ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img

Full Model for the Two-Step Polarization Curves of Hydrogen Evolution, Measured on RDEs in Dilute Acid Solutions

  • María de Jesús Gálvez-Vázquez
    María de Jesús Gálvez-Vázquez
    Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
  • Vitali Grozovski
    Vitali Grozovski
    Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
  • Noémi Kovács
    Noémi Kovács
    Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    Department of Physical Chemistry, Eötvös Loránd University of Budapest, Pázmány Péter sétány 1/A, H-1117 Budapest, Hungary
  • Peter Broekmann*
    Peter Broekmann
    Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    *E-mail: [email protected]. Tel: +41316314317 (P.B.).
  • , and 
  • Soma Vesztergom*
    Soma Vesztergom
    Department of Physical Chemistry, Eötvös Loránd University of Budapest, Pázmány Péter sétány 1/A, H-1117 Budapest, Hungary
    *E-mail: [email protected]. Tel: +36204612429 (S.V.).
Cite this: J. Phys. Chem. C 2020, 124, 7, 3988–4000
Publication Date (Web):January 18, 2020
https://doi.org/10.1021/acs.jpcc.9b11337

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

  • Open Access

Article Views

3947

Altmetric

-

Citations

LEARN ABOUT THESE METRICS
PDF (3 MB)

Abstract

Polarization curves of the hydrogen evolution reaction (HER), recorded on rotating disk electrodes (RDEs) in mildly acidic solutions, usually show a “two step” behavior. That is, two exponentially rising segments (the first commonly assigned to H+, the second to water reduction) are separated by a limiting current plateau. Here, we devise an analytical model for the full polarization curve by assuming that HER proceeds according to a quasireversible two-electron reaction, H+ + H2O + 2e ⇌ H2 + OH, obeying the Erdey-Grúz–Volmer–Butler equation. Our model is able to reproduce the two step behavior of polarization curves and can also be used for the fitting of measured currents over a broad range of pH, rotation rate, and electrode potential, on both Au and on Pt. We show that the length of the limiting current plateaus measured on RDEs for HER is inversely related to the electrocatalytic activity of the electrode and that at a given rotation rate a linear relationship exists between the plateau length and the bulk solution pH. By analyzing this relationship, we can estimate kinetic parameters, even in cases where the transport performance of the RDE would otherwise not be sufficient to measure well-defined kinetic currents at low overpotentials.

Introduction

ARTICLE SECTIONS
Jump To

The electrochemical hydrogen evolution reaction (HER) is regarded as a straightforward way of transferring electrical energy to a chemical one, enabling the storage of electricity gained from renewable sources like hydro and solar plants. (1) The development of efficient catalysts for HER has thus become a subject of intensive research. Hydrogen evolution gains, however, further importance, as HER is an almost inevitable side reaction of cathodic electrode processes occurring in aqueous environments. For example, in the electroreduction of CO2 (2,3) or the deposition of base metals, (4−6) HER often appears as a parasitic reaction.
It is usually claimed (7) that in acidic solutions the overall hydrogen evolution reaction can be described as
(R1)
while in neutral or alkaline media the reaction is written as
(R2)
The exact mechanism of the above reactions, including the identification of the rate-determining step and the pH dependency, is still a matter of debate. While according to Reaction R2, HER can also occur, at a moderate rate, in solutions that are neutral or even alkaline, when hydrogen production is the primary goal of the electrode process, usually acidic conditions are applied.
Under acidic conditions, Reaction R1 is known to proceed quickly on certain transition metals (e.g., on platinum (8−14)) and less quickly on others (e.g., on gold (15−20)). The catalytic performance of these metals can be explained by Sabatier’s principle; (21) that is, the catalytic rate follows a volcano trend with the binding energy of the metal and hydrogen.
To obtain meaningful kinetics in acidic solutions, it is essential to compensate for the effect of mass transport, as it can have a decisive role in the observed current/overpotential characteristics, especially for kinetically facile reactions. (21) It is usually assumed (11,20) that experimentally measured currents for HER can be described, at least at certain overpotential intervals, by the Erdey-Grúz–Volmer–Butler equation:
(1)
While current/potential curves recorded at high transport rates on a gold rotating disk electrode (RDE) can be made subject to analysis based on eq 1, (20) on more facile catalysts like Pt, no kinetic currents can be determined due to the presence of severe diffusion hindrance even at vigorous stirring. As noted by Zheng et al. (14) and by the group of Gasteiger, (12,13) on a Pt surface, the rate of charge transfer becomes so fast (with respect to that of mass transfer) that essentially a thermodynamic equilibrium is established.
The pH dependency of the rate of HER is often related to a reactant switching from H+ to water molecule (from Reaction R1 to R2). (21−27) This reactant switching can be observed, for example, on the polarization curves of HER recorded at rotating disk electrodes (28) immersed into mildly acidic solutions, exhibiting a “two step” behavior, as shown by Figure 1.

Figure 1

Figure 1. Polarization curve of an RDE, showing two hydrogen evolution steps with an intermittent limiting current plateau section.

Figure 1 demonstrates that in mildly acidic conditions HER is kinetically controlled at low cathodic overpotentials, yielding cathodic currents that rise exponentially with the applied cathodic overpotential (“first kinetic control section”). At higher cathodic overpotentials, the mass transport of H+ becomes rate determining and a limiting current plateau is attained. Cathodic currents higher than the limiting current of H+ reduction can only be achieved by applying extremely negative potentials (see the “second kinetic control section” in Figure 1). At such potentials, it is usually assumed that apart from H+ reduction (Reaction R1), also the reduction of water molecules (Reaction R2) contributes to HER; thus, a further exponential increase of the cathodic current can be seen, following the limiting plateau section.
The described two step behavior of HER polarization curves was first noticed as early as 1956 by Nagel and Wendler (29) and it was studied more recently by the groups of Tobias, (22) Mayrhofer, (23,24) Bruckenstein, (30) Arenz, (31) Pereira, (32) and the present authors. (27) Finding an adequate model to describe the shape of the two step polarization curves is, however, difficult due to complications arising from an interplay of the mass transport of the diffusing species (H+ and OH) and a bulk chemical reaction, the autoprotolysis of water:
(R3)
To write proper kinetic equations for HER, Reactions R1R3 all have to be taken into account, in a scheme suggested by Figure 2a. This was attempted before by Hessami and Tobias, (22) who used the equidiffusivity approximation (i.e., the assumption that the diffusion coefficients of H+ and OH ions are equal in the solution) to solve combined reaction–diffusion–convection equations to obtain pH profiles. The approach of Mayrhofer et al. (23,24) was based on a different, finite diffusion layer-based approximation, where the diffusion layer thickness was determined by a weighted average of the two diffusion coefficients. None of these two methods were concerned, however, about the kinetics of HER. In the work of Hessami and Tobias, (22) the pH profiles were parametrized by the current of HER (and no potential dependence of this current was analyzed), while in the works of Mayrhofer et al., (23,24) the approximation of full reversibility was used (that is, the authors assumed that the near-surface pH depends linearly on the applied potential, as dictated by Nernst’s equation).

Figure 2

Figure 2. Reaction schemes for HER. Two separate reactions are shown in (a) for acidic and neutral to alkaline solutions. A combined scheme is shown in (b) for near-neutral solutions. The autoprotolysis of water is part of both schemes.

In a recent work of our group, (27) we attempted to model HER by taking into account two strictly irreversible reactions, the reduction of H+ and that of water molecules, both following Erdey-Grúz–Volmer–Butler kinetics. In ref (27), we developed a digital simulation-based modeling approach to HER and we presented an approximative analytical model that could well describe polarization curves at various values of pH and rotation rates. In this model, we used an assumption that the diffusivity of OH ions exceeds that of H+ and that thus at high current densities the near-surface solution layer does not turn alkaline but neutral instead. The resulting model could be used to describe HER polarization curves measured on nickel electrodes.
The model described in ref (27) followed the reaction scheme shown in Figure 2a and it thus contained altogether four variable (fittable) kinetic parameters: two reaction rate and two charge-transfer coefficients, each describing the reduction of H+ and that of H2O molecules, respectively. We noticed, however, a strong correlation between the fitted parameters, which suggested that the reduction of this model would still be possible.
In this present paper, we aim to develop an analytical model that can well describe the polarization curves of HER, recorded on rotating disk electrodes immersed into mildly acidic solutions, by taking into consideration a quasireversible charge-transfer reaction, which represents the combination or Reactions R1 and R2 and that contains an inherent coupling, as a result of autoprotolysis, Reaction R3, between the concentrations of H+ and OH ions. From a mathematical point of view, this model, represented by Figure 2b, is simpler than the one previously described, (27) as it contains only two variable kinetic parameters (a single reaction rate and a single charge-transfer coefficient).
In what follows, we will give a brief description of the model and then present how it can be used for the estimation of kinetic parameters of HER on two chosen model electrodes (gold and platinum). We will demonstrate that the different lengths of the limiting current plateaus observed on these metals can be used as a direct measure of electrocatalytic hindrance.

Theory

ARTICLE SECTIONS
Jump To

Thermodynamic Considerations

Although Reactions R1 and R2 (the formation of H2 either by the reduction of H+ ions or by that of water molecules) are seemingly different, from a thermodynamic point of view, both processes lead to the same equilibrium conditions. The standard potentials are ER1 = 0 V and ER2 = −0.8277 V vs standard hydrogen electrode (SHE) (33) and assuming equilibrium conditions, we can use Nernst’s equation to relate the potentials of electrode Reactions R1 and RRR2 to the aH+ and aOH activities of H+ and OH ions, respectively, as well as to the ΦH2 fugacity of hydrogen gas:
(2)
and
(3)
while from the equilibrium condition ER1 = ER2 it follows that
(4)
Assuming that the autoprotolysis of water, Reaction R3, exactly determines the product of H+ and OH activities as the ionic product of water Kw, we get to Kw = 1.019 × 10–14 using the standard potential values mentioned before. It can thus be seen that Reactions R1R3 are not independent from each other and they all have to be considered when describing the thermodynamics of hydrogen evolution. Somewhat contradicting this statement, in the formal kinetic treatment of HER, it is still common to treat Reactions R1 and R2 as separate processes, each valid in its respective pH regime. In what follows, we aim to develop a kinetic treatment that can describe hydrogen evolution on an electrode surface near which the pH shifts, depending on the applied electrode potential, from acidic to alkaline values.

Kinetic Considerations

As a first step, we have to write kinetic rate equations for HER that take both H+ and OH ions into account. Probably the most straightforward possibility of doing this is to sum Reactions R1 and R2, in a manner illustrated by Figure 2b, to get
(R4)
Reaction R4 contains the H+ (or the H3O+) ion as a reactant and the OH ion as a product. Assuming that the reaction can proceed in both directions, the current density j yielded by the reaction can be expressed as a sum of cathodic (jc) and anodic (ja) terms:
(5)
Assuming that Reaction R4 follows the Erdey-Grúz–Volmer–Butler equation, the cathodic and anodic current densities may be formally expressed as
(6a)
and
(6b)
In eqs 6a and 6b, αc and αa are charge-transfer coefficients, k′ is a reaction rate coefficient, and the c0 terms stand for near-surface concentrations. The eqs 6a, 6b can be simplified by utilizing the usual assumption that αa + αc = n = 2 (the number of electrons involved in Reaction R4). (34) Further assuming that the near-surface concentrations of water and of H2 molecules can be treated as unit constants—the latter, at least, in a solution saturated with H2, assuming that the dissolved H2 concentration shows no significant pH dependence—we can introduce another reaction rate coefficient k in place of k′ as
(7)
This turns eqs 6a and 6b into
(8a)
and
(8b)
In eqs 6a, 6b, 8a, and 8b, E° denotes a potential value where no net current flows in the case when cH+0 = cOH0. E° can thus be expressed as
(9)
using Nernst’s equation (where we again assumed a unity fugacity of H2).
Provided that the activity coefficients of H+ and OH ions can both be considered unity and that the autoprotolysis Reaction R3 is fast enough so that it always maintains the equilibrium constraint that
(10)
independent of space and of time, eqs 8a, 8b can further be simplified to the following form:
(11a)
and
(11b)
where c = 1 mol dm–3 is the standard concentration.
Note that to get from eqs 6a and 6b to eqs 11a and 11b, we made use of eq 9, where, by definition, ER1 = 0 V vs SHE. Thus, the electrode potential E, appearing in equation set 11a, 11b, is also to be referenced to SHE.
Equations 11a and 11b determine the current of the electrode reaction provided that the near-surface concentration of hydrogen ions, cH+0, is known. On rotating disk electrodes (RDEs), stationary currents can be measured for HER and deriving a mathematical expression for cH+0 becomes possible as described below.

Problem of Transport

Provided that (i) mass transfer occurs only by means of diffusion and convection and other means of transport (e.g., migration) can be ignored and (ii) that the diffusion coefficients DH+ and DOH are constants, independent of the concentrations and of spatial coordinates, the condition of stationarity can be expressed in the form of the following equation:
(12)
Here, vz denotes the axial (z direction) component of the stationary fluid flow under the RDE that can be approximated (35) as
(13)
where a ≈ 0.51023, (28) ω denotes the angular velocity of rotation, and ν the kinematic viscosity of the solution.
If we now express the concentration of OH ions from eq 10 as and plug this into eq 12, we arrive to a nonlinear ordinary differential equation, describing stationary H+ concentration profiles under the RDE. This equation has no analytical solution; assuming, however, that the diffusion coefficients of H+ and OH ions are equal—and this assumption is not uncommon— (22)the differential equation can be linearized in the form:
(14)
where is the diffusion coefficient of H+ and OH ions (assumed to be equal) and the function
(15)
was introduced to describe the difference of H+ and OH concentrations as a function of the z distance measured from the electrode surface. Equation 14 is a linear, second-order ordinary differential equation with a known solution:
(16)
In eq 16, Δc0 and Δc denote near-surface (z = 0) and bulk (z → ∞) concentration differences (see eq 15) and Qs(x) is the regularized incomplete gamma function (36) (with s = 1/3) defined as
(17)
From eq 16, the current density can be expressed as
(18)
where
(19)
is the generalized Nernstian diffusion layer thickness that contains the gamma function (36) defined by the integral
(20)

Modeling the Polarization Curve of an RDE

The current density expressed by eq 18 is equal to that given in eq 5, with the partial current densities defined by eqs 11a and 11b. This yields the following expression for cH+0:
(21)
where the θ(α) function (a dimensionless, potential dependent, combined kinetic parameter) is defined as
(22)
Finally, the equation of a polarization curve recorded on an RDE can be expressed by combining eqs 18 and 21 to
(23)
Equation 23 is the final result of this theoretical treatment. It is an analytical formula for the current density of an RDE on which hydrogen is evolved at a given pH, rotation rate (f), and electrode potential (E). The parameters of this model are listed in Table 1, where it can be seen that the model relies only on two kinetic parameters, the k reaction rate coefficient and the αc charge-transfer coefficient of Reaction R4. Note that the presented model is very robust, as it considers only the effect of charge transfer, that of mass transport occurring by diffusion and convection, and the effect of the autoprotolysis. Although no mechanistic details (such as those of H adsorption to the surface) are considered, we will see that this robust model delivers well when used for the fitting of experimentally obtained data.
Table 1. Description of Symbols Used
symbolmeaningformula or typical value(s)
basic physicochemical parameters
pHpH of the bulk of the solution3
frotation rate of the RDE625 min–1
kreaction rate coefficient for Reaction R41 μm s–1
αccharge-transfer coefficient for Reaction R41
Ttemperature298.15 K
diffusion coefficient of H+ and OH ions, considered equal10–4 cm2 s–1
νkinematic viscosity of the solution8.917 × 10–7 m2 s–1
constants
FFaraday’s constant96 485.3 C mol–1
RRegnault’s constant8.314 J mol–1 K–1
cstandard concentration1 mol dm–3
aKármán’s constant (28)0.51023
Γ(1/3)see eq 202.67894
ER1standard potential of Reaction R10 V vs SHE
ER2standard potential of Reaction R2–0.8277 V vs SHE
Kwautoprotolysis constant of water1.019 × 10–14
derived quantities
cH+H+ concentration in the bulk of the solution10–pHc
Δcdifference of H+ and OH concentrations in the bulk
ωangular frequency of rotationf
δNgeneralized Nernstian diffusion layer thickness, see eq 19
θ(α)combined kinetic parameter, see eq 22
cH+0near-surface H+ concentration, see eq 21
jcurrent density of the RDE, see eq 23
jcat,H+catalytic current density of H+ reduction, see eq 24
j0exchange current density, see eq 25
limiting current density, see eq 26
jcat,H2Ocatalytic current density of water splitting, see eq 29
jrevNernstian (reversible) current density, see eq 31

Experimental

ARTICLE SECTIONS
Jump To

The rotating disk electrodes used in this study were obtained from Metrohm (Herisau, Switzerland). The diameter of the disk electrodes was (3.00 ± 0.05) mm, embedded into a PEEK shaft of 10 mm outer diameter. The geometric surface area was used for calculating current densities. Prior to the experiments, the electrodes were dipped, for a few moments, into Caro’s acid and were then rinsed abundantly with ultrapure water (Milli-Q by Merck Millipore, R = 18.2 MΩ cm, used for the preparation of solutions as well).
The studied solutions of different pH were prepared by diluting calculated amounts of a 0.1 mol dm–3 HClO4 (70%, Merck, Suprapure) stock solution and solid NaClO4 (99.99%, trace metals basis, Merck) with ultrapure water, by keeping the ionic strength at a fixed value of 0.1 mol dm–3. The pH of the solutions was measured by a calibrated Metrohm 914 pH meter.
A lab-made three-electrode glass cell was used for the experiments. For measurements on Au, a large Au foil was used as a counter electrode, while for measurements on Pt, we applied a large surface area Pt foil. All measurements were carried out by using a Hg|Hg2SO4|K2SO4 (sat) reference electrode (Radiometer Analytical XR200, connected to the main chamber through a Luggin capillary). Electrode potentials in the paper are reported with reference to the standard hydrogen electrode (SHE); for the potential shift, the value of EHg/Hg2SO4 = 650 mV was used.
Prior to measurements, the solution in the cell was deaerated with a pure Ar (5N, Alphagaz) flow for 15 min and saturated by hydrogen (5N, Alphagaz). The electrodes were submerged to the electrolyte solution without potential control, and the equilibrium potential Eeq was determined by measuring the open-circuit potential. Polarization curves presented in the paper were recorded point by point by steady-state current measurements at given electrode potentials and rotation rates, according to the following sequence (see Figure 3): the electrode potential and the rotation rate were set, and the current was measured until it reached a stationary value. Then, the potential was set back to Eeq and a rotation rate of f = 4500 min–1 was applied for some seconds to remove any accumulated H2 from the surface. The current measurement was then repeated with other potential and rotation rate settings. The measured data were IR-corrected postexperimentally—the solution resistance was determined by means of high-frequency impedance measurements. During all measurements, the solutions were kept saturated with H2 by continuous but slow purging (that did not interfere with the hydrodynamics of rotation).

Figure 3

Figure 3. Experimental protocol for the measurement of HER polarization curves on RDEs. The stationary current of the RDE is measured at given electrode potential (E) and rotation rate (f) values (green periods). Between the measurements, the potential control is switched off and the RDE rotated quickly, to remove accumulated bubbles (red periods).

The measurements were automated by using an Autolab PGSTAT128N potentiostat in connection with a Metrohm Autolab rotator unit and by the application of the Nova v2.1 software.

Results

ARTICLE SECTIONS
Jump To

Experimentally obtained polarization curves are shown in Figure 4 for both the gold and the platinum RDE. These curves were recorded in mildly acidic solutions (2.0 ≲ pH ≲ 3.6) and at different rotation rates (400 < f < 2500). As shown by the figures, the measured polarization curves can well be fitted using eq 23. Note that during the fitting we varied only three parameters (, k, and αc), while the other parameters were fixed at values listed in Table 1. Also note that the optimization was carried out by considering all data points shown either for gold or for platinum Figure 4 and not in a curve-by-curve manner. Even under such strict conditions, the calculated values (shown by the green curves) match reasonably well the measured ones (shown by the red dots).

Figure 4

Figure 4. Experimentally obtained polarization curves (red dots) on an Au (a) and on a Pt (b) RDE, showing a two step behavior. In each panel (at different values of pH), the cathodic current density increases as the rotation rate f is set to values of 400, 625, 900, 1225, 1600, 2025, and 2500 min–1. The green curves are created by fitting the model described by eq 23 globally, that is, for all pH and rotation rate values, and by optimizing only three parameters (αc, k, and ). Determined confidence intervals (at 95% statistical certainty) for the fitted parameters are αc = 0.486 ± 0.067, and for gold and αc = 0.643 ± 0.037, and for platinum. Other parameter values (not optimized) are shown in Table 1.

Optimized values of αc, k, and are given in the caption of Figure 4 for Au and Pt. As can be seen, the optimized values match well with the diffusion coefficient of H+ ion known from the literature. (31) The determined values of αc are also in good agreement with those found in the literature (37) and, as we will show later, the determined k and α values can be recombined to exchange current densities in the expected range for both gold (38) and platinum. (11)
A notable difference between the polarization curves of gold and platinum, as can be seen in Figure 4, is in the length of their limiting current plateaus, which is substantially shorter in the case of Pt. This, as we will see below, can be explained by the reaction rate coefficient k on platinum being 5 orders of magnitude higher compared with gold, and by that the determined αc values for the two metals also differ.

Discussion

ARTICLE SECTIONS
Jump To

General Behavior of the Model; Polarization Curve Segments

As could be seen in Figure 4, the model function given by eq 23 can describe well the two step behavior of HER polarization curves. For the parameter values listed in Table 1, a calculated polarization curve is shown in Figure 5 (thick gray curve).

Figure 5

Figure 5. Polarization curve (thick gray) calculated using eq 23 and the parameter values of Table 1, shown in two different representations: with linear axis scaling in (a) and on a Tafel plot in (b). The three different segments of the curves, marked by the dashed lines, can be approximated by eqs 2430, as discussed in the text. The contour map in the background of (a) shows the variation of pH as a function of the distance measured from the electrode surface at each disk potential. (Details of calculating pH profiles using the presented model are discussed later, cf. to Figure 9).

First of all, it can be seen that the modeled polarization curves clearly exhibit three distinct parts: (i) a starting exponential rise, assigned to the charge-transfer-controlled reduction of H+ ions denoted by jcat,H+ and plotted by a dashed red curve in Figure 5; (ii) an almost horizontal plateau section (jlim), shown by a dashed black line, where the current is limited by the rate of transport of H+; and, finally, (iii) another exponential rise that we may assign to a charge-transfer-controlled reduction of H2O, denoted by jcat,H2O and is shown by the green dashed curve in Figure 5.
The model is able to describe a smooth transition between the aforementioned limiting cases. Formulae for the current density for each limiting case can be derived as follows.

Catalytic Current of H+ Reduction, jcat,H+

In this limiting case, the current is controlled by the catalytic reduction of (i.e., charge transfer to) H+ ions. If the cathodic potential is far enough from the equilibrium potential Eeq (where the rate of the opposite reaction, hydrogen oxidation, is negligible) but still not very negative (so that the transport of H+ ions still does not become rate limiting), current can be determined by assuming that ω → ∞ and that . This turns eq 23 into
(24)
The current density calculated from eq 24 is plotted by the red dashed curve in Figure 5. As shown by the Tafel representation in Figure 5b, this section of the polarization curve is characterized by a Tafel slope of . Equation 24 is also useful for expressing the exchange current density j0 by evaluation at
(25)
The variation of exchange current densities on pH is shown for gold and platinum in Figure 6. These data agree well with values found in the literature. (11,38)

Figure 6

Figure 6. Exchange current densities j0 in the studied pH range, calculated by using eq 25, based on the kinetic parameters determined by nonlinear fitting in Figure 4 for gold and platinum.

Limiting Current of H+ Reduction, jlim

In this limiting case, shown by the black dashed line in Figure 5, the current density is governed solely by the transport of H+ ions from the bulk of the solution to the electrode surface. A formula for the limiting current can be provided within the framework of the presented model by assuming that the near-surface concentration difference of H+ and OH ions (Δc0) equals zero in eq 18. Then
(26)
where for fairly acidic solutions, we can assume that ΔccH+ and thus
(27)
The Tafel slope corresponding to this flat plateau is, as shown in Figure 5b, ∞.

Case of Mixed Charge-Transfer/Transport Control for H+ Reduction

It can be shown that in this region of mixed control, where the gray curve in Figure 5 already leaves jcat,H+ but still does not attain the limiting current plateau, our model yields current densities that fully match those calculated from the Koutecký–Levich equation (39) and thus for this “first transition” section of the polarization curve
(28)

Charge-Transfer-Controlled H2O Reduction

When the potential is very negative, a series expansion of the current, as given by eq 23, around −∞ for the term to the first order yields the following equation
(29)
Note that, as expected, this equation is independent of cH+. This current is shown by the dashed green curve in Figure 5; as seen in the Tafel representation, Figure 5b, the corresponding Tafel slope is .

Mixed Charge-Transfer Control of H2O Reduction and Transport Control of H+ Reduction

It can be shown that at the potential regime between these two segments, i.e., in the “second transition section”, the current can be described by the equation
(30)
In what follows, we will analyze the effect of varying certain parameters (αc and k) on the calculated polarization curves.

Parameter Dependencies

Dependence on αc

Polarization curves calculated based on eq 23 show a strong dependence on the value of the charge-transfer coefficient αc, as illustrated by Figure 7a,b.

Figure 7

Figure 7. Effect of varying the parameters αc (a, b) and k (c, d) on the calculated polarization curves. Values assumed are shown on the graph; other parameters are given in Table 1. Polarization curves are shown in two different representations: with linear axis scaling (a, c) and on a Tafel plot (b, d).

With respect to the definition of the charge-transfer coefficient αc, we emphasize that since the presented model is built on Reaction R4, a two-electron reaction, we assumed that αc + αa = 2; (34) thus, based on this definition, αc = 1 represents symmetry.
Note in Figure 7b that the Tafel slopes of the first and second exponentially increasing segments are and , in accordance with what was said earlier about these segments. Also note that as a result of the two differing Tafel slopes, the length of the transport-limited plateau is heavily influenced by the value of αc: smaller αc values result in longer plateaus.

Dependence on k

The dependence of the polarization curves on the reaction rate coefficient k is illustrated by Figure 7c,d. Note that as shown by the figure, the polarization curves tend toward a reversible curve (indicated in Figure 7c,d by a dashed blue line) if we increase the value of k. Indeed, in the k → ∞ limit, the current of eq 23 reduces to
(31)
Note that we get to this very same expression of the reversible current if we solve eq 12 by assuming that the value of Δc0 is determined by Nernst’s equation (that is, if we utilize a Nernstian boundary condition). As expected, eq 31 contains no kinetic parameters as it is a consequence of thermodynamic and transport-related considerations.
Note in Figure 5 that the “first steps” of the polarization curves calculated for finite k values tend to achieve the reversibility limit quite easily. In the case of αc = 1, the first steps of the polarization curves calculated for k ≥ 10–2 m s–1 are practically indistinguishable from the reversible current. This agrees well with the experimental observations and the argumentation of Gasteiger et al. who have warned that in such cases no conclusions with respect to the kinetics of HER should be drawn from RDE experiments. (12) We note, however, that for the “second step” of the polarization curves, it takes far higher (in fact, unrealistically high, k ≥ 1 m s–1) rate coefficients to match the reversibility case. This finding offers a new perspective for the interpretation of RDE polarization curves, as will be discussed below.

Kinetic Parameters from Plateau Lengths

As we saw before, HER polarization curves exhibit plateaus of different lengths when measured on metals of different electrocatalytic activities (Figure 4); this behavior is reproduced by the presented model. Although the fitting of the whole model to a set of measurements is computationally not difficult—eq 23 is, after all, an analytical expression—and fitting the full model is always favorable, it seems worthy to analyze plateau lengths and, in particular, their dependence on the pH, in the hope that this analysis will make kinetic parameters accessible more easily.
As the term plateau length is, however, not well-defined, it seems easier to introduce the concept of breakdown overpotential (ηbr) as a quantity of similar meaning. We define ηbr, as illustrated by Figure 8a, as the overpotential at which the measured current, following the second current step, reaches the value of 2jlim.

Figure 8

Figure 8. (a) Concept of the breakdown overpotential ηbr illustrated on a polarization curve. (b) Dimensionless breakdown overpotentials plotted as a function of pH for gold and platinum, determined from measured data (dots). The lines were created by linear fitting to the measured data, using a pH2 weighting. The acquired slopes and intercepts were used according to eq 32 to calculate the kinetic parameters shown in Table 2. Chosen rotation rate: 625 min–1.

Keeping in mind that , we can obtain the following formula for the value of ηbr from eq 30:
(32)
That is, following a standard Levich analysis for the determination of , we can plot the normalized breakdown potentials (measured at a chosen, fixed rotation rate) as a function of pH; we can then perform linear fitting to these data (shown in Figure 8b for a rotation rate of 625 min–1 for gold and platinum) and determine αc and k values from the slope and intercept. As shown in Table 2, the results are acceptable but not as reliable as those of full-model fitting.
Table 2. Kinetic Parameters of HER on Pt and on Au, in Mildly Acidic Solutionsa
systemαc
AuFigure 4–5.10 ± 0.260.486 ± 0.067
 Figure 8–5.2 ± 1.40.50 ± 0.11
PtFigure 40.024 ± 0.0500.643 ± 0.037
 Figure 80.2 ± 1.60.63 ± 0.16
a

Parameters were determined either by the fitting of the full model to all measured data, Figure 4, or by plotting limiting current plateau lengths as a function of pH, Figure 8.

pH Profiles

The presented model may not only be found useful to fit HER polarization curves but it may also give predictions on how the local pH changes in the vicinity of a rotating electrode at which H2 evolves. The basis of such predictions is eq 16. In it, expressing Δc0 with the aid of the current density j by using eq 23, we arrive to the following equation describing the variation of the concentration differences of H+ and OH as a function of the distance measured from the electrode surface:
(33)
Concentration difference profiles, calculated based on eq 33, are shown in Figure 9a. Provided that pH is known, eq 33 permits the direct calculation of pH profiles as well; for that we need to recall the definition of Δc in eq 15. The pH(z) profile can then be calculated as
(34)
Some example pH profiles are shown in Figure 9b.

Figure 9

Figure 9. (a) Normalized concentration difference profiles as a function of the normalized distance, for some chosen values of the normalized current density (shown in the figure). (b) An example for pH profiles calculated for various normalized current values, assuming that pH = 3.

It follows from eqs 33 and 18 that as the cathodic current increases, pH values measured in the vicinity of the electrode will rise. Exactly at j = jlim, the pH at the electrode surface reaches neutrality (pH0 ≈ 7). When a cathodic current density higher than jlim is forced through the electrode surface, the near-electrode solution region gets alkaline, yet at a given distance, the pH drops suddenly to acidic values (see Figure 9b).
The distance at which the aforementioned drop occurs (i.e., the distance of the neutrality point zneut) depends on how much the current density exceeds the limiting current density and it can be expressed as
(35)
where we denoted by Qs–1(x) the inverse of the regularized incomplete gamma function Qs(x), defined by eq 17.
In Figure 10, zneut is plotted as a function of the normalized current density . The zneut quantity can be interpreted as a measure of how deep the near-electrode solution layer becomes alkaline, as shown by Figure 9. Note in Figure 10 that true alkalination (i.e., the appearance of a pH ≳ 7 region) can only occur if and that at high current densities, the depth of alkalination can exceed the diffusion layer thickness δN.

Figure 10

Figure 10. Full black curve: the distance of neutrality (normalized to the diffusion layer thickness δN) as a function of the current density normalized to . Note that the function, eq 35, is not defined for . Dashed gray curve: an estimate for the neutrality distance based on an analytical solution assuming that DOHDH+. (27)

Concluding Remarks; Comparison to Other Works

The presented model seems suitable to describe polarization curves of HER measured at two electrodes of different electrocatalytic activities (Au and Pt). The extracted parameters are in agreement with some shown in the literature before, (11,20) yet at this point, we would like to emphasize limitations of the treatment and to make a brief comparison to earlier works on the topic of HER.
An important issue, to which we would like to direct the readers attention, is the equidiffusivity assumption that we utilized. This assumption, although not unprecedented in the literature (see the works of Tobias as an example (22)), poses some limitations to the validity of our analysis. Avoiding use of the assumption is, unfortunately, not possible, as it is required for eq 12 to be analytically solvable. The equidiffusivity assumption means that we consider the diffusion coefficients of H+ and OH ions equal, appearing as a common coefficient in our equations. We are aware (and the reader should also be) that this approximation is only valid on an order of magnitude level (while in reality, DH+ is supposed to be about 2 times higher than DOH). The equidiffusivity assumption, as also pointed out by Tobias, (22) renders the concentration of OH at and under the disk surface to be underestimated. That the model equations can still be used to obtain good apparent fits of the measured polarization curves is probably attributed to the fact that currents measured at high cathodic overpotentials show only a moderate (order of 1/3) dependence on (cf. eq 29).
At this point, it seems worthy to make a comparison between the model presented here and one of our previous attempts at modeling polarization curves of HER. In ref (27), we constructed analytical approximations to an otherwise digital simulation-based model where we assumed that the diffusion coefficient of OH is not only equal to but actually much higher than that of H+. This assumption also led to a fittable model function, but it predicted that the near-surface solution region, instead of getting alkaline, would get neutral up to a certain depth. Of course, under these circumstances, the “distance of neutrality” was higher than the one obtained here: instead of the formation of a thin alkaline layer, we assumed the formation of a thicker, however, neutral layer. In Figure 10, we show a comparison between the two approaches.
A further difference between the model presented here and the one we described before (27) is that while the previous one was dealing with two irreversible reactions (namely, the reduction of H+ ions and that of water molecules), the present model is built on a combined, quasireversible, reaction (Reaction R4). The present model thus performs better compared with the previous one in two ways: (i) it provides good fits with only two (instead of four) kinetic parameters, and (ii) it is also able to model a quasiequilibrium (that is, the k → ∞ case). Although within the framework of the present model it is still possible to distinguish two reactions (see our discussion of polarization curve segments assigned to H+ reduction and to water splitting), we emphasize here that such distinctions are all but arbitrary and the treatment presented here is built on a single charge-transfer reaction, Reaction R4.
In comparison to the models of some other authors, (22−24) a clear advantage of the model presented here is that it contains kinetic parameters and does not rely on the use of the Nernstian boundary condition. This condition would predict that the near-surface pH is exactly determined by, and directly proportional to, the electrode potential. Instead of using a Dirichlet (i.e., Nernstian) boundary condition, our model utilizes a Neumann condition (40) and thus allows the calculation of kinetic currents.
Although the model seems to fit experimental data on a broad scale (Figure 4), here we would like to draw the attention of the readers to another limitation, which arises from the very fact that we assumed the validity of the Erdey-Grúz–Volmer–Butler equation for Reaction R4. We thus ignored any effects related to surface kinetics and that the parameters k and αc can be potential dependent. (41) This simplification was, however, necessary to describe the experimentally observed two step behavior of HER, visible at high cathodic overpotentials. Consequently, the k and αc parameters that we determined here can be considered valid primarily at high overpotential. Although some variations of these parameters may occur—and this is probably responsible for the fits of Figure 4 not being perfect at low overpotentials—the overall fits are still satisfactory.

Conclusions

ARTICLE SECTIONS
Jump To

We have developed a new model that is able to describe HER as it occurs on RDEs immersed into mildly acidic solutions, where the polarization curves show a two step behavior. The model is centered around a single reaction, Reaction R4, that contains both H+ ions (as a reactant) and OH ions (as a product). We assumed that the Erdey-Grúz–Volmer–Butler equation applies for this reaction and that the diffusion coefficient of the two reacting species (H+ and OH) are equal. On the basis of these assumptions, we managed to solve the differential equations governing the system; the resulting analytical model could be used for the fitting of experimentally obtained polarization curves. By varying only three model parameters, we achieved good fits over a relatively broad range of pH and rotation rates for both Au and Pt RDEs.
A very important implication of the model is that the plateau lengths seen on RDE polarization curves are (inversely) related to electrocatalytic activity. We showed that at fixed rotation rates, a linear relationship exists between the plateau length and the bulk solution pH. By analyzing this relationship, we can get a good estimate of the kinetic parameters k and αc, even in cases where the transport performance of the RDE is not sufficient to measure well-defined kinetic currents using the standard Koutecký–Levich analysis. (12,41)
Within the presented framework, it is also possible to model the variation of pH as a function of distance measured from the electrode surface. This result may become useful if we study HER as a side reaction of, for example, metal deposition processes where local pH rises can have unwanted effects on the deposit.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Authors
  • Authors
    • María de Jesús Gálvez-Vázquez - Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    • Vitali Grozovski - Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
    • Noémi Kovács - Department of Chemistry and Biochemistry, University of Bern, Freiestrasse 3, CH-3012 Bern, SwitzerlandDepartment of Physical Chemistry, Eötvös Loránd University of Budapest, Pázmány Péter sétány 1/A, H-1117 Budapest, Hungary
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

Support by the CTI Swiss Competence Center for Energy Research (SCCER Heat and Electricity Storage) is gratefully acknowledged. P.B. acknowledges financial support from the Swiss National Foundation (grant No. 200020–172507). S.V. acknowledges support from the National Research, Development and Innovation Office of Hungary (NKFIH grant Nos. PD124079 and K129210). M.d.J.G.-V. and N.K. acknowledge the financial support by the Swiss Government Excellence Scholarships for Foreign Scholars (ESKAS).

References

ARTICLE SECTIONS
Jump To

This article references 41 other publications.

  1. 1
    Wang, H.; Gao, L. Recent Developments in Electrochemical Hydrogen Evolution Reaction. Curr. Opin. Electrochem. 2018, 7, 714,  DOI: 10.1016/j.coelec.2017.10.010
  2. 2
    Ooka, H.; Figueiredo, M. C.; Koper, M. T. M. Competition between Hydrogen Evolution and Carbon Dioxide Reduction on Copper Electrodes in Mildly Acidic Media. Langmuir 2017, 33, 93079313,  DOI: 10.1021/acs.langmuir.7b00696
  3. 3
    Cave, E. R.; Shi, C.; Kuhl, K. P.; Hatsukade, T.; Abram, D. N.; Hahn, C.; Chan, K.; Jaramillo, T. F. Trends in the Catalytic Activity of Hydrogen Evolution during CO2 Electroreduction on Transition Metals. ACS Catal. 2018, 8, 30353040,  DOI: 10.1021/acscatal.7b03807
  4. 4
    Schlesinger, M.; Paunovic, M., Eds., Modern Electroplating, 5th ed.; Wiley: New York, 2010.
  5. 5
    Ritzert, N. L.; Moffat, T. P. Ultramicroelectrode Studies of Self-Terminated Nickel Electrodeposition and Nickel Hydroxide Formation upon Water Reduction. J. Phys. Chem. C 2016, 120, 2747827489,  DOI: 10.1021/acs.jpcc.6b10006
  6. 6
    Wu, J.; Wafula, F.; Branagan, S.; Suzuki, H.; van Eisden, J. Mechanism of Cobalt Bottom-Up Filling for Advanced Node Interconnect Metallization. J. Electrochem. Soc. 2019, 166, D3136D3144,  DOI: 10.1149/2.0161901jes
  7. 7
    Horányi, G. Electrochemical Dictionary;Bard, A. J.; Scholz, F.; Inzelt, G., Eds.; Springer Verlag: Heidelberg, 2008; p 343.
  8. 8
    Bockris, J. O. M.; Ammar, I. A.; Huq, A. K. M. S. The Mechanism of the Hydrogen Evolution Reaction on Platinum, Silver and Tungsten surfaces in Acid Solutions. J. Phys. Chem. A. 1957, 61, 879886,  DOI: 10.1021/j150553a008
  9. 9
    Bagotzky, V. S.; Osetrova, N. V. Investigations of Hydrogen Ionization on Platinum with the Help of Micro-Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1973, 43, 233249,  DOI: 10.1016/S0022-0728(73)80494-2
  10. 10
    Tavares, M. C.; Machado, S. A. S.; Mazo, L. H. Study of Hydrogen Evolution Reaction in Acid Medium on Pt Microelectrodes. Electrochim. Acta 2001, 46, 43594369,  DOI: 10.1016/S0013-4686(01)00726-5
  11. 11
    Neyerlin, K. C.; Gu, W.; Jorne, J.; Gasteiger, H. A. Study of the Exchange Current Density for the Hydrogen Oxidation and Evolution Reactions. J. Electrochem. Soc. 2007, 154, B631B635,  DOI: 10.1149/1.2733987
  12. 12
    Sheng, W.; Gasteiger, H. A.; Shao-Horn, Y. Hydrogen Oxidation and Evolution Reaction Kinetics on Platinum: Acid vs Alkaline Electrolytes. J. Electrochem. Soc. 2010, 157, B1529B1536,  DOI: 10.1149/1.3483106
  13. 13
    Durst, J.; Siebel, A.; Simon, C.; Hasché, F.; Herranz, J.; Gasteiger, H. A. New insights into the electrochemical hydrogen oxidation and evolution reaction mechanism. Energy Environ. Sci. 2014, 7, 22552260,  DOI: 10.1039/C4EE00440J
  14. 14
    Zheng, J.; Sheng, W.; Zhuang, Z.; Xu, B.; Yan, Y. Universal Dependence of Hydrogen Oxidation and Evolution Reaction Activity of Platinum-group Metals on pH and Hydrogen Binding Energy. Sci. Adv. 2016, 2, e1501602  DOI: 10.1126/sciadv.1501602
  15. 15
    Pentland, N.; Bockris, J. O. M.; Sheldon, E. Hydrogen Evolution Reaction on Copper, Gold, Molybdenum, Palladium, Rhodium, and Iron. J. Electrochem. Soc. 1957, 104, 182,  DOI: 10.1149/1.2428530
  16. 16
    Ives, D. J. G. Some Abnormal Hydrogen Electrode Reactions. Can. J. Chem. 1959, 37, 213221,  DOI: 10.1139/v59-028
  17. 17
    Brug, G. J.; Sluyters-Rehbach, M.; Sluyters, J. H.; Hemelin, A. The Kinetics of the Reduction of Protons at Polycrystalline and Monocrystalline Gold Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1984, 181, 245266,  DOI: 10.1016/0368-1874(84)83633-3
  18. 18
    Conway, B. E.; Bai, L. State of Adsorption and Coverage by Overpotential-Deposited H in the H2 Evolution Reaction at Au and Pt. Electrochim. Acta 1986, 31, 10131024,  DOI: 10.1016/0013-4686(86)80017-2
  19. 19
    Khanova, L. A.; Krishtalik, L. I. Kinetics of the hydrogen evolution reaction on gold electrode. A new case of the barrierless discharge. J. Electroanal. Chem. 2011, 660, 224229,  DOI: 10.1016/j.jelechem.2011.01.016
  20. 20
    Kahyarian, A.; Brown, B.; Nesić, S. Mechanism of the Hydrogen Evolution Reaction in Mildly Acidic Environments on Gold. J. Electrochem. Soc. 2017, 164, H365H374,  DOI: 10.1149/2.1061706jes
  21. 21
    Obata, K.; Stegenburga, L.; Takanabe, K. Maximizing Hydrogen Evolution Performance on Pt in Buffered Solutions: Mass Transfer Constrains of H2 and Buffer Ions. J. Phys. Chem. C 2019, 123, 2155421563,  DOI: 10.1021/acs.jpcc.9b05245
  22. 22
    Hessami, S.; Tobias, C. W. In-Situ Measurement of Interfacial pH Using a Rotating Ring-Disk Electrode. AIChE J. 1993, 39, 149162,  DOI: 10.1002/aic.690390115
  23. 23
    Katsounaros, I.; Meier, J. C.; Klemm, S. O.; Topalov, A. A.; Biedermann, U. P.; Auinger, M.; Mayrhofer, K. J. J. The Effective Surface pH during Reactions at the Solid-Liquid interface. Electrochem. Commun. 2011, 13, 634637,  DOI: 10.1016/j.elecom.2011.03.032
  24. 24
    Auinger, M.; Katsounaros, J. C.; Meier, I.; Klemm, S. O.; Biedermann, P. U.; Topalov, A. A.; Rohwerder, M.; Mayrhofer, K. J. J. Near-surface ion distribution and buffer effects during electrochemical reactions. Phys. Chem. Chem. Phys. 2011, 13, 1638416394,  DOI: 10.1039/c1cp21717h
  25. 25
    Shinagawa, T.; García-Esparza, A. T.; Takanabe, K. Insight on Tafel Slopes from a Microkinetic Analysis of Aqueous Electrocatalysis for Energy Conversion. Sci. Rep. 2015, 5, 13801  DOI: 10.1038/srep13801
  26. 26
    Strmcnik, D.; Uchimura, M.; Wang, C.; Subbaraman, R.; Danilovic, N.; van der Vliet, D.; Paulikas, A. P.; Stamenkovic, V. R.; Markovic, N. M. Improving the Hydrogen Oxidation Reaction Rate by Promotion of Hydroxyl Adsorption. Nat. Chem. 2013, 5, 300306,  DOI: 10.1038/nchem.1574
  27. 27
    Grozovski, V.; Vesztergom, S.; Láng, G. G.; Broekmann, P. Electrochemical Hydrogen Evolution: H+ or H2O Reduction? A Rotating Disk Electrode Study. J. Electrochem. Soc. 2017, 164, E3171E3178,  DOI: 10.1149/2.0191711jes
  28. 28
    Vesztergom, S. Encyclopedia of Interfacial Chemistry: Surface Science and Electrochemistry;Wandelt, K., Ed.; Elsevier: Amsterdam, 2018; pp 421444.
  29. 29
    Nagel, K.; Wendler, F. Die Wasserstoffelektrode als zweifache Elektrode. Z. Elektrochem. 1956, 60, 10641072,  DOI: 10.1002/bbpc.19560600924
  30. 30
    Kanzaki, Y.; Tokuda, K.; Bruckenstein, S. Dissociation Rates of Weak Acids Using Sinusoidal Hydrodynamic Modulated Rotating Disk Electrode Employing Koutecky-Levich Equation. J. Electrochem. Soc. 2014, 161, H770H779,  DOI: 10.1149/2.0221412jes
  31. 31
    Wiberg, G. K. H.; Arenz, M. On the Influence of Hydronium and Hydroxide Ion Diffusion on the Hydrogen and Oxygen Evolution Reactions in Aqueous Media. Electrochim. Acta 2015, 159, 6670,  DOI: 10.1016/j.electacta.2015.01.098
  32. 32
    Carneiro-Neto, E. B.; Lopes, M. C.; Pereira, E. C. Simulation of Interfacial pH Changes during Hydrogen Evolution Reaction. J. Electroanal. Chem. 2016, 765, 9299,  DOI: 10.1016/j.jelechem.2015.09.029
  33. 33
    Vanýsek, P. CRC Handbook of Chemistry and Physics, 93rd ed.; Haynes, W. M., Ed.; Chemical Rubber Company: Boca Raton FL, 2012.
  34. 34
    Guidelli, R.; Compton, R. G.; Feliu, J. M.; Gileadi, E.; Lipkowski, J.; Schmickler, W.; Trasatti, S. Defining the Transfer Coefficient in Electrochemistry: An Assessment (IUPAC Technical Report). Pure Appl. Chem. 2014, 86, 245258,  DOI: 10.1515/pac-2014-5026
  35. 35
    Bard, A. J.; Faulkner, L. R. Electrochemical Methods. Fundamentals and Applications; John Wiley & Sons: New York, 2001.
  36. 36
    Bronstein, I. N.; Semendjajew, K. A.; Musiol, G.; Muehlig, H. Taschenbuch der Mathematik; Verlag Harri Deutsch: Frankfurt am Main, 2008.
  37. 37
    Haghighat, S.; Dawlaty, J. M. pH Dependence of the Electron-Transfer Coefficient: Comparing a Model to Experiment for Hydrogen Evolution Reaction. J. Phys. Chem. C 2016, 120, 2848928496,  DOI: 10.1021/acs.jpcc.6b10602
  38. 38
    Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Kitchin, J. R.; Chen, J. G.; Pandelov, S.; Stimming, U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23J26,  DOI: 10.1149/1.1856988
  39. 39
    Koutecky, J.; Levich, V. G. The use of a rotating disk electrode in the studies of electrochemical kinetics and electrolytic processes. Zh. Fiz. Khim. 1958, 32, 15651575
  40. 40
    Cheng, A. H.-D.; Cheng, D. T. Heritage and early history of the boundary element method. Eng. Anal. Boundary Elem. 2005, 29, 268302,  DOI: 10.1016/j.enganabound.2004.12.001
  41. 41
    Kucernak, A. R.; Zalitis, C. General Models for the Electrochemical Hydrogen Oxidation and Hydrogen Evolution Reactions: Theoretical Derivation and Experimental Results under Near Mass-Transport Free Conditions. J. Phys. Chem. C 2016, 120, 1072110745,  DOI: 10.1021/acs.jpcc.6b00011

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 10 publications.

  1. Tomohiro Fukushima, Motoki Fukasawa, Kei Murakoshi. Unveiling the Hidden Energy Profiles of the Oxygen Evolution Reaction via Machine Learning Analyses. The Journal of Physical Chemistry Letters 2023, 14 (30) , 6808-6813. https://doi.org/10.1021/acs.jpclett.3c01596
  2. Brian M. Tackett, David Raciti, Nicholas W. Brady, Nicole L. Ritzert, Thomas P. Moffat. Potentiometric Rotating Ring Disk Electrode Study of Interfacial pH during CO2 Reduction and H2 Generation in Neutral and Weakly Acidic Media. The Journal of Physical Chemistry C 2022, 126 (17) , 7456-7467. https://doi.org/10.1021/acs.jpcc.2c01504
  3. Abhijit Dutta, Iván Zelocualtecatl Montiel, Kiran Kiran, Alain Rieder, Vitali Grozovski, Lukas Gut, Peter Broekmann. A Tandem (Bi2O3 → Bimet) Catalyst for Highly Efficient ec-CO2 Conversion into Formate: Operando Raman Spectroscopic Evidence for a Reaction Pathway Change. ACS Catalysis 2021, 11 (9) , 4988-5003. https://doi.org/10.1021/acscatal.0c05317
  4. Yu Wang, Tairan Wang, Shuyu Bu, Jiaxiong Zhu, Yanbo Wang, Rong Zhang, Hu Hong, Wenjun Zhang, Jun Fan, Chunyi Zhi. Sulfolane-containing aqueous electrolyte solutions for producing efficient ampere-hour-level zinc metal battery pouch cells. Nature Communications 2023, 14 (1) https://doi.org/10.1038/s41467-023-37524-7
  5. Tamás Pajkossy. Past and present of electrochemical science in Hungary. Journal of Solid State Electrochemistry 2023, 27 (7) , 1747-1754. https://doi.org/10.1007/s10008-023-05410-3
  6. Guoping Gao, Lin-Wang Wang. A potential and pH inclusive microkinetic model for hydrogen reactions on Pt surface. Chem Catalysis 2021, 1 (6) , 1331-1345. https://doi.org/10.1016/j.checat.2021.10.006
  7. Karolina Chat-Wilk, Ewa Rudnik, Grzegorz Włoch, Piotr Osuch. Importance of anions in electrodeposition of nickel from gluconate solutions. Ionics 2021, 27 (10) , 4393-4408. https://doi.org/10.1007/s11581-021-04166-y
  8. Soma Vesztergom, Abhijit Dutta, Motiar Rahaman, Kiran Kiran, Iván Zelocualtecatl Montiel, Peter Broekmann. Hydrogen Bubble Templated Metal Foams as Efficient Catalysts of CO 2 Electroreduction. ChemCatChem 2021, 13 (4) , 1039-1058. https://doi.org/10.1002/cctc.202001145
  9. Noémi Kovács, Vitali Grozovski, Pavel Moreno-García, Peter Broekmann, Soma Vesztergom. A Model for the Faradaic Efficiency of Base Metal Electrodeposition from Mildly Acidic Baths to Rotating Disk Electrodes. Journal of The Electrochemical Society 2020, 167 (10) , 102510. https://doi.org/10.1149/1945-7111/ab975b
  10. Bibhudatta Nanda, Manila Mallik. Production of Copper Powder by Electrodeposition with Different Equilibrium Crystal Shape. Transactions of the Indian Institute of Metals 2020, 73 (8) , 2113-2119. https://doi.org/10.1007/s12666-020-02015-6
  • Abstract

    Figure 1

    Figure 1. Polarization curve of an RDE, showing two hydrogen evolution steps with an intermittent limiting current plateau section.

    Figure 2

    Figure 2. Reaction schemes for HER. Two separate reactions are shown in (a) for acidic and neutral to alkaline solutions. A combined scheme is shown in (b) for near-neutral solutions. The autoprotolysis of water is part of both schemes.

    Figure 3

    Figure 3. Experimental protocol for the measurement of HER polarization curves on RDEs. The stationary current of the RDE is measured at given electrode potential (E) and rotation rate (f) values (green periods). Between the measurements, the potential control is switched off and the RDE rotated quickly, to remove accumulated bubbles (red periods).

    Figure 4

    Figure 4. Experimentally obtained polarization curves (red dots) on an Au (a) and on a Pt (b) RDE, showing a two step behavior. In each panel (at different values of pH), the cathodic current density increases as the rotation rate f is set to values of 400, 625, 900, 1225, 1600, 2025, and 2500 min–1. The green curves are created by fitting the model described by eq 23 globally, that is, for all pH and rotation rate values, and by optimizing only three parameters (αc, k, and ). Determined confidence intervals (at 95% statistical certainty) for the fitted parameters are αc = 0.486 ± 0.067, and for gold and αc = 0.643 ± 0.037, and for platinum. Other parameter values (not optimized) are shown in Table 1.

    Figure 5

    Figure 5. Polarization curve (thick gray) calculated using eq 23 and the parameter values of Table 1, shown in two different representations: with linear axis scaling in (a) and on a Tafel plot in (b). The three different segments of the curves, marked by the dashed lines, can be approximated by eqs 2430, as discussed in the text. The contour map in the background of (a) shows the variation of pH as a function of the distance measured from the electrode surface at each disk potential. (Details of calculating pH profiles using the presented model are discussed later, cf. to Figure 9).

    Figure 6

    Figure 6. Exchange current densities j0 in the studied pH range, calculated by using eq 25, based on the kinetic parameters determined by nonlinear fitting in Figure 4 for gold and platinum.

    Figure 7

    Figure 7. Effect of varying the parameters αc (a, b) and k (c, d) on the calculated polarization curves. Values assumed are shown on the graph; other parameters are given in Table 1. Polarization curves are shown in two different representations: with linear axis scaling (a, c) and on a Tafel plot (b, d).

    Figure 8

    Figure 8. (a) Concept of the breakdown overpotential ηbr illustrated on a polarization curve. (b) Dimensionless breakdown overpotentials plotted as a function of pH for gold and platinum, determined from measured data (dots). The lines were created by linear fitting to the measured data, using a pH2 weighting. The acquired slopes and intercepts were used according to eq 32 to calculate the kinetic parameters shown in Table 2. Chosen rotation rate: 625 min–1.

    Figure 9

    Figure 9. (a) Normalized concentration difference profiles as a function of the normalized distance, for some chosen values of the normalized current density (shown in the figure). (b) An example for pH profiles calculated for various normalized current values, assuming that pH = 3.

    Figure 10

    Figure 10. Full black curve: the distance of neutrality (normalized to the diffusion layer thickness δN) as a function of the current density normalized to . Note that the function, eq 35, is not defined for . Dashed gray curve: an estimate for the neutrality distance based on an analytical solution assuming that DOHDH+. (27)

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 41 other publications.

    1. 1
      Wang, H.; Gao, L. Recent Developments in Electrochemical Hydrogen Evolution Reaction. Curr. Opin. Electrochem. 2018, 7, 714,  DOI: 10.1016/j.coelec.2017.10.010
    2. 2
      Ooka, H.; Figueiredo, M. C.; Koper, M. T. M. Competition between Hydrogen Evolution and Carbon Dioxide Reduction on Copper Electrodes in Mildly Acidic Media. Langmuir 2017, 33, 93079313,  DOI: 10.1021/acs.langmuir.7b00696
    3. 3
      Cave, E. R.; Shi, C.; Kuhl, K. P.; Hatsukade, T.; Abram, D. N.; Hahn, C.; Chan, K.; Jaramillo, T. F. Trends in the Catalytic Activity of Hydrogen Evolution during CO2 Electroreduction on Transition Metals. ACS Catal. 2018, 8, 30353040,  DOI: 10.1021/acscatal.7b03807
    4. 4
      Schlesinger, M.; Paunovic, M., Eds., Modern Electroplating, 5th ed.; Wiley: New York, 2010.
    5. 5
      Ritzert, N. L.; Moffat, T. P. Ultramicroelectrode Studies of Self-Terminated Nickel Electrodeposition and Nickel Hydroxide Formation upon Water Reduction. J. Phys. Chem. C 2016, 120, 2747827489,  DOI: 10.1021/acs.jpcc.6b10006
    6. 6
      Wu, J.; Wafula, F.; Branagan, S.; Suzuki, H.; van Eisden, J. Mechanism of Cobalt Bottom-Up Filling for Advanced Node Interconnect Metallization. J. Electrochem. Soc. 2019, 166, D3136D3144,  DOI: 10.1149/2.0161901jes
    7. 7
      Horányi, G. Electrochemical Dictionary;Bard, A. J.; Scholz, F.; Inzelt, G., Eds.; Springer Verlag: Heidelberg, 2008; p 343.
    8. 8
      Bockris, J. O. M.; Ammar, I. A.; Huq, A. K. M. S. The Mechanism of the Hydrogen Evolution Reaction on Platinum, Silver and Tungsten surfaces in Acid Solutions. J. Phys. Chem. A. 1957, 61, 879886,  DOI: 10.1021/j150553a008
    9. 9
      Bagotzky, V. S.; Osetrova, N. V. Investigations of Hydrogen Ionization on Platinum with the Help of Micro-Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1973, 43, 233249,  DOI: 10.1016/S0022-0728(73)80494-2
    10. 10
      Tavares, M. C.; Machado, S. A. S.; Mazo, L. H. Study of Hydrogen Evolution Reaction in Acid Medium on Pt Microelectrodes. Electrochim. Acta 2001, 46, 43594369,  DOI: 10.1016/S0013-4686(01)00726-5
    11. 11
      Neyerlin, K. C.; Gu, W.; Jorne, J.; Gasteiger, H. A. Study of the Exchange Current Density for the Hydrogen Oxidation and Evolution Reactions. J. Electrochem. Soc. 2007, 154, B631B635,  DOI: 10.1149/1.2733987
    12. 12
      Sheng, W.; Gasteiger, H. A.; Shao-Horn, Y. Hydrogen Oxidation and Evolution Reaction Kinetics on Platinum: Acid vs Alkaline Electrolytes. J. Electrochem. Soc. 2010, 157, B1529B1536,  DOI: 10.1149/1.3483106
    13. 13
      Durst, J.; Siebel, A.; Simon, C.; Hasché, F.; Herranz, J.; Gasteiger, H. A. New insights into the electrochemical hydrogen oxidation and evolution reaction mechanism. Energy Environ. Sci. 2014, 7, 22552260,  DOI: 10.1039/C4EE00440J
    14. 14
      Zheng, J.; Sheng, W.; Zhuang, Z.; Xu, B.; Yan, Y. Universal Dependence of Hydrogen Oxidation and Evolution Reaction Activity of Platinum-group Metals on pH and Hydrogen Binding Energy. Sci. Adv. 2016, 2, e1501602  DOI: 10.1126/sciadv.1501602
    15. 15
      Pentland, N.; Bockris, J. O. M.; Sheldon, E. Hydrogen Evolution Reaction on Copper, Gold, Molybdenum, Palladium, Rhodium, and Iron. J. Electrochem. Soc. 1957, 104, 182,  DOI: 10.1149/1.2428530
    16. 16
      Ives, D. J. G. Some Abnormal Hydrogen Electrode Reactions. Can. J. Chem. 1959, 37, 213221,  DOI: 10.1139/v59-028
    17. 17
      Brug, G. J.; Sluyters-Rehbach, M.; Sluyters, J. H.; Hemelin, A. The Kinetics of the Reduction of Protons at Polycrystalline and Monocrystalline Gold Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1984, 181, 245266,  DOI: 10.1016/0368-1874(84)83633-3
    18. 18
      Conway, B. E.; Bai, L. State of Adsorption and Coverage by Overpotential-Deposited H in the H2 Evolution Reaction at Au and Pt. Electrochim. Acta 1986, 31, 10131024,  DOI: 10.1016/0013-4686(86)80017-2
    19. 19
      Khanova, L. A.; Krishtalik, L. I. Kinetics of the hydrogen evolution reaction on gold electrode. A new case of the barrierless discharge. J. Electroanal. Chem. 2011, 660, 224229,  DOI: 10.1016/j.jelechem.2011.01.016
    20. 20
      Kahyarian, A.; Brown, B.; Nesić, S. Mechanism of the Hydrogen Evolution Reaction in Mildly Acidic Environments on Gold. J. Electrochem. Soc. 2017, 164, H365H374,  DOI: 10.1149/2.1061706jes
    21. 21
      Obata, K.; Stegenburga, L.; Takanabe, K. Maximizing Hydrogen Evolution Performance on Pt in Buffered Solutions: Mass Transfer Constrains of H2 and Buffer Ions. J. Phys. Chem. C 2019, 123, 2155421563,  DOI: 10.1021/acs.jpcc.9b05245
    22. 22
      Hessami, S.; Tobias, C. W. In-Situ Measurement of Interfacial pH Using a Rotating Ring-Disk Electrode. AIChE J. 1993, 39, 149162,  DOI: 10.1002/aic.690390115
    23. 23
      Katsounaros, I.; Meier, J. C.; Klemm, S. O.; Topalov, A. A.; Biedermann, U. P.; Auinger, M.; Mayrhofer, K. J. J. The Effective Surface pH during Reactions at the Solid-Liquid interface. Electrochem. Commun. 2011, 13, 634637,  DOI: 10.1016/j.elecom.2011.03.032
    24. 24
      Auinger, M.; Katsounaros, J. C.; Meier, I.; Klemm, S. O.; Biedermann, P. U.; Topalov, A. A.; Rohwerder, M.; Mayrhofer, K. J. J. Near-surface ion distribution and buffer effects during electrochemical reactions. Phys. Chem. Chem. Phys. 2011, 13, 1638416394,  DOI: 10.1039/c1cp21717h
    25. 25
      Shinagawa, T.; García-Esparza, A. T.; Takanabe, K. Insight on Tafel Slopes from a Microkinetic Analysis of Aqueous Electrocatalysis for Energy Conversion. Sci. Rep. 2015, 5, 13801  DOI: 10.1038/srep13801
    26. 26
      Strmcnik, D.; Uchimura, M.; Wang, C.; Subbaraman, R.; Danilovic, N.; van der Vliet, D.; Paulikas, A. P.; Stamenkovic, V. R.; Markovic, N. M. Improving the Hydrogen Oxidation Reaction Rate by Promotion of Hydroxyl Adsorption. Nat. Chem. 2013, 5, 300306,  DOI: 10.1038/nchem.1574
    27. 27
      Grozovski, V.; Vesztergom, S.; Láng, G. G.; Broekmann, P. Electrochemical Hydrogen Evolution: H+ or H2O Reduction? A Rotating Disk Electrode Study. J. Electrochem. Soc. 2017, 164, E3171E3178,  DOI: 10.1149/2.0191711jes
    28. 28
      Vesztergom, S. Encyclopedia of Interfacial Chemistry: Surface Science and Electrochemistry;Wandelt, K., Ed.; Elsevier: Amsterdam, 2018; pp 421444.
    29. 29
      Nagel, K.; Wendler, F. Die Wasserstoffelektrode als zweifache Elektrode. Z. Elektrochem. 1956, 60, 10641072,  DOI: 10.1002/bbpc.19560600924
    30. 30
      Kanzaki, Y.; Tokuda, K.; Bruckenstein, S. Dissociation Rates of Weak Acids Using Sinusoidal Hydrodynamic Modulated Rotating Disk Electrode Employing Koutecky-Levich Equation. J. Electrochem. Soc. 2014, 161, H770H779,  DOI: 10.1149/2.0221412jes
    31. 31
      Wiberg, G. K. H.; Arenz, M. On the Influence of Hydronium and Hydroxide Ion Diffusion on the Hydrogen and Oxygen Evolution Reactions in Aqueous Media. Electrochim. Acta 2015, 159, 6670,  DOI: 10.1016/j.electacta.2015.01.098
    32. 32
      Carneiro-Neto, E. B.; Lopes, M. C.; Pereira, E. C. Simulation of Interfacial pH Changes during Hydrogen Evolution Reaction. J. Electroanal. Chem. 2016, 765, 9299,  DOI: 10.1016/j.jelechem.2015.09.029
    33. 33
      Vanýsek, P. CRC Handbook of Chemistry and Physics, 93rd ed.; Haynes, W. M., Ed.; Chemical Rubber Company: Boca Raton FL, 2012.
    34. 34
      Guidelli, R.; Compton, R. G.; Feliu, J. M.; Gileadi, E.; Lipkowski, J.; Schmickler, W.; Trasatti, S. Defining the Transfer Coefficient in Electrochemistry: An Assessment (IUPAC Technical Report). Pure Appl. Chem. 2014, 86, 245258,  DOI: 10.1515/pac-2014-5026
    35. 35
      Bard, A. J.; Faulkner, L. R. Electrochemical Methods. Fundamentals and Applications; John Wiley & Sons: New York, 2001.
    36. 36
      Bronstein, I. N.; Semendjajew, K. A.; Musiol, G.; Muehlig, H. Taschenbuch der Mathematik; Verlag Harri Deutsch: Frankfurt am Main, 2008.
    37. 37
      Haghighat, S.; Dawlaty, J. M. pH Dependence of the Electron-Transfer Coefficient: Comparing a Model to Experiment for Hydrogen Evolution Reaction. J. Phys. Chem. C 2016, 120, 2848928496,  DOI: 10.1021/acs.jpcc.6b10602
    38. 38
      Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Kitchin, J. R.; Chen, J. G.; Pandelov, S.; Stimming, U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23J26,  DOI: 10.1149/1.1856988
    39. 39
      Koutecky, J.; Levich, V. G. The use of a rotating disk electrode in the studies of electrochemical kinetics and electrolytic processes. Zh. Fiz. Khim. 1958, 32, 15651575
    40. 40
      Cheng, A. H.-D.; Cheng, D. T. Heritage and early history of the boundary element method. Eng. Anal. Boundary Elem. 2005, 29, 268302,  DOI: 10.1016/j.enganabound.2004.12.001
    41. 41
      Kucernak, A. R.; Zalitis, C. General Models for the Electrochemical Hydrogen Oxidation and Hydrogen Evolution Reactions: Theoretical Derivation and Experimental Results under Near Mass-Transport Free Conditions. J. Phys. Chem. C 2016, 120, 1072110745,  DOI: 10.1021/acs.jpcc.6b00011

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect