Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Chen Et Al 2016 Electric Field Effects in Electrochemical Co2 Reduction

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Letter

pubs.acs.org/acscatalysis

Electric Field Effects in Electrochemical CO2 Reduction


Leanne D. Chen,†,‡ Makoto Urushihara,†,§ Karen Chan,†,‡ and Jens K. Nørskov*,†,‡

SUNCAT Center for Interface Science and Catalysis, Department of Chemical Engineering, Stanford University, Stanford, California
94305−5025, United States

SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, Menlo Park, California 94025, United
States
§
Central Research Institute, Mitsubishi Materials Corporation, 1002−14 Mukohyama, Naka, Ibaraki 311−0102, Japan
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via INDIAN INST OF TECH HYDERABAD on July 18, 2024 at 17:51:24 (UTC).

ABSTRACT: Electrochemical reduction of CO2 has the potential to reduce


greenhouse gas emissions while providing energy storage and producing chemical
feedstocks. A mechanistic understanding of the process is crucial to the discovery of
efficient catalysts, and an atomistic description of the electrochemical interface is a
major challenge due to its complexity. Here, we examine the CO2 → CO
electrocatalytic pathway on Ag(111) using density functional theory (DFT)
calculations and an explicit model of the electrochemical interface. We show that the
electric field from solvated cations in the double layer and their corresponding
image charges on the metal surface significantly stabilizes key intermediates*CO2 and *COOH. At the field-stabilized sites, the
formation of *CO is rate-determining. We present a microkinetic model that incorporates field effects and electrochemical
barriers from ab initio calculations. The computed polarization curves show reasonable agreement with experiment without fitting
any parameters.
KEYWORDS: CO2 reduction, field effects, density functional theory

fficient electrochemical reduction of carbon dioxide (CO2)


E would simultaneously mitigate its effect on the climate,
provide a sustainable, carbon-neutral means to store chemical
In this Letter, we use density functional theory (DFT) and
an explicit model of the electrochemical interface to investigate
the energetics of the CO2 → CO pathway on Ag(111) and
energy, and produce feedstocks of carbon-containing molecules compare this to vacuum calculations using the CHE model. We
for the chemical industry.1 One of the simplest reduced show that, under realistic electrochemical conditions, the
products that can be made from CO2 is carbon monoxide adsorbed CO2 intermediate is significantly stabilized by local
(CO), which requires only two proton−electron transfers. electric fields exerted by solvated cations and their correspond-
ing image charges at the electrode | electrolyte interface. We
Noble metals such as silver and gold have been shown to be the
develop a microkinetic model that incorporates electric field
best catalysts for this process. However, the process still effects and electrochemical barriers determined through ab
requires a sizable overpotential and it has been hypothesized initio calculations, and show that this model accounts
that the mechanism is limited by the endergonic reaction reasonably for the experimentally observed polarization
energy of the first intermediate: adsorbed CO2.2−6 Several curves.7,8
experimental studies have also shown a pH effect, in which the We study the following pathway:
overpotential for this reaction at pH 14 is lower than that at pH
CO2 (g) + * → *CO2 (1)
7.7−9
A mechanistic understanding of the CO2 reduction process is *CO2 + H 2O + e− → *COOH + OH− (2)
crucial to the design of more active catalysts. An ab initio
− −
investigation of the kinetics is very challenging due to the *COOH + H 2O + e → *CO + H 2O + OH (3)
complexity of the electrochemical interface, where an accurate
description of all components (ions, solvent, interfacial charge, *CO → CO(g) + * (4)
and potential) is required.10−14 Recent developments in where a lone asterisk (*) denotes a surface site and an * symbol
methodology have allowed the study of charged surfaces,15,16 before a molecule denotes an adsorbed species.
and many approaches have been henceforth used in modeling Reaction 1 is sometimes written as an electron transfer step,
electrochemical reactions. These include the computational since the adsorption of CO2 induces a polarization of the
hydrogen electrode (CHE),17 the surface charging (SC)
method,18,19 the linearized Poisson−Boltzmann (PB) equa- Received: August 11, 2016
tion,19,20 and the self-consistent continuum solvation (SCCS) Revised: September 9, 2016
scheme.21,22 Published: September 22, 2016

© 2016 American Chemical Society 7133 DOI: 10.1021/acscatal.6b02299


ACS Catal. 2016, 6, 7133−7139
ACS Catalysis Letter

electron density, such that the adsorbed molecule becomes Figure 2a shows the calculated free-energy diagram at pH 0
partially negatively charged. However, if this were a true for all intermediates in the CO2 → CO pathway on Ag(111)
electron transfer step, the reaction energy should be dependent and Pt(111) at the thermodynamic equilibrium potential (U0 =
on the work function (or electrochemical potential) of the −0.11 V at pH 0)6 versus the reversible hydrogen electrode
surface, and we find that not to be the case. This is because (RHE). These energies were obtained under vacuum at zero
there is hybridization between the CO2 and metallic states electric field with the appropriate solvation corrections. The
during the adsorption process, such that the charge transfer is effect of applied potential are included through the CHE
not directly a function of the potential. We show that by model, where the chemical potential of an electron changes
evaluating the CO2 adsorption energy on a Pt(111) surface, with the RHE potential via −eU.17 Within this model, the
with different coadsorbates leading to substantial variation in electrostatic potential is global and affects the energy of the
the work function, shown in Figure 1. [Note: the Pt(111) electrons in the electrode (in an explicit solvation picture, the
field effect is local and affects the energy of adsorbates outside
the surface). The CHE model then allows us to separate the
effect of electrochemical potential from other physical
influences at the electrochemical interface.
Typically, a barrier of 0.75 eV is considered to be the upper
limit for facile kinetics, bceause it yields a turnover frequency of
∼1 s−1 at room temperature. On Pt(111), the activation of CO2
is facile, even at room temperature. The formation of *CO2 as
an intermediate before the coupled proton−electron transfer is
only moderately uphill in free energy. However, CO production
is still slow on Pt(111) due to the strong adsorption of CO,
which limits *CO desorption.9,26,27 On Ag(111), on the other
hand, both *CO2 and *COOH intermediates are high in free
energy, and *CO desorption is facile at the equilibrium
potential.
Figure 2b shows the free-energy diagram for Ag(111) at
−0.11 vs RHE that includes an applied field of −1.0 V/Å. As
discussed in the following, this local field strength is typical of
solvated cations at the double layer and the corresponding
energy is obtained via eq 5 (presented later in this work). The
field leads to a significant shift in the free energy of *CO2 (and,
to a lesser extent, *COOH), such that the first step is no longer
thermodynamically difficult. Therefore, we expect substantial
field effects on the CO2 → CO kinetics over Ag and other
weakly adsorbing surfaces. For more reactive surfaces such as
Pt, this effect should be considerably smaller, since *CO2 is
already quite stable there and therefore any field stabilization
will not be seen in the kinetics (see Figure 2a).
Now, we address field effects on adsorbed CO2 in detail.
Field effects on adsorption energies have been studied
extensively in the literature, both to explain alkali promotion
in heterogeneous catalysis28−30 and as a basis for understanding
Figure 1. Geometry and energetics of CO2 adsorption. surface electrochemistry.31−36 Field effects on an adsorbate are
strongly dependent on its dipole moment and polarizability:
many adsorbates show rather small field effects, such as the
surface is considered here, since the adsorption of CO2 is common intermediates in the oxygen reduction and evolution
favorable, even in the absence of an electric field (ΔE < 0). reactions.34,35 An exception is the C−C coupling reaction
Note that we show ΔG in Figure 1b, which is slightly positive energy, where significant effects have been found on important
from the enthalpic and entropic contributions (see the intermediates.33
Supporting Information (SI) for details of the electronic From a modeling perspective, the field effect can be
structure calculations). The side of the slab without any investigated in two ways, via (1) applying an external dipole
adsorbates retains the bare Pt(111) work function (5.8 eV) potential under vacuum, and (2) explicitly considering
from our calculations. Although molecular adsorption on adsorption at an Ag(111) | water interface (see Figures S3−
Pt(111) can vary significantly with the choice of exchange- S5 in the SI for calculation details). We examine the effects of
correlation functional, the Bayesian error estimation functional potassium (K+) and 1-ethyl-3-methylimidazolium (EMIM+) in
with van der Waals correlation (BEEF-vdW) used in the the current study, since experiments have suggested that these
current study has been shown to compare favorably with ions lower the overpotential for CO production.7,8,37,38 Alkali
experimental data.23,24 A 4 × 4 cell was used to ensure adequate cations have been hypothesized to adsorb to the electrode
separation (a minimum of 6 Å) between CO2 and the surface7,39 or interact with surface-adsorbed species non-
coadsorbate, to prevent interaction. These work function covalently.40,41 Surface X-ray scattering data have suggested
changes agree well with a previous study on halide adsorption some alkali and alkaline-earth cations to be solvated and not
to the Pt(111) surface.25] adsorbed at the electrode surface.42 EMIM+ has been shown to
7134 DOI: 10.1021/acscatal.6b02299
ACS Catal. 2016, 6, 7133−7139
ACS Catalysis Letter

Figure 2. Free-energy diagram for CO2(g) to CO(g) without (left) and with (right) the effects of electric field on Ag(111) at the equilibrium
potential of the overall reaction (−0.11 V vs RHE). The same pathway on Pt(111) is shown to contrast CO2 and CO adsorption energies for the two
metals in the absence of an applied electric field.

Figure 3. (a) Plot showing how an applied electric field affects the energetics of CO2 adsorption at 0 V vs SHE, showing field-stabilized energies
obtained with both vacuum and explicit solvation methods. (b) Depiction of the solvated structures representative of cation-containing water layers.

have a high coverage on metals, because of the strong van der comparison of the two datasets, since adsorbates in the vacuum
Waals interactions.43 EMIM+ has also been hypothesized to calculations experience no stabilization from hydrogen bonding
catalyze the CO2 → CO mechanism through specific chemical to the solvent.
bonding; however, the exact nature of this remains unclear,44,45 The energy of an adsorbate in an electric field is given by
and we focus only on the field effect in the present study.
In Figure 3a, we show the adsorption energy as a function of α ,2
E = E0 + μ , −
field strength for *CO2. The adsorbate adopts a bent 2 (5)
configuration on the surface, signifying sharing of electron where E represents the potential energy; E0 is the adsorption
density with the metal.46 The energies obtained under vacuum energy in the absence of an electric field; μ and α are the
through applying an external potential are indicated by gray intrinsic dipole moment and polarizability values, respectively,
circles, while the energies obtained with explicit solvent and for the adsorbate; and , represents the electric field.47 The
solvated K+ and EMIM+ ions are indicated by purple and blue equation of best fit for the curve in Figure 3a gives μ = 0.96 e Å
polygons, respectively (for determination of the electric field and α = 0.76 e Å2 V−1. Generally, energies obtained through
with the latter method, see Figures S4 and S5 in the SI). The calculations under vacuum and with explicit solvation agree,
solvated ions are located ∼4 Å from the metal surface and are suggesting that the stabilization from the solvated ions is due to
not specifically adsorbed. The numbers refer to distinct solvent the electric field and not specific chemical bonding. All solvent
configurations for both ions (see Figure S2 in the SI). Figure 3b configurations exert similar field strengths in the vicinity of
shows an example of the solvated cation structures used in our *CO2 and thus result in similar stabilizing effects.
calculations. An average solvation energy (see eq S3 in the SI) In contrast to classical models of the electrochemical
has been applied to the vacuum calculations to allow for the interface, the field effect is localized on the atomic scale. Figure
7135 DOI: 10.1021/acscatal.6b02299
ACS Catal. 2016, 6, 7133−7139
ACS Catalysis Letter

4 shows the cation-induced electric fields taken from a z-slice r4 = k4θCO − k −4pCO θ
* (9)
near the center of adsorbed CO2 with K+ solvated at the surface
where k is the rate constant and θ the coverage of a specific site
denoted in the subscript. The potential-dependent rate-
constant ki for the ith step is

kBT ⎛ E0 ⎞ ⎡ eβ (U − U 0) ⎤
ki = exp⎜⎜ −
a, i
⎟⎟ exp⎢ − i i

h ⎝ B ⎠
k T ⎢
⎣ k BT ⎥⎦ (10)

where β is the transfer coefficient, U the applied potential, and


E0a the reaction barrier at zero applied potential (the rate
constant for a chemical step would simply omit the second
exponent containing the potential dependence).
We consider two sitesone stabilized by field and one
unaffectedwhere the total current density is the sum of the
two contributions (for a list of equations, see the section
entitled “Microkinetic Modeling” in the SI). As for the
concentration of ions near the surface, we assume the
concentration of solvated K+ in the double layer to be equal
to that in the bulk solution at the pzc and evaluate the following
Figure 4. Electric field distribution taken at a z-slice near the center of expression:
*CO2 in a 6 × 6 unit cell with a K+ coverage of 1/36. One cell,
pzc
including metal atoms (parallelogram and solid circles in light gray), is θion = c iondHA site (11)
outlined to illustrate the ion coverage. Solvent atoms are represented
in dashed light gray circles. K+ (large circles) and *CO2 in its vicinity where θpzcion is the coverage of ions at the pzc, cion the
are outlined in black for emphasis. concentration of the ion in bulk solution (1 M, e.g. KOH), dH
the Helmholtz layer thickness (∼3 Å), and Asite the area of a
(see eqs S4 and S5 in the SI for details) at a coverage of 1/36. surface site from our DFT calculations (7.7 Å2). This
This ion coverage corresponds roughly to a potential of −0.5 V expression gives an ionic coverage of 1.4% in the Helmholtz
vs SHE, which is calculated with eq 12 (presented in the plane at the pzc. Our simulations were done with a static water
following section). Where the electrochemical interface is structure, and in a physical system these field-stabilized sites
explicitly simulated, accurately evaluating the electrochemical could migrate with solvent fluctuations. However, the overall
potential has been a challenge.48,49 One of the main problems is concentration of these sites should be constant on average.
that the solvent dipoles, which should average out to zero in a Diffuse layer effects can dominate close to the pzc, although
real physical system at room temperature,50 contribute much more negative of this potential, the concentration of
significantly to the work function obtained in our simulations. cations increases according to a simple Helmholtz model:
Therefore, no straightforward mapping between the true
electrochemical potential and the work function obtained θion = C H(U − Upzc) (12)
from our simulations exists. Hence, in the present work, we
have used a classical approach by referencing the potential of where CH is the Helmholtz capacitance (∼50 μF cm−2)51 and
zero charge (pzc) and relating the ion coverage to a realistic Upzc is the potential of zero charge (−0.46 V vs SHE) for
potential using eq 12. Ag(111).52 However, we note that the increase in cation
Although the field is large in the vicinity of the K+ ion (on concentration at more reducing potentials has little effect on
the order of −1 V/Å), it decays to ∼0 V/Å outside a radius of the kinetics, since the exponential terms dominate in the overall
∼5 Å. Therefore, only *CO2 in the immediate vicinity of the rate expression.
ion would be affected by the induced field. The atomistic We expect the rates of steps (1), (2), and (4) to be largely
picture also differs from a macroscopic model where the energy determined by thermochemistry, since the activation barriers
of the adsorbate varies continuously with an applied field: here, for gas-phase adsorption and proton−electron transfers to
the field is binaryeither “on” or “off”and is always on the oxygen are generally small,53,54 and we have found the
order of −1 V/Å in the “on” sites. We find the average field adsorption of CO2 to have a barrier of <0.1 eV (see Figure
over the entire cell (at this z-slice) to be −0.36 V/Å, which S7 in the SI). Our DFT calculations suggest that step (3) is
would correspond roughly to that expected from a classical rate-determining with an intrinsic barrier of 0.7 eV, and solvent
Helmholtz model for an interfacial potential drop of ∼1 V and reorganization adds an additional 0.35 eV to this barrier.55 The
a Helmholtz layer thickness of ∼3 Å. value of β was determined to be 0.64 from the surface charge
We now formulate a simple microkinetic model that (electrons summed over all Ag atoms in the metallic slab and all
accounts for the effects of electric field. The rate equations adsorbates, from which the number of electrons of the neutral
for all elementary steps are given as follows: system was subtracted) of the transition state via Bader
analysis.56,57 We note that there is a typical 0.2 eV uncertainty
r1 = k1pCO θ − k −1θCO2 (6) associated with our DFT reaction and transition-state
2 *

− energies.23 We have also checked the value of β with hybrid


r2 = k 2θCO2 − k −2[OH ]θCOOH (7) calculations58 and found the agreement to be within 0.1.
Assuming the rest of the steps are at quasi-equilibrium, they are
r3 = k 3θCOOH − k −3[OH−]θCO (8) governed by their equilibrium constants:
7136 DOI: 10.1021/acscatal.6b02299
ACS Catal. 2016, 6, 7133−7139
ACS Catalysis Letter

⎛ ΔGi ⎞
K i = exp⎜ − ⎟
⎝ kBT ⎠ (13)
where the ΔG value of each step is shown in the free-energy
diagram. We can then solve for the overall rate analytically.
Now we contrast the free-energy landscapes at pH 7 and pH
14. In Figure 5, the energy of each species containing a

Figure 6. Computed (solid lines) polarization curves at pH 7 (orange)


and pH 14 (blue), compared to experimental results (filled circles)
adapted from refs 7 and 8. Pressures for CO2 and CO are 1 bar and 1
mbar, respectively. Dotted lines mark the equilibrium potential for
CO2 → CO at each pH (−0.52 V at pH 7 and −0.94 V at pH 14). The
overpotential (distance between the equilibrium and onset potentials)
at pH 7 is shown to be much larger than the overpotential at pH 14.
Figure 5. Free-energy diagram for the field-stabilized pathway at the
equilibrium potential at pH 7 and pH 14.
from DFT calculations and no fitted parameters, our kinetic
hydroxide ion is shifted upward by 0.059 eV per pH unit away model reproduces the onset potential in the experimental data
from 0, and the energy of each species containing an electron is with reasonable agreement, as shown in the inset of Figure 6. At
shifted by an equal −eU to maintain equilibrium. The dotted larger current densities, transport limitations can contribute and
lines represent a typical solvent reorganization barrier of 0.35 these effects are not captured in the current model. The
eV.55 We see that the rate-determining step (see Figure S8 in dependence of electrochemical rates on the absolute potential
the SI for the transition-state calculation) has a larger barrier at rather than the RHE potential have led to the hypothesis that
pH 7 than that at pH 14. There are no contributions from the potential-determining step is an electron transfer decoupled
configurational entropy to the transition-state energy. This is from the corresponding proton transfer.61−63 Our computed
because the transition state is too unstable to have any polarization curves suggest that this dependence can simply
appreciable concentration according to transition-state theory arise from step 3, where the barrier decreases as a function of
(TST).59 Thus, the energy of this state is dependent only on absolute potential.
the absolute potential, −βeU. In summary, we have shown that local cation-induced fields
Figure 6 compares the computed current densities from our have a substantial stabilizing effect on adsorbed CO2. These
kinetic model to the experimentally measured polarization cations shift the free-energy landscape significantly for the CO2
behavior7,8, using the energies shown in Figure 5. The → CO reaction on weakly adsorbing metals such as Ag so that
conversion factor is given as follows: CO2 adsorption is facile, and the *COOH → *CO step
becomes potential-limiting instead. Theoretical and exper-
⎛ z ⎞ imental polarization curves show reasonable agreement at
rmacroscopic = ⎜⎜ ⎟⎟rmicroscopic experimental onset potentials. The implication is that electric
⎝ ρsite q ⎠ (14) field effects should be included to model CO2 reduction in
where z is the number of electrons transferred, ρsite the site general, since many intermediates are expected to possess large
density (given in cm−2), q the charge of an electron (given in dipoles or polarizabilities whose energies would be affected
mC), and rmicroscopic the rate obtained from the kinetic model significantly by an electric field. These dipole and polarizability
(given in s−1). Consequently, the resulting macroscopic rate has values can be calculated under vacuum and added as an
units of mA cm−2. The equilibrium potential U0 for CO2 → CO additional term in the CHE model, since they agree well with
shifts vs SHE, according to the Nernst equation:60 explicit solvent calculations. Further work will focus on field
effects on CO2 reduction intermediates and transition states
U0 = U0pH = 0 − 0.059pH (15) beyond CO, which could have a significant impact on the
resulting kinetics.


where UpH=0
0 is the equilibrium potential at pH 0. The
equilibrium potential for the reaction under discussion then
changes from −0.11 at pH 0 to −0.52 V at pH 7 and −0.94 V
ASSOCIATED CONTENT
at pH 14. These equilibrium potentials are indicated by dotted *
S Supporting Information

lines in Figure 6. The Supporting Information is available free of charge on the


Several studies in the literature have shown that the ACS Publications website at DOI: 10.1021/acscatal.6b02299.
overpotential for this pathway is considerably greater at pH Details of DFT calculations, solvation correction, field
7, compared to that at pH 14.7−9 Using only values obtained quantification, and microkinetic modeling (PDF)
7137 DOI: 10.1021/acscatal.6b02299
ACS Catal. 2016, 6, 7133−7139
ACS Catalysis Letter

■ AUTHOR INFORMATION
Corresponding Author
(21) Andreussi, O.; Dabo, I.; Marzari, N. J. Chem. Phys. 2012, 136,
064102.
(22) Andreussi, O.; Marzari, N. Phys. Rev. B: Condens. Matter Mater.
*E-mail: norskov@stanford.edu. Phys. 2014, 90, 245101.
Notes (23) Wellendorff, J.; Lundgaard, K. T.; Møgelhøj, A.; Petzold, V.;
The authors declare no competing financial interest. Landis, D. D.; Nørskov, J. K.; Bligaard, T.; Jacobsen, K. W. Phys. Rev.


B: Condens. Matter Mater. Phys. 2012, 85, 235149.
(24) Gautier, S.; Steinmann, S. N.; Michel, C.; Fleurat-Lessard, P.;
ACKNOWLEDGMENTS Sautet, P. Phys. Chem. Chem. Phys. 2015, 17, 28921−28930.
The authors gratefully acknowledge support from the Office of (25) Gossenberger, F.; Roman, T.; Forster-Tonigold, K.; Groß, A.
Basic Energy Sciences of the U.S. Department of Energy to the Beilstein J. Nanotechnol. 2014, 5, 152−161.
SUNCAT Center for Interface Science and Catalysis. L.D.C. (26) Taguchi, S.; Aramata, A.; Enyo, M. J. Electroanal. Chem. 1994,
372, 161−169.
acknowledges financial support from the Natural Sciences and (27) Hansen, H. A.; Varley, J. B.; Peterson, A. A.; Nørskov, J. K. J.
Engineering Research Council of Canada for the CGS-D3 Phys. Chem. Lett. 2013, 4, 388−392.
fellowship. This material is based, in part, on work performed (28) Ertl, G.; Lee, S. B.; Weiss, M. Surf. Sci. 1982, 114, 527−545.
by the Joint Center for Artificial Photosynthesis, a DOE Energy (29) Nørskov, J. K.; Holloway, S.; Lang, N. D. Surf. Sci. 1984, 137,
Innovation Hub, supported through the Office of Science of the 65−78.
U.S. Department of Energy, under Award No. DE-SC0004993. (30) Strongin, D. R.; Somorjai, G. A. J. Catal. 1988, 109, 51−60.
This research used resources of the National Energy Research (31) Holloway, S.; Nørskov, J. K. J. Electroanal. Chem. Interfacial
Scientific Computing Center, a DOE Office of Science User Electrochem. 1984, 161, 193−198.
Facility supported by the Office of Science of the U.S. (32) Wasileski, S. A.; Koper, M. T. M.; Weaver, M. J. J. Am. Chem.
Department of Energy (under Contract No. DE-AC02- Soc. 2002, 124, 2796−2805.
(33) Montoya, J. H.; Shi, C.; Chan, K.; Nørskov, J. K. J. Phys. Chem.
05CH11231). We thank Dr. Joseph H. Montoya for stimulating
Lett. 2015, 6, 2032−2037.
discussions and Chihiro Urushihara for assistance with creating (34) Hyman, M. P.; Medlin, J. W. J. Phys. Chem. B 2005, 109, 6304−
figures.


6310.
(35) Karlberg, G. S.; Rossmeisl, J.; Nørskov, J. K. Phys. Chem. Chem.
REFERENCES Phys. 2007, 9, 5158−5161.
(1) Whipple, D. T.; Kenis, P. J. A. J. Phys. Chem. Lett. 2010, 1, 3451− (36) Mamatkulov, M.; Filhol, J. S. Phys. Chem. Chem. Phys. 2011, 13,
3458. 7675−7684.
(2) Hori, Y.; Murata, A.; Kikuchi, K.; Suzuki, S. J. Chem. Soc., Chem. (37) Rosen, B. A.; Salehi-Khojin, A.; Thorson, M. R.; Zhu, W.;
Commun. 1987, 728−729. Whipple, D. T.; Kenis, P. J. A.; Masel, R. I. Science 2011, 334, 643−
(3) Chandrasekaran, K.; Bockris, J. O. Surf. Sci. 1987, 185, 495−514. 644.
(4) Bockris, J. O.; Wass, J. C. J. Electrochem. Soc. 1989, 136, 2521− (38) Sun, L.; Ramesha, G. K.; Kamat, P. V.; Brennecke, J. F. Langmuir
2528. 2014, 30, 6302−6308.
(5) Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electrochim. Acta (39) Hori, Y.; Suzuki, S. Bull. Chem. Soc. Jpn. 1982, 55, 660−665.
1994, 39, 1833−1839. (40) Strmcnik, D.; Kodama, K.; van der Vliet, D.; Greeley, J.;
(6) Hori, Y. Modern Aspects of Electrochemistry; Springer: New York, Stamenkovic, V. R.; Marković, N. M. Nat. Chem. 2009, 1, 466−472.
2008; pp 89−189. (41) Stoffelsma, C.; Rodriguez, P.; Garcia, G.; Garcia Araez, N.;
(7) Thorson, M. R.; Siil, K. I.; Kenis, P. J. A. J. Electrochem. Soc. 2013, Strmcnik, D.; Marković, N. M.; Koper, M. T. M. J. Am. Chem. Soc.
160, F69−F74. 2010, 132, 16127−16133.
(8) Ma, S.; Luo, R.; Moniri, S.; Lan, Y.; Kenis, P. J. A. J. Electrochem. (42) Strmcnik, D.; van der Vliet, D. F.; Chang, K. C.; Komanicky, V.;
Soc. 2014, 161, F1124−F1131. Kodama, K.; You, H.; Stamenkovic, V. R.; Marković, N. M. J. Phys.
(9) Kuhl, K. P.; Hatsukade, T.; Cave, E. R.; Abram, D. N.; Kibsgaard, Chem. Lett. 2011, 2, 2733−2736.
J.; Jaramillo, T. F. J. Am. Chem. Soc. 2014, 136, 14107−14113. (43) Urushihara, M.; Chan, K.; Shi, C.; Nørskov, J. K. J. Phys. Chem.
(10) Taylor, C. D.; Wasileski, S.; Filhol, J.-S.; Neurock, M. Phys. Rev. C 2015, 119, 20023−20029.
B: Condens. Matter Mater. Phys. 2006, 73, 165402. (44) Snuffin, L. L.; Whaley, L. W.; Yu, L. J. Electrochem. Soc. 2011,
(11) Steinmann, S. N.; Michel, C.; Schwiedernoch, R.; Sautet, P. 158, F155−F158.
Phys. Chem. Chem. Phys. 2015, 17, 13949−13963. (45) Lau, G. P. S.; Schreier, M. R.; Vasilyev, D.; Scopelliti, R.; Grätzel,
(12) Lespes, N.; Filhol, J.-S. J. Chem. Theory Comput. 2015, 11, M.; Dyson, P. J. J. Am. Chem. Soc. 2016, 138, 7820−7823.
3375−3382. (46) Bendavid, L. I.; Carter, E. A. J. Phys. Chem. C 2013, 117, 26048−
(13) Goodpaster, J. D.; Bell, A. T.; Head-Gordon, M. J. Phys. Chem. 26059.
Lett. 2016, 7, 1471−1477. (47) Petrucci, R. H.; Herring, F. G.; Madura, J. D.; Bissonnette, C.
(14) Xiao, H.; Cheng, T.; Goddard, W. A., III; Sundararaman, R. J. General Chemistry: Principles and Modern Applications; Prentice Hall:
Am. Chem. Soc. 2016, 138, 483−486. Upper Saddle River, NJ, 2007.
(15) Lozovoi, A. Y.; Alavi, A. Phys. Rev. B: Condens. Matter Mater. (48) Tripković, V.; Björketun, M. E.; Skúlason, E.; Rossmeisl, J. Phys.
Phys. 2003, 68, 245416. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 115452.
(16) Otani, M.; Sugino, O. Phys. Rev. B: Condens. Matter Mater. Phys. (49) Björketun, M. E.; Zeng, Z.; Ahmed, R.; Tripković, V.; Thygesen,
2006, 73, 115407. K. S.; Rossmeisl, J. Chem. Phys. Lett. 2013, 555, 145−148.
(17) Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; (50) Roman, T.; Groß, A. Catal. Today 2013, 202, 183−190.
Kitchin, J. R.; Bligaard, T.; Jónsson, H. J. Phys. Chem. B 2004, 108, (51) Schmickler, P. D. W.; Santos, E. Interfacial Electrochemistry;
17886−17892. Springer: Berlin, Heidelberg, 2010.
(18) Steinmann, S. N.; Michel, C.; Schwiedernoch, R.; Filhol, J.-S.; (52) Frumkin, A. N.; Petrii, O. A.; Damaskin, B. B. Comprehensive
Sautet, P. ChemPhysChem 2015, 16, 2307−2311. Treatise of Electrochemistry; Springer: Boston, MA, 1980; pp 221−289.
(19) Steinmann, S. N.; Sautet, P. J. Phys. Chem. C 2016, 120, 5619− (53) Shi, C.; O’Grady, C. P.; Peterson, A. A.; Hansen, H. A.;
5623. Nørskov, J. K. Phys. Chem. Chem. Phys. 2013, 15, 7114−7122.
(20) Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, (54) Shi, C.; Chan, K.; Yoo, J. S.; Nørskov, J. K. Org. Process Res. Dev.
T. A.; Hennig, R. G. J. Chem. Phys. 2014, 140, 084106. 2016, 20, 1424−1430.

7138 DOI: 10.1021/acscatal.6b02299


ACS Catal. 2016, 6, 7133−7139
ACS Catalysis Letter

(55) Hansen, H. A.; Viswanathan, V.; Nørskov, J. K. J. Phys. Chem. C


2014, 118, 6706−6718.
(56) Henkelman, G.; Arnaldsson, A.; Jónsson, H. Comput. Mater. Sci.
2006, 36, 354−360.
(57) Tang, W.; Sanville, E.; Henkelman, G. J. Phys.: Condens. Matter
2009, 21, 084204.
(58) Krukau, A. V.; Vydrov, O. A.; Izmaylov, A. F.; Scuseria, G. E. J.
Chem. Phys. 2006, 125, 224106.
(59) Eyring, H. J. Chem. Phys. 1935, 3, 107.
(60) Bard, A. J.; Faulkner, L. R. Electrochemical Methods:
Fundamentals and Applications; John Wiley & Sons, Inc.: New York,
1980.
(61) Koper, M. T. M. Chem. Sci. 2013, 4, 2710−2723.
(62) Koper, M. T. M. Phys. Chem. Chem. Phys. 2013, 15, 1399−1407.
(63) Katsounaros, I.; Chen, T.; Gewirth, A. A.; Marković, N. M.;
Koper, M. T. M. J. Phys. Chem. Lett. 2016, 7, 387−392.

7139 DOI: 10.1021/acscatal.6b02299


ACS Catal. 2016, 6, 7133−7139

You might also like