Está en la página 1de 10

Electrochimica Acta 44 (1999) 45734582

Electrochemical oxidation of hydrogen peroxide at


platinum electrodes. Part IV: phosphate buer dependence
Simon B. Hall a,*, Emad A. Khudaish a, Alan L. Hart b
a
Institute of Fundamental Sciences, Massey University, Private Bag 11222, Palmerston North, New Zealand
b
Sensor Group, AgResearch, Grasslands Research Centre, Palmerston North, New Zealand
Received 25 January 1999; received in revised form 26 April 1999

Abstract

The electrochemical oxidation of H2O2 at platinum rotating disc electrodes and microelectrodes was studied as a
function of phosphate buer concentration in the range 0100 mM and pH from pH 4 to pH 10. The results were
interpreted in terms of development of a surface binding site for H2O2 from a precursor site through interaction
with H2 PO4 from the electrolyte. In the absence of phosphate an alternative binding site mechanism was evident.
The precursor site was shown to be inhibited by protons at low pH producing an inactive site. # 1999 Elsevier
Science Ltd. All rights reserved.

Keywords: Platinum; Hydrogen peroxide; Oxidation; Rotating disc electrode; Mechanism

1. Introduction Wilson [11] considered a similar mechanism in which


H2O2 acted directly to reduce an oxidized surface site
A mechanism for the oxidation of hydrogen per- without rst binding to it.
oxide at platinum electrodes has recently been reported We have developed optimizing techniques to test the
by us [1,2] to account for the saturation kinetics validity of our model to the response as a function of
observed in this reaction. Oxidation of H2O2 at plati- electrode rotation rate and [H2O2] [1], potential [2] and
num is of interest since it forms the amperometric re- temperature [12]. An important feature of this mechan-
sponse in many sensor devices incorporating enzyme ism and that of sputtered Pd/Au on carbon [9] is the
electrodes [38]. The mechanism is a modication of binding site on the surface of the platinum and palla-
that reported by Johnston et al. [9] for oxidation of dium respectively. Support for the existence of an oxi-
H2O2 at sputtered Pd/Au on carbon. The main dized binding site not only arises from the t of the
sequence in our mechanism involves reversible binding mechanism to our experimental data, but also from
of H2O2 to a surface binding site, a redox process earlier work by others showing that oxidation of H2O2
within the surface complex to give O2 and a reduced is favored on oxidized Pt surfaces [10,13,14]. This was
binding site, and oxidation of the reduced site to com- based upon linking the potential at which H2O2 rst
plete the catalytic cycle and form the amperometric re- oxidizes to that for platinum oxide lm formation [15].
sponse. Lingane and Lingane [10] and Hickling and In part II of our study [2] we examined the potential
dependence of the reaction at 208C and proposed that
the number of binding sites per unit area increases
* Corresponding author. Fax: +64-63505682. markedly with potential above +244 mV vs. Ag/AgCl,
E-mail address: s.b.hall@massey.ac.nz (S.B. Hall) increasing to a maximum ca. +650 mV. This was also

0013-4686/99/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 3 - 4 6 8 6 ( 9 9 ) 0 0 1 8 3 - 8
4574 S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582

found to be the case for all experiments in the range


5358C [12].
The buer employed in our previous studies has
been conned to 0.1 M phosphate buer at pH 7.3.
This was adopted as standard since these conditions
are widely used in enzyme-based sensor devices [16
22]. In this study we extended our investigations to a
wider range of phosphate concentrations and pH. This
has led us to modify and augment our model to
explain events at the surface of platinum electrodes at
other than the commonly used physiological con-
ditions. In particular, our view of the binding site has
changed.

2. Experimental

All chemicals and electrochemical equipment were as


detailed in our previous studies [1,2,12]. Two series of
phosphate buer solutions were prepared. The rst,
with total phosphate concentration in the range 1100
mM with pH 7.3 maintained throughout, whilst the
second series maintained total phosphate concentration
at 0.1 M but with pH varying over the range pH 410.
All studies were performed at 20 2 0.18C.
Steady-state measurements at platinum rotating disc
electrodes (0.1195 cm2, Bioanalytical Systems, IN,
USA) were made using the regimen of rotation rates
and anodic potentials established in previous studies Fig. 1. Steady-state response at a platinum rotating disc elec-
[1,2]. Full iR compensation was employed for all rotat- trode at a xed rotation rate (o=4000 rpm) and potential
ing disc electrode experiments using the method of He (E=+584 mV vs. Ag/AgCl) as a function of bulk hydrogen
and Faulkner [23] incorporated in the Bioanalytical peroxide concentration for a range of [PO3
4 ]tot at pH 7.3: (Q)

100B/W potentiostat. 2 mM, (R) 5 mM, () 10 mM, (q) 25 mM, (.) 50 mM, (+)
75 mM and (r) 100 mM.
Platinum microelectrodes (10 mm diameter,
Bioanalytical Systems, IN, USA) were employed to
examine low phosphate concentration buers so as to 3. Results and discussion
avoid iR artifacts. Prior to making steady-state
measurements these electrodes were polished with 1 A selection of steady-state responses (at a single ro-
mm diamond slurry and 0.05 mm alumina (both of tation rate and potential) of a Pt rotating disc elec-
Bioanalytical Systems, IN, USA) and then conditioned trode to H2O2 for a range of phosphate buer
electrochemically in the same manner as that for rotat- concentrations, [PO34 ]tot, is shown in Fig. 1. The ro-
ing disc electrodes detailed in our previous work [1,2]. tation rate, potential and temperature dependence have
Flow injection analysis (FIA) was performed on a been described in our previous papers [1,2,12]. Two
Flow Solution 3000 device (Alpkem, OR, USA) main features are identied in Fig. 1. First, the rate of
equipped with a single peristaltic pump and a 20 ml H2O2 oxidation (at any given potential, rotation rate
sample loop. The carrier was 0.1 M phosphate buer and [H2O2]) increases with buer concentration. This
at pH 7.3 at a ow rate of 2502 10 ml/min. The sol- cannot be attributed to changes in iR artifacts since
ution stream from the FIA system was connected to a full iR compensation was employed throughout these
cross-ow thin-layer ow cell (Bioanalytical Systems, experiments. Secondly, the saturation described in our
IN, USA) with an internal volume of 2.9 ml tted with previous papers [1,2] is approached at lower [H2O2] for
a 0.110 cm2 platinum working electrode. The potential lower phosphate buer concentrations. These features
of the working electrode was controlled by a may be interpreted in terms of our mechanism for
Bioanalytical Systems CV-27 potentiostat, the output H2O2 oxidation [1,2]. The oxidation process is inter-
of which was connected to an analogue to digital con- preted to take place via formation of a complex
verter housed in the Flow Solution 3000. formed between H2O2 and a PtII species acting as a
S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582 4575

is the fractional surface coverage of the complex


Pt(OH)2H2O2.
Both new features exhibited in Fig. 1 may be
accounted for if N were to increase with [PO34 ]tot. In
Part III of our study [12] we proposed the existence of
some precursor site, PtPS, to account for the potential
dependence of N (increasing with more anodic poten-
tial) with an incompletely dened equilibrium given by

KBS
PtPS
)
*
PtOH2 : 5

We now wish to consider whether phosphate could be


involved as a reactant in this equilibrium. It is also
possible that the rate constant k2 may increase with
[PO3
4 ]tot. However, this could not fully account for the
buer concentration dependence since if N were to
remain constant while k2 increases, then increasing
[PO3
4 ]tot would result in saturation occurring to the
same extent with [H2O2] in each curve in Fig.1
(although the current at saturation would be greater at
higher [PO34 ]tot).
To assist with interpretation of the buer concen-
tration dependence a selection of the data presented in
Fig. 1 is shown as a function of [PO34 ]tot in Fig. 2. In
addition to displaying the previously noted trend that
rate increases with [PO34 ]tot, a number of signicant
features are evident in this view of the data. First, the
relative increase in response with [PO3 4 ]tot is greater
Fig. 2. A selection of data from Fig. 1 plotted as a function for higher [H2O2]. For example, for 5 mM H2O2 the
of [PO3 response increases by a factor of 1.9 from 25 mM to
4 ]tot at pH 7.3 for xed [H2O2]bulk: (^) 5 mM, (Q) 12
mM, (R) 19 mM, () 26 mM, (q) 33 mM and (.) 40 mM. 100 mM phosphate, whilst for 40 mM H2O2 the factor
is 3.5 increase over the same range of [PO3 4 ]tot. We
ascribe this to a balance of equilibria between for-
binding site at the electrode surface,
mation of the binding site from some precursor and
K1 the binding of H2O2 to the active site according to
H2 O2 PtOH 2
)
*
PtOH2  H2 O2 : 1 reaction (1). Secondly, the increase in rate with
[PO3
4 ]tot does not follow the expected saturation kin-
This complex undergoes internal charge transfer etics if reversible formation of a binding site from a
precursor site were to occur according to
k2
PtOH2  H2 O2
4
Pt 2H2 O O2 2
KBS
PtPS Hx PO4x3 )

*
PtBS , 6
and the surface Pt site oxidizes to reform the binding
site, provide the amperometric signal and complete the where Hx PO4x3 is the particular phosphate species
catalytic cycle, involved in the interaction and we now propose to
k3
employ PtBS as the label for the binding site, instead of
PtOH2 2H 2e :
Pt 2H2 O
4 3 Pt(OH)2, in this and all subsequent expressions.
The large increase in rate over the [PO3 4 ]tot range
At potentials more anodic than +270 mV vs. Ag/AgCl 75100 mM may be suggestive of further interaction of
we have demonstrated that reaction (2) is the rate lim- phosphate with the binding site,
iting process [2] so that the overall rate is given by
KBS2
PtBS Hx POx3 *
4 )PtBS2 , 7
j k2 NyPtOH2 H2 O2 , 4

where N is the total number of surface sites (occupied, where it is conceivable that the oxidation state for Pt
unoccupied and reduced) per unit area and yPtOH2 H2 O2 in PtBS2 is higher than that for PtBS.
4576 S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582

A quantitative analysis is not readily obtainable in


the present study. Where quantitative analysis was
undertaken previously [1,2,12] it was necessary to
include consideration of a number of inhibiting reac-
tions. We proposed that the reaction product O2 inter-
feres with the oxidation of H2O2 by competitive
inhibition of the binding site according to

K4
PtBS O2
)
*
PtBS  O2 , 8

where PtBSO2 represents the species previously labeled


Pt(OH)2O2. The proton also inhibits through uncom-
petitive inhibition of the occupied site (now written as
PtBSH2O2),

K5
PtBS H2 O2 H
)
*
PtBS  H2 O2  H : 9

The overall rate equation was shown to be

j k2 NK1 H2 O2 =
10
f1 K4 O2 K1 H2 O2 1 K5 H k2 =k3 g:

If the simplest case for phosphate dependence accord-


ing to reaction (6) were considered (ignoring the possi-
bility of higher phosphate species) then the equilibrium
constant for binding site formation would be given by

yPtBS
KBS , 11 Fig. 3. Steady-state response at a platinum microelectrode at
Hx POx3
4 yPtPS 5 mM H2O2 as a function of potential for a range of [PO3
4 ]tot
at pH 7.3: (^) 0 mM, (w) 1 mM, (Q) 2 mM, (R) 5 mM, ()
where yPtBS and yPtPS are the fractional surface cov- 10 mM, (q) 25 mM, (.) 50 mM, (+) 75 mM and (r) 100
erages for the binding and precursor sites respectively. mM.
Provided the solution pH is held constant, the con-
centration of the particular phosphate species involved
in binding site formation could be assumed to be We strongly believe that quantitative analysis of the
directly proportional to the total phosphate concen- experimental data to aord rate and equilibrium con-
tration. Furthermore, the bulk concentration, stants in the manner taken previously [1,2] is not
[PO3
4 ]tot,bulk, is appropriate since this remains con- appropriate in this situation. In our previous work
stant. The equilibrium may now be written
[1,2,12] we used SIMPLEX optimization techniques
yPtBS for the iterative solution of the rate polynomials that
0
KBS 3
, 12 arise from Eq. (10) once diusion eects are accounted
PO4 tot,bulk yPtPS
for. A similar polynomial expression could be ident-
where K'BS would be expected to exhibit pronounced ied for Eq. (13) that incorporates the involvement of
pH dependence. phosphate. It is our view that meaningful results
Applying steady-state conditions to all surface would not be obtained from such an analysis. The new
1
species and solving for rate, yields term (KBS0
[PO3
4 ]tot,bulk) is likely to dominate the con-
tribution from the terms K4[O2] and K5[H+] leading to

!
 1
j k2 NK1 H2 O2 1 K4 O2 0 K1 H2 O2 1 K5 H k2 =k3 : 13
K BS PO3
4 tot,bulk
S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582 4577

Fig. 4. A selection of data from Fig. 3 plotted as a function Fig. 5. Steady-state response at a platinum microelectrode as
of [PO3 a function of bulk hydrogen peroxide concentration in water
4 ]tot at pH 7.3 for xed potentials vs. Ag/AgCl: (Q)
264 mV, () 360 mV, (r) 456 mV, (w) 552 mV, (q) 648 mV (with no phosphate buer or supporting electrolyte) for a
and (.) 744 mV. range of potentials vs. Ag/AgCl: (Q) 264 mV, () 360 mV,
(r) 456 mV, (w) 552 mV, (q) 648 mV and (.) 744 mV.

low precision results for K4 and K5. This situation


would not be ameliorated by varying rotation rate issue since quantitative analysis was deemed to be
since there are no mass-transport considerations for futile and furthermore, the decrease in O2 and H+ in-
phosphate. Another major concern is that the possi- hibition aorded by hemispherical diusion is advan-
bility of higher phosphate species forming as described tageous in clarifying the dependence on buer
in reaction (7) would prevent reliable optimization. concentration.
Linked to our hesitation to undertake optimization The response of a Pt microelectrode for xed [H2O2]
is our nal observation that we make for the data pre- and varying [PO3 4 ]tot as a function of potential is
sented in Fig. 2. If simple formation of the binding site shown in Fig. 3. Two distinct patterns of response are
were to occur according to reaction (6) and the species evident, depending on the total phosphate concen-
PtBS were requisite for any oxidation of H2O2, then a tration. For all experiments with [PO34 ]tot>5 mM a
response would not be expected in the absence of maximum rate develops (or appears to be approached)
phosphate. Extrapolation of the data presented in with increasing anodic potential. The current at the
Fig. 2 does not satisfactorily indicate whether this is maximum appears to increase with increasing [PO3 4 ]tot
the case. We were unable to conduct rotating disc elec- and the open circuit potential is common at E = 244
trode experiments for [PO3 4 ]tot < 2 mM whilst main- mV vs. Ag/AgCl, consistent with that reported by us
taining full iR compensation due to increased previously for rotating disc experiments [2]. The
electrolyte resistance. It is for this reason that this increase in maximum response with [PO3 4 ]tot is consist-
[PO34 ]tot region was explored using microelectrodes ent with our observation in the present work that
where iR artifacts do not prevail. Furthermore, the phosphate is incorporated in the binding site. The data
loss of rotation rate dependence information is not an for experiments with [PO34 ]tot < 5 mM exhibit a mark-
4578 S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582

edly dierent prole. Here, exponential increases in re-


sponse with potential are observed, with higher re-
sponse for lower [PO3 4 ]tot. The open circuit potential is
shifted anodically to E = 282 mV vs. Ag/AgCl in the
absence of phosphate.
The exponential potential dependence is such that at
the highest potential explored the rate of oxidation is
greatest in the absence of phosphate. In general a mini-
mum response at each potential is found in the range 1
mM < [PO3 4 ]tot < 2 mM. These observations indicate
that phosphate is not absolutely necessary for the oxi-
dation of H2O2 to proceed. The minima are more
readily seen by presenting the data (in Fig. 3) as a
function of [PO34 ]tot. These relationships are shown in
Fig. 4 for a selection of anodic potentials. At low po-
tentials (E < 458 mV vs. Ag/AgCl) minima are not
observed at low [PO3 4 ]tot, whilst at higher anodic po-
tentials progressively more pronounced minima occur
in the range 1 mM < [PO3 4 ]tot < 2 mM. We propose
that this is due to two distinct mechanisms for oxi-
dation of H2O2 at platinum. At physiological concen-
trations of phosphate the mechanism described by us
previously [1,2] predominates. As the concentration of
phosphate changes to lower levels an alternative mech-
anism begins to operate, in particular at high anodic
potentials (where the reaction proceeds more rapidly
than in the presence of phosphate). Furthermore, the
phosphate-free mechanism is impaired by low concen-
trations of phosphate and it is only when sucient
binding sites involving phosphate are formed that Fig. 6. Steady-state response at a platinum microelectrode at
higher oxidation rates occur. In contrast, at low poten- xed potential (E=+584 mV vs. Ag/AgCl) with 100 mM
tials the phosphate-mediated mechanism is more rapid citrate buer at pH 7.3 maintained throughout for a range of
at all concentrations of phosphate. [PO3
4 ]tot: (^) 0 mM, (w) 1 mM, (Q) 2 mM, (R) 5 mM, ()
10 mM, (q) 25 mM, (.) 50 mM and (+) 75 mM.
The nature of the phosphate-free mechanism on Pt
microelectrodes was examined in the H2O2 range 025
mM and for potentials 232 to 744 mV vs. Ag/AgCl dence obtainable from Fig. 5 to support any
using staircase potentiometry and the data are shown hypothesis for accompanying uncompetitive inhibition
in Fig. 5. In a similar manner to that reported by us by H+ since mass-transport variation by rotation rate
for phosphate buer systems (here and in earlier work) is not available.
saturation kinetics are observed at all potentials in There is the possibility that the phosphate depen-
Fig. 5 with greater response at more anodic potentials. dence of both rotating disc and microelectrodes arise,
It is important to note that no response is observed at in part, from diminished buering capacity of the elec-
all potentials in the absence of H2O2 indicating that trolyte adjacent to the electrode surface as [PO3 4 ]tot
the potential dependence exhibited in the phosphate- decreases. The predicted consequence of any failure of
free curve in Fig. 3 cannot be attributed to oxidation buering would be to exacerbate the uncompetitive in-
of the solvent. In a similar manner to the initial quali- hibition by the proton according to reaction (9). The
tative analysis in our rst study [1] linear regions in signicance of electrolyte buer capacity on the phos-
Hanes [24] plots of [H2O2]/j vs. [H2O2] were found. phate dependence was assessed by conducting a second
This is consistent with a binding site mechanism oper- series of microelectrode experiments with the inclusion
ating also in the phosphate-free system with maximum of 0.1 M citrate buer at pH 7.3 throughout whilst
rate (at saturation) and binding constants, KM, increas- varying [PO34 ]tot. The response of a Pt microelectrode
ing with potential. Exact MichaelisMenten kinetics to [PO3
4 ]tot in this mixed phosphatecitrate buer is
are not exhibited in that lower than expected rates are shown in Fig. 6 as a function of [H2O2]. At all anodic
found at low [H2O2], a phenomenon we previously potentials examined the trend shown in Fig. 6 is main-
demonstrated was consistent with competitive inhi- tained with saturation kinetics exhibited with a greater
bition of the binding site by O2 [1]. There is no evi- response for higher [PO3 4 ]tot and no minima at low
S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582 4579

Fig. 7. A series of responses for a platinum electrode in a Fig. 8. Steady-state responses (Q) at a platinum microelec-
thin-layer ow cell at a xed potential (E=+584 mV vs. Ag/ trode at a xed potential (E=+584 mV vs. Ag/AgCl) with
AgCl) connected to an FIA system for two series of test sol- [PO3
4 ]tot=100 mM and [H2O2]=14 mM for a range of buer
utions at pH 7.3. (a) 5 replicates of 10 mM H2O2 and 50 mM pH's. The smooth dashed curve is the best t for Eq. (16) to
phosphate. (b) 5 replicates of 10 mM H2O2 and 100 mM the experimental data.
phosphate. 100 mM phosphate carrier, ow rate was 250210
ml/min.
electrode to H2O2 (not shown here) is similar to that
reported by us for rotating disc electrodes [1].
[PO3
4 ]tot. This is consistent with our interpretation However, lower current densities and more pro-
that phosphate is involved in a binding site and the nounced saturation kinetics are exhibited due to the
dependence cannot be attributed solely to a loss of relatively low solution velocity across the face of the
buer capacity at low [PO3 4 ]tot. electrode in the thin-layer ow cell (akin to low ro-
Two further observations may be made on the data tation rates in rotating disc experiments) resulting in
shown in Fig. 6. First, the current obtained at any poor removal of the inhibiting species O2 and H+. Of
given phosphate and H2O2 concentration is greater in greater interest is the ability to investigate a series of
the presence of citrate indicating that the enhanced increasing and decreasing [PO3 4 ]tot with xed [H2O2] in
buer capacity may play some part. Secondly, in the a rapid sequence of experiments. One such sequence is
absence of phosphate, the oxidation process does not shown in Fig. 7. Here the relative response to 10 mM
revert to the high rate shown in Figs. 3 and 4. This H2O2 in 50 and 100 mM phosphate buer is displayed.
may indicate that citrate inhibits the rapid phosphate- First, a series of 5 replicate 50 mM phosphate sol-
free mechanism to some extent. utions were passed through the thin-layer ow cell fol-
The reversibility of an electrode (within a single pol- lowed by 5 100 mM buer, 5 50 mM buer and
ishing and pretreatment regime) to a series of test sol- then nally 5 100 mM buer.
utions with varying phosphate concentration was The extent of any carry over of H2O2 is negligible
determined. This was achieved using a thin-layer ow since the response returns to zero after every entrained
cell containing a planar Pt electrode connected to an volume of sample is swept out of the thin layer cell by
FIA system. The response of the thin-layer ow cell the carrier (0.1 M phosphate buer). The average re-
4580 S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582

tial and [H2O2] is shown in Fig. 8. The maximum oxi-


dation rate occurs at pH 6.8, decreasing markedly
under more basic or acidic conditions so that half the
maximum response is attained at pH 6.0 and 7.5.
The potential dependence for each pH at a xed
[H2O2] is shown in Fig. 9. The proles for pH >6.4
are similar to the 100 mM phosphate curve in Fig. 3
(although the magnitude of the response varies as
shown in Fig. 8), with maxima developing ca. +600
mV vs. Ag/AgCl. The potential dependence for pH
<6, however, is similar to the exponential proles for
the low [PO34 ]tot curves in Fig. 3, with the response
decreasing markedly with decreasing pH. The potential
dependence is further complicated by the open-circuit
potential shifting to more anodic potentials with
decreasing pH. This is consistent with the linear sweep
voltammetry experiments for H2O2 on Pt reported by
Guilbault and Lubrano [25] where similar shifts were
found. Interpolation of the present data, however,
does not reveal clear Nernstian behavior of the open-
circuit potential.
We interpret the results presented in Figs. 8 and 9 in
two parts. At all pHs we propose that H2 PO 4 is the
species required for binding site formation. The rapid
drop in response above pH 6.8 is consistent with the
decrease in [H2 PO 4 ] given pKa2=7.3 [26]. Thus, reac-
tion (6) may now be written
KBS
Fig. 9. Steady-state responses at a platinum microelectrode
PtPS H2 PO *
4 )PtBS : 14
for [PO3
4 ]tot=100 mM and [H2O2]=14 mM as function of
potential (vs. Ag/AgCl) for a range of pH's: (^) 4.1, (q) 5.3, This is consistent with the FTIR studies of phosphate
(R) 6.0, (Q) 6.4, (.) 6.8, (w) 7.3, (+) 7.5, (r) 7.8, () 8.2, adsorption on Pt at pH 6.8 reported by Ye et al. [27]
(^) 9.2 and () 10.1. where it was found that H2 PO 4 was the species inter-
acting with the surface.
The decrease in response at pH <6.8 is greater than
sponse for replicates at each phosphate concentration
that expected if the uncompetitive inhibition by the
demonstrates that the eect of varying phosphate is re-
proton according to reaction (9) were to be the sole
versible and indicates that the kinetics for attaining
pH eect. We predict that the bulk pH must fall below
steady-state conditions for Eq. (6) are relatively rapid.
pH 3 before even a 5% decrease would occur based
The spread in peak maximum values in each set of
upon the kinetic and equilibria data reported pre-
replicates relates to the reproducibility of the FIA sys-
viously [12]. This marked decrease in response over the
tem. The ow cell is not compensated for iR eects
range pH 6.8 to pH 5 can also not be attributed to
since the solution resistance is inherently varying
decrease in [H2PO4-] at low pH since pKa1=2.15 [26].
during the sequence of test solutions. We have deter-
We propose that this low pH inhibition mode must act
mined, however, that the solution resistance in this cell
to hinder the formation of the binding site in reaction
is too small to be the cause for the 50 mM phosphate
(14). The simplest possibility is that the precursor site,
having lower response than the 100 mM.
PtPS, undergoes protonation to give an inactive site,
The steady-state response of a Pt microelectrode to
H2O2 in a range of phosphate buers with KH
[PO3 PtPS H
)
*
PtPS  H : 15
4 ]tot=0.1 M and varying pH was determined. In
general, for pH >6.4 the H2O2 concentration depen-
dence proles were similar to those shown for a rotat- We ascribe the exponential potential dependence in
ing disc electrode at pH 7.3. At lower pH, however, Fig. 9 at pH <6.4 to a progressive decrease in KH
the saturation kinetics proles are not as distinct with with more anodic potential.
development of sigmoidal responses. A selection of Incorporation of this equilibrium into a mass-bal-
these responses as a function of pH for a xed poten- ance expression for all surface species (excluding
S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582 4581

Fig. 10. Summary of the mechanism for oxidation of hydrogen peroxide at both the phosphate-mediated binding site, PtBS, and
phosphate-free precursor site, PtPS. The mechanism above the dashed line is that which predominates in phosphate buer under
physiological conditions as in papers IIII [1,2,12]. Below the dashed line the mechanism is that found in phosphate free systems.

PtBSH2O2H+ and PtBSO2 since we assume these have smooth dashed curve in Fig. 8 and is for
negligible contribution under microelectrode con- KH=2.5  103 mol1 m3 and KBS=0.1 mol1 m3.
ditions) yields a modied rate equation. There is good qualitative agreement which indicates
that the proposed inhibition of the precursor site is not
j k2 NK1 H2 O2 KBS H2 PO
4 = inconsistent with the experimental data.
16
1 KH H K1 H2 O2 KBS H2 PO
4 :
4. Conclusions
The predicted form of this response can be compared
with the experimental data using the value for K1 The data presented in this paper indicate that the
reported previously [12] and [H2PO4 ] calculated from mechanism proposed in our previous studies [1,2,12]
pKa1, pKa2 and pKa3 for phosphate [26]. The optimized was not complete. The binding site was labeled
t of Eq. (16) to the experimental data is shown as a Pt(OH)2 in our previous work. We now recognize that
4582 S.B. Hall et al. / Electrochimica Acta 44 (1999) 45734582

this is an inappropriate label given the involvement of [5] V.G. Prabhu, L.R. Zarapkar, R.G. Dhaneshwar,
H2 PO 4 in forming the active site. This does not
Electrochim. Acta 26 (1981) 725.
detract, however, from the utility of this mechanism in [6] G. Jonsson, L. Gorton, Electroanalysis 1 (1989) 465.
describing the oxidation of H2O2 at Pt in phosphate [7] Y. Zhang, G.S. Wilson, J. Electroanal. Chem. 345
buers at xed concentration and pH. We have (1993) 253.
[8] J.R. Woodward, R.B. Spokane, Commercial biosensors,
demonstrated there is a relationship between phos-
in: G. Ramsay (Ed.), Chemical Analysis Monograph
phate buer concentration and the number of binding
Series, vol. 148, John Wiley, New York, 1998, p. 227.
sites at the electrode surface and hence the overall rate, [9] D.A. Johnston, M.F. Cardosi, D.H. Vaughan,
with saturation kinetics exhibited at all [PO3 4 ]tot. In Electroanalysis 7 (1995) 520.
the absence of phosphate, saturation kinetics are also [10] J.J. Lingane, P.J. Lingane, J. Electroanal. Chem. 5
observed for H2O2 oxidation and at high anodic poten- (1963) 411.
tials this mechanism operates at a higher rate than in [11] A. Hickling, W. Wilson, J. Electrochem. Soc. 98 (1951)
the presence of phosphate. H2 PO 4 has been inferred to 425.
be the species involved in the formation of the binding [12] S.B. Hall, E.A. Khudaish, A.L. Hart, Electrochim. Acta
site based upon the pronounced pH dependence exhib- 44 (1999) 2455.
ited at pH >6.8. A new form of proton inhibition of [13] J.J. Lingane, J. Electroanal. Chem. 2 (1961) 296.
the precursor site has also been proposed to account [14] L. Miller, J. Electroanal. Chem. 16 (1968) 531.
for the depression in response at pH <6.8. The modi- [15] F.C. Anson, J.J. Lingane, J. Am. Chem. Soc. 79 (1957)
ed mechanism is summarized in Fig. 10. 4901.
[16] L. Gorton, Anal. Chim. Acta 178 (1985) 247.
[17] G.J. Moody, G.S. Sanghera, J.D.R. Thomas, Analyst
111 (1986) 605.
Acknowledgements [18] D.S. Bindra, G.S. Wilson, Anal. Chim. Acta 61 (1989)
2566.
The authors wish to acknowledge the Massey [19] D. Wijesuriya, M.S. Lin, G.A. Rechnitz, Anal. Chim.
University Research and Research Equipment Fund Acta 234 (1990) 453.
(MURF and MUREF 1996 and 1997) for contri- [20] E. Csoregi, L. Gorton, G. Marko-Verga, Anal. Chim.
butions to the funding for this work. Acta 273 (1993) 59.
[21] Q. Chi, S. Dong, Anal. Chim. Acta 278 (1993) 17.
[22] F. Mizutani, S. Yabuki, S. Izjima, Electroanalysis 7
References (1995) 706.
[23] P. He, L.R. Faulkner, Anal. Chem. 58 (1986) 517.
[1] S.B. Hall, E.A. Khudaish, A.L. Hart, Electrochim. Acta [24] C.S. Hanes, Biochem. J. 26 (1932) 1406.
43 (1998) 579. [25] G.G. Guilbault, G.J. Lubrano, Anal. Chim. Acta 64
[2] S.B. Hall, E.A. Khudaish, A.L. Hart, Electrochim. Acta (1973) 439.
43 (1998) 2015. [26] M.A. Habib, J.O'M. Bockris, J. Electrochem. Soc. 132
[3] J.E. Harrar, Anal. Chem. 35 (1963) 893. (1985) 108.
[4] H. Lundback, G. Johanson, Anal. Chim. Acta 128 [27] S. Ye, H. Kita, A. Aramata, J. Electroanal. Chem. 333
(1981) 141. (1992) 299.

También podría gustarte