Understanding the charge transfer processes at solid oxide fuel cell (SOFC) electrodes is critical to designing more efficient and robust materials. Activation losses at SOFC electrodes have been widely attributed to the ambipolar migration of charges at the mixed ionic–electronic conductor–gas interface. Empirical Butler–Volmer kinetics based on the transition state theory is often used to model the current–voltage relationship, where charged particles transfer classically over an energy barrier. However, the hydrogen oxidation/water electrolysis reaction H2(g) + O2− ⇌ H2O(g) + 2e− must be modeled through concerted electron and proton tunneling events, where we unify the theory of the electrostatic surface potential with proton-coupled electron transfer kinetics. We derive a framework for the reaction rate that depends on the electrostatic surface potential, adsorbate dipole moment, the electronic structure of the electron donor/acceptor, and vibronic states of the hydrogen species. This theory was used to study the current–voltage characteristics of the Ni/gadolinium-doped ceria electrode in H2/H2O(g), where we find excellent validation of this novel model. These results yield the first reported quantification of the solvent reorganization energy for an SOFC material and suggest that the three-phase boundary mechanism is the dominant pathway for charge transfer at cermet electrodes.
I. INTRODUCTION
The solid oxide fuel cell (SOFC) is a highly efficient chemical-to-electrical energy conversion technology compatible with both existing fuel (e.g., natural gas) and future fuel (e.g., renewably sourced hydrogen) infrastructures.1–3 These devices work through electrochemical RedOx reactions at the air and fuel electrodes, whereby chemical energy is used to drive external work. The Faradaic reactions at the fuel electrode can be given simply as H2(g) + O2− ⇌ H2O(g) + 2e− (elementary steps are listed in Table I).4,5 The hydrogen oxidation/water hydrolysis reaction may occur via the two-phase boundary (2 PB) [Fig. 1(a)] or the three-phase boundary (3 PB) [Fig. 1(b)]. At the 2 PB, two neighboring polarons in the CeO2 surface will combine with two neighboring protons. This process is driven by a combination of the electrostatic potential and the concentration of electronic defects at the interface between mixed ionic–electronic conductors (MIEC) and the gas phase.2 Alternatively, in the presence of an electrocatalytic metallic phase, the electron–proton recombination may occur via a 3 PB mechanism. Electron density migrates from a polaronic state in the oxide phase to the metal; meanwhile, the proton is reduced as it migrates to the metal surface. The electrostatic nature of the 3 PB process is less well studied.
Elementary steps for the two-phase boundary (2 PB) and three-phase boundary (3 PB) reaction mechanisms written in the Kröger–Vink notation. For water adsorption, H2O(g) consumes an oxygen vacancy and an oxygen site , forming two hydroxyls on the CGO surface. This is followed by the rate limiting PCET step. At the 2 PB, two hydroxyls and two polarons form H2(g) at the surface of the MIEC. At the 3 PB, one hydroxyl and one polaron combine at a free site on the nickel surface (), forming an adsorbed hydrogen on the nickel surface (HNi). At the 3 PB, there is an additional hydrogen desorption step.
Elementary step . | Two-phase boundary . | Three-phase boundary . |
---|---|---|
Water adsorption | ||
PCET | ||
Hydrogen desorption | 2HNi ⇌ H2(g) + 2* |
Elementary step . | Two-phase boundary . | Three-phase boundary . |
---|---|---|
Water adsorption | ||
PCET | ||
Hydrogen desorption | 2HNi ⇌ H2(g) + 2* |
Schematic illustration of the charge transfer step of the water hydrolysis reaction at the Ni/CGO surface. (a) Via 2 PB pathway, where two protons and two polarons combine at the surface of the MIEC. (b) 3 PB pathway, where the polaron migrates into the metal phase and the proton is reduced at the metal surface forming a neutral hydrogen state. Red spheres and purple isosurfaces represent protons and polarons, respectively. The purple lines represent the electric field created by the intrinsic dipole moment of the surface hydroxyl. PT and ET represent proton transfer and electron transfer processes, respectively.
Schematic illustration of the charge transfer step of the water hydrolysis reaction at the Ni/CGO surface. (a) Via 2 PB pathway, where two protons and two polarons combine at the surface of the MIEC. (b) 3 PB pathway, where the polaron migrates into the metal phase and the proton is reduced at the metal surface forming a neutral hydrogen state. Red spheres and purple isosurfaces represent protons and polarons, respectively. The purple lines represent the electric field created by the intrinsic dipole moment of the surface hydroxyl. PT and ET represent proton transfer and electron transfer processes, respectively.
Gadolinium-doped ceria (CGO) displays considerably fast oxygen transport kinetics and high electronic conductivity under reducing conditions.11–15 To enhance the overall electronic conductivity of the electrode and to increase the density of reactive 3 PB, it is beneficial to mix a metallic phase with the MIEC to create a cermet electrode, i.e., nickel/gadolinium-doped ceria (Ni/CGO).1,2,6,16–19 The operation of the Ni/CGO electrode under electrical bias has been studied for decades; however, a unifying model for hydrogen electro-oxidation or water electrolysis has not yet been agreed upon.18,20–30 Most charge transfer models utilize the empirical Butler–Volmer equation, which is convenient for fitting current–voltage data at low overpotentials.31,32 However, the Butler–Volmer equation suggests that the electrostatic potential driving the charge transfer reaction is located at the electrode–electrolyte interface. In our previous work, we have illustrated that the electrostatic potential modulated by the state of the electrode is located at the surface of the MIEC.1,2 The driving force for charge transfer should therefore be determined by the electrostatic potential at the MIEC–gas interface.
Herein, we build on these previous findings to model the kinetics of electron and proton transfer at MIEC–gas interfaces. By accounting for nuclear quantum effects of the transferring proton, the charge transfer at SOFC electrodes can be described within a concerted proton-coupled electron transfer (PCET) framework. Therefore, in this study, we aim to combine the theory of the electrostatic surface potential with the vibronically nonadiabatic PCET theory to develop a unifying kinetic model for charge transfer at MIEC–gas interfaces.1,2,33–36 This model is adapted to both 2 PB and 3 PB rate expressions that are fit to experimental data to discern the PCET mechanism at MIEC–gas interfaces.
II. THEORY
A. Thermodynamics
B. Electrode kinetics
C. Electron transfer
III. PROTON-COUPLED ELECTRON TRANSFER
By treating the proton quantum mechanically, we derive an expression for rates at the MIEC–gas interface in the framework of proton-coupled electron transfer (PCET). We will begin this section by summarizing the general theory of PCET through the derivation of vibronic states and charge transfer mechanisms. After which, we will look at where PCET is applied to electrochemistry specifically before integrating PCET into the framework of the MIEC–gas interface.
A. Fundamental concepts of PCET theory
PCET reactions are prevalent in many biological, chemical, and electrochemical processes.34–37,57,62–70 A general reaction square scheme for heterogeneous electrochemical PCET involving the transfer of one electron and one proton is illustrated in Figs. 2(a)–2(c).67 Because the transferring electron and proton are treated quantum mechanically within this theoretical treatment, such a PCET reaction can be described in terms of the four diabatic states shown on the corners of the surface. As Figs. 2(a)–2(c) imply, the net PCET mechanism could be sequential with initial electron transfer (ET) or proton transfer (PT) producing a thermodynamically stable intermediate, or concerted, without forming such an intermediate.57
(a)–(c), Reaction scheme illustrating the heterogeneous proton-coupled electron transfer process, where the electrode participates directly in forming and breaking chemical bonds. H+, e−, and M* represent a proton, electron, and free metal electrode surface site, respectively, H represents the hydrogen atom, MH+ represents a proton adsorbed onto the metal surface, and MH represents a hydrogen adsorbed onto the metal surface. The contour map details the excess chemical potential energy landscape for (a) ET limited reduction, (b) PCET limited reduction, and (c) PT limited reduction. The planes represent the ET reaction coordinates and the PT reaction coordinate.39 The axis in white plots represent the 1D chemical potential landscape explored by overall PCET reaction. RET is the electron transfer (ET) step, RPT is the PT step, and RPCET is the PCET step. (d) Schematic illustration of the excess chemical potential landscape explored by the PCET process. The left-hand panel illustrates the reactant proton potential with its corresponding proton vibrational wave functions along the proton coordinate, while the right-hand panel illustrates the product proton potential with its corresponding proton vibrational wave functions. The middle panel shows a set of stacked PCET Marcus parabolas corresponding to the oxidized (blue) and reduced (red) diabatic states of the reaction. The splitting between the stacked free energy curves corresponds to energy level splitting between the sets of reactant and product vibronic states (μ and ν, respectively). represents the excess free energy difference between the ground vibronic states of the reactants and products, and λ represents the reorganization energy.
(a)–(c), Reaction scheme illustrating the heterogeneous proton-coupled electron transfer process, where the electrode participates directly in forming and breaking chemical bonds. H+, e−, and M* represent a proton, electron, and free metal electrode surface site, respectively, H represents the hydrogen atom, MH+ represents a proton adsorbed onto the metal surface, and MH represents a hydrogen adsorbed onto the metal surface. The contour map details the excess chemical potential energy landscape for (a) ET limited reduction, (b) PCET limited reduction, and (c) PT limited reduction. The planes represent the ET reaction coordinates and the PT reaction coordinate.39 The axis in white plots represent the 1D chemical potential landscape explored by overall PCET reaction. RET is the electron transfer (ET) step, RPT is the PT step, and RPCET is the PCET step. (d) Schematic illustration of the excess chemical potential landscape explored by the PCET process. The left-hand panel illustrates the reactant proton potential with its corresponding proton vibrational wave functions along the proton coordinate, while the right-hand panel illustrates the product proton potential with its corresponding proton vibrational wave functions. The middle panel shows a set of stacked PCET Marcus parabolas corresponding to the oxidized (blue) and reduced (red) diabatic states of the reaction. The splitting between the stacked free energy curves corresponds to energy level splitting between the sets of reactant and product vibronic states (μ and ν, respectively). represents the excess free energy difference between the ground vibronic states of the reactants and products, and λ represents the reorganization energy.
B. PCET at electrochemical interfaces
C. PCET at the CeO2–H2/H2O interface
(a) Polaron coverage, (b) hydroxyl coverage, and (c) electrostatic surface potential shift on a Sm0.2Ce0.8O1.9–gas interface as a function of reservoir potential at 500 °C in 1:8:4 H2O:H2:Ar. Gray circles represent data collected by APXPS measurements (Feng et al).10 Polaron coverage is fit to Eq. (43); hydroxyl coverage and electrostatic surface potential shift is fit to Eq. (42).1 (d) Current density as a function of overpotential for the Ni/CGO electrode in full cell (0.5 bar H2, 0.5 bar H2O, 600 °C). Gray circles represent experimentally observed current densities derived from electrochemical impedance spectroscopy (Williams et al.),78 black lines represent the Butler–Volmer model given by Eq. (14) (RMSE = 0.166), red lines represent the 2PB-PCET model given by Eq. (44) (RMSE = 0.213), and blue lines represent the 3PB-PCET model given by Eq. (47) (RMSE = 0.074).
(a) Polaron coverage, (b) hydroxyl coverage, and (c) electrostatic surface potential shift on a Sm0.2Ce0.8O1.9–gas interface as a function of reservoir potential at 500 °C in 1:8:4 H2O:H2:Ar. Gray circles represent data collected by APXPS measurements (Feng et al).10 Polaron coverage is fit to Eq. (43); hydroxyl coverage and electrostatic surface potential shift is fit to Eq. (42).1 (d) Current density as a function of overpotential for the Ni/CGO electrode in full cell (0.5 bar H2, 0.5 bar H2O, 600 °C). Gray circles represent experimentally observed current densities derived from electrochemical impedance spectroscopy (Williams et al.),78 black lines represent the Butler–Volmer model given by Eq. (14) (RMSE = 0.166), red lines represent the 2PB-PCET model given by Eq. (44) (RMSE = 0.213), and blue lines represent the 3PB-PCET model given by Eq. (47) (RMSE = 0.074).
The transition states requires a Ce atom to occupy the cation site at the 3 PB. Therefore, we can express the activity coefficient of the transition state as γTS = (1 −x)−1, where x is the concentration of aliovalent dopants since the dopants cannot participate in RedOx reactions.39 When conditioning Eq. (47) to the electrochemical impedance spectroscopy data, we find an excellent fit in Fig. 3(d). When compared to the Butler–Volmer equation [Eq. (14)] and 2PB-PCET equation [Eq. (44)], we find that the 3PB-PCET model fits the experimental data far better. This suggests that the mechanism involves PCET at the 3 PB and that the kinetics of this PCET reaction can be accurately captured using a nonadiabatic model. It is important to note that the vibronic coupling and Boltzmann population are difficult to calculate. Goldsmith et al. illustrated limited change in the contribution from each vibronic coupling relative to the exponentially scaled electrostatic component.75 Therefore, we assume that we do not observe their effects in the current–voltage experiment. From this fitting, we were able to calculate the reorganization energy λ = 18kBT (1.35 eV at 873 K). This is large relative to that estimated for lithium ion intercalation into LiFePO4 and graphite .31,76 DFT calculations carried out by Castleton et al. approximated the diabatic barrier for polaron migration in bulk CeO2−δ.77 The result that agreed closest to our system was the local density approximation + U (LDA + U), which yielded λLDA+U = 1.28 eV. However, there was a wide range of results as the barrier was heavily dependent on the functional used; moreover, the polaron migration process is not entirely representative of the PCET process.
Interesting, the electrostatic driving force in Eqs. (47) and (44) is the electrostatic surface potential located at the MIEC–gas interface, even if the PCET event occurs at the 3 PB.2 The origin of this phenomenon is rooted in the dipole–dipole interactions, whereby the coverage of adsorbates at the 3 PB will influence the adsorption thermodynamics across the MIEC–gas interface. Under water electrolysis mode, PCET is driven by the charged particles (adsorbates and polarons) seeking to establish thermodynamic equilibrium with the gas phase, resulting in the release of hydrogen, relaxation of the electrostatic surface potential, and a spatial region free of charged particles. When PCET occurs at the 2 PB, a charge-free region will be filled as the charged particles reorganize to maximize configurational entropy and minimize columbic interactions. The same principle applies to the 3 PB mechanism, where a charge-free region close to the 3 PB will stimulate reorganization across the entire MIEC surface, resulting in the relaxation of the electrostatic surface potential at the CeO2–H2/H2O interface. We can account for the asymmetry observed in the current–voltage plot in Fig. 3(d) to dipole–dipole depolarization. This occurs when a dipole is exposed to the electrostatic field of all other dipoles, shrinking in the hydroxyl bond length, thereby reducing the size of the electric field normal to the surface. Dipole–dipole depolarization increased at negative overpotential as the coverage of hydroxyls on the electrode surface increases [Fig. 3(b)].1,2,16 A decrease in the adsorbate dipole moment reduces the magnitude of the surface potential shift and therefore increases the resistance of the PCET process at negative overpotentials relative to positive overpotentials.
IV. CONCLUSION
In this work, the origin and physics of the activation overpotential for the Ni/CGO fuel electrode has been explained. Previously, we demonstrated the effects of nonequilibrium thermodynamics on the concentration of electronic defects and the adsorbate induced electrostatic surface potential. Here, we derive a kinetic relationship between the overpotential and surface potential via the 2 and 3 PB mechanism. In the framework of the proton-coupled electron transfer (PCET) theory, electronic and protonic species were treated as quantum particles whereby the probability of a concerted tunneling event was controlled by an external voltage. To account for the electronic structure of the metallic phase, variation in the density of states was neglected to give an analytical expression for the hydrolysis reaction via the 3 PB. Using the current–voltage data, we found the best fit was given by the 3 PB PCET equation vs the Butler–Volmer equation and the 2 PB PCET equation. This suggested that the dominant pathway for charge transfer at the Ni/CGO electrode includes the electrocatalytic metallic phase. Understanding the mechanism of ambipolar charge transfer at the electrode–gas interface is a significant consideration for the design and operational conditions of SOC electrodes as the strength of the intrinsic dipole moment of the adsorbed gas species is shown to have a profound effect on the magnitude of the activation overpotential. The theory of the electrostatic surface potential put forward in this work is not confined to PCET processes at SOFC electrodes and can be extended to other fields where charge transfer occurs at a solid–gas interface, such as nitrogen reduction and in lithium air batteries.
SUPPLEMENTARY MATERIAL
Supplementary material notes: derivations of the electrostatic surface potential, affinity of the water electrolysis reaction, and the de Donder relation.
ACKNOWLEDGMENTS
This work was supported by Ceres Power Ltd. I.D.S. and S.J.S. acknowledge the EPSRC for funding through the award of Grant No. EP/R002010/1. R.E.W. acknowledges start-up funding from the Case School of Engineering at Case Western Reserve University.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Nicholas J. Williams: Conceptualization (lead); Data curation (lead); Methodology (lead); Writing – original draft (lead). Robert E. Warburton: Formal analysis (supporting); Validation (supporting); Writing – review & editing (supporting). Ieuan D. Seymour: Visualization (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Alexander E. Cohen: Visualization (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Martin Z. Bazant: Supervision (equal); Writing – review & editing (supporting). Stephen J. Skinner: Supervision (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article and its supplementary material.