Exchange Current Density 2
Exchange Current Density 2
Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta
A R T I C L E I N F O A B S T R A C T
Article history: Using a high mass transport floating electrode technique with an ultra-low catalyst loading (0.84–
Received 2 March 2015 3.5 mgPt cm2) of commonly used Pt/C catalyst (HiSPEC 9100, Johnson Matthey), features in the hydrogen
Received in revised form 30 June 2015 oxidation reaction (HOR) and hydrogen evolution reaction (HER) were resolved and defined, which have
Accepted 30 June 2015
rarely been previously observed. These features include fine structure in the hydrogen adsorption region
Available online 14 July 2015
between 0.18 < V vs. RHE < 0.36 V vs. RHE consisting of two peaks, an asymptotic decrease at potentials
greater than 0.36 V vs. RHE, and a hysteresis above 0.1 V vs. RHE which corresponded to a decrease in the
cathodic scan current by up to 50% of the anodic scan. These features are examined as a function of
hydrogen and proton concentration, anion type and concentration, potential scan limit, and temperature.
We provide an analytical solution to the Heyrovsky–Volmer equation and use it to analyse our results.
Using this model we are able to extract catalytic properties (without mass transport corrections; a
possible source of error) by simultaneously fitting the model to HOR curves in a variety of conditions
including temperature, hydrogen partial pressure and anion/H+ concentration. Using our model we are
able to rationalise the pH and hydrogen concentration dependence of the hydrogen reaction. This model
may be useful in application to fuel cell and electrolyser simulation studies.
ã 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
1. Introduction of 5.7 A cm2Geo for a 10.15 mgPt cm2 catalyst loading (60% Pt/C
HiSPEC 9100 catalyst, Alfa Aesar). Assuming this peak current
The hydrogen oxidation reaction (HOR), along with the density was the absolute maximum geometric current density
hydrogen evolution reaction (HER) is one of the most studied (jGeo,max) for the technique, a corresponding mass transport
reactions of modern science due to its simple reaction of one coefficient (kMT) equal to 58 cm s1 was attained; equivalent to a
hydrogen molecule going to two protons, while releasing two rotation rate of 4 109 rpm on the RDE. kMT was calculated by
electrons. This reaction also gains considerable attention as the combining the Faraday equation with Fick’s law
anode reaction of a fuel cell. However, due to its facile nature, the
kMT ¼ J=cH2 (1)
kinetics have often been obscured by the relatively slow H2
transport capabilities of the electrochemical cell. This is evident Where
from room temperature (or room temperature corrected, where
J ¼ jGeo;max =nF (2)
the activation energy was also stated) exchange current densities
spanning three orders of magnitude in the literature, j0 = 0.2 – J is the hydrogen flux, n = 2 for the HOR and cH2is the concentration
60 mA cm2 [1–11]. of hydrogen, taken as the saturation concentration in Nafion at
A previous paper by this group [12] showed ultra-high mass 5.1 107 mol cm3 [13]; due to the likelihood that a thin layer of
transport can be achieved on a floating electrode. This technique liquid or Nafion covers the catalyst in this gas diffusion electrode
combined gaseous transport through a porous support and an [14,15]. This assumes the properties of the Nafion thin film layer
ultra-thin catalyst layer to obtain a geometric peak current density remain the same as bulk Nafion.
It is worth noting that the real kMT of the floating electrode
technique is likely to be much greater as Fig. 10 in [12] shows that
* Corresponding author. Tel.: +44 20 75945831. the geometric current density is still increasing linearly with
E-mail address: [email protected] (A.R. Kucernak).
http://dx.doi.org/10.1016/j.electacta.2015.06.146
0013-4686/ ã 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
764 C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776
catalyst loading, with no sign of bulk mass transport effects, From these benefits, the floating electrode gives an
let alone a mass transport limitation. idealised environment to measure the HOR/HER, defining features
kMT is also likely to be very high for a fuel cell [11,16,17]. which have rarely been observed. These features include fine
However, with a comparatively larger catalyst loading than the structure in the hydrogen adsorption region between 0.18 and
floating electrode, the rapid increase in current with over-potential 0.36 V vs. RHE consisting of two peaks, an asymptotic decrease at
for the HOR means the working limit of the potentiostat can be potentials greater than 0.36 V vs. RHE, and a hysteresis above 0.1 V
reached at small over-potentials. For example, to reach a current vs. RHE. This paper explains these features in terms of surface
density of 600 mA cm2Spec (as stated as the jspec,max in [12]), with adsorbed species and edge and facet sites on the catalyst. In
an equivalent PEFC loading of 0.2 mgPt cm2Geo and a catalyst metal addition, intrinsic catalyst properties are extracted using a
area of 89 m2 g1 (as the HiSPEC 9100 used in this study), the numerical model to fit the HOR curves in a variety of conditions
geometric current density would be over 100 A cm2Geo. Even including temperature, hydrogen partial pressure and anion/H+
when reducing the catalyst loading, the area of a fuel cell electrode concentration.
causes the absolute current density to be large. Neyerlin et al. [16]
reached currents of 3 A cm2Geo at only 60 mV vs. RHE with a
25 cm2 electrode at a 35 mgPt cm2 catalyst loading, after correct- 2. Experimental
ing for iR losses. The effect of uncompensated resistance prevents
the measurement of the HOR at higher over-potentials and could A commercial 60 wt.% Pt/C catalyst (HiSPEC 9100, Alfa Aesar,
incur large errors from their correction. metal area of 89 m2 g1 [18]) and high purity gases (N2, H2 and O2 at
The floating electrode technique uses small quantities of >5.8 N from Air Products) and acids were used; perchloric acid
catalyst (sub 10 mgPt cm2) deposited uniformly and homo- from VWR (Merc Suprapur) and GFS chemicals (Veritas double
geneously across a small catalyst area (typically 1 mm radius distilled) were both used with negligible difference in performance
spot, i.e. 0.0314 cm2 geometric area) which gives three advantages: (not shown) and sulfuric acid from VWR (Aristar grade). The
such a thin catalyst layer allows all the catalyst to see an equivalent electrodes were made up as in [12]. Polycarbonate track etched
environment, giving access to electrochemical fine structure (e.g. membranes (Sterlitech, PCTF0447100, 0.4 mm pores) were coated
peaks) which would otherwise be blurred in an electrode with a with 100 nm of gold by sputter deposition, and after deposition of
gradient of conditions across or through its catalyst layer; the the catalyst, the pores were hydrophobised with an amorphous
absolute current remains low (mA’s equivalent to A cm2 fluoropolymer (Teflon AF 2400, 2.1 mg cm2Geo). The catalyst was
geometric current), even at high over-potentials, reducing the deposited via the vacuum filtered catalyst (VFC) method described
possible errors from correcting for iR effects and local Joule heating in [12] to achieve uniform and homogeneous catalyst spots at
effects; and small amounts of catalyst are needed for one test, ultra-low catalyst loadings (mgPt cm2) from a dilute catalyst ink.
making it beneficial to test novel catalysts produced in small The ink makeup was optimised to use small quantities of catalyst
quantities. This allows the floating electrode to be an advantageous (typically 1 mg), especially useful for novel catalysts synthesised in
technique for testing novel fuel cell catalysts ex-situ while small quantities. A stock ink containing a catalyst to solvent ratio of
retaining the high mass transport conditions expected in a PEFC. 1:10 (mg:ml); the solvent contained 50% butyl acetate (Sigma,
Additionally, a benefit over the RDE technique is that as the anhydrous grade) and 50% from a mix of propan-2-ol (VWR,
electrolyte is stationary, convective transport of environmental Normapur analytical reagent) and the PFSA solution (DuPont
contaminants (chloride etc.) is significantly reduced, leading to DE521 Nafion solution, 5 wt%) to give a catalyst to PFSA ratio of 1:1
better performance in the presence of even vanishingly small (volume). An aliquot of this stock catalyst ink was then diluted to
quantities of solution impurities. 500 ml with a 50:50 mix of butyl acetate and propan-2-ol to
Fig. 1. Voltammogram of the HOR and HER on HiSpec 9100 60% Pt/C catalyst at a loading of 0.84 mgPt cm2. 101 kPa H2, 4 mol dm3 HClO4, 298 K, 10 mV s1. The ordinate axis
corresponds to the specific current density (left), geometric current density (first right) and mass activity (second right), refer to Eq. (25) for relation between the units.
C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776 765
produce the required loading (see reference [12] to correct for loss The Epeak,low of the HOR at 0.18 V vs. RHE has a current density of
of catalyst through the pores). 0.51 A cm2Spec, corresponding to a turnover number of 1200H2
The performances of the floating electrodes were measured molecules per surface platinum atom per second. Previously, we
with a Gamry Reference 600 potentiostat in a water jacketed three showed that the peak current density remains constant at
electrode electrochemical cell utilizing a Pt counter electrode and a 0.6 0.06 A cm2Spec with no decrease in current density as the
RHE in a Luggin capillary configuration. These electrodes were catalyst loading increased from 0.72 to 10.15 mgPt cm2, showing
floated on liquid electrolytes with the reactant gas supplied to the no bulk mass transport limitations [12]. The HER current density at
catalyst through the hydrophobised pores, allowing high mass the cathodic limit is greater than ten times the current density of
transport. 4 mol dm3 perchloric acid was used to minimise the the HOR limiting current density, and as the transportation of H2
uncompensated resistance at the high currents achieve from the away from the electrode looks adequate (i.e. no bubble formation);
HOR, unless otherwise stated. Conditioning the electrodes by it is logical to assume this electrode is capable of supplying H2 to
switching between a HOR scan (100% H2, 50 mV s1, 0.1–1 V vs. the electrode at the same rate. This adds further confirmation that
RHE) and an ORR scan (100% O2, 50 mV s1, 1.1–0.3 V vs. RHE), with the peak current density is related to an electrokinetic or
an intermediate N2 purge, 10 times or until the scan was repeatable adsorption step at the Pt surface such as a hydrogen adsorption
at 25 C was found to improve the repeatability of the measure- limitation or a local hydrogen or proton transport limitation;
ment across electrodes, while no loss of performance was observed possibly caused by a thin layer of Nafion or water as discussed in
due to degradation. The cell was held at the relevant temperature [15].
using a re-circulator (Polyscience digital temperature controller,
0.1 C). Gas mixtures were obtained using gas flow controllers 3.1. HOR reaction
(Bronkhurst EL flow series). Uncompensated resistances were
corrected using the high frequency intercept of impedance As our purpose in this paper is to consider some aspects of the
measurements at a range of voltages. hydrogen oxidation reaction close to the equilibrium potential, we
To measure the potential of zero total charge (PZTC), CO will used a simplified model which neglects both the potential
displacement was recorded by dosing the electrode with CO at a dependence of adsorption site density, and assumes Heyrovsky–
range of potentials while recording the current, a method Volmer kinetics.
described in detail in [19,20]. This technique assumes CO is a
kMT;gas
surf
strong adsorbate, able to displace all of the previously adsorbed Hbulk
2 ! H2 (5)
species, however, the technique cannot distinguish between
species of the same charge. As CO is a neutral species, CO
displacement causes a quantity of charge associated with the k1
Pt þ Hsurf ? H3 Oþ;surf þ Pt Had þ e (6)
desorption (e) of the previous species to be released, as shown in 2
k1
Eqs. (3) and (4) for hydrogen and anion (A) species with partial
charge transfer (d), respectfully.
k2
Pt Had þ H2 Osurf ? H3 Oþ;surf þ Pt þ e (7)
Pt H þ CO ! Pt CO þ Hþ þ e (3) k2
Pt Að1dÞ þ CO þ de ! Pt CO þ A
kMT;liq
(4) H3 Oþ;surf ? H3 Oþ;bulk (8)
kMT;liq
Where kMT,gas and kMT,liq are the mass transport coefficients in gas
3. Results and discussion and liquid respectively. We will assume for simplicity that that
mass transport of reactant and products are fast (although see
Fig. 1 shows a typical CV of the HOR (Expanded in the inset) above), and so we only consider Eqs. (6) and (7). Although we do
using the ultra-low catalyst loading floating electrode technique not explicitly include water within the kinetic equations as it acts
(loading = 0.84 mgPt cm2). The scan also extends into the HER as a solvent in our system, it might be important to include this
region. The HOR shows fine structure in the hydrogen adsorption parameter under fuel cell systems, especially when operating at
region between 0.18 and 0.36 V vs. RHE consisting of two peaks low relative humidities. Eqs. (6) and (7) can be analytically solved
(herein called EPeak,low and EPeak,high for the peak at the lower and under steady state conditions, and we will provide a full derivation
higher potential, respectively), an asymptotic decrease at poten- in a future paper, however an abbreviated derivation follows. The
tials greater than 0.36 V vs. RHE, and a hysteresis above 0.1 V vs. hydrogen coverage, under steady-state conditions can be found by
RHE which corresponds to a decrease in the cathodic scan current finding the solution to the equation
by up to a half in comparison to the anodic scan; as reported duHad
previously [12,15]. For the HER, the current density reached ¼ k1 aH2 surf 1 uHad k1 aH Oþ;surf u Had k2 uHad
dt 3
8 A cm2Spec at a potential of -0.33 V vs. RHE with no sign of a þ k2 aH Oþ;surf 1 uHad
3
limitation, or current disruption from bubble formation. Such high ¼0 (9)
current densities without bubble formation, and hence potential
disruption of the catalyst layer, are unparalleled and are typically where uHad is the surface coverage of adsorbed hydrogen. Solution
the limiting factor for measuring the HER on the RDE. Instead, the of (9) provides
voltammogram remained smooth; confirming the fast mass k1 aHsurf þ k2 aH þ;surf
3O
transport of H2 gas from the catalyst that this floating electrode uHad ¼ 2
(10)
k1 aHsurf þ k2 þ k1 aH þ;surf þ k2 aH þ;surf
technique is capable of. The current density at the cathodic scan 2 3O 3O
KaH þ;surf K >> 1 or K << 1). A formula similar to Eq. (20) is often used in
3O
previous literature without appreciation of the conditions under
F
f ¼ (15) which it is applicable. Under these simplifying conditions
RT
R ðT T Þ
Ea 1 1
Where f = F/RT, with their usual meanings. Substituting into Eqs. F
jðhÞ ffi j0;obs he h af
(10) and (11) provides the potential dependence of hydrogen RT
g e
coverage and current density
j0;obs ¼ 4Fk0;obs aH surf
2
aH Oþ;surf
3
(21)
eq
eq k2 ð1aÞf h
k1 aH2 surf eaf n þ K e
uHad ðhÞ ¼ eq eq eq
eq
k2
(16)
ðk1 aH2 surf þ k2 Þeaf h þ ðKk1 aH2 surf þ Þeð1aÞf h It is important to note that if the above criteria are not met, then
K
this simplified form is not applicable, and inconsistent results will
be obtained. One example of such inconsistent results will be non-
linearity in the reactant stoichiometry plots (Log(jo) vs log
eq eq
j¼F ð1 u Had ðhÞÞ þ k2 uHad ðhÞ eaf !
k1 aH surf
h
2
eq (areactant)), although in practice many ignore such nonlinearities
k and attempt to force a linear fit.
Kk1 aH surf u Had ðhÞ þ 2 ð1 uHad ðhÞÞ eð1aÞf h Þ
eq
(17)
2 K In the above analysis, we have for simplicity assumed that
the number of available sites for adsorption is 1 uHad . In fact,
It can be seen that j depends in a complicated way
eq eq as we will show below, the situation is somewhat more
onK; k1 ; k2 ; aH surf and (through K), aH Oþ;surf . For instance, when complicated and that the number of available sites for adsorption,
2 3
the Heyrovsky step is fast compared to the Volmer step, then a first especially at higher potentials may be much less than this term
order dependence in hydrogen concentration is expected. If we suggests. A more accurate approach may be to write this term as
assume that both Heyrovsky and Volmer step show the same 1 uSpectator u Had where the extra term is associated with the
activation energy then we can include the temperature depen- species that adsorb more strongly than the hydrogen on the
dence through the Arrhenius equation (assuming a standard surface. The change in aH2 is the activity of dissolved hydrogen and
temperature of 298 K for T*)
C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776 767
Where aH2 and PH2 are the saturation activity and partial pressure
of pure hydrogen at standard conditions.
Fig. 2a) shows the anodic scans of the HOR in pure H2 and at
different H2 partial pressures between 1 p(H2) 101 kPa in N2.
The x-axis is plotted in terms of over-potential (h), denoting that
the potential scale has been corrected for the shift in equilibrium
potential (Ee) for the set conditions through the Nernst equation, in
the rearranged form of Eq. (23).
pffiffiffiffiffiffiffi !
2:303RT aH 2 aH 2 O
Ee ¼ E0 log (23)
nF aH3 Oþ
3.3. Anion Species adsorb to a small extent, especially as the perchlorate ion is at a
high concentration (4 mol dm3) in this experiment. This conclu-
To test the effect of anion adsorption on the HOR, two sion is also supported by recent work by Omura et al. [26] who
electrolytes commonly used for their so called specifically and through quartz crystal microbalance and IR spectroscopy see
non-specifically adsorbed anions were used; sulfuric and indications of perchlorate adsorption at higher potentials. The
perchloric acid. Anion interaction at the electrode surface has potential dependence of this inhibition is explored further in the
been shown to have a strong displacement and blocking effect on sections Platinum PZTC and Hysteresis of the HOR. In these
the ORR; the stronger the adsorption, the greater the blocking electrodes, which utilise a small quantity of PFSI binder, it might be
effect [22]. As the electrocatalyst is bound with Nafion ionomer, we expected that some effect of anion adsorption from the relatively
may also expect some effect due to the teathered sulfonate groups strongly adsorbing sulfonic acid groups of the PFSI might have
which have been previously shown to lead to strong anion some contribution to the shutdown of hydrogen oxidation at
adsorption effects [13,23]. higher potential. However, due to the low coverage of the sterically
Fig. 3 shows the HOR using the floating electrode in both hindered sulfonate groups limited surface coverage (estimated at
electrolytes. Close to the origin and going up to EPeak,low, the activity 0.1 monolayers on Pt(111)) [23], it is unlikely to be the major
of the electrode in the two acids appears very similar. The potential cause the 5 fold decrease in activity observed between 0.2 and
of EPeak,low in the two electrolytes are within 20 mV of each other. 0.8 V vs. RHE here.
Therefore, in the region below EPeak,low, no effect of anion Bagotzky and Osetrova [1] observed HOR inhibition from
adsorption was observed and we conclude the exchange current Br and I anion adsorption which shows the greater the
density can be measured in either acid with no significant anion adsorption strength, the greater the kinetic inhibition.
difference. This is in agreement with Rau et al. [24] when using Although we cannot exclude the possibility of some very small
polycrystalline Pt with no perflorosulfonic ionomer (PFSI). The lack amount of free chloride in our electrolyte (a contaminant in
of interference is most probably due to the potential being below even the most stringently prepared perchloric acid), we
the PZTC, where anion adsorption is minimal, as discussed below. have performed dosing experiments which suggest that the
This is also shown in Table 1 and Fig. 5 below, where the exchange level is less than 0.4 mmol dm3. Furthermore, a tremendous
current density is the same in sulfuric acid as it is in perchloric acid benefit of our approach compared to RDE experiments is the
for equivalent acid concentrations. absence of forced convection. Hence the transport of adventi-
As the potential passes EPeak,low, however, there begins to be a tious poisons is much slower (ca. 100-fold) compared to RDE
significant difference between the two electrolytes. While the HOR measurements where the diffusion layer is much thinner
activity decays in both solutions, in ionomer/perchloric acid, a (ca. 1 mm) compared to the boundary layer in quiescent
5 fold decrease in activity is seen between 0.2 and 0.8 V vs. RHE electrolyte (ca. 100 mm).
(before oxide formation is thought to begin), but in sulfuric acid, Hence there is a clear indication that the HOR is affected by
the stronger adsorbing anion, the magnitude of Epeak,high is anion adsorption above the EPeak,low. The reduced activity for the
significantly suppressed, and the HOR remains suppressed well HOR covers the entire potential region of the ORR, and agrees with
into high potentials (>0.8 V vs. RHE); leading to a 10 fold decrease the reduced activity of the ORR seen in sulfuric acid in comparison
in activity between 0.2 and 0.8 V vs. RHE. The extra suppression to perchloric acid. In this sense, Wiberg et al. [27] correlated the
over perchloric acid can also be seen clearly on the reverse scan. inhibition of the HOR with oxide coverage (generally >0.8 V vs.
This suggests both electrolyte systems cause perturbation to the RHE) on a polycrystalline Pt disk RDE and applied this surface
HOR, with the stronger anion giving a stronger perturbation. Kita blocking coverage to the ORR. The floating electrode technique
et al. [25] performed experiments on single crystal platinum should allow the HOR to be used as a probe for the state of the Pt
electrodes in static electrolytes and observed a decrease in activity surface across the entire ORR potential window.
at higher potentials (ca, 0.5 V vs. RHE), with a more pronounced
decrease or stronger inhibition for sulphuric acid then perchloric 3.4. Proton Activity
acid, similar to in Fig. 3. They proposed that as perchloric acid is
not a specifically adsorbing anion, the HOR inhibition in perchloric Fig. 4 and Table 1 show the effect of electrolyte pH and ionic
acid must be down to reorientation of surface water molecules in strength on the activity of the HOR in a) perchloric acid and b)
the double layer. In contrast we consider that perchlorate does sulfuric acid between 0.1–4 mol dm3. A decrease in the gradient
Table 1
Properties of the HOR at 298 K, 101 kPa H2 in different electrolytes. Exchange current densities calculated using Eq. (21).
Electrolyte conc./mol dm3 pH j0*/mA cm2 EPeak,low/V vs. RHE RHFR y/V Predicted Rsolnz/V
HClO4
4 0.6 128 0.18 1.2 1.4
1.59 0.2 92 0.24 2.4 2.1
0.63 0.2 82 0.30 5.1 4.2
0.25 0.6 50 0.40 11 10
0.1 1 32 0.54 25 26
HClO4 NaClO4
4 0 0.6 150 0.16 3.1 2.8
0.4 3.6 0.4 38 0.31 3.6 5.9
H2SO4
4 0.6 138 0.20 1 1.3
2 0.3 118 0.20 1.5 1.7
0.5 0.3 70 0.27 4.8 4.8
0.1 14 28 0.51 17 23
*Calculation of “simple” exchange currents is discussed in Section 3.5, using Eq. (21). No correction made for water vapour.
yHigh frequency resistance measured using electrochemical impedance spectroscopy.
*Resistance (R) calculated using conductivity values from [28] for HClO4 and H2SO4 and [42] for NaClO4, the distance between the WE and RE (l) is 0.5 cm apart from the mixed
electrolyte section (HClO4 + NaClO4) in which the gap was 1 cm. For the mixed electrolyte, it is a combined resistance assuming negligible effect of ionic activity coefficients.
C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776 769
Fig. 4. Electrolyte concentration effect: Voltammograms of the HOR with a Fig. 5. The dependence of pH on the logarithm of the exchange current density of
1.4 mgPt cm2 Pt/C catalyst run in a range of electrolyte concentrations of both the different electrodes presented in Table 1, giving a gradient of about 0.4.
perchloric and sulphuric acid. 101 kPa H2, 10 mV s1, 298 K. Voltammograms are iR Temperature: 298 K, 101 kPa H2.
corrected using electrochemical impedance data.
across the polarisation region and peak shift to higher potentials is 2) An effect of proton concentration on the exchange current
clearly observed in both solutions. As there was no supporting density
electrolyte in these experiments, the resistance also increased with At low current densities, the production of protons is not so
reducing acid concentration, and was corrected using the high rapid, therefore the decreased gradient is not likely to be related to
frequency intercept from impedance spectra. To check that the iR a local pH change alone. Fig. 5 shows the pH dependence of the
correction wasn't introducing a systematic error, experimental logarithm of exchange current density in the electrolytes listed in
spectra were collected for an electrode placed in 4 mol dm3 Table 1, from pH 0.6 to 1. An almost linear dependence of
perchloric acid and 0.4 mol dm3 perchloric acid plus 3.6 mol dm3 exchange current density with pH in both sulfuric acid and
sodium perchlorate electrolytes. For these two electrolytes the perchloric acid is seen with a gradient of close to 0.4, very similar
perchlorate anion concentration was the same, and the resistance to the gradient of 0.5 presented by Bagotzky and Osetrova [1]. The
did not change too significantly (results shown in Table 1). For the derivation of the HOR/HER (Eqs. (9)–(20)) in terms of a Heyrovsky–
0.4 mol dm3 perchloric acid with 3.6 mol dm3 sodium perchlo- Volmer reaction gives a dependence on hydrogen ion concentra-
rate support, both the exchange current density and peak potential tion, and that phenomenological dependence will depend on the
correlated well to results in 0.4 mol dm3 perchloric acid without precise values of the electrochemical rate constants (see below).
sodium perchlorate. In addition, the theoretical resistance was Durst et al. and Sheng et al. [31,32] also observed a reduction in
calculated assuming a distance between the working and reference exchange current density with pH and offered an alternative
electrode of 0.5 cm and the molar conductivity of the respective discussion linking it to a change in hydrogen binding energy (HBE)
electrolytes from [28], which also correlated well with the (proportional to the change in activation energy through the
experimentally measured resistances. Bronsted–Evans–Polanyi relationship [33]). While a change in the
We postulate that two effects dominate the CVs as the proton Pt PZTC with pH was also considered (discussion on PZTC below), it
concentration decreases: has generally been found to decrease in potential with increasing
1) A change in local proton concentration pH [34], while here EPeak,low increases with pH; therefore it is
As calculated above, 1200H2 molecules are oxidised per surface unlikely to be the cause of this increasing EPeak,low.
platinum atom per second at the peak current density. This rapid As will be shown below, the effect is adequately described by
production in protons local to the catalyst surface is likely to have the full analysis of the Heyrovsky–Volmer model provided above.
an effect on the local proton activity. A local pH change would shift
the Nernstian reversible potential and appear as an apparent
increase in polarisation, shifting the peak to higher potentials
(appearing like a resistance), and is likely to be the cause of the
peak broadening. In effect this is a mass transport effect, but of the
product, not the reactant. While this change in local pH should be
prevented by the principle of charge neutrality and possibly even
more so in our system where the presence of PFSI in the catalyst
layer should create a fixed layer of sulfonate anions at the surface
which would likely prevent large pH, we believe it cannot be
discounted. Bagotzky and Osetrova [1] observed this effect when
using a very thin film of electrolyte across a microelectrode. They
concluded that transport of protons away from the catalyst surface
was limited by the thin film and removed the effect by increasing
the film thickness. However, a thicker film caused a larger barrier
for H2 diffusion through which would decrease measured peak
current densities. The opposite effect has been recognised at
microelectrodes during the HER, leading to proton depletion near
the electrode, and hence an increase in local solution resistance,
and enhanced proton migration [29,30]. In this study, most results Fig. 6. Variation of horwith temperature effects. 2.2 mgPt cm2HiSpec 9100 60 % Pt/
C electrode in 4 mol dm3HClO4, 1bar H2, 10 mV s1, CE = Pt wire and RE = RHE. Inset
used a high acidity electrolyte (4 mol dm4 or pH 0.6) to reduce shows an Arrhenius plot of j0 after correction for increased water vapour partial
any change in local pH. pressure with temperature.
770 C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776
g
Table 2 P H2
“Raw” exchange current densities, and exchange current densities corrected by the
Dlog j0 = PH
2
hydrogen partial pressure variation due to the presence of water vapour for the HOR Ea ¼ 2:303R (24)
D1=T
on HiSpec 9100 60 % Pt/C across a range of temperatures in 4 mol dm3 HClO4.
T/ C 5 15 25 30 40 50 60 eq
The rate constant k0 can be related to the exchange current
j0,T /mA cm2 80 116 144 154 192 224 238
density by normalising to standard conditions. Due to the
g
P H2 80 118 148 160 206 254 294 additional observed change in exchange current density by proton
j0;T = PH
*/mA cm2
2 activity for acidic conditions, Eq. (21) relates the exchange current
*exchange current density corrected for H2O vapour partial pressure, g = 1. density through activation energy (Ea) at 25 C, partial pressure
fraction of 1 and pH 0.
3.5. Exchange Current Density and Activation Energy
3.5.1. Full fit of experimental data set
The micro-polarisation region of an anodic scan at different Recent developments in measuring the HOR and HER at high
temperatures in 4 mol dm3 HClO4 is shown in Fig. 6. As the mass transports have promoted a number of attempts to model the
temperature increased, the gradient increased. Additionally, the hydrogen oxidation and evolution reaction in the literature. For
scan cuts the x-axis at slightly negative potentials, which could be a instance Chen and Kucernak [8] considered whether Tafel–Volmer
sign of proton concentration varying with respect to the reference or Heyrovski–Volmer kinetics adequately explained single Pt
electrode, as discussed above. As each curve was scanned from the particle microelectrode results whereas others [35–37] have
HER where protons will have been previously removed, the local considered a potential dependent switch from Tafel–Volmer to
pH will have shifted more alkaline. This shift of the local Heyrovski–Volmer kinetics. While both models can fit the mass
equilibrium potential by 0.7 mV equates to a local pH change of transport limited curve from the RDE and the shoulder in the curve
0.05 pH units. The cathodic scan also does not pass through the observed by Chen and Kucernak [8], both models in their current
origin, but passes at a more positive potential, showing a more forms are not capable of predicting the double peak with a dip in
acidic local pH (results not shown). This effect cannot be accounted current in-between as observed here and in other papers [12,17].
for by the capacitance of the electrode as the equivalent capacitive Clearly from the results above, there is also a significant anion
currents are about 20-fold smaller than the currents measured in surface blocking effect and pH effect which controls the
the presence of hydrogen. electrokinetics of the HOR and any model which is going to
Due to the short potential range of the polarisation curve reproduce the experimental curves needs to include these effects.
(10 mV), linearization of the electrokinetic equations are possible To outline some of the elements needed for such a model, the
and Eq. (19) (or Eq. (21) under suitable limiting conditions) can be potential of zero total charge and potential dependence of anion
used to calculate j0 for each temperature, shown in Table 2. This adsorption are discussed below.
requires the further assumption that through the potential range, However, for the HOR or HER on the anode of a fuel cell or the
the proton activity remains constant (i.e. proton concentration is cathode of an electrolyser, the over-potential typically remains
not perturbed too much by the reaction). As this series of below 10 mV. Therefore, in this paper we provide two different
experiments were completed in the same electrolyte strength and numerical fits to all the data presented across a polarisation range
the current densities are still relatively low, the proton activity is of 10 mV and a variety of experimental conditions. We also show
likely to remain relatively constant. The mass transport correction that these models may be extended to about 50 mV without
factor, which was present in [8], has been removed from Eq. (19) introduction of too much error. Such numerical fits may be useful
and (21) as; 1) with a peak current density of 600 mA cm2Spec at in providing models for fuel cell anodes and electrolyser cathodes.
25 C, the current was less than 10% at 10 mV over-potential and 2) The fits utilise either Eq. (19) or (21). Fits utilising the latter
the peak current density is assumed to be caused by a combination approximate the approaches previously used in the literature in
of the kinetics getting faster until an adsorption/local diffusion which it is assumed that the concentration dependence of protons
limiting current density was reached at larger over-potentials, and and hydrogen can be separated from an extrinsic rate constant,
ufree decreasing due to site blocking species, rather than a mass whereas the fit utilising Eq. (19) is the full explicit Heyrovsky–
transport limitation; i.e., the geometric mass transport limitation is Volmer approach. In both cases the potential was corrected for the
likely to be above this. The peak current density is discussed in small shift in pH which led to the scan not passing through the
detail in the Platinum PZTC section. The exchange current density origin (shift <1 mV in all cases). The fitting parameters when using
eq eq
at 25 C is similar in both sulphuric and perchloric acid (see Fig. 5). Eq. (19) were K; k1 ; k2 ; Ea . In this case there are no stoichiometric
In Fig. 6 it can be seen that between 50 C and 60 C the rate of coefficients for protons and hydrogen as their activity is included
improvement in HOR performance with temperature decreases. explicitly within the kinetic equation. An “effective” stoichiometric
This is due to reduced hydrogen partial pressure associated with coefficient may be calculated by examining how the exchange
increased water vapour partial pressure. To correct for this, the gas current density calculated from the fitted parameters varies as a
phase in the catalyst layer and down the pores of the polycarbonate function of the reactant concentration within the appropriate
track etched membrane was assumed to be at equilibrium with the concentration range. The fitting parameters when using Eq. (21)
were k0 , Ea, e and g . Both fits were performed on all datasets
eq
saturation vapour pressure of water for that corresponding
temperature. This is very plausible considering the catalyst layer simultaneously (20 data sets). The operating conditions for the
is in contact with an aqueous electrolyte. The effect is corrected for datasets are described in a footnote to Table 3.
in the third row in Table 2 assuming a hydrogen reaction exponent, The specific current density (jNorm.Ptsurf.) can be converted to
g, of unity, as discussed in the hydrogen partial pressure absolute current (i), geometric current density (jGeo) or mass
measurements above. activity (jMass) through Eq. (25).
Using the Arrhenius equation, adapted for the exchange current
i ¼ jNorm:Pt surf: A SAPt LPt ¼ jGeo A ¼ jMass A LPt (25)
density with water vapour and hydrogen partial pressure
correction (Eq. (23)), the activity can be plotted vs. the inverse
temperature as in the inset in Fig. 6, giving the gradient as the These terms are linked through the electrode area (A), the
activation energy at 18 kJ mol1. platinum loading (LPt) and the catalyst’s metal area (SAPt). The
C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776 771
Table 3
Summary of values experimentally measured in this paper under standard conditions, aH+ = 1; aH2 = 1; T = 298 K. Global fits were determined by simultaneously fitting 20 data
sets using either Eq. (21) (Global Fit 1) or Eq. (19) (Global fit 2). Hydrogen partial pressures were corrected for water vapour. Grey cells denote the parameters used in the fitting
process.
Table 4
A comparison of literature obtained exchange current densities and activation energies, T close to 298 K.
can all the catalyst be assured to be well supplied with hydrogen electrode surface plus the charge density transferred during the
and therefore fully active. adsorption process equals zero [19,20,38]. Due to the interaction
Table 4 also shows activation energies quoted in the literature; with the electrolyte, different electrolytes or electrolyte concen-
scattered between 10–43 kJ mol1, with the value reported here trations cannot directly be compared, but only used as a guide.
lying between these values. This medium activation energy means Fig. 9 shows the resultant current flow from CO displacement of
that the exchange current density is less affected by temperature the adsorbed species on the HiSpec 9100 Pt/C with a particle size of
then recently reported values [9–11]. Extrapolating the exchange 2.4 nm [18] coated in Nafion at different potentials in 4 mol dm3
current density for an acid concentration range of 1–4 mol dm3 to perchloric acid. The noise in the graph occurs as the electrode is
a more typical operating temperature of a fuel cell at 80 C gives a floating on the electrolyte surface and is slightly disrupted by the
range of 260–440 mA cm2Spec. This range is close 470–600 mA purge of nitrogen or CO; which in this case is bubbled through the
cm2Spec reported by Neyerlin et al. [16] (using their equation with solution. As the potential increases, the charge associated with the
aa + ab = 1 which is in the same form as Eq. (21)), 489 mA cm2Spec displacement current moves from a positive to a negative current,
reported by Wang et al. [11] and 405 mA cm2Spec by extrapolating showing a shift in adsorbed species from cations to anions, inset in
the exchange current density at 25 C from Sun et al. [10] using Fig. 9. The x-intercept shows the PZTC is 0.24 0.01 V vs. RHE. This
their activation energy. However it is much lower than the value is shifted negative in comparison to the PZTC of polycrystal-
770 mA cm2Spec reported by Wesselmark et al. [17]. Note that Sun line platinum at 0.285 V vs. RHE in 0.1 mol dm3 HClO4 [20].
et al. and Wesselmark et al. have a factor of two difference due to However, studies on single crystals [19] have shown that the PZTC
the method of calculation, which would make their values larger of polycrystalline Pt is an average of the surface facets; Pt(11 0), Pt
than those reported here. This still leaves a large variation in (111) and Pt(1 0 0); in 0.1 mol dm3 HClO4 they are 0.23, 0.34 and
exchange current densities reported at both low (20 C) and high 0.43 V vs. RHE, respectively. With the particle size of 2.4 nm, a
(80 C) temperature, however, with the introductions of higher cuboctahedron particle would contain Pt(1 0 0) and Pt(111) facets;
mass transport techniques, higher exchange current densities are however, this would not account for the decrease in PZTC to 0.24 V
generally being reported. Also, while most results have a vs. RHE. Climent et al. [39,40] showed that an increased fraction of
comparable catalyst as used here (Pt/C), others used different step density to surface facets caused a negative shift in PZTC and
catalysts such as a nanostructured thin film PtCoMn [11] and a therefore these step sites have a lower PZTC (they measured a Pt
sputtered Pt layer catalyst [17] and therefore the activities could be (11 0) step to have a PZTC of 0.15 V vs. RHE). Mayrhofer et al. [20]
substantially different. This still leaves a large variation in applied this to nano-particles, showing that as the particle size
exchange current densities reported at both low (20 C) and high decreased (or ratio of edges to facets increases), the PZTC shifted
(80 C) temperature, however, with the introductions of higher negative in potential. Our PZTC value is close to Mayrhofer et al.’s
mass transport techniques, higher exchange current densities are value reported for a 1 nm Pt particle size under a thin film of Nafion
generally being reported. and in 0.1 mol dm3 HClO4; while both catalyst layers are bound
with Nafion, the use of different electrolyte concentrations make
3.6. Platinum PZTC the results not directly comparable, as discussed above.
Comparing the PZTC to the HOR in Fig. 1, it could be expected
As discussed above, the HOR curve was not affected by anion that the decrease in current density coincides with the adsorption
species below EPeak,low, while above EPeak,low, a large anion affect of anions acting as site blocking species, reducing the active Pt
was observed. The adsorption of anion (and cation) species are surface area (assuming that the rate limiting step is the adsorption
strongly dependent on the charge at the surface of the electrode (Pt of hydrogen and therefore site dependent). However, the measured
in this case). The potential of zero free charge (PZFC) is a commonly PZTC falls between the potentials of the two peaks. This suggests
used potential where the surface of the electrode has a zero total that the two peaks in the HOR are two different sites (edges and
charge, first postulated by Frumkin and Petrii [38], and is generally facets) with different PZTC’s and the measured PZTC corresponds
measured in ultra-high vacuum. In an electrolyte, the ions in the to the ratio of the two respective sites (as for polycrystalline Pt with
solution interact with the surface, and therefore the PZFC cannot its different facets). As the step density has been shown to lower
be measured. Instead, a potential of zero total charge (PZTC) can be the PZTC, this suggests the edge sites would have a PZTC below
measured which is the potential where the total charge at the 0.24 V vs. RHE, and we propose that this is where EPeak,low appears
at 0.2 V vs. RHE. While the facets would be made of Pt(1 0 0) and Pt
(111) facets and in 0.1 mol dm3 HClO4 they have a PZTC of
0.34 and 0.43 V vs. RHE, and we propose this is where Epeak,high sits
at 0.36 V vs. RHE. Therefore, below the PZTC for the respective
surface site, Hads is the dominant surface adsorbed species and the
HOR increases with over-potential. This is supported by the fact
that there is no effect of anion species below EPeak,low in Fig. 3.
Above the PZTC of the respective surface, the onset of competi-
tively adsorbing anions hinders the reaction, causing a current
decay. This implies that the ratio of the peak heights would change
with facets to edge ratio or particle size; the authors are currently
exploring this hypothesis.
While the two peaks in Fig. 1 have been observed on Pt
elsewhere [17] in high mass transport techniques, the decay in
mass transport above EPeak,high has not been widely presented at
such low potentials, typically because the HOR remains under
mass transport limitations until high potentials are exceeded. The
current decay has mainly been studied using the RDE, occurring at
Fig. 9. Current measured from CO displacement on 2.4 nm Pt nanoparticles
supported on carbon in 4 mol dm3 HClO4 at 298 K. CE = Pt wire, RE = RHE. The inset
>0.8 V vs. RHE and rationalised by the formation of oxides.
shows the charge density with respect to potential, with the PZTC at 0.24 0.01 V However, as the current is mass transport limited on the RDE
vs. RHE (intercept of the x-axis). between 0.1–0.8 V vs. RHE, the potential at which the current
774 C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776
Fig.10. Effect of scan limit on two Pt/C electrodes: a) shows the change in anodic limit with a 0.7 mgP cm2 electrode in 0.5 mol dm3 HClO4 and b) shows the change in
cathodic limit with a 1.9 mgPt cm2 electrode in 4 mol dm3 HClO4. Both run with the conditions H2, 298 K, 10 mV s1, CE = Pt wire and RE = RHE.
begins to decay is masked. In fact, the decay in current in our is reversed and only when the potential goes below the PZTC again,
results above 0.8 V vs. RHE in Fig. 1 is very similar to the decay in does the current density start to recover. As the upper potential
current density seen in the RDE once j < jMT (this can be seen limit increases further, the hysteresis increases (i.e. current on the
clearly in the inset of Fig. 9 in [12]). cathodic scan decreases), possibly showing that this anion
A suitable comparison of the HOR can be made with Ru in the adsorption proceeds on the surface as the potential is increased.
RDE. Due to its order of magnitude lower activity, this reaction can This has the effect of blocking the HOR, and remains on the surface
be measured without becoming mass transport limited. In Fig. 2 of until the potential is scanned below the PZTC for the specific site.
[41], a broad peak was observed at 0.2 V vs. RHE and attributed to At higher potentials, this anion adsorption will change to oxide
a kinetic limitation in agreement with the discussion here for Pt. A formation, further perturbing the HOR.
single peak was observed, but as they were using polycrystalline For the cathodic scan limit window opening experiment
Ru, the edge contribution is likely to be minimal. After the peak (Fig. 10b), the scans start at 1.1 V vs. RHE and the lower scan
current density, the current decayed rapidly to almost zero at 0.4 V limit reduces in potential from 1 to 0.1 V vs. RHE. The scans
vs. RHE. The authors assigned this decay to oxide adsorption, with between 1.1 V vs. RHE and EPeak,low all show little hysteresis, with
Ru well known to have an onset of oxide formation at low the last scan showing no hysteresis labelled. There is a slight
potentials which assists in the oxidation of carbonaceous species hysteresis in the high potential region (0.6–0.9 V), and this seems
(e.g. CO). This decay is very similar in shape to the decay observed to be associated with oxide formation and reduction. Therefore, a
for Pt here, where the current density decays 5 fold from consistent behaviour is observed for the site blocking species
500 mA cm2Spec at 0.2 V to <100 mA cm2Spec at 0.8 V in Fig. 1. perturbing the current density down to a potential of 0.2 V.
As described above, on platinum we ascribe this decay to anion However, all of these scans follow the lower cathodic scan,
adsorption (sulfonate chains of the PFSI and perchlorate ions) as suggesting that a percentage of the sites remain irreversibly
oxide formation is not considered to occur until higher potentials, blocked and only when the potential decreases to EPeak,low (or the
and the onset of the current decay seems to be coincident with the PZTC of the edges), do these sites become re-activated. This is
PZTC. For the case of Ru however, specifying whether anion shown by the gradual increase in current density (and hysteresis)
adsorption or oxide species dominate in the decay is challenging, of the anodic scan as the lower potential limit moves from EPeak,low
and is likely to be a mixture of both. to 0.1 V vs. RHE; only upon going below 0 V vs. RHE does the
current density fully recover to jMax. This partial deactivation of the
3.7. Hysteresis of the HOR HOR occurs across the entire potential region of the ORR, and it is
intriguing to consider whether the HOR could be used as a probe
Fig. 10 shows the effect of a) the anodic limit and b) the cathodic for the state of the Pt surface during the ORR.
limit on two window opening scans for the HOR. The precise shape
of the curves in the two window opening experiments is slightly 4. Conclusion
different, due to the effect of electrolyte concentration as discussed
above, with Fig. 10a) and b) run in 0.5 and 4 mol dm3 perchloric Using the floating electrode technique, kinetic parameters have
acid, respectively. The anodic scan limit shows a “humming bird” been extracted from the HOR with no mass transport correction
shape, where the scan starts at 0.01 V vs. RHE for all scans and the factor. This is made possible through a mixture of the high mass
upper limit changes from 0.1 to 1.1 V vs. RHE. The last scan without transport capability and an optimised low loading catalyst layer.
hysteresis (labelled) has an upper potential of 0.22 V vs. RHE for The high mass transport capability has enabled rapid gas transport
this CV and is just before EPeak,low. The next scan, with upper towards the catalyst layer (HOR) or away from the catalyst layer
potential of 0.33 V vs. RHE, displays a hysteresis loop. As discussed (HER). Maximum current densities of 0.6 and 8 A cm2Spec which
in the Platinum PZTC section, after EPeak,low, anion adsorption corresponds to 1200 and 19,000 H2 molecules per surface platinum
occurs on the Pt edge sites, reducing the current density. In atom per second were observed for the HOR and HER, respectively.
addition, the current density carries on decreasing as the potential The low catalyst loading enables homogeneous conditions across
C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776 775
the catalyst layer (removing internal polarisation gradients), [4] J. Maruyama, M. Inaba, K. Katakura, Z. Ogumi, Z.-i. Takehara, Influence of
allowing definition such as the two peaks observed in the HOR. Nafion1 film on the kinetics of anodic hydrogen oxidation, Journal of
Electroanalytical Chemistry 447 (1998) 201–209.
For both perchloric and sulphuric acid, negligible anion effect [5] J. Zhou, Y. Zu, A.J. Bard, Scanning electrochemical microscopy: Part 39. The
was seen for the HOR below EPeak,low. These potentials are below proton/hydrogen mediator system and its application to the study of the
the PZTC where little anions adsorption is expected and therefore electrocatalysis of hydrogen oxidation, Journal of Electroanalytical Chemistry
491 (2000) 22–29.
we conclude the exchange current density can be measured in [6] C.G. Zoski, Scanning Electrochemical Microscopy: Investigation of Hydrogen
either perchloric or sulphuric acid (for a set pH). Above EPeak,low, Oxidation at Polycrystalline Noble Metal Electrodes, The Journal of Physical
a > 5 fold decrease in activity occurred between 0.2 and 0.8 V vs. Chemistry B 107 (2003) 6401–6405.
[7] J.X. Wang, S.R. Brankovic, Y. Zhu, J.C. Hanson, R.R. Adži c, Kinetic
RHE (before the onset of oxide formation), showing anion Characterization of PtRu Fuel Cell Anode Catalysts Made by Spontaneous Pt
adsorption perturbates the reaction. While the Pt/Nafion/ Deposition on Ru Nanoparticles, Journal of The Electrochemical Society 150
perchloric acid environment had a 5 fold decrease, the Pt/ (2003) A1108–A1117.
[8] S. Chen, A. Kucernak, Electrocatalysis under Conditions of High Mass
Nafion/sulphuric acid had a 10 fold decrease, showing the stronger
Transport: Investigation of Hydrogen Oxidation on Single Submicron Pt
adsorbing sulfuric acid has a greater blocking effect. Particles Supported on Carbon, The Journal of Physical Chemistry B 108 (2004)
Our data suggests that proton generation during the HOR 13984–13994.
cannot be ignored as it was found to cause a shift in the local pH [9] C. Song, Y. Tang, J.L. Zhang, J. Zhang, H. Wang, J. Shen, S. McDermid, J. Li, P.
Kozak, PEM fuel cell reaction kinetics in the temperature range of 23–120 C,
towards higher current densities (causing Epeak,low to shift to Electrochimica Acta 52 (2007) 2552–2561.
positive potentials) and the pH of the acid was found to alterthe [10] Y. Sun, J. Lu, L. Zhuang, Rational determination of exchange current density for
exchange current density (in the pH range studied). hydrogen electrode reactions at carbon-supported Pt catalysts, Electrochim.
Acta 55 (2010) 844–850.
We have developed and analytically solved the full Heyrovsky- [11] X. Wang, R.K. Ahluwalia, A.J. Steinbach, Kinetics of Hydrogen Oxidation
Volmer equation and find that it adequately explains our and Hydrogen Evolution Reactions on Nanostructured Thin-Film Platinum
experimental results in that Alloy Catalyst, Journal of The Electrochemical Society 160 (2013)
F251–F261.
[12] C.M. Zalitis, D. Kramer, A.R. Kucernak, Electrocatalytic performance of fuel cell
a) It correctly predicts the effective reaction order for protons and reactions at low catalyst loading and high mass transport, Physical Chemistry
hydrogen under standard conditions; Chemical Physics 15 (2013) 4329–4340.
[13] J. Jiang, A. Kucernak, Investigations of fuel cell reactions at the composite
b) it correctly predicts the reaction order at high over potentials
microelectrode|solid polymer electrolyte interface. I. Hydrogen oxidation at
c) it is capable of replicating electrochemical performance in a the nanostructured Pt|NafionÂ1 membrane interface, J. Electroanal. Chem.
dataset comprising of almost three orders of magnitude of 567 (2004) 123–137.
[14] T.A. Greszler, D. Caulk, P. Sinha, The Impact of Platinum Loading on Oxygen
hydrogen partial pressure; more than an order order of
Transport Resistance, Journal of The Electrochemical Society 159 (2012)
magnitude of hydrogen concentration and a temperature range F831–F840.
from 278–333 K. [15] C.M. Zalitis, D. Kramer, J. Sharman, E. Wright, A.R. Kucernak, Pt Nano-Particle
Performance for PEFC Reactions at Low Catalyst Loading and High Reactant
Mass Transport, ECS Transactions 58 (2013) 39–47.
The linearised form of the model may be of significant relevance [16] K.C. Neyerlin, W. Gu, J. Jorne, H.A. Gasteiger, Study of the Exchange Current
to fuel cells and electrolysers. A further improvement to this work Density for the Hydrogen Oxidation and Evolution Reactions, Journal of The
would be to derive a model to fit the HOR across the entire Electrochemical Society 154 (2007) B631–B635.
[17] M. Wesselmark, B. Wickman, C. Lagergren, G. Lindbergh, Hydrogen oxidation
potential region. To do this, understanding of the anion effects at reaction on thin platinum electrodes in the polymer electrolyte fuel cell,
high potential are needed. Furthermore, this model should be Electrochem. Commun. 12 (2010) 1585–1588.
considered for studying hydrogen oxidation under alkaline [18] Certificate of Analysis, Correspondence with Johnson Matthey Fuel Cells,
16 Dec 2009.
conditions. [19] V. Climent, R. Gomez, J.M. Orts, A. Rodes, A. Aldaz, J.M. Feliu, Electrochemistry,
In addition, we have presented evidence that the two peaks Spectroscopy, and Scanning Tunneling Microscopy Images of Small Single-
correspond to two different surface sites, edges or Pt atoms with Crystal Electrodes, in: A. Wieckowski (Ed.), Interfacial Electrochemistry
Theory, Experimental, and Application, MarcelDekker, Inc., New York, 1999,
low coordination numbers (EPeak,low) and facets (EPeak,high) and that pp. 463.
the experimentally measured PZTC is a contribution of the two. [20] K.J.J. Mayrhofer, B.B. Blizanac, M. Arenz, V.R. Stamenkovic, P.N. Ross, N.M.
Further work is currently been completed with different Pt particle Markovic, The Impact of Geometric and Surface Electronic Properties of Pt-
Catalysts on the Particle Size Effect in Electrocatalysis, The Journal of Physical
sizes (changing the ratio of edges to facets) to explore the change in
Chemistry B 109 (2005) 14433–14440.
the ratio of the peak heights. [21] C.M. Zalitis, A.R. Kucernak, Manuscript in preparation.
[22] K.L. Hsueh, E.R. Gonzalez, S. Srinivasan, Electrolyte effects on oxygen reduction
kinetics at platinum: A rotating ring-disc electrode analysis, Electrochimica
Acknowledgements Acta 28 (1983) 691–697.
[23] R. Subbaraman, D. Strmcnik, V. Stamenkovic, N.M. Markovic, Three Phase
Interfaces at Electrified Metal-Solid Electrolyte Systems 1. Study of the Pt(hkl)-
The authors would like to thank the U.K. Engineering and Nafion Interface, The Journal of Physical Chemistry C 114 (2010) 8414–8422.
Physical Sciences Research Council (EPSRC) for funding through [24] M.S. Rau, M.R. Gennero de Chialvo, A.C. Chialvo, Kinetic study of the hydrogen
Hydrogen to Fuel Cells (H2FC SUPERGEN) Flexible Funding Award oxidation reaction on Pt over the complete overpotential range, Journal of
Power Sources 229 (2013) 210–215.
EPSRC Grant Ref. EP/J016454/1; EP/G030995/1 – Supergen Fuel Cell [25] H. Kita, Y. Gao, T. Nakato, H. Hattori, Effect of hydrogen sulphate ion on the
Consortium – Fuel cells – Powering a Greener Future – CORE; and hydrogen ionization and methanol oxidation reactions on platinum single-
EP/K503733/1 – EPSRC Impact Acceleration Research Grant. crystal electrodes, Journal of Electroanalytical Chemistry 373 (1994)
177–183.
[26] J. Omura, H. Yano, M. Watanabe, H. Uchida, Electrochemical Quartz Crystal
References Microbalance Analysis of the Oxygen Reduction Reaction on Pt-Based Electrodes.
Part 1: Effect of Adsorbed Anions on the Oxygen Reduction Activities of Pt in HF,
[1] V.S. Bagotzky, N.V. Osetrova, Investigations of hydrogen ionization on HClO4, and H2SO4 Solutions, Langmuir 27 (2011) 6464–6470.
platinum with the help of micro-electrodes, Journal of Electroanalytical [27] G.K.H. Wiberg, M. Arenz, Establishing the potential dependent equilibrium
Chemistry and Interfacial Electrochemistry 43 (1973) 233–249. oxide coverage on platinum in alkaline solution and its influence on the
[2] W. Vogel, L. Lundquist, P. Ross, P. Stonehart, Reaction pathways and poisons— oxygen reduction, Journal of Power Sources 217 (2012) 262–267.
II: The rate controlling step for electrochemical oxidation of hydrogen on [28] E.W. Washburn, Electrical Conductivity of Aqueous Solutions, International
Pt in acid and poisoning of the reaction by CO, Electrochim. Acta 20 (1975) Critical Tables of Numerical Data, Physics, Chemistry and Technology (1st
79–93. Electronic Edition), Knovel, 1926–1930.
[3] N.M. Markovi c, B.N. Grgur, P.N. Ross, Temperature-Dependent Hydrogen [29] K. Aoki, A. Baars, A. Jaworski, J. Osteryoung, Chronoamperometry of strong
Electrochemistry on Platinum Low-Index Single-Crystal Surfaces in Acid acids without supporting electrolyte, Journal of Electroanalytical Chemistry
Solutions, The Journal of Physical Chemistry B 101 (1997) 5405–5413. 472 (1999) 1–6.
776 C.M. Zalitis et al. / Electrochimica Acta 176 (2015) 763–776
[30] K. Aoki, A. Tokida, Resistance of solution without supporting electrolyte under [37] J.X. Wang, T.E. Springer, R.R. Adzic, Dual-Pathway Kinetic Equation for the
the reduction of HCl, Electrochimica Acta 45 (2000) 3483–3488. Hydrogen Oxidation Reaction on Pt Electrodes, Journal of The Electrochemical
[31] J. Durst, A. Siebel, C. Simon, F. Hasche, J. Herranz, H.A. Gasteiger, New insights Society 153 (2006) A1732–A1740.
into the electrochemical hydrogen oxidation and evolution reaction [38] A.N. Frumkin, O.A. Petrii, Potentials of zero total and zero free charge of
mechanism, Energy & Environmental Science 7 (2014) 2255–2260. platinum group metals, Electrochimica Acta 20 (1975) 347–359.
[32] W. Sheng, Z. Zhuang, M. Gao, J. Zheng, J.G. Chen, Y. Yan, Correlating hydrogen [39] V.c. Climent, R. Gómez, J.M. Feliu, Effect of increasing amount of steps on the
oxidation and evolution activity on platinum at different pH with measured potential of zero total charge of Pt(111) electrodes, Electrochimica Acta 45
hydrogen binding energy, Nat Commun 6 (2015) . (1999) 629–637.
[33] J.K. Nørskov, T. Bligaard, A. Logadottir, S. Bahn, L.B. Hansen, M. Bollinger, H. [40] V. Climent, G.A. Attard, J.M. Feliu, Potential of zero charge of platinum stepped
Bengaard, B. Hammer, Z. Sljivancanin, M. Mavrikakis, Y. Xu, S. Dahl, C.J.H. surfaces: a combined approach of CO charge displacement and N2O reduction,
Jacobsen, Universality in Heterogeneous Catalysis, Journal of Catalysis 209 Journal of Electroanalytical Chemistry 532 (2002) 67–74.
(2002) 275–278. [41] H.A. Gasteiger, N.M. Markovic, P.N. Ross, H2 and CO Electrooxidation on Well-
[34] E. Gileadi, S.D. Argade, J.O.M. Bockris, The Potential of Zero Charge of Platinum Characterized Pt, Ru, and Pt-Ru. 1. Rotating Disk Electrode Studies of the Pure
and Its pH Dependence, The Journal of Physical Chemistry 70 (1966) 2044– Gases Including Temperature Effects, The Journal of Physical Chemistry 99
2046. (1995) 8290–8301.
[35] M.R. Gennero de Chialvo, A.C. Chialvo, Hydrogen diffusion effects on the [42] G.J. Janz, B.G. Oliver, G.R. Lakshminarayanan, G.E. Mayer, Electrical
kinetics of the hydrogen electrode reaction. Part I. Theoretical aspects, Physical conductance, diffusion, viscosity, and density of sodium nitrate, sodium
Chemistry Chemical Physics 6 (2004) 4009–4017. perchlorate, and sodium thiocyanate in concentrated aqueous solutions, The
[36] P.M. Quaino, J.L. Fernández, M.R. Gennero de Chialvo, A.C. Chialvo, Hydrogen Journal of Physical Chemistry 74 (1970) 1285–1289.
oxidation reaction on microelectrodes: Analysis of the contribution of the [43] H.A. Gasteiger, J.E. Panels, S.G. Yan, Dependence of PEM fuel cell performance
kinetic routes, Journal of Molecular Catalysis A: Chemical 252 (2006) 156–162. on catalyst loading, Journal of Power Sources 127 (2004) 162–171.