Sunteți pe pagina 1din 8

Electrochimica Acta 50 (2005) 5594–5601

Kinetics of the hydrogen evolution reaction on Fe–Mo film


deposited on mild steel support in alkaline solution
N.R. Elezović a , V.D. Jović a , N.V. Krstajić b,∗
a Center for Multidisciplinary Studies, University of Belgrade, 11030 Belgrade, Serbia and Montenegro
b Faculty of Technology and Metallurgy, University of Belgrade, 11000 Belgrade, Serbia and Montenegro

Received 12 November 2004; received in revised form 21 January 2005; accepted 5 March 2005
Available online 17 May 2005

Abstract

The mechanism and kinetics of the hydrogen evolution reaction were studied in 1.0 mol dm−3 NaOH solution on Fe–Mo alloy elec-
trodes prepared by electrodeposition at constant current densities from a pyrophosphate bath. A series of electrode containing 34–59 at.%
Mo was prepared. Electrodes displayed porous character, and electrochemical impedance spectroscopy was used to characterized real
surface area. It was found that within the whole potential region the mechanism of the HER is a consecutive combination of the
Volmer step, followed dominantly by a rate controlling Heyrovsky step, while the contribution of the parallel Tafel step is negligible.
The kinetic parameters of the HER were determined. With an increase in the molybdenum content, the electrodes become more ac-
tive, and an increase in the real surface area is observed. The main factor influencing the electrode activity seems to be the real surface
area.
© 2005 Elsevier Ltd. All rights reserved.

Keywords: Hydrogen evolution; Fe–Mo electrodes; Alkaline solution; Impedance; Mechanism

1. Introduction substitution by more active catalysts based on combination


of Co or Ni with Mo up to now is not successful. The main
The hydrogen evolution reaction (HER) is of technologi- reason is probably the fact that Co or Ni ions present in the
cal importance for such processes as water electrolysis, chlo- solution at very low concentration as corrosion product cat-
rate and chlorine production. For these applications materials alytically degrade active chlorine (ClO− , HClO) which lead
with low overpotential are needed. to substantial current losses [13].
Miles [1] suggested that a combination of two metals from In this study, we report the results regarding the HER on
the two brunches of volcano curve could result in enhanced some Fe–Mo coatings prepared by the electrochemical de-
catalytic activity. Jakšić [2] showed that a combination of Ni position from suitable pyrophosphate bath [14] on mild steel
or Co with Mo could result in a substantial enhancement of substrate. The main purpose is to study the mechanism and
the HER. kinetics in order to determine surface coverage by adsorbed
Numerous studies thoroughly investigated the kinetics and hydrogen and to understand the source of the electrode ac-
mechanism of the HER at Ni–Mo [3–12] alloys of various tivity.
compositions, obtained by suitable baths. All these electrodes
are characterized by having a high roughness factor.
However, despite extensive studies, in the case of chlorate 2. Theory
cell process, mild steel is a traditional cathode material and its
Generally, the mechanism of the HER in aqueous acid
∗ Corresponding author. solutions is treated as a combination of three basic steps, two
E-mail address: nedeljko@elab.tmf.bg.ac.yu (N.V. Krstajić). electrochemical and one chemical:

0013-4686/$ – see front matter © 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2005.03.037
N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601 5595

k backward reactions of steps 2 and 3 could be neglected be-


H3 O + + M + e ⇔1
MHads + H2 O (1)
k−1 cause they have no significant influence on the reaction rates
of the second (Eq. (4b)) and third step (Eq. (4c)) at overpo-
followed by
tential η ≤ −60 mV, and the steady state coverage (ΘH ), of
k the intermediate Hads species is the following function of the
MHads + H3 O+ + e ⇔2
M + H 2 + H2 O (2)
k−2 corresponding rate constants:

and/or
(k1 + k−1
 + k ) +
2 (k1 + k−1
 +k  )2 + 8k  k
2 1 3
k ΘH = − (8)
2MHads ⇔3
2M + H2 (3) 4k3
k−3
and the pseudocapacitance, C␾ is given by
The rates of corresponding steps are given by
  q1 dΘH
β1 Fη C = (9)
v1 = k1 (1 − ΘH )exp − k−1 ΘH dη
RT
  The presence of the electrochemical rate constants in Eq. (8)

× exp (1 − β1 ) = k1 (1 − ΘH ) − k−1

ΘH (4a) clearly indicates the rather complex dependence of ΘH on
RT
the electrode potential.
  Theoretically, six variables, i.e. four independent chemi-
β2 Fη cal rate constants of the three basic steps and two symmetry
v2 = k2 ΘH exp − − k−2 (1 − ΘH )
RT factors of the electrochemical steps (β1 and β2) can describe
  the mechanism of the HER on the electrode in the corre-

× exp (1 − β2 ) = k2 ΘH − k−2

(1 − ΘH ) (4b) sponding solution. However, it can be reasonably assumed
RT
for elementary electrode reactions, that β1 = β2 = 0.5. With
this assumption the problem is reduced to the determination
v3 = k3 ΘH
2
− k−3 (1 − ΘH )2 (4c) of four independent variables, i.e. k1 , k−1 , k2 , k3 .
The Faradaic impedance of the HER is described by the
In these rate laws surface concentrations of the interme- following equation [15]:
diate Hads and the corresponding free adsorption sites at  −1
the electrode are given as the coverage (ΘH ), or surface
1 1 1
sites concentrations (1 − ΘH ). The chemical rate constants = +  (10)
ki (mol cm−2 s−1 ) have nominally been written to include Z̃f Rct 1
+ jωCp
Rp
H3 O+ ion concentration, in the 0.5 mol dm−3 HClO4 solu-
tion. The equation is written directly from Armstrong’s equivalent
Under open circuit conditions all three reaction steps are circuit [16].
in equilibrium with net reaction rates equal to zero so that The equivalent circuit elements: Rct , Rp , Cp are complex
functions of the kinetic parameters and are given [17] by
v1 = v2 = v3 = 0 (5)    
−1 ∂v1 ∂v2
In this case and according to Eq. (5) only four out of six rate Rct = F +
∂E ΘH ∂E ΘH
constants are independent parameters, i.e. k1 , k2 , k−1 , and k3 ,  2
while k−2 = (k1 k2 )/k−1 , and k−3 = k12 k3 /k−1
2 . βF
= [k1 (1 − ΘH ) − k−1

ΘH + k2 ΘH ] (11a)
The steady-state kinetics of the HER at constant current RT
density are characterized by the conditions of the charge bal-
ance (with rates in mol cm−2 s−1 ): Generally, it is reasonable to assume that more than one
j of the basic steps is involved in the mechanism of the HER,
r0 = = −(v1 + v2 ) (6) which means that the value of the experimental Tafel slope
F
is determined by the reciprocal of the Faradaic resistance
and the mass balance with respect to the intermediate Hads : (i.e. the sum of charge transfer and pseudo resistances)
q1 dΘH (Rct + Rp )−1 , where the pseudo resistance is presented by
r1 = = v1 − v2 − 2v3 (7)
F dt
R20
q1 is the charge necessary for a monolayer coverage by ad- Rp = − (11b)
Rct + R0
sorbed hydrogen.
When a particular value of the mass balance, i.e. r1 = 0 and the parameter R0 is defined as follows:
is set, the steady-state coverage can be calculated, assuming     
F 2 τp ∂v1 ∂v2
that the Langmuir adsorption isotherm is operable. Calcu- R−1 = + (11c)
lations showed that the rate constants k−2 and k−3 of the
0
q1 ∂ΘH E ∂ΘH E
5596 N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601

ethanol for 5 min, then etched in 25 wt.% HCl for 2 min. Af-
ter this procedure samples were washed with distilled water,
dried and weighted and then immersed in the solution for
Fe–Mo alloy electrodeposition. Fe–Mo film was deposited
on one side of the electrode. The inactive walls of the sup-
port were protected from the solution by the alkaline resistant
epoxy resin. After deposition samples were washed, dried
and weighted again to determine the mass of the alloy. All
solutions were made using distilled-deionized water and an-
alytical grade chemicals.
Fe–Mo alloys were deposited to a constant charge of
36 C cm−2 at three different current densities: 100 mA cm−2
(Fe–Mo100), 50 mA cm−2 (Fe–Mo50) and 20 mA cm−2
Fig. 1. CPE equivalent circuit used for the HER.
(Fe–Mo20), from the plating bath with the following compo-
sition: 9 g dm−3 of FeCl3 , 40 g dm−3 Na2 MoO4 , 75 g dm−3
with the time constant τ p equal to NaHCO3 and 45 g dm−3 Na4 P2 O7 at 60 ◦ C. A Pt mesh,
        placed parallel to the cathode, was used as a counter elec-
F ∂v3 ∂v2 ∂v1 trode during electrodeposition and electrolyte was moder-
τp−1 = 2 + − (11d)
q1 ∂ΘH E ∂ΘH E ∂ΘH E ately stirred with the magnetic stirrer.
The counter electrode was a platinum sheet of 5 cm2 geo-
However, on solid electrodes, the double-layer capaci-
metric area.
tance is substituted by a constant phase element (CPE) its
The reference electrode was the Hg|HgO electrode in
impedance is given as [18]:
1.0 mol dm−3 NaOH, held at a constant temperature of 25 ◦ C.
1 All potentials are referred to the standard hydrogen electrode
ZCPE = (12)
T (jω)α scale (SHE). The calculated value of the equilibrium poten-
tial of the HER in 1.0 mol dm−3 solution of NaOH (pH 13.86)
where T is capacitance parameter (in F cm−2 sα−1 ) and α ≤ 1 at 25.0 ◦ C is −0.818 V.
is a parameter characterizing rotation of the complex plane
impedance plot. The double-layer capacitance may be esti- 3.3. Measurements
mated from [18]:
1−α Tafel lines were recorded using potentiostatic steady-state
α
T = Cdl (R−1 −1
 + Rct ) (13)
voltammetry, point by point at 60 s intervals, in the range of
This is the so-called CPE model and it is presented in potential from −1.12 to −0.82 V, using a PAR 273 potentio-
Fig. 1. It predicts the formation of two depressed semicircles stat, with good reproducibility of measurement.
on the complex–plane plots. Whenever the potential of the WE approached approxi-
mately −1.1 V (or when current densities were close to, or
above approximately 0.1 A cm−2 ) it was found that the un-
3. Experimental compensated solution resistance was significant. Therefore,
the IR drop was systematically determined in all measure-
3.1. Cell and chemicals ments, using ac impedance methods. All data presented in
this article are corrected for the IR drop.
A conventional three-compartment cell was used. The Simultaneously with the Tafel lines, electrochemical
working electrode (WE) compartment was separated by frit- impedance spectra of the WE at selected constant potentials
ted glass discs from the other two compartments. The WE were determined, using a PAR 273 potentiostat, together with
compartment was jacketed and thermostated during measure- a PAR 5301 lock-in-amplifier, controlled through a GPBI
ments at 25.0 ◦ C using an ultrathermostat. All measurements PC2A interface. The fast Fourier transformation (FFT) tech-
were performed in 1.0 mol dm−3 solution of NaOH (Spec- nique was used in the impedance measurements in the fre-
trograde, Merck), prepared in deionized water. quency region below 5 Hz. So, both impedance spectra in
The WE compartment was saturated with purified hydro- the complex plane and corresponding Bode diagrams were
gen at standard pressure during measurements. obtained in the frequency range from 50 mHz to 100 kHz,
at the constant potentials of the WE. In all measurements
3.2. Electrodes above 5 Hz, ten frequency points per decade were taken. The
real and imaginary components of the impedance spectra in
All samples were deposited on mild steel substrates having the complex plane were analyzed using the nonlinear least
a surface area of 1 cm2 . The steel substrates were first sand squares (NLSs) fitting program to determine the parameters
blasted using 50 ␮m particles, degreased in NaOH-saturated of the proposed equivalent circuit. The rate constants were
N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601 5597

calculated by minimizing the residual (S) of the sum of each Table 1


experimental datum (the dc polarization measurements and The influence of the current density for deposition of Fe–Mo coatings, jdep ,
on current efficiency, ηi , and alloy composition
ac impedance results).
The real electrode surface area was evaluated from in situ Electrode jdep Mo (at.%) Fe (at.%) Current efficiency,
(mA cm−2 ) ηi (%)
measurements of double-layer capacitance, determined from
initial potential–decay slopes (dη/dt) following interruption Fe–Mo20 20 34.0 66.0 36.8
Fe–Mo50 50 41.8 58.2 32.0
of significant, overpotential (η)-dependent, Faradaic currents Fe–Mo100 100 59.3 40.7 29.8
passing across the electrode/electrolyte interface. Potential-
relaxation transients were recorded digitally by computer
over five to six decades of time using a very fast electronic Table 2
switcher. Kinetic parameters for the HER obtained from the polarization curves in
1.0 mol dm−3 NaOH at 25 ◦ C
Scanning electron microscopy (SEM-JOEL 840) was
used to characterize the as-deposited surfaces and en- Electrode Fe–Mo20 Fe–Mo50 Fe–Mo100
ergy dispersive X-ray spectroscopy (EDS) to determine b1 (mV) 36 35 37
alloy composition. Selected deposits were mounted in b2 (mV) 127 124 125
cross-section, polished and examined by optical microscopy. j0 (× 106 A cm−2 ) 1.8 4.2 20.4
j (× 103 A cm−2 ) (η = −0.2 V) 14.6 32.7 59.3

4. Results and discussion current densities (examined with the naked eye) seemed to
be less rough.
4.1. Composition and morphology of the Fe–Mo alloys
4.2. Polarization measurements
The influence of the current density (potential) on current
efficiency and alloy composition is presented in Table 1. As The polarization curves obtained on different Fe–Mo
can be seen the content of Mo increases, while the content of electrodes are shown in Fig. 3, and Table 2 presents
Fe decreases with increasing average current density. the kinetics parameters j0 and Tafel slopes, b, for those
In Fig. 2 are shown typical top view of Fe–Mo alloy (a) electrodes. Polarization curves were recorded after holding
and a cross section of the deposit of a thickness of about the electrodes at a constant cathodic current density of
20 ␮m (b). As can be seen morphology of Fe–Mo alloys is 100 mA cm−2 for about 1 h. The electrode containing
characterized by the presence of micro-cracks, with some 59.3 at.% molybdenum (Fe–Mo100) is found to have the
of them being up to 2 ␮m wide (Fig. 2(b)). It is important lowest hydrogen overpotential, and the highest exchange
to note that the surface of the samples deposited at lower current density, j0 ≈ 2 × 10−5 A cm−2 . The polarization

Fig. 2. (a) Scanning electron micrographs of the Fe–Mo electrode surface (1) Fe–Mo100 (59.3% Mo); 2) Fe–Mo50 (41.8% Mo); (3) Fe–Mo20 (34% Mo).
Magnification 1000×. (b) Cross-section of the Fe–Mo100 deposit of a thickness of about 20 ␮m.
5598 N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601

Fig. 4. Impedance spectra in the complex plane for the HER on Fe–Mo100
electrode at four constant overpotential within the Tafel region of the po-
larization curve. Circled points are experimental data and solid lines are
calculated using the NLS fitting procedure.
Fig. 3. Tafel polarization curves for the HER on Fe–Mo electrodes in
1.0 mol dm−3 NaOH at 25 ◦ C.

curves are characterized by two Tafel slopes at all electrodes,


≈−40 mV at lower overpotential region and ≈−120 mV at
high overpotential region.

4.3. ac Impedance

Typical complex plane plots for an electrode containing


59.3 at.% Mo (Fe–Mo100) are shown in Fig. 4 and Bode
plots in Fig. 5. Two overlapped semicircles was found for
all Fe–Mo electrodes at all overpotentials. Data presented in
Figs. 4 and 5 are interpreted by NLS fitting procedure [19]
to determine the elements of the equivalent circuit, given in
Fig. 1. The values of all parameters obtained by this proce-
Fig. 5. Bode plot for the HER on Fe–Mo electrodes in 1.0 mol dm−3 NaOH
dure, which are necessary to calculate the rate constants in the at 25 ◦ C. Solid lines are calculated using the NLS fitting procedure using the
mechanism of the HER are presented in Table 3. The ohmic corresponding values for q1 , assigned in figure.
resistance of the solution was R = 0.69 cm2 .
low and Eq. (13) cannot be used to estimate the average Cdl .
4.4. Surface roughness For determination of double-layer capacity the open circuit
potential decay method was used. In this technique, poten-
Using the NLS fit of the experimental impedance data to tial decay is monitoring after the opening of the circuit. The
the CPE model parameters characterizing the double-layer overpotential–logarithm of time curve for Fe–Mo100 elec-
may be obtained. From the parameters T and α it is possible trode is presented in Fig. 6. It is evident that the electrode does
to determine an average double-layer capacitance. However, not relax to the equilibrium potential even after 10 seconds
the value of the parameter α, which is close to 0.5 is too indicating the presence of a large pseudocapacitance. Open

Table 3
Values of the approximated elements of the CPE equivalent circuit for the HER at Fe–Mo100 in 1.0 mol dm−3 NaOH at 25.0 ◦ C, taken at the constant
overpotentials
−η (V) Rct ( cm2 ) Error (%) Rp ( cm2 ) Error (%) Cp (F cm−2 ) Error (%) T (F cm−2 s␣−1 ) Error (%) α Error (%)
0.055 2.04 5.2 56.3 3.9 0.046 8.1 0.13 8.5 0.54 3.5
0.064 2.03 4.8 35.9 3.5 0.048 6.6 0.14 8.1 0.54 3.2
0.074 1.33 4.6 12.8 4.9 0.070 6.6 0.19 6.7 0.49 2.1
0.087 1.31 5.5 7.67 4.8 0.080 6.2 0.16 5.9 0.49 1.9
0.120 1.05 5.3 2.64 5.6 0.12 6.1 0.16 4.9 0.53 1.7
0.128 0.99 6.5 2.14 5.7 0.13 5.9 0.16 4.6 0.53 1.6
0.150 0.86 6.1 1.32 7.1 0.14 5.7 0.16 4.4 0.52 1.5
0.164 0.78 6.5 0.95 7.0 0.16 5.6 0.15 4.3 0.43 1.4
N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601 5599

Fig. 6. Overpotential, η, against log time, t plot for the HER at Fe–Mo100
electrode recordered after interruption of polarization at overpotential Fig. 7. η against (−dη/dt) plots for the HER at Fe–Mo100 electrode obtained
η = −0.44 V in 1.0 mol dm−3 NaOH at 25 ◦ C. by the numerical differentiation of initial part of η–t curve presented in this
figure.
circuit potential decay can be can be presented by the follow-
ing equation [20]:
j
C + Cdl = dη
(14)
dt

where C is an pseudocapacitance and j is the measured


steady-state current density at the overvoltage η. In order to
analyze this equation the derivative dη/dt must be determined
from the experimental η–t curves. Corresponding η–dη/dt
curves obtained after numerical differentiation is shown in
Fig. 7. At more negative overpotential region, where the pseu-
docapacitance is negligible, the double-layer capacitance can
be determined. The corresponding C–η curves for Fe–Mo
electrodes obtained using Eq. (14) are presented in Fig. 8.
The double-layer capacities Cdl , increase with an increase in
the molybdenum content (Table 4).
From the double-layer capacitances the real electrode sur- Fig. 8. The capacitance, C, against overpotential plots for the HER at Fe–Mo
electrodes in 1.0 mol dm−3 NaOH at 25 ◦ C.
face area and the surface roughness (R) can be estimated, if the
double-layer capacitance for a smooth surface is known. Tak- tained are included in Table 4. They indicate that the surface
ing into account the value of ca. 20 ␮F cm−2 used earlier in roughness increases with increase of molybdenum content in
the literature [21,22], the values of the surface roughness ob- the alloy.
However, one should keep in mind that if the high fre-
Table 4 quency data produced a distorted semicircle in complex plane
Double-layer capacitance values, surface roughness, and intrinsic activities impedance plot with the parameter α ∼ 0.5, the short time po-
for Fe–Mo electrodes in 1.0 mol dm−3 NaOH at 25 ◦ C
tential decay curves (presented in Fig. 8) are also affected by
Electrode Cdl (mF cm−2 ) R j0 R−1 (×10−8 A cm−2 ) this problem, because there is an equivalence between time
Fe–Mo20 2.5 125 1.4 and frequency data, which means that it is not possible ac-
Fe–Mo50 7 350 1.2 curately to determine the double-layer capacitance of these
Fe–Mo100 11 550 3.7
electrodes even with the open circuit potential decay method.

Table 5
Values of the rate constants (in mol cm−2 s−1 ) for individual steps of the HER on Fe–Mo electrodes in 1.0 mol dm−3 NaOH at 25 ◦ C
Electrode k1 k−1 k2 k−2 a k3 k−3 a
Fe–Mo20 (34 at.% Mo) 1.5 × 10−8 1.0 × 10−6 2.1 × 10−9 3 × 10−11 1 × 10−8 2.3 × 10−12
Fe–Mo50 (41.8 at.% Mo) 3 × 10−8 2 × 10−6 3 × 10−9 4.5 × 10−11 3 × 10−8 6.8 × 10−12
Fe–Mo100 (59.2 at.% Mo) 6 × 10−8 5 × 10−6 6 × 10−9 5 × 10−11 6 × 10−8 4.4 × 10−12
a k−2 = (k1 k2 )/k−1 ; k−3 = k12 k3 /k−1
2 .
5600 N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601

Fig. 10. Simulated Θ–η curves, numerically calculated for the set of rate
Fig. 9. Experimental (circled points) and simulated of: (1) Tafel plot; (2) η constants for the HER at Fe–Mo electrodes in 1.0 mol dm−3 NaOH solution
vs. log [1/(Rct + Rp )]; (3) η vs. log (1/Rp ); (4) η vs. log (1/Rct ) for the HER at 25 ◦ C.
at Fe–Mo100 in 1.0 mol dm−3 NaOH solution.

Fig. 11. Overpotential dependence of the theoretically calculated individual rates for the Volmer, Heyrovsky and Tafel steps occurring simultaneously with the
HER on Fe–Mo electrodes in 1.0 mol dm−3 NaOH solution at 25 ◦ C. Theoretical polarization curve, obtained by summing individual curves is represented by
full lines. Experimental data are presented by circled points.
N.R. Elezović et al. / Electrochimica Acta 50 (2005) 5594–5601 5601

4.5. Rate constants kinetics of the electrodeposited Fe–Mo alloy electrodes. The
surface roughness estimated using the double-layer capac-
The rates constants of the individual steps for the HER itance ratio shows that real surface area increases with in-
was determined from polarization measurements and fitted crease in the molybdenum content. This factor is principally
parameters values of CPE equivalent circuit using NLS fitting responsible for the increased activity.
procedure [19]. The rate constants of the forward and backward re-
The estimated values of the rate constants are presented actions of Volmer, Heyrovsky and Tafel steps were esti-
in the Table 5, for the apparent surface area of 1 cm2 . mated by a non-linear fitting method of the experimen-
The complex–plane impedance diagrams calculated from tal results obtained from polarization and impedance mea-
the evaluated k1 values are presented by the solid lines in surements. The HER proceeds through Volmer–Heyrovsky
Figs. 4 and 5. mechanism and reaction rate is controlled by Heyrovsky
Conway and coworkers [17] have shown that at suffi- step.
ciently high η when ΘH is almost constant and for Lang-
muiran adsorption, the theoretical separation between η–j
and the simulated η–log (Rct + Rp )−1 plots (curves 1 and 2 Acknowledgement
in Fig. 9) is equal to log (βF/RT). With β = 0.5 and T = 298 K,
log (βF/RT) = 1.29, which is close to the observed separation This paper has been supported by the Ministry of Sci-
between the two lines of 1.31. This analysis proves the Lang- ence and Environmental Protection, Republic of Serbia, un-
muiran adsorption of H for the HER at Fe–Mo electrodes. der Contract No. 1825.
Using the values of the rate constants from Table 5, and
Eq. (8) it is possible to calculate the dependence of ΘH on the
overpotential of the Fe–Mo electrodes, which are presented References
in Fig. 10. It is interesting to note, that this dependence is
almost the same for all Fe–Mo electrode. [1] M.H. Miles, J. Electroanal. Chem. 29 (1984) 1539.
[2] M.M. Jakšić, Electrochim. Acta 29 (1984) 1539.
The calculated values of the maximum charge of the WE,
[3] B.E. Conway, L. Bai, M.A. Sattar, J. Hydrogen Energy 12 (1987)
q1 , obtained by fitting of Bode plots (Fig. 5) are close to 607.
15 mC cm−2 for Fe–Mo100, 5 mC cm−2 for Fe–Mo50 and [4] I.A. Raj, K.I. Vasu, J. Appl. Electrochem. 22 (1992) 471.
3 mC cm−2 for Fe–Mo20 electrodes. [5] B.E. Conway, L. Bay, D.F. Tessier, J. Electroanal. Chem. 161 (1984)
The reaction rates of the Volmer, Heyrovsky and Tafel 39.
[6] C. Fan, D.L. Piron, P. Paradis, Electrochim. Acta 39 (1994)
steps in the mechanism of the HER and the resulting total
2715.
current value are all independently shown in Fig. 11, as calcu- [7] B.E. Conway, L. Bai, J. Chem. Soc., Farady Trans. 81 (1985)
lated from the corresponding ki values, in a wide overpotential 1841.
region. In this figure, all the rate values, vi (mol cm−2 s−1 ) [8] I.A. Raj, V.K. Venkatesan, Int. J. Hydrogen Energy 12 (1988)
are shown in terms of current density, Fvi (A cm−2 ). 215.
[9] C. Fan, D.L. Piron, A. Sleb, P. Paradis, J. Electrochem. Soc. 141
From Fig. 11, one can conclude that the reaction
(1994) 382.
mechanism is a consecutive combination of Volmer and [10] J.D. Schmotz, J. Balej, J. Appl. Electrochem. 19 (1989) 519.
Heyrovsky step and that Heyrovsky step prevails over Tafel [11] B.E. Conway, L. Bay, M.A. Sattar, Int. J. Hydrogen Energy 12 (1987)
step in low overpotential region at all investigated Fe–Mo 607.
electrodes. Reaction rate is controlled by the Heyrovsky [12] A. Lasia, A. Rami, J. Electroanal. Chem. 294 (1990) 123.
[13] H. Yoshida, T. Akazawa, T. Haneda, Denki Kagaku 41 (1973)
reaction because of the much smaller k2 value as compared
68.
with the k1 and k3 values. [14] L.O. Case, A. Krohn, J. Electrochem. Soc. 105 (1958) 512.
The intrinsic activity, that is the rate constant of the [15] D.A. Harrington, B.E. Conway, Electrochim. Acta 32 (1987)
limiting step per real surface area or the exchange current 1703.
density per real surface are, k2 /R, or j0 /R, was estimated [16] R.D. Armstrong, M. Henderson, J. Electroanal. Chem. 39 (1972)
81.
(see Table 4). Intrinsic activities of Fe–Mo20 (34 at.% Mo)
[17] L. Bai, D.A. Harrington, B.E. Conway, Electrochim. Acta 32 (1987)
and Fe–Mo50 (41.8 at.% Mo) are similar and smaller than 1713.
Fe–Mo100 (59.3 at.% Mo). [18] G.J. Brug, A.L.G. Van Der Eeden, M. Sluyters-Rehbach, J.H.
Sluyters, J. Electroanal. Chem. 176 (1984) 275.
[19] N. Krstajić, M. Popović, B. Grgur, M. Vojnović, D. Šepa, J.
Eletroanal. Chem. 512 (2001) 16.
5. Conclusions
[20] R. Simpraga, L. Bai, B.E. Conway, J. Appl. Electrochem. 25 (1995)
628.
Polarization measurements and the ac impedance mea- [21] A. Lasia, A. Rami, J. Appl. Electrochem. 22 (1992) 376.
surements were used to determine the mechanism and the [22] L. Chen, A. Lasia, J. Electrochem. Soc. 138 (1991) 3321.

S-ar putea să vă placă și