You are on page 1of 11

CHEMCATCHEM

MINIREVIEWS

DOI: 10.1002/cctc.201402669

Nanostructured Metallic Electrocatalysts for Carbon


Dioxide Reduction
Qi Lu,[a, b] Jonathan Rosen,[a] and Feng Jiao*[a]

The recent development of nanostructured electrocatalysts for alytic properties and potential challenges of nanostructured
CO2 reduction has attracted much attention because they ex- electrocatalysts are summarized. The behavior of nanosized
hibit unique properties compared to their bulk counterparts. In catalysts in ionic liquid electrolytes is also discussed. Ideas for
this minireview, the latest studies on electrocatalytic CO2 re- the design of next-generation electrocatalysts by taking the
duction with a focus on the advances of nanostructured metal- advantages of nanostructured materials are proposed.
lic electrocatalysts are reviewed and discussed. The distinct cat-

Introduction

The reduction in greenhouse CO2 emissions from fossil fuel uti- One of the most difficult challenges is that a key intermediate
lization is critical for human society.[1] Ideally, one would like to species CO2C , formed by an electron transfer to a CO2 mole-
convert CO2 produced in power plants, refineries, and petro- cule, requires a highly negative potential of 1.90 V versus the
chemical plants to liquid fuels or useful chemicals using renew- standard hydrogen electrode (SHE) in aqueous media under
able energy as the energy input.[2] Over the past few decades, standard conditions to drive the reaction.[6] Following this key
catalytic CO2 reduction has been investigated actively using reaction step, there are several proton-assisted multiple-elec-
thermochemical,[3] photochemical,[4] and electrochemical ap- tron-transfer processes that are more favorable to occur at po-
proaches.[5] The thermochemical CO2 conversion through a re- tentials more positive than that.[7] In principle, CO2 could be
forming process requires not only high reaction temperatures converted into a variety of C1 compounds, such as CO, for-
and pressures but also an equivalent amount of hydrogen as mate, methanol, and methane, through a proton-assisted cata-
the reducing agent, which could be problematic energetically lytic process. As a result of the similarity of their proton-cou-
to implement in large-scale applications. In the case of photo- pled reduction potentials, the product selectivity becomes
chemical processes, a few systems have been reported to be a cause for concern. In addition, electrochemical CO2 reduction
photocatalytically active for CO2 reduction, although the selec- in an aqueous electrolyte has to compete with the hydrogen
tivity and production rate of these systems are still too low to evolution reaction (HER), which has a similar equilibrium po-
be economically valid. Conversely, electrochemical CO2 reduc- tential as the CO2 reduction reaction but requires a much
tion has become attractive because it has several characteristic lower overpotential to occur. Therefore, ideal electrocatalysts
advantages over other approaches. For example, the reaction for CO2 reduction must be able to reduce CO2 at low overpo-
can be conducted under ambient conditions and the reaction tentials but high reaction rates (i.e., currents) while hydrogen
rate can be controlled easily by tuning the external bias (i.e., evolution is suppressed to yield the desired products
overpotential); the products produced at different electrodes selectively.
can be separated naturally using individual reaction chambers, Bulk metals have been studied intensively as CO2 reduction
which minimizes the cost associated with postreaction electrocatalysts over the past few decades. The properties of
separation. bulk metallic CO2 reduction catalysts have been reviewed in
As CO2 is a fully oxidized and thermodynamically stable mol- detail by Hori.[7] Recent efforts have been devoted to the de-
ecule, this solution imposes several major technical challenges. velopment of nanostructured metallic electrocatalysts because
they possess several attractive qualities over bulk catalysts. The
clearest advantage is that nanostructured materials usually
provide much more surface active sites than bulk materials be-
[a] Dr. Q. Lu, J. Rosen, Prof. Dr. F. Jiao
cause of their enhanced surface areas. The catalytic activity of
Center for Catalytic Science and Technology (CCST)
Department of Chemical and Biomolecular Engineering heterogeneous catalysts is usually proportional to the number
University of Delaware of surface active sites, which makes nanostructuring an effec-
150 Academy Street, Newark, DE 19716 (USA) tive approach to boost performance. In addition, nanostruc-
E-mail: jiao@udel.edu
tured electrocatalysts have been shown to improve catalytic
[b] Dr. Q. Lu
stability.[8] This is likely because of the increased tolerance to
Department of Chemical Engineering
Columbia University heavy metal impurities in the electrolyte, which is a common
500 W. 120th Street, Mudd 801, New York, NY 10027 (USA) problem for the stability of electrocatalysts.[9] Although a ppm

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &1&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

level of impurities is sufficient to poison a bulk catalyst, nano-


structured electrocatalysts are able to accommodate impurities
or contamination much better because of their large surface
area. Moreover, the nanostructured catalyst surface usually
contains a large portion of edge/low-coordinated sites, which
have catalytic behavior that is very different from flat surface/
fully coordinated sites in their bulk counterparts.[10] These dis-
tinct properties of nanostructured catalysts have been re-
viewed for several important electrocatalytic reactions, such as
oxygen reduction,[11] hydrogen evolution,[12] and oxygen evolu-
tion,[13] which is not the case for electrochemical CO2
reduction.
This article focuses mainly on the recent development of
nanostructured metallic catalysts for electrochemical CO2 re-
duction, which includes the synthesis of specific metallic nano-
structures, their unique catalytic properties, and their reaction
mechanisms. Future strategies are proposed for the design of
highly selective and efficient CO2 reduction electrocatalysts
and the integration of the catalyst into high-performance CO2
electrolyzers. Notably, some important studies on nonmetal-
lic[14] and homogeneous CO2 reduction electrocatalysts[15] are
excluded because they are outside of the scope of this article.

Nanostructured Ag Catalysts
Historically, the metallic Ag surface has shown a good selectivi-
ty to convert CO2 to CO in an aqueous electrolyte. Importantly,
such reduction activity occurs at an overpotential smaller than
that of most other monometallic surfaces.[7] This is most likely
because 1) the CO2 reduction reaction is favored on the Ag
surface because of its appropriate binding energy to COOH
and CO[16] and 2) the competing HER is not favored because
the surface binding energy to atomic H is too low.[17] In 1994,
Hori et al. reported that a polycrystalline Ag surface is able to
reduce CO2 to CO with an efficiency of approximately 80 % at
0.97 V versus the reversible hydrogen electrode (RHE), and the
rest of the electrons were used in the reduction of hydrogen
and the formation of a trace amount of formate.[18] The authors
studied the product selectivity and adsorption of CO2C in dif-
ferent electrolytic media and proposed a reaction scheme in Figure 1. a) Tafel plot of the partial current density to each product. Indica-
tions for the low-, intermediate-, and high-overpotential regions are shown
an aqueous electrolyte as shown in Equations (1)–(4):
above the plot. b) Faradaic efficiency for each product as a function of po-
tential. Reproduced with permission from Ref. [18].
CO2 þe þ* ! * CO2 C  ð1Þ
* CO2 C  þHþ ðaqÞ ! * COOH ð2Þ
þ 
* COOHþH ðaqÞ þe ! * COþH2 O ð3Þ reduce CO2 to CO is approximately 1.0 to 1.2 V (vs. RHE), in
which the electrochemical driving force is sufficient for the re-
* CO ! COþ* ð4Þ
action but not large enough to promote significant HER, and
3) trace amounts of methanol, ethanol, and methane can be
in which * denotes either a surface-bound species or formed at potentials lower than 1.2 V (vs. RHE; Figure 1). No-
a vacant catalytically active site. However, a systematic catalytic tably, methanol and ethanol have never been reported to form
study over a wide potential window was not provided. Recent- on an Ag surface. The formation of CC bonds in ethanol is
ly, Hatsukade et al. performed a more complete evaluation of particularly intriguing in the electrochemical reduction of CO2,
a polycrystalline Ag surface in aqueous media at ambient con- and only a few surfaces have shown such a capability. The au-
ditions using a highly sensitive custom reactor.[19] Their results thors proposed that *CO might receive another pair of e and
show that: 1) the data obtained by Hori et al. agree well with H+(aq) to form more stable *COH intermediates that can be fur-
a modern examination, 2) the optimal potential window to ther reduced, although additional studies are required to eluci-

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &2&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

date the mechanism. Nevertheless, this work established the The reduction of CO2 to CO on an Ag surface is a two-elec-
baseline of CO2 reduction on a polycrystalline Ag surface in an tron-transfer process. One electron is first transferred to a CO2
aqueous medium. molecule to form a CO2C intermediate species absorbed on
To utilize the Ag surface to construct high-performance, the metallic surface. Subsequently, another electron is trans-
practical electrocatalysts to convert CO2 to CO, Lu et al. devel- ferred to the CO2C anion with two protons to form one CO
oped a dealloy process to synthesize a nanoporous Ag (np-Ag) and one H2O molecule. Previous studies indicate that the first
catalyst with a monolithic structure and highly curved inner step proceeds at a much more negative potential than the fol-
surfaces (Figure 2 a–c).[8b] The resulting catalytic materials ach- lowing steps, and it is therefore, rate-determining for the
whole process. This is supported by the Tafel slope of
132 mV dec1 for polycrystalline Ag, close to the calculated
value of approximately 120 mV dec1.[20] However, for np-Ag,
the Tafel slope of 58 mV dec1 reveals a fast initial electron-
transfer step before a successive non-electron-transfer, rate-de-
termining step,[20, 21] which suggests strongly that the np-Ag
surface is able to stabilize the CO2C intermediate better than
a flat surface.
On the highly curved np-Ag surface, stepped surfaces such
as Ag(2 11) and Ag(11 0) are much more prevalent than on the
bulk polycrystalline surface. It is exceedingly possible that the
large amount of low-coordinated Ag atoms on the np-Ag sur-
face are able to reduce the activation energy of CO2 to CO2C
conversion. A demonstration was performed using DFT calcula-
tions to compare the free-energy change of each proton–elec-
tron pair transfer on various model Ag surfaces. Ag surfaces
with steps and edges, namely, Ag(2 11) and Ag(11 0), exhibit
significantly lower free-energy changes for the first proton-cou-
pled electron transfer in CO2 reduction [a combination of
Eqs. (1) and (2), to result in *COOH] than the closed packed
surfaces [Ag(111) and Ag(1 0 0); Figure 2 e]. A similar trend is
observed for the *CO step, as Ag(2 11) exhibits a lower overall
free-energy change that avoids the overbinding of *CO. Nota-
bly, such theoretical calculations are consistent with the experi-
mental study of Hoshi et al.[22] on several single-crystalline Ag
Figure 2. a) Schematic diagram of a nanopore of the Ag electrocatalyst with
surfaces, although the catalytic behavior of an Ag surface with
a highly curved internal surface. b) Scanning electron micrograph of np-Ag.
c) Corresponding high-resolution transmission electron micrograph with visi- an index higher than (11 0) was not evaluated.
ble lattice fringes. Inset: the Fourier transform shows that the np-Ag is com-
posed of an extended crystalline network. d) Overpotential versus CO pro-
duction partial current density on polycrystalline Ag and np-Ag. e) Free- Nanostructured Au and Cu Catalysts
energy diagrams for the electrochemical reduction of CO2 to CO on flat
[Ag(111) and Ag(1 0 0)] and edge [Ag(2 2 1) and Ag(11 0)] surfaces. Repro- A metallic Au surface is currently the most efficient catalytic
duced with permission from Ref. [8b]. surface to convert CO2 to CO selectively. Although the scarcity
of Au may prevent implementation on a large scale, it remains
an ideal material for fundamental study because of its high ac-
ieved an exceptionally high activity of approximately tivity and stable chemical and electrochemical properties. Zhu
20 mA cm2 with over 90 % selectivity at a mediocre potential et al. performed an interesting combined theoretical and ex-
of 0.6 V (vs. RHE) in an aqueous environment, which is the new perimental modeling study based on Au nanoparticles (Au
benchmark performance for Ag-based electrocatalysts. After NPs).[23] The authors synthesized a series Au NPs with different
the differentiation of contributions from the enhanced surface particle sizes of 4, 6, 8, and 10 nm. These particles were poly-
area and the specific surface of np-Ag, it was discovered that crystalline with crystallite diameters of 2.0, 2.3, 4.0, and 5.9 nm,
the intrinsic activity of the np-Ag surface is at least 20 times respectively. In an electrocatalytic study, the 8 nm Au NPs ex-
higher than that of the polycrystalline surface. To understand hibited a better gravimetric activity than their larger or smaller
the origins of such an improvement, Tafel analysis was per- counterparts (Figure 3 a). The DFT study on different crystal
formed on both the np-Ag surface and polycrystalline Ag sur- faces and a 13-atom Au cluster (Au13) suggested that 1) the
face to gain mechanistic insights. Although the Tafel slope of edge sites on Au NP surfaces favored CO2 reduction and 2) the
the polycrystalline Ag surface was 132 mV dec1, the np-Ag corner sites favored the HER. The authors examined the rela-
surface exhibited a dramatic decrease in Tafel slope to tionship between the density of catalytically active surface
58 mV dec1, which further proved its intrinsically better per- sites and the Au cluster size (Figure 3 b). It was found that the
formance (Figure 2 d). 8 nm Au NPs with a 4 nm crystallite diameter are small enough

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &3&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

Notably, among the 16 products observed, five of them have


not been reported previously. It can be seen that 1) in the low
overpotential region (0.6 to 0.8 V vs. RHE), only the four
major products, namely, CO, formate, methane, and ethylene
are observed, 2) in a higher overpotential region (0.9 to
1.15 V vs. RHE), more than 10 additional C1–C3 products are
observed, and 3) at very high overpotentials (1.15 V vs. RHE),
all products are suppressed expect for methane and the HER is
promoted (Figure 4). On the basis of these results, the authors
proposed several hypotheses to explain the formation of the
wide variety of high-C products: 1) the different surface sites
on polycrystalline Cu might be geared to different products,
2) the same type of surface sites were able to catalyze the for-
mation of different C1 oxygenates that could be accessed and
coupled for CC bonds at high overpotentials, and 3) the de-
hydroxylation of surface intermediates that involve 2 H+ and
2 e could be repeated to form a wide range of products. On
the basis of these hypotheses, the authors proposed several
plausible pathways for the production of C2 and C3 species.
Certainly, further experimental work is needed to elucidate the
CO2 reduction mechanisms on a polycrystalline Cu surface
fully.
Interestingly, if the Cu surface was implemented as a nano-
scale material, its ability to convert CO2 was harmed greatly by
a significant promotion of the HER. Reske et al. examined the
catalytic behavior of Cu nanoparticles (Cu NPs) with different
particle sizes that ranged from approximately 2–15 nm.[26] A
significant increase of H2 selectivity can be seen on Cu NPs in
comparison with that of the bulk material (Figure 5). It is also
evident that hydrocarbon formation is suppressed on the
nanoscale material, whereas the CO selectivity is enhanced. As-
sisted by a modeling study, the authors considered that such
a change originated from the stronger surface binding to Had
Figure 3. a) Potential-dependent Faradaic efficiency of Au nanoparticles in and CO2 intermediates (*CO2C or *COOH) because of the pres-
the electrocatalytic reduction of CO2 to CO. b) Density of adsorption sites ence of low-coordinate atoms, which could make the Heyrov-
(yellow, light orange, dark orange, and red symbols for (111), (0 0 1), edge,
or corner on-top sites, respectively) on closed-shell cuboctahedral Au clus-
sky mechanism for HH bond formation the more dominant
ters versus the cluster diameter. The weight fraction of Au bulk atoms is pathway for the HER. Nevertheless, more investigations may
marked by gray dots. Reproduced with permission from Ref. [23]. be required to better understand the size-dependent mecha-
nism. For instance, a more detailed analysis of Cu NP catalysts
with a particle size of 15 nm and bulk foil can be valuable to
bridge the gap between bulk and nanoscale materials.
to provide a near-optimum number of edge sites that are par- Similar observations were made for other Cu nanostructures.
ticularly active for CO2 reduction into CO and minimize the Recently, Sen et al. described the synthesis of a 3 D Cu nano-
number of corner sites active for the HER. foam (Figure 6 a) that comprised nanoscale dendritic walls and
Metallic Cu is the only known material able to convert CO2 was characterized as an electrocatalyst for CO2 reduction.[27]
electrochemically in an appreciable amount into a variety of The results presented in Figure 6 b show that the HER is more
hydrocarbons.[24] However, Cu is not an ideal catalytic material significantly promoted than that on a bulk Cu surface
for CO2 reduction because 1) it usually requires large overpo- (Figure 4) as in the case of Cu NPs. Although some enhance-
tentials of above 1 V, 2) it deactivates on short time scales, and ment in HCOOH formation was achieved, the production of
3) its selectivity is too poor for practical applications.[1a] Never- C1C3 species that were unique to the bulk Cu surface was
theless, an understanding of the unique ability of Cu to cata- suppressed nearly to trace amounts.
lyze hydrocarbon formation would aid the design of future cat- It is clear that the implementation of the Cu surface on
alysts that are active at a low overpotential and have good a nanoscale is not trivial. The influence of material dimensions
product control. Recently, Kuhl et al. re-evaluated the polycrys- on the CO2 reduction mechanism is not understood. An experi-
talline Cu surface as a CO2 catalyst using modern characteriza- mental modeling study with well-defined material dimensions
tion techniques and a highly sensitive electrochemical cell to coupled with in situ spectroscopy characterization can be an
establish a baseline for better mechanistic understanding.[25] effective approach to provide deeper insights.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &4&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

Oxide-Derived Metallic
Catalysts
One method employed recently
to improve the performance of
metallic catalysts with great suc-
cess was the oxidation and sub-
sequent electrochemical reduc-
tion of bulk metals such as Cu,
Au, and Sn. This process is able
to modify the catalytic surface
on the nanoscale and creates
more active surface reaction
sites. The intentional oxidation
and reduction of a polycrystalline
Cu surface was first proposed by
Frese[28] who reported increased
methanol production on oxi-
dized Cu electrodes compared
to that on bulk Cu. However, sta-
bility issues were not resolved
and complete product analysis
was not undertaken. More re-
cently Le et al. confirmed the re-
sults of Frese by showing the in-
creased methanol activity of
electrodeposited copper oxide
electrodes compared to air-oxi-
dized electrodes and anodized
electrodes.[29] The presence of
a transient CuI state, which de-
grades quickly during electro-
chemical reduction, contributed
to the increased methanol pro-
duction. Most recently, Li et al.
showed the enhanced stability
and CO2 reduction activity of
Figure 4. Faradaic efficiency for each product as a function of potential for major, intermediate, and minor prod- polycrystalline Cu foils oxidized
ucts. Reproduced with permission from Ref. [25].
intentionally at much higher oxi-
dation temperatures than those
used in previous studies. After
a thermal treatment at 500 8C for 12 h in air, the Cu foil exhibit-
ed a nanowire structure that exhibited very coarse surfaces
(Figure 7 a).[8a] With an in situ reduction during electrolysis, the
resulting Cu catalyst showed a reduced surface from CuO (ac-
cording to X-ray photoelectron spectroscopy; XPS) and pro-
duced formate and CO primarily at much lower potentials than
the unmodified counterparts. The catalytic activity and selectiv-
ity were also significantly improved (Figure 7 b). Such an en-
hancement is intriguing because 1) it has been shown already
that a decrease of the dimensions of Cu materials to the nano-
scale decreases the overall performance and 2) unlike the
curved Ag surfaces, the Tafel slope of the oxide-derived Cu
was similar to that of bulk Cu (  120 mV dec1), which indi-
Figure 5. Faradic selectivity of reaction products during CO2 electroreduction
cates that the formation of CO2C intermediates is still the rate-
on Cu NPs compared to that of a bulk foil. Lines are for guidance only. Con-
ditions: 0.1 m KHCO3, E = 1.1 V vs. RHE, 25 8C. Reproduced with permission determining step. Hypothetically, it is possible that the pres-
from Ref. [26]. ence of CuI on the surface, which is difficult to distinguish

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &5&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

Figure 7. a) SEM image of a polycrystalline Cu foil annealed at 500 8C for


Figure 6. a) SEM image of a Cu foam deposited electrochemically. b) The
12 h. b) The corresponding Faradaic efficiencies for the production of CO
corresponding product distribution as a function of applied potential. Repro-
and HCO2H as a function of applied potentials. Reproduced with permission
duced with permission from Ref. [27].
from Ref. [8a].

from Cu0 by XPS, or surface defect sites may be responsible for


the increase in activity, which likely improves the formation of mate at high overpotentials. Recently, Zhang et al. showed the
the CO2C intermediates and suppresses the HER activity. It is enhanced CO2 reduction activity of SnO2 nanoparticles reduced
difficult, however, to draw conclusions about any improvement electrochemically to formate selectively (Figure 9 a).[30] The au-
in the activity per site without further characterization of such thors concluded that the SnO2 particle size played a major role
a catalytic surface because the increase in the specific surface in the surface-species binding energy, and a maximum CO2 re-
area of the catalyst is much larger than the improvement in its duction activity was achieved using 5 nm particles. In addition,
activity. Chen and Kanan examined the performance of electrodeposit-
This technique has also been applied to polycrystalline Au ed Sn/SnOx nanoclusters as CO2 reduction catalysts.[21a] The role
foils with great success. Recently, Chen et al. reported a highly of SnOx was to help stabilize the CO2 intermediate. Compared
active electrocatalyst prepared from the intentional oxidation to bulk Sn, the higher density Sn/SnOx sites were able to en-
and reduction of Au foils by electrochemical processes.[21b] hance the catalyst performance with higher selectivity and cur-
After an anodic oxidation with an AC bias and a subsequent rent density (Figure 9 b) and to suppress the HER. However, it
electrochemical reduction at constant potential, a thick layer remains unclear upon which site the CO2 reduction actually
(  1 mm) of Au nanoparticles was formed on the surface of the occurs, namely, the interface between Sn and SnOx or simply
foil (Figure 8 a). The resulting surface-modified Au catalyst the SnOx surface.
showed enhanced CO production at much lower overpoten-
tials than bulk Au (Figure 8 b). The enhanced performance was
Nanostructured Electrocatalysis in Ionic
a result of the improved stabilization of the CO2 intermediate
Liquids
in addition to an increase in surface area, which is similar to
the case of nanoporous Ag. The mechanism of such an im- Another route to reduce the energy barrier effectively in elec-
provement could be understood from the modeling study un- trochemical CO2 reduction is to introduce certain ionic liquid
dertaken by Zhu et al.,[23] which was discussed in the previous (IL) molecules to the electrolyte. Inspired by the fact that 1-
section. butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide
Another CO2 reduction material of interest is Sn, the bulk (BMIM-NTf2) is able to reduce the overpotential in the O2 re-
form of which has the ability to reduce CO2 primarily to for- duction reaction by forming a complex with the O2C inter-

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &6&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

Figure 8. a) SEM image of the surface of oxide-derived Au nanoparticle film.


b) CO production partial current density versus overpotential on oxide-de- Figure 9. a) Dependence of total current densities (red squares) and Faradaic
rived Au and polycrystalline Au. Reproduced with permission from efficiencies (blue circles) on the applied electrolysis potential for formate
Ref. [21b]. production on reduced nano-SnO2 (  5 nm) loaded on carbon black
(  30 nm). b) Comparison of CO2 reduction catalysis for Sn foil and in situ
deposited Sn/SnOx thin-film electrodes for the Faradiac efficiency of CO and
HCO2H production at various applied potentials. Reproduced with permis-
mediate, Rosen et al. proposed that CO2 reduction might also sion from Refs. [31]and [21a].
be facilitated if certain IL molecules could form a complex with
the CO2C intermediate to reduce the energy barrier (Fig-
ure 10 a).[31] The authors examined Ag nanoparticle (> 100 nm) proximately 2 mA cm2 to above 10 mA cm2 (Figure 11 b). A
catalysts in an 18 mol % 1-ethyl-3-methylimidazolium tetra- further increase of the H2O content promoted the HER rapidly
fluoroborate (EMIM-BF4) solution by using a sandwich-type and caused the decrease of both CO efficiency and current
flow-cell reactor. The CO2 reduction started at an overpotential density. This result shows that the addition of an appropriate
as low as 0.17 V (1.5 V in applied cell voltage) with a Faradaic amount of water to EMIM-BF4 can accelerate the desired CO2
efficiency near 100 %. If the cell voltage increased from 1.5 to reaction, which is likely because the hydrolysis of BF4 provides
2.5 V (0.17 to 1.17 V in overpotential), the turnover number in- greater proton availability as evident in the change of electro-
creased from 0.8 to 1.35 s1, which corresponds to current den- lyte pH (Figure 11 a). It is also evident that the EMIM cation
sities of approximately 0.61–1.08 mA cm2, and the CO produc- could effectively inhibit the H2 evolution expected upon the
tion was maintained at a very high efficiency of > 96 % addition of water, even at high water concentrations.
(Figure 10 b). Salehi-Khojin et al. investigated the effect of the Ag particle
To follow up, Rosen et al. studied the influence of the H2O size in CO2 reduction using an IL electrolyte.[33] The authors
content in the EMIM-BF4 solution electrolyte on the catalytic prepared a series of Ag nanoparticle catalysts with different
performance.[32] The authors employed the same system as in particle sizes (1–200 nm) but identical mass loading. The study
their previous work except for using smaller Ag nanoparticles was conducted in a three-electrode electrochemical cell
(40 nm). At a cell voltage of 2.5 V, in nearly dry IL electrolyte, equipped with a Ag wire as the reference electrode. The elec-
the Faradaic efficiency of CO and H2 were both very low (Fig- trolyte employed was EMIM-BF4 with a low water content of
ure 11 a), which the authors believed was because of current approximately 75 mm. The CO2 reduction current density in-
losses such as leakage. However, as the water content in- creased significantly as the particle size decreased from 200 to
creased up to approximately 90 mol %, the CO efficiency in- 5 nm (Figure 12). This is expected because a smaller particle
creased to nearly 100 % without any promotion of the HER size tends to expose a larger surface area and denser stepped
(Figure 11 a), and the total current density increased from ap- surface sites. However, as the particle size decreased further to

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &7&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

Figure 11. Effect of the addition of H2O to EMIM-BF4 on a) the electrolyte pH


and the faradaic efficiency of CO2 reduction to CO and H2 and b) the flow-
cell current density at a cell voltage of 2.5 V. Reproduced with permission
Figure 10. a) A schematic of how the free energy of the system changes from Ref. [33].
during the reaction CO2+2 H++2 eÐCO+H2O in water or acetonitrile (solid
line) or EMIM-BF4 (dashed line). b) Plot of the Faradaic efficiency of the pro-
cess to form the desired CO and the undesired H2 and the turnover rate as
a function of applied cell potential. Reproduced with permission from
Ref. [32].

1 nm, the current density decreased by more than one order


of magnitude. The authors considered that this is likely to be
caused by a change of binding energy to certain intermediates
during the reduction reaction as they witnessed a clear change
of binding energy to hydroxyl and sulfate groups in a con-
trolled study. However, according to Zhu et al., it could also be
because the ratio of edge sites (active for CO2 reduction) to
corner sites (active for HER) is no longer optimal at such
a finite size.[23]
An IL electrolyte may also change the selectivity in the CO2
reduction reaction. Bi was reported to be a formate-selective
catalyst for CO2 reduction in an aqueous environment.[34] Re-
Figure 12. Current density for CO (0.75 V vs. SHE, electrolyte purged with
cently, DiMeglio and Rosenthal synthesized a rosebud-type Bi CO2) and H2 formation (1.14 V vs. SHE, electrolyte purged with Ar) as
thin film on a glassy carbon electrode (GCE) using electro- a function of Ag particle size. Reproduced with permission from Ref. [34].
chemical deposition (Figure 13 a) and characterized its CO2 re-
duction activity in an IL-containing electrolyte.[35] Interestingly, 4 mA cm2 was achieved with a CO efficiency of approximately
Bi became a CO-selective catalyst in the IL-containing electro- 93 % at 1.95 V versus SCE in the IL-containing electrolyte. The
lyte. The enhancement of activity after the introduction of the Tafel analysis shows a slope of 139 mV dec1, which indicates
IL is clear (Figure 13 b). A current density of approximately that the initial electron transfer to form CO2C is still the rate-

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &8&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

Summary and Outlook


The latest progress on nanostructured metallic electrocatalysts
for CO2 reduction was reviewed. As the scope of this article is
limited to nanostructured metallic electrocatalysts, other im-
portant systems, such as homogeneous and nonmetallic elec-
trocatalysts, are not discussed.
To reveal the important role of nanostructuring, important
recent discoveries of nanoarchitectured Ag, Au, and Cu cata-
lysts and oxide-derived metallic catalysts were reviewed. Un-
surprisingly, all of these nanostructured metallic catalysts ex-
hibited enormous promise to achieve high-performance CO2
reduction. On the basis of the results presented here, we be-
lieve that under-coordinated metal atoms on a highly nano-
structured surface play an important role to facilitate the elec-
trocatalytic conversion of CO2 to reduced species. In other
words, atoms that sit on the edge or on a curved surface
behave differently from those on a flat surface. This can be the
main reason that nanostructured catalysts exhibit distinct be-
haviors compared to bulk catalysts as shown in the cases of
nanoporous and nanoparticle metallic catalysts, although it
may be debatable for universal implementation. The impor-
tance of surface nanostructuring was further proved by com-
putational modeling, in which the surface sites on a nanosized
particle had much lower activation barriers for the initial reduc-
tion of CO2 molecules.
In addition to aqueous electrolytes, which are the most
common for electrochemical CO2 reduction, studies of nano-
structured Ag and Bi catalysts in electrolytes based on ionic
liquids (ILs) were also discussed. The combination of a nano-
Figure 13. a) SEM image of a Bi thin film deposited on a GCE. b) CV traces structured electrocatalyst and an IL shows many advantages
recorded for Bi-deposited and bare GCEs in MeCN that contained 20 mm
over aqueous electrolyte systems, such as a decreased energy
EMIM-BF4. Inset: Bi-deposited GCE in MeCN without the IL. Reproduced with
permission from Ref. [36]. barrier for CO2 reduction, enhanced CO2 solubility, and sup-
pressed hydrogen evolution reaction. However, potential chal-
lenges of IL-based electrolytes have also been identified and
determining step. To follow up, Medina-Ramos et al. improved discussed.
the cost-effectiveness of the system 1) by eliminating the need For an ideal CO2-reduction electrocatalyst, the following criti-
for the expensive tertabutylammonuim tetraflourophosphate cal properties have to be satisfied: 1) highly active under low
(TBAPF4) by using BMIM-OTf as the only additive to MeCN and overpotentials to minimize the cost of electricity, 2) highly se-
2) by depositing Bi in situ from an organometallic Bi3+ complex lective to the production of the desired chemicals with no re-
(Bi(OTf)3) at a less negative potential.[36] In these two studies, quirement of postreaction separation, 3) highly stable for long-
the authors calculated the standard equilibrium potential of term commercial applications, and 4) based on abundant ele-
CO2/CO in IL/MeCN and used the value to estimate the overpo- ments with a minimum capital cost. The nanostructured cata-
tentials. However, the actual equilibrium potential may shift if lysts discussed in this article are undoubtedly approaching the
the test environment deviates from the standard conditions, targets. Therefore, it is urgent to explore nanostructured cata-
which has been a concern in aqueous electrolytes if a RHE lysts for further improvement because they offer opportunities
scale is used. to access unique catalytic properties that cannot be found in
ILs can definitely facilitate the CO2 reduction reaction not bulk catalysts.
only by forming a possible complex of their cations and CO2C The performance of nanostructured metallic catalysts may
to reduce the reaction energy barrier but also by enhancing be enhanced by the introduction of bimetallic surfaces and bi-
the reaction rate because of the greater solubility of CO2.[37] metallic core–shell structures. Bimetallic electrocatalysts offer
However, the use of an IL-based electrolyte is not a simple unprecedented opportunities for process optimization.[38] The
panacea. The complicated chemical and electrochemical properties of a bimetallic catalyst can be tuned between the
nature of an IL electrolyte system requires a more sophisticat- two types of metal species located on the same surface or on
ed analysis. For instance, carbon-trace CO2 may be necessary different core–shell layers. As a result, bimetallics are able to
to distinguish the products from the possible decomposition access catalytic properties that cannot be mimicked by mono-
of the IL. metallics. Importantly, it may be possible to design a bimetallic

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &9&
These are not the final page numbers! ÞÞ
CHEMCATCHEM
MINIREVIEWS www.chemcatchem.org

catalyst that can effectively assist the further reduction and [12] C. G. Morales-Guio, L.-A. Stern, X. Hu, Chem. Soc. Rev. 2014, 43, 6555 –
protonation of adsorbed *CO so that liquid-fuel production 6569.
[13] H. Dau, C. Limberg, T. Reier, M. Risch, S. Roggan, P. Strasser, ChemCatCh-
from CO2 can be feasible. However, bimetallics, especially their em 2010, 2, 724 – 761.
nanostructured materials, have been understudied as CO2 re- [14] a) B. Kumar, M. Asadi, D. Pisasale, S. Sinha-Ray, B. A. Rosen, R. Haasch, J.
duction electrocatalysts to date. Abiade, A. L. Yarin, A. Salehi-Khojin, Nat. Commun. 2013, 4, 2819 – 2819;
This review article has been written in the hope to stimulate b) M. Asadi, B. Kumar, A. Behranginia, B. A. Rosen, A. Baskin, N. Repnin,
D. Pisasale, P. Phillips, W. Zhu, R. Haasch, R. F. Klie, P. Kral, J. Abiade, A.
future research interest in the development of efficient electro- Salehi-Khojin, Nat. Commun. 2014, 5, 4470 – 4470.
catalysts with new nanostructures and compositions for CO2 [15] a) D. S. Laitar, P. Muller, J. P. Sadighi, J. Am. Chem. Soc. 2005, 127,
reduction. 17196 – 17197; b) D. C. Grills, Y. Matsubara, Y. Kuwahara, S. R. Golisz,
D. A. Kurtz, B. A. Mello, J. Phys. Chem. Lett. 2014, 5, 2033 – 2038; c) M. D.
Sampson, A. D. Nguyen, K. A. Grice, C. E. Moore, A. L. Rheingold, C. P.
Acknowledgements Kubiak, J. Am. Chem. Soc. 2014, 136, 5460 – 5471.
[16] H. A. Hansen, J. B. Varley, A. A. Peterson, J. K. Norskov, J. Phys. Chem.
Lett. 2013, 4, 388 – 392.
The authors are grateful for financial support from the National [17] W. C. Sheng, M. Myint, J. G. G. Chen, Y. S. Yan, Energy Environ. Sci. 2013,
Science Foundation Faculty Early Career Development (CAREER) 6, 1509 – 1512.
program (Award No. CBET-1350911). [18] Y. Hori, H. Wakebe, T. Tsukamoto, O. Koga, Electrochim. Acta 1994, 39,
1833 – 1839.
[19] T. Hatsukade, K. P. Kuhl, E. R. Cave, D. N. Abram, T. F. Jaramillo, Phys.
Keywords: electrochemistry · energy conversion · ionic Chem. Chem. Phys. 2014, 16, 13814 – 13819.
[20] E. Gileadi, Electrode Kinetics for Chemists, Engineers, and Materials Scien-
liquids · nanostructures · reduction
tists, VCH, New York, 1993.
[21] a) Y. Chen, M. W. Kanan, J. Am. Chem. Soc. 2012, 134, 1986 – 1989; b) Y.
[1] a) M. Gattrell, N. Gupta, A. Co, J. Electroanal. Chem. 2006, 594, 1 – 19;
Chen, C. W. Li, M. W. Kanan, J. Am. Chem. Soc. 2012, 134, 19969 – 19972.
b) N. S. Lewis, D. G. Nocera, Proc. Natl. Acad. Sci. USA 2006, 103, 15729 –
[22] N. Hoshi, M. Kato, Y. Hori, J. Electroanal. Chem. 1997, 440, 283 – 286.
15735; c) J. Kim, T. A. Johnson, J. E. Miller, E. B. Stechel, C. T. Maravelias,
[23] W. Zhu, R. Michalsky, O. Metin, H. Lv, S. Guo, C. J. Wright, X. Sun, A. A.
Energy Environ. Sci. 2012, 5, 8417 – 8429.
Peterson, S. Sun, J. Am. Chem. Soc. 2013, 135, 16833 – 16836.
[2] a) R. Angamuthu, P. Byers, M. Lutz, A. L. Spek, E. Bouwman, Science
[24] A. A. Peterson, F. Abild-Pedersen, F. Studt, J. Rossmeisl, J. K. Norskov,
2010, 327, 313 – 315; b) Z. F. Chen, J. J. Concepcion, M. K. Brennaman, P.
Energy Environ. Sci. 2010, 3, 1311 – 1315.
Kang, M. R. Norris, P. G. Hoertz, T. J. Meyer, Proc. Natl. Acad. Sci. USA
[25] K. P. Kuhl, E. R. Cave, D. N. Abram, T. F. Jaramillo, Energy Environ. Sci.
2012, 109, 15606 – 15611; c) L. Hammarstrçm, S. Hammes-Schiffer, Acc.
2012, 5, 7050 – 7059.
Chem. Res. 2009, 42, 1859 – 1860.
[26] R. Reske, H. Mistry, F. Behafarid, B. R. Cuenya, P. Strasser, J. Am. Chem.
[3] a) W. C. Chueh, C. Falter, M. Abbott, D. Scipio, P. Furler, S. M. Haile, A.
Soc. 2014, 136, 6978 – 6986.
Steinfeld, Science 2010, 330, 1797 – 1801; b) B. Xu, Y. Bhawe, M. E. Davis,
[27] S. Sen, D. Liu, G. T. R. Palmore, ACS Catal. 2014, 4, 3091 – 3095.
Chem. Mater. 2013, 25, 1564 – 1571; c) D. Arifin, V. J. Aston, X. Liang,
[28] K. W. Frese, J. Electrochem. Soc. 1991, 138, 3338 – 3344.
A. H. McDaniel, A. W. Weimer, Energy Environ. Sci. 2012, 5, 9438 – 9443.
[29] M. Le, M. Ren, Z. Zhang, P. T. Sprunger, R. L. Kurtz, J. C. Flake, J. Electro-
[4] a) A. J. Morris, G. J. Meyer, E. Fujita, Acc. Chem. Res. 2009, 42, 1983 –
chem. Soc. 2011, 158, E45 – E49.
1994; b) E. E. Barton, D. M. Rampulla, A. B. Bocarsly, J. Am. Chem. Soc.
[30] S. Zhang, P. Kang, T. J. Meyer, J. Am. Chem. Soc. 2014, 136, 1734 – 1737.
2008, 130, 6342 – 6344; c) C. Wang, Z. Xie, K. E. deKrafft, W. Lin, J. Am.
[31] B. A. Rosen, A. Salehi-Khojin, M. R. Thorson, W. Zhu, D. T. Whipple,
Chem. Soc. 2011, 133, 13445 – 13454.
P. J. A. Kenis, R. I. Masel, Science 2011, 334, 643 – 644.
[5] a) E. E. Benson, C. P. Kubiak, A. J. Sathrum, J. M. Smieja, Chem. Soc. Rev.
[32] B. A. Rosen, W. Zhu, G. Kaul, A. Salehi-Khojin, R. I. Masel, J. Electrochem.
2009, 38, 89 – 99; b) M. Rakowski Dubois, D. L. Dubois, Acc. Chem. Res.
Soc. 2013, 160, H138 – H141.
2009, 42, 1974 – 1982; c) C. Costentin, M. Robert, J.-M. Saveant, Chem.
[33] A. Salehi-Khojin, H.-R. M. Jhong, B. A. Rosen, W. Zhu, S. Ma, P. J. A. Kenis,
Soc. Rev. 2013, 42, 2423 – 2436; d) M. Jitaru, D. A. Lowy, M. Toma, B. C.
R. I. Masel, J. Phys. Chem. C 2013, 117, 1627 – 1632.
Toma, L. Oniciu, J. Appl. Electrochem. 1997, 27, 875 – 889.
[34] S. Komatsu, T. Yanagihara, Y. Hiraga, M. Tanaka, A. Kunugi, Denki Kagaku
[6] H. A. Schwarz, R. W. Dodson, J. Phys. Chem. 1989, 93, 409 – 414.
1995, 63, 217 – 224.
[7] Y. Hori, Modern Aspects of Electrochemistry, Vol. 42, Springer, New York,
[35] J. L. DiMeglio, J. Rosenthal, J. Am. Chem. Soc. 2013, 135, 8798 – 8801.
2008.
[36] J. Medina-Ramos, J. L. DiMeglio, J. Rosenthal, J. Am. Chem. Soc. 2014,
[8] a) C. W. Li, M. W. Kanan, J. Am. Chem. Soc. 2012, 134, 7231 – 7234; b) Q.
136, 8361 – 8367.
Lu, J. Rosen, Y. Zhou, G. S. Hutchings, Y. C. Kimmel, J. G. Chen, F. Jiao,
[37] M. B. Shiflett, B. A. Elliott, S. R. Lustig, S. Sabesan, M. S. Kelkar, A. Yokoze-
Nat. Commun. 2014, 5, 3242 – 3242.
ki, ChemPhysChem 2012, 13, 1806 – 1817.
[9] Y. Hori, H. Konishi, T. Futamura, A. Murata, O. Koga, H. Sakurai, K.
[38] a) N. M. Marković, T. J. Schmidt, V. Stamenkovic, P. N. Ross, Fuel Cells
Oguma, Electrochim. Acta 2005, 50, 5354 – 5369.
2001, 1, 105 – 116; b) H. Ataee-Esfahani, L. Wang, Y. Nemoto, Y. Yamau-
[10] a) J. Kibsgaard, Z. Chen, B. N. Reinecke, T. F. Jaramillo, Nat. Mater. 2012,
chi, Chem. Mater. 2010, 22, 6310 – 6318; c) C. Wang, M. Chi, D. Li, D.
11, 963 – 969; b) C. Chen, Y. Kang, Z. Huo, Z. Zhu, W. Huang, H. L. Xin,
Strmcnik, D. van der Vliett, G. Wang, V. Komanicky, K.-C. Chang, A. P.
J. D. Snyder, D. Li, J. A. Herron, M. Mavrikakis, M. Chi, K. L. More, Y. Li,
Paulikas, D. Tripkovic, J. Pearson, K. L. More, N. M. Markovic, V. R. Sta-
N. M. Markovic, G. A. Somorjai, P. Yang, V. R. Stamenkovic, Science 2014,
menkovic, J. Am. Chem. Soc. 2011, 133, 14396 – 14403.
343, 1339 – 1343.
[11] a) S. Sharma, B. G. Pollet, J. Power Sources 2012, 208, 96 – 119; b) M. K.
Jeon, C. H. Lee, G. I. Park, K. H. Kang, J. Power Sources 2012, 216, 400 – Received: August 21, 2014
408; c) C.-H. Cui, S.-H. Yu, Acc. Chem. Res. 2013, 46, 1427 – 1437. Published online on && &&, 0000

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &10&
These are not the final page numbers! ÞÞ
MINIREVIEWS
Small is beautiful: Electrochemical CO2 Q. Lu, J. Rosen, F. Jiao*
reduction is an attractive approach to
&& – &&
convert CO2 produced in power plants,
refineries, and petrochemical plants to Nanostructured Metallic
liquid fuels or useful chemicals. Recent Electrocatalysts for Carbon Dioxide
progress in nanostructured metallic cat- Reduction
alysts has exhibited tremendous prom-
ise for such realization. This review
takes a closer look at those studies, and
future research directions are proposed
and discussed.

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemCatChem 0000, 00, 1 – 11 &11&
These are not the final page numbers! ÞÞ

You might also like