Recent Developments in Faradaic Bioelectrochemistry

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Electrochimica Acta 45 (2000) 2623 – 2645

www.elsevier.nl/locate/electacta

Recent developments in faradaic bioelectrochemistry


Fraser A. Armstrong a,*, George S. Wilson b
a
Inorganic Chemistry Laboratory, Uni6ersity of Oxford, South Parks Road, Oxford OX1 3QR, UK
b
Department of Chemistry, Uni6ersity of Kansas, Lawrence, KS 66045, USA

Papers received in Newcastle, 20 December 1999

Abstract

Progress in the area of faradaic bioelectrochemistry over the last decade has been reviewed. Electrochemical and
spectroscopic studies of the protein – electrode interface are discussed along with the use of electron transfer
promoters, self-assembled monolayers (SAMs), and surfactant films. Voltammetric techniques are described that
permit rapid, in-situ measurement of formal potentials while also providing kinetic resolution on the sub-millisecond
time scale. Some enzymes, especially those of analytical interest, must be coupled or ‘wired’ to the electrode using
mediators in order to achieve rapid electron transfer. Supramolecular structures designed to facilitate electron transfer
and to immobilize co-factors are described. The use of electrochemically-based biosensors for in-vivo measurements
is documented. © 2000 Elsevier Science Ltd. All rights reserved.

Keywords: Protein voltammetry; Wired enzymes; In vivo electrochemistry; Chemically modified electrodes; Electron transfer kinetics

1. Introduction trons and finally, we document the development of


biosensors to be used especially in monitoring events
More than 200 years have passed since the death of occurring about single cells, in tissue slices or in intact
Luigi Galvani, who could certainly be considered a brain.
father of bioelectrochemistry, so the field is hardly a The use of voltammetric methods to study redox
new development. Because of the broad range of activ- proteins and their active sites has gained increasing
ities we have decided to focus only on faradaic bioelec- attention. Following the early pioneering studies of the
trochemistry, meaning that important biological 1970s and 1980s, voltammetric methods have become
phenomena such as membrane potentials and currents routine tools for determining formal potentials of redox
resulting from ion transport across membranes will not centers in small proteins. It is now well established that,
be discussed. Instead, we focus on three areas that have with suitable electrodes, small electron carriers such as
been especially prominent in the 1990s. First we note cytochromes, ferredoxins and blue Cu proteins give
the extensive use of voltammetric techniques to extract reversible, electrochemistry without the need for small
essential physicochemical data concerning the kinetics mediators. Progress is being made with increasingly
and energetics of protein redox reactions. Second, we complex systems, including multi-centered enzymes,
consider fundamental and practical considerations in and it is becoming acknowledged that the electrochemi-
linking or ‘wiring’ otherwise electroinactive enzymes to cal response of active sites provides a useful diagnostic
electrodes so that they can efficiently transport elec- signal, in many ways analogous to that provided by
various spectroscopic methods. The electrode replaces
physiological redox partners, providing instead a driv-
* Corresponding author. Tel.: + 44-1865-270841; fax: + 44-
1865-272690. ing force that is variable in the potential and time
E-mail address: fraser.armstrong@chem.ox.ac.uk (F.A. domains to energize reactions and a sensor to measure
Armstrong) the response.

0013-4686/00/$ - see front matter © 2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 3 - 4 6 8 6 ( 0 0 ) 0 0 3 4 2 - X
2624 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

2. The protein– electrode interface features can be probed. The various types of protein–
electrode interface currently under study are depicted in
Little more than 20 years ago, the idea that dynamic Fig. 1.
electrochemical methods could be applied to protein Interactions between protein and electrode may be
molecules as large as cytochrome c (let alone much tailored to be weak or strong. Weak interactions might
larger enzymes) was regarded with scepticism. The no- ideally give rise to diffusion controlled voltammetry of
table pioneers of this era were the groups led by Niki solution species, whereas with strong interactions the
(voltammetry of cytochrome c3 on mercury) [1], Hill experiment may address just a stable protein film, ide-
(Au modified with a monolayer of organic ‘promoter’) ally a monolayer. Following the pioneering work by the
[2] and Kuwana (cytochrome c at a metal oxide elec- groups of Hill and Taniguchi on the modification of
trode) [3]. In each case it was established that important gold electrodes with organic adsorbates that yield diffu-
information on a protein’s redox properties could be sion-controlled electrochemistry of cytochrome c [5,6],
obtained from experiments in which direct electron further efforts have been made to understand how these
exchange took place between the electrode and the ‘promoters’ work. In particular, do they provide tran-
active site. Since that time, many important develop- sient interaction sites for the protein or do they prevent
ments have taken place in the design and structural blocking by contaminants? It appears that both these
characterization of electrode surfaces that are active factors are important. Lateral interactions among the
with proteins [4]. The misleading dogma that protein adsorbate molecules are important and lead to self
adsorption always presents an undesirable problem has assembled monolayers (SAMs), the X-functionality
given way to more constructive ideas in which the (nowadays usually a thiol) being bound to the metal (as
adsorbed state actually provides the key to studying thiolate) and the Y-functionality interacting with the
and understanding these complex systems, many of solution and protein molecule. Selectivity of modified
which are intimately associated with interfaces (e.g. electrodes for particular proteins can be tuned by
membranes) in their natural environment. Accordingly, changing the functionality Y; thus whereas cytochrome
efforts have been made to achieve strong adsorption of c prefers acidic groups such as COO(H) which can
proteins on electrodes modified in such a way as to interact more favorably with the lysine-dominated re-
preserve native properties or provide a well-defined gion around the exposed heme edge, acidic proteins
environment in which electron-transfer or spectroscopic such as plastocyanin or ferredoxins respond best if Y is

Fig. 1. Cartoon depicting the various types of electrode surfaces for which protein voltammetry is commonly observed. Shown are
a metal electrode modified with an ‘XY’ SAM, a metal oxide electrode, a pyrolytic graphite ‘edge’ electrode often used in
conjunction with mobile co-adsorbates such as aminocyclitols, and an electrode modified with a surfactant layer in which protein
molecules are embedded.
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2625

a basic group such as NH2 (H+) [7,8]. Although it has bare gold or indium oxide, or at a rotating modified
often been assumed that the adsorbate itself is stable, gold electrode. Assuming that the surface is not blocked
recent work has shown that 4-mercaptopyridine (or so as to generate inactive sites, these observations are
bis(4-pyridyl) disulfide, which adsorbs as 4-mercaptopy- predicted by a model that takes into account the effect
ridine decomposes slowly at gold to give adsorbed of the potential drop that occurs across the adsorbed
elemental sulfur, thereby inactivating the surface for layer.
protein interaction [9]. The loss of activity gives Important advantages are gained if proteins can be
changes in waveform similar to those expected on the immobilized at the electrode surface, giving an elec-
basis of the ‘microscopic model’, as active areas of the troactive film that is ideally of monolayer coverage.
electrode become progressively blocked [10]. Armstrong and co-workers have termed this approach
The microscopic model accounts for the rounded and ‘protein film voltammetry’ to encompass a variety of
drawn-out shapes of voltammograms attributed to ‘ir- different configurations [15]. Without the need to have
reversible’ electrochemistry, in terms of electron trans- protein molecules free in bulk solution, minuscule sam-
fer that is in fact reversible but occurs only at isolated ple quantities can be used, and there is greatly increased
active zones on the electrode, i.e. those sites that are resolution of thermodynamic and kinetic information.
suitably functionalized to interact with the protein Referring back to Fig. 1, immobilization can be
molecule. It has found fairly wide acceptance, with achieved in several ways. Most simply, the electrostatics
support stemming from studies on electrodes coated can be optimised. For example, at low ionic strength,
with lipids and presenting only ‘pinholes’ to allow cytochrome c adsorbs strongly at gold electrodes
interaction with proteins such as cytochrome c [11]. The modified with v-carboxylate alkanethiol SAMs, as ex-
effect of blocking protein interaction sites has also been pected if this is governed by electrostatic interactions
examined using mixed SAMs comprising active ‘X– Y’ involving the lysine residues surrounding the exposed
adsorbates diluted with non-active co-adsorbates that heme edge [16]. Stable hydrophilic electrode materials
lack the appropriate ‘Y’-functionality [12]. With solu- such as metal oxides or polished carbon (glassy carbon
tions of cytochrome c, the classical, peak shaped or pyrolytic graphite edge-PGE, which has acidic CO
voltammograms which are exhibited at gold electrodes functionalities [17] also have a good capability for
modified with just 3-mercaptopropionic acid or bis(4- electrostatically controlled adsorption of proteins with-
pyridyl) disulfide (which is bound as 4-mercaptopy- out denaturation or loss of biochemical activity
ridine) become increasingly sigmoidal as n-alkylthiols [15,18,19]. Proteins have even been shown to be elec-
are added to the monolayer. Interestingly, with 3-mer- troactive when immobilized in carbon nanotubes [20].
captoethanoic acid, the effect is much more marked as For the more acidic proteins, adsorption may require or
n is increased from 3 to 18, whereas with bis(4-pyridyl) be enhanced by the inclusion of co-adsorbates such as
disulfide it is the shorter chain n-alkylthiols that are neomycin, polymyxin or other polyamines. Although
more effective in blocking the surface. It is proposed the structures of the resulting electrode surfaces are not
that this is due to the way the co-adsorbates interdis- well understood, carbon and metal oxides offer certain
perse on the surface, with interaction between bis(4- advantages over SAM-modified metal electrodes: for
pyridyl) disulfide and long chain n-alkylthiols being example, pyrolytic graphite provides a very wide poten-
sufficiently weak so as to leave open large areas of tial window, while metal oxides are optically transpar-
adsorbed bis(4-pyridyl) disulfide suitable for cy- ent and permit UV-visible spectral changes to be
tochrome c interaction. Relevant to this are recent observed as the surface-bound protein is redox-cycled
studies on how stabilities of n-alkyl SAMs depend on [21]. Protein molecules may also be attached to elec-
chain length; notably shorter chains produce a less trodes by covalent bonding. Dong and co-workers have
stable layer because the hydrophobic effect is smaller reported the use of carbodiimide coupling to link the
[13]. normally membrane-bound enzyme cytochrome c oxi-
Further refinements in the analysis and interpretation dase to a Au/3-mercaptopropionic acid SAM [22]. The
of diffusion-controlled protein voltammograms have ‘chemisorbed’ enzyme shows reversible voltammetry
been reported. Honeychurch and Rechnitz have consid- due to one or more active sites, and can transfer
ered the influence on the appearance of cyclic voltam- electrons on to cytochrome c in solution. Hill and
mograms arising from the requirement that for co-workers achieved adsorption of Azurin on gold by
SAM-modified electrodes, electron exchange must oc- engineering a cysteine residue into the structure, The
cur at the outer extremity of the SAM and not the resulting AuS bond orients the enzyme for electron
electrode surface itself [14]. For cytochrome c solution exchange and restricts its lateral movement on the
voltammetry at a bis(4-pyridyl) disulfide-modified gold electrode [23].
electrode, the peak separation is always greater than 57 Various spectroscopic studies have been undertaken
mV; further, the diffusion coefficient is smaller than to examine the structural integrity of proteins adsorbed
that obtained from studies at simple electrodes such as at electrodes. For heme proteins, the Raman spectrum
2626 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

is a sensitive probe of spin state and hence of any protein molecules are obtained by several methods.
disruption to the active site. Surface-enhanced reso- The simplest way is to evaporate a chloroform solu-
nance Raman spectroscopy (SERRS) studies show tion of surfactant onto the electrode, which is typi-
that oxidized and reduced cytochrome c remain in cally pyrolytic graphite, gold or platinum. The
their native conformations when adsorbed at a silver electrode is then placed in a solution of the protein
electrode coated with a monolayer of bis(4-pyridyl) which then diffuses into the film. Another method is
disulfide or 4-mercaptopyridine, the depth of which is to mix an aqueous vesicle dispersion of surfactant
short enough to allow resonance transfer to the with protein solution then evaporate this onto the
protein [24]. By contrast, at a bare silver electrode, electrode. These procedures give multiple layers, and
the spin-state marker bands of adsorbed cytochromes for gold or platinum deposited on a quartz crystal,
are shifted, signalling conformational changes that the depth of the layer can gauged by QCM measure-
may be relevant to function. Exploiting this further, ments. Initial studies were carried out with myo-
Hildebrandt and co-workers have reported an applica- globin, a heme-containing oxygen binding protein that
tion of time-resolved SERRS to measure the kinetics displays only slow electron transfer at bare electrode
of electron transfer and coupled conformational surfaces [32]. Rusling’s group found that myoglobin
changes in a bacterial counterpart of cytochrome c, (Mb) gives reversible voltammetry when contained in
cytochrome c552, adsorbed at a silver surface [25]. a DDAB film at basal plane graphite, a discovery that
Techniques such as ellipsometry and quartz crystal has been elegantly exploited (see below) by Farmer
microbalance (QCM) measurements are widely used and co-workers [33]. The approach was subsequently
to quantitate the progress of adsorption of proteins extended to cytochrome P450, enabling direct detec-
and supporting nanostructure on electrode surfaces tion of redox transitions of the heme group [34]. Re-
[26]. The question of how protein molecules are ori- duction potentials are sensitive to the nature of the
ented at the electrode has provoked some interesting surfactant, i.e. whether DDAB or DMPC are used,
experiments, such as that reported by Wilson and co- and also (as expected) to the presence of CO. The
workers who used antibodies to study the potential- enzyme film also catalyzes reduction of O2 and reduc-
dependence of orientation of adsorbed cytochrome c3 tive dechlorination of trichloroacetic acid. These films
[27]. Orientation distributions of molecules adsorbed behave electrochemically as thin layers within which
on various surfaces have been studied by a combina- the protein molecules diffuse freely, although the rea-
tion of absorption linear dichroism and emission an- son for this high mobility is not clear. As liquid crys-
isotropy [28]. The results indicate that for protein tals, it is easy to envisage a very dynamic medium
adsorption at SAMs there is a quite a broad distribu- with alkyl chains moving around like a solvent.
tion of orientations, so that the protein molecules are There is substantial spectroscopic evidence to sug-
rather disordered. This is relevant to the problem of gest that these proteins retain most if not all of their
peak broadening due to dispersion, noting that, ide- native structure when immobilized in these surfactant
ally, all the molecules in a monolayer should behave films. Rusling and co-workers have employed UV-visi-
identically. This aspect has been identified in different ble, EPR and reflectance FT-IR to establish the in-
kinds of experiment and modeled [29,30]. In cyclic tegrity of myoglobin and cytochrome P450 in films
voltammetry, reversible electrochemistry for an immo- deposited on suitable substrates [35]. Amide IR bands,
bilized sample is characterized by a peak half-height which reflect backbone CO stretching and are sensi-
width of 90.6/n mV at 25°C. This value varies with tive to protein conformation, do not alter significantly
temperature (width 8T/K), or interactions between when Mb is transferred from aqueous solution to the
sites (cooperative or anticooperative). However it can DDAB film.
also be increased by dispersion, i.e. inhomogeneity in Boussaad and Tao have used atomic force mi-
thermodynamics or kinetics. croscopy (AFM) to examine the interactions of
Another way to prepare electroactive films of proteins (myoglobin or cytochrome c) and DDAB
protein molecules is to immobilize them in surfactant with a pristine, highly-ordered pyrolytic graphite
films, which resemble lipid bilayers and multilayers. (HOPG) surface [36]. Cytochrome c adsorbs ran-
Rusling and co-workers have developed procedures domly, rapidly giving a complete layer. By contrast,
for preparing surfactant films into which proteins are adsorption of Mb from solution proceeds slowly, but
embedded. These may be multi-layer assemblies within eventually results in chain-like structures about 60 nm
which protein molecules diffuse, but do not escape to long consisting of blobs about 6 nm wide. At this
the bulk aqueous phase [31]. The surfactants are wa- point in time, voltammetry reveals a reversible redox
ter-insoluble, with two long-chain alkyl chains; exam- couple with a potential close to that expected for
ples being didodecyldimethylammonium bromide myoglobin. If the electrode is coated with DDAB,
(DDAB), dimyristoyl-phosphatidyl-cholate (DMPC) adsorption of Mb is much more random, suggesting
and dihexadecylphosphate (DHP). Films containing that strong DDAB – protein interactions break up the
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2627

inter-protein interactions responsible for aggregation at reductive wave as the scan rate is increased. The pres-
the bare electrode. The corresponding voltammetry is ence of a bilayer is supported by QCM measurements,
less reversible than at the bare electrode, most likely while the catalytic enhancement that is observed with
reflecting the increased distance over which electron cytochrome c in solution suggests that the enzyme is
transfer must occur. The structure of the DDAB layer oriented with its binding site for this natural partner
was also found to be potential dependent, with the directed away from the electrode. However, the charge
liquid crystal phase only being favored above 0 V. that is passed during potential sweeps is far greater
These results are somewhat at variance with previous than expected for a monolayer, prompting the sugges-
conclusions that Mb is inactive at a bare graphite tion that the excess charge is capacitative and arises
electrode, but nevertheless provide an alternative and from a redox-coupled conformational change that in-
revealing perspective. creases ion transport in the lipid bilayer. This is an
Rusling, Lvov and co-workers have also examined interesting idea, since redox-driven conformational
the properties of films prepared using an alternating changes are believed to be functionally important in
layer-by-layer adsorption strategy [37]. Stacked bilayers this enzyme’s role as a proton pump.
can be formed starting from a layer of mercapto-
propanesulfonic acid on gold or quartz, followed by
successive layers of protein (Mb, cytochrome P450), 3. What factors determine ET rates at
polyanion/DNA, protein, etc. The progress of assembly protein– electrode interfaces?
can be monitored using a QCM. These films do not
support protein mobility and most of the resulting The most revealing experiments in this area have
electrochemical activity stems from proteins closest to been carried out with SAM-modified electrodes since
the electrode surface. Using this method, it has recently these provide the best defined surfaces. As with non-
proved possible to detect active sites in a photosyn- protein systems, it is expected that ET rates for protein-
thetic reaction center [38]. SAM-electrode systems depend on a donor-acceptor
The scope for developing membrane-mimetic surfaces distance, and indeed electrochemical studies with func-
appears very promising. Salamon and co-workers re- tionalized (X– Y) n-alkanethiol SAMs have confirmed
ported studies on two membrane bound proteins, that rates depend on the length of the spacer. Bowden
spinach chloroplast cytochrome f and beef heart mito- and co-workers used cyclic voltammetry to determine
chondrial cytochrome c oxidase integrated into a phos- electron exchange rate constants (k0) between cy-
phatidylcholine bilayer coated on a tin-doped indium tochrome c and gold across medium-length
oxide electrode [39]. Voltammetric signals due to re- HS(CH2)n COOH SAMs [16], whereas Niki and co-
versible redox couples were obtained in each case, one workers studied chain lengths as short as n =2 using ac
from cytochrome f at 365 mV and two from cy- potential-modulated UV-visible electroreflectance [41].
tochrome c oxidase, at 250 and 380 mV. These values With this method, ET rates are obtained from the
are quite close to reported values, the latter correspond- frequency dependence of the reflected light, and it is
ing to the CuA and cytochrome a electron-transfer worth noting that it can be applied to systems such as
centers of the terminal oxidase, although no catalytic ‘blue’ Cu proteins which do not have such intense
activity was described. The shapes of the voltam- optical transitions as hemes [42]. With the longer spac-
mograms suggest that these proteins are immobilized in ers 9 B nB 11, rate constants depend exponentially on
the bilayer. chain length, consistent with a through-bond tunneling
Hawkridge and co-workers have developed an inter- mechanism and a decay factor of 1.1 per CH2− group.
esting way to immobilize cytochrome c oxidase, based This result is in excellent agreement with results ob-
on the formation of a submonolayer of octadecylmer- tained for simple redox species attached to alkanethiol
captan (OM) on a metal surface (Ag deposited on Au) SAMs [43,44]. As the chain length is shortened, rate
[40]. This serves as the template/anchor for assembling constants do not increase as expected but instead ap-
a bilayer that is completed by amphiphiles (L-phos- proach a limiting value; accordingly it is proposed that
phatidylethanolamine or L-phosphatidylcholine). The electron exchange becomes gated by changes in confor-
cytochrome c oxidase/membrane mimetic electrode is mation/orientation of the protein. The implication is
prepared by entrapping an aliquot of solution contain- that the most stable orientation does not correspond
ing detergent (deoxycholate)-solubilised enzyme and with the most favorable orientation for electron trans-
amphiphile against the OM-modified electrode surface fer, so that some rate-limiting adjustment is required.
with dialysis membrane, the principle being that as the However, studies carried out on n-alkyl SAMs termi-
detergent dialyzes out, the enzyme becomes incorpo- nated with a covalently attached small-molecule redox
rated into the OM/amphiphile bilayer. The immobilized couple have indicated that complications arise here also
enzyme exhibits voltammetry in which a broad oxida- as the chain length is shortened [45]. Whatever the
tive wave observed at low scan rates is joined by a reason for kinetic limitations, it is clear that rate con-
2628 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

stants well in excess of 1000 s − 1 are readily attainable 4. Studies of protein redox reactions using
with proteins whose active sites lie quite close to the voltammetric techniques
surface that contacts the electrode. Rate constants of
this magnitude are observed also in studies of proteins An obvious advantage of voltammetric methods over
adsorbed at graphite electrodes [46]. the more conventional and widespread use of mediated
Evidence that the rates of electron transfer depend in potentiometric titrations is that a large amount of data
a very subtle way on how the protein interacts with the can be accumulated in a short time-an obvious asset in
electrode is provided by Bowden and co-workers, who such laborious tasks as measuring the variation of
have reported an intriguing result obtained for the reduction potentials with pH. Freedom from the depen-
voltammetry of horse heart and yeast cytochrome c dence on chemical titrants and mediators or the need
adsorbed at mixed SAMs [47]. With horse heart cy- for a characteristic spectral feature further facilitates
tochrome c, the ET rate constant obtained using a experiments that are not readily undertaken with tradi-
mixed SAM derived from equimolar amounts of tional methods. Some of these aspects will now be
HS(CH2)10COOH and HS(CH2)8OH is about 5-fold described before dealing with some of the more recent
faster than one comprised only of HS(CH2)10COOH. developments in which the protein is interrogated as a
By contrast, with yeast cytochrome c, the pure film on the electrode.
HS(CH2)10COOH SAM gives only a very low rate, but First, unlike potentiometric studies, there is no re-
there is a 2500-fold rate enhancement when switching quirement for a distinctive and unambiguous change in
to the mixed HS(CH2)10COOH/HS(CH2)7OH mono- some spectroscopic parameter-for example, an EPR
layer. The two proteins have similar ET reorganisation signal that may be observable only at cryoscopic tem-
energies and surface charge distribution, but there are peratures and thus require rapid freezing. A problem
large differences in amino acid composition that could with freeze quenching is that re-equilibration of elec-
trons can occur during the freezing process, particularly
affect the ET coupling to the electrode.
if quenching from elevated temperatures [52]. By con-
Continuing this theme, the question of why the elec-
trast, direct electrochemical methods allow in situ mea-
trochemical rates observed for myoglobin depend so
surements of reduction potentials. This aspect is
much on the electrode environment is unclear. Here we
exemplified in measurements of the temperature depen-
are comparing heterogeneous rate constants (cm s − 1)
dence of the reduction potential of the ferredoxin from
among different studies in which myoglobin is diffusing
Pyrococcus furiosus [52– 54]. This organism is a hyper-
in the aqueous or surfactant film phase. Rusling and
thermophile found naturally near deep sea hydrother-
co-workers have suggested that the high activity ob-
mal vents at temperatures exceeding 100°C; an
served in surfactant films is due, at least in part, to
understanding of its electron-transport chains thus re-
inhibition of adsorption of deactivating contaminants
quires knowledge of the thermodynamic properties of
[48]. Another possibility concerns the nature of the individual components measured under these extreme
ligation to iron and how this affects the ET reorganisa- conditions. Earlier potentiometric titrations, monitored
tion energy. Hawkridge and co-workers showed that by EPR spectroscopy after freezing equilibrated sam-
the CN−-ligated form displays much more reversible ples, had shown a non-linear variation of reduction
voltammetry than the native aquo-Fe(III) form at an potential at high temperature, with a discontinuity oc-
indium tin oxide electrode [49]. Variations also exist curring at 80°C [53]. A change in conformation was
between myoglobins from different species; the protein proposed, and this was discussed in terms of its rele-
from horse heart shows a higher rate constant than the vance for an organism that grows at these elevated
recombinant protein from sperm whale [50]. Arm- temperatures. However, recent in situ voltammetric
strong’s group later showed that a significant increase studies (at a glassy carbon electrode) have revealed that
in the rate constant for myoglobin electrochemistry at a no such discontinuity occurs [52,54].
pyrolytic graphite edge electrode could be achieved A second example is the ability to measure redox
simply by replacing the distal histidine (H64) by amino thermodynamics at high pressures. The variation of
acids such as glycine or leucine [51]. The structures of reduction potentials with pressure yields V 0, the change
many of these H64 mutants have been crystallographi- in molar volume between oxidized and reduced forms,
cally characterized. Histidine-64 stabilizes the H2O lig- and hence the difference in compressibility ( 8 #V 0/
·
and that is bound to Fe(III) but not Fe(II), and also #P). Sligar and co-workers used a high-pressure
links this coordinated H2O to bulk solvent water via a voltammetric cell to study cytochrome c up to pressures
chain of H-bonds; the mutants break this connection. of 5 kbar [55]. Over the range from ambient to 5 kbar,
Consequently, placing myoglobin in the more hydro- the potential increased by 76 mV, enabling the authors
phobic environment of a surfactant film can disrupt to quantify independent observations that cytochrome
these interactions and lower the reorganization energy c(II) has a more compact structure than cytochrome
for electron transfer. c(III). Smith and coworkers have applied voltammetric
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2629

very negative reduction potential and was once thought


to be of the ‘HiPIP’ type ([4Fe – 4S]3 + /2 + ). Until re-
cently, the so-called ‘clostridial 8Fe’ ferredoxins were
the only well studied class of small proteins containing
two [4Fe – 4S] clusters, and invariably these have very
similar reduction potentials, at around − 400 mV.
Through voltammetric methods, and again through the
ability to probe active sites with very negative reduction
potentials, it has recently become clear that there is a
distinct class of naturally occurring 8Fe ferredoxins in
which one of the [4Fe – 4S] clusters has a much lower
potential which prevents its reduction by dithionite
[59,60].
Even when very negative potentials do not pose a
problem, iron-sulfur clusters are often difficult to study
by conventional methods, their broad and uncharacter-
istic absorption spectra providing only poor signatures
for real-time measurements at ambient temperature.
The situation is particularly unsatisfactory if dealing
with labile, air-sensitive species. Voltammetry provides
an attractive solution to these problems, a good exam-
ple being the ability to engage and monitor reactions
occurring at [3Fe – 4S] clusters. As shown in Fig. 2,
[3Fe – 4S] clusters display two signals, one due to the
well-studied [3Fe – 4S] + /0 couple and one at very nega-
tive potential which integrates to double intensity and is
assigned to the [3Fe – 4S]0/2 − couple [61,62]. The latter
Fig. 2. Voltammograms of films of 7Fe ferredoxins on a PGE signal has the narrow width expected for cooperative
electrode at 0°C. Azotobacter 6inelandii (scan rate 20 mV s − 1, two-electron transfers and the variation of reduction
pH 7.0). Desulfo6ibrio africanus (scan rate 191 mV s − 1, pH potentials with pH shows that 2 – 3 H+ are also
7.0) and Sulfolobus acidocaldarius (scan rate 10 mV s − 1, pH transferred.
7.4). In each case an electroactive coverage of approximately
Proton transfer is not a widely acknowledged prop-
one monolayer is obtained in the presence of polymyxin as
erty of Fe– S clusters, but [3Fe – 4S] clusters appear to
co-adsorbate. Signals A%, B% and C% refer to the redox couples
[3FE – 4S] + /0, [4FE – 4S]2 + / + and [3FE – 4S]0/2, respectively. be an interesting exception. The two signals serve as a
diagnostic feature of the presence of [3Fe – 4S] clusters
in proteins, and this has been exploited in studies of the
methods to study both high-temperature and high-pres-
potential-dependent transformations to cubane adducts
sure effects on rubredoxins (Rd, active site = Fe(SR)4)
[4Fe – 4S] or [M3Fe– 4S] that are observed for the ferre-
from a mesophile (Clostridium pasteurianum) and hy-
doxins from Desulfo6ibrio africanus and P. furiosus
perthermophile (P. furiosus) [56]. The effects are appre-
[62– 65]. These redox-coupled reactions (see below) are
ciable, for example the reduction potential of P.
conveniently induced and monitored in proteins ad-
furiosus Rd decreases from 31 mV at 25°C to −93 mV sorbed at a pyrolytic graphite edge electrode. Several
at 95°C, while reduction potentials increase slightly novel clusters have been ‘synthesised’ in this way, in-
with pressure. cluding those incorporating Tl, Cu and Pb. The result-
Highly reducing species with reduction potentials ing [M3Fe– 4S] clusters undergo reversible ligand
that are too negative to access with conventional chem- exchange, which presumably occurs at the new subsite
ical titrants such as dithionite can be studied by voltam- M [66].
metric methods. The result of this is that there are now Coupling between electron transfer and chemical/
many examples of proteins containing a [4Fe – 4S]2 + / + conformational changes is conveniently depicted in
cluster with reduction potentials more negative than terms of a square-scheme such as that shown (Scheme
− 600 mV. In particular, voltammetry has made possi- 1), in which the species are interconnected by electro-
ble the extensive characterisation of crystallographi- chemical (E) and chemical (C) transformations [66].
cally-defined mutants of the 7Fe ferredoxin from The latter include protein/prosthetic group conforma-
Azotobacter 6inelandii in which the environments of the tional changes, and transfer of ions (including protons).
[3Fe – 4S] and [4Fe – 4S] cluster are altered systemati- Whereas equilibrium methods like potentiometry re-
cally [57,58]. The [4Fe – 4S] cluster in this protein has a veal only the thermodynamics of these systems, dy-
2630 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

namic methods provide kinetic information. In biology, activates the enzyme to catalyze peroxide reduction
electron transfer drives chemical reactions; equally im- [72].
portant, perhaps, are ‘gated’ electron-transfer reactions, The redox-coupling experiments outlined above in-
in which electron-transfer rates are determined (gated) volved free protein molecules undergoing reversible dif-
by the preceding chemical process. Voltammetry is a fusion-controlled electrode reactions. Alternatively,
powerful tool for resolving and quantifying these com- coupled chemistry can be studied in adsorbed proteins.
plex reactions, and several examples involving cy- Immobilising the protein sample creates important ad-
tochromes and Fe– S clusters have recently been vantages in terms of sample size and response time, and
reported. even cyclic voltammetry can provide kinetic resolution
The redox properties of cytochromes depend greatly on the millisecond timescale or lower, and allow the
on the nature of the axial ligation to iron. In particular, detection and measurement of electron-transfer gating
since Fe(II) may prefer different ligands to Fe(III), [73].
electron transfer can drive changes in axial coordina- Many redox reactions in proteins are coupled to
tion and vice versa, and the resulting kinetics can be proton transfer, and indeed this forms the basis of
observed and deconvoluted by electrochemical meth- energy conservation whereby transfer of electrons down
ods. Studies by Haladjian and co-workers [67], then a thermodynamic gradient is used to pump protons
Barker and Mauk [68] revealed the ligand interchange uphill so that energy is stored as a proton gradient.
kinetics accompanying redox cycling of mitochondrial Some of the details of how this can occur at an active
cytochrome c. Under mild alkaline conditions, the axial site have been revealed by a study of the [3Fe – 4S]
methionine ligand to Fe(III) is easily replaced by com- cluster in A. 6inelandii 7Fe ferredoxin [73]. One-electron
peting ligands, and cyclic voltammograms recorded on reduction (to give [3Fe – 4S]0) is coupled to the uptake
appropriate timescales (i.e. scan rates of the order of a of a single proton: this further example of proton
few volts per second) reveal the interconversion kinet- transfer at Fe– S clusters is significant because high-res-
ics. Interestingly, this type of problem can be tackled by olution crystal structure studies on the different oxida-
other electrochemically-based methods, notably tion and protonation states reveal (a) that the cluster is
Hawkridge and co-workers used double potential step buried with no access for water molecules, and (b) that
chronoellipsometry, in which circular dichroism is mon- a carboxylate group from an aspartate (D15) is located
itored through an optically transparent indium oxide close to the cluster on the protein surface. The mecha-
electrode [69]. Feinberg and co-workers have recently nism of proton transfer between the cluster and solvent
studied a mutant form of yeast cytochrome c, in which has been determined using protein film voltammetry in
phenylalanine-82, situated close to Met-80, is replaced conjunction with site-directed mutagenesis. As indi-
by histidine [70]. This directs the formation of a well- cated in Fig. 4, data analysis procedures (‘trumpet
defined Fe(III) state in which His-82 replaces the origi- plots’) compare the scan rate dependence of voltammet-
nal Met-80 ligand present in Fe(II). The results are ric peak positions to those expected for suitable kinetic
depicted in Fig. 3. Recently Mauk and co-workers have models. With the D15N mutant, proton transfer is very
been able to detect states formed in mutants in which slow, and oxidation of [3Fe – 4S]0 is gated by release of
one or more of the lysine residues that normally replace H+. The native protein exhibits much faster proton
Met-80 in the Fe(III) form are replaced by alanine [71]. transfer, and rate constants vary with pH in accordance
Even more complex transformations occur in the with the proton being transferred by the D15
di-heme enzyme cytochrome c peroxidase from P. deni- carboxylate.
trificans. Here, cyclic voltammetry displays how reduc- As mentioned above, [3Fe – 4S] clusters are unique in
exhibiting further redox chemistry in the form of a
tion of one heme induces a change at the other, and
two-electron/two-proton reduction of [3Fe – 4S]0 to
yield a novel ‘hyper-reduced’ state in which all three
iron atoms are formally 2+ [61]. The physiological
significance of this reaction is unclear. However the
achievement of formal oxidation levels ranging from
all-Fe(III) to all-Fe(II) is unprecedented for Fe– S clus-
ters, as is the coupling to multiple proton transfer. The
process occurs at potentials of around − 700 mV at pH
and may involve some rearrangement of the cluster
environment. Recent experiments using fast scan cyclic
voltammetry reveal that the hyper-reduced 2-state can
undergo a very fast and reversible two-electron/two-
proton oxidation, yielding a state that differs from the
Scheme 1. normal ‘0’ level into which it converts within a second
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2631

Fig. 3. Top. Experimental cyclic voltammogram (circles) and simulation (solid line) for Phe82His yeast Iso-1-cytochrome c (in
solution) obtained using a gold electrode modified with a monolayer of bis(4-pyridyl) disulfide. Scan rate 50 mV s − 1. Bottom.
Square scheme showing redox-coupled interchange of Met-80 and the residue His-82 that is introduced. The voltammogram is
simulated using an equilibrium constant KR/R% = 3.6 × 105 (i.e. R favored) and k (R%“ R) =10 s − 1. Adapted from original figure in
reference [70], with permission.
2632 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

Interesting voltammetry has been observed for the


iron storage protein ferritin. This protein has a giant
cavity which can accommodate over 4000 Fe(III) min-
eralized as a hydrated oxide. Although it is widely
accepted that uptake and release of iron involves reduc-
tion to Fe(II), details of the processes involved are yet
to be established. Zapien and coworkers observed that
ferritin adsorbs at a tin-doped indium oxide electrode
to give films of submonolayer coverage, thus providing
the opportunity to control and observe iron mobiliza-
tion [75]. The voltammetry is not straightforward; in
the first reductive scan, two large reduction peaks ap-
pear at potentials of approximately − 100 and − 300
mV versus SHE, and correspond to two different forms
of the protein. Subsequent cycles show just single oxi-
dation and reduction peaks at an average potential of 0
mV. However, in the presence of EDTA, the voltam-
meric peaks vanish after only one cycle, showing that
iron is released and then complexed. This establishes
the capability of voltammetry to probe this complex
chemistry and further studies should provide some in-
teresting alternative perspectives on this important
problem.
For electron-transport enzymes, an important crite-
Fig. 4. Plots of peak position as a function of scan rate rion for voltammetric experiments should be that elec-
(logarithmic scale) for proton-coupled reduction and reoxida- trons exchanged with the electrode should be translated
tion of the buried [3FE – 4S] + /0 cluster of the D15N mutant of into catalytic action. Without this, there must be doubt
Azotobacter 6inelandii ferredoxin, in which proton transfer is about whether or not the enzyme is catalytically active
retarded. The detailed kinetics and energetics can be analyzed
or able to function with the electrode as a redox
by simulation of the shapes of the ‘trumpet plots’. In this case
partner. A theoretical analysis of electrocatalytic wave-
the black diamonds correspond to pH 8.34 (pH pK, thus no
protonation occurs) while the gray diamonds correspond to forms due to adsorbed enzymes has been described, the
pH 5.50 (pH  pK). The solid line is a simulation with pK= aim being to recognize different modes of kinetic con-
6.9, and the proton release rate koff is 2.5 s − 1. The shapes are trol and extract mechanistic information [76]. The inter-
explained as follows using the scheme shown above. At pH\ face consisting of electrode, enzyme and electrolyte is
pK, no proton transfer occurs and the peaks separate in an considered as a three-resistor series through which cata-
approximately symmetric fashion. By contrast, at pH  pK, lytic current flows. Three limiting cases have been con-
the reduced cluster becomes protonated and three regimes are sidered, corresponding to control by interfacial electron
observed: (a) at low scan rate, the peaks lie close together at transfer (least useful), substrate mass transport, and
the formal reduction potential which is at a more positive enzyme catalysis (most useful).
value because protonation is coupled to electron transfer; (b) Farmer and co-workers have discovered that myo-
at intermediate scan rates, electron transfer to the cluster is
globin entrapped in surfactant films catalyzes interest-
followed by proton transfer (E1CR), but re-oxidation is pre-
ing transformations of NO and NO2− [33]. In this
vented because the proton ‘off’ rate is too slow and gates the
reaction, so that no return peak is observed; (c) at fast scan confined and controllable state, myoglobin provides an
rates the electron is ‘re-called’ from the reduced cluster before excellent small-protein model system for examining the
protonation can occur, thus giving peaks that coincide in NO metabolizing activities of certain heme enzymes.
position with the high pH experiment, i.e. protonation is Some results are shown in Fig. 5.
decoupled (E1 only). Reference [46]. Under N2, Mb confined within a DDAB film at
basal-plane graphite displays two reversible redox cou-
ples, assigned to Fe(III)/(II) and Fe(II)/(I) transforma-
tions of the active site. The protein appears to diffuse
[73]. These results are supported by recent experiments within this film, as noted also by the Rusling group.
carried out at − 85°C in which the protein coated When N2 is replaced by NO, the voltammetry changes
electrode is transferred to a cryosolvent (70% dramatically, with the appearance of a catalytic reduc-
methanol), enabling this chemistry to be observed at tion peak at − 0.75 V versus SCE which is dependent
slow scan rates [74]. The results suggest that the cluster on the partial pressure of NO but independent of pH
is performing disulfide chemistry, although its physio- over the range 5.5– 10. Controlled potential electrolysis
logical relevance is unclear. with mass spectrometric analysis of the head gas reveals
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2633

that the major product is N2O. A film of pre-formed Mb confined to these surfactant films is also active in
MbNO cycled under a N2 atmosphere shows a re- oxygenation reactions. Under aerobic conditions, large
versible couple at − 0.87 V, which persists indefinitely catalytic reduction currents are observed, which are
at high pH provided the potential is cycled sufficiently ascribed to the reduction of coordinated O2 to produce
fast to outrun an irreversible chemical reaction which H2O2 [77]. However, the oxy-intermediate that is pro-
results in the reappearance of Fe(III)/(II) and Fe(II)/(I) duced in this reaction is active also in oxygenating
signals. Piecing these observations together and draw- substrates such as styrene. Styrene oxidation is not
ing on known reactions of NO species in solution, the rate-determining in the electrochemical mechanism but
authors have modeled the MbNO electrochemistry in the transformations are clearly evident from glc analy-
terms of the following hypothesis and scheme (Scheme sis of bulk electrolysis products.
2) the result of which is the reduction of NO to give For adsorbed redox couples, half-height widths vary
N2O. as 1/n, while peak heights depend on n 2, where n is the
Production of NO2 is accounted for by side reactions number of electrons transferred in the process. The
including reduction of the Fe(II) nitrosyl Mb which is significance of this relationship is that cooperative two-
electron active sites appear much more prominent in
followed by irreversible dissociation of reactive NO−.
voltammograms and can quite easily be diagnosed. This
Mb-DDAB films catalyze reduction of NO2− in reac-
is illustrated in the voltammetry of yeast cytochrome c
tions that produce NH3, NH2OH, N2 and N2O [33].
peroxidase, an enzyme that catalyses reduction of hy-
drogen peroxide by cytochrome c and which has long
been a valuable system for studying oxygen activation
by hemes. Adsorption of cytochrome c peroxidase
(Fe(III) resting state) at a pyrolytic graphite edge elec-
trode from dilute ice-cold 20 mM phosphate buffer
gives rise to a prominent signal at + 750 mV versus
SHE (pH 5.4) [78]. The signal consists of oxidation and
reduction peaks that each have half widths below 80
mV, and it is therefore assigned to a cooperative two-
electron couple, i.e. oxidation of Fe(III) to a state that
is equivalent to ‘Fe(V)’. The peaks transform to a
sigmoidal catalytic reduction wave when H2O2 is
added. Studies of the rotation rate and peroxide con-
centration dependence yielded a turnover number com-
parable to that reported from solution studies, and it
was concluded that the oxidized form of the redox
couple is either ‘compound I’ (Fe(IV) and radical) or a
closely related species. The electron-exchange kinetics
of adsorbed cytochrome c peroxidase were also mea-
sured using staircase cyclic voltammetry, although in-
terpretation of the value of approximately 6 s − 1 is
complicated by the fact that two electrons are trans-
ferred [79]. Comparisons have been made with a mu-
tant W51F in which the distal tryptophan (not the same
tryptophan that houses the radical) changed to pheny-
lalanine. The mutant has a more positive reduction
potential ( +883 mV at pH 5.4) and higher catalytic
activity, but is less stable at the electrode. The result
showed that this tryptophan plays a significant role in
stabilizing the Fe(IV) heme, possibly by creating favor-
able interactions with the bound oxide ion.
The flavin group provides a further illustration of
Fig. 5. Catalytic reduction of NO by myoglobin incorporated
how cooperative two-electron centers may be particu-
in a surfactant film. Cyclic voltammograms in deaerated pH
7.4 phosphate buffer, 0.1 M NaBr, scan rate 500 mV s − 1: (a)
larly ‘visible’ to voltammetric methods. Normally, FAD
graphite electrode cast with a DDAB film; (b) myoglobin or FMN cofactors in enzymes are detected and studied
incorporated into DDAB film; (c) as for (b) but solution via their visible absorption spectrum (400– 500 nm) or
saturated with NO. Modified from original figure in reference occasionally through EPR if a one-electron radical is
[33]b, with permission. stabilized. Flavocytochrome c3, a periplasmic fumarate
2634 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

Scheme 2.

reductase from the marine bacterium Shewanella turnover’ signals consist of (a) oxidation and reduction
frigidimarina, provides an interesting example of an envelopes covering the range 0.1 to −0.2 V, each
enzyme in which the FAD is very difficult to observe dominated by a prominent peak (the covalently-bound
because its spectrum is obscured by the intense Soret FAD) which overlays the weaker features due to two
bands from four heme groups; however, it is rendered Fe– S clusters, and (b) a weaker feature at − 0.3 V
clearly visible by voltammetry. The enzyme adsorbs which is assigned to a [4Fe – 4S] cluster having a singu-
from dilute solution at a pyrolytic graphite edge elec- larly negative reduction potential. Addition of fumarate
trode to give a dense coverage equivalent to an elec- gives rise to an intense sigmoidal catalytic reduction
troactive monolayer [80]. The resulting voltammetric wave whose amplitude is dependent on rotation rate.
signal consists of complex reduction and oxidation As mass-transport control is relaxed, i.e. at high fu-
peaks enveloping the contributions of all redox centers. marate concentrations and high rotation rates, the cata-
As shown in Fig. 6, the FAD cofactor is clearly visible
as a sharp and prominent component of this envelope.
Despite not being covalently bound, the FAD does not
dissociate from the adsorbed enzyme. The adsorbed
enzyme is very active: when fumarate is added, the
complex envelope collapses to a sigmoidal wave the
magnitude of which is very dependent on electrode
rotation rate.
The ability to probe activity as a function of poten-
tial makes it possible to detect even the subtlest of
changes in rate that may accompany redox transforma-
tions of centers in an enzyme. This may indicate a
regulatory activity for certain centers or help confirm
their involvement in catalytic electron relays. An inter-
esting example is the fumarate reductase from E. coli.
This is a membrane-bound enzyme consisting of four
subunits, two of which are catalytic and membrane-ex-
trinsic (i.e. protrude from the membrane surface) while Fig. 6. Baseline corrected voltammogram for a film of flavocy-
the other two are membrane-intrinsic anchors and tochrome c3 (Shewanella frigidimarina) adsorbed at a PGE
provide a binding site for the natural electron donor electrode in the presence of polymyxin as co-adsorbate. Tem-
menaquinol. A soluble form consisting only of the perature 0°C, pH 6.0. The FAD cofactor is clearly observed
above the redox transitions due to the four heme groups. The
membrane-extrinsic subcomplex can be obtained, either
different components have been resolved (gray lines within
by chemical resolution or genetic engineering. This envelope) and the resulting overall simulation is overlaid on
adsorbs strongly at a rotating disk pyrolytic graphite the experimental result with excellent agreement. When fu-
edge electrode and yields informative voltammetry marate is added to the solution and the electrode is rotated,
[81,82]. In the absence of fumarate, the observed ‘non- the signals transform into a strong catalytic wave.
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2635

lytic wave exhibits a second boost in current close to ported by Bianco and Haladjian, but details on the
the potential of the [4Fe – 4S] cluster. This shows that enzyme properties were not revealed [87]. More re-
despite its low potential, this center is somehow in- cently, the all-Fe enzyme from Megasphaera elsdenii
volved in catalytic electron transfer. has been studied [88]. The enzyme adsorbs at a glassy
Mitochondrial complex II (succinate – ubiquinone ox- carbon electrode, and from the relative catalytic cur-
idoreductase) is structurally similar to fumarate reduc- rents obtained in either direction it was possible to
tase (two membrane anchors and two determine electrochemically the degree to which it is
membrane-extrinsic subunits containing a covalently- biased to operate in the direction of H2 production. The
bound FAD and three Fe– S centers) and can also be NiFe enzyme from Chromatium 6inosum has recently
prepared as a membrane-extrinsic soluble form (succi- been studied; in this case ‘non-turnover’ signals are
nate dehydrogenase) which adsorbs on a pyrolytic revealed quite clearly in the CO-inhibited enzyme,
graphite edge electrode. In this state it is catalytically which are tentatively assigned to the [3Fe – 4S] cluster
active, but the response is not stable and the coverage is and the two [4Fe – 4S] inhibited enzyme, that are tenta-
too low to observe signals from the active sites. How- tively assigned to the two [4Fe – 4S] clusters (each at
ever, the short-lived activity is informative [83]. Experi- approximately − 0.3 V) and the [3Fe – 4S] cluster at
ments conducted with a 1:1 mixture of fumarate and − 0.03 V [89]. Catalytically, the enzyme shows a clear
succinate produce a characteristic current response: suc- preference for H2 oxidation over proton reduction.
cessive cycles show succinate oxidation at high poten- Oxidation of H2 (10% atmosphere) is mass-transport
tial and fumarate reduction at low potential, each controlled, as revealed by strong sigmoidal currents
decreasing in amplitude but crossing at isosbestic po- which are very dependent on electrode rotation rate,
tentials in each direction. The average values of these and the estimated turnover number is in excess of 1500
potentials (at which there is no net reaction) corre- s − 1, considerably higher than values obtained by con-
sponds to the formal reduction potential of fumarate. ventional dye reduction assays in solution. Significantly,
Catalytic waveforms at different pH values are deter- the H2 oxidation activity is very high at potentials
mined by computing difference voltammograms, i.e. below that of the [3Fe – 4S] cluster, so that this site must
subtracting an early cycle from a later one. The result- remain reduced throughout turnover.
ing profiles show: (a) that at pH values above 7.4 the
enzyme is actually biased to operate in the direction of
fumarate reduction, and (b) that although fumarate 5. Mediated protein electron transfer
reductase activity is high, it is limited to just a narrow
region of potential. As the potential is made more Most of the proteins discussed above have as their
negative, intending to drive fumarate reduction faster, biological function the transfer of electrons to other
the activity shuts down. Because this resembles negative proteins. They typically function in ordered structures
resistance, the behavior was termed the ‘tunnel diode’ such as mitochondria and the redox active centers are
effect [81]. The decrease in activity can be fitted to a generally accessible to the outer surface of the protein.
two-electron reaction that occurs close to the potential In contrast, the redox reactions catalyzed by a class of
region expected for the FAD. This suggests that in its enzymes known as oxidoreductases involve small
FAD-reduced form, the enzyme favours a less-active molecules. Thus one of the two required substrates
conformation which is avoided when the steady-state (redox couples) for the enzyme will be designated the
potential is maintained at a higher value. In support of substrate and other the co-factor or co-substrate. Be-
this, non-electrochemical assays showed a negative or- cause the catalytic center of the enzyme is buried well
der in viologen concentration, i.e. the rate increases as within the protein, communication between this center
the reductant is consumed [84]. Succinate dehydroge- and an electrode serving as a source or sink for elec-
nase from E. coli gives the same result as the beef heart trons is frequently poor or non-existent. Because oxi-
enzyme despite there being a significant difference in doreductases selectively catalyze reactions of analytical
the isoelectric point and only 50% amino acid similarity importance, it is of interest to efficiently couple the
(mostly confined to those residues believed to lie in the enzyme electron transfer to the electrode, which then
active site) [85]. These results argue against, but do not effectively serves as one of the substrates. With a few
rule out the possibility, that the ‘tunnel diode’ effect is exceptions such as horseradish peroxidase (HRP)
due instead to a potential dependent re-orientation. [90,91], heterogeneous electron transfer rates for en-
Armstrong and co-workers used this system to explore zymes are vanishingly low so that the enzyme redox
how voltammetry can be used to observe and resolve activity must be coupled to the electrode using media-
H/D isotope effects in electron-transport catalysis [86]. tors whose redox states are alternatively recycled at the
Hydrogenases have featured in several reports, and electrode, at adjacent mediators, and the enzyme active
useful information has been obtained. An early exam- site. If successfully accomplished, the rate of recycling
ple of faradaically monitored catalytic activity was re- of the redox state of the enzyme then becomes propor-
2636 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

tional to the rate of the enzymatic reaction and there- principle, the sensor response would no longer be sensi-
fore to the concentration of the substrate of interest. tive to fluctuations in S%ox. This condition can only be
Such a device is known as an enzyme electrode and has met if the rate of electron transfer via the ‘wired’ route
been widely employed as an analytical tool, especially is high compared to the rate of reaction of the enzyme
for the determination of glucose and lactate. There are with endogenous oxygen. Most mediator-based sensors
two broad classes of oxidoreductases: oxidases that use involving oxidases do not meet this requirement and
oxygen as a co-substrate and dehydrogenases that typi- exhibit parasitic effects of oxygen on their response
cally use NAD or NADP as a co-factor. It should be which in some cases are as large as 70% [94]. Willner’s
said at the outset that critical properties of the enzyme group has elegantly addressed the question of whether
electrode as a practical device (sensitivity, stability, size, it is possible to ‘wire’ an enzyme with sufficently high
selectivity and the medium in which measurements will efficiency to prevent competition with oxygen [95]. This
be made, etc.) will determine whether the device is objective was apparently met by removing the non-co-
useful or not. One important goal is to make the sensor valently bound FAD redox center from the enzyme
‘reagentless’, meaning that nothing has to be added to (GOx) and reconstituting the enzyme on a tether con-
the medium for proper functioning. This requirement is sisting of cystamine chemisorbed on a gold surface, a
a major reason for the overwhelming popularity of pyrroloquinoline quinone (PQQ) link, and FAD. The
oxidases, because many media have sufficient oxygen transfer of electrons from the FAD to the PQQ and
present that none need be added. The reaction sequence thence to the electrode was sufficiently rapid that a high
is: turnover number (maximum rate for the enzymatic
reaction) was obtained (900 9 150 s − 1), which com-
Sred + Eox ? Sox + Ered (1) pares favorably with a value of about 1000 s − 1 for the
S%ox + Ered ? S%red + Eox (2) enzyme in solution. There is only a monolayer of
enzyme on the electrode surface, thus aiding efficient
If, for example, glucose oxidase (GOx) is the enzyme communication. Such a device might have, however,
employed, then Sred, Sox, S%ox and S%red correspond to only limited stability because of the small amount of
glucose, gluconic acid, oxygen, and hydrogen peroxide, enzyme present. Heller’s group has examined a variety
respectively. The goal is to make reaction 1 rate limit- of approaches to ‘wiring’ of enzymes with the goal of
ing, and this necessarily requires a large excess of finding a mediator that reacts rapidly with the enzyme,
oxygen (S%ox). The overall rate of the reaction is moni- but has also a low potential. In the case of oxidases, the
tored either by following the rate of consumption of formal potential of the flavin redox center is low
oxygen (requires two measurements: one in the presence enough that a relatively weak oxidant with a potential
of enzyme, one in its absence) or the formation of in the range of + 30 to −100 mV versus AgCl/Ag
hydrogen peroxide. The advantage of the former reference would be sufficient to drive reaction 2 to
method is the selectivity that can be achieved using a completion without at the same time effecting the elec-
gas permeable membrane, the disadvantage being the trocatalytic oxidation of species such as ascorbate. The
added complexity of the device. In the latter case an approach has involved the covalent attachment of re-
applied potential of around 600 mV versus AgCl/Ag dox centers such as [Os(bpy)2Cl] + /2 + [96] or [Os(4,4%-
reference electrode is needed, and in this region numer- dimethoxy-2,2% bipyridine)2Cl] + /2 + (E o% = 0.035 V
ous endogenous species such as ascorbic acid, uric acid, versus SCE [96] to polymeric hydrogels. The low poten-
and a variety of biogenic amines are electroactive. A tial of the mediator reduces but does not eliminate the
major and largely successful solution to this problem is electrocatalytic oxidation of ascorbate. Although the
the use of permselective membranes to exclude interfer- hydrogels are somewhat fluidic, the redox centers are
ing species. The electrochemistry of hydrogen peroxide still not mobile enough to compete favorably with
at physiological pH at a platinum electrode is actually freely diffusing oxygen. Some success in reducing oxy-
quite complicated and has been the subject of recent gen dependence has been achieved by employing mem-
study [92,93]. The rate of electron transfer depends on branes that will ‘salt out’ oxygen so that it does not
the potential-dependent formation of Pt(II) sites on the reach the enzyme layer [97].
electrode surface. A number of other mediators have been tried includ-
Mediated electron transfer has therefore attracted ing ferrocene derivatives [98], phenoxazines, phenothi-
considerable attention not only for reasons of funda- azines, and Wurster’s salts [99] and the conclusions are
mental interest but also to make a more practical consistent: mediated systems frequently show poor sta-
enzyme electrode. Mediators should be distinguished bility when operated continuously for extended (hours –
from electrode modifiers which are not themselves elec- days) periods of time. In general, it appears that the
troactive at the potential of the protein electron transfer demise of the mediator itself, through leaching or chem-
reaction. If reaction 2 above could be replaced by one ical degradation, is the primary problem [100]. A solu-
involving coupling or ‘wiring’ to the electrode, then, in tion has been the dissolution of the mediator in carbon
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2637

paste to provide a largely renewable supply. Wang has where few endogenous species in biological fluids are
proposed this strategy for ‘storing’ oxygen [101]. This electroactive, the disadvantage is that HRP can non-
approach is acceptable provided that there is no prob- specifically catalyze the oxidation of various organic
lem with soluble mediator leaking into the sample. On molecules present in the medium. Direct electron trans-
the other hand, electrochemically-based ‘fingerstick’ fer of HRP has also been demonstrated at suitably
strips which aid diabetic patients in monitoring blood prepared carbon electrodes [90].
glucose use non-covalently immobilized mediators such Several fundamental studies have enhanced under-
as ferrocene and ferricyanide, but such devices need standing of electron transfer between mediators and
only to give a reliable response for perhaps 30 s [102]. enzymes: investigation of single electron donors and
Enzymes have also been adsorbed on or trapped within acceptors (largely ferrocene derivatives) shows that
electrodeposited conducting polymers [103,104]. their reactivities do not follow simple outer sphere
One mediator system that yields surprisingly stable electron transfer [108]. Mikkelsen and co-workers [109]
sensors is the so-called organic salt electrode. This is have developed the theory for intramolecular electron
based on the use of an electrode coated with the transfer involving the multiple attachment of ferrocene
organic salt tetrahydrofulvalene (TTF) tetra- derivatives to GOx. The highest rate of intramolecular
cyanoquinodimethane (TCNQ). First reported by electron transfer was observed for a ferrocene car-
Kulys [105] in the 1980s it became the focus of signifi- boxylic acid (0.9 s − 1). If the rate of reaction with
cant study. It is thought to mediate electron transfer oxygen with the enzyme is assumed to be 2 × 106 M − 1
with GOx, resulting from the corrosion of the electrode s − 1, then the ‘relay rate’ would have to be ] 5 ×103
to produce a layer of soluble cationic TTF+ close to s − 1 if the sensor is placed in an air saturated solution
the electrode surface that can reoxidize the enzyme ([O2] =240 mM). The observed rates are far short of
according to reaction 2 [106]. It has been suggested that this limit, again emphasizing the influence of oxygen on
TTF+ is also not stable [99], thus the apparent sensor the relay process. The theory for electron transfer to
stability is probably due to the replenishment by corro- enzymes through mediator layers has also been devel-
sion of the TTF+ available for electrocatalysis. Khan oped [110,111].
[107] has investigated in some detail the importance of There are over 300 enzymes that employ either
the morphology of the organic electrode surface in NAD(P) or NAD(P)H as a co-factor, and for this
defining the sensitivity and stability of a glucose sensor reason the issues surrounding both the immobilization
constructed from this material. A high rate of mediator of such enzymes (dehydrogenases) and the recycling of
efficiency was achieved due in part to the nature of the the NAD/NADH couple become important. Many
TTF+TCNQ− salt deposit that was dendritic and electrochemical studies of NADH oxidation have been
shaped like a tree. This could have caused efficient carried out following the classical work of Elving and
corrosion with the result that the sensor showed very co-workers in the 1970s [112]. The objective has been to
little dependence on oxygen. Ascorbate is a potential find a modified electrode that will reduce the overpo-
interferent to the operation of sensors using this system, tential for NADH oxidation and prevent or minimize
but in this case the interference was about 1% at 10 adsorption of oxidation products on the electrode sur-
mM glucose. The apparent Michaelis constants for face. A number of mediators including phenoxazine
glucose, that are again indicative of efficient mediation, derivatives [113], catechols [114] bipyridines, metal
were in the range of 20 – 60 mM, depending on the complexes [115], and hydroxybenzaldehydes [116] have
applied potential. been used for this purpose. A significant advantage of
Another approach to the facilitation of electron the latter materials, especially 3,4 dihydroxybenzalde-
transfer is the use of colloidal gold particles [91]. HRP hyde, is the property of improved stability and the
is immobilized on gold particles (30 nm diameter) possibility of electrodeposition to form a chemically
which are then deposited on an electrode surface. By modified electrode. The pseudo first order rate con-
comparison, efficient coupling of the enzyme to a stants for reaction with NADH are in the range of
smooth gold electrode is not observed, due no doubt to 2.6× 103 M − 1 s − 1. The above functionalities have
the ability of the colloidal gold particles to interact been attached to polymers [117] that are then deposited
intimately with the protein. The authors postulate that on the electrode surface. There has been considerable
only the monolayer of gold particles in direct contact discussion over the years as to whether the oxidation of
with the electrode is really effective and that the addi- NADH involves hydride transfer (one step) [118] or
tion of mediators to ‘connect’ subsequent layers does whether instead the process involves three steps (elec-
not appear to have any effect. Hydrogen peroxide tron/proton/electron) but within one complex [119].
produced by reaction 2 above oxidizes the HRP and a The goal of the chemically-modified electrode is to
potential of 0 V versus AgCl/Ag is required to reduce lower the barrier to electron transfer and prevent the
the HRP to its native state. The advantage of this dimerization of the one-electron oxidation product. The
arrangement is that the applied potential is in a range quinone-like structures offer the possibility of meeting
2638 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

Fig. 7. Schematic of Integrated lactate dehydrogenase enzyme electrode. Reference [122]a with permission.

such requirements. When NAD is used as a co-factor it tor. The stability is moderate and the apparent KM for
is generally soluble and freely diffusing. Numerous lactate is about 3.6 mM. [120]. The oxidation of
attempts have been made to immobilize both the en- NADH can be carried out bioelectrocatalytically using
zyme and the co-factor without notable success. How- an enzyme (diaphorase), and a variety of quinones or
ever, Willner’s group has developed a scheme whereby flavins have been employed as mediators [121].
the NAD is first immobilized via a PQQ-containing
tether that serves, in effect, as a template for lactate
dehydrogenase binding. After the enzyme is in place, it 6. Preparation of ordered protein layers
is then cross-linked with glutaraldehyde to stabilize the
layer. The ordered structure is shown schematically in The desire to couple electron transfer from an en-
Fig. 7. This is the first clear example of a fully inte- zyme layer to an electrode has additionally focused
grated and ‘reagentless’ system where NAD is a cofac- attention on the ordered immobilization of protein. The
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2639

strategy is to build up successive layers of enzyme so any other proteins present, selectivity is not a major
that the resulting structure can be studied systemati- problem. In passing from the cell culture to the brain
cally. Four approaches have been employed: (1) Build- slice or intact brain as the measurement medium, sev-
ing of successive layers of enzyme by reaction of the eral issues arise that must be considered. First, the
enzyme with bifunctional linkers (tethers) [122]; (2) Use insertion of the sensor into tissue creates tissue damage,
of biotinylated enzyme linked in successive layers using which is minimized as the diameter of the sensor de-
avidin [122– 124]; (3) Linking of successive layers using creases. Implantation can result in the creation of a
electrostatic immobilization [110]; (4) Linking of succes- liquid layer immediately surrounding the sensor that
sive layers using antibodies [125]. Fundamental studies has diffusional and other properties different from the
based on these ordered to quasi-ordered structures have tissue itself. Second, species diffuse through tissue nec-
employed freely diffusing mediators as well as media- essarily more slowly, with the result that diffusion
tors covalently attached to the enzyme [126]. The an- coefficients can be a factor of 3 – 4 smaller in tissue than
choring of the tether to the electrode can block the in aqueous solution. This issue has been dealt with in
surface thus reducing the effective electrode surface the context of transport of choline in rat brain [132].
area available for amperometric detection of substrates. Finally, the sensor may lose sensitivity compared to its
This can be serious problem especially when using in-vitro value as a result of its contact with biological
microelectrodes. As a solution to this problem, Kuhr fluid. The observed signal due to neurotransmitter re-
and co-workers [127] have developed techniques for lease may be distorted due to the finite response time of
preparing nanostructure domains (photopatterned lines the sensor. If the response time is known it is possible
about 5 mm wide) on carbon electrodes in which alter- to deconvolute the signal, taking the sensor properties
nate lines are ‘bare’ electrode and tether anchor into account [133]. Electrochemical measurements in
(biotin). the brain have been recently reviewed [134].
Much of the research involving in-vivo electrochem-
istry has exploited the electroactivity of small molecules
7. In-vivo electrochemistry such as catecholamines. Many, if not most, relevant
species, are not electroactive, and therefore they must
The advances in electrochemical detection of species be determined indirectly. Because of its importance to
in biological milieu in the 1990s were based on a desire the treatment of diabetes, glucose sensors have become
to ask more detailed questions about the real-time the ‘holy grail’ in the biosensor area. There are many
fluctuations of biologically relevant analytes. This ne- approaches being taken to the monitoring of blood
cessitated the development of sensors with dimensions glucose, but at present electrochemical detection ap-
in the 1 – 10 mm range and with response times (90% pears the most tractable. Although the electrochemi-
maximum response) of a few seconds or less. Wightman cally-based glucose sensor has been known since the
and co-workers [128] observed current transients at 1960s [135], it was not until the mid-1980s that feasibil-
microelectrodes that they attributed to the single vesicu- ity of in-vivo monitoring was demonstrated [136]. This
lar release of epinephrine and norepinephrine from approach [137] has involved a needle sensor (enzyme
isolated adrenal medullary cells. The signal, obtained electrode of about 250 mm diameter) that can be im-
under constant potential amperometric conditions, is planted in the subcutaneous tissue of the forearm or
manifested as a series of spikes ranging from 0.2 to 2 abdomen for periods of 4 – 7 days, after which they are
pcoulombs (1– 10 attomol catecholamine detected). One removed and replaced. Such a sensor is shown in Fig. 8.
micron diameter electrodes were used to monitor exocy- The sensing element is built on a Pt– Ir wire substrate,
tosis from a vesicle whose diameter is B200 nm. This followed by a permselective membrane to exclude elec-
work established that it was possible to directly moni- troactive interferences such as ascorbate, urate, and
tor and time resolve secretion events at the single-cell acetaminophen. Next the enzyme is immobilized, and
level. Although the quantal hypothesis for neurotrans- this is followed by an external membrane having the
mitter release is well accepted, caution must be taken in property of being highly permeable to oxygen while
interpretation of results. The actual time-dependent exhibiting only limited permeability to glucose. These
events are convoluted by diffusion from the vesicle to properties make it possible to prepare a sensor with a
the point of observation by the sensor. Catecholamine linear response to glucose over the required range (2–
exocytosis in PC12 cells was monitored by Chen, Luo 20 mM glucose) and response characteristics largely
and Ewing [129]. Using a carbon electrode modified independent of oxygen partial pressure. The short-term
with ruthenium oxide/cyanoruthenate developed by implants give useful results but may need to be cali-
Cox and Gray [130] Kennedy and co-workers [131] brated one to two times per day due to limited stability
were able to detect exocytosis of insulin from beta cells. while the long-term implants currently suffer from de-
The sensor is not very specific for insulin, which it does, struction of the implanted instrument package by the
however, oxidize rapidly, but because there are few if remarkably corrosive elements within the tissue. An
2640 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

important alternative approach might involve a chroni- Electrochemical detection has proven extremely valu-
cally implanted device that will be expected to function able in monitoring inorganic species that play an im-
reliably for 6 – 12 months [138]. portant role in biological systems. These include
Other sensors have also been developed for brain oxygen, hydrogen peroxide, superoxide, and nitric ox-
studies including glucose [139], glutamate [140] and ide. Oxygen is ideally suited to electrochemical detec-
lactate [141]. The sensors implanted in the brain must tion and such methodology was reported in the
be smaller, faster in response than those in the subcuta- physiology literature as early as 1942 [144]. There is
neous tissue, and free of interferences. For example, the currently significant interest in oxygen in studies of
glutamate sensor must detect 1 mM glutamate in the oxidative stress [145]. For such studies it is necessary to
presence of 300 – 400 mM ascorbate. The latter elec- know not simply concentrations, but fluxes in or out of
troactive interferent can be removed in two ways: with cells. One such approach is the ‘self-referencing’ elec-
a permselective membrane or by using ascorbate oxi- trode [146]. The concept is to make measurements of
dase, an enzyme that reduces oxygen to water and not
the current due to oxygen at points outside a cell in
peroxide, thus making it useful for ascorbate elimina-
culture that fall within the cell concentration gradient.
tion. Other tissues have also been examined as the
By physically moving the 1 – 3 mm diameter electrode at
monitoring of lactate release in cardiac tissue has been
about 0.3 Hz over a distance of 5 mm, two values of the
demonstrated [142]. If oxidases are employed, it is
oxygen concentration can be measured. From this in-
essential to establish that low oxygen levels do not
affect sensor response. Progress in sensor design and formation and an electrode calibration, the concentra-
methodology has emphasized the value of time-depen- tion gradient can be calculated. It is presumed that
dent data that an electrochemical sensor can provide. electrode motion does not disturb the gradient. The
Of perhaps even greater importance is the possibility of movement of the electrode has another advantage: it
obtaining simultaneously data on different analytes. By allows electrode noise to be eliminated through phase
monitoring glucose, lactate and oxygen simultaneously sensitive detection. Electrochemistry has provided some
in the rat brain, it was possible to demonstrate the of the greatest excitement of the decade due to the
elevation of lactate levels that might otherwise have possibility of measuring nitric oxide, voted the
been confused with ischemia (oxygen deficiency) [141]. Molecule of the Year by Science. This very reactive
Multiple analyte measurements have also been re- species cannot be easily monitored by alternative meth-
ported, some combining optical and electrochemical ods. Malinski [147] reported on an electrode modified
detection [143]. by a Ni(II) porphyrin. Although the oxidative electro-

Fig. 8. Schematic of implantable glucose sensor used for continuous monitoring in humans (see reference [137]).
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2641

chemistry is quite complicated, it appears that an an- purine (adenosine and inosine) release during cardiac
odic current proportional to NO can be obtained to ischemia. To accomplish this detection it was necessary
detection limits in the range of about 10 nM. The to employ a cascade of three enzymes: adenosine
biochemistry of this versatile molecule has been dis- deaminase, nucleotide phosphorylase, and xanthine ox-
cussed in the context of electrochemical measurements idase. By determining the charge generated from the
[148]. Using a 30 mm platinum disk electrode and oxidation of peroxide, the final product of the serial
double differential pulse amperometry, it was possible enzymatic reactions, femtomole levels of adenosine
to selectively measure hydrogen peroxide in the rat could be detected. The detection limit is about 120
striatum in the presence of dopamine, ascorbic acid, amol. The technology lends itself easily to rapid
and uric acid with a detection limit of 0.03 mM [149]. screening techniques as employed within the pharma-
A two-enzyme system (superoxide dismutase and cata- ceutical industry. Its virtue, the small volume of liquid,
lase) was used as part of a 50 mm disk microelectrode may produce concentrations of the analyte significantly
to measure hyperoxia (superoxide) in the striatum of a higher than physiological levels, thus leading to experi-
freely moving rat [150]. mental artifacts. Similarly, the effect of incubation of a
Thus far relatively little attention has been paid to single cell in elevated enzyme levels may also affect cell
acquisition of detailed understanding of how sensor behavior. Presumably many, if not most, of these
performance is affected by contact with tissue and how difficulties can be resolved by appropriate control ex-
the tissue is affected by the presence of the sensor. It is periments.
understood that the initial implantation of sensors can Although the electrochemistry of nucleic acids has
destroy capillaries and surrounding cells and this inva- been studied for many years [157], there has been
sion triggers the wound healing process the first step of increased interest in oligonucleotide detection for diag-
which involves protein adsorption on the sensor. With nostic purposes [158]. The technique of oscillographic
this in mind, membrane materials have been developed polarography has been revisited to permit the selective
with ‘water-like surfaces’, that tend to minimize ad- detection of pyrimidine-based nucleotide using a cop-
sorption [151]. Biocompatibility studies on biosensors per electrode and decoupling of the charging and
have been limited, and much needs to be learned about faradaic current in the frequency domain [159].
the interfacial reactions which occur.

9. Bioelectrochemistry in the new millennium


8. New electrochemical methodology as applied to
bioelectrochemistry The advances of the last 10 years have been signifi-
cant and have focused strongly on the understanding
New technology had a significant impact on bioelec- of the interface between the electrode and the
trochemistry during the 1990s. Bard’s group extended molecules reacting with it. We can certainly expect
scanning tunneling microscopy to the use of the probe nanotechnology to heavily influence developments. As
as an electrode (scanning electrochemical microscopy the interfacial chemistry is better understood it will be
(SECM)) to detect the activity of glucose oxidase on a possible to more effectively promote biocatalysis and
non-conducting surface [152]. The technique has subse- to further develop biochemical fuel cells [160]. Several
quently been extended to other applications such as groups have reported on the development of microelec-
the detection of guard cells on a plant leaf [153] or the trochemical enzyme transistors [161,162]. Amplification
detection of carcinoembryonic antigen (CEA) mi- is achieved because peroxide generated by an enzy-
crospotted on a plate and reacted with a CEA– matic reaction alters the conductivity of a poly(aniline)
horseradish peroxidase conjugate [154]. The method film [162]. Lower detection limits are possible than
can detect as few as  100 molecules in a 20 mm spot. with conventional sensors because the products of the
SECM has also been used to pattern active enzyme on reaction can accumulate. Miniaturized bioelectronic
a surface. This is accomplished by deactivating enzyme devices are also expected to proliferate. As more sen-
by oxidation of Cl− or Br− at the enzyme surface sors are developed, it will be increasingly possible to
[155]. simultaneously monitor several key analytes in real
With increased emphasis on measurements involving time. Finally, we can expect to see bioelectrochemical
single cells, it has proven useful to exploit microfabri- techniques used by a growing range of scientists.
cation and micromachining techniques to produce
wells containing electrodes that can accommodate sin-
gle cells. Using this format, Bratten and co-workers Acknowledgements
[156] developed a 200 mm diameter well with electrodes
in the bottom, with a depth of 20 mm and a total FAA acknowledges financial support from the UK
volume of  600 pL. The objective was to measure Research Councils (EPSRC, BBSRC), Wellcome Trust,
2642 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

Leverhulme Trust, Royal Society and St. John’s Col- [27] K.L. Egodage, B.S. DeSilva, G.S. Wilson, J. Am.
lege, Oxford. The partial support of the National Insti- Chem. Soc. 119 (1997) 5295.
tutes of Health, (US) to GSW, grants DK30718 and [28] P.L. Edmiston, J.E. Lee, S.-S. Cheng, S.S. Saavedra, J.
DK55297, is gratefully acknowledged. Am. Chem. Soc. 119 (1997) 560.
[29] T.M. Nahir, E.F. Bowden, J. Electroanal. Chem. 410
(1996) 9.
[30] R.A. Clark, E.F. Bowden, Langmuir 13 (1997) 559.
References
[31] J.F. Rusling, Acc. Chem. Res. 31 (1998) 363.
[32] J.F. Rusling, A.-E.F. Nassar, J. Am. Chem. Soc. 115
[1] K. Niki, T. Yagi, H. Inokuchi, K. Kimura, J. Am.
Chem. Soc. 101 (1979) 3335. (1993) 11891.
[2] M.J. Eddowes, H.A.O. Hill, J. Chem. Soc. Chem. [33] (a) R. Lin, M. Bayachou, J. Greaves, P.J. Farmer, J.
Commun. 771 (1977). Am. Chem. Soc. 119 (1997) 12689; (b) M. Bayachou,
[3] P. Yeh, T. Kuwana, Chem. Lett. 1145 (1977). R. Lin, W. Cho, P.J. Farmer, J. Am. Chem. Soc. 120
[4] F.A. Armstrong, Struct. Bond. 72 (1990) 137. (1998) 9888.
[5] P.M. Allen, H.A.O. Hill, N.J. Walton, J. Electroanal. [34] Z. Zhang, A.-E.F. Nassar, Z. Lu, J.B. Schenckman,
Interface Electrochem. 178 (1984) 69. J.F. Rusling, J. Chem. Soc. Faraday Trans. 93 (1997)
[6] Y. Taniguchi, K. Toyosawa, H. Yamaguchi, K. Ya- 1769.
sukouchu, J. Chem. Soc. Chem. Commun. 1032 [35] A.-E.F. Nassar, Z. Zhang, N. Hu, J.F. Rusling, T.F.
(1982). Kumosinski, J. Phys. Chem. B 101 (1997) 2224.
[7] H.A.O. Hill, D.J. Page, N.J. Walton, D.J. Whitford, J. [36] S. Boussaad, N.J. Tao, J. Am. Chem. Soc. 120 (1999)
Electroanal. Interface Electrochem. 187 (1987) 315. 4510.
[8] L.S. Wong, V.L. Vilker, W.T. Yap, Y. Reipa, Lang- [37] Y.M. Lvov, Z. Lu, J.B. Schenckman, X. Zu, J.F.
muir 11 (1995) 4818. Rusling, J. Am. Chem. Soc. 120 (1998) 4073.
[9] B.D. Lamp, D. Hobara, M.D. Porter, K. Niki, T.M. [38] J. Kong, Z. Lu, Y.M. Lvov, R.Z.B. Desamero, H.A.
Cotton, Langmuir 13 (1997) 736. Frank, J.F. Rusling, J. Am. Chem. Soc. 120 (1998)
[10] A.M. Bond, Anal. Proc. 29 (1992) 132.
7371.
[11] Z. Salamon, G. Tollin, Bioelectrochem. Bioenerg. 26
[39] Z. Salamon, J.T. Hazzard, G. Tollin, Proc. Natl.
(1991) 321.
Acad. Sci. USA 90 (1993) 6420.
[12] Y. Sato, F. Mizutani, J. Electroanal. Chem. 438 (1997)
[40] J.D. Burgess, M.C. Rhoten, F.M. Hawkridge, Lang-
99.
muir 14 (1998) 2467.
[13] K.J. Stevenson, M. Mitchell, H.S. White, J. Phys.
Chem. 102 (1998) 1235. [41] Z.Q. Feng, S. Imabayashi, T. Kakiuchi, K. Niki, J.
[14] M.J. Honeychurch, G.A. Rechnitz, J. Phys. Chem. 101 Chem. Soc. Faraday Trans. 93 (1997) 1367.
(1997) 7472. [42] A.K. Gaigalus, G. Niaura, J. Colloid Interface Sci.
[15] F.A. Armstrong, J. Hirst, H.A. Heering, Chem. Soc. 193 (1997) 60.
Rev. 26 (1997) 169. [43] C.E.D. Chidsey, Science 251 (1991) 919.
[16] S. Song, R.A. Clark, E.F. Bowden, M.J. Tarlov, J. [44] H.O. Finklea, L. Liu, M.S. Ravenscroft, S. Punturi, J.
Phys. Chem. 97 (1993) 6564. Phys. Chem. 100 (1996) 18852.
[17] F.A. Armstrong, P.A. Cox, H.A.O. Hill, V.J. Lowe, [45] J.F. Smalley, S.W. Feldberg, C.E.D. Chidsey, M.R.
B.N. Oliver, J. Electroanal. Chem. Interface Elec- Linford, M.D. Newton, Y-P. Liu, J. Phys. Chem. 99
trochem. 217 (1987) 331. (1995) 13141.
[18] J.L. Willit, E.F. Bowden, J. Phys. Chem. 94 (1991) [46] J. Hirst, J.L.C. Duff, G.N.L. Jameson, M.A. Kemper,
8241. B.K. Burgess, F.A. Armstrong, J. Am. Chem. Soc. 120
[19] A. Manjaqui, J. Haladjian, P. Bianco, Electrochim. (1998) 7085.
Acta 35 (1990) 177. [47] A.E. Kasmi, J.M. Wallace, E.F. Bowden, S.M.
[20] J.J. Davis, M.L.H. Green, H.A.O. Hill, Y.C. Leung, Binet, R.J. Linderman, J. Am. Chem. Soc. 120 (1998)
P.J. Sadler, A.V. Xavier, S.C. Tsang, Inorg. Chim. 225.
Acta 272 (1998) 261.
[48] A.-E.F. Nassar, W.S. Willis, J.F. Rusling, Anal. Chem.
[21] M. Collinson, E.F. Bowden, Anal. Chem. 64 (1992)
67 (1995) 2386.
1470.
[49] B.C. King, F.M. Hawkridge, B.M. Hoffman, J. Am.
[22] J. Li, G. Cheng, S. Dong, J. Electroanal. Interface
Chem. Soc. 114 (1992) 10603.
Electrochem. 416 (1996) 97.
[23] J.J. Davis, C.M. Halliwell, H.A.O. Hill, G.W. Canters, [50] M. Tominaga, T. Kumagai, S. Takita, I. Taniguchi,
M.C. van Amsterdam, M.P. Verbeet, New J. Chem. 22 Chem. Lett. 1771 (1993).
(1998) 1119. [51] B.R. Van Dyke, P. Saltman, F.A. Armstrong, J. Am.
[24] D. Hobara, K. Niki, C. Zhou, G. Chumanov, T.M. Chem. Soc. 118 (1996) 3490.
Cotton, Colloids. Surf. B 93 (1994) 241. [52] P.L. Hagedoorn, M.C.P.F. Driessen, M. Van den
[25] S. Lecomte, H. Wackerbarth, T. Soulimane, G. Buse, Bosch, I. Landa, W.R. Hagen, FEBS Lett. 440 (1998)
P. Hildebrandt, J. Am. Chem. Soc. 120 (1998) 7381. 311.
[26] V. Reipa, A.K. Gaigalas, V.L. Vilker, Langmuir 13 [53] J.-B. Park, C. Fan, B.M. Hoffman, M.W.W. Adams,
(1997) 3508. J. Biol. Chem. 266 (1991) 19351.
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2643

[54] P.S. Breraton, M.F.J.M. Verhagen, Z.H. Zhou, [80] K.L. Turner, M.K. Doherty, H.A. Heering, F.A. Arm-
M.W.W. Adams, Biochemistry 37 (1998) 7351. strong, G.A. Reid, S.K. Chapman, Biochemistry 38
[55] M.T. Cruanes, K.K. Rodgers, S.G. Sligar, J. Am. (1999) 3302.
Chem. Soc. 114 (1992) 9660. [81] A. Sucheta, B.A.C. Ackrell, B. Cochran, F.A. Arm-
[56] L.D. Gillès de Pélichy, E.T. Smith, Biochemistry 38 strong, Nature 356 (1992) 361.
(1999) 7874. [82] H.A. Heering, J.H. Weiner, F.A. Armstrong, J. Am.
[57] S.E. Iismaa, A.E. Vázquez, G.M. Jensen, P.J. Chem. Soc. 119 (1997) 11628.
Stephens, J.N. Butt, F.A. Armstrong, B.K. Burgess, J. [83] J. Hirst, A. Sucheta, B.A.C. Ackrell, F.A. Armstrong,
Biol. Chem. 266 (1991) 21563. J. Am. Chem. Soc. 118 (1996) 5031.
[58] B. Shen, D.R. Jollie, C.D. Stout, T.C. Diller, F.A. [84] B.A.C. Ackrell, F.A. Armstrong, B. Cochran, A.
Armstrong, C.M. Gorst, G.N. La Mar, P.J. Stephens, Sucheta, T. Yu, FEBS Lett. 326 (1993) 92.
B.K. Burgess, J. Biol. Chem. 269 (1994) 8564. [85] H.R. Pershad, J. Hirst, B. Cochran, B.A.C Ackrell,
[59] H.S. Gao-Sheridan, H.R. Pershad, F.A. Armstrong, F.A. Armstrong, Biochim. Biophys. Acta (1999) in
B.K. Burgess, J. Biol. Chem. 273 (1998) 5514. press.
[60] P. Kyritsis, O. Hatzfeld, T.A. Link, J.-M. Moulis, J. [86] J. Hirst, B.A.C. Ackrell, F.A. Armstrong, J. Am.
Biol. Chem. 273 (1998) 15404. Chem. Soc. 119 (1997) 7434.
[61] J.L.C. Duff, J.L.J. Breton, J.N. Butt, F.A. Armstrong, [87] P. Bianco, J. Haladjian, J. Electrochem. Soc. 139
A.J. Thomson, J. Am. Chem. Soc. 118 (1996) 8593. (1992) 2428.
[62] S.E.J. Fawcett, D. Davis, J.L. Breton, A.J. Thomson, [88] J.N. Butt, M. Filipiak, W.R. Hagen, Eur. J. Biochem.
F.A. Armstrong, Biochem. J. 335 (1998) 357. 245 (1997) 116.
[63] J.N. Butt, F.A. Armstrong, J. Breton, S.J. George, [89] H.R. Pershad, J.L.C. Duff, H.A. Heering, E.C. Duin,
S.P.J. Albracht, F.A. Armstrong, Biochemistry 38
A.J. Thomson, E.C. Hatchikian, J. Am. Chem. Soc.
(1999) in press.
113 (1991) 6663.
[90] E. Csöregi, L. Gorton, G. Marko-Varga, Anal. Chim.
[64] J.N. Butt, A. Sucheta, F.A. Armstrong, J. Breton, A.J.
Acta 273 (1993) 59.
Thomson, E.C. Hatchikian, J. Am. Chem. Soc. 113
[91] J. Zhao, R.W. Henkens, J. Stonehuerner, J.P. O’Daly,
(1991) 8948.
A.L. Crumbliss, J. Electroanal. Chem. 327 (1992) 109.
[65] J.N. Butt, S.E.J. Fawcett, J. Breton, A.J. Thomson,
[92] Y. Zhang, G.S. Wilson, J. Electroanal. Chem. 345
F.A. Armstrong, J. Am. Chem. Soc. 119 (1997) 9729.
(1993) 253.
[66] J.N. Butt, A. Sucheta, F.A. Armstrong, J. Breton, A.J.
[93] S.B. Hall, E.A. Khudaish, A.L. Hart, Electrochim.
Thomson, E.C. Hatchikian, J. Am. Chem. Soc. 115
Acta 43 (1997) 579; S.B. Hall, E.A. Khudaish, A.L.
(1993) 1413.
Hart, Electrochim. Acta 43 (1998) 2015; S.B. Hall,
[67] J. Haladjian, R. Pilard, P. Bianco, P.-A. Serre,
E.A. Khudaish, A.L. Hart, Electrochim. Acta 44
Bioelectrochem. Bioenerg. 9 (1982) 91.
(1999) 2455.
[68] P.D. Barker, A.G. Mauk, J. Am. Chem. Soc. 114
[94] G. Kenausis, C. Taylor, I. Katakis, A. Heller, J.
(1992) 3619.
Chem. Soc. Faraday Trans. 92 (1996) 4131.
[69] X. Yuan, S. Sun, F.M. Hawkridge, J.F. Chlebowski, I. [95] I. Willner, V. Heleg-Shabtai, R. Blonder, A.F. Büch-
Taniguchi, J. Am. Chem. Soc. 112 (1990) 5380. mann, A. Heller, J. Am. Chem. Soc. 118 (1996) 10321.
[70] B.A. Feinberg, X. Liu, M.D. Ryan, A. Schejter, [96] R. Rajagopalan, A. Aoki, A. Heller, J. Phys. Chem.
C. Zhang, E. Margoliash, Biochemistry 37 (1998) 100 (1996) 3719.
13091. [97] T.J. Ohara, R. Rajagopalan, A. Heller, Anal. Chem.
[71] F.I. Rosell, J.C. Ferrer, A.G. Mauk, J. Am. Chem. 66 (1994) 2451.
Soc. 120 (1998) 11234. [98] H. Bu, A.M. English, S.R. Mikkelsen, Anal. Chem. 68
[72] H. Lopes, G.W. Pettigrew, I. Moura, J.J.G. Moura, J. (1996) 3951.
Biol. Inorg. Chem. 3 (1998) 632. [99] J. Kulys, T. Buch-Rasmussen, K. Bechgaard, V. Razu-
[73] J. Hirst, G.N.L. Jameson, J.W.A. Allen, F.A. Arm- mas, J. Kazlauskaite, J. Marcinkeviciene, J.B. Chris-
strong, J. Am. Chem. Soc. 120 (1998) 11994; J. Hirst, tensen, H.E. Hansen, J. Mol. Catal. 91 (1994) 407.
F.A. Armstrong, Anal. Chem. 70 (1998) 5062. [100] P.N. Bartlett, S. Booth, D.J. Caruana, J.D. Kilburn,
[74] J.P. McEvoy, F.A. Armstrong, J. Chem. Soc. Chem. C. Santamaria, Anal. Chem. 69 (1997) 734.
Commun. (1999) in press. [101] J. Wang, F. Lu, J. Am. Chem. Soc. 120 (1998) 1048.
[75] R.J. Cherry, A.J. Bjornsen, D.C. Zapien, Langmuir 14 [102] T.P. Henning, D.D. Cunningham, in: G. Ramsay
(1998) (1971). (Ed.), Commercial Biosensors: Applications to Clinical,
[76] H.A. Heering, J. Hirst, F.A. Armstrong, J. Phys. Bioprocess, and Environmental Samples, ch. 1, Wiley,
Chem. B 102 (1998) 6889. New York, 1998.
[77] A. Onuoha, X. Hu, J.F. Rusling, J. Am. Chem. Soc. [103] P.N. Bartlett, J.M. Cooper, J. Electroanal. Chem. 362
119 (1997) 3979. (1993) 1.
[78] M.S. Mondal, D.B. Goodin, F.A. Armstrong, J. Am. [104] W. Schuhmann, in: T. Cass, F.S. Ligler (Eds.), Immo-
Chem. Soc. 120 (1998) 6270. bilized Biomolecules in Analysis: A Practical Ap-
[79] H.A. Heering, M.S. Mondal, F.A. Armstrong, Anal. proach, ch. 11, Oxford University Press, New York,
Chem. 71 (1999) 174. 1998.
2644 F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645

[105] N.K. Cenas, J. Kulys, J. Bioelectrochem. Bioenerg. 8 [129] T.C. Chen, G. Luo, A.G. Ewing, Anal. Chem. 66 (1994)
(1981) 103. 3031.
[106] B.S. Hill, C.A. Scolari, G.S. Wilson, J. Electroanal. [130] J.A. Cox, T.J. Gray, Anal. Chem. 61 (1989) 2462.
Chem. 252 (1988) 125. [131] L. Huang, H. Shen, M.A. Atkinson, R.T. Kennedy,
[107] G.F. Khan, M. Ohwa, W. Wernet, Anal. Chem. 68 Proc. Natl. Acad. Sci. USA 92 (1995) 9608.
(1996) 2939. [132] Q. Xin, R.M. Wightman, Brain Res. 776 (1997) 126.
[108] C. Bourdillon, C. Demaille, J. Moiroux, J-M. Savéant, J. [133] M.A. Bunin, R.M. Wightman, Methods Enzymol. 296
Am. Chem. Soc. 115 (1993) 2. (1998) 689.
[109] A. Badia, R. Carlini, A. Fernandez, F. Battaglini, S.R. [134] R.D. O’Neill, J.P. Lowry, M. Mas, Crit. Rev. Neuro-
Mikkelsen, A.M. English, J. Am. Chem. Soc. 115 (1993) biol. 12 (1998) 69.
7053. [135] L.C. Clark Jr., C. Lyons, Ann. New York Acad. Sci. 102
[110] J. Hodak, R. Etchenique, E.J. Calvo, K. Singhal, P.N. (1962) 29; S.J. Updike, G.P. Hicks, Nature 214 (1967)
Bartlett, Langmuir 13 (1997) 2708.
986.
[111] C. Danilowicz, R. Etchenique, E.J. Calvo, L. Diaz,
[136] M. Shichiri, Y. Yamasaki, N. Hakui, H. Abe, Lancet 2
Anal. Chem. 68 (1996) 4186.
(1982) 1129.
[112] J. Moiroux, P.J. Elving, Anal. Chem. 50 (1978) 1056.
[137] V. Poitout, D. Moatti-Sirat, G. Reach, Y. Zhang, G.S.
[113] B. Persson, L. Gorton, J. Electroanal. Chem. 292 (1990)
Wilson, F. Lemonnier, J-C. Klein, Diabetologia 36
115.
(1993) 658.
[114] H. Jaegfeldt, T. Kuwana, G. Johansson, J. Am. Chem.
Soc. 105 (1983) 1805; C. Ueda, D.C.S. Tse, T. Kuwana, [138] B.J. Gilligan, R.K. Rhodes, M.C. Shults, S.J. Updike,
Anal. Chem. 54 (1982) 850. Diabetes Care, 17 (1994) 882; J.C. Armour, J.Y. Lu-
[115] Q. Wu, M. Maskus, F. Pariente, F. Tobalina, V.M. cisano, B.D. McKean, D.A. Gough, Diabetes 39 (1990)
Fernández, E. Lorenzo, H.D. Abruña, Anal. Chem. 68 1519.
(1996) 3688. [139] Y. Hu, G.S. Wilson, J. Neurochem. 68 (1997) 1745.
[116] F. Pariente, F. Tobalina, M. Darder, E. Lorenzo, H.D. [140] Y. Hu, K.M. Mitchell, F.M. Albahadily, E.K.
Abruña, Anal. Chem. 68 (1996); E. Lorenzo, F. Pari- Michaelis, G.S. Wilson, Brain Res. 659 (1994) 117; M.R.
ente, L. Hernández, F. Tobalina, M. Darder, Q. Wu, M. Ryan, J.P. Lowry, R.D. O’Neill, Analyst 122 (1997)
Maskus, H.D. Abruña, Biosensors Bioelectron. 13 1419.
(1998) 319. [141] Y. Hu, G.S. Wilson, J. Neurochem. 69 (1997) 1484.
[117] B. Persson, H.L. Lan, L. Gorton, Y. Okamoto, P.D. [142] S.A.M. Marzouk, V.V. Cosofret, R.P. Buck, H. Yang,
Hale, L.I. Boguslavsky, T.A. Skotheim, Biosens. W.E. Cascio, S.S.M. Hassan, Anal. Chem. 69 (1997)
Bioelectron. 8 (1993) 81; B. Persson, H.S. Lee, L. Gor- 2646.
ton, T. Skotheim, P. Bartlett, Electroanalysis 7 (1995) [143] Q. Xin, R.M. Wightman, Anal. Chem. 70 (1998) 1677.
935. [144] P.W. Davies, F. Brink, Rev. Sci. Instr. 13 (1942) 524.
[118] B.W. Carlson, L.L. Miller, J. Am. Chem. Soc. 107 [145] S. Arbault, P. Pantano, J.A. Jankowski, M. Vuillaume,
(1985) 479. C. Amatore, Anal. Chem. 67 (1995) 3382.
[119] A. Anne, J. Moiroux, J-M. Savéant, J. Am. Chem. Soc. [146] S.C. Land, D.M. Porterfield, R.H. Sanger, P.J.S. Smith,
115 (1993) 10224. J. Exp. Biol. 202 (1999) 211.
[120] A. Bardea, E. Katz, A.F. Bückmann, I. Willner, J. Am. [147] Z. Taha, F. Kiechle, T. Malinski, Biochem. Biophys.
Chem. Soc. 119 (1997) 9114. Res. Commun. 188 (1992) 734.
[121] Y. Ogino, K. Takagi, K. Kano, T. Ikeda, J. Electroanal. [148] T. Malinski, S. Mesaros, P. Tomboulian, Methods Enzy-
Chem. 396 (1995) 517. mol. 268 (1996) 58.
[122] (a) I. Willner, E. Katz, B. Willner, Electroanalysis 9
[149] H. Yokoyama, N. Tsuchihashi, N. Kasai, T. Matsue, I.
(1997) 965; (b) W. Jin, F. Bier, U. Wollenberger, F.
Uchida, N. Mori, H. Ohya-Nishiguchi, H. Kamada,
Scheller, Biosens. Bioelectron. 10 (1995) 823.
Biosens. Bioelectron. 12 (1997) 1037.
[123] P. Pantano, T.H. Morton, W.G. Kuhr, J. Am. Chem.
[150] H. Yokoyama, N. Kasai, T. Matsue, N. Tsuchihashi, I.
Soc. 113 (1832) 1991.
Ushida, T. Ogata, N. Mori, H. Ohya-Nishiguchi, H.
[124] N. Anicet, C. Bourdillon, J. Moiroux, J-M. Savéant, J.
Kamada, J. Brain Sci. 23 (1997) 181.
Phys. Chem. B 102 (1998) 9844.
[125] C. Bourdillon, C. Demaille, J. Moiroux, J-M. Savéant, [151] C.A.P. Quinn, R.E. Connor, A. Heller, Biomaterials, 18
Acc. Chem. Res. 29 (1996) 529. (1997) 1665; K. Ishihara, H. Oshida, Y. Endo, T. Ueda,
[126] N. Anicet, A. Anne, J. Moiroux, J-M. Savéant, J. Am. A.Watanabe, N. Nakabayashi, J. Biomed. Mater. Res.
Chem. Soc. 120 (1998) 7115. 26 (1992) 1543.
[127] N. Dontha, W.B. Nowall, W.G. Kuhr, Anal. Chem. 69 [152] D.T. Pierce, P.R. Unwin, A.J. Bard, Anal. Chem. 64
(1997) 2619; W.B. Nowall, D.O. Wipf, W.G. Kuhr, (1992) 1795.
Anal. Chem. 70 (1998) 2601; W.B. Nowall, N. Dontha, [153] M. Tsionsky, Z.G. Cardon, A.J. Bard, R.B. Jackson,
W.G. Kuhr, Biosens. Bioelectron. 13 (1998) 1237. Plant Physiol. 113 (1997) 895.
[128] R.M. Wightman, J.A. Jankowski, R.T. Kennedy, K.T. [154] H. Shiku, T. Matsue, I. Uchida, Anal. Chem. 68 (1996)
Kawagoe, T.J. Schroeder, D.J. Leszczyzyn, J.A. Near, 1276.
E.J. Dilberto Jr., O.H. Viveros, Prod. Natl. Acad. Sci. [155] H. Shiku, T. Takada, H. Yamada, T. Matsue, I. Uchida,
USA 88 (1991) 10754. . Anal. Chem. 67 (1995) 312.
F.A. Armstrong, G.S. Wilson / Electrochimica Acta 45 (2000) 2623–2645 2645

[156] C.D.T. Bratten, P.H. Cobbold, J.M. Cooper, Anal. [159] P. Singhal, W.G. Kuhr, Anal. Chem. 69 (1997) 3552.
Chem. 70 (1998) 1164. [160] I. Willner, E. Katz, F. Patolsky, A.F. Bückmann, J.
[157] E. Paleček, Electroanalysis 8 (1996) 7. Chem. Soc. Perkin Trans. 2 (1998) 1817.
[158] S.R. Mikkelsen, Electroanalysis, 8 (1996) 15; J. Wang, [161] T. Matsue, M. Nishizawa, T. Sawaguchi, I. Uchida, J.
G. Rivas, X. Cai, E. Paleček, P. Nielsen, D. Chem. Soc. Chem. Commun. (1991) 1029.
Grant, M. Ozsoz, M. Flair, Anal. Chim. Acta 347 [162] P.N. Bartlett, P.R. Birkin, J.H. Wang, F. Palmisano,
(1997) 1. G. De Benedetto, Anal. Chem. 70 (1998) 3685.

You might also like