A., Weiner, J. H., and Armstrong, F. A. (1997) J. Am. Chem.Soc. 119, 11628–11638

36. Heering, H. A., Hirst, J., and Armstrong, F. A. (1998) J. Phys. Chem. B 102,6889–6902

37. Armstrong, F. A. (1999) Bioc. Soc. Tran. 27, 206–21038. Jeuken, L. J., Jones, A. K., Chapman, S. K., Cecchini, G., and Armstrong,

F. A. (2002) J. Am. Chem. Soc. 124, 5702–571339. Armstrong, F. A. (2005) Curr. Opinion Chem. Biol. 9, 110–11740. Srikanth, S., Marsili, E., Flickinger, M. C., and Bond, D. R. (2008) Biotech-

nol. Bioeng. 99, 1065–107341. Marsili, E., Rollefson, J. B., Baron, D. B., Hozalski, R. M., and Bond, D. R.

(2008) Appl. Environ. Microbiol. 74, 7329–733742. Richter, H., Nevin, K. P., Jia, H. F., Lowy, D. A., Lovley, D. R., and Tender,

L. M. (2009) Energy Env. Sci. 2, 506–51643. Dumas, C., Basseguy, R., and Bergel, A. (2008) Electrochim. Acta 53,

5235–524144. Fricke, K., Harnisch, F., and Schroder, U. (2008) Energy Environ. Sci. 1,

144–14745. Hau, H. H., Gilbert, A., Coursolle, D., and Gralnick, J. A. (2008) Appl.

Environ. Microbiol. 74, 6880–688646. Fourmond, V., Hoke, K., Heering, H. A., Baffert, C., Leroux, F., Bertrand,

P., and Leger, C. (2009) Bioelectrochemistry 76, 141–14747. Armstrong, F. A., Heering, H. A., and Hirst, J. (1997) Chem. Soc. Rev. 26,

169–17948. Hirst, J., Duff, J. L. C., Jameson, G. N. L., Kemper,M. A., Burgess, B. K., and

Armstrong, F. A. (1998) J. Am. Chem. Soc. 120, 7085–709449. Laviron, E. (1979) J. Electroanal. Chem. 101, 19–2850. Bard, A. J., and Faulkner, L. R. (2001) Electrochemical Methods, 2nd Ed.,

John Wiley and Sons51. Armstrong, F. A., Camba, R., Heering,H. A., Hirst, J., Jeuken, L. J. C., Jones,

A. K., Leger, C., and McEvoy, J. P. (2000) Faraday Disc. 116, 191–20352. Murphy, J. N., and Saltikov, C. W. (2007) J. Bacteriol. 189, 2283–229053. Hirst, J., Sucheta, A., Ackrell, B. A., and Armstrong, F. A. (1996) J. Am.

Chem. Soc. 118, 5031–503854. Busalmen, J. P., Esteve-Nunez, A., and Feliu, J. M. (2008) Env. Sci. Technol.

42, 2445–245055. Busalmen, J. P., Esteve-Nunez, A., Berna, A., and Feliu, J.M. (2008)Angew.

Chem. Int. 47, 4874–487756. Eggleston, C. M., Voros, J., Shi, L., Lower, B. H., Droubay, T. C., and

Colberg, P. J. S. (2008) Inorg. Chim. Acta 361, 769–77757. Baymann, F., Barlow, N. L., Aubert, C., Schoepp-Cothenet, B., Leroy, G.,

and Armstrong, F. A. (2003) FEBS Lett. 539, 91–9458. Xiong, Y., Shi, L., Chen, B.,Mayer,M.U., Lower, B. H., Londer, Y., Bose, S.,

Hochella, M. F., Fredrickson, J. K., and Squier, T. C. (2006) J. Am. Chem.Soc. 128, 13978–13979

59. Bonneville, S., Behrends, T., Van Cappellen, P., Hyacinthe, C., and Roling,W. F. M. (2006) Geochim. Cosmochim. Acta 70, 5842–5854

60. Blanford, C. F., and Armstrong, F. A. (2006) J. Solid State Electrochem. 10,826–832

61. Xing, D., Zuo, Y., Cheng, S., Regan, J. M., and Logan, B. E. (2008) Environ.Sci. Technol. 42, 4146–4151

Direct Electron Transfer by Shewanella

OCTOBER 16, 2009 • VOLUME 284 • NUMBER 42 JOURNAL OF BIOLOGICAL CHEMISTRY 28873

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from

Page 10: ElectrochemicalMeasurementofElectronTransferKinetics by ... · When peak heights had stabilized (typically within 6 h), working electrodes were washed to remove loosely attached cells

Daniel Baron, Edward LaBelle, Dan Coursolle, Jeffrey A. Gralnick and Daniel R. Bond MR-1oneidensis

ShewanellaElectrochemical Measurement of Electron Transfer Kinetics by

doi: 10.1074/jbc.M109.043455 originally published online August 6, 20092009, 284:28865-28873.J. Biol. Chem. 

  10.1074/jbc.M109.043455Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

  http://www.jbc.org/content/284/42/28865.full.html#ref-list-1

This article cites 59 references, 16 of which can be accessed free at

by guest on January 13, 2021http://w

ww

.jbc.org/D

ownloaded from