Sie sind auf Seite 1von 10

Perspective

pubs.acs.org/JPCL

Catalysts and Reaction Pathways for the Electrochemical Reduction


of Carbon Dioxide
Ruud Kortlever, Jing Shen, Klaas Jan P. Schouten, Federico Calle-Vallejo, and Marc T. M. Koper*
Leiden Institute of Chemistry, Leiden University, PO Box 9502, 2300 RA Leiden, The Netherlands
ABSTRACT: The electrochemical reduction of CO2 has
gained signicant interest recently as it has the potential to
trigger a sustainable solar-fuel-based economy. In this
Perspective, we highlight several heterogeneous and molecular
electrocatalysts for the reduction of CO2 and discuss the
reaction pathways through which they form various products.
Among those, copper is a unique catalyst as it yields
hydrocarbon products, mostly methane, ethylene, and ethanol,
with acceptable eciencies. As a result, substantial eort has
been invested to determine the special catalytic properties of
copper and to elucidate the mechanism through which
hydrocarbons are formed. These mechanistic insights, together
with mechanistic insights of CO2 reduction on other metals and molecular complexes, can provide crucial guidelines for the
design of future catalyst materials able to eciently and selectively reduce CO2 to useful products.

he electrocatalytic reduction of carbon dioxide has


attracted the interest of electrochemists and inorganic
chemists for decades as it can facilitate a sustainable lowtemperature redox cycle for energy storage and conversion.1,2
However, major issues still need to be resolved before the
reduction of CO2 to fuels becomes appealing for technological
applications. The main problems holding back electrocatalytic
CO2 reduction are the high overpotentials needed and the poor
product selectivity and faradaic eciency. The high overpotentials and poor product selectivities are the result of
inappropriate adsorption energies of key reaction intermediates.35 The low faradaic eciencies are due to the competition
with the hydrogen evolution reaction (HER), which takes place
in the same range of potentials as CO2 reduction.6,7 Thus, new
catalysts need to be developed that increase the product
selectivity and eciency of electrocatalytic CO2 reduction while
simultaneously lowering the overpotentials.
In recent years, numerous research articles on electrochemical CO2 reduction have focused on the determination of
the reaction mechanism, both experimental8,9 and theoretical,35,10 as more mechanistic insight is expected to lead to
better, tailor-made catalytic systems. In this Perspective, we will
discuss the current trends in understanding and designing
electrocatalysts for CO2 reduction and the pathways through
which various products are formed, focusing on heterogeneous
electrocatalysts and (immobilized) metal complexes. We will
also outline the implications of these recent mechanistic
insights for the design of selective and ecient catalysts for
CO2 reduction.
Theoretical Considerations on the Electroreduction of CO2. The
electrochemical reduction of carbon dioxide can be viewed as a
multiple protonelectron reaction leading to various products
P and water
2015 American Chemical Society

In this Perspective, we will discuss


the current trends in understanding and designing electrocatalysts for CO2 reduction and
the pathways through which
various products are formed,
focusing on heterogeneous electrocatalysts and (immobilized)
metal complexes.
kCO2 + n(H+ + e) P + mH 2O

(1)

The most typical products P in aqueous media, the


coecients k, n, and m, and the equilibrium potentials are
shown in Table 1. Oxalate (C2O42, k = 2, n = 2, m = 0, E0 =
0.59 V versus NHE) has also been detected in signicant
amounts, though primarily in aprotic solvents, for which the
quoted redox potential does not apply. Note that for all
reactions in which an equal number of protons and electrons is
transferred (which corresponds to the vast majority of the
reactions in aqueous media), the reversible hydrogen electrode
(RHE) is the most sensible reference scale, and it will be used
throughout this Perspective.
The thermodynamic theory of multiple protonelectron
transfer predicts that for all reactions in which only 2 electrons
(that is n = 2) are transferred, catalysts will exist that are able to
Received: July 20, 2015
Accepted: September 24, 2015
Published: September 24, 2015
4073

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

Table 1. Main Products of the Electrochemical Reduction of CO2a

product name and formula

E0 (V versus RHE)

carbon monoxide, CO
formic acid, HCOOH
formaldehyde, HCHO
methanol, CH3OH
methane, CH4
ethanol, CH3CH2OH
ethylene, C2H4

1
1
1
1
1
2
2

2
2
4
6
8
12
12

1
0
1
1
2
3
4

0.10
0.20 (for pH < 4); 0.20 + 0.059[pH-4] (for pH > 4)
0.07
0.02
0.17
0.09
0.08

The coecients k, n, and m in eq 1 are provided in each case together with the standard equilibrium potentials.

passing that Cu surfaces with square symmetry, namely, (100)


facets, are able to break certain scaling relations during CO2
reduction to C2 products due to ensemble eects.23
An additional consideration that needs to be taken into
account in the design of suitable electrocatalysts for CO2
reduction is the role of proton-coupled electron transfer.
Most theoretical/computational studies on heterogeneous
electrocatalysts35,22 make use of the computational hydrogen
electrode (CHE) approach to model proton-coupled electron
transfers.24 Within this simple and convenient model, it is
assumed that at every step in the mechanism, concerted
(simultaneous) protonelectron transfer takes place. Importantly, pH and potential eects are not included directly in the
simulations; instead, they are added externally in a linear
fashion. This implies that changes in pH do not lead to actual
changes in activity on the RHE scale as the energies of all steps
are shifted proportionally. In practice, however, many protoncoupled electron-transfer reactions, such as eq 1, show
signicant pH dependence on the RHE scale. In recent
theoretical work, we have shown that this is typically due to the
decoupling of proton and electron transfer at some stage in the
reaction mechanism, for instance, in an elementary step that
includes only electron transfer but not proton transfer.25
Whenever this occurs, the overall reaction rate becomes pHdependent on the thermodynamically relevant RHE scale.
Below, we will show that these considerations are highly
relevant for electrocatalytic CO2 reduction, not only in
optimizing reactivity but also in optimizing selectivity, for
instance, with respect to H2 evolution, which is normally an
important reaction but undesired in the context of CO2
reduction.
Carbon Dioxide Reduction on Copper. Ever since Hori made
his landmark discovery in 1985 that copper has the unique
ability to electrochemically reduce CO2 to hydrocarbons such
as methane and ethylene with good faradaic eciencies in
comparison to other catalysts,26 substantial eort has been
invested to understand the special reactivity of copper for this
reaction.27 It was shown in early work that CO is a key
intermediate in the formation of hydrocarbons from the
reduction of CO2 on copper,28 which is now widely accepted
in the literature. However, proposing a conclusive mechanism
for the reduction of CO2 on copper is challenging, as illustrated
by the recent observation of 16 dierent products formed from
CO2.29 Besides methane and ethylene, these products include a
broad mix of aldehydes, ketones, carboxylic acids, and alcohols,
out of which 12 are C2 or C3 species, showing the complexity of
this reaction. In Figure 1 the potential dependence of these
products is shown, as published recently by Jaramillo et al.29 So
far, detailed mechanistic pathways have been suggested for the
formation of C1 and C2 products, to be discussed below. The
origin of the many other products remains largely unclear,

catalyze the corresponding redox reactions in both directions,


namely, both reduction and oxidation, with negligible overpotentials.11 We will refer to this situation as reversible
catalysis. For instance, hydrogen oxidation and hydrogen
evolution catalyzed on platinum electrodes are good examples
of reversible catalysis. It is worth noting that reversible catalysts
satisfy the Sabatier principle, which means that they bind key
reaction intermediates in an optimal fashion. Optimality is
dened in this context as a compromise in binding strength;
too strong binding leads to catalyst poisoning, while too weak
binding prohibits the commencement of the reaction. For twoelectron transfer reactions, there is typically only a single
intermediate, and hence, the identication of the optimal
catalyst becomes a simple one-dimensional optimization
problem with the ideal catalyst being on top of the so-called
activity volcano. These ideas go back to Parsons12 and
Gerischer13 but have been revived recently14 owing to the
ability to calculate accurate binding energies of presumed
catalytic intermediates by rst-principles quantum chemical
calculations, typically employing the density functional theory
(DFT) formalism. Detailed discussions of the use of such
calculations for the design of (electro)catalysts can be found in
the literature.15,16
Redox reactions involving the transfer of more than two
electrons (and protons) will take place through more than one
catalytic intermediate, which ultimately causes their catalytic
irreversibility.11 This is because the binding energies of certain
adsorbed intermediates follow linear energetic scaling relationships,1719 which often locate the top of activity volcano plots
far from the equilibrium potentials. The best-known example is
the universal energy dierence of 3.2 eV between the *OOH
and *OH intermediates during oxygen reduction and oxygen
evolution, which is considerably larger than the ideal value of
2.46 eV, required for reversible catalysis.11 In the multielectron
transfer reduction of CO2, a similar universal energetic scaling
between *CO and *HCO intermediates leads to an overpotential for the six- or eight-electron reduction of CO2 to
methanol or methane.20
Universal scaling relations are the result of the existence of
similar chemical bonds between various adsorbed species and
dierent catalytic surfaces, for instance, through a single bond
pairing a single valence electron.18 This is the case for
oxygenates (*OR, where R can be H, OH, CH3, CH2CH3,
etc.),19 which imposes substantial intrinsic overpotentials for
the reduction of CO2: the energetic scaling of *OH with
*OCH3 in the pathways to CH4 or CH3OH;4,5 the scaling of
*OH and *OCHCH2 for the reduction to C2H4;3 and the
scaling between *OCHCH2, *OCHCH3, and *OCH2CH3 for
the reduction to CH3CH2OH.3 Therefore, materials where
dierent bonding modes exist need to be designed and
synthesized if scaling relations are to be broken.2022 Note in
4074

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

toward formic acid is thus separate from the hydrocarbon


pathway, which must go through carbon monoxide.30,31 On the
other hand, a carbene species (*CH2) on the surface, formed
from *CO, was proposed to be a common intermediate for the
production of methane and ethylene formation. Methane is
then formed by double protonelectron transfer to the carbene
intermediate, while ethylene is formed by either dimerization of
*CH2 species or CO insertion in a FischerTropsch-like step,
which has also been suggested to be the pathway for the
formation of alcohols.32
In contrast with this carbene mechanism, Peterson et al.
have performed DFT calculations of the reduction of CO2 to
methane on Cu(211) surfaces (which contain short (111)
terraces separated by (100)-like steps), suggesting that in the
thermodynamically most favorable pathway, the second CO
bond is broken only at a late stage of the mechanism (see
Figure 2a).4 After the initial formation of *CO, there is
subsequent hydrogenation to *HCO, *H2CO, and *H3CO
(methoxy), and this methoxy intermediate is reduced to CH4
and *O, which is nally reduced to H2O. Ethylene is formed by
dimerization of HxCO species and subsequent deoxygenation.10
Note that this mechanism includes *H2CO and *H3CO, which
cannot explain the fact that formaldehyde (CH2O) is reduced
only to methanol (CH3OH) and that methanol cannot be
reduced to methane.8 The inclusion of kinetic barriers in the
theoretical analysis made by Nie et al. on Cu(111)5 leads to a
substantially dierent mechanism for the formation of CH4,
which is in better agreement with various experimental
observations, namely, the deactivation of the catalyst while
producing methane presumably by coking30,31,33 and the
reduction of formaldehyde to methanol and not to methane.8
Note that within their analysis, Nie et al. predict the formation
of C2H4 to occur via coupling of *CH2 moieties.
It is important to note that despite their dierent features,
none of the above mechanisms are able to explain three
important experimental observations. The rst observation was
made by Hori et al.,32 namely, that the formation of methane
from CO shows a dierent pH dependence from the formation
of ethylene. The second observation is that during CO2
reduction, often ethylene formation takes place at less negative
potentials without any simultaneous formation of methane. The
third observation links these two observations to the copper
surface structure, based on our own experiments with copper
single-crystal electrodes;34,35 on Cu(111), the reduction of CO

Figure 1. Current eciencies of the products of CO2 reduction on Cu


electrodes in 0.1 M KHCO3 (pH 6.8) as a function of potential for
major (top), intermediate (middle), and minor (bottom) products.
Reproduced from ref 29 with permission from The Royal Society of
Chemistry.

though we emphasize that they occur in such small amounts


that they are only detectable with NMR spectroscopy. Formic
acid is one of the products of CO2 reduction on Cu. Early
mechanistic studies found that formic acid cannot be reduced
to other products, suggesting that the mechanistic pathway

Figure 2. Pathways for the electrochemical production of methane from CO2 on Cu electrodes from (a) a thermodynamic analysis, adapted from ref
4, and (b) a combined thermodynamic and kinetic analysis from ref 5. Species in black are adsorbates, while those in red are reactants or products in
solution. The two mechanisms are identical up to the second protonelectron transfer, that is, until *CO is formed, after which they form (a)
*CHO and (b) *COH.
4075

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

Figure 3. Possible reaction pathways for the electrocatalytic reduction of CO2 to products on transition metals and molecular catalysts: (a) pathways
from CO2 to CO, CH4 (blue arrows), CH3OH (black arrows), and HCOO (orange arrows); (b) pathways from CO2 to ethylene (gray arrows) and
ethanol (green arrows); (c) pathway of CO2 insertion into a metalH bond yielding formate (purple arrows). Species in black are adsorbates, while
those in red are reactants or products in solution. Potentials are reported versus RHE, while RDS indicates rate-determining steps and (H+ + e)
indicates steps in which either concerted or separated protonelectron transfer takes place.

to both methane and ethylene is observed to take place


simultaneously. On the other hand, on Cu(100), a second
pathway is observed that forms ethylene at lower potentials
without methane formation, especially at high (alkaline) pH, in
agreement with Horis observation on the pH dependence of
ethylene formation.32 Earlier work by Hori and co-workers also
showed a strong structural dependence for the reduction of
CO2 and CO on Cu; methane is preferentially formed on
Cu(111), while ethylene is the main product on Cu(100).36
Therefore, the combined evidence suggests the existence of
structure-sensitive and pH-dependent pathways on Cu. On the
basis of this conclusion and taking into account other
experimental observations,8 as well as our own DFT results
on the reduction of CO on Cu(100),3 we propose the
comprehensive mechanism for the reduction of CO2 on copper
shown schematically in Figure 3. In this mechanism, a
distinction is made between the pathway leading to methane
(C1 pathway) and the pathway leading to ethylene (C2
pathway). In the C1 pathway, the CO intermediate is rst
reduced to a formyl species (*CHO) or a *COH species,
which is further reduced to methane. The mechanism assumes
an early breaking of the second CO bond (as was shown
earlier in Figure 2b), though we emphasize that a late breaking

of this CO bond may be more thermodynamically favorable


according to the DFT calculations of Peterson et al. (see Figure
2a).4 Dimerization of the intermediates in this C1 pathway may
also yield ethylene at high applied overpotentials.10 It should be
noted that this pathway toward ethylene is believed to be the
pathway that produces the bulk of the amount of ethylene once
current densities reach 10 mA cm2 and takes place on both
Cu(100) and Cu(111).
In the C2 pathway, the key CC bond-making step at low
overpotentials is a CO dimerization step mediated by electron
transfer rendering a *C2O2 intermediate. Proton transfer
happens only af ter the formation of the negatively charged
adsorbed CO dimer.3,37 Note that this dimerization is the ratedetermining step for the reduction of CO (RDS2 in Figure 3).
Such a decoupling resulting in rst electron then proton
transfer explains why the CO reduction prefers alkaline media.
The CC bond making by the reductive coupling of CO is
somewhat similar to the reductive carbonyl coupling to olens
as originally suggested by McMurry.38 Our DFT calculations
have shown that the formation of the negatively charged CO
dimer is most stable on square arrangements of four surface
atoms,23 explaining the observed preferential formation of
ethylene on Cu(100) at low potentials. At this point, it is
4076

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

Carbon Dioxide Reduction on Other Metals. Recently, Peterson


and Nrskov have outlined the requirements for optimal
catalysts for the reduction of CO2 to methane.20 According to
their work, it is essential that a catalyst eciently catalyzes the
protonation of *CO to *COH or *CHO while simultaneously
showing poor activity for the HER, in order to optimize
methane production. Metals that bind *CO weakly will not
produce methane as CO will desorb before further reduction
can take place, while metals that bind *CO strongly still require
large overpotentials to reduce it to methane because of the
highly unfavorable thermodynamic conditions needed for the
formation of *CHO or *COH. Copper is near the top of the
volcano, indicating that the *CO adsorption strength on copper
is suitable for the production of methane. The binding energies
of *CO on gold and silver are too weak; therefore, they mostly
produce CO and are not able to produce any hydrocarbons.
Metals such as Pt, Pd, and Ni bind CO too strongly.
Consequently, they will be poisoned by *CO, which cannot
be removed from the surface. However, small amounts of
methane have been observed experimentally on Pd and Ni
electrodes.1,49
In earlier reports, the production of hydrocarbons and
alcohols was mostly assumed to be a unique property of Cu
that was absent for other transition metals. However, Jaramillo
et al. recently showed that the production of methane and
methanol is more general than initially thought. They have
shown the production of methane on Fe, methane and
methanol on Au, Zn, Ni, and Pt, and methane and methanol
(and ethanol) on Ag, albeit in small amounts.9,50 Their
proposed mechanism for the formation reduction of CO2 to
methane and methanol on Ag shows similarities to the
mechanism that we proposed for the formation of the same
products on Cu in Figure 3.8,9 Adsorbed CO is rst formed,
and then it is either desorbed or further transformed into a
formyl (*CHO) species or a *COH species on the surface.
These intermediates then react to form methane and/or
methanol.
Transition metals mainly produce carbon monoxide but also
formate.51 It is believed that the formation of CO or formic acid
depends on the initial binding mode of the rst intermediate of
CO2 reduction, as is illustrated in Figure 3a. The precursor that
leads to the formation of CO binds to the catalyst via the C
atom, that is, the carboxyl intermediate (*COOH). In DFT
simulations and in the mechanism shown in Figure 3a,
*COOH is assumed to be formed through a concerted
protonelectron transfer to CO2. However, in some of the
older experimental literature, the formation of *COOH was
considered to take place via the formation of a CO2 radical,
which was suggested to adsorb at copper and gold electrode
surfaces.49,52 The existence of this anionic intermediate implies
the decoupling of proton and electron transfer (see Figure 3a)
and is, therefore, expected to show a dierent pH dependence
from the concerted pathway. A potential example will be
discussed below in the section on molecular catalysts. On the
other hand, the catalytic intermediate that leads to formate or
formic acid is expected to bind to the catalyst through (one of)
the oxygen atoms, either in a monodentate or bidentate fashion,
such that the C atom is available for hydrogenation (see Figure
3a). It is not clear whether this intermediate forms through a
reaction with *H via a CO2 insertion reaction into the metal
hydrogen bond or if the intermediate can be formed through
direct protonation with H+ from solution.

important to note that Bard and co-workers were able to reduce


CO in liquid ammonia to C2O22 using Pt, Ni, and Hg
electrodes, which suggests that CC coupling indeed involves
electron transfer without any proton transfer.39 It is also
assumed that this CO dimerization is the key step in the
recently observed selective reduction of CO to ethanol on
oxide-derived copper electrodes.40
Nanoparticulate copper electrodes have recently gained
signicant attention due to reports that oxide-derived copper
nanoparticles can reduce the onset potentials for both formic
acid and CO compared to polycrystalline copper electrodes.41,42 Furthermore, it was shown that on a roughened
copper-nanoparticle-covered electrode, prepared by electrodeposition of copper, the relative selectivity toward ethylene
over methane could be increased.43,44 Both the oxide-derived
nanoparticles and the electrodeposited nanoparticles showed an
increased stability in comparison to polycrystalline copper,
which generally shows fast deactivation. These observations are
attributed to the surface structure of the nanoparticles as these
tend to contain more defect sites such as kinks and steps. For
instance, a recent study by VerdaguerCasadevall et al. has
linked the activity change for oxide-derived copper catalysts for
the reduction of CO to the number of metastable sites that
bind CO strongly and can exist at grain boundaries.45 However,
we have shown that the initial crystal orientation of Cu2O lms
before reduction has only a minor eect on the product
selectivity and deactivation, while the oxide layer thickness has
a signicant eect.46 Therefore, we hypothesize that the
selectivity change is not only associated with the surface
structure but also (or perhaps even rather) with the increase of
the eective surface area, which in turn has a substantial eect
on the local pH in the boundary layer near the electrode under
reductive conditions.33,46 To support this hypothesis, we have
shown that a low buer capacity of the electrolyte favors
ethylene formation on copper nanoparticles, while increasing
the buer capacity leads to an increase in the selectivity toward
methane.44 This was recently conrmed by Varela et al., who
demonstrated that the activity and selectivity of a smooth
copper electrode can also be tuned by changing the bicarbonate
concentration of the electrolyte.47 Following the reaction
mechanism in Figure 3, this can be explained by the dierent
pH gradients near the electrode surface, with a smaller pH
gradient near the electrode with increasing buer strength. A
low buering capacity leads to a high local pH near the
electrode, yielding an enhanced production of ethylene at such
low buer capacities, in agreement with the earlier observation
that CO can selectively be reduced in the C2 pathway to
ethylene in alkaline media, separate from the C1 pathway
producing both methane and ethylene. Interestingly, when the
selectivity toward methane is increased, a deactivation eect is
observed that is not present when ethylene is the major
product.44 This implies that large amounts of poisoning
intermediates form on copper nanoparticles in the C1 pathway,
and such intermediates do not take part in the C2 pathway.
The eect of Cu nanoparticle size for the reduction of CO2
was recently studied, showing that the reduction of CO2 on Cu
nanoparticles of 2 nm and smaller is signicantly enhanced.48
This increase in activity however is due to an increase in H2 and
CO production, while the eciency toward hydrocarbons
vanishes. These catalytic properties are believed to be due to an
increase in undercoordinated sites, which enhance both the
HER and the reduction of CO2 to CO.
4077

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

transfer reaction, a feature that had previously been


demonstrated only for immobilized electroactive enzymes.67
Role of the Electrolyte. The nature of the electrolyte plays an
important role in electrocatalytic systems for CO2 reduction.
Solvent and pH control the (relative) concentrations of the
reactants, CO2 and H+ or *H. Furthermore, solvent, pH, and
the presence of certain cations or anions can stabilize reaction
intermediates or inhibit their formation and are speculated to
aid directly in the choice of certain reaction pathways.1
Therefore, the use of suitable electrolytes at the right pH
values can be an important step toward more active and
selective electrocatalytic processes.
One important parameter to control proton availability is the
pH of the electrolyte. Given that CO2 forms bicarbonate and
carbonate when it is purged through alkaline solutions,
electrolytes for CO2 reduction are limited to the neutral and
acidic pH range. However, because both the CO2 reduction and
the HER consume protons (or generate OH, alternatively) at
the electrode surface, the local pH near the electrode is typically
higher than that in the bulk. As a result, CO2 near the electrode
can react with OH, forming bicarbonate or carbonate,
depending on the local pH in the proximity of the electrode.
Besides, intermediates of the CO2 reduction such as CO nd a
local pH at the electrode that diers from the bulk pH, which
according to Figure 3 may have signicant consequences on the
selectivity of the reaction. This is because the C1 and C2
pathways have distinctly dierent pH dependence. Although
most studies use buered solutions as electrolytes (either
bicarbonate buers or phosphate buers) to counteract the
overall changes in pH, near the electrode, there might be
signicant pH dierences with respect to the bulk electrolyte.
Gupta et al. have illustrated this phenomenon with some simple
calculations.1,68 To a rst approximation, the local pH at the
electrode is determined by the bulk pH of the electrolyte, its
buer strength, as well as the local current density owing. In
turn, the local current density depends on the local surface
roughness. We believe that this eect is often overlooked when
explaining observed changes in reaction selectivity with changes
in surface structure.46
Several studies have also probed the role of the anions and
cations in the electrolyte.69,70 It has been hypothesized that
cations play a role in electrochemical CO2 reduction either by
specically adsorbing on the electrode, thus inuencing the
potential at the outer Helmholtz plane, or by delivering water
molecules from their solvation shell to the electrode, thus
inuencing the hydrogen coverage on the electrode. Additionally, one could imagine that dierent cations may have dierent
interactions with (negatively charged) intermediates of the
reaction. On Cu electrodes, an increase in the selectivity for the
production of ethylene over methane has been observed with
an increase in the Stokes radii of the hydrated cations, Li+ >
Na+ > K+ > Cs+.69,70 Note that Hori and Murata also reported
hindrance of the HER by using dierent alkaline cations and
obtained the highest selectivity toward hydrocarbons with K+.70
Furthermore, Kyriacou et al. have recently shown that the rate
of reduction can be inuenced by using multivalent cations as
supporting electrolytes.69 A consistent understanding of the
role of cations remains elusive, but they are appealing to tune
the catalysts reactivity and selectivity.
Most studies in electrochemical CO2 reduction have focused
on aqueous electrolytes. However, CO2 dissolves poorly in
water (0.034 M), leading to low concentrations of CO2 in a
saturated aqueous electrolyte. Besides, HER is a prominent

On p-block metals such as In, Sn, Hg, and Pb, formic acid
and formate are the main products.51,53 The mechanistic
pathway by which formate and formic acid are formed on these
metals has been speculated to proceed via a (weakly adsorbed)
CO2 radical that reacts with water to form formate or formic
acid.53 Formic acid and formate can be formed with high
selectivity on these metals because they are poor catalysts for
the competing HER. However, in order to reach these high
selectivities, high overpotentials are needed as the redox
potential for the formation of the CO2 radical is 1.9 V
(versus NHE).51 It should be noted that recently some
improvements have been reported, showing high selectivities
for CO2 reduction at lower overpotentials on nanostructured
Pb and Sn electrodes.54,55 However, we expect that very low
(near-zero) overpotentials for the reduction of CO2 to
formate/formic acid require the stabilization of an adsorbed
formate intermediate as illustrated in the top part of Figure 3.
As was the case for Cu, some studies have probed the eect
of nanoparticles size on the reduction of CO2 on other
metals.56,57 A decrease in the size of Au nanoparticles caused an
increase in the observed current density and a decrease in the
faradaic eciency toward CO. This is believed to be caused by
an increase of undercoordinated sites, as was also hypothesized
for Cu nanoparticles.56 Tuning the size of Pd nanoparticles
showed that an increase in faradaic eciency toward CO can be
achieved by decreasing the particle size. In this case, the
increase in CO production was ascribed to an increase in corner
and edge sites, which display a higher activity toward CO2
reduction to CO.57
A promising area of research is the design of bimetallic
catalysts for the reduction of CO2. Early work showed that the
selectivity of a metal electrode for CO or formic acid can be
altered by the addition of adatoms on the surface.49 Sakata et al.
studied the eects of alloying Cu with Ni, Sn, Pb, Zn, Cd, and
Ag.58 They showed that alloying can have a (benecial) eect
on the onset potentials for the formation of products from CO2
and that some alloys are able to make products that both metals
separately cannot produce in detectable amounts. A Cu/Ni
alloy, for instance, produces methanol at low overpotentials
with a faradaic eciency of 5%, while Cu and Ni separately
produce methanol only in negligible amounts.58 Moreover,
theoretical studies are also enriching the state of the art in the
design of more energetically ecient and chemically selective
bimetallic catalysts for the reduction of CO2. Approaches based
on copper alloys and the so-called isolated active sites aim at
retaining or mimicking the unique reducing features of Cu
while trying to either hinder the HER or reduce the onset
potential.59,60
CuAu alloys and Cu overlayers on Pt have also been
investigated for the production of CO and hydrocarbons,
respectively.6164 For CuAu alloys, it was shown that the
activity and selectivity toward CO production can be tuned by
controlling the electronic and geometric properties,61,62 while
for Cu overlayers on Pt, dierences in the surface strain allowed
tuning the selectivity toward methane or ethylene.64
The potential of bimetallic catalysts is further illustrated by
our own recent work on a PtPd alloy.65,66 This catalyst was
shown to produce formic acid from CO2 with an onset
potential very close to the corresponding equilibrium potential,
with high faradaic eciencies at high current density and very
moderate overpotentials, and with good stability. Furthermore,
this alloy is able to reversibly reduce CO2 to formic acid and
oxidize formic acid to CO2, as expected for a two-electron4078

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

metal centers, either HCOOH is produced as an end product


without consecutive reaction or CO is formed as a precursor to
the more reduced products CH3OH and CH4. MCOOH was
assumed as the rst and only key intermediate. For the
activation of CO2, their calculations suggest a competition
between *H and *COOH adsorption, with *H being more
strongly bound. In order to overcome *H poisoning, they
suggest to reduce CO instead of CO2, and their DFT results
predict Rh centers to be the most active catalysts. Nielsen et al.
have also utilized DFT calculations to investigate the potential
intermediates and the reaction mechanism of the electrochemical reduction of CO2 to CO on a cobalt porphyrin in
water.85 They suggest that CO2 adsorption on cobalt centers is
likely to take place simultaneously with electron transfer to
form a [CoIPCO2]2 intermediate (CoP denoting the Coporphyrin ring), which will be subsequently protonated by a
neighboring H3O+ and the motion of a water molecule toward
the CO 2 group to form the next intermediate
[CoIIPCOOH]. Recent work by Varela et al. on a metaldoped nitrogenated carbon, which is a solid-state analogue of
macrocyclic complexes, showed an independence on the nature
of the metal dopant for the activity of CO2 reduction toward
CO.86 Therefore, it is believed that the production of CO on
this material occurs on the nitrogen functionalities. Methane
production is believed to be dependent on the strength of the
interaction between CO and the metal center, which
determines if CO can be protonated before desorption.
Cobalt complexes have been found to be the most eective
immobilized molecular electrocatalysts for CO2 reduction.
Kapusta and Hackerman found formate as the main product on
a Co phtalocyanine deposited on a carbon electrode, with small
amounts of methanol.87 From Tafel plots, they concluded that
the transfer of the rst electron to CO2 is the rate-determining
step. However, other papers using Co macrocycles report CO
as the main product of electrochemical CO2 reduction in
water.81,83 Recently, our group has obtained novel insights into
the electrochemical reduction of CO2 catalyzed by an
immobilized Co protoporphyrin by combining cyclic voltammetry with online electrochemical mass spectrometry at
dierent values of pH. The pH-dependent formation of CO
indeed suggests the formation of a CO2 anion bound to the
Co macrocycle, with Co presumably being in the CoI state.88
This CO2 anion is then protonated by water, rather than by
H+ as suggested by Nielsen and Leung,85,89 explaining why at
less acidic conditions (pH = 3), CO formation can reach up to
60% faradaic eciency compared to <1% at pH = 1. The
further reduction of CO has a slow rate, resulting in a small but
measurable production (up to 3% faradaic eciency) of the
eight-electron-transfer product CH4, which appears to take
place through concerted electronproton transfer in acidic
media. Smaller amounts of formic acid (and methanol) were
also observed in acidic media, but formic acid was not an
intermediate for the further reaction. The suggested reactions
and pathway for CO2 reduction on the Co-porphyrin are also
contained in Figure 3.
In this Perspective, we have discussed catalysts, mechanistic
features, and reaction pathways for the electrochemical
reduction of CO2. Understanding the mechanisms by which
products are formed can provide crucial insights necessary for
eective catalyst design strategies. Substantial eort has been
invested recently in understanding the mechanism of CO2
reduction on Cu, and several theoretical and experimental
mechanisms exist. The formation of products from CO2 on

reaction in aqueous media and usually accounts for a major part


of the faradaic eciency in CO2 reduction.6,7 To avoid these
problems, solvents with increased solubility of CO2 and lower
proton concentration can be used. In aprotic solvents, such as
dimethyl sulfoxide, N,N-dimethylformamide, propylene carbonate, or acetonitrile, oxalic acid/oxalate has been observed as
the main product on metals such as Pb, In, Sn, and Hg, while
CO and carbonate are the main products on Pt, Pd, Au, and
Zn.51 The production of CO and carbonate can be explained by
a disproportionation reaction between two CO2 molecules,
while oxalic acid/oxalate is made by the reductive coupling of
two CO2 molecules.51,71
Mixtures of water and nonaqueous solvents lead to a broader
product spectrum. Hori et al. employed water/acetonitrile
mixtures to control proton availability,72 showing that the
product selectivity for CO2 reduction on Pt is dependent on the
amount of water in the electrolyte. Without water, oxalic acid is
formed; low amounts of water lead to the formation of formic
acid, and high amounts of water favor hydrogen evolution.
Controlling proton availability by adjusting the amount of
water in an organic electrolyte can thus be an eective way to
control the faradaic eciency and selectivity for CO2 reduction.
Ionic liquids have recently gained signicant attention as
electrolytes for the reduction of CO2. Yu and co-workers have
shown that the ionic liquid EMIMBF3Cl is able to bind CO2 via
a Lewis base adduct and is active in the reduction of CO2.73
Remarkably, this ionic liquid system is able to reduce CO2 to
CO with very high faradaic eciencies at low overpotentials.74
This has been explained by the stabilization of the CO2
intermediate by the cation of the ionic liquid.
Carbon Dioxide Reduction on (Immobilized) Metal Complexes.
There have been intensive investigations focused on molecular
complexes with metal centers as catalysts for the electrochemical CO2 reduction. Most of the existing work on these
materials employs catalysts dissolved in the electrolyte with
their oxidation state recycled on an inert electrode, and this
type of work has been reviewed extensively in recent
years.2,7577 In this Perspective, we will limit ourselves to
molecular catalysts immobilized on electrodes, so as to remain
within the context of heterogeneous catalysis. Moreover, the
eventual application of molecular catalysts in an electrochemical
device will likely involve their immobilization on a surface.
In 1974, Meshitsuka et al. rst utilized nickel or cobalt
phthalocyanine dip coated onto a graphite electrode to catalyze
the electrochemical reduction of CO2.78 In later work, dierent
kinds of metal complexes were employed, such as metal
porphyrins,79 cyclams,80,81 and phthalocyanines.82,83 The twoelectron-transfer products, namely, CO, HCOOH, and oxalic
acid (H2C2O4) are the main products, depending on the
catalyst and the experimental conditions, while the observation
of the six- and eight-electron transfer products CH3OH and
CH4 is quite rare.
Although intensive experimental and theoretical studies have
been performed to understand the mechanism of electrochemical CO2 reduction on metal complexes, the detailed
mechanistic pathways are still uncertain. The most probable
rst intermediates are a CO2 radical bound to the catalyst,
which is subsequently protonated to the MCOOH
intermediate (where M is the metal center of the complex),
or a CO2 inserted into metalH bond to form MOOCH.
Tripkovic et al. have used the CHE approach to study
theoretically the electrochemical reduction of CO2 and CO on
metal-functionalized porphyrin-like graphene.84 With dierent
4079

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

other metals and metal complexes remains largely unclear.


Furthermore, parameters such as electrolyte composition and
pH can change the outcome of the reaction substantially. We
believe that negatively charged intermediates, such as the
adsorbed CO2 anion intermediate and the adsorbed (CO)2
anionic dimer, which play a crucial role in some of the
mechanisms for C1 and C2 formation, are of special interest as
these intermediates are particularly sensitive to pH and solvent
eects.

Catalysis Society Award 2015 and the EFCATS Best Ph.D. Thesis
Award 2015. Currently, he is working at Avantium.
Federico Calle-Vallejo studied chemical engineering at Colombia
(20022007), received his Ph.D. in Denmark in Nrskovs group
(20072011), and was a postdoc in Kopers group (Leiden, 2011
2013) and Sautets group (Lyon, 20132015). He is currently in
Leiden with a Veni grant for young researchers and is interested in
modeling (electro)catalytic processes using DFT.
Marc Koper has been Professor of Fundamental Surface Science at
Leiden University since 2005. His research interests focus on a
fundamental molecular-level understanding of electrocatalytic reactions by combining electrochemical, spectroscopic, and theoretical
techniques. http://casc.lic.leidenuniv.nl/

We believe that negatively


charged intermediates, such as
the adsorbed CO2 anion intermediate and the adsorbed (CO)2
anionic dimer, which play a
crucial role in some of the
mechanisms for C1 and C2 formation, are of special interest as
these intermediates are particularly sensitive to pH and solvent
eects.

ACKNOWLEDGMENTS
This work is supported by NanoNextNL, a micro and
nanotechnology consortium of the Government of The
Netherlands and 130 partners, and the National Research
School Combination Catalysis (NRSC-C). J.S. acknowledges
the award of a grant of the Chinese Scholarship Council
(CSC). F.C.V. acknowledges funding by The Netherlands
Organization for Scientic Research (NWO), Veni Project
Number 722.014.009.

We have discussed some design strategies for the development of ecient and selective catalysts for the electrocatalytic
reduction of CO2. One of these strategies is the alloying of
dierent metals in order to optimize the binding energies of
reaction intermediates like *CO. Alloying might also be good
for breaking scaling relations between similarly adsorbed
intermediates as they cause intrinsic adsorption overpotentials.
Additionally, product selectivity can be inuenced by the pH of
the electrolyte and the use of nonaqueous electrolytes instead
of aqueous electrolytes. Therefore, optimal CO2 reduction
systems require the conjunction of a catalyst with suitable
adsorption properties and an electrolyte with benecial eects
on the catalytic activity and selectivity. We believe that such
delicate catalystelectrolyte interplay can be achieved by means
of a precise, holistic, and consistent understanding of the CO2
reduction mechanism.

REFERENCES

(1) Hori, Y. Electrochemical CO2 Reduction on Metal Electrodes;


Vayenas, C., Ed.; Springer: New York, 2008; Vol. 42.
(2) Rakowski DuBois, M.; DuBois, D. L. Development of Molecular
Catalysts for CO2 Reduction and H2 Production/Oxidation. Acc.
Chem. Res. 2009, 42, 19741982.
(3) Calle-Vallejo, F.; Koper, M. T. M. Theoretical Considerations on
the Electroreduction of CO to C2 Species on Cu(100) Electrodes.
Angew. Chem., Int. Ed. 2013, 52, 72825.
(4) Peterson, A. A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.;
Norskov, J. K. How Copper Catalyzes the Electroreduction of Carbon
Dioxide into Hydrocarbon Fuels. Energy Environ. Sci. 2010, 3, 1311
1315.
(5) Nie, X.; Esopi, M. R.; Janik, M. J.; Asthagiri, A. Selectivity of CO2
Reduction on Copper Electrodes: the Role of the Kinetics of
Elementary Steps. Angew. Chem., Int. Ed. 2013, 52, 245962.
(6) Whipple, D. T.; Kenis, P. J. A. Prospects of CO2 Utilization via
Direct Heterogeneous Electrochemical Reduction. J. Phys. Chem. Lett.
2010, 1, 34513458.
(7) Gattrell, M.; Gupta, N.; Co, A. Electrochemical Reduction of
CO2 to Hydrocarbons to Store Renewable Electrical Energy and
Upgrade Biogas. Energy Convers. Manage. 2007, 48, 12551265.
(8) Schouten, K. J. P.; Kwon, Y.; van der Ham, C. J. M.; Qin, Z.;
Koper, M. T. M. A New Mechanism for the Selectivity to C1 and C2
Species in the Electrochemical Reduction of Carbon Dioxide on
Copper Electrodes. Chem. Sci. 2011, 2, 1902.
(9) Hatsukade, T.; Kuhl, K. P.; Cave, E. R.; Abram, D. N.; Jaramillo,
T. F. Insights into the Electrocatalytic Reduction of CO2 on Metallic
Silver Surfaces. Phys. Chem. Chem. Phys. 2014, 16, 138149.
(10) Montoya, J. H.; Peterson, A. A.; Norskov, J. K. Insights into C-C
Coupling in CO2 Electroreduction on Copper Electrodes. ChemCatChem 2013, 5, 737742.
(11) Koper, M. T. M. Thermodynamic Theory of Multi-Electron
Transfer Reactions: Implications for Electrocatalysis. J. Electroanal.
Chem. 2011, 660, 254260.
(12) Parsons, R. The Rate of Electrolytic Hydrogen Evolution and
the Heat of Adsorption of Hydrogen. Trans. Faraday Soc. 1958, 54,
10531063.
(13) Gerischer, H. Mechanismus der Elektrolytischen Wasserstoffabscheidung und Adsorptionsenergie von Atomarem Wasserstoff. Bull.
Soc. Chim. Belg. 1958, 67, 506527.

AUTHOR INFORMATION

Corresponding Author

*E-mail: m.koper@chem.leidenuniv.nl.
Notes

The authors declare no competing nancial interest.


Biographies
Ruud Kortlever obtained his Masters degree in Chemistry at Leiden
University and is currently a Ph.D. student in the group of Marc Koper
at Leiden University. His research is focused on the development of
ecient and selective nanostructured catalysts for the electrochemical
reduction of CO2.
Jing Shen obtained her Masters degree in Physical Chemistry at the
University of Science and Technology of China and is currently a
Ph.D. student in the group of Marc Koper at Leiden University. Her
research is focused on the electrochemical reduction of CO2 on metal
complexes.
Klaas Jan Schouten obtained his M.Sc. cum laude in 2009. He
continued with Prof. Marc Koper for a Ph.D. project on electrochemical CO2 reduction. His thesis has been awarded with the Dutch
4080

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

CO2 Electroreduction by Process Conditions. ChemElectroChem 2015,


2, 354358.
(34) Schouten, K. J.; Qin, Z.; Gallent, E. P.; Koper, M. T. Two
Pathways for the Formation of Ethylene in CO Reduction on SingleCrystal Copper Electrodes. J. Am. Chem. Soc. 2012, 134, 98647.
(35) Schouten, K. J. P.; Perez Gallent, E.; Koper, M. T. M. Structure
Sensitivity of the Electrochemical Reduction of Carbon Monoxide on
Copper Single Crystals. ACS Catal. 2013, 3, 12921295.
(36) Hori, Y.; Takahashi, I.; Koga, O.; Hoshi, N. Selective Formation
of C2 Compounds from Electrochemical Reduction of CO2 at a Series
of Copper Single Crystal Electrodes. J. Phys. Chem. B 2002, 106, 15
17.
(37) Montoya, J. H.; Shi, C.; Chan, K.; Norskov, J. K. Theoretical
Insights into a CO Dimerization Mechanism in CO2 Electroreduction.
J. Phys. Chem. Lett. 2015, 6, 20322037.
(38) McMurry, J. E.; Fleming, M. P. A New Method for the
Reductive Coupling of Carbonyls to Olefins. Synthesis of B-Carotene.
J. Am. Chem. Soc. 1974, 96, 47084709.
(39) Uribe, F. A.; Sharp, P. R.; Bard, A. J. Electrochemistry in Liquid
Ammonia - Part VI. Reduction of Carbon Monoxide. J. Electroanal.
Chem. Interfacial Electrochem. 1983, 152, 173182.
(40) Li, C. W.; Ciston, J.; Kanan, M. W. Electroreduction of carbon
monoxide to liquid fuel on oxide-derived nanocrystalline copper.
Nature 2014, 508, 504507.
(41) Li, C. W.; Kanan, M. W. CO2 Reduction at Low Overpotential
on Cu Electrodes Resulting from the Reduction of Thick Cu2O Films.
J. Am. Chem. Soc. 2012, 134, 72317234.
(42) Qiao, J.; Jiang, P.; Liu, J.; Zhang, J. Formation of Cu
Nanostructured Electrode Surfaces by an AnnealingElectroreduction
Procedure to Achieve High-Efficiency CO2 Electroreduction. Electrochem. Commun. 2014, 38, 811.
(43) Tang, W.; Peterson, A. A.; Varela, A. S.; Jovanov, Z. P.; Bech, L.;
Durand, W. J.; Dahl, S.; Norskov, J. K.; Chorkendorff, I. The
Importance of Surface Morphology in Controlling the Selectivity of
Polycrystalline Copper for CO2 Electroreduction. Phys. Chem. Chem.
Phys. 2012, 14, 7681.
(44) Chen, C. S.; Handoko, A. D.; Wan, J. H.; Ma, L.; Ren, D.; Yeo,
B. S. Stable and Selective Electrochemical Reduction of Carbon
Dioxide to Ethylene on Copper Mesocrystals. Catal. Sci. Technol. 2015,
5, 161168.
(45) Verdaguer-Casadevall, A.; Li, C. W.; Johansson, T. P.; Scott, S.
B.; McKeown, J. T.; Kumar, M.; Stephens, I. E.; Kanan, M. W.;
Chorkendorff, I. Probing the Active Surface Sites for CO Reduction on
Oxide-Derived Copper Electrocatalysts. J. Am. Chem. Soc. 2015, 137,
980811.
(46) Kas, R.; Kortlever, R.; Milbrat, A.; Koper, M. T.; Mul, G.;
Baltrusaitis, J. Electrochemical CO2 Reduction on Cu2O-Derived
Copper Nanoparticles: Controlling the Catalytic Selectivity of
Hydrocarbons. Phys. Chem. Chem. Phys. 2014, 16, 1219412201.
(47) Varela, A. S.; Kroschel, M.; Reier, T.; Strasser, P. Controlling the
Selectivity of CO2 Electroreduction on Copper: The Effect of the
Electrolyte Concentration and the Importance of the Local pH. Catal.
Today 2015, DOI: 10.1016/j.cattod.2015.06.009.
(48) Reske, R.; Mistry, H.; Behafarid, F.; Roldan Cuenya, B.; Strasser,
P. Particle Size Effects in the Catalytic Electroreduction of CO2 on Cu
Nanoparticles. J. Am. Chem. Soc. 2014, 136, 697886.
(49) Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electroctrocatalytic Process of CO Selectivity in Electrochemical Reduction of
CO2 at Metal Electrodes. Electrochim. Acta 1994, 39, 18331839.
(50) Kuhl, K. P.; Hatsukade, T.; Cave, E. R.; Abram, D. N.;
Kibsgaard, J.; Jaramillo, T. F. Electrocatalytic Conversion of Carbon
Dioxide to Methane and Methanol on Transition Metal Surfaces. J.
Am. Chem. Soc. 2014, 136, 1410713.
(51) Jitaru, M.; Lowry, D. A.; Toma, M.; Toma, B. C.; Oniciu, L.
Electrochemical Reduction of Carbon Dioxide on Flat Metallic
Cathodes. J. Appl. Electrochem. 1997, 27, 875889.
(52) Noda, H.; Ikeda, S.; Yamamoto, A.; Einaga, H.; Ito, K. Kinetics
of Electrochemical Reduction of Carbon-Dioxide on a Gold Electrode

(14) Greeley, J.; Jaramillo, T. F.; Bonde, J.; Chorkendorff, I. B.;


Norskov, J. K. Computational High-Throughput Screening of
Electrocatalytic Materials for Hydrogen Evolution. Nat. Mater. 2006,
5, 909913.
(15) Norskov, J. K.; Bligaard, T.; Rossmeisl, J.; Christensen, C. H.
Towards the Computational Design of Solid Catalysts. Nat. Chem.
2009, 1, 3746.
(16) Calle-Vallejo, F.; Koper, M. T. M. First-Principles Computational Electrochemistry: Achievements and Challenges. Electrochim.
Acta 2012, 84, 311.
(17) Abild-Pedersen, F.; Greeley, J.; Studt, F.; Rossmeisl, J.; Munter,
T.; Moses, P.; Skulason, E.; Bligaard, T.; Nrskov, J. Scaling Properties
of Adsorption Energies for Hydrogen-Containing Molecules on
Transition-Metal Surfaces. Phys. Rev. Lett. 2007, 99, 016105.
(18) Calle-Vallejo, F.; Martnez, J. I.; Garca-Lastra, J. M.; Rossmeisl,
J.; Koper, M. T. M. Physical and Chemical Nature of the Scaling
Relations between Adsorption Energies of Atoms on Metal Surfaces.
Phys. Rev. Lett. 2012, 108, 116103.
(19) Calle-Vallejo, F.; Loffreda, D.; Koper, M. T. M.; Sautet, P.
Introducing Structural Sensitivity into Adsorption-Energy Scaling
Relations by Means of Coordination Numbers. Nat. Chem. 2015, 7,
403410.
(20) Peterson, A. A.; Nrskov, J. K. Activity Descriptors for CO2
Electroreduction to Methane on Transition-Metal Catalysts. J. Phys.
Chem. Lett. 2012, 3, 251258.
(21) Halck, N. B.; Petrykin, V.; Krtil, P.; Rossmeisl, J. Beyond the
Volcano Limitations in ElectrocatalysisOxygen Evolution Reaction.
Phys. Chem. Chem. Phys. 2014, 16, 136828.
(22) Hansen, H. A.; Varley, J. B.; Peterson, A. A.; Nrskov, J. K.
Understanding Trends in the Electrocatalytic Activity of Metals and
Enzymes for CO2 Reduction to CO. J. Phys. Chem. Lett. 2013, 4, 388
392.
(23) Li, H.; Li, Y.; Koper, M. T.; Calle-Vallejo, F. Bond-Making and
Breaking Between Carbon, Nitrogen, and Oxygen in Electrocatalysis. J.
Am. Chem. Soc. 2014, 136, 15694701.
(24) Nrskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.;
Kitchin, J. R.; Bligaard, T.; Jonsson, H. Origin of the Overpotential for
Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108,
1788617892.
(25) Koper, M. T. M. Theory of Multiple ProtonElectron Transfer
Reactions and its Implications for Electrocatalysis. Chem. Sci. 2013, 4,
27102723.
(26) Hori, Y.; Kikuchi, K.; Suzuki, S. Production of CO and CH4 in
Electrochemical Reduction of CO2 at Metal Electrodes in Aqueous
Hydrogencarbonate Solution. Chem. Lett. 1985, 11, 16951698.
(27) Gattrell, M.; Gupta, N.; Co, A. A Review of the Aqueous
Electrochemical Reduction of CO2 to Hydrocarbons at Copper. J.
Electroanal. Chem. 2006, 594, 119.
(28) Hori, Y.; Murata, A.; Takahashi, R.; Suzuki, S. Electroreduction
of CO to CH4 and C2H4 at a Copper Electrode in Aqueous Solutions
at Ambient Temperature and pressure. J. Am. Chem. Soc. 1987, 109,
50225023.
(29) Kuhl, K. P.; Cave, E. R.; Abram, D. N.; Jaramillo, T. F. New
Insights into the Electrochemical Reduction of Carbon Dioxide on
Metallic Copper Surfaces. Energy Environ. Sci. 2012, 5, 7050.
(30) Hori, Y.; Murata, A.; Takahashi, R. Formation of Hydrocarbons
in the Electrochemical Reduction of Carbon Dioxide at a Copper
Electrode in Aqueous Solution. J. Chem. Soc., Faraday Trans. 1 1989,
85 (8), 23092326.
(31) Cook, R. L.; MacDuff, R. C.; Sammells, A. F. Evidence for
Formaldehyde, Formic Acid, and Acetaldehyde as Possible Intermediates during Electrochemical Carbon Dioxide Reduction at
Copper. J. Electrochem. Soc. 1989, 136, 19821984.
(32) Hori, Y.; Takahashi, R.; Yoshinami, Y.; Murata, A. Electrochemical Reduction of CO at a Copper Electrode. J. Phys. Chem. B
1997, 101, 70757081.
(33) Kas, R.; Kortlever, R.; Ylmaz, H.; Koper, M. T. M.; Mul, G.
Manipulating the Hydrocarbon Selectivity of Copper Nanoparticles in
4081

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

The Journal of Physical Chemistry Letters

Perspective

in Phosphate Buffer Solutions. Bull. Chem. Soc. Jpn. 1995, 68, 1889
1895.
(53) Chaplin, R. P. S.; Wragg, A. A. Effects of Process Conditions and
Electrode Material on Reaction Pathways for Carbon Dioxide
Electroreduction with Particular Reference to Formate Formation. J.
Appl. Electrochem. 2003, 33, 11071123.
(54) Chen, Y.; Kanan, M. W. Tin Oxide Dependence of the CO2
reduction Efficiency on Tin Electrodes and Enhanced Activity for Tin/
Tin Oxide Thin-Film Catalysts. J. Am. Chem. Soc. 2012, 134, 1986
1989.
(55) Zhang, S.; Kang, P.; Meyer, T. J. Nanostructured Tin Catalysts
for Selective Electrochemical Reduction of Carbon Dioxide to
Formate. J. Am. Chem. Soc. 2014, 136, 17347.
(56) Mistry, H.; Reske, R.; Zeng, Z.; Zhao, Z. J.; Greeley, J.; Strasser,
P.; Cuenya, B. R. Exceptional Size-Dependent Activity Enhancement
in the Electroreduction of CO2 over Au Nanoparticles. J. Am. Chem.
Soc. 2014, 136, 164736.
(57) Gao, D.; Zhou, H.; Wang, J.; Miao, S.; Yang, F.; Wang, G.;
Wang, J.; Bao, X. Size-Dependent Electrocatalytic Reduction of CO2
over Pd Nanoparticles. J. Am. Chem. Soc. 2015, 137, 428891.
(58) Watanabe, M.; Shibata, M.; Kato, A.; Azuma, M.; Sakata, T.
Design of Alloy Electrocatalysts for CO2 Reduction. J. Electrochem. Soc.
1991, 138, 33823389.
(59) Hirunsit, P.; Soodsawang, W.; Limtrakul, J. CO2 Electrochemical Reduction to Methane and Methanol on Copper-Based
Alloys: Theoretical Insight. J. Phys. Chem. C 2015, 119, 82388249.
(60) Karamad, M.; Tripkovic, V.; Rossmeisl, J. Intermetallic Alloys as
CO Electroreduction CatalystsRole of Isolated Active Sites. ACS
Catal. 2014, 4, 22682273.
(61) Christophe, J.; Doneux, T.; Buess-Herman, C. Electroreduction
of Carbon Dioxide on Copper-Based Electrodes: Activity of Copper
Single Crystals and CopperGold Alloys. Electrocatalysis 2012, 3,
139146.
(62) Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Synergistic
geometric and electronic effects for electrochemical reduction of
carbon dioxide using goldcopper bimetallic nanoparticles. Nat.
Commun. 2014, 5, 4948.
(63) Varela, A. S.; Schlaup, C.; Jovanov, Z. P.; Malacrida, P.; Horch,
S.; Stephens, I. E. L.; Chorkendorff, I. CO2 Electroreduction on WellDefined Bimetallic Surfaces: Cu Overlayers on Pt(111) and Pt(211). J.
Phys. Chem. C 2013, 117, 2050020508.
(64) Reske, R.; Duca, M.; Oezaslan, M.; Schouten, K. J. P.; Koper, M.
T. M.; Strasser, P. Controlling Catalytic Selectivities during CO2
Electroreduction on Thin Cu Metal Overlayers. J. Phys. Chem. Lett.
2013, 4, 24102413.
(65) Kortlever, R.; Balemans, C.; Kwon, Y.; Koper, M. T. M.
Electrochemical CO2 Reduction to Formic Acid on a Pd-Based Formic
Acid Oxidation Catalyst. Catal. Today 2015, 244, 5862.
(66) Kortlever, R.; Peters, I.; Koper, S.; Koper, M. T. M.
Electrochemical CO2 Reduction to Formic Acid at Low Overpotential
and with High Faradaic Efficiency on Carbon-Supported Bimetallic
PdPt Nanoparticles. ACS Catal. 2015, 5, 39163923.
(67) Armstrong, F. A.; Hirst, J. Reversibility and Efficiency in
Electrocatalytic Energy Conversion and Lessons from Enzymes. Proc.
Natl. Acad. Sci. U. S. A. 2011, 108, 1404954.
(68) Gupta, N.; Gattrell, M.; MacDougall, B. Calculation for the
Cathode Surface Concentrations in the Electrochemical Reduction of
CO2 in KHCO3 Solutions. J. Appl. Electrochem. 2006, 36, 161172.
(69) Schizodimou, A.; Kyriacou, G. Acceleration of the Reduction of
Carbon Dioxide in the Presence of Multivalent Cations. Electrochim.
Acta 2012, 78, 171176.
(70) Murata, A.; Hori, Y. Product Selectivity Affected by Cationic
Species in Electrochemical Reduction of CO2 and CO at a Cu
Electrode. Bull. Chem. Soc. Jpn. 1991, 64, 123127.
(71) Amatore, C.; Saveant, J.-M. Mechanism and Kinetic Characteristics of the Electrochemical Reduction of Carbon Dioxide in Media of
Low Proton Availability. J. Am. Chem. Soc. 1981, 103, 50215023.

(72) Tomita, Y.; Teruya, S.; Koga, O.; Hori, Y. Electrochemical


Reduction of Carbon Dioxide at a Platinum Electrode in AcetonitrileWater Mixtures. J. Electrochem. Soc. 2000, 147, 41644167.
(73) Snuffin, L. L.; Whaley, L. W.; Yu, L. Catalytic Electrochemical
Reduction of CO2 in Ionic Liquid EMIMBF3Cl. J. Electrochem. Soc.
2011, 158, F155.
(74) Rosen, B. A.; Salehi-Khojin, A.; Thorson, M. R.; Zhu, W.;
Whipple, D. T.; Kenis, P. J.; Masel, R. I. Ionic Liquid-Mediated
Selective Conversion of CO2 to CO at Low Overpotentials. Science
2011, 334 (6056), 6434.
(75) Saveant, J.-M. Molecular Catalysis of Electrochemical Reactions.
Mechanistic Aspects. Chem. Rev. 2008, 108, 23482378.
(76) Appel, A. M.; Bercaw, J. E.; Bocarsly, A. B.; Dobbek, H.; DuBois,
D. L.; Dupuis, M.; Ferry, J. G.; Fujita, E.; Hille, R.; Kenis, P. J.; et al.
Frontiers, Opportunities, and Challenges in Biochemical and Chemical
Catalysis of CO2 Fixation. Chem. Rev. 2013, 113, 662158.
(77) Schneider, J.; Jia, H.; Muckerman, J. T.; Fujita, E.
Thermodynamics and Kinetics of CO2, CO, and H+ Binding to the
Metal Centre of CO2 Reduction Catalysts. Chem. Soc. Rev. 2012, 41,
203651.
(78) Meshitsuka, S.; Ichikawa, M.; Tamaru, K. Electrocatalysis by
Metal Phtalocyanines in the Reduction of Carbon Dioxide. J. Chem.
Soc., Chem. Commun. 1974, 158159.
(79) Sonoyama, N.; Kirii, M.; Sakata, T. Electrochemical Reduction
of CO2 at Metal-Phophyrin Supported Gas Diffusion Electrodes under
High Pressure CO2. Electrochem. Commun. 1999, 1, 213216.
(80) Beley, M.; Collin, J. P.; Ruppert, R.; Sauvage, J. P.
Electrocatalytic Reduction of CO2 by Ni cyclam2+ in Water; Study
of the Factors Affecting the Efficiency and the Selectivity of the
Process. J. Am. Chem. Soc. 1986, 108, 74617467.
(81) Fisher, B.; Eisenberg, R. Electrocatalytic Reduction of CO2 by
Using Macrocycles of Ni and Co. J. Am. Chem. Soc. 1980, 102, 7361
7363.
(82) Magdesieva, T. V.; Yamamoto, T.; Tryk, D. A.; Fujishima, A.
Electrochemical Reduction of CO2 with Transition Metal Phthalocyanine and Porphyrin Complexes Supported on Activated Carbon
Fibers. J. Electrochem. Soc. 2002, 149, D89D95.
(83) Lieber, C. M.; Lewis, N. S. Catalytic Reduction of CO2 at
Carbon Electrodes Modified with Cobalt Phtalocyanine. J. Am. Chem.
Soc. 1984, 106, 50345035.
(84) Tripkovic, V.; Vanin, M.; Karamad, M.; Bjorketun, M. E.;
Jacobsen, K. W.; Thygesen, K. S.; Rossmeisl, J. Electrochemical CO2
and CO Reduction on Metal-Functionalized Porphyrin-like Graphene.
J. Phys. Chem. C 2013, 117, 91879195.
(85) Nielsen, I. M. B.; Leung, K. Cobalt-Porphyrin Catalyzed
Electrochemical Reduction of Carbon Dioxide in Water. 1. A Density
Functional Study of Intermediates. J. Phys. Chem. A 2010, 114, 10166
10137.
(86) Varela, A. S.; Ranjbar Sahraie, N.; Steinberg, J.; Ju, W.; Oh, H.
S.; Strasser, P. Metal-Doped Nitrogenated Carbon as an Efficient
Catalyst for Direct CO2 Electroreduction to CO and Hydrocarbons.
Angew. Chem., Int. Ed. 2015, 54, 1075810762.
(87) Kapusta, S.; Hackerman, N. Carbon Dioxide Reduction at a
Metal Phthalocyanine Catalyzed Carbon Electrode. J. Electrochem. Soc.
1984, 131, 15111514.
(88) Shen, J.; Kortlever, R.; Kas, R.; Birdja, Y.; Diaz-Morales, O.;
Kwon, Y.; Ledezma-Yanez, I.; Schouten, K. J. P.; Mul, G.; Koper, M. T.
M. Electrocatalytic Reduction of Carbon Dioxide to Carbon Monoxide
and Methane at an Immobilized Cobalt Protoporphyrin. Nat.
Commun. 2015, 6, 8177.
(89) Leung, K.; Nielsen, I. M. B.; Sai, N.; Medforth, C.; Shelnutt, J. A.
Cobalt-porphyrin catalyzed electrochemical reduction of carbon
dioxide in water. 2. Mechanism from first principles. J. Phys. Chem.
A 2010, 114, 1017410184.

4082

DOI: 10.1021/acs.jpclett.5b01559
J. Phys. Chem. Lett. 2015, 6, 40734082

Das könnte Ihnen auch gefallen