You are on page 1of 8

Perspective

pubs.acs.org/JPCL

Activity Descriptors for CO2 Electroreduction to Methane on


Transition-Metal Catalysts
Andrew A. Peterson†,§ and Jens K. Nørskov*,†,‡

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

SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, Menlo Park, California, United States
§
Center for Atomic-scale Materials Design, Department of Physics, Technical University of Denmark, DK-2800 Lyngby, Denmark
Downloaded via RAJIV GANDHI INST PETROLEUM TECHLGY on March 19, 2023 at 06:33:36 (UTC).

*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The electrochemical reduction of CO2 into hydrocarbons and alcohols


would allow renewable energy sources to be converted into fuels and chemicals. However,
no electrode catalysts have been developed that can perform this transformation with a low
overpotential at reasonable current densities. In this work, we compare trends in binding
energies for the intermediates in CO2 electrochemical reduction and present an activity
“volcano” based on this analysis. This analysis describes the experimentally observed
variations in transition-metal catalysts, including why copper is the best-known metal
electrocatalyst. The protonation of adsorbed CO is singled out as the most important step
dictating the overpotential. New strategies are presented for the discovery of catalysts that
can operate with a reduced overpotential.

U nlike renewable electricity, which can come from several


sources (hydro, solar, wind, geothermal, and others),
renewable carbon-based fuels are currently only produced from Beyond the criterion of selectivity
biomass. The only other possibility to produce such fuels is by over the hydrogen evolution
chemically reducing carbon dioxide; if CO2 can be electro-
chemically reduced to hydrocarbons, then any of the renewable reaction, no systematic criteria for
electricity sources can also be used to create fuels, simulta-
neously addressing storage issues for these often intermittent effective CO2 reduction catalysts have
resources.1 Over the past 3 decades, many heterogeneous and been proposed. It is the purpose of
homogeneous electrocatalysts have been screened for their
effectiveness in this reaction and the related photoelectrochem- the present Perspective to develop
ical process (artificial photosynthesis),2−18 and none have been
found to be more effective than pure metallic copper for its
such criteria.
repeatable demonstration to produce high (>50%) Faradaic
yields of hydrocarbons at reasonable (∼5 mA cm−2) current be comparatively poor HER activity in the presence of CO2.
densities. The high selectivity of copper for hydrocarbons in However, beyond this criterion of selectivity over the HER, no
CO2 electroreduction has sparked much interest in uncovering
systematic criteria for effective CO2 reduction catalysts have
its mechanism.11−16,19,20 However, the overpotential necessary
to achieve this reaction on copper electrodes is prohibitively been proposed. It is the purpose of the present Perspective to
high, on the order of 1 V,2 which dwarfs the overpotential of develop such criteria.
the oxygen evolution reaction,21,22 the typical counter electrode Many of the trends in CO2 reduction by heterogeneous
reaction for these systems. electrocatalysts can be observed from the pioneering
From a purely thermodynamical point of view, CO2 reduc- experimental work of Hori and co-workers.2,10 In Table 1, we
tion to the simplest hydrocarbon, CH4, should be possible at a have highlighted a subset of these electrocatalysts. This table
standard potential of +0.17 V versus the reversible hydro- captures many of the major trends in CO2 reduction, including
gen electrode, or RHE. Because, by definition, hydrogen evolu- selectivity toward hydrocarbons (Cu and, to a lesser extent, Ni
tion becomes thermodynamically possible at 0 V versus RHE
(at any pH), CO2 reduction will be in competition with the Received: November 4, 2011
hydrogen evolution reaction (HER)23 at all negative potentials, Accepted: January 5, 2012
and a key criterion for selective catalysts in CO2 reduction will Published: January 5, 2012

© 2012 American Chemical Society 251 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258
The Journal of Physical Chemistry Letters Perspective

Table 1. Faradaic Yields in CO2 Reduction on Face- shown in Figure 1, which is bound by stoichiometry to involve
Centered Cubic (fcc) Metal Electrodes, As Reported by Hori the transfer of eight proton−electron pairs, resulting in seven
et al.,10 for Experiments at 5 mA cm−2 Current Density in a adsorbed intermediates. In both the previous analysis and in the
0.1 M KHCO3 Buffer at 18.5°C current work, we focus on the limiting potential (UL) at which
each elementary step of a reaction becomes exergonic (or
Faradaic yield, %
downhill in free energy) as a simple measure of the potential
V vs hydrocarbons/ dependence of the electrochemical rate. For an example
organicsa
electrode RHE CO HCOOH H2 total
elementary hydrogenation reaction, A* + H+ + e− → B*
Ni −1.09 2.1 0.0 1.4 88.9 92.4 (where A* and B* are generic adsorbed species), the limiting
Cu −1.05 72.3 1.3 9.4 20.5 103.5 potential can be found as a function of the chemical potentials
Pd −0.81 2.9 28.3 2.8 26.2 60.2 (μ) of the bound species
Ag −0.98 0.0 81.5 0.8 12.4 94.6
Pt −0.68 0.0 0.0 0.1 95.7 95.8 μ[B*] − μ[A*] − μ[H+ + e−](U = 0 VRHE)
UL = −
Au −0.75 0.0 87.1 0.7 10.2 98.0 e
a
Predominantly CH4, but also including C2H4, C2H5OH, C3H7OH, C2H6, 0V
−ΔG elem
C3H5OH, CH3CHO, and C2H5CHO. =
e (1)
and Pd), the production of CO (Au and Ag), and the lack of where e is the (positive) charge of an electron and the chemical
strong catalytic activity in CO2 reduction, evolving H2 instead potential of the proton−electron pair at 0 V can be found from
(Pt and Ni). These electrode materials are all late transition the computational hydrogen electrode model.19,24 This is equiv-
metals that can be represented in the face-centered cubic (fcc) alent to the negative Gibbs free-energy change of the elemen-
crystal structure; this uniformity allows us to perform a system- tary reaction divided by the electronic charge. The chemical
atic analysis of the factors determining the electrochemical potentials of the bound species A* and B* can be found from
activity and selectivity toward hydrocarbons. We focus here on standard statistical mechanics treatment of our calculated
the simplest hydrocarbon methane, but the analysis is quite binding energies (EB) of these adsorbed species. It has been
general. Our analysis is based on an extensive set of density found that such an analysis correctly predicts the onset poten-
functional theory (DFT) calculations of the key adsorbates in tial for the formation of different products in CO2 reduction
CO2 reduction on these surfaces, conducted for this study. over Cu surfaces19,20 and that it provides a good description of
In previous work, we proposed19,20 a mechanism by which trends in reactivity of different metals for the oxygen reduction
copper electrocatalysts reduce CO2 to methane; this route is reaction.25

Figure 1. Adsorption energy scaling. The proposed pathway from CO2 to CH4 on copper surfaces is shown at the top of the figure, and the
calculated adsorption energies of the key bound intermediates on fcc (211) facets are shown in the two lower figures. (More tightly bound
adsorbates correspond to more negative binding energies.) The adsorption energies of those adsorbates binding to the surface through carbon can be
correlated and are plotted against the binding energy of CO in the left plot. Similarly, the adsorption energies of those adsorbates binding through
oxygen can be correlated and are plotted versus the binding energy of OH in the right plot. Ir and Rh are included as additional fcc metals to those
shown in Table 1 in order to provide more points for the scaling relations.

252 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258


The Journal of Physical Chemistry Letters Perspective

Figure 2. Limiting potentials (UL) for elementary proton-transfer steps in the mechanism of Figure 1. Each line is the calculated potential at which
the indicated elementary reaction step is neutral with respect to free energy, as a function of the electrocatalysts’ carbon or oxygen affinity, EB[CO]
or EB[OH]. Each elementary reaction involves the addition of a proton−electron pair (H+ + e−) to the reactant. The equilibrium potential for the
overall electrochemical reduction of CO2 to CH4 is +0.17 V versus RHE, which is indicated on the figure. Thermodynamics limits any real catalyst to
operate at potentials more negative than the equilibrium potential; therefore, the theoretical overpotential as a function of EB[CO] or EB[OH] can
be represented by the distance between the equilibrium line and the most-negative limiting potential line (highlighted in gray). The elementary steps
in this figure are limited to the steps in the mechanism of Figure 1 and exclude CH2O* → OCH3*, which has dependence on both EB[CO] and
EB[OH]. The CH2O* → OCH3* transformation can be found in the Supporting Information. The EB[CO] and EB[OH] values for each metal are
shown on the plot, along with their overall predicted value of UL based on these scaling relations.

Given that there are seven intermediates in the reduction of CO2 → COOH*, are the most negative of the reaction steps
CO2 to CH4, such an analysis involves an understanding of the shown in Figure 1; thus, for materials that follow this same
problem in seven-dimensional space of adsorbate−surface interac- reaction network, they would be expected to dictate the
tion energies. However, as seen in Figure 1, roughly the first half of overpotential requirement of the reaction. Copper is seen to sit
the intermediates interact with the catalyst surface through a car- near the top of this “volcano”-type relation.
bon atom, and the second half interact through an oxygen atom. A similar plot is shown on the right-hand side of Figure 2 for
Scaling relations associated with the d-band theory of adsorption the steps that involve oxygen-bound intermediates. The ele-
suggest there may be strong correlations within the carbon-bound mentary UL lines on this plot are much less negative; it is only for
and oxygen-bound species.26 Using this principle, we are able materials that have very high affinities for OH that the potential of
to successfully correlate the adsorption energies of the carbon- clearing OH from the surface becomes comparable to the
bound species to EB[CO] and those of the oxygen-bound species limitations set by the CO* → CHO* reaction. This does indicate
to EB[OH], as shown in Figure 1. For example, the binding that for materials that exhibit strong OH binding (more negative
energy of CHO can be found from the linear equation in Figure 1 EB[OH]), OH removal will likely limit the catalysts’ effectiveness;
however, for most materials, the potential-limiting step will be set
EB[CHO] = 0.88EB[CO] + 2.03 eV (2) by the CO* → CHO* transformation discussed above.
A striking feature of Figure 2 is the severity of the limiting
These scaled affinity relations allow us to reduce the dimen- potential needed to protonate CO* to CHO*. This UL is not
sionality of this reaction network from seven to two, thus making only the strongest limitation, it is also the least sensitive
the search for trends within these transition-metal catalysts trac- to changes in EB[CO]; that is, the line is nearly horizontal.
table. In this way, the limiting potentials for each of the elemen-
tary steps can be estimated for each surface as a function of that
surface’s affinity for CO and OH. Full mathematical details are
available in the Supporting Information. The magnitude and insensitivity of
The elementary limiting potentials that scale with EB[CO]
are shown in the left pane of Figure 2. Each solid line on this the energetics of this reaction to the
figure shows the limiting potential (UL) of an elementary proto- electronic properties of the catalyst
nation reaction for the reaction network of Figure 1. Because
UL is the potential at which the elementary reaction becomes material can explain both the reason
exergonic, each limiting potential gives a first-order indication
of the electrical potential at which that elementary step begins for the large overpotential require-
to have an appreciable rate. The equilibrium potential is also ment and the dearth of materials that
plotted in Figure 2. The difference between each UL and the
equilibrium potential gives a first-order estimate of the over- have been found that can catalyze
potential requirement for that elementary step; thus, the most this reaction with lower overpoten-
negative UL line at any CO adsorption energy dictates the
theoretical overpotential. This is indicated as the gray area of tials than Cu.
the plot. The two bottom-most UL lines, CO* → CHO* and
253 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258
The Journal of Physical Chemistry Letters Perspective

The magnitude and insensitivity of the energetics of this that produce CHO*; the most limiting of these steps is shown,
reaction to the electronic properties of the catalyst material can which is the protonation of CH* to form CH2*. For materials
explain both the reason for the large overpotential requirement that form CHO*, other downstream pathways are possible;
and the dearth of materials that have been found that can these later steps will be important to selectivity15 but should
catalyze this reaction with lower overpotentials than Cu. The not affect the overall theoretical overpotential. The most
root cause of this insensitivity can be seen in Figure 1, in which significant is the subsequent protonation of CHO* to CH2O*
the CO and CHO binding energies scale with a slope of 0.88, or CHOH*. In both cases, this reaction appears to require less
meaning that a surface that stabilizes CHO will stabilize CO by overpotential than the protonation of CO*.
a similar amount. Thus, the binding energies are coupled, and Over most of the region of the plot, the CHO* route is
the relative binding energies between the two change only favored over the COH* route. However, at low values of
slightly with catalyst material. EB[CO], the route to COH* begins to become possible. This
Of course, electrocatalysts are not bound to follow the reac- upturn suggests the possibility that materials with a high affinity
tion mechanism shown in Figure 1. Other mechanistic possibi- for carbon may have reduced overpotentials; however, some
lities are included in Figure 3, focusing on the mechanisms that attributes of materials in this region may prevent this from
being realized. Materials that bind carbon strongly will lead to
high coverages when in actual electrochemical operation; this
will weaken the binding strength of each incremental adsorbate,
thus effectively shifting the binding strength of these materials
to the right on the plot. Additionally, many of the materials that
bind carbon strongly also bind oxygen strongly; therefore, remov-
ing OH from the surface may prove to be problematic, as seen for
materials at the left edge of the EB[OH] plot in Figure 2.
The experimentally observed trends in CO2 reduction can be
deduced from Figure 3. Au and Ag sit to the right of the CO
desorption line, agreeing with the experimental observations in
Table 1, as well as numerous studies showing CO formation
from CO2 on Au10,34−41 and Ag10,34,42,43 electrodes.
Copper sits atop the volcano but still requires an extreme
overpotential in order to overcome the CO to CHO hydro-
genation step. The theoretically predicted overpotential is in
agreement with experiment, as shown in earlier studies.19,20
Experimentally, the use of CO as a reactant has been found to
produce a similar product spectrum as CO2, and the potential
Figure 3. Limiting potentials (UL), allowing for competitive reaction requirement to reach a current density of 5 mA cm−2 in 0.1 M
mechanisms. Free-energy neutrality lines are plotted as in Figure 1, but
KHCO3 buffer is nearly identical to that of CO2 (although
multiple pathways are possible. Competitive pathways are shown as
lines of the same color, and the more favorable route is shown as a canceling effects of an increased solution pH and decreased
solid line, while the less favorable route is shown dotted. Also shown is mass transport likely play a role),10,33 indicating that the
the binding energy at which adsorbed CO is in equilibrium with potential-limiting step comes after the adsorption of CO on
gaseous CO (at a partial pressure of 1%); the CO would be predicted the catalyst surface.
to desorb at binding energies weaker than this line. As in Figure 1, the The materials in Figure 3 that are known to have significant
EB[CO] values are shown for each metal. The open circles correspond activity in hydrogen evolution are marked with open circles; the
to surfaces that have an experimental exchange current density in the HER would be expected to dominate over any CO2 reduction
hydrogen evolution reaction greater than 10−4 A cm−2;27−29 in these in these materials. However, in these materials, particularly Pt
materials, the hydrogen evolution reaction would be expected to and Ni, the onset of H2 evolution is significantly higher than
dominate over CO2 reduction.
that in the CO2-free system. This can easily be understood as a
poisoning of the catalyst surface by adsorbed CO. The CO is
proceed through a CO* intermediate. (This is based on the
bound very tightly to these materials, and the only means of
experimental11,15,30−33 and theoretical19 findings that CO,
removing it under these conditions is to react it off, which
rather than HCOOH, is an intermediate in hydrocarbon
requires significant potentials. Some of the CO can be seen to
production on Cu; however, this does neglect the possibility react off as CH4 in the case of Ni in Table 1, which was brought
that a formate intermediate could produce more reduced to potentials slightly more negative than that of Cu, in
products than formic acid.) A vertical equilibrium line shows agreement with the more negative potentials needed by the
the binding energy at which gas-phase CO (at a partial pressure volcano relation in Figure 3. In the case of Pt, the hydrogen
of 1%) is more stable than adsorbed CO*; materials to the right evolution current density in Table 1 precluded the potential
of this line would be expected to liberate any produced CO from being raised more negative than −0.68 V versus RHE;
rather than further react it to hydrocarbons. CO* can be however, in experiments in which Pt was taken to more
protonated to form COH* rather than CHO*, which may negative potentials, some hydrocarbons were observed.42 This
change the reaction pathway for the strongest binding is in agreement with a large body of experimental evidence, in
materials. Lines for both of these reactions are shown in the which CO has been observed to poison the electrocatalyst
same color, with the most favorable reaction at any value of surface in CO2 electroreduction experiments on Pt44−48 and
EB[CO] shown as a solid line and the less favorable shown as a Ni49,50 surfaces.
dotted line. For materials that protonate CO* to form COH*, Under these highly reducing conditions, Pd would be ex-
the downstream pathway will be different than that for those pected to form a hydride due to the high effective H fugacity,51
254 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258
The Journal of Physical Chemistry Letters Perspective

which will likely weaken its affinity for carbon52 and shift it to
the right (weaker CO binding) in Figure 3. Experimental
evidence of this occurring can be seen in Table 1, in which a
Faradaic balance was not closed. This behavior has been
observed and exploited numerous times in electrochemical CO2
reduction.54−59 Nevertheless, this study captures the trends that
Pd weakly produces CH4 but is in major competition with the
hydrogen evolution reaction and the formation of a hydride.

Crucially, effective catalysts must be


capable of efficiently catalyzing the
protonation of adsorbed CO to
adsorbed CHO or COH and exhibit
simultaneous poor activity for the
competitive hydrogen evolution
reaction.

Crucially, effective catalysts must be capable of efficiently


catalyzing the protonation of adsorbed CO to adsorbed CHO or
COH and exhibit simultaneous poor activity for the competitive
hydrogen evolution reaction. Regardless of the downstream
pathway after the formation of CHO or COH, products that are
more reduced than CO, such as hydrocarbons and alcohols, will be
produced as long as the subsequent products desorb from the
electrocatalyst surface. We observe that Cu exhibits a slightly better-
than-average ability to perform the hydrogenation of adsorbed CO,
but this is only in conjunction with its relatively poor activity at the
hydrogen evolution reaction that this works, as exhibited by it
having among the most negative potentials in Table 1.
For an electrocatalytic system to be more effective at reduc-
ing CO2 to CH4, a means to protonate CO at a less negative Figure 4. Suggested decoupling strategies exploiting the geometric
potential must be found; in other words, the binding energy differences between adsorbed CHO and adsorbed CO.
EB[CHO] must be strengthened (made more negative) relative
to EB[CO]. For materials at the right side of the volcano of CO*. However, the element with high oxygen affinity may
Figure 3, a possible strategy is to increase the CO(g) partial change the potential-determining step, such that clearing of
pressure, either by letting product gases build up or by OH* from those sites may limit UL. Thus, a balance must
employing a series of electrochemical reactors. However, for be found between these two effects.
materials over the broadest range of carbon-binding energies, • Ligand stabilization. In an analogous concept to alloying,
CO and CHO will both be bound to the surface, and their an electrophilic chemical ligand can be applied in homo-
binding energies must be decoupled. We can use our under- geneous catalysis, in which a ligand can be constructed
standing of surface reactivity to suggest general means in which that is geometrically capable of interacting with the
the binding energetics of CO and CHO may be decoupled, planar geometry of CHO* but not that of the linear CO*.
possibly leading to electrocatalysts with substantially improved This partial bond formation with CHO* will act to
overpotentials relative to that of Cu. Several strategies are listed stabilize the complex, thus lowering the energy of CHO*
below and shown in Figure 4. relative to that of CO*. Many other homogeneous
• Alloying with metals with higher oxygen af f inity. CO tends catalysis techniques may prove beneficial in performing
to bind to surfaces in an upright geometry with respect to this reaction more effectively.60−62
the carbon atom. In contrast, CHO tends to bind in a • Tethering. Homogeneous and heterogeneous approaches
planar geometry with respect to the carbon atom, with can be combined through tethering of a ligand to an
the oxygen not extending directly away from the surface electrode surface, such that the adsorbates (CO* and
as is the case with CO*. (See also Figure 1.) Catalyst CHO*) bind to the electrode surface, but only the planar
materials have varying degrees of affinity for oxygen and CHO* can interact with the tethered ligand. The reverse
for carbon; alloying an element that has high oxygen is also possible, in which the adsorbates bind to the
affinity into an electrocatalyst may allow CHO* to bind ligand and the CHO* interacts with the surface.
to the surface through both the carbon and oxygen atoms, • Addition of promoters. Promoters can change the relative
thus increasing its stability without affecting the stability of binding strength of adsorbates through electronic effects
255 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258
The Journal of Physical Chemistry Letters


Perspective

(ligand effects, induced fields) and through structural effects REFERENCES


(geometry disruption, binding through promoter atoms) or (1) Lewis, N. S.; Nocera, D. G. Powering the Planet: Chemical
through combinations of these effects. It is likely that many Challenges in Solar Energy Utilization. Proc. Natl. Acad. Sci. U.S.A.
promoted catalysts will not follow the same correlation of 2006, 103, 15729−15735.
binding energies of CO and CHO, as is observed on the (2) Hori, Y. Electrochemical CO2 Reduction on Metal Electrodes. In
pure transition metals in this study. The addition of pro- Modern Aspects of Electrochemistry; Springer: New York, 2008; Vol. 42,
moters that are resistant to electrochemical reduction may Chapter 3, pp 89−189.
lead to new possibilities, and one such possibility for geo- (3) Benson, E. E.; Kubiak, C. P.; Sathrum, A. J.; Smieja, J. M.
metrical stabilization is shown in Figure 4, in which the Electrocatalytic and Homogeneous Approaches to Conversion of CO2
promoter atom acts to make the adsorbed promoter− to Liquid Fuels. Chem. Soc. Rev. 2009, 38, 89−99.
(4) Roy, S. C.; Varghese, O. K.; Paulose, M.; Grimes, C. A. Toward
adsorbate complex adsorb in a bidentate manner. Solar Fuels: Photocatalytic Conversion of Carbon Dioxide to
• Hydrogen bond stabilization/solvent ef fects. Due to the Hydrocarbons. ACS Nano 2010, 4, 1259−1278.
differences in binding geometries, hydrogen bond donors (5) Darensbourg, D. J. Chemistry of Carbon Dioxide Relevant to Its
near the surface may interact with CO* and CHO* in Utilization: A Personal Perspective. Inorg. Chem. 2010, 49, 10765−
different fashions. Also, a hydrogen bond acceptor may 10780.
work to preferentially stabilize COH* (with a strong (6) Barton Cole, E.; Lakkaraju, P. S.; Rampulla, D. M.; Morris, A. J.;
hydrogen-bonding group, OH) over CO*. The hydrogen Abelev, E.; Bocarsly, A. B. Using a One-Electron Shuttle for the
binding difference may be achieved by changes in the Multielectron Reduction of CO2 to Methanol: Kinetic, Mechanistic,
electrolyte or by the presence of hydrogen-bonding ligands and Structural Insights. J. Am. Chem. Soc. 2010, 132, 11539−11551.
attached to the catalyst surface. (7) Le, M.; Ren, M.; Zhang, Z.; Sprunger, P. T.; Kurtz, R. L.; Flake,
J. C. Electrochemical Reduction of CO2 to CH3OH at Copper Oxide

■ METHODS
Electronic structure calculations were undertaken using the
Surfaces. J. Electrochem. Soc. 2011, 158, E45−E49.
(8) Rakowski Dubois, M.; Dubois, D. L. Development of Molecular
Electrocatalysts for CO2 Reduction and H2 Production/Oxidation. Acc.
Chem. Res. 2009, 42, 1974−1982.
DACAPO plane wave implementation of density functional theory, (9) Savéant, J. M. Molecular Catalysis of Electrochemical Reactions.
using the (211) facet of a periodic fcc crystal and the RPBE Mechanistic Aspects. Chem. Rev. 2008, 108, 2348−2378.
exchange−correlation functional.63 Full details are available in the (10) Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electrocatalytic
Supporting Information. Process of CO Selectivity in Electrochemical Reduction of CO2 at


Metal Electrodes in Aqueous Media. Electrochim. Acta 1994, 39,
ASSOCIATED CONTENT 1833−1839.
(11) DeWulf, D. W.; Jin, T.; Bard, A. J. Electrochemical and Surface
*
S Supporting Information Studies of Carbon Dioxide Reduction to Methane and Ethylene at
The computational methodolgy, the additional scaling relations Copper Electrodes in Aqueous Solutions. J. Electrochem. Soc. 1989,
necessary to create Figure 3, a comparison to direct 136, 1686−1691.
hydrogenation reactions, and a two-dimensional volcano plot (12) Frese, K., Jr. Electrochemical Reduction of CO2 at Solid
Electrodes. In Electrochemical and Electrocatalytic Reactions of Carbon
illustrating transformation of CH2O* to OCH3*. This material Dioxide; Sullivan, B.; ; Krist, K., Guard, H., Eds.; Elsevier: New York,
is available free of charge via the Internet at http://pubs.acs.org.


1993; Chapter 6, pp 145−216.
(13) Gattrell, M.; Gupta, N.; Co, A. A Review of the Aqueous
AUTHOR INFORMATION Electrochemical Reduction of CO2 to Hydrocarbons at Copper.
J. Electroanal. Chem. 2006, 594, 1−19.
Biographies (14) Hori, Y.; Takahashi, R.; Yoshinami, Y.; Murata, A. Electro-
chemical Reduction of CO at a Copper Electrode. J. Phys. Chem. B
Andrew A. Peterson is an Assistant Professor at Brown University. He 1997, 101, 7075−7081.
conducted this work while a postdoctoral scholar at Stanford. Andrew (15) Schouten, K. J. P.; Kwon, Y.; van der Ham, C. J. M.; Qin, Z.;
also worked with Jens Nørskov as a postdoc at the Technical Koper, M. T. M. A New Mechanism for the Selectivity to C1 and C2
University of Denmark and received his Ph.D. and bachelors from Species in the Electrochemical Reduction of Carbon Dioxide on
MIT and the University of Minnesota, respectively. Copper Electrodes. Chem. Sci. 2011, 2, 1902−1909.
(16) Chaplin, R.; Wragg, A. Effects of Process Conditions and
Jens K. Nørskov is the Leland T. Edwards Professor in Chemical Electrode Material on Reaction Pathways for Carbon Dioxide
Engineering at Stanford University and is the Director of the Electroreduction with Particular Reference to Formate Formation.
SUNCAT Center for Interface Science and Catalysis at Stanford and J. Appl. Electrochem. 2003, 33, 1107−1123.
the SLAC National Accelerator Laboratory. Previously, Jens directed (17) Delacourt, C.; Ridgway, P. L.; Kerr, J. B.; Newman, J. Design of
the Center for Atomic-scale Materials Design at the Technical an Electrochemical Cell Making Syngas (CO + H2) from CO2 and
University of Denmark. H2O Reduction at Room Temperature. J. Electrochem. Soc. 2008, 155,


B42−B49.
(18) Whipple, D. T.; Kenis, P. J. A. Prospects of CO2 Utilization via
ACKNOWLEDGMENTS Direct Heterogeneous Electrochemical Reduction. J. Phys. Chem. Lett.
We appreciate insights from Professor Clifford Kubiak and 2010, 1, 3451−3458.
Dr. Kyle A. Grice of the University of California at San Diego. (19) Peterson, A. A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.;
Nørskov, J. K. How Copper Catalyzes the Electroreduction of Carbon
This material is based on work supported by the Air Force Dioxide into Hydrocarbon Fuels. Energ. Environ. Sci. 2010, 3, 1311−
Office of Scientific Research through the MURI program under 1315.
AFOSR Award No. FA9550-10-1-0572. The authors also (20) Durand, W. J.; Peterson, A. A.; Studt, F.; Abild-Pedersen, F.;
acknowledge support from the Catalysis for Sustainable Energy Nørskov, J. K. Structure Effects on the Energetics of the Electro-
(CASE) initiative, which is funded by the Danish Ministry of chemical Reduction of CO2 by Copper Surfaces. Surf. Sci. 2011, 605,
Science, Technology, and Innovation. 1354−1359.

256 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258


The Journal of Physical Chemistry Letters Perspective

(21) Trasatti, S. Electrocatalysis in the Anodic Evolution of Oxygen (42) Noda, H.; Ikeda, S.; Oda, Y.; Imai, K.; Maeda, M.; Ito, K.
and Chlorine. Electrochim. Acta 1984, 29, 1503−1512. Electrochemical Reduction of Carbon Dioxide at Various Metal
(22) Dau, H.; Limberg, C.; Reier, T.; Risch, M.; Roggan, S.; Strasser, P. Electrodes in Aqueous Potassium Hydrogen Carbonate Solution. Bull.
The Mechanism of Water Oxidation: From Electrolysis via Homo- Chem. Soc. Jpn. 1990, 63, 2459−2462.
geneous to Biological Catalysis. ChemCatChem 2010, 2, 724−761. (43) Yano, H.; Shirai, F.; Nakayama, M.; Ogura, K. Electrochemical
(23) Marković, N. M.; Ross, P. N. Surface Science Studies of Model Reduction of CO2 at Three-Phase (Gas−Liquid−Solid) and Two-
Fuel Cell Electrocatalysts. Surf. Sci. Rep. 2002, 45, 117−229. Phase (Liquid−Solid) Interfaces on Ag Electrodes. J. Electroanal.
(24) Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Chem. 2002, 533, 113−118.
Kitchin, J. R.; Bligaard, T.; Jonsson, H. Origin of the Overpotential for (44) Giner, J. Electrochemical Reduction of CO2 on Platinum
Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, Electrodes in Acid Solutions. Electrochim. Acta 1963, 8, 857−865.
17886−17892. (45) Breiter, M. On the Nature of Reduced Carbon Dioxide.
(25) Greeley, J.; Stephens, I. E. L.; Bondarenko, A. S.; Johansson, Electrochim. Acta 1967, 12, 1213−1218.
T. P.; Hansen, H. A.; Jaramillo, T. F.; Rossmeisl, J.; Chorkendorff, I.; (46) Beden, B.; Bewick, A.; Razaq, M.; Weber, J. On the Nature of
Nørskov, J. K. Alloys of Platinum and Early Transition Metals as Reduced CO2: An IR Spectroscopic Investigation. J. Electroanal. Chem.
Oxygen Reduction Electrocatalysts. Nat. Chem. 2009, 1, 552−556. 1982, 139, 203−206.
(26) Abild-Pedersen, F.; Greeley, J.; Studt, F.; Rossmeisl, J.; Munter, T.; (47) Nikolic, B.; Huang, H.; Gervasio, D.; Lin, A.; Fierro, C.; Adzic, R.;
Moses, P.; Skúlason, E.; Bligaard, T.; Nørskov, J. Scaling Properties of Yeager, E. Electroreduction of Carbon Dioxide on Platinum Single Crystal
Adsorption Energies for Hydrogen-Containing Molecules on Transition- Electrodes: Electrochemical and in situ FTIR Studies. J. Electroanal.
Metal Surfaces. Phys. Rev. Lett. 2007, 99, 16105. Chem. 1990, 295, 415−423.
(27) Trasatti, S. Work Function, Electronegativity, and Electro- (48) Aramata, A.; Enyo, M.; Koga, O.; Hori, Y. FT-IR Spectrometry
chemical Behaviour of Metals: III. Electrolytic Hydrogen Evolution in of the Reduced CO2 at Pt Electrode and Anomalous Effect of Ca2+
Acid Solutions. J. Electroanal. Chem. 1972, 39, 163−184. Ions. Chem. Lett. 1991, 20, 749−752.
(28) Bockris, J. O.; Reddy, A. K. N. Modern Electrochemistry; Plenum (49) Hori, Y.; Koga, O.; Aramata, A.; Enyo, M. Infrared Spectros-
Press: New York, 1970; Vol. 2. copic Observation of Adsorbed CO Intermediately Formed in the
(29) Eberhardt, D.; Santos, E.; Schmickler, W. Hydrogen Evolution Electrochemical Reduction of CO2 at a Nickel Electrode. Bull. Chem.
on Silver Single Crystal Electrodes-First Results. J. Electroanal. Chem. Soc. Jpn. 1992, 65, 3008−3010.
1999, 461, 76−79. (50) Koga, O.; Matsuo, T.; Yamazaki, H.; Hori, Y. Infrared
(30) Hori, Y.; Murata, A.; Takahashi, R.; Suzuki, S. Electroreduction Spectroscopic Observation of Intermediate Species on Ni and Fe
of Carbon Monoxide to Methane and Ethylene at a Copper Electrode Electrodes in the Electrochemical Reduction of CO2 and CO to
in Aqueous Solutions at Ambient Temperature and Pressure. J. Am. Hydrocarbons. Bull. Chem. Soc. Jpn. 1998, 71, 315−320.
Chem. Soc. 1987, 109, 5022−5023. (51) Frieske, H.; Wicke, E. Magnetic Susceptibility and Equilibrium
(31) Kim, J. J.; Summers, D. P.; Frese, K. W. J. R. Reduction of CO2
Diagram of PdHn. Ber. Bunsen-Ges. Phys. Chem. 1973, 77, 48−52.
and CO to Methane on Cu Foil Electrodes. J. Electroanal. Chem. 1988,
(52) Johansson, M.; Skúlason, E.; Nielsen, G.; Murphy, S.; Nielsen, R.;
245, 223−244.
Chorkendorff, I. Hydrogen Adsorption on Palladium and Palladium
(32) Cook, R. L.; MacDuff, R. C.; Sammells, A. F. Evidence for
Hydride at 1 bar. Surf. Sci. 2010, 604, 718−729.
Formaldehyde, Formic Acid, and Acetaldehyde as Possible Inter-
(53) Ikeda, S.; Takagi, T.; Ito, K. Selective Formation of Formic Acid,
mediates during Electrochemical Carbon Dioxide Reduction at
Oxalic Acid, and Carbon Monoxide by Electrochemical Reduction of
Copper. J. Electrochem. Soc. 1989, 136, 1982−1984.
Carbon Dioxide. Bull. Chem. Soc. Jpn. 1987, 60, 2517−2522.
(33) Hori, Y.; Murata, A.; Takahashi, R. Formation of Hydrocarbons
(54) Ayers, W.; Farley, M. Carbon Dioxide Reduction with an
in the Electrochemical Reduction of Carbon Dioxide at a Copper
Electric Field Assisted Hydrogen Insertion Reaction. Catalytic
Electrode in Aqueous Solution. J. Chem. Soc. Faraday Trans. 1 1989,
85, 2309−2326. Activation of Carbon Dioxide; ACS Symposium Series 363; American
(34) Hori, Y.; Kikuchi, K.; Suzuki, S. Production of CO and CH4 in Chemical Society: Washington, DC, 1988; pp 147−154.
Electrochemical Reduction of CO2 at Metal Electrodes in Aqueous (55) Azuma, M.; Hashimoto, K.; Watanabe, M.; Sakata, T.
Hydrogencarbonate Solution. Chem. Lett. 1985, 14, 1695−1698. Electrochemical Reduction of Carbon Dioxide to Higher Hydro-
(35) Azuma, M.; Hashimoto, K.; Hiramoto, M.; Watanabe, M.; carbons in a KHCO3 Aqueous Solution. J. Electroanal. Chem. 1990,
Sakata, T. Carbon Dioxide Reduction at Low Temperature on Various 294, 299−303.
Metal Electrodes. J. Electroanal. Chem. 1989, 260, 441−445. (56) Ohkawa, K.; Hashimoto, K.; Fujishima, A.; Noguchi, Y.;
(36) Hori, Y.; Murata, A.; Kikuchi, K.; Suzuki, S. Electrochemical Nakayama, S. Electrochemical Reduction of Carbon Dioxide on
Reduction of Carbon Dioxides to Carbon Monoxide at a Gold Hydrogen-Storing Materials: Part 1. The Effect of Hydrogen
Electrode in Aqueous Potassium Hydrogen Carbonate. J. Chem. Soc., Absorption on the Electrochemical Behavior on Palladium Electrodes.
Chem. Commun. 1987, 728−729. J. Electroanal. Chem. 1993, 345, 445−456.
(37) Fujihira, M.; Noguchi, T. A Highly Sensitive Analysis of (57) Ohkawa, K.; Noguchi, Y.; Nakayama, S.; Hashimoto, K.;
Electrochemical Reduction Products of CO2 on Gold by New Fujishima, A. Electrochemical Reduction of Carbon Dioxide on
Differential Electrochemical Mass Spectroscopy (DEMS). Chem. Hydrogen-storing Materials: Part 4. Electrochemical Behavior of the
Lett. 1992, 21, 2043−2046. Pd Electrode in Aqueous and Nonaqueous Electrolyte. J. Electroanal.
(38) Kedzierzawski, P.; Augustynski, J. Poisoning and Activation of Chem. 1994, 369, 247−250.
the Gold Cathode during Electroreduction of CO2. J. Electrochem. Soc. (58) Yoshitake, H.; Takahashi, K.; Ota, K.i. Electrochemical
1994, 141, L58−L60. Reduction of CO2 on Hydrogen-Enriched and Hydrogen-Depleted
(39) Noda, H.; Ikeda, S.; Yamamoto, A.; Einaga, H.; Ito, K. Kinetics Surfaces. J. Chem. Soc., Faraday Trans. 1994, 90, 155−159.
of Electrochemical Reduction of Carbon Dioxide on a Gold Electrode (59) Iwakura, C.; Takezawa, S.; Inoue, H. Catalytic Reduction of
in Phosphate Buffer Solutions. Bull. Chem. Soc. Jpn. 1995, 68, 1889− Carbon Dioxide with Atomic Hydrogen Permeating through
1895. Palladized Pd Sheet Electrodes. J. Electroanal. Chem. 1998, 459,
(40) Ohmori, T.; Nakayama, A.; Mametsuka, H.; Suzuki, E. Influence 167−169.
of Sputtering Parameters on Electrochemical CO2 Reduction in (60) Narayanan, B. A.; Amatore, C.; Kochi, J. K. Reduction of Metal
Sputtered Au Electrode. J. Electroanal. Chem. 2001, 514, 51−55. Carbonyls via Electron Transfer. Formation and Chain Decomposition
(41) Stevens, G. B.; Reda, T.; Raguse, B. Energy Storage by the of Formylmetal Intermediates. Organometallics 1986, 5, 926−935.
Electrochemical Reduction of CO2 to CO at a Porous Au Film. (61) Cutler, A. R.; Hanna, P. K.; Vites, J. C. Carbon Monoxide and
J. Electroanal. Chem. 2002, 526, 125−133. Carbon Dioxide Fixation: Relevant C1 and C2 Ligand Reactions

257 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258


The Journal of Physical Chemistry Letters Perspective

emphasizing (η5-C5H5)Fe-Containing Complexes. Chem. Rev. 1988,


88, 1363−1403.
(62) Miller, A. J. M.; Labinger, J. A.; Bercaw, J. E. Reductive Coupling
of Carbon Monoxide in a Rhenium Carbonyl Complex with Pendant
Lewis Acids. J. Am. Chem. Soc. 2008, 130, 11874−11875.
(63) Hammer, B.; Hansen, L. B.; Nørskov, J. K. Improved
Adsorption Energetics within Density-Functional Theory Using
Revised Perdew−Burke−Ernzerhof Functionals. Phys. Rev. B 1999,
59, 7413−7421.

258 dx.doi.org/10.1021/jz201461p | J. Phys. Chem. Lett. 2012, 3, 251−258

You might also like