A Review of Catalysts For The Electroreduction of CO2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

Chem Soc Rev

View Article Online


REVIEW ARTICLE View Journal | View Issue

A review of catalysts for the electroreduction of


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

carbon dioxide to produce low-carbon fuels


Cite this: Chem. Soc. Rev., 2014,
43, 631
Jinli Qiao,b Yuyu Liu,*c Feng Hong*a and Jiujun Zhang*de

This paper reviews recent progress made in identifying electrocatalysts for carbon dioxide (CO2) reduction to
produce low-carbon fuels, including CO, HCOOH/HCOO, CH2O, CH4, H2C2O4/HC2O4, C2H4, CH3OH,
CH3CH2OH and others. The electrocatalysts are classified into several categories, including metals, metal alloys,
metal oxides, metal complexes, polymers/clusters, enzymes and organic molecules. The catalyts’ activity,
product selectivity, Faradaic efficiency, catalytic stability and reduction mechanisms during CO2 electroreduction
have received detailed treatment. In particular, we review the effects of electrode potential, solution–electrolyte
Received 9th September 2013 type and composition, temperature, pressure, and other conditions on these catalyst properties. The challenges
DOI: 10.1039/c3cs60323g in achieving highly active and stable CO2 reduction electrocatalysts are analyzed, and several research directions
for practical applications are proposed, with the aim of mitigating performance degradation, overcoming addi-
www.rsc.org/csr tional challenges, and facilitating research and development in this area.

1. Introduction high energy demands, CO2 conversion and utilization seems to be a


more attractive and promising solution. Normally, CO2 conversion
Carbon dioxide (CO2) is the most notorious greenhouse gas, can be achieved by chemical methods,9–16 by photocatalytic and
released by both natural and artificial processes. It is also a electrocatalytic reduction,17–25 and by a few other means.26–28
necessary material for the growth of all earth’s plants and for many However, at the present time, certain barriers still hinder the
industrial processes.1–4 In an ideal scenario, the CO2 produced on practical application of CO2 capture, conversion, and utilization.
Earth should be balanced with what is consumed, so that the These barriers include (1) the high costs of CO2 capture, separation,
level of CO2 remains constant to maintain environmental stability. purification, and transportation to user sites; (2) the high energy
Unfortunately, with the intensification of human industrial requirements for CO2 chemical/electrochemical conversion;
activities, this balance has gradually been disrupted, leading (3) limitations in market size and investment incentives; (4) lack
to more CO2 production and making global warming a pressing of industrial commitment to enhance CO2-based chemicals; and
issue. Therefore, reducing CO2 production and converting CO2 (5) insufficient socio-economic driving forces.4,29 Despite such
into useful materials seems to be necessary, indeed critical, for challenges, CO2 capture, conversion, and utilization is still
environmental protection, and various governments worldwide recognized as a feasible and promising cutting-edge area of
have signaled their concern by increasing their investment in exploration in energy and environmental research.
research to address the CO2 issue. In recent years, CO2 conversion using electrochemical cata-
The different proposed technologies follow one of two major lysis approaches has attracted great attention for its several
approaches: to capture and geologically sequestrate CO2, or to advantages: (1) the process is controllable by electrode poten-
convert CO2 into useful low-carbon fuels.5–8 In today’s world of tials and reaction temperature; (2) the supporting electrolytes
can be fully recycled so that the overall chemical consumption
a
College of Chemistry, Chemical Engineering & Biotechnology, Donghua University, can be minimized to simply water or wastewater; (3) the
2999 Ren’min North Road, Shanghai 201620, P. R. China. electricity used to drive the process can be obtained without
E-mail: f [email protected]; Tel: +86-21-67792649
b
generating any new CO2—sources include solar, wind, hydro-
College of Environmental Science and Engineering, Donghua University,
2999 Ren’min North Road, Shanghai 201620, P. R. China
electric, geothermal, tidal, and thermoelectric processes; and
c
Multidisciplinary Research on the Circulation of Waste Resources, (4) the electrochemical reaction systems are compact, modular,
Graduate School of Environmental Studies, Tohoku University, Aramaki, on-demand, and easy for scale-up applications.8 However,
aza Aoba 6-6-11, Aoba-ku, Sendai 980-8579, Japan. challenges remain, such as the slow kinetics of CO2 electro-
E-mail: [email protected]; Tel: +81-90-6008-9342
d
reduction, even when electrocatalysts and high electrode reduction
Research Institute of Donghua University, Shanghai 201620, P. R. China
e
NRC Energy, Mining & Environment, National Research Council of Canada,
potential are applied; the low energy efficiency of the process, due
4250 Wesbrook Mall, Vancouver, B.C. V6T 1W5, Canada. to the parasitic or decomposition reaction of the solvent at high
E-mail: [email protected]; Fax: +1 604 221 3001; Tel: +1 604 221 3087 reduction potential; and high energy consumption. Researchers

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 631
View Article Online

Review Article Chem Soc Rev

have recognized that the biggest challenge in CO2 electroreduction (CH2CH2), ethanol (CH3CH2OH), as well as others. The thermo-
is low performance of the electrocatalysts (i.e., low catalytic activity dynamic electrochemical half-reactions of CO2 reduction and
and insufficient stability). their associated standard electrode potentials are listed in
Electrochemical reduction of CO2 can proceed through two-, Table 1.30 Note that the reactions listed in Table 1 are thermo-
four-, six-, and eight-electron reduction pathways in gaseous, dynamic, only indicating each reaction’s tendency and possi-
aqueous, and non-aqueous phases at both low and high tem- bility but giving no indication of the reaction’s kinetics, such as
peratures. The major reduction products are carbon monoxide rate and mechanism. In addition, the standard potentials listed in
(CO), formic acid (HCOOH) or formate (HCOO) in basic solution, Table 1 are for aqueous solutions only; the potential values in non-
oxalic acid (H2C2O4) or oxalate (C2O42 in basic solution), form- aqueous solutions are different from those listed in Table 1.31 The
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

aldehyde (CH2O), methanol (CH3OH), methane (CH4), ethylene kinetics of CO2 electroreduction involve very complicated reaction

Dr Jinli Qiao is a Professor, PhD Dr Yuyu Liu is an Associate


Supervisor and Scientific Core- Professor at the Graduate School
Competency Leader at Donghua of Environmental Studies,
University, China. She received Tohoku University, Japan. Dr Liu
her PhD in Electrochemistry received his PhD in Environmental
from Yamaguchi University, Engineering from Yamaguchi
Japan, in 2004 before joining the University, Japan, in 2003. He
National Institute of Advanced then worked at Kyushu Environ-
Industrial Science and Technol- mental Evaluation Association,
ogy (AIST), Japan, as a research Osaka Institute of Technology,
scientist. From 2004–2008, she Tokyo University of Agriculture
carried out 7 fuel cell projects and Technology, and Yokohama
Jinli Qiao including two NEDO projects of Yuyu Liu National University as Post-
Japan on the development of doctoral and Research Fellow.
novel proton-conducting membranes, new binders for MEA fabri- Dr Liu has more than 10 years of experience in the environment
cation, and non-platinum catalysts. Since March 2008, she has science and technology, particularly in the areas of air quality
carried out a total of 8 projects funded by Chinese Government monitoring, water and soil research, and their associated
including the National Natural Science Foundation of China. instrument development. He is also a member of the Japan
Dr Qiao has more than 20 years of scientific research experience, Society on Water Environment and Japan Society of Material
particularly in the area of electrochemical material development Cycles and Waste Management.
and energy storage and conversion.

Dr Feng Hong is a Professor, Dr Jiujun Zhang is a Principal


PhD Supervisor and Scientific Research Officer and Technical
competency Leader at the Core Competency Leader at the
College of Chemistry, Chemical National Research Council of
Engineering and Biotechnology, Canada Energy, Mining & Environ-
Donghua University, China. He ment Portfolio (NRC-EME).
received his PhD in engineering Dr Zhang received his BS and
from Nanjing Forestry University MSc in electrochemistry from
in 1998. He then went to Sweden Peking University in 1982 and
for his postdoctoral research at 1985, respectively, and his PhD
the Lund Institute of Technology/ in electrochemistry from Wuhan
Lund University, and then joined University in 1988. He then
Feng Hong Karlstad University as a senior Jiujun Zhang carried out three terms of
research scientist. His research postdoctoral research at the
interests include biomaterials, electrochemical energy materials California Institute of Technology, York University, and the
and fuel cells. Prof. Hong is a research proposal referee for University of British Columbia. Dr Zhang has over 30 years of
several programs hosted by the National Natural Science scientific research experience, particularly in the area of
Foundation of China, and International Cooperation Funding of electrochemical energy storage and conversion. He is also the
the Ministry of Science and Technology of China. Adjunct Professor at the University of British Columbia, the
University of Waterloo, Peking University, and Donghua University.

632 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

Table 1 Selected standard potentials of CO2 in aqueous solutions (V vs. compounds, such as metal complexes. This is probably because
SHE) at 1.0 atm and 25 1C, calculated according to the standard Gibbs energies these metals have vacant orbits and active d electrons, which are
of the reactants in reactions. Reprinted with permission from ref. 30. Copyright
believed to be able to energetically facilitate the bonding between
r 1985 CRC Press
the metal and the CO2 for adduct formation and also facilitate the
Electrode potentials desorption of the reduction product(s), as discussed below.
Half-electrochemical (V vs. SHE) under
thermodynamic reactions standard conditions
2.1 Transition metals and related electrocatalysts
CO2(g) + 4H+ + 4e = C(s) + 2H2O(l) 0.210
CO2(g) + 2H2O(l) + 4e = C(s) + 4OH 0.627 2.1.1 Titanium. Normally, titanium (Ti) on its own has no
significant catalytic activity towards CO2 electroreduction.38
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

CO2(g) + 2H+ + 2e = HCOOH(l) 0.250


CO2(g) + 2H2O(l) + 2e = HCOO(aq) + OH 1.078 However, TiO2 has shown some activity in both photocatalysis
CO2(g) + 2H+ + 2e = CO(g) + H2O(l) 0.106 and electrocatalysis for CO2 electroreduction.25,39 When TiO2 is
CO2(g) + 2H2O(l) + 2e = CO(g) + 2OH 0.934 used as an electrocatalyst, thin TiO2 films or mixtures of TiO2
CO2(g) + 4H+ + 4e = CH2O(l) + H2O(l) 0.070 and other metal oxides are usually deposited on a Ti electrode
CO2(g) + 3H2O(l) + 4e = CH2O(l) + 4OH 0.898
substrate to catalyze CO2 electroreduction. For example, in the
CO2(g) + 6H+ + 6e = CH3OH(l) + H2O(l) 0.016
preparation of a CO2 catalytic electrode, Monnier et al.40,41 pre-
CO2(g) + 5H2O(l) + 6e = CH3OH(l) + 6OH 0.812
pared TiO2, TiO2–Ru (or RuO2), and TiO2–Pt thin film electrodes
CO2(g) + 8H+ + 8e = CH4(g) + 2H2O (l) 0.169
CO2(g) + 6H2O(l) + 8e = CH4(g) + 8OH 0.659 by thermal deposition on titanium rods. Bandi42 deposited
2CO2(g) + 2H+ + 2e = H2C2O4(aq) 0.500 metallic oxide mixtures (including RuO2, TiO2, MoO2, Co3O4,
2CO2(g) + 2e = C2O42(aq) 0.590 and Rh2O3) on titanium foil. Cueto and Hirata43 prepared a
2CO2(g) + 12H+ + 12e = CH2CH2(g) + 4H2O(l) 0.064 TiO2–indium-tin oxide (ITO) thin-film glass electrode via a
2CO2(g) + 8H2O(l) + 12e = CH2CH2(g) + 12OH 0.764 fixed-potential bulk electrolysis process, in which 1-butyl-3-
2CO2(g) + 12H+ + 12e = CH3CH2OH(l) + 3H2O(l) 0.084 methylimidazolium tetrafluoroborate (BMImBF4, an ionic liquid)
2CO2(g) + 9H2O(l) + 12e = CH3CH2OH(l) + 12OH 0.744 was used as both solvent and supporting electrolyte. Several
material characterization techniques were employed to analyze
the coated electrode. X-ray diffraction (XRD) results showed
mechanisms, and the reaction rates are very slow, even in the
that TiO2 films were suitable candidates for electrocatalytic
presence of electrocatalysts. In general, the catalysts currently being
processes since no phase changes could be observed (Fig. 1).
employed are still not active enough. Furthermore, in some cases,
XPS measurements revealed a significant dissociation of CO2
the product of the electroreduction is not a single species but a
into chemisorbed CO32. Furthermore, observation using Auger
mixed product containing several component species (for example,
electron spectroscopy (AES) indicated strong interactions between
it could be a mixture of C, CO, HCOOH, H2C2O4, CH2O, CH3OH,
TiO2 and CO2 or CO32 and revealed a decrease in carbon content
CH4, CH2CH2, CH3CH2OH, and so on). The number of species and
after the CO2 electroreduction process (Fig. 2). Recently, electro-
the amount of each species that is present are factors strongly
catalytic synthesis of low-density polyethylene (PE) from CO2 on a
dependent on the kind and selectivity of the electrocatalyst
nanostructured TiO2 (ns-TiO2) film electrode was carried out by
employed and what electrode potential is applied. This suggests that
the currently employed electrocatalysts have insufficient catalytic
selectivity and stability. In most cases, these catalysts can survive
for fewer than 100 hours,32–34 which is far below the requirements
for practical use and technological commercialization. Therefore,
unsatisfactory catalysis, including low catalytic activity, selectivity,
and stability, is the biggest challenge in CO2 electroreduction. In the
past several decades, almost all the efforts in CO2 electroreduction
studies have been focused on research and development of electro-
catalysts to overcome the above challenges.35,36 Hence, while several
review articles related to CO2 reduction have been published,35–37 a
comprehensive review specifically focusing on electrocatalysts for
CO2 electroreduction is definitely necessary to facilitate research and
development in this area.

2. Elements and compounds used as


electrocatalysts for CO2
electroreduction
Fig. 1 XRD pattern of a TiO2–ITO thin film (a) before and (b) after a CO2
The most commonly explored electrocatalysts for CO2 electro- bulk electrolysis reduction reaction in BMImBF4. Reprinted with permis-
reduction are transition metal elements and their associated sion from ref. 43. Copyright r 2006 Springer.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 633
View Article Online

Review Article Chem Soc Rev


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 2 XPS (C1s, Ti2p, and O1s chemical states) and AES spectrum of a TiO2–ITO thin film: (a) before electrochemical reduction of CO2; (b) CO2
saturated in BMImBF4; and (c) after 19 hours of CO2 bulk electrolysis reduction reaction in BMImBF4. PQRE: Pt(s), Aux: Pt(s).43 Reprinted with permission
from ref. 43. Copyright r 2006 Springer.

controlled potential electrolysis in a mixture of H2O and EMImBF4 for the selective reduction of CO2. Due to the unique structures
at room temperature and ambient pressure.44 The ns-TiO2 film of these catalysts, some selectivity towards the desired products
appeared to be remarkably efficient and selective for the electro- has been achieved. For example, Qu et al.46 loaded RuO2
chemical reduction of CO2 when EMImBF4 was the solvent,45 onto TiO2 nanotubes (NTs) or nanoparticles (NPs) to form
as EMImBF4 maintained a high concentration of CO2 at the RuO2–TiO2(NTs) or RuO2–TiO2(NPs), which were then coated
electrode surface. According to the mechanism described in onto a Pt electrode for the electrocatalytic reduction of CO2. The
reactions (1)–(5), high pressure in the nanopores of the ns-TiO2 potentiostatic electrolysis of CO2 on the RuO2–TiO2(NT) coated
film can lead to polymerization of: CH2 to form PE.44 Pt electrode showed the selective formation of methanol with a
current efficiency of up to 65.5%, much better than was
CO2 + TiIII - CO2*(adsorbed on ns-TiO2 film) + TiIV (1)
achieved on the RuO2–TiO2(NP) coated Pt electrode.
Ti IV
+ e - Ti
 III
(2) 2.1.2 Molybdenum, chromium, and tungsten. Metallic
electrodes of Cr, Mo, and W do not appear to show significant
CO2*(ads) + H+ + e - CO(ads) + OH (3)
activity towards CO2 electroreduction. For example, Noda et al.38
CO(ads) + 4H + 4e -: CH2(ads) + H2O
+ 
(4) tested electrodes of these metals at 1.6 V vs. Ag/AgCl in KCl-
saturated 0.1 M KHCO3 aqueous solution at 298 1C, and did not
:CH2(ads) - [CH2CH2]n (5)
observe significant catalytic activity for the electrochemical
Indeed, TiO2 and carbon nanotubes have been explored as reduction of CO2. However, in an earlier study,47 researchers
catalyst supports in the synthesis of nanostructured electrocatalysts using molybdenum metal electrodes in the electrolysis of

634 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

CO2 electrocatalysts.56 Sende et al.57 electropolymerized several


4-v-tpy and 6-v-tpy complexes of transition metals (including Cr)
onto glassy carbon electrodes (GCEs) for CO2 electroreduction.
They found that in CO2-saturated 0.10 M aqueous NaClO4
solution, their electropolymerized [Cr(4-v-tpy)2]2+ film was at
0.86 V vs. Ag/AgCl, which was more positive than for Fe, Ni, Ru,
and Os complexes in the potential range of 1.10 to 1.22 V vs.
Ag/AgCl; it also yielded a considerable amount of formaldehyde
(CH2O), up to 87% at 1.10 V (far higher than the 39% and 28%
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

achieved with Co and Fe complexes, respectively). The turnover


number (TON) for catalyzed CO2 electroreduction was up to
6100 for the [Cr(4-v-tpy)2]2+ complex, about 50% lower than for
both the [Co(4-v-tpy)2]2+ (11 000) and [Fe(4-v-tpy)2]2+ (15 000)
complexes. In all cases, virtually the only reaction product
detected was CH2O, demonstrating the catalysts’ high selectivity.
It is worth mentioning that Sende et al. also explored one Cr salt
(diaquabis(oxalato)-chromate(III) (K[Cr(III)(C2O4)2(H2O)2]) as a
homogeneous catalyst for the electrochemical reduction of CO2
Fig. 3 Cyclic voltammetry for a Mo electrode in (A) N2 and (B) CO2-
as an Everitt’s salt, i.e., KFeII[FeII(CN)6]-modified platinum gauze
saturated 0.2 M Na2SO4 aqueous solution (pH 4.2). Electrodes (3.4 cm2) electrode.
were pretreated with HCl and allowed to remain at open circuit for 1 hour Regarding tungsten-related electrocatalysis for CO2 electro-
before scanning. Sweep rates were (A) 0.5 to 10 and (B) 0.5 to 6 V min1. reduction, Reda et al.58 obtained a tungsten-containing formate
CO was formed at open circuit; CO2 reduction (Mo + 2CO2 - MoO2 +
dehydrogenase enzyme (FDH1) from Syntrophobacter fumaroxidans
2CO) and molybdenum oxidation (Mo + 2H2O - MoO2 + 4e + 4H+, E =
0.57 V vs. SCE at pH 4.2) occurred simultaneously. Reprinted with
and adsorbed it on a freshly polished, pyrolytic graphite edge
permission from ref. 47. Copyright r 1986 Elsevier. electrode to catalyze CO2 reduction. As either a homogeneous
or a heterogeneous catalyst, FDH1 catalyzed the electrochemical
reduction of CO2 in 20 mM Na2CO3 (pH 6.5) to produce only
CO2-saturated 0.2 M Na2SO4 solution (pH 4.2) at 20 1C in formate, with a reduction rate more than two orders of magni-
the potential range of 0.7 to 0.8 V vs. saturated calomel tude greater than that achieved with other known catalysts for
electrode (SCE) observed a Faradaic efficiency above 50%, with the same reaction. Although the mechanism of CO2 activation
the main product being methanol. Even in 0.05 M H2SO4 at and reduction in Cr-containing enzymes such as FDH1 is not
20 1C, methanol was obtained at 0.57 to 0.67 V vs. SCE with a
Faradaic efficiency of up to 46%. Moreover, cyclic voltammetric
(CV) measurements indicated that the corrosion of molybdenum
metal to molybdenum dioxide (MoO2) might be the source of
electrons for the electroreduction of CO2 (Fig. 3). In the late
1990s, Mo-containing catalysts were found to be useful in the
reduction of chlorate, bromate, and iodide anions.48–52 The
electrochemistry of molybdenum and molybdenum oxides has
been described in detail.53 Nakazawa et al.54 employed two
types of iron–sulfur clusters, [Fe4S4(SR)4]2 (R = C6H5CH2 and
(CH3)3C) and [M2Fe6S8(SCH2CH3)9]3 (M = Mo or W), to catalyze
the electrochemical reduction of CO2. It was observed that the
reduction potential was shifted by about 0.7 and 0.5 V, respec-
tively, in the positive potential direction compared to what
occurred without any catalyst, demonstrating the catalytic effect
of these two catalysts. Bandi et al.42 thermally decomposed a
mixed metal oxide consisting of 25% RuO2, 30% MoO2, and
45% TiO2 on Ti foil to catalyze CO2 electroreduction to methanol,
but no striking performance was observed.
Regarding Cr-related catalysts, Ogura and Yoshida55 employed Fig. 4 Proposed mechanism for CO2 electroreduction catalyzed by
(III) W-containing formate dehydrogenase enzyme (FDH1), in which two
Cr –TPPCl (TPP = 5,10,15,20-tetraphenylporphyrin) for CO2
electrons are transferred from the cathode to the active site to reduce
electroreduction in dimethylformamide (DMF) solutions and
CO2 to formate, forming a C–H bond. Conversely, when formate is
compared the results with those for Co(II)–TPP, Ni(II)–TPP, oxidized, the two electrons are transferred from the active site to the
Fe(III)–TPPCl, and Fe(II)–TPP. Cr complexes with 4-v-tpy and electrode. Reprinted with permission from ref. 58. Copyright r 2008 Proc.
6-v-tpy (v-tpy = vinyl-terpyridine) have also been reported for Natl. Acad. Sci. U. S. A.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 635
View Article Online

Review Article Chem Soc Rev

fully understood, Fig. 4 shows a proposed bioelectrocatalytic reduction compared to those catalyzed by fac-Re(CO)3(bpy)Cl in
CO2 reduction reaction. In this mechanism, CO2 is known to solution. CO and oxalate (on the pure poly-Re(CO)3(vbpy)Cl film)
insert into the W–H bond in [HW(CO)5] to form a stable were found to be the main products. The results of electrochemical
formate adduct. This W–H insertion reaction is also seen in kinetic studies of poly-Re(CO)3(vbpy)Cl showed that as the film
homogeneous organometallic catalysis.59 thickness was increased, the film’s rate-determining step could be
2.1.3 Rhenium and manganese. Normally, rhenium com- changed from (1) the chemical reaction between reduced Re and
plexes have quite promising catalytic activity. In the present CO2 to (2) electron transport to the catalytic sites.
paper’s context, tricarbonyl rhenium(I) complexes are a large Schrebler et al.68 investigated the electrochemical reduction of
group of catalysts for catalyzing the photochemical and electro- CO2 on a Re film electrodeposited onto a polycrystalline Au
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

chemical reduction of CO2 to CO.60–62 support in a CH3OH solution with 0.1 M LiClO4 under atmo-
Lehn et al.61 studied the homogeneous catalysis of CO2 spheric pressure of CO2. The CO2 electroreduction displayed a
electroreduction using 9.0  104 M Re(I)(bpy)(CO)3Cl (namely, Tafel slope of 2RT/F, suggesting that the first electronation of
Lehn catalyst; bpy = 2,2 0 -bipyridine) in 0.1 M (CH3CH2)4NCl the CO2 molecule to form CO2* was the rate-determining step. It
DMF–H2O (9 : 1) solutions at 25 1C on a GCE with a potential at was found that the product distribution was strongly dependent
1.25 V vs. normal hydrogen electrode (NHE), which was much on the electrode potential at which the electrolysis was carried
less negative than 2.0 V in the absence of the catalyst. With out, as well as on the hydrodynamic conditions. Under stirred
the catalyst, the current efficiency reached 91%. The catalyst CV conditions the Faradaic efficiency of CO production was 87%
showed a reversible one-electron reduction process at 1.25 V at 1.35 V, whereas under quiescent conditions, the Faradaic
vs. NHE, forming [Re(0)(bpy)(CO)3Cl]. O’Toole et al.63 electro- efficiency of CO and CH4 production was 57% and 10%, respec-
polymerized another kind of tricarbonyl rhenium complex, tively. Schrebler et al. also prepared Re and Cu–Re microalloy
Re(CO)3(vbpy)Cl (vbpy = 4-vinyl-4 0 -methyl-2,2 0 -bipyridine), on a polypyrrole (PPy) modified Au electrodes for the electrochemical
Pt electrode to form a polymeric film that was used to hetero- reduction of CO2 in CH3OH solution with the same composition.
geneously catalyze the electrochemical reduction of CO2 to CO. Higher Faradaic efficiencies for CH4 were obtained at 1.35 V,
Results indicated that the heterogeneous catalyst could enhance with Au–PpyRe at 34% and Au–PpyCu–Re at 31%. Importantly,
the TON 20–30 times more than was observed in the homogeneous both Re and Cu–Re alloy could be highly dispersed on the PPy
case with Re(CO)3(bpy)Cl catalyst. When the Re(CO)3(vbpy)Cl films, and the amount and selectivity of CO, CH4, and H2 were
complex was coated on metallic Pt or on p-Si and polycrystalline independent of the hydrodynamic conditions of the solution. The
thin films of p-WSe2 semiconducting electrodes, the TONs were fac-Re(CO)3(vbpy)Cl was also electropolymerized onto a meso-
as high as ca. 600 and 450, respectively, for CO2 electroreduction porous TiO2 film coated on a SnO2-doped glass electrode for
to CO.64 Cosnier et al.65,66 investigated the effects that electrode CO2 electroreduction.69 The nanoporous nature of TiO2 allowed
material, film thickness, and the structure of bipyridyl ligands an increase in the 2D number of redox sites per surface area and
had on the catalytic activity, stability, and current efficiency hence achieved a significant enhancement in catalytic yield. In an
for CO2 electroreduction catalyzed by electropolymerized fac- effort to improve the catalytic activity of CO2 electrocatalysts, Cheung
Re(L)(CO)3Cl complexes (L = pyrrole-substituted bpy) on metallic et al.70 recently electropolymerized a poly-Re(CO)3(k2-N,N-PPP)Cl
Pt and carbon felt electrodes (Fig. 5).65,66 The coatings obtained film onto a GCE. Their results showed that the modified
by electropolymerization of a monomer containing one pyrrole electrode also exhibited electrocatalytic activity for the reduction
group (L1) seemed to be as stable as those prepared with of CO2 to CO.
monomers containing two pyrrole groups (L2), whereas poly- To understand the mechanism, Re(CO)3LCl complexes with
Re(L)(CO)3Cl films (L3, L4, and L5) with lower reduction different bpy ligands that contain different substitutions on the
potentials were more stable but less active towards the electro- benzene ring have been employed as example catalysts for CO2
chemical reduction of CO2. electroreduction. A systematic study of the electrochemical reduction
O’Toole et al.67 co-electropolymerized cis-[(bpy)2Re(vpy)2]2+ with of CO2 catalyzed by 1.0 mM Re(CO)3LCl complexes (L = bpy, dcbpy,
fac-Re(CO)3(vbpy)Cl or fac-[Re(CO)3(vbpy)CH3CN]+ on electrodes to dmbpy, 4,40 -di-tert-butyl-bpy, or 4,40 -dimethoxy-bpy) in CH3CN +
form thin polymeric films (heterogeneous catalysts). Metal sites on 0.1 M tetrabutylammonium hydroxide (TBAH) on a glassy carbon
these films showed increased reactivity and stability toward CO2 working electrode revealed that the electron donating/withdrawing

Fig. 5 Structure of electropolymerized fac-Re(L)(CO)3Cl complexes. L = pyrrole-substituted bpy. Reprinted with permission from ref. 65. Copyright r
1986 Elsevier.

636 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

substituents in the 4,40 positions of the bipyridine ligand had a [Re(bpy)(CO)3]* + CO2 - Re(bpy)(CO)3CO2 (10)
significant effect on CO2 electroreduction.71 The catalytic activity was
increased (with less negative reduction potentials) in the order Re(bpy)(CO)3CO2 + CO2 + 2e
OCH3 o C(CH3)3 o CH3 o H o COOH. When L = bpy, dcbpy, - Re(bpy)(CO)3 + CO32 + CO (11)
dmbpy, and 4,4 0 -di-tert-butyl-bpy, the Re complex gave the best
[ fac-Re(bpy)(CO)3]2 + 2e 2 2[Re(bpy)(CO)3]
 
(12)
catalytic activity, even 2.2 times higher than that of Lehn catalyst.
Recently, fac-(5,5 0 -bisphenylethynyl-2,2 0 -bipyridyl)Re(CO)3Cl was [Re(bpy)(CO)3] + CO2 - [Re(bpy)(CO)3CO2] (13)
explored as a catalyst for CO2 electroreduction. The results
showed a 6.5-fold increase in the current density of CO2 to CO at [Re(bpy)(CO)3CO2] + A + e
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

1.75 V vs. NHE compared to without CO2, and the Faradaic - [Re(bpy)(CO)3] + CO + [AO]
efficiency for CO production was around 45%.72 In this study,
(A = an oxide ion acceptor) (14)
the supporting electrolyte solution was CH3CN with 0.1 M
(n-CH3(CH2)3)4NPF6. In their experimental observation of the above CO2 electro-
Regarding the solution pH effect, Wong et al.34 tested the effect reduction mechanism, Johnson et al.74 reported the reaction
of four weak Brönsted acids, CF3CH2OH, C6H5OH, CH3OH, and products using an infrared spectroelectrochemical (IR-SEC)
H2O, on CO2 electroreduction catalyzed by [Re(CO)3(bpy)(py)]2+ in method, employing an optically transparent thin-layer electro-
CH3CN. They found that the addition of these acids could chemical cell. The catalysts used were [Re(CO)3(bpy)P(OEt)3]+,
enhance the rate of the catalytic process as well as improve the [Re(CO)3(bpy)CH3CN]+, Re(CO)3(bpy)Cl, and Re(CO)3(bpy)CF3SO3).
catalyst’s lifetime, suggesting that the acidity of Brönsted acid They confirmed that the [Re(CO)3(bpy)Cl]* radical was only
could increase the efficiency of the reduction process. attacked by CO2 to form [Re(bpy)(CO)3CO2]* after the dissocia-
Since the discovery of the Lehn catalyst ( fac-Re(bpy)(CO)3Cl),61 tion of Cl, and that [Re(CO)3(dmbpy)Cl]* tended to form the
the mechanism of CO2 reduction has also been explored [Re(CO)3(dmbpy)]* radical for CO2 reduction. Scheiring et al.75
through chemical synthesis, electrochemical, and spectro- conducted a mechanism study using electron paramagnetic
scopic measurements.69–71,73 Sullivan et al.73 proposed catalytic resonance spectroscopy (EPR); the catalysts employed were
pathways, including an initial one-electron reduction of the paramagnetic Re(CO)3(bpy)X complexes (X = Cl, CF3SO3,
catalyst, which then catalyzed CO2 reduction to form CO, as CH3O, H, tetrahydrofuran, CH3CN, CO, HCO2, HCO3, and
expressed in reactions (6)–(14): CH3C(O)). Furthermore, sum frequency generation (SFG) spectro-
scopy and DFT calculations indicated that Re(CO)3Cl(dcbpy) could
fac-Re(bpy)(CO)3Cl + e " [Re(bpy)(CO)3Cl]* (6)
bind to a rutile TiO2 surface through the –COOH groups of dcbpy
[Re(bpy)(CO)3Cl]* - [Re(bpy)(CO)3]* + Cl (7) in bidentate or tridentate linkage motifs, and the Re atom was
exposed to the solution in a configuration suitable for the catalysis
[Re(bpy)(CO)3]* + S - [Re(bpy)(CO)3S]* of CO2 electroreduction.76 Table 2 summarizes all the mono, bi,
(S = solvent molecule) (8) and tridentate ligands employed in Re complexes.
Manganese generally does not catalyze the electrochemical
2[Re(bpy)(CO)3]* - [ fac-Re(bpy)(CO)3]2 (9) reduction of CO2. However, some catalytically active Mn carbonyl

Table 2 Mono-, bi-, and tridentate ligands employed in Re complexes

Pyridine (py)34
[4, vinyl-]. 4-Vinyl-pyridine (vpy)67

2,2 0 -Bipyridine (bpy)61,62,71


[4, –CH2(CH2)3R; 4 0 , –CH3].65 (L1)66 R: pyrrolyl
[4,4 0 , –CH2R]. (L2)66
[4, –CH2CH(CH2R)2; 4 0 , –CH3]. (L3)66
[4,4 0 , –COOCH2(CH2)5R]. (L4)66
[4,4 0 , –COOCH(CH2R)2]. (L5)66
[5, –CH3; 5 0 , vinyl-]. 4-Vinyl-4 0 -methyl-bpy (vbpy)63,64,67,69
[4,4 0 , –OCH3]. 4,4 0 -Dimethoxy-bpy71
[4,4 0 , –C(CH3)3]. 4,4 0 -Di-tert-butyl-bpy71
[5,5 0 , C6H5CRC–]. 5,5 0 -Bisphenylethynyl-bpy72
[5,5 0 , –COOH]. 4,4 0 -Dicarboxyl-bpy (dcbpy)71,76
[4,4 0 , –CH3]. 4,4 0 -Dimethyl-bpy (dmbpy)71

[R, –CH2CH2CH2-N(pyrrole)]. N-(3-N,N 0 -Bis(2-pyridyl)propylamino)pyrrole (PPP)70

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 637
View Article Online

Review Article Chem Soc Rev

complexes, i.e., [Mn(bpy)(CO)3]+, [Mn(dmbpy)(CO)3]+,77 and [Mn(bpy- Table 3 Simplified reaction scheme for CO2 reduction by Fe(0)-porphyr-
t-Bu)(CO)3]+,78 recently were found to have some activity toward ins (por = porous, AH = acid). Reprinted with permission from ref. 86. Copyright
r 2012 Science
CO2 reduction. X-ray crystallography of [Mn(bpy-tBu)(CO)3]
(ref. 78) showed a five-coordinate Mn center, similar to its [(por)Fe(I)] + e " [(por)Fe(0)]2
rhenium analogue and to the IR-SEC of Mn(bpy-tBu)(CO)3Br. [(por)Fe(0)]2 + CO2 + 2AH - [(por)Fe(II)CO] + H2O + 2A (K)
[(por)Fe(II)CO] + [(por)Fe(0)]2 - [(por)Fe(I)] + CO (K 0 { K)
2.1.4 Iron, cobalt, and nickel
2.1.4.1 Fe/Co/Ni metal electrodes. In general, the electro- CO2 + 2AH + 2e - CO + H2O + 2A
chemical reduction of CO2 on electrodes of Group 8–10 metals, If pK(CO2+H2O) { pKAH:
2A + 2(CO2 + H2O) - 2AH + 2CO3H
such as Fe, Co, and Ni, in aqueous solutions under ambient
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

conditions produces H2 and CO or other hydrocarbons. However, 3CO2 + H2O + 2e - CO + 2CO3H
these metals have high activities for the hydrogenation of CO and/or
CO2 in heterogeneous catalytic reactions (e.g., the Fischer–Tropsch
reaction). At 30 atm, the electrochemical reduction of CO2 on in a DMF solution with tetraalkylammonium (TAA) salts, three
Fe electrodes at a constant current density of 120 mA cm2 complexes of Fe-porphyrins, including Fe-tetraphenylporphyrin
produced HCOOH with a Faradaic efficiency of B60%.79 (FeTPP), showed electrocatalytic activity for CO2 reduction,
Several factors that might have affected the reduction process mainly producing CO, with a Faradaic yield of over 94%. The
were proposed, including the diffusion of molecular CO2 to addition of weak Brönsted acids to the solution, such as
the electrode surface and the suppression of hydrogen evolu- n-propanol, 2-pyrrolidone, or CF3CH2OH,84,85 considerably improved
tion by CO adsorption on the electrode surface.80 In the electro- the reduction process catalyzed by Fe(0)–TPP. The catalytic
chemical reduction of CO2 at Ni electrodes in aqueous media, currents and the lifetime of the catalyst were both increased
H2 as well as some small hydrocarbons, such as CH4, C2H4, without significant hydrogen evolution. In contrast, the HCOOH
and C2H6, was produced. Under high CO2 pressure, the Faradaic yield decreased as the acidity of the acid synergist was increased,
efficiency for CO2 reduction on Ni electrodes could be increased and finally became negligible when CF3CH2OH was added.
by raising the CO2 pressure, lowering the temperature, and Most recently, Costentin et al.86 found that modification of
polarizing the electrode potential at a more negative FeTPP through the introduction of phenolic groups in all ortho
potential.81 It was suggested that hydrocarbons were formed and ortho’ positions of phenyl groups—to form iron 5,10,15,20-
on Fe, Co, and Ni electrodes through pathways similar to tetrakis(2 0 6 0 -dihydrolphenyl) porphyrin, i.e., FeTDHPP, and
the Fischer–Tropsch reaction of thermal catalysis.81 In CO2 iron 5,10,15,20-tetrakis(2 0 6 0 -dimethoxyphenyl) porphyrin, i.e.,
electroreduction catalyzed by Fe and Ni metals, CO adsorption FeTDDMPP—could considerably speed up the rate of CO2
on these two electrodes was observed by IR spectroscopy, electroreduction to CO by an electrogenerated iron(0) complex
and the results suggested that the catalytic activity was on a GCE. The catalyst manifested a CO Faradaic yield above
strongly related to the bonding between CO and the metal 90% through 50 million turnovers over 4 hours of electrolysis at
surface.82 low overpotential and 0.465 V, with no observed degradation.
The reason for the enhanced activity appeared to be the high
2.1.4.2 Fe/Co/Ni complex catalysts. To date, a large number local concentration of protons associated with the phenolic
of studies have been conducted on CO2 electroreduction cata- hydroxyl substituents. The stoichiometric reactions involved in
lyzed by iron, cobalt, and nickel complexes. Both the type of the the catalytic reduction of CO2 to CO by Fe(0)-porphyrins are
metal and the structure of the ligands play important roles in presented in Table 3. Using carbon nanotubes (CNTs) for
the catalytic behavior of these complexes. support, an FeTPPCl (FeP-CNT)-modified GCE for CO2 electro-
Fe/Co/Ni complexes of porphyrin. In CO2 electroreduction reduction was studied using CV and CO2 electrolysis in 0.1 M
catalyzed by metal porphyrin (M–P)-modified electrodes (M = NaHCO3 solution.87 The FeP-CNT exhibited a less negative cathode
Fe(II, III), Co(II), Ni(II), and Cr(III); P = tetraphenylporphyrin) potential and much higher reaction rate than pure FeP or CNT
in DMF solutions,55 methanol production can be expressed electrodes. By adding FDH, NADH, and methyl viologen to the
as follows: electrolyte, HCOOH became the only product, and the concentra-
tions of HCOOH formed at the cathode followed the order
CO2 + 6M–P + 6H+ - CH3OH + 6M–P+ + H2O (15) FeP-CNT > FeP > CNT. This improved catalytic activity was
attributed to synergistic catalysis and direct electron transfer
M–P + e - M–P
+ 
(16)
between Fe-porphyrin and CNTs, demonstrated using both
It has been reported that Fe(II) porphyrin yields efficient, electrochemical impedance spectroscopy (EIS) and electron
CO-selective, and durable catalysts for CO2 electroreduction when paramagnetic resonance (EPR) analysis. The p–p interaction
it is reduced to Fe(0)-porphyrin by two successive electrons between the porphyrin ring and sidewalls of the CNTs reduced
during the reduction process. In an early study, Takahashi the electron density around the Fe nuclei in FeTPPCl, which
et al.83 found that electrodes modified by Fe-meso-tetracarboxy- expanded the macrocyclic conjugated structure of FeTPPCl and
phenyl porphyrin (Fe-mTCPP) and Fe-tetraphenylporphine further increased the potential for CO2 reduction.87
sulfonate (Fe-sTPP) had insignificant catalytic activity towards Regarding cobalt complexes as catalysts, Co porphyrin cata-
CO2 electroreduction in aqueous electrolytic solutions. However, lyzed the electroreduction of CO2, producing CO in high yield.88

638 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

ligand had a strong steric effect on the catalytic activity of the


CoTPP core. Recently, catalyst-modified diamond surfaces,
dubbed ‘‘smart’’ electrodes, have been explored, with reports
of good stability and electrocatalytic activity for CO2 electro-
reduction to CO in CH3CN.94 In this research, a catalytically
active Co complex was covalently attached to a B-doped, p-type
conductive diamond. In addition, some fluorinated derivatives
of CoTPP (the phenyl rings of the CoTPP were substituted with
electron withdrawing groups, such as F and CF3) have also
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

been explored as catalysts for CO2 electroreduction, yielding


positive results.95
It should be noted that the catalytic behavior of a metal
porphyrin for CO2 electroreduction can be affected by many
factors. Even the same catalyst may display different activity
when used as a homogeneous catalyst rather than a hetero-
geneous catalyst.55,96 CO2 pressure and solution temperature
also have a significant effect on CO2 electroreduction. For
Fig. 6 Cyclic voltammograms of the Co(II)–TPP complex at different
example, when CO2 pressure was increased from 1 to 20 atm,
modification stages of the glassy carbon electrode in pH 6.8 phosphate
buffer solution under a N2 atmosphere; py = aminopyridine; sweep rate = the current efficiencies of CO2 electroreduction by Fe- and
50 mV s1. Reprinted with permission from ref. 90. Copyright r 1997 Co-meso-TPP-supported gas diffusion electrodes (GDEs) were
Elsevier. increased by up to 97.45% and 84.6%, respectively.97 According
to the literature, the strength with which a catalyst bonds to
the electrode surface has significant effects on the catalyst’s
In the presence of aquopentacyanoferrate(II) or 2-hydroxyl-1- activity,98 as has been demonstrated by photoelectron spectro-
nitrosonaphthalene-3,6 disulphonatocobalt(II) as the homogeneous scopy (PES) and scanning tunneling spectroscopy studies, as well
catalyst, the Pt plate electrodes modified by CoIITPP, NiIITPP, as quantum-chemical calculations based on density functional
FeIIITPPCl, or FeIITPP showed decreasing CO2 reduction activity theory (DFT).
in an aqueous solution of methanol plus 0.1 M KCl (pH 3.5),55
Fe/Co/Ni complexes of phthalocyanine. Metal phthalocyanine
indicating that the CoIITPP complex, as a heterogeneous
(MPc) complexes seem to have less catalytic activity toward CO2
catalyst, certainly had higher catalytic activity. Using a GCE
electroreduction than metal porphyrin complexes. For example,
modified by CoIITPP or CoTPP-py-NHCO, the current efficiency
iron phthalocyanine tetrasulfonate (FeTSPc), as distinct from
of CO2 electroreduction to CO in an aqueous phosphate buffer
CoTSP and NiTSP, did not show remarkable homogeneous
solution (a mixture of l/15 M NaH2PO4 and l/15 M Na2HPO4) was
catalytic activity for CO2 electroreduction in Clark–Lubs buffer
92% at 1.1 V vs. SCE. A CoTPP-py-NHCO-modified GCE also
solution.99 CoPc and NiPc complexes coated on a graphite
showed high catalytic stability, with an overall TON exceeding 107.89
electrode showed some catalytic activity for CO2 electroreduction,
An amine cation radical method was explored to coordinate a
with HCOOH being the predominant product in aqueous solu-
CoII–TPP complex directly onto a GCE to catalyze CO2 electro-
tions (pH 3–7).100–102 The electrocatalytic reduction of CO2 in an
reduction, and the results were quite promising (Fig. 6).90
aqueous electrolyte catalyzed by a graphite electrode coated with
Compared to a GCE physically attached to Co(II)–TPP, this
a CoPc/poly-4-vinylpyridine (PVP) film was also studied, and
GCE chemically bonded to Co(II)–TPP yielded enhanced CO2
electroreduction.
In the exploration of new catalysts, 5,10,15,20-tetrakis
(4-methoxyphenyl)porphyrinato cobalt(II) (CoTMPP) adsorbed
on a nanoporous activated carbon fiber (n-ACF) electrode was
found to be an effective catalyst, yielding CO with current
efficiencies of up to 70%.91 The catalytic activity of a binuclear
Co complex, Co(I)TMPyP–MTPPS (TMPyP = a,b,g,d-tetrakis-
(1-methylpyridinium-4-yl)porphyrin p-toluenesulfonate; TPPS =
tetraphenylporphine tetrasulfonic acid; M = metal), was also
tested, and significant catalytic activity was observed.92 The
researchers believed that the catalytic activity should be
strongly dependent on the central metal (M) of TPPS, which
could serve as an electron mediator.92 Imaoka et al.93 prepared
phenylazomethine dendrimers to bear CoTPP cores as homo- Fig. 7 Proposed mechanism for CoPc in an aqueous medium in the
geneous catalysts for CO2 electroreduction on a GCE in a DMF presence and absence of CO2. Reprinted with permission from ref. 103.
solution containing a strong Lewis acid. They observed that the Copyright r 1995 Elsevier.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 639
View Article Online

Review Article Chem Soc Rev

enhanced catalytic activity was observed compared to what cycles, no significant deactivation was observed (the turnover
occurred with a pure CoPc catalyst.103 As shown in Fig. 7, the frequency was B103 mol of CO produced per mole of nickel
peripheral N atom on the Pc ring can be protonated by complex in 1 hour).112 A study of the reduction mechanism on
the addition of a proton following the first reduction. The mercury, using CV, polarography, and electrocapillarity, indi-
second reduction generates the active species for the catalytic cated that the adsorbed complex Ni(I)-cyclam on Hg was the
reduction of both proton (upward arrow) and CO2 (downward active catalyst.113 Theoretical calculations performed by
arrow), generating H2 and CO, respectively, at pH 4.4. When Sakaki114 provided a reasonable description of the catalytic
CoPc was incorporated into a PVP film coated on a graphite mechanism, based on the oxidation states of the adsorbed
electrode, the polymer-incorporated CoPc showed high catalytic complexes ([Ni-cyclam]2+ and [Ni-cyclam]+), which were strongly
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

activity towards CO2 electroreduction, with high selectivity dependent on the electrode potentials. Regarding the mecha-
from CO2 to CO, where PVP functioned as a coordinative and nism, Fujihira113 observed that the desorption of Ni(I)-cyclam
weakly basic agent.104 The coordination of PVP to CoPc was gradually slowed down with the formation of nonadsorbable
considered to have caused the increase in electron density on Ni(I)-cyclam-CO by the reaction of Ni(I)-cyclam in solution with
the central metal ion, facilitating the formation of the inter- electrocatalytically generated CO from CO2. In the CO2 reduction
mediate. Cobalt octabutoxyphthalocyanine (CoPc(BuO)8) process, Ni(cyclam)2+ was only weakly adsorbed over a limited
coated on a graphite electrode was also explored as a catalyst potential range, and in quantities substantially less than one
for CO2 reduction; the results showed that this catalyst had monolayer, whereas Ni(cyclam)+ could be strongly adsorbed on
higher activity and selectivity from CO2 to CO than non-sub- the electrode surface over a wide potential range.115 However,
stituted CoPc.105 Under typical conditions at pH 4.4, the most the high electrocatalytic activity of the Ni(cyclam)2+ complex
active and selective CO2 reduction was achieved at 1.30 V vs. for the reduction of CO2 at a static mercury electrode was
Ag/AgCl; the production selectivity of CO/H2 was reported to severely diminished in the presence of CO when the mercury
be B4.2. In addition, Zhang et al.106 employed a rotating electrode was not stirred.116 The cause of the decrease in
ring (platinum)-disk (graphite) electrode to analyze CO2 electro- activity was proposed to be an insoluble complex of Ni(O) and
reduction to CO, catalyzed by N,N0 ,N00 ,N00 -tetramethyltetra-3,4- Ni(cyclam)CO, which formed during the reduction of CO2.
pyridoporphyrazinocobalt(II) coated on a graphite disk and To explore this mechanism, in situ analyses of the products of
protected by a Nafion film. In a microbial electrolysis cell, cobalt CO2 electroreduction catalyzed by Ni-cyclam were carried out
tetra-amino phthalocyanine (CoTAPc) coated on multiwalled using differential electrochemical mass spectroscopy during CV
CNTs was reported to produce high current efficiency for on an amalgamated-gold mesh electrode.117
HCOOH production.107 A binuclear nickel complex, Ni2(biscyclam)4+, was found to
Besides phthalocyanine ligands, other kinds of hexaaza- have similar catalytic activity to that of Ni(cyclam)2+ for CO2
macrocycle ligands, such as phenanthroline and bipyridine, electroreduction. In Ni2(biscyclam)4+, two Ni atoms are indirectly
and their complexes with Co(II), Ni(II), and Cu(II) metal centers, linked. When DMF with low water content was used as a solvent,
have also been explored as CO2 electroreduction catalysts.108 high Faradaic yields of HCOO were observed (up to 75%) in
For example, CV and UV-vis spectroscopy have been employed addition to CO.118 Methyl substitution of the amines on the cyclam
to study the electroreduction of CO2 using, as an electrocatalyst, ring was also explored as a ligand for the Ni complex when used
hexaazamacrocycles derived from the condensation of 1,10- as a CO2 reduction electrocatalyst, and the result showed some
phenanthroline and its Co(II) complex, dissolved in DMF solution. catalytic effects. The Ni(II) complex of N-hydroxyethylazacyclam
The results showed that the ligand had no catalytic activity, (i.e. 3-(20 -hydroxyethyl)-1,3,5,8,12-penta-azacyclotetradecane nickel(II)
whereas its cobalt complex showed electrocatalytic activity toward perchlorate) appeared to be more active than unsubstituted
the reduction of CO2, generating CO and HCOOH.109 Ni(cyclam).119 Recently, Schneider et al.120 explored a series of
materials that are structurally similar to [Ni(cyclam)]2+ to test
Fe/Co complexes of corroles. Some iron and cobalt complexes
them as electrocatalysts for CO2 reduction at a mercury pool
with corroles have also been explored as CO2 electroreduction
working electrode in aqueous solution.120 Both [Ni(HTIM)]2+
catalysts. For example, Ph3PCoIII(tpfc) (tpfc = 5,10,15-tris-
(HTIM = C-RRSS-2,3,9,10-tetramethyl-1,4,8,11-tetraazacyclotetra-
(pentafluorophenyl)corrole), ClFeIV(tpfc), and ClFeIV(tdcc)
decane) and [Ni(DMC)]2+ (DMC = C-meso-5,12-dimethyl-1,4,8,11-
(tdcc = 5,10,15-tris(2,6-dichlorophenyl)corrole) dissolved in
tetraazacyclotetradecane) showed better electrocatalytic activities
CO2-saturated CH3CN solution were tested, using CV, for their
than Ni(cyclam)2+. Schneider et al. suggested that (1) the cata-
catalytic activity towards CO2 electroreduction; the results
lyst’s geometry should be suitable for its adsorption onto the
indicated that the CoI and FeI complexes were both effective
mercury electrode surface and (2) there should be electronic
catalyst centers.110
effects of methyl groups or cyclohexane rings on the cyclam
Ni complexes of cyclams. Nickel complexes of cyclam (i.e., backbone. Additional observations have also been made about
1,4,8,11-tetraazatetradecane) and Ni(cyclam)2+ have been recog- the influence of methyl substitution on these catalysts’ activities
nized as highly selective catalysts for the electroreduction of (Fig. 8).120,121 Abba et al. found that the structural features of the
CO2 to CO in aqueous solution on a mercury cathode (as the cyclam and azacyclam framework played an important role in
working electrode).111,112 The stability of such catalysts was also the enhanced catalytic efficiency of Ni-cyclam derivatives for CO2
found to be remarkable. Even after thousands of catalytic electroreduction. Even small deviations from such a geometrical

640 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

the macrocycle methyl groups could destabilize the adducts,


depending on the nature of the individual complex catalyst.
With a Ni complex, it was reported that both RRSS-NiIIHTIM(ClO4)2
(HTIM = 2,3,9,10-tetramethyl-1,4,8,11-tetraazacyclotetradecane)
and NiIIDMC(ClO4)2 (DMC = C-meso-5,12-dimethyl-l,4,8,11-tetra-
azacyclotetradecane) showed electrocatalytic activity for CO2
reduction. However, the latter was even more active than
Ni(cyclam)2+ catalysts, whereas the former was not.126 In addition,
for Ni complex catalysts, the issue of catalyst poisoning during CO2
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

electroreduction seems to be a concern. For example, this issue


was reported when researchers employed isomers of a tetraaza-
macrocyclic Ni(II) complex in solutions saturated with argon, CO,
and CO2.125–127
Some Ni-macrocyclic complexes were found to have catalytic
activity for CO2 electroreduction from CO2 to oxalate.128 For
example, with a homogeneous electron-transfer rate constant
of B105 M1 s1, the complex Ni-Etn(Me/COOEt)-Etn was
tested as a selective catalyst for CO2 electroreduction to yield
oxalate. In this test, electrolysis experiments were carried out
using 1.5  105 mol of the Ni complex as the catalyst in
CO2-saturated CH3CN + 0.25 M Bu4NClO4 solution. The overall
reaction mechanism was interpreted in terms of an outer-
sphere electron-transfer reaction followed by dimerization of
CO2 radical anions. In Table 4, the structure of the Fe/Co/Ni
complexes and their tetradentate ligands for CO2 electro-
chemical reduction catalysts are summarized.
Fe/Co/Ni complexes with tridentate ligands. Iron, cobalt, and
nickel complexes with tridentate ligands such as 40 -vinyl-2,2 0 :6 0 ,200 -
Fig. 8 (A) Cyclic voltammograms of 1.0 mM Ni(cyclam)2+ in 0.1 M KCl(aq) terpyridine (v-tpy) can also be employed as catalysts for CO2
(GC electrode; 100 mV s1 scan rate). (B) CVs of 1.0 mM Ni(cyclam)2+, electroreduction.57,129–132 The structures of these catalysts are
Ni(DMC)2+, and Ni(TMC)2+ in a 0.08 M tetrabutylammonium hexafluoro- provided in Table 5-(T1). An iron complex, [Fe(4-v-tpy)2]2+, electro-
phosphate electrolyte (1 : 4 water–acetonitrile (CH3CN); GC electrode;
polymerized onto an electrode yielded a TON in excess of 15 000.
100 mV s1 scan rate). Reprinted with permission from ref. 120. Copyright
r 2012 American Chemical Society. It was observed that heterogeneous catalysis with v-tpy (electro-
polymerized on the electrode) gave a higher catalytic activity than
homogeneous catalysis (dissolved in solution).132 Heterogeneous
arrangement caused the electrocatalytic effect to be drastically catalysis by iron complexes of 4-v-tpy and 6-v-tpy, which were
reduced or completely lost.122 electropolymerized onto a GCE, was very selective, yielding only
For CO2 electroreduction catalyzed by Ni-cyclam complexes formaldehyde.57 The geometric structure and degree of conjuga-
on non-mercury electrodes, a GCE was used as the electrode tion of the ligand, as well as the orientation and fixation of the
substrate on which a Ni-cyclam complex catalyst was coated, macrocyclic metal complexes on the electrode surface, were found
together with a Nafion film123 or with a poly-(allylamine) to be important for catalytic activity and stability in CO2 electro-
(PALA) backbone, and some effective electrocatalytic activity reduction.133 Furthermore, iron, cobalt, and nickel complexes of
was observed.124 The orientation of the nickel-cyclam complex some diacetylpyridine-derived tridentate ligands also appear to be
on the electrode surface was found to be critical to the catalytic effective electrocatalysts for CO2 reduction.134
effect.123
Fe/Co/Ni with bidentate ligands. Fe, Co, and Ni complexes
Ni/Co complexes of tetraazamacrocycles. Some nickel and with bidentate ligands such as H2dophen,135 H2salen,136 and
cobalt complexes of tetraazamacrocycles, such as Co(I)-14- salophen137 have also been employed to electrocatalyze CO2
membered tetraazamacrocycles, can also catalyze the electro- reduction. In addition, polypyrrole Co(II) Schiff-base complexes
chemical reduction of CO2 to CO and H2 in either CH3CN or were electropolymerized on a polished GCE to catalyze CO2
water solution, with current efficiencies of greater than 90%.125 reduction.138 The corresponding structures of the above catalysts
In the reduction process, charge transfer from Co to bonded are given in Table 5-(T2). However, the most studied systems
CO2 was found to be an important factor in stabilizing the CO2 seemed to be iron-based complexes. For example, electrolysis
adducts. The H-bonding interactions between the bound CO2 of CO2 using (Fe(dophen)Cl)22HCON(CH3)2 and Fe(dophen)-
and amine macrocycle N–H protons might serve to additionally (N-MeIm)2ClO4 as catalysts at 2.0 V vs. a ferrocenium–
stabilize the adduct in some cases, while steric repulsion by ferrocene reference electrode gave a mixture of CO, HCOO

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 641
View Article Online

Review Article Chem Soc Rev

Table 4 Fe/Co/Ni complexes and their tetradentate ligands for CO2 electrochemical reduction catalysts

Ligand core structure Substitution on the ligand core structure Fe/Co/Ni-complexes


1,4,7,10-H are substituted by 4-carboxyphenyl Fe-meso-tetra(carboxyphenyl)porphine83
1,4,7,10-H are substituted by 4-sulfophenyl Fe-tetraphenylporphine sulfonate83
1,4,7,10-H are substituted by phenyl Fe–L;32,33,55,84–86,97,98 Co–L;55,89,90,95,97,98
Ni–L,55,96 L = tetraphenylporphyrin (TPP)
1,4,7,10-H are substituted by 2 0 6 0 - Fe–L;86 Co–L,91 L = 5,10,15,20-tetrakis-
dihydroxylphenyl (2 0 6 0 -dihydroxylphenyl)-porphyrin
1,4,7,10-H are substituted by 2 0 6 0 - Fe-5,10,15,20-tetrakis(2 0 6 0 -dimethoxyphenyl)-
dimethoxyphenyl porphyrin86
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

1,4,7,10-H are substituted by 4-methoxyphenyl Co-5,10,15,20-tetrakis (4-methoxyphenyl)-


porphyrin (CoTMeOPP)133
(1) Porphyrin 1,4,7,10-H are substituted by p-ethynylphenyl Co-5,10,15,20-tetra(p-ethynylphenyl)-porphyrin94
1,4,7-H are substituted by C6F3H2, C6H4CF3, Co–L,95 L = T3FPP, T3CF3PP, and TF5PP
and C6F5, respectively

Fe–L,100 Co–L,100–104,133 Ni–L,100,102


Unsubstituted
L = phthalocyanine
(1 or 2)H, (3 or 4)H, (5 or 6)H, and (7 or 8)H M-phthalocyanine tetrasulfonate;
are substituted by ‘‘SO3H’’ M = Fe, Co, Ni99
1,2,3,4,5,6,7,8H are substituted by 8 of Co-octabutoxyphthalocyanine105
‘‘nCH3CH2CH2CH2’’
(1 or 2)H, (3 or 4)H, (5 or 6)H, and (7 or 8)H M-tetrakis aminophthalocyanine,
are substituted by ‘‘NH2’’ M = Fe, Co, Ni371
Naphthalocyanato cobalt(II)133

(2) Phthalocyanine

(C6H5)3PCo(III)L; ClFe(IV)L, L = 5,10,15-


1,2,3 H are substituted by pentafluorophenyl
tris(pentafluorophenyl)corrole110
1,2,3 H are substituted by 2,6-dichlorophenyl ClFe(IV)L, L = 5,10,15-tris(2,6-
dichlorophenyl)corrole110
Hydrophobic vitamin B12133

(3) Corrole

Ni(cyclam)Cl2111–116,118,120–124,126
Other (N–)H substituted cyclams Ni–L,119–121,123,136 L= (N–)H substituted
cyclams

[Ni2(biscyclam)]4+ (ref. 118)

Biscyclam
(4) 1,4,8,11-Tetraazacyclotetradecane (cyclam)
Ref. 119

(5) 14-Membered tetraazamacrocycles Co,125 Ni125,127,128

— Ni,108 Co108,109

(6) Hexaazamacrocycle (ref. 109)

(major product) and C2O42.135 In this catalysis process, the in situ FTIR measurements, where the formation of an Fe–Z1-
Fe(I) species seemed to play an important role. As well, the rate CO2 intermediate led to the production of CO. The formation of
of CO2 reduction was enhanced by adding CF3CH2OH or oxalate was attributed to the dimerization of two reduced CO2
CH3OH, as a proton source, to the electrolyte. Both iron carbonyl molecules. The overall reaction mechanism was proposed to be
and iron formato species were detected as intermediates by as follows:

642 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Chem Soc Rev

Table 5 (T1) Fe/Co/Ni complexes and their tetradentate ligands for CO2 electrochemical reduction catalysts. (T2) Fe/Co/Ni complexes and their bidentate ligands for CO2 electroreduction catalysts

(T1)

This journal is © The Royal Society of Chemistry 2014


2,2 0 :6 0 ,200 -Terpyridine (typ) [M(tpy)2], 6-Vinylterpyridine (6-v-tpy)57 [M(6-v-tpy)2],
M = Fe,Co,Ni131,132 4-Vinylterpyridine (4-v-tpy) [M(4-v-tpy)2], M = Fe,Co,Ni
4 0 -Vinyl-2,2 0 :6 0 ,200 -terpyridine (v-tpy)
M = Fe,Co,Ni57
[M(v-tpy)2], M = Fe,Co,Ni129–132

Dimethylaminomethylphenylsulfanyl-
phenylamine (dapa) [M(dapa)2], 2,3,5,6-Tetra-2-pyridylpyrazine [M(tppz)2], n,n,n 0 ,n 0 -Tetrakis(2-pyridylmethyl)ethane-
M = Fe,Co,Ni131 M = Fe,Co,Ni131 1,2-diamine (tpen)131 [M(tpen)], M = Co,Ni 2,4,6-Tris(2-pyridyl)-1,3,5-triazine
(tptz)213, [M(tptz)2], M = Fe,Co131

6-Bis-[1-(phenylimino)- 2,6-Bis[1-(2-methyl-
ethyl]pyridine (DAPA), 2,6-Bis[1-(benzylimino)-ethyl]pyridine pyridylimino)ethyl]-
(DAPB), Fe,Co,Ni134 2,6-Bis[1-(2-ethylpyridyl- Vinyl-DAPA,
Fe,Co,Ni134 pyridine (DAPMP),
imino)-ethyl]pyridine Br-DAPA Fe,Co,Ni134 Fe,Co,Ni134
Fe,Co134
(DAPEP), Fe,Ni134
View Article Online

Chem. Soc. Rev., 2014, 43, 631--675 | 643


Review Article
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Review Article

Table 5 (continued)

(T2)

644 | Chem. Soc. Rev., 2014, 43, 631--675


N,N 0 -Bis(salicylidene)-o-phenylenediamine N,N 0 -Bis(salicylidene)ethylenediamine (H2salen)136
2,9-Bis(2-hydroxyphenyl)-1,10-phenanthroline (H2dophen)135
(salophen)137

Complex of Co and 2-(3-pyrrole-l-yl-propylimino-methyl)-


phenol138
Ni(COD)2 = nickel dicyclooctadiene166
[Fe4S4(SR)4]2 (ref. 54,139,140)

Polyaniline (PAn)159,161
View Article Online

This journal is © The Royal Society of Chemistry 2014


Chem Soc Rev
View Article Online

Chem Soc Rev Review Article

FeIII(dophen)Cl + e " FeII(dophen) + Cl (17) Carbon nanotube supported metal catalysts. Unlike the
Fe/Co/Ni metal electrodes or their complex catalysts mentioned
FeII(dophen) + e " [FeI(dophen)] (18) above, Fe nanoparticles supported on carbon nanotubes
[Fe (dophen)] + e " [Fe (dophen)]
I   I 2
(19) (Fe/CNTs) were found to show high electrocatalytic activity
toward CO2 reduction to small liquid fuels. Under conventional
liquid-phase operations, C2H4 was found to be the main
product of the electrocatalytic reduction of CO2 at high opera-
tion potentials,7,142 whereas in the gas phase, isopropanol was
the main product.143,144 In this case, Fe/CNTs even showed
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

better activity than Pt/CNTs but were unstable. The explanation


for their degradation was thought to be that the electrolyte,
especially K ions, reacted with iron particles, causing their
(20) dissolution and migration. In the case of Pt/CNTs, K ions could
cover the Pt particles, causing deactivation. Preliminary tests
(21)
indicated that Fe–Co/CNT catalysts were more stable. When
CNTs were further doped with nitrogen to form N-doped CNTs
(NCNTs), a supported FeOx catalyst (FeOx/NCNTs) yielded
[Fe(dophen)(CO)]+ + e + S " Fe(dophen)S + CO (22) enhanced electrocatalytic activity and selectivity compared to
(S is a solvent molecule) FeOx deposited on pristine or oxidized CNTs.145 These iron
nanoparticles supported on NCNTs were able to selectively
[Fe(dophen)(CO2)] + S " Fe(dophen)S + CO2* (23) reduce CO2 to isopropanol. To study the mechanism, the researchers
employed microcalorimetry to determine the chemisorption
2CO2* - C2O42 (24)
sites of adsorbed CO2; the results showed that NCNTs could

The mechanism of HCOO formation was proposed to be as cause the formation of small nanoparticles on which there were
follows: reversible sites (120 kJ mol1). This approach had, in fact,
already been utilized in the photoelectrocatalytic synthesis of
[Fe(dophen)] + AH " Fe(dophen)H + A (25)
solar fuels from CO2.144,146 Furthermore, Bocarsly et al.147
employed two N-containing heteroaromatics (imidazole and
(26)
pyridine) as homogeneous ‘‘aromatic amine catalysts’’ in the
photoelectrochemical reduction of CO2 at some illuminated
(27) iron pyrite (FeS2) electrodes. In their study of the catalysis
mechanism of a series of imidazole derivatives, CV measurements
In the above mechanisms expressed by reactions (21) to (27), were carried out (over a scan rate range of 5 to 200 mV s1), and
HA is the proton source.135 they found that imidazole could reduce CO2 to a mixture of CO
and HCOOH at a moderate potential, while pyridine selectively
2.1.4.3 Other Fe/Co/Ni-containing catalysts produced formic acid.
Fe4S4 cluster catalysts. Cubane-type Fe4S4 clusters139 are a
unique type of iron complexes in which the ‘‘Fe4 active site’’ Fe complexes with CO and CN ligands. Rail and Berben148
structure plays a significant role in the electron-transfer reac- reported that under appropriate conditions, Et4N[Fe4N(CO)12]
tions for CO2 electrocatalytic reduction (see Table 5-T2). In this could be a catalyst for both the hydrogen evolution reaction
regard, Tezuka et al.140 reported the electroreduction of CO2 cata- (HER) and CO2 electroreduction at 1.25 V vs. SCE using a GCE.
lyzed by two tetranuclear iron–sulfur clusters, [Fe4S4(SCH2C6H6)4]2 Everitt’s salt (ES) coated on a Pt plate electrode was also found
and [Fe4S4(SC6H6)4]2, in DMF solution. Their results showed to have activity for CO2 electroreduction to methanol149–151 in
that without the catalysts, oxalate was predominantly formed in the presence of either pentacyanoferrate(II) (Na3FeII(CN)5H2O) or
CO2 reduction, together with small quantities of formate and CO, amminepentacyanoferrate(II) (Na3FeII(CN)5NH3). The proposed reac-
whereas formate was obtained preferentially in the presence of tion mechanism was that the ES (Prussian blue reduction product)
the catalyst clusters. It was found that CO2 could be generated was first reduced from Prussian blue (PB) (KFeIII[FeII(CN)6]) on
through electron transfer from the reduced clusters to CO2 in the the cathode to start the reduction by transferring electrons to
bulk solution. In another study of CO2 electroreduction catalyzed CO2. In another study, it was found that CO2 could also be
by [Fe4S4(SCH2C5H6)4]2, it was found that the cubane structure reduced to CH3OH at an ES-mediated electrode in the presence
of the cluster could be collapsed, generating two main products, of the 1,2-dihydroxybenzene-3,5-disulfonate (tiron) ferrate(III)
C6H5COO and HCOO.54 Adding C6H5CH2SH could preserve complex and ethanol.152 The proposed mechanism was as
the cluster structure for a long time, producing more C6H5COO follows. First, a weak coordination bond was formed between
than HCOO. Recently, Yuhas et al.141 obtained enhanced electro- the central metal and ethanol, followed by the insertion of CO2
catalytic activity for CO2 electroreduction catalyzed by both ternary to form an intermediate Fe(III)–tiron–ethyl formate complex;
Ni–Fe4S4 and Co–Fe4S4-based biomimetic chalcogels.141 this complex was then finally reduced by ES, with the

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 645
View Article Online

Review Article Chem Soc Rev

consumption of protons in the solution, to CH3OH and the


initial metal complex:

CO2 + 6ES + 6H+ " CH3OH + 6PB + 6K+ + H2O (28)

6PB + 6K+ + 6e = 6ES (29)

Although catalyzed CO2 electroreduction to CH3OH was also


Fig. 9 Schematic of CODH-catalyzed CO2 electroreduction to CO. Rep-
feasible with both quinone-derivative-coated Pt and stainless rinted with permission from ref. 173. Copyright r 2003 American
steel electrodes,153 the current efficiency (>50%) was much lower Chemical Society.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

than what the ES-coated Pt plate electrode yielded (>80%).154 The


high performance of this ES catalyst was simultaneously con-
firmed for the photocatalytic conversion of CO and CO2,155 and Simón-Manso et al.166 synthesized several kinds of dinuclear
the same catalyst was used to produce CH3OH fuel from CO2 for nickel(0) complexes with dppa as a bridging ligand, expressed as
a direct liquid fuel cell.156 Ni2(m-dppa)2(m-CNR)(CNR)2 (R = Me(CH3), n-Bu(CH3CH2CH2CH2),
or 2,6-(CH3)2C6H3). The products of CO2 reduction electrocatalyzed
Prussian blue-based catalysts. Based on their observations by these complexes were mainly CO and CO32, with a small
regarding the catalytic activity of the Co-2-hydroxy-l-nitro- amount of HCOO formed when residual water was present.
sonaphthalene-3,6-disulfonate complex in the homogeneous The steric effect of such large multinuclear Ni complex mole-
catalysis of CO2 electroreduction using an ES-mediated elec- cules should be a major factor in the efficiency of CO2 electro-
trode in CH3OH,157 Ogura et al.158 carried out a further study reduction.165 Typical candidates are dinickel complexes [Ni2L2–6]4+
using a PB-polyaniline (PAn) dual film coated Pt/Co-2-hydroxy-l- or pentaazamacrocycles with (CH3OHCH2)n bridges (n = 2, 3, 4, 6)
nitrosonaphthalene-3,6-disulfonate complex electrode in or a p-xylylenediamine linkage (L6). The redox potentials were
aqueous solution158 (the structure is presented in Table 5-T2). remarkably constant, but the current peak separations increased,
The PB film was first electrodeposited on a Pt plate from an reflecting slower electron transfer due to the steric effect. The
aqueous ferric ferricyanate solution; the PAn film was then catalytic currents increased slightly as the linking chain length
deposited on a thin PB-coated electrode by repeated potential increased, due to improved stereochemical constraints.
cycling in KCl solution (pH 1) containing 0.1 M C6H5NH2.
Lactic acid, ethanol, acetone, and methanol were detected at Enzyme catalysts. Shin et al.173 reported using carbon monoxide
a low overpotential (0.6 V vs. SCE) under ambient conditions. dehydrogenase (CODH) from Moorella thermoacetica to catalyze
Ogura et al.159 then developed Fe–L (L = 4,5-dihydroxybenzene- CO2 electroreduction to CO with a current efficiency of B100% at
1,3-disulfonate and 2-hydroxy-l-nitrosonaphthalene-3,6-disulfonate) 0.57 V vs. NHE in a 0.1 M phosphate buffer solution (pH 6.3).
complex-immobilized PAn/PB-modified electrodes. The results of CODH can in fact catalyze microbial interconversion between CO
CO2 reduction at dual-film Pt electrodes modified with and without and CO2;174 Fig. 9 presents a schematic of the reaction pathway.
Fe(II) complexes confirmed that the Fe-4,5-dihydroxybenzene-1,3- There are three types of CODH: Mo-CODH, in which the
disulfonate complex-immobilized PAn/PB-modified electrode active site is a CuMo-pterin; Ni-CODH, in which the active site
yielded high catalytic activity and product selectivity for lactic acid. is a [Ni4Fe-5S] cluster; and Ni-CODH/ACS, in which Ni-CODH is
It was found that CO2 could be reduced at the active centers existing part of a larger complex structure for CO2 reduction. Parkin
in the coated film as well as at the electrode/solution interface. et al.175 studied rapid and efficient electrocatalytic CO2/CO
Interestingly, in CO2 electroreduction catalyzed by Fe(II) complex- interconversion by carboxydothermus hydrogenoformans CO
immobilized PAn/PB/Pt electrodes, formic acid was also one of dehydrogenase I (Ch Ni-CODH; see Fig. 10), which was attached
the products, in addition to lactic acid.160 The formation process onto a pyrolytic graphite ‘‘edge’’ (PGE) electrode. This approach
for products such as lactic acid, formic acid, methanol, ethanol, provided a way to study the electrocatalytic activity of enzymes
and so on was examined using FTIR reflection spectra.161–163 under strict potential control; at the same time, it set a standard
for future studies of CO and CO2 electrochemical conversion.
Multinuclear nickel complexes. Unlike iron and cobalt, nickel 2.1.5 Pt group metals
can form multinuclear complexes, which have also been explored 2.1.5.1 Pt group metal electrodes. Pt group metals are well-
as catalysts for CO2 electroreduction.164,165 Ni atoms were linked known catalysts for CO2 electroreduction. CO2, CO, and H2
with each other either directly164,166 or indirectly.118,165,167 It has adsorption on Pt group metal-based electrode surfaces differs
been found that in other multinuclear complexes, metal atoms depending on the metal, leading to different catalytic activities/
are almost all indirectly linked.168–172 Lee et al.167 studied the stabilities and product selectivities.176–181
electrocatalytic activity of a multinuclear nickel(II) complex,
Ni3(L)(ClO4)6 (L = 8,80 ,800 -(2,20 ,200 -nitrilotriethyl)-tris(1,3,6,8,10,13,15- Ruthenium. Normally, Ru metal shows high catalytic activity
heptaazatricyclo[11.3.1.13,15]octadecane)), towards CO2 reduction in the gas phase conversion of CO2 to CH4.181 In the aqueous
and compared these results with those from catalysis by a mono- electrochemical process of CO2 reduction, supported Ru sponge
nuclear complex, such as [Ni(cyclam)]2+, in CH3CN–H2O (9 : 1, v/v). electrodes did not show high catalytic activity.38,178 However, due
Unfortunately, the catalytic efficiencies (TON) of [Ni3(L)]2+ were to their high stability, Ru electrodes were used for long-term CO2
both lower than those of [Ni(cyclam)]2+ and monometallic [Ni(3)]2+. reduction at a constant potential.177

646 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 11 FTIR spectrum of adsorbed CO on a Pt electrode formed during


the electrochemical reduction of CO2 in the CH3CN electrolyte with a
Fig. 10 Structure of Ch CODH II showing the two subunits (blue and water concentration of 10.5 mM. Potential with respect to Fe/Fc+: +1.42
gray) and the arrangement of Fe–S clusters (black) that relay electrons to and 0.08 V. Reprinted with permission from ref. 195. Copyright r 2000
and from the [Ni4Fe-5S] active sites (red). Reprinted with permission from American Chemical Society.
ref. 175. Copyright r 2007 American Chemical Society.

was increased by increasing the CO2 pressure, whereas the


Palladium. The electrocatalytic activity of Pd for CO2 efficiency for hydrogen formation was decreased. In a water-
reduction was first investigated in a 1.0 M NaHCO3 solution.182 free electrolyte, CO2 electroreduction on a Pt electrode yielded
HCOOH and CO (main products) and small amounts of hydro- oxalate as the main product. Since there was no water, H2
carbons (from methane to hexane) resulted from an electrolysis evolution was not a concern. In this case, the CO2 that diffused
process catalyzed by a Pd electrode in CO2-saturated KHCO3 to the electrode might have been readily reduced to CO2, with
aqueous solution.183,184 It was observed that the current yield of the CO2 then reacting with CO2 to form oxalate. By increasing
CO could be increased substantially by increasing the pressure; the H2O concentration in the solution, H2 evolution was
when the electrode potential was held at 1.8 V vs. Ag/AgCl in 0.1 M enhanced, leading to the production of HCOO and CO rather
KHCO3, the yield went from 5.3% at 1.0 atm to 57.9% at than oxalate, which was confirmed by in situ FTIR reflection
50 atm.184 The evolution of H2 can be suppressed by hydrogen absorption spectroscopy (Fig. 11).193–195
absorption on the Pd surface, and this absorbed hydrogen can Regarding CO2 electroreduction catalyzed by supported Pt
react with the adsorbed reaction intermediates to change the nanoparticle catalysts, Centi et al.196 employed carbon-supported
electrocatalytic activity.185 Hydrogenated Cu-modified Pd electro- Pt nanoparticles (Pt/C) as the catalyst to convert CO2 to long
des also showed higher catalytic activities, producing HCOOH, carbon-chain hydrocarbons (>C5) at room temperature and
CH4, and CH3OH.186,187 It was found that CO2 electroreduction on ambient atmospheric pressure in a continuous flow cell. Feng
Pd electrodes could occur at potentials higher than the reversible et al.197 prepared a three-dimensional porous nanostructured
hydrogen potential, suggesting that adsorbed hydrogen atoms electrode, composed of nanoporous CuPt composites, for CO2
might take part in the slow stage of the electroreduction of HCO3 electroreduction in ionic liquid BMIMBF4.197 When Pt nano-
(CO2) to HCOO.188 Ohkawa et al.189 studied the CO2 electro- particles were supported on either calcia-stabilized zirconia
reduction reaction on a Pd electrode in a non-aqueous CH3CN (Pt/CSZ) or MnO2 (Pt/MnO2) to form a high-temperature CO2
solution and compared the results to those obtained in an reduction catalyst at 300–900 1C, up to 100% selectivity for
aqueous solution; they demonstrated that the concurrent paraformaldehyde (PFA) production was achieved.198 More recently,
desorption of hydrogen could lead to enhanced catalytic activity Pt/C–TiO2 and Pt-Pd/C–TiO2 based nanocomposite cathodes
for CO2 electroreduction. were employed to catalyze the electrochemical reduction of
CO2 to CH4 and isopropanol.199 In addition, electrochemical
Platinum. CO2 electroreduction on a Pt electrode surface was conversion of methanol and CO2 to dimethyl carbonate (DMC)
studied early on by Eggins and McNeill,190 in water, dimethyl was realized at a graphite-Pt electrode in a dialkylimidazolium
sulfoxide (DMSO), CH3CN, and propylene carbonate solutions, ionic liquid (1-benzyl-3-methylimidazolium chloride) methanol
respectively. By applying differential electrochemical mass spectro- system without any other additives.200
metry (DEMS), Brisard et al.191 investigated the mechanism of CO2
electroreduction catalyzed by polycrystalline Pt in acidic media. The Single-crystal surfaces of Pt group metals. The surface struc-
results showed that the main product was methanol. tures of Pt single-crystal electrodes have long been found to have a
A gas-diffusion electrode with Pt electrocatalyst was also significant influence on catalytic activity for CO2 reduction.201,202 A
employed for CO2 reduction under high pressure (o50 atm); difference was observed between Pt(111) and Pt(110) single-crystal
a Faradaic efficiency of 46% was obtained at a current density electrodes when researchers investigated the dynamic process of
of 900 mA cm2, and CH4 and CH3CH2OH were found to be the adsorbed CO formation from CO2 and adsorbed hydrogen; the rate
major products.192 The Faradaic efficiency for CH4 formation of CO formation on Pt(110) was more than 10 times higher than

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 647
View Article Online

Review Article Chem Soc Rev

on Pt(111).202 Hoshi et al.203 confirmed the order of activity for film on a Si(111) wafer was also used as the electrode for the
CO2 reduction: Pt(S)-[n(111)  (111)] > Pt(S)-[n(100)  (111)] > electrochemical reduction of CO2.210 The main reduction pro-
Pt(S)-[n(111)  (100)] (where S represents a single-crystal electrode) ducts obtained in acidic and neutral media were HCOOH and
in 0.1 M HClO4. They further found that CO2 reduction rates on CH3OH, with efficiencies of 40 and 7.7%, respectively. Qu et al.46
Pt(S)-[n(110)–(100)] electrodes (n = 2, 9) were higher than those on loaded RuO2–TiO2 nanotubes and RuO2–TiO2 nanoparticles,
Pt(110), and the rates on Pt(S)-[n(100)  (110)] (n = 2, 3, 9) were respectively, onto Pt electrodes for CO2 electroreduction. Com-
over twice as high as on Pt(S)-[n(100)  (111)].204 In addition, pared with electrodes coated with RuO2 or with RuO2–TiO2
Pt(210) (= Pt(S)-[2(100)  (110)]) yielded the highest rate of CO2 nanoparticles, the electrodes coated with RuO2–TiO2 nanotubes
reduction. This remarkably high activity might derive from the kink had a higher electrocatalytic activity for the conversion of CO2 to
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

site characteristics of Pt(S)-[n(110)  (100)] and Pt(S)-[n(100)  CH3OH, with a current efficiency of up to 60.5%, suggesting that
(110)]. The order of activity for the stepped surfaces was Pt(110) > the nanotube structure might be important in achieving high
Pt(S)-[n(111)  (111)] > Pt(S)-[n(111)  (100)] > Pt(S)-[n(111)  efficiency and selectivity for CO2 electroreduction.
(100)] > Pt(100) > Pt(111).205 The most active site in the stepped
2.1.5.3 Pt group metal complex catalysts
surfaces was derived from the pseudo-4-fold bridged site in
Pt group metal complexes with tetradentate ligands. Becker
Pt(S)-[n(111)  (111)]. Kinked stepped surfaces showed a higher
et al.96 explored several Pd-tetradentate complexes, including
activity than unkinked ones. The electrocatalytic activity of Pt
PdII–porphyrins (PdII–TPP) and PdII-2,3,7,8,12,13,17,18-octaethyl-
single-crystal electrodes towards CO2 reduction decreased in
porphyrin (PdII–OEP) complexes, as electrocatalysts for CO2
the order Pt(210) > Pt(310) > Pt(510).206
reduction. CV results showed that only PdII–TPP and PdII–OEP
More recently, when evaluating 12 Bi–, Se–, Te–, and Sb–Pt(hkl)
displayed electrocatalytic activity towards CO2 reduction, pro-
electrodes for the electrocatalysis of CO2 reduction in both acid and
ducing oxalic acid. Sende et al.57 found that the reduction
neutral aqueous media, Sanchez-Sanchez et al.207 found that only
potential of CO2 at electrodes modified with electropolymerized
Bi–Pt(111), Te–Pt(111), and Sb–Pt(100) electrodes showed a visible
films of [Fe(4-v-tpy)2]2+, [Ru(4-v-tpy)2]2+, and [Os(4-v-tpy)2]2+
current increase in the presence of CO2, whereas Se–Pt(100) and
tended to become more negative from Fe to Os (1.10, 1.20,
Te–Pt(100) had lower catalytic activities than the corresponding
and 1.22 V, respectively), which is consistent with their positions
unmodified Pt(100) electrode. The surface structure effects of Rh, Pd,
in the periodic table (the first, second, and third rows).
and Ir single-crystal electrodes were also examined, as shown in
Table 6. It was found that the rate of CO2 electroreduction on a Pd Pt group metal complexes of polypyridine. Bolinger et al.211
single-crystal electrode was strongly dependent on the crystal orien- studied CO2 electroreduction using [Ru(tpy)(dppene)Cl]+ (dppene =
tation. The rate of CO2 reduction at 0.5 V vs. RHE on Pd(110) was cis-1,2-bis(diphenylphosphino)ethylene) or cis-[Rh(bpy)2(TFMS)2]+
two orders of magnitude higher than on Pt(110).208 (TFMS = trifluoromethanesulphonate anion) as the target cata-
lysts. Two Ru complexes—[Ru(tpy)(bpy)(S)]2+ (S = solvent) and
2.1.5.2 Pt group metal oxide (mixture) catalysts. In an early
Ru(tpy)(Mebin-py)(S)2+ (Mebim-py = 3-methyl-1-pyridylbenz-
study on the electroreduction of CO2 on various conductive oxide
imidazol-2-ylidene)—and their catalytic activities in CO2 electro-
mixtures (RuO2, TiO2, MOO2, Co3O4, and Rh2O3), two metal oxide
reduction were recently reported.212 In addition, Rh(III) complexes
electrodes—i.e., RuO2 (35, mole percentage) + TiO2(65) and RuO2(20) +
of tptz have also been explored and showed effective catalytic
Co3O4(10) + SnO2(8) + TiO2(62)—showed high current efficiencies for
properties in CO2 electroreduction.213
methanol production when the electrode potential was controlled
Regarding transition metal polyphosphine complexes as
near the equilibrium potential of hydrogen evolution in a solution of
electrocatalysts for CO2 reduction, Slater and Wagenknecht214
0.2 M Na2SO4 (pH 4) saturated with CO2.42 Later, Bandi and Kühne209
investigated Rh(diphos)2Cl (diphos = 1,2-bis((diphenyl-
investigated the electrocatalytic activities of mixed Ru/Ti oxide electro-
phosphino)ethane2) in CH3CN. DuBois and Miedaner215 also
des (titanium sheets); their results indicated that the overpotential for
observed the catalytic activity of M(PhP(CH2CH2PPh2)2)L(BF4)2
H2 evolution increased with increasing TiO2 content.
(M = Pd, Pt, Ni) for CO2 reduction to CO in acidic CH3CN
In a comparison study of CO2 electroreduction in 0.5 M
solutions. Pd complexes (L = CH3CN, P(OMe)3, PEt3, P(CH2OH)3,
NaHCO3 solution, three electrodes—Ru, Cu–Cd-modified Ru,
or PPh3) exhibited significant CO2 catalytic activity, while Pt
and Cu–Cd-modified RuOX + IrOX—were used for electrolysis
complexes (L = PEt3) and Ni complexes (L = P(OMe)3 and PEt3)
for 8 hours while the potential was held at 0.8 V vs. SCE. Both
did not. In another report, DuBois et al.216 confirmed the electro-
methanol and acetone were produced.177 A RuO2-coated diamond
catalytic activity of [Pd(tridentate)(CH3CN)](BF4)2 complexes in
acidic DMF or CH3CN solutions; they found that if one or more
Table 6 Orders of catalytic activity of CO2 electroreduction catalyzed by of the phosphorus atoms of the tridentate ligand were substituted
single-crystal electrode surfaces of Pt group metals with a nitrogen or sulfur heteroatom, the resulting complexes
would not show significant catalytic activity towards CO2
Rh Rh(110) > Rh(100) > Rh(111) (0.1 M HClO4)272
Rh(100) > Rh(110) > Rh(111) (0.5 M H2SO4)272 reduction. By comparing the rate constants of catalysts with
Pd Pd(110) > Pd(111) > Pd(100) (0.1 M HClO4)208 different alkyl and aryl substituents on the terminal phosphorus
Ir Ir(110) c Ir(100) = Ir(111) (no reactivity)274 atoms, they found that the reaction rate of Pd(I) intermediates with
Pt Pt(S)-[n(110)–(100)] > Pt(110) > Pt(S)-[n(111)  (111)] >
Pt(S)-[n(100)  (111)] > Pt(S)-[n(111)  (100)] > Pt(100) > CO2 was increased by increasing the electron-donating ability of the
Pt(111);202,203,205 Pt(210) > Pt(310) > Pt(510)206 R groups, and that the steric interactions were of less importance.

648 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

Steffey et al.217 synthesized and characterized many types of Pd


complexes containing tridentate ligands with PXP (X = C, N, O,
S, As) donor sets—for example, the complex Pd(PCP)(PEt3)(BF4)
(where PCP is 2,6-bis((diphenylphosphino)methyl)phenyl)—and
also evaluated their catalytic activities towards CO2 electroreduc-
tion. They found that Pd(PCP)(CH3CN)(BF4) exhibited significant
catalytic currents in the presence of acid and CO2. The dependence
of the catalytic current on CO2 and acid concentrations was
consistent with the formation of a hydroxycarbonyl intermediate,
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

which decomposed in the presence of acid to form H2 as the catalytic


product. Raebiger et al.171 explored a bimetallic Pd complex with
bis(triphosphine) ligand (C6H4(P(CH2CH2P(C6H11)2)2)2) and found
that this complex could electrocatalyze the reduction of CO2 to CO
in acidic DMF solutions with a significantly higher TON than
that achieved with other previous monometallic, bimetallic,
and dendritic complexes of this catalyst class.
Pt group metal complexes with bidentate ligands. Polypyridyl Fig. 12 Schematic of the CO2 electroreduction mechanism catalyzed by
complexes of the second- and third-row transition metals have also the cis-[Ru(bpy)2(CO)H](PF6) complex. Reprinted with permission from
been reported to act as efficient homogeneous catalysts for the ref. 220. Copyright r 1991 American Chemical Society.
electrocatalytic reduction of CO2. Ishida et al.218,219 conducted
research on the controlled potential electrocatalysis of CO2
cis-(Cl)-Ru(bpy)(CO)2Cl2 and cis-(CO)-Ru(bpy)(CO)2(C(O)OMe)Cl
reduction in a saturated H2O–DMF (9 : 1, v/v) solution containing
complexes, catalyzed CO2 electroreduction could lead to a
Ru(bpy)2(CO)22+ or Ru(bpy)2(CO)Cl+ at an electrode potential of
Ru–Ru dimer when a chloride ion was lost, while the trans(Cl)-
1.5 V vs. SCE, and observed different product selectivity at different
Ru(bpy)-(CO)2Cl2 complex could form a polymeric film of
pH values. At pH 6.0, both CO and H2 were the main products,
[Ru(bpy)(CO)2]n.222 This suggested that a choice could be made
whereas at pH 9.5, HCOO was produced. They considered
between homogeneous and heterogeneous systems using the same
[Ru(byp)2(CO)COO]+, i.e., [Ru(byp)2(CO)COO]0, to be an important
experimental setup simply by changing the stereochemistry of
intermediate formed through the following reactions:
the precursor. Collombdunandsauthier et al.223 also confirmed
[Ru(bpy)2(CO)2]2+ + OH - [Ru(byp)2(CO)COOH]+ (30) the catalytic activity of [Ru(II)(bpy)(CO)2]n polymeric thin films.
The (mononuclear) [(bpy)2Ru(dmbbbpy)](PF6)2 and (dinuclear)
[Ru(byp)2(CO)COOH]+ + OH - [Ru(byp)2(CO)COO]+ + H2O
[(bpy)2Ru(dmbbbpy)Ru(bpy)2](PF6)4 (dmbbbpy = 2,20 -bis(1-methyl-
(31)
benzimidazol-2-yl)-4,4 0 -bipyridine) was synthesized for CO2
Moreover, the authors also found that the percentages of electroreduction by Ali et al.224 In CO2-saturated CH3CN, CO2
HCOO, CO, and H2 produced during the reduction were reduction catalyzed by [(bpy)2Ru(dmbbbpy)](PF6)2 yielded a
largely dependent on the pKa.218 For example, the current current efficiency of 89% for HCOOH in the presence of H2O
efficiency of HCOO formation was increased by increasing the (B2.5%) and 64% for C2O42 in the absence of H2O; in a similar
pKa value and reached 84.3% in the presence of Me2NHHCl. experiment, the current efficiencies for [(bpy)2Ru(dmbbbpy)-
Pugh et al.220 studied the catalytic activity of the complex cis- Ru(bpy)2](PF6)4 were 90% and 70%, respectively.
[Ru(bpy)2(CO)H](PF6) (i.e., cis-[Ru(bpy)2(CO)H]+) and, by means Tanaka and Mizukawa225 observed the highly selective formation
of FTIR spectroscopy, observed the cis-[Ru(bpy),(CO)H]+, cis- of ketones (current efficiency of 20%) in the electrocatalytic reduc-
[Ru(bpy)2(CO)(OC(O)H)]+, and cis-[Ru(bpy)2(CO)(NCCH3)]2+ spe- tion of CO2 by Ru(bpy)(napy)2(CO)22+ (napy = 1,8-naphthyridine-KN)
cies in the solutions at the end of the electrolysis period. They in the presence of (CH3)4NBF4. Only CH3C(O)CH3 and CO32
suggested a different mechanism of HCOO formation that were formed, according to the following reaction:
included the formation of cis-[Ru(bpy)2(CO)H]0 and the inser-
tion of CO2, as shown in Fig. 12.
Chardonnoblat et al.221 prepared a ‘‘[RuII(bpy)(CO)2]n’’ polymeric
film electrode by the electrochemical reduction of a mono- (32)
(bipyridine) complex, such as RuII(bpy)(CO)2(Cl)2 or RuII(bpy)-
(CO)2(CH3CN)2, for CO2 electrocatalytic reduction. Exhaustive For structural and spectroscopic characterization, Tanaka
electrolysis at 1.55 V vs. SCE produced CO with a current et al.226 employed Ru(II) complexes [Ru(bpy)2(CO)L] (L = CH3,
efficiency of 97% but yielded only 3% current efficiency for C(O)H, and C(O)CH3) as the model catalysts for a multi-step
formate production. By comparing the electrochemical behaviors CO2 reduction study using IR, Raman, 13C-NMR, and single-
of isomers of Ru(bpy)(CO)2Cl2 (trans(Cl)-Ru(bpy)-(CO)2Cl2 and crystal X-ray crystallography. To probe the mechanism, they
cis-(Cl)-Ru(bpy)(CO)2Cl2) and the behaviors of cis-(CO)- further prepared a series of [Ru(bpy)2(CO)L]n+ (L = CO2,
Ru(bpy)(CO)2(C(O)OMe)Cl complexes, they found that for the C(O)OH, CO, CHO, CH2OH, CH3, and C(O)CH3; n = 0–2) and

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 649
View Article Online

Review Article Chem Soc Rev

determined their molecular structures by X-ray analyses.227 HCOOH yields of 10–20%, while the current efficiency for hydrogen
They did so because they thought that these complexes might evolution was 31–54%. RhCl(CO)(PPh3)2 and IrCl(CO)(PPh3)2 (PPh3 =
be the reaction intermediates in the multi-electron reduction of triphenylphosphine) showed different electrocatalytic behaviors in
CO2 in protic media. DMF solution. IrCl(CO)(PPh3)2 was found to be an efficient homo-
Not all ruthenium complexes of 2,2 0 -bipyridine have been geneous catalyst for CO2 electroreduction to CO and HCOOH.234
found to be electrocatalytically active toward CO2 reduction.215 In addition, the Pd–organophosphine dendrimer complex was
For example, Begum and Pickup228 experimentally confirmed that tested as a CO2 reduction catalyst.235
Ru(2,20 -bipyridine)2[2-(2-pyridyl)benzothiazole]2+ was a highly active
2.1.5.4 Other catalysts containing Pt group metals. Other
catalyst whereas 1-methylbenzimidazole analogue was not.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

complexes containing Pt group metals have also been explored as


Bolinger et al.229 investigated the electrocatalytic reduction
CO2 reduction electrocatalysts.236–240 A typical candidate was an Ir
of CO2 by 2,2 0 -bipyridine complexes of both Rh and Ir
complex, for example, Ir(Z5-C5Me5))2(Ir(Z4-C5Me5)CH2CN)(m3-S)2,
(cis-[M(bpy)2Cl2]+ or cis-[M(bpy)2(TFMS)2]+, where TFMS is a
as reported by Tanaka et al.237 They used it to catalyze CO2
trifluoromethanesulfonate anion). It was observed that cis-
reduction to produce oxalate in CO2-saturated CH3CN solution.
[Rh(bpy)2X2]+ (X = Cl or TFMS) gave a current efficiency as
Some water-stable iridium dihydride complexes supported by
high as 80% for CO2 electroreduction to formate at a potential
PCP-type pincer ligands in CH3CN–water mixtures were also
of 1.55 V vs. SCE in CH3CN solution. The proton source for
explored as catalysts for CO2 reduction.240
the formate formation was considered to be [(n-Bu)4N](PF4) via
2.1.6 Copper, silver, gold, zinc, cadmium, and mercury
the Hofmann degradation:
2.1.6.1 Cu/Ag/Au/Zn/Cd/Hg metal electrodes. Several studies
[Rh(bpy)2] + CO2 - [Rh(bpy)2CO2] (33) using metal electrodes for CO2 electroreduction in aqueous KHCO3
solution reported that CH4 was predominantly the product at
[Rh(bpy)2CO2] + (n-Bu)4N+
a Cu cathode, CO at Ag and Au cathodes, and HCOO at a Cd
- [Rh(bpy)2CO2H] + H3CCH2CHQCH2 + (n-Bu)3N (34) cathode,38,241 as reviewed by Jitaru et al.242 and Gattrell et al.243
[Rh(bpy)2CO2H] - [Rh (bpy)2] + HCOO
I + 
(35) Cu electrodes. The metallic Cu electrodes thus far developed
can be classified into several types: bulk Cu electrode, Cu-
The electrocatalytic activity of cis-[Os(bpy)2(CO)H](PF4) in
electrodeposited GCE, in situ electrodeposited Cu electrode,241,244,245
the reduction of CO2 was observed by Bruce et al.230,231 They
and Cu-coated GDE.246 Aside from low hydrocarbons such as
found that under anhydrous conditions, CO was the dominant
CH4, C2H4, CO, HCOOH, alcohols (methanol, CH3CH2OH,
product. However, in the presence of water, 25% HCOO could
and CH3CH2CH2OH), and esters, some relatively high
be formed. In the reduction process, [Os(byp)2(CO)H] was
hydrocarbons—such as paraffins and olefins containing up to
considered an important intermediate that could be coordi-
6 carbon atoms—can also be formed using Cu electrodes.38,241,247–251
nated by CO2 to produce CO and HCOO.
The Faradaic efficiencies of these products were largely depen-
Nallas and Brewer170 explored a new family of catalysts for CO2
dent on temperature, type and concentration of electrolytes,
electroreduction, namely, two mixed-metal trimetallic complexes:
electrode potential, pH, crystal surface, and even the purity of
{[(bpy)2Ru(BL)]2IrCl2}5+ (BL = 2,3-bis(2-pyridyl)quinoxaline) and
the cathode material (Cu). For example, in the electrolysis of
2,3-bis(2-pyridyl)benzoquinoxaline. The two remote Ru centers
CO2-saturated 0.5 M KHCO3 aqueous solution with a 99.999%
served to tune the redox properties of the central catalytically
Cu sheet cathode, Hori et al.247 found that increasing the
active IrIII(BL)2Cl2 core. These catalysts should represent a new
temperature (0 to 40 1C) caused the Faradaic efficiency of
class of systems in which the redox properties of catalytic sites
CH4 production to drop rapidly from 65% to nearly zero, while
can be altered through remote metal coordination and variation
that of C2H4 gradually increased by up to 20%. In addition, they
without changing the coordination environment of the catalytic
carried out experiments using a range of electrolyte strengths,
iridium site.
from 0.03 to 1.5 M.249 Using voltammetric, coulometric, and
Some Pt group metal complexes with bidentate ligands,
chronopotentiometric measurements, they observed that CO
such as Ir complexes, have also been explored for CO2 electro-
was predominantly formed at potentials more positive than
chemical reduction. For example, two [Ir2(dimen)4]Y2 com-
1.2 V vs. NHE, while hydrocarbons (e.g., CH4 and C2H4) and
plexes (dimen = 1,8-diisocyanomenthane; Y = (PF6) and
alcohols (e.g., CH3CH2OH and CH3CH2CH2OH) were produced
[B(C6H5)2]) were studied using infrared spectroelectrochemistry;
in greater abundance below 1.3 V, where the Faradaic effici-
the results indicated that [Ir2(dimen)4]2+ first accepted two elec-
ency of CO dropped. In fact, CO was a reaction intermediate
trons to form [Ir2(dimen)4]0, then reacted with CO2 and H2O to
that strongly adsorbed on the cathode, interfering with hydrogen
form two main products, formate and bicarbonate.232
formation. In KCI, K2SO4, KCIO4, and dilute HCO3 solutions,
Pt group metal complexes with monodentate ligands. Hossain C2H4 and alcohols were found to be the main products,
et al. reported the electrocatalytic reduction of CO2 on either a whereas CH4 was preferentially produced in relatively concen-
GCE or a Pt electrode using a series of PdL2Cl2 complexes (L = trated HCO3 and K2HPO4 solutions. In nonaqueous solutions,
substituted pyridine and pyrazole) as catalysts in CH3CN contain- the electrochemical reduction of CO2 on a 99.999% Cu wire in
ing 0.1 M tetraethylammonium perchlorate.233 These catalysts CH3OH at 20–25 1C and 40 atm primarily yielded CO when
(L = pyrazole, 4-methylpyridine, and 3-methylpyrazole) gave tetrabutylammonium (TBA) salts (TBABF4 and TBAClO4) were

650 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

particles in CH3OH and conducted the electrochemical reduction


of CO2 with Pb and Zn electrodes. Results showed that without
the addition of the copper particles, only HCOOH and CO could
be detected, but after their addition, hydrocarbons were formed.
The Faradaic efficiencies of CH4 and C2H4 rose gradually as the
amount of Cu particles in the solution was increased, while the
currency efficiencies of HCOOH and CO decreased. When
B-370 Cu (99.9% pure) was employed at the cathode instead
of high-purity Cu,257 the Faradaic efficiency of CH4 was found
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

to be lower. In Hori’s study,248 which used pure Cu (99.999%)


and different electrolytes—including KCl (0.1/0.5/1.5 M),
KHClO4 (0.1 M), K2SO4 (0.1 M), and K2HPO4 (0.1, 0.5 M)—the
Faradaic efficiency of C2H4 reached 38.2–48.1%, while that of CH4
was 11.5–17.0%. Kim et al.258 tried to enhance the formation rate of
CH4; the highest rates they obtained were 8  105 mol cm2 h1
Fig. 13 Current–potential characteristics of the electrochemical (22 1C, 17 mA cm2) and 1.1  105 mol cm2 h1 (0 1C,
reduction of CO2 in a CO2 + methanol medium with various concentra- 23 mA cm2) at 2.0 and 2.3 V vs. SCE, respectively, on Cu foil
tions of TBABF4 as the supporting electrolyte. Concentrations of TBABF4: in 0.5 M KCO3 (pH 7.6). They found that the formation rate of CH4
33 mM (dotted line); 66 mM (dashed line); 0.1 M (solid line). Reprinted with
was even higher if the electrode surface was prepared by cleaning it
permission from ref. 252. Copyright r 1995 Elsevier.
using HCl rather than HNO3 or by oxidation in air. Indeed,
electrode surface conditions can have a major effect on an electro-
used as supporting electrolytes.252 The formation of HCOOCH3 de’s catalytic activity in CO2 electroreduction. For example, some
became predominant when lithium salts (LiBF4, LiClO4, and significant performance differences were found between rough
NH4ClO4) were used. The current–potential curves for various and smooth electrodes as well as between thermally and non-
concentrations of TBABF4 showed a large cathodic current and thermally treated electrodes.259 Ohta et al.260 found that under
a shoulder wave (Fig. 13). Since the magnitude of the shoulder ultrasonic irradiation, the production rates of CH4, HCOOH, and
wave was dependent on the concentration of TBABF4, this wave CO were greatly affected.260 Cook et al.244,261 employed an in situ
was attributed to the reduction of TBA+, presumably to TBA : electrodeposited Cu electrode (glass carbon substrate with in situ
deposited Cu) in 0.5 M KHCO3 aqueous solution. Although the
NR4+ + e -  NR4 (36) main products were still CH4 and C2H4, Cu purity and morphology
were found to be crucial for promoting a high rate of CO2
CO2 +  NR4 -  CO2 + NR4+ (37)
reduction. They achieved a Faradaic efficiency of 71.3% for C2H4
At 2.3 to 4.0 V vs. Ag/AgCl, the electrochemical reduction of and CH4 production by employing a Cu-based GDE.246
CO2 at a 99.98% Cu electrode in a CH3OH solution containing Recently, a Cu nanoparticle-covered electrode was reported
80 mM LiOH supporting salt at about 30 1C, CH4, C2H4, CO, to give better selectivity towards hydrocarbons than the surfaces
and HCOOH were the main products.253 The best current of electropolished and argon-sputtered copper electrodes.262
efficiency for CH4 (the main product) was 63% at 4.0 V vs. The copper nanoparticles were formed in two steps. In the first
Ag/AgCl. When Kaneco et al.254 studied CO2 electroreduction at step, the potential at the electropolished copper electrode was
a 99.98% Cu electrode in a CH3OH solution with 80 mM CsOH scanned between 0.6 and +1.15 V vs. RHE at 20 mV s1 under
supporting salt at about 30 1C, CH4, C2H4, C2H6, CO, and N2-saturated KClO4. In the second step, the copper was redepo-
HCOOH were the products. The maximum Faradaic efficiency sited on the electrode surface; this was performed under CO2-
of C2H4 (the main product) was 32.3% at 3.5 V vs. Ag/AgCl. saturated KClO4 with a constant bias of 1.3 V vs. RHE for
The C2H4/CH4 current efficiency ratio was in the range 2.9–7.9. 20 minutes. Scanning tunneling microscopy (STM), scanning
It was thought that small cations, such as Li+ and Na+, might electron microscopy (SEM) (Fig. 14), and CV (Fig. 15) were
not easily adsorb on the electrode surface due to their strong employed to compare various treated electrodes. A few layers
hydration, while a less hydrated, bulky cation, such as Cs+, of Cu nanoparticles with sizes of 50–100 nm covered the Cu
might preferentially adsorb on the cathode, giving a less surface, creating a surface area 2–3 times greater than the geo-
hydrated electrode surface. On such a surface, the conversion metric surface area of the Cu electrode. However, CV measure-
of the intermediate CuQCH2 to CH2CH2 might occur more ments in CO2 (pH 6.0) showed that the current density of the
easily than on a more hydrated one, producing more CH2CH2 nanoparticle-covered surface (at 0.75 V vs. RHE) was 10 times
in the presence of Cs+. Under a CO2 pressure of 10 atm at about higher than that of the electropolished surface, indicating that
30 1C in CH3OH with 0.5 M CsOH supporting salt, when the surface morphology can contribute more to current density
potential was increased from 3.5 to 2.0 V, the Faradaic than just the effect of increased surface area. The morphological
efficiencies of CO and C2H4 dropped slowly from 84% to 40% effect was explained by the roughened surface having a greater
and 5% to 4%, respectively, but those of HCOOH, CH4, and H2 abundance of undercoordinated sites; this was demonstrated
slowly increased.255 Kaneco et al.256 suspended 1 mm copper by DFT calculations. In addition, electrochemical quartz crystal

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 651
View Article Online

Review Article Chem Soc Rev


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 14 Scanning electron microscopy images for three types of surfaces: (a) electropolished, (b) copper nanoparticle covered, and (c) sputtered.
Reprinted with permission from ref. 262. Copyright r 2012 The Royal Society of Chemistry.

microbalance (EQCM)—a high-resolution (Bnanograms) mass 3D cathode of 30# mesh tinned-copper was used as the cathode.
sensing technique—has also been used to probe electroformed With currents of 1–8 A, feeding gas phase CO2 concentrations
and electroreduced products on a copper electrode in aqueous of 16–100 vol%, and operating times of 10–180 minutes,
solutions containing NaHCO3 and Na2CO3.263 a current efficiency of 86% was achieved for HCOO. The
Li and Oloman264 investigated the electroreduction of CO2 in efficiency was dependent on current density and CO2 pressure.
a laboratory bench-scale continuous reactor in which a flow-by In a study of copper-catalyzed CO2 electroreduction, Kuhl
et al.265 reported an experimental methodology that allows for
product identification and quantification with unprecedented
sensitivity. Among all the possible products, CH4 and C2H4 had
the largest current efficiencies; the remaining products were
oxygenates and other C2 and C3 species. The researchers offered
some possible reaction pathways to account for the production
of all the C2 and C3 species observed (see Fig. 16).
Regarding the catalytic stability of CO2 electroreduction
on Cu electrodes, several other factors have been identified,
including CO adsorption,266 electrode purity,257 the formation
of carbon deposits,267 and the presence of other surface-
poisoning species.268,269 The carbon deposited film seems to
be a major factor in the irreversible degradation of the elec-
trode surface. For example, when the electrochemical reduction
of CO2 to both CH4 and C2H4 was conducted in aqueous 0.5 M
KHCO3 solution at a constant potential of 2.00 V vs. SCE,
a black film formed on the surface of the Cu (99.999%) cathode.267
XPS and AES studies indicated that this film was graphitic carbon
formed by CO2 reduction through HCOO. Graphitic carbon
deposit was also found to cause a decline in Cu electrodes’ catalytic
activity.270,271 To resolve this problem, the researchers developed a
new method that could selectively convert CO2 to C2H4 at the three-
phase (gas/liquid/solid) interface on a CuIBr confined Cu-mesh
electrode in an aqueous solution of KBr. The conversion per-
centages of CO2 and H2 were found to be about 90% and 2%,
respectively. It was suggested that the immobilized CuIBr, acting
as a heterogeneous catalyst, offered some adsorption sites
for reduction intermediates such as CO and carbene (H2C:).
The mechanism can be expressed as follows:

CO2(g) + 2H+ + 2e - CO(g) + H2O (38)

Fig. 15 Cyclic voltammograms (CVs) of the formation of copper nano- CuBr(s) + CO(g) - CuBr  CO (39)
particles in 0.1 M KClO4 purged with N2 at pH 10.5. (a) CVs of the
electropolished copper surface, copper nanoparticle covered surface, CuBr  CO + 4H+ + 4e - CuBr  :CH2 + H2O (40)
and sputtered copper surface in 0.1 M KClO4 purged with (b) CO2 (c)
and N2. The current density is normalized by the geometric area of the 2CuBr  :CH2 - CuBr  :CH2QCH2 + CuBr(s) (41)
electrode surface. The overpotentials are corrected for ohmic resistance
between the working and reference electrodes. Reprinted with permission Similar to Pt-group metals such as Rh,272 Pd,208,273 Ir,274
201
from ref. 262. Copyright r 2012 The Royal Society of Chemistry. Pt, and Ag,275 single-crystal copper electrode surfaces

652 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 16 Proposed reaction pathways for C2 and C3 products with enol-like surface intermediates. Arrows between overlapping circles indicate the
changes between the enol, keto, and diol forms of each product. Arrows between non-overlapping circles indicate the electrochemical reduction steps
involved in the addition of 2H+ and 2e. For simplification, product names are intended to refer to all forms of the product. Reprinted with permission
from ref. 265. Copyright r 2012 The Royal Society of Chemistry.

can normally enhance the electrocatalytic activity in CO2 elec- requirements was consistent with experimental measurements. Liu
troreduction by introducing steps and kinks into atomically flat et al.283 studied the electroreduction of CO2 to CO on Fe, Co, Ni,
surfaces. For example, Hori et al.276–278 investigated the electro- and Cu surfaces using a DFT method involving three reaction steps:
chemical reduction of CO2 at various types of copper single- adsorption of CO2, decomposition of CO2, and desorption of
crystal electrodes in 0.1 M KHCO3 aqueous solution and found CO. Both the binding energies and the reaction energies were
that the reaction selectivity could be greatly altered by changing calculated. They found that the reaction energies and the total
the crystal orientation. The major product with electrodes reaction energy barrier were strongly dependent on the type of
based on (100) terrace surfaces (i.e., Cu(S)-[n(100)  (111)] electrode metal.
and Cu(S)-[n(100)  (110)]) was C2H4. The formation of CH4
Ag and Au electrodes. Similar to Cu, both Ag and Au
was promoted at Cu(111) or by the introduction of (111) or
electrodes also show considerable catalytic activity towards CO2
(110) step atoms to the (100) basal plane. A Cu(S)-[n(111) 
electroreduction when appropriate electrolytes are employed.284
(111)] electrode yielded high amounts of C2+ substances
In 0.1 KHCO3 aqueous solution at 25 1C, the Faradaic efficien-
(i.e., substances containing more than two carbon atoms), while
cies for CO production from CO2 at Ag(99.98%) and Au(99.95%)
a (110) electrode derived from Cu(S)-[n(111)  (111)] uniquely
electrodes at 1.6 V vs. Ag/AgCl saturated with KCl were found to
produced high yields of CH3COOH, CH3CHO, and C2H5OH.
be 64.7% and 81.5%, respectively.38 A recent study251 reported
CO seems to be the key intermediate in CO2 electroreduction
that when Ag-coated nanoporous Cu composites (NPC) were
to CH4 and CH2CH2 on Cu electrodes.279,280 Schouten et al.280
employed in CO2 reduction in BMImBF4, the electrosynthesis
observed two reaction pathways: a C1 pathway leading to CH4
of dimethyl carbonate (DMC) could be realized. The highest yield
formation on single-crystal Cu(111) electrodes and a C2 path-
of DMC was 80%; this was attributed to the electrodes’ high
way leading to ethylene formation on Cu(100) electrodes.281
surface area, open porosity, and high efficiency.
The authors also proposed a mechanism based on reactions as
A sputtering deposition technique was used to prepare a Au
a function of potential, using online mass spectrometry com-
electrode for CO2 reduction.285 The results indicated that Ar pressure
bined with mechanisms suggested in the literature.
had an effect on the Au surface’s geometrical structure and surface
Regarding theoretical studies of CO2 electroreduction on Cu
area, resulting in different CO2 reduction potentials in both KCl and
electrodes, some researchers have carried out studies to further
KHCO3 solutions. A porous Au film electrode (a 200–260 nm Au film
our fundamental understanding of the effects of face-centered
deposited on a porous hydrophilic polymer membrane) was also
Cu facets—such as cubic fcc(111), fcc(100), and fcc(211)—on the
prepared by vapor deposition for the electrochemical reduction of
energetics of CO2 electroreduction. Durand et al.282 reported that
CO2 to CO in 99.99% KHCO3 aqueous solution.286 Recently, ligand-
the intermediates in CO2 reduction could be mostly stabilized by
protected Au25(SC2H4Ph)18 clusters were explored to promote the
the (211) facet, followed by fcc(100) and fcc(111). This implied that
catalytic electroreduction of CO2, and high catalytic activity in the
the (211) facet should be the most active surface in producing CH4,
conversion of CO2 to CO was achieved.287
as well as the by-products H2 and CO. HCOOH production might
be mildly enhanced on the more closely packed surfaces (i.e., (111) Zn, Cd, and Hg electrodes. Due to the high overpotentials
and (100)). Their theoretical prediction of the trends for voltage of hydrogen evolution on Zn, Cd, and Hg electrodes, they

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 653
View Article Online

Review Article Chem Soc Rev

were considered as suitable electrodes for CO2 reduction. prepared a CuO–Zn composite electrode by pressing a mixture of
In 0.1 M KHCO3 aqueous solution, the electrochemical reduction Zn particles (B7 mm) and CuO or Cu2O powder (B100 nm) for
of CO2 on Zn, Hg, and Cd electrodes was found to be very selective CO2 reduction.296 It was found that without the copper oxide
for the formation of HCOOH and CO, with HCOOH seemingly the particles, only HCOOH and CO were formed, whereas with the
only product at the Hg cathode. On metallic Zn, Cd, and Hg CuO–Zn composite electrode, hydrocarbons such as CH4 and
electrodes, the Faradaic efficiencies of HCOOH were measured to C2H4 could be obtained. The maximum formation efficiencies of
be 20%, 39%, and 94%, while those of CO were 39.6%, 14.4%, CH4 and C2H4 were 7.5% and 6.8%, respectively. Le et al.142
and undetectable, respectively.38 Shibata et al.288 employed a examined the catalytic activity of an electrode electrodeposited
Cd-loaded GDE to reduce CO2 and nitrite ions with various as a cuprous oxide thin film and found that the Faradaic
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

catalysts and found that the maximum current efficiency of efficiency for CH3OH production was 38%. They believed that
urea ((NH2)CO) formation was about 55% at 1.0 V vs. SCE. Cu(I) species should play a critical role in selectivity for CH3OH.
In an organic solution such as DMSO, CO2 electroreduction was More recently, Li and Kanan297 prepared Cu as Cu2O layers for
also studied using both Au and Hg electrodes.289,290 CO2 reduction. The Cu2O layers formed at 500 1C exhibited a
It should be mentioned that several decades ago, a mercury large surface roughness, resulting in the electrochemically active
pool electrode was considered the best electrode for CO2 surface area (ECSA) of a reduced electrode being 480 times larger
reduction.291 The results showed that in the neutral pH range, than that of a polycrystalline Cu electrode. This improved ECSA
all the current was consumed in the production of HCOOH, resulted in a 0.5 V lower overpotential for CO2 reduction than
while in acid solutions, both HCOOH and H2 were produced. on a polycrystalline Cu electrode. An interesting result was
The mechanism was proposed to be as follows: that the SEM showed a dense array of rods with diameters of
100–1000 nm on the electrode surface. These rods were the
CO2 + e - CO2(ads) (42)
outermost portion of a thick Cu2O layer coating the electrode.
CO2(ads) + H2O - HCO3(ads) + OH (43) However, these rods were not necessary for efficient CO2 reduction.
It was observed that the Cu particles formed by reducing mm-thick
HCO3(ads) + e - HCOO
 
(44)
(B3 mm) Cu2O films at the potentials at which CO2 reduction
H+ + e - Hads (45) occurred could catalyze the reduction of CO2 to CO and
HCOOH with high Faradaic efficiencies and exceptionally low
H + CO2 - HCO2
+
(46)
overpotentials (Fig. 17). At high overpotentials, this Cu particle
H + H + + e - H 2 (47) electrode could produce C2 hydrocarbons exclusively. To obtain

HCO2 + H + e - HCOOH
+ 
(48)

With a Hg electrode, the formation of malate (OOCCH-


(OH)CH2COO) in the process of CO2 reduction was also
observed in aqueous solutions containing quaternary ammo-
nium salts.292 Since the observed coulombic yield was more
than 100%, the overall reaction was believed to consist of not
only reaction (49) but also possibly (50):

(49)

(50)

However, with other supporting electrolytes, such as NaHCO3,


NaH2PO4–Na2HPO4, NaCl, NaClO4, Na2SO4, LiHCO4, and KHCO3
and their mixtures, the favorable product was found to be
HCOO, and the Faradaic efficiency of HCOOH formation
increased with increasing CO2 pressure.293,294 In 0.5 M KHCO3
aqueous solution, the Faradaic efficiency of HCOOH produc-
tion was as high as 100% at a CO2 pressure of 20 atm.294

2.1.6.2 Cu and Au oxide-related catalysts


Cu oxides. Cu oxides have been explored to catalyze
Fig. 17 SEM images (a) before and (b) after CO2 reduction electrocatalysis
CO2 electroreduction. For example, Chang et al.295 deposited
at 0.5 V vs. RHE; (c) Faradaic efficiencies for CO and HCO2H vs. potential
as-prepared Cu2O particles onto a carbon cloth electrode for on polycrystalline Cu and Cu annealed at 500 1C for 12 hours. Reprinted
CO2 reduction and carried out CV measurements; the results with permission from ref. 297. Copyright r 2012 American Chemical
showed that CH3OH was the predominant product. Ohya et al.296 Society.

654 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

insight into the mechanistic pathway(s) for CO2 reduction, Tafel CO2-saturated 0.01 M TBATFB–DMF solution (TBATFB = tetra-
plots were analyzed. The plot for annealed Cu was found to be linear butylammonium tetrafluoroborate) by Kumar et al.305 It was
over the range of overpotentials from 0.05 to 0.3 V with a slope of 116 observed that this catalyst has high selectivity towards CO2
mV dec1. This slope was consistent with a rate-determining reduction to oxalic acid, and the highly active site was believed
initial electron transfer to CO2 to form a surface-adsorbed to be a Cu(I) species (Fig. 18). Cu-based Perovskite-type
CO2* intermediate, which suggested that the Cu surfaces A1.8A 0 0.2CuO4 (A = La, Pr, and Gd; A 0 = Sr and Th) were also
formed by reducing thick Cu2O layers could enable the for- explored for catalyzing CO2 electroreduction.306 For example,
mation of this CO2* intermediate while suppressing H2O when these kinds of catalysts were incorporated into a GDE, the
reduction. Regarding the understanding of fundamentals, Wu cumulative Faradaic efficiencies of CH3OH, CH3CH2OH, and
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

et al.298 investigated the adsorption of CO2, H2CO3, HCO3, and CH3CH2CH2OH reached 40% in La1.8Sr0.2CuO4 GDE/0.5 M KOH
CO32 on a Cu2O(111) surface by first-principles calculations aqueous solution under ambient conditions.
based on DFT at the B3LYP hybrid functional level, in which the
Cu2O(111) surface was modeled using an embedded cluster Au oxide catalysts. Chen et al.307 reduced Au oxide films to
method. It was concluded that on the surface, H2CO3 was Au nanoparticles on electrodes for CO2 reduction. High selec-
dissociated into an H+ ion and an HCO3 ion, which was the tivity from CO2 to CO in water at overpotentials as low as
only activated CO2 species on the surface. 140 mV was observed. The high catalytic activity was thought to
be due to the dramatically increased stabilization of the CO2*
Cu organic frameworks and Cu-based Perovskite-type catalysts. intermediate on the surfaces of the oxide-derived Au electrodes.
Metal organic frameworks with crystalline ordered structures, The proposed mechanisms are shown in Fig. 19.
extra-high porosity, high thermal stability, as well as adjustable
chemical functionality have been pursued for many purposes, 2.1.6.3 Cu/Ag/Au/Zn/Cd/Hg alloy catalysts. In CO2 electro-
including gas-storage applications.299–304 Recently, Cu3(BTC)2 reduction, metal alloys such as Cu alloys can exhibit both high
(BTC = 1,3,5-benzenetricarboxylate), a Cu-based metal organic electrocatalytic activity and product selectivity, and they yield
framework, was explored as an electrode for CO2 reduction in a products quite different from those produced on a pure Cu

Fig. 18 Cyclic voltammograms for (a) Cu3(BTC)2-coated GC (the red line represents GC background); (b) CV 1 = bare GC, 2 = bare GC in the presence of
CO2, 3 = Cu3(BTC)2-coated GC, and 4 = Cu3(BTC)2-coated GC in the presence of CO2, in a solution containing 0.01 M TBATFB–DMF, at a scan rate of 50
mV s1; (c) FTIR spectra of (I) oxalic acid (authentic) and (II) oxalic acid (synthesized); (d) GC–MS spectrum of the bulk electrolysis product of CO2 at the
Cu3(BTC)2 coated electrode surface. Reprinted with permission from ref. 386. Copyright r 2012 Elsevier.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 655
View Article Online

Review Article Chem Soc Rev

In organic solutions, for example, a methanol with 0.1%


KHCO3 and 0.05 M Bu4NBF4 solution, the Cu-2,9,16,23-tetra-
tert-butyl-phthalocyanine (CuTBPC) complex coated on a graphite
rod catalyzed CO2 reduction to CH4 with maximum current
efficiencies of 29–35%, while an ‘‘electropolymerized’’ tetraamino-
substituted Cu-monophthalocyanine-coated electrode gave current
efficiencies as high as 75–85%.312 However, when a Cu-2,9,16,23-
tetraaminophthalocyanine (CuTAPC) complex-coated electrode
was used, the main product was found to be HCOOCH3, with
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 19 Proposed mechanisms for CO2 reduction to CO on polycrystal- small amounts of CO and CH4. The formation pathway for
line Au and oxide-derived Au. Reprinted with permission from ref. 307. HCOOCH3 was assumed to be CO2 - CO2 - HCOOH -
Copyright r 2012 American Chemical Society. HCOOCH3. In addition, Cu(II) complexes formed with other
tetradentate ligands, such as dinaphthotetraaza[14]annulene and
5,5 0 -terpyridinophane macrocycles, have also been explored as
electrode, where CH4 and C2H4 formation predominates. When
catalysts for CO2 electroreduction.313–315
using Cu alloy electrodes, HCOOH, CO, and CH3OH were found
to be the products at less negative potentials and almost at the 2.1.6.5 Other catalysts. Dinuclear Cu complexes have also
reversible potentials of their formation.308,309 In 0.05 M KHCO3 shown activity towards CO2 electroreduction.168,169,172,316 For
aqueous solution, Cu–Ni alloys could produce CH3OH and example, Field et al.168,169 synthesized one dinuclear Cu complex,
HCOOH at onset potentials of 0.38 V and 0.5 V vs. SHE, res- Cu2(m-PPh2bipy)2(CH3CN)2(PF6)2 (PPh2bipy = 6-diphenyl-
pectively, which were quite a bit lower than the 0.90 V required phosphino-2,2 0 -bipyridyl), by treating Cu(CH3CN)4PF6 with
for CO production. Cu–Sn and Cu–Pb produced HCOOH and CO 6-diphenylphosphino-2,2 0 -bipyridine in CH3CN solution, and
with an enhanced reaction rate at their reversible potentials. used it as the catalyst for homogeneously catalyzing the electro-
Both Cu single-crystal and Cu–Au alloy (CunAu100n, n = 99, 90, chemical reduction of CO2 to selectively produce CO and CO32
80, 50, respectively) electrodes were also explored for CO2 in 0.1 M tetra-N-butylammonium perchlorate–CH3CN solution.
electroreduction in aqueous KH2PO4–K2HPO4 buffer solution, Two sequential single-electron transfers to [Cu2(m-PPh2bipy)2-
and H2, CO, CH4, C2H4, and trace C2H6 were found to be the (CH3CN)2]2+ were observed at E1/2(2+/+) = 1.35 V and E1/2(+/0) =
products.310 For the Cu single-crystal electrode, the fraction of 1.53 V vs. SCE, respectively. Thus, two possible routes were
CH4 in the product mixture was increased while that of CO was proposed, as follows:
decreased, in the order Cu(poly) o Cu(100) o Cu(111). For the
Cu–Au alloy electrode, the fraction of CO production increased [Cu2(m-PPh2bipy)2(CH3CN)2]2+ + e
markedly with increasing Au content, while the fraction of CH4 - [Cu2(m-PPh2bipy)2(CH3CN)2]+ (51)
gradually diminished. Amongst all the examined electrodes, the
Au50Cu50 alloy appeared to be the most efficient catalyst for the [Cu2(m-PPh2bipy)2(CH3CN)2]+ + e
conversion of CO2 into carbon-containing gaseous products. - [Cu2(m-PPh2bipy)2(CH3CN)2]0 (52)
Schizodimou and Kyriacou311 employed a Cu88Sn6Pb6 alloy
cathode for CO2 reduction to investigate the effects of the [Cu2(m-PPh2bipy)2(CH3CN)2]0 + CO2 - product (53)
supporting electrolyte and cathode potential on the reduction Indeed, the PPh2bipy ligand offers the dual advantage
rate; the results showed that the fractions of H2, CO, CH3OH, of coordinated bipyridine and bridging phosphines. The
HCOOH, CH4, CH3CHO, and C2H6 in the production mixture p*-unsaturation of the bipy component of the PPh2bipy ligand
changed with the type of electrolyte and the cathode potential. could provide the ability to shuttle electrons in and out of a
closed-shell d10–d10 binuclear complex. Kauffman et al.287 recently
2.1.6.4 Cu/Ag/Au/Zn/Cd/Hg complex catalysts. The Cu/Ag/Au/ electrocatalyzed the reduction of CO2 to CO using ligand-protected
Zn/Cd/Hg complexes that so far have been explored for cata- [Au25(SC2H4Ph)18] clusters in aqueous solution at 1.0 V vs.
lyzing CO2 electroreduction are those formed by metal cations RHE. The efficiency was approximately 100%, while the rate was
with tetradentate ligands. In an early study, two Ag(II) porphyrin 7–700 times higher than for larger Au catalysts and 10–100 times
complexes, i.e., Ag–2,3,7,8,12,13,17,18-OEP and Ag–TPP, were higher than for current state-of-the-art processes.
employed to homogeneously catalyze the reduction of CO2 at a
glassy carbon cathode in CH2Cl2 + 0.1 M tetrabutylammonium 2.2 Aluminum, gallium, indium, and thallium
fluoroborate (TBAF) solution. It was found that only AgII(OEP) 2.2.1 Al metal electrodes and Al-containing catalysts. Al has
displayed electrocatalytic activity, and oxalic acid was the main been tried for the catalysis of CO2 electroreduction, but unfortu-
product.96 Recently, Cu- and Zn-meso-TPP supported GDEs were nately the catalytic activities have been extremely low.38,290 For
fabricated for CO2 reduction in 0.5 M KHCO3 aqueous solution by example, it was reported that in the electrochemical reduction of
Sonoyama et al.;97 their results showed that the current efficiencies CO2 by a metallic Al electrode at 1.6 V vs. Ag/AgCl in 0.1 M
for CO and HCOOH formation could be increased and H2 evolution KHCO3 aqueous solution, the Faradaic efficiencies for CH4, C2H4,
depressed by increasing the CO2 pressure from 1 to 20 atm. and C2H6 production were as low as 0.58%, 0.04%, and 0.11%,

656 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

respectively, versus 99% for H2 evolution.38 In another experiment, Weaver326 tried to fix CO2 by combining a p-InP electrode with a
at a constant potential of 2.2 V vs. SCE in a CO2-saturated 0.05 M biological catalyst (a formate dehydrogenase enzyme); they observed
KHCO3 aqueous solution, the sum of the typical current efficien- a 2e reduction of CO2 to HCOOH. Kaneco et al.327 carried out a
cies of CH4, C2H4, C2H6, and HCOOH was less than 1%, com- series of experiments on the photoelectrochemical reduction
pared with 95.7% for H2.290 of CO2 at a p-InP electrode in a methanol-based electrolyte
2.2.2 Ga electrodes and Ga-containing catalysts. In a photo- (nonaqueous media). In CO2-saturated 80 mM LiOH–methanol
electrochemical cell, p-type Ga-containing semiconductors—i.e., solution, they observed the maximum current efficiencies to be
p-gallium phosphide (p-GaP),317,318 p-gallium arsenide (p-GaAs),319 B40% for CO and B30% for HCOOH in the potential range of
and n-gallium arsenide (n-GaAs)318—were explored as catalysts; 2.2 to 2.5 V vs. Ag/AgCl.327 They also found that metal-
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

high selectivity for CH3OH formation (a 6e reaction) was observed modified InP electrodes could give different product selectivity.328
but only at some exceptionally high overpotentials. Recently, For example, on Pb-, Ag-, Au-, and Cu-modified InP electrodes,
Bocarsly et al.320 reported the highly selective reduction of CO2 the main reduction products of CO2 were CO and HCOOH
to CH3OH at illuminated p-GaP photoelectrodes, with near (the maximum current efficiency of CO was 80.4% on Ag-InP);
100% Faradaic efficiency for underpotentials greater than in comparison, a Pd-modified electrode yielded only CO, and
300 mV at 0.52 V vs. SCE. a Ni-modified electrode produced hydrocarbons with low
2.2.3 In metal electrodes and In-containing catalysts. Faradaic efficiencies (0.7% for CH4 and 0.2% for CH2CH2).
Normally, CO2 electroreduction at an indium metal electrode Recently, they tried to catalyze CO2 reduction by suspending
in aqueous media predominantly produces formate,241,321,322 Cu particles (1 mm diameter) in 0.10 M NaOH–methanol
but in nonaqueous solutions, the main product is CO.322 solution.329 After the addition of Cu particles, the current effi-
Kapusta and Hackermant321 reported CO2 reduction to formate ciencies for methane and ethylene improved.
ions in a 0.5 M HCOOH + 0.5 M HCOONa solution with high 2.2.4 Tl metal electrodes and Tl-containing catalysts. Tl metal
current efficiency (about 95%), although the overall power electrodes in aqueous electrolytes have been found to favor the
efficiency was low due to the high overpotential of the reaction. formation of formic acid, while in nonaqueous solutions, oxalic
Ikeda et al.322 also observed the electrochemical reduction of acid has been the dominant product.242,322 Unfortunately, the
CO2 in 0.1 M tetraethylammonium perchlorate (TEAP) aqueous literature contains only these two cited studies of Tl metal
electrolytes at potentials of 1.8–2.4 V vs. Ag/AgCl; the current electrodes as cathode catalysts for CO2 electroreduction.
efficiencies for HCOOH formation were in the range of 80–90%.322
However, in 0.1 M TEAP/propylene carbonate (nonaqueous 2.3 Tin and lead
electrolytes) at 2.0 to 2.2 V vs. Ag/AgCl, the current efficien- 2.3.1 Sn metal electrodes and Sn-containing catalysts. Sn
cies were up to 95%.323 Todoroki et al.294 achieved a Faradaic metal electrodes were reported to be most active towards CO2
efficiency of 100% under 60 atm of CO2. They found that the electroreduction in aqueous electrolytes, producing HCOO.
efficiency of HCOOH formation became higher at high cathodic However, in nonaqueous electrolytes, the predominant product
current densities (or more negative potentials), and that at less was CO, with small amounts of formic acid, oxalic acid, and
negative potentials, HCOOH formation could be suppressed glyoxalic acid.242,322 Some early studies indicated that Sn metal
and CO formation became relatively predominant. Mizuno working electrodes could catalyze CO2 electrochemical reduction
et al.324 also found that the Faradaic efficiency for HCOOH in aqueous inorganic salt solutions to exclusively produce
was about 100% at 20–60 1C, compared with 44.5% at 100 1C. HCOOH with a current efficiency as high as B95%.241,321
Recently, Narayanan et al.325 studied the conversion of CO2 to However, during the reduction reaction, organometallic com-
formate in an alkaline polymer electrolyte membrane cell in plexes formed on the electrode surface, accelerating the rate of
which In powder-coated porous carbon paper was the cathode. hydrogen evolution and leading to poor reaction efficiency.321
Three different aqueous solutions (CO2-saturated deionized Increasing the temperature also can cause a decrease in Faradaic
H2O, 1 M NaHCO3, and 1 M Na2CO3) were used as the cathode efficiency for formic acid production and an increase for
feed. The instantaneous Faradaic efficiency of formate production hydrogen.324 Li and Oloman330,331 developed a scale-up reactor
achieved in NaHCO3 solution was as high as 80% (although it system for CO2 electroreduction, in which a granulated tin
decreased to B10% over a period of 1.0 hour). Carbon mass cathode (99.9 wt% Sn) and a feed gas of 100% CO2 were used.
transport was found to be the limiting factor for Faradaic efficiency. The results showed that the granulated tin cathode yielded
However, this mass transport limitation could be mitigated using better performance than the tinned-copper mesh cathode
a high bicarbonate concentration or high carbon dioxide reported in their previous communications, in terms of both
pressure. In their experiments, high Faradaic efficiency could current efficiency and stability.264,330 The formate current effi-
be maintained during continuous operation, even at moderately ciencies were up to 91%. When tin was electrodeposited on a
high current densities. GDE in a zero-gap cell for CO2 electroreduction, the electrode
An In-containing semiconductor (p-type indium phosphide, showed good stability.332 Chen and Kanan333 prepared a thin-
p-InP) electrode was used in the electrolysis of CO2-saturated film catalyst by simultaneously electrodepositing Sn0 and SnOx
Na2SO4 aqueous solutions; the photocurrent densities for onto a Ti electrode. Using an H-type cell reactor and CO2-
CH3OH formation were found to be 60–100 mA cm2, and the saturated aqueous NaHCO3 solution, the Sn0/SnOx catalyst
current efficiencies were found to be 40–80%.319 Parkinson and exhibited up to eight-fold higher partial current density and

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 657
View Article Online

Review Article Chem Soc Rev


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 21 Tafel plots of the products by electrochemical reduction of CO2


at a Pb electrode in methanol at 15 1C. HCOOH, |; CO, J; H2, K.
Fig. 20 Tafel plots for SnB, SnG, and SnGDL, obtained from the corres- Catholyte/anolyte: 300 mM dm3 KOH in methanol. Reprinted with
ponding voltammograms. Reprinted with permission from ref. 334. Copy- permission from ref. 338. Copyright r 1998 Elsevier.
right r 2013 Elsevier.

efficiency of B63%). Normally, the Faradaic efficiency for CO2


Table 7 Tafel parameters obtained from the Tafel plots. bc1 represents
reduction rises with increasing electrolyte concentration.
the slope for the lower overpotential region and bc2 for the high over-
potential region. Reprinted with permission from ref. 334. Copyright r 2.3.2 Pb metal electrodes and Pb-containing catalysts.
2013 Elsevier In general, lead, glassy carbon, mercury, platinum, and gold
are popular electrode materials for CO2 electroreduction.190
SnB SnG SnGDL
The earliest CO2 electroreduction was probably conducted on a
bc1 (mV) 116 180 185 lead cathode at 1.5 to 2.2 V vs. NHE in quaternary ammonium
bc2 (mV) 430 480 458
j0 (A cm2) 1.2  106 2.8  105 1.6  104
salt aqueous solutions, with a carbonate/bicarbonate buffer being
Eeq (V) 0.10 0.030 0.053 used to maintain the pH at 8.3.336 Pb metal electrodes in aqueous
electrolytes favored the formation of formic acid, while in non-
aqueous electrolytes, oxalic acid was the dominant product.322
four-fold higher Faradaic efficiency for CO2 reduction, more On a Pb electrode, the Faradaic efficiency could be increased by
than a Sn electrode coated with a native SnOx layer. They increasing the CO2 pressure and decreasing the temperature.294
suggested that metal–metal oxide composite materials were For example, CO2 electroreduction on a Pb electrode in a
promising catalysts for sustainable fuel synthesis. Recently, CO2-satuarated 0.05 M KHCO3 solution at 0 1C gave a higher
Surya Prakash et al.334 compared three kinds of Sn electrodes for current efficiency than at higher temperature.290 At a high pres-
CO2 reduction in aqueous NaHCO3 solution: a Sn-powder-decorated sure of B50 atm and a high temperature of B80 1C, using Pb
gas diffusion layer (SnGDL) electrode, a Sn metal disc electrode granule electrodes in a fixed-bed reactor fed with aqueous 0.2 M
(SnB), and a Sn powder-coated graphite (SnG) electrode. The K2CO3 electrolyte solution, the only reaction product was found to
exchange current densities ( j0, A cm2) of CO2 reduction on these be HCOOH, with a maximum Faradaic efficiency of 94% at 1.8 V
electrodes, determined by Tafel plots from current–voltage curves, vs. SCE.337 If the reduction was conducted at ambient temperature
showed a five-fold increase on the SnGDL electrode as compared and pressure, CO and methane were also produced.338 The Tafel
to that on the SnG electrode (Fig. 20 and Table 7), although the plots from the current–voltage curves for HCOOH and CO forma-
SnG electrode showed a j0 value two orders of magnitude higher tion (Fig. 21) show a linearity in the entire potential range studied,
than on the SnB electrode. The maximum current density indicating that the electroreduction of CO2 to formic acid and CO
obtained during electrolysis was 27 mA cm2 at 1.6 V vs. was not limited by mass transfer.
NHE, with 70% Faradaic efficiency for the formation of for- Eneau-Innocent et al.339 employed techniques such as cyclic
mate, which probably is one of the highest values to be found in voltammetry, chronoamperometry, and in situ infrared reflectance
the literature on Sn electrodes at ambient pressure. spectroscopy to investigate the catalytic activity of lead electrodes
The electrolyte also has an effect on CO2 reduction on Sn-based towards CO2 electrodimerization in 0.2 M tetraethylammonium
electrodes. For example, Wu et al.335 found that both SO42 and perchlorate–propylene carbonate (TEAP–PC) solution, and found
Na+ gave higher Faradaic and energy efficiencies (as high as B95% that CO was not produced while oxalate was the main product. The
for 0.1 M Na2SO4 at a potential of 1.7 V vs. SCE), while HCO3 CV measurements for CO2 reduction in aqueous medium on a lead
and K+ yielded a higher rate of HCOOH production (0.5 M KHCO3 plate in a filter-press cell showed that when the pH of the cathodic
was found to be the optimal electrolyte for obtaining a high pro- solution was 8, CO2 existed predominantly in the form of HCO3,
duction rate of HCOOH, which reached over 3.8 mmol min1 cm2 and HCOOH was the exclusive product, with high Faradaic yields of
at a potential of 2.0 V vs. SCE while maintaining a Faradaic 65–90%.340 The main reaction can be expressed as follows:

658 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

HCO3 + H2O + 2e - HCOO + 2OH (54)

Subramanian et al.341 designed a flow type electrochemical


membrane reactor to improve CO2 mass transfer in electro-
catalytic reduction in a potassium phosphate buffer solution.
The anode and cathode chambers in this reactor were separated
by a composite perfluoropolymer cation exchange membrane
(Nafions 961 and Nafions 430). With a lead-coated cathode,
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

a maximum current efficiency of 93% for formate formation


was achieved.

2.4 Alkaline metals and alkaline earth metals


Neither alkaline metals nor alkaline earth metals can be used
as electrodes for CO2 catalytic reduction catalysts due to their
instability in electrochemical systems. However, their salts,
commonly used as supporting electrolytes in electrochemical
cells for CO2 electroreduction, exhibit different effects on product
selectivity, reduction reaction rate, and even catalyst stability.
To study the effect of electrolytes on CO2 electroreduction,
Murata and Hori342 tested several aqueous 0.1 M MHCO3
solutions, where M = Li+, Na+, K+, and Cs+. They found that
H2 evolution prevailed over CO2 reduction in the Li+ electrolyte,
whereas CO2 reduction was favorable in Na+, K+, and Cs+
solutions. In addition, they also observed that the magnitudes
of the C2H4/CH4 ratio were in the order Li+ > Na+ > K+ > Cs+,
indicating that the current efficiency for the formation of C2H4 Fig. 22 Proposed mechanism for the enhancement of Mg2+ in CO2
electroreduction catalyzed by Fe(0)TPP catalyst on a Cu electrode in
was apparently related to the size of the cation radius, which
DMF + TAA solution. Reprinted with permission from ref. 32. Copyright
should have a strong effect on the outer Helmholtz plane (OHP) r 1991 American Chemical Society.
potential of the electrode–electrolyte interface. The same con-
clusion was also reached by Kyriacou and Anagnostopoulos.343
In their study on CO2 electroreduction catalyzed by Fe(0)TPP in Table 8 Typical Faradaic efficiencies of CH4 and C2H2 production in the
DMF with tetraalkylammonium (TAA) salts as the supporting electrochemical reduction of CO2 at a Cu electrode. Reprinted with
permission from ref. 348. Copyright r 2007 Springer
electrolyte, adding Mg2+ ions into the solution dramatically
improved the rate of CO2 reduction to CO. The catalyst’s Faradaic efficiency (%)
stability was also improved, and a Faradaic efficiency of over CH3OH H2O
94% was achieved. The mechanism is presented in Fig. 22.
Cation of supporting salts CH4 C2H4 CH4 C2H4
Other salts have also been used as supporting electrolytes in
CH3OH solution to study their effect on CO2 electroreduction Li 63 14.7 26 4
Na 63 17.6 19 11
on a Cu electrode. They include Li salts (LiBF4, LiClO4, K 16 37.5 16 14
LiCl, LiBr, LiI, LiClO4, and CH3COOLi),242,252 sodium salts Rb 4.6 31.0 — —
(CH3COONa, NaCl, NaBr, NaI, NaSCN, and NaClO4),344 potassium Cs 4.1 32.7 15 13
salts (CH3COOK, KBr, KI, and KSCN),345 cesium salts (CH3COOCs,
CsCl, CsBr, CsI, and CsSCN),346 and alkali hydroxides such as
LiOH,253 NaOH,347 KOH, RbOH,348 and CsOH.347 Table 8 lists 2.5 Other catalysts
some typical cations of supporting electrolytes and their effects Besides the metals and metal complexes mentioned above,
on the Faradaic efficiency of CO2 electroreduction. organic molecules can also be mediators and catalysts for
Interested in the effect of multivalent supporting cations, CO2 electroreduction. Oh and Hu349 have recently published a
Schizodimou and Kyriacou311 investigated CO2 electroreduction review of the literature on this subject.
on a Cu(88)–Sn(6)–Pb(6) alloy cathode in 1.5 M HCl and found Conducting polymer electrodes. Conducting polymer electrodes
that the reduction rate slightly increased in the presence of have also been developed for the heterogeneous electrocatalysis
divalent cations such as Mg2+, Ca2+, and Ba2+. They also observed of CO2 reduction. For example, Koleli et al.350 developed a poly-
that the enhancement in the reduction rate was dependent on aniline (PAn) electrode. In methanol solution, the maximum
the charge number of the supporting electrolyte cation—the Faradaic efficiencies were found to be 12% for formic acid and
higher the charge number, the greater the rate enhancement, in the 78% for acetic acid. Aydin et al.351 developed a polypyrrole (PPy)
order Na+ o Mg2+ o Ca2+ o Ba2+ o Al3+ o Zr4+ o Nd3+ o La3+. electrode. In the electrocatalytic reduction of CO2 under high

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 659
View Article Online

Review Article Chem Soc Rev


Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Fig. 23 Proposed overall mechanism for pyridinium-catalyzed CO2 reduction to three products: formic acid, formaldehyde, and methanol. Reprinted
with permission from ref. 354. Copyright r 2010 American Chemical Society.

pressure in CH3OH at an overpotential value of 0.4 V vs.


Ag/AgCl, the maximum Faradaic efficiencies at 20 bar were 1.9,
40.5, and 62.2% for HCHO, HCOOH, and CH3COOH, respec-
tively. Smith et al.352 synthesized two organic polymers based
on repeating benzimidazole and pyridine–bipyridine units,
respectively. The pyridine-based polymer exhibited stable elec-
trochemical behavior, while the bipyridine-based polymer gave
large catalytic currents for CO2 reduction in CH3CN containing
1% H2O.
Aromatic amine catalysts. Seshadri et al.353 found that the
pyridinium cation and its substituted derivatives could be
effective and stable homogeneous electrocatalysts for the multiple-
electron, multiple-proton reduction of CO2 to methanol at low
potentials, with Faradaic yields of up to 30%. The selectivity
for pyridinium-catalyzed methanol production was found to Fig. 24 Two potential routes for the formation of PyCOOH0 in the
increase significantly under photoelectrocatalytic conditions. homogeneous phase. Reprinted with permission from ref. 357. Copyright
Cole et al.354 observed CO2 reduction to HCOOH, HCHO, and r 2013 American Chemical Society.
CH3OH at 0.58 V vs. SCE when using a Pt disk electrode in a
10 mM aqueous solution of pyridine (Py) at pH 5.3. In the
Py–p-GaP system. Based on ab initio quantum chemical calcula-
proposed overall mechanism, with an inner-sphere-type electron
tions, they identified PyCOOH0 as an important intermediate,
transfer, the pyridinium radical was believed to play a role in
whose formation was believed to be the rate-determining step for
the reduction, as shown in Fig. 23. Ertem et al.355 proposed that
CO2 reduction to CH3OH. As shown in Fig. 24, in the homo-
the pyridinium cation ‘‘PyH+’’ could undergo a one-electron
geneous phase, the formation of PyCOOH0 proceeds by two
reduction, forming hydrogen atoms adsorbed on the Pt surface,
potential routes. In Route 1, Py is protonated to PyH+, which
‘‘Pt–H’’. This ‘‘Pt–H’’ was susceptible to electrophilic attack
is then reduced to PyH0, which reduces CO2 to form PyCOOH0.
by CO2, leading to a two-electron proton-coupled hydride
In Route 2, Py and CO2 are combined to form PyCO2, which is
transfer reaction:
reduced to PyCOO and, finally, protonated to PyCOOH0.
CO2 + Pt–H + PyH+ + e - Py + Pt + HCOOH (55) Bocarsly et al.147 compared the catalytic activity of imidazole
and pyridine in the photoelectrochemical reduction of CO2 at
In a theoretical study of CO2 electroreduction in the pre- illuminated iron pyrite (FeS2) electrodes. The aqueous electrolyte
sence of a pyridinium cation and its substituted derivatives, was composed of 10 mM catalyst and 0.5 M KCl solution, and a
Keith and Carter356 employed first-principles quantum chemistry 350–1350 nm light beam with an intensity of 890 mW cm2 was
and the thermodynamic energies of various pyridine-derived used to illuminate the electrode surface. The mechanism of
intermediates, as well as energy barrier heights for key homo- imidazole-based catalysis was investigated using CV measurements
geneous reaction mechanisms. They predicted that the actual (over a scan rate range of 5 mV s1 to 200 mV s1) to analyze the
form of the co-catalyst was not the long-proposed pyridinyl catalytic activity of a series of imidazole derivatives toward CO2
radical in solution, but was more probably a surface-bound reduction. Results indicated that imidazole could reduce CO2
dihydropyridine species. Lim et al.357 investigated the mecha- to a mixture of CO and HCOOH at a moderate potential, while
nism of homogeneous CO2 reduction by pyridine (Py) in the pyridine selectively catalyzed the production of HCOOH.

660 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

Radical anion catalysts. Gennaro et al.358,359 found that the


anion radicals of certain aromatic esters (i.e., phenyl benzoate
and methyl benzoate) and nitriles (i.e., benzonitrile) possessed
remarkable catalytic activity toward CO2 reduction to oxalate. The
reduction pathway was simply described as follows: (1) A (aromatic
ester or nitrile) was first reduced to A (an aromatic radical anion);
(2) A transferred the electron to CO2, forming CO2*; and (3) two
CO2* dimerized to give oxalate.
Ionic liquid (catalysts). CO2 has been dissolved in the ionic
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

liquid EMIMBF3Cl and electrochemically reduced at ambient


pressure and room temperature.360 The BF3Cl anion catalyzed
the CO2 reduction by forming a Lewis acid–base adduct,
BF3–CO2. The B–Cl bond was relatively weak in BF3Cl.
Thermogravimetric analysis (TGA) results showed that the
Fig. 26 Steady-state electrochemical kinetics visualized using current–
decomposition of EMIMBF3Cl required two steps. The first
voltage curves. When both the oxidized and the reduced forms of a redox-
included the breaking of the B–Cl bond and then the release of active species are present, a reversible electrochemical reaction (one with
BF3 at about 220 1C; the second involved the breaking of other a large exchange current density) produces a single sigmoidal wave (blue)
bonds which started at about 330 1C. that cuts (without inflection) through the zero-current axis at the equili-
Enzyme catalysts. Two and a half decades ago, Yoneyama brium potential (Eeq) and achieves a potential-independent limiting current
in either direction at a relatively low overpotential. Conversely, if the
and coworkers361 used isocitrate dehydrogenase (IDH) as an
exchange current density is low, the current is negligible around Eeq, and
electrocatalyst for the fixation of CO2 to isocitric acid (HOOCCH2- two sigmoidal waves (red), one for either direction, are separated in
CH(COOH)CH(OH)COOH) in oxoglutaric acid (HOOCCH2CH2- potential, emerging from the baseline with an exponential dependence
C(QO)COOH). The reaction occurred selectively, with current on potential. A substantial overpotential is required to match the current
efficiencies approaching 100% at 0.95 V vs. SCE in 0.2 M produced by the reversible system. Reprinted with permission from
ref. 365. Copyright r 2011 National Academy of Sciences.
(HOCH2)3CNH2 (tris buffer, pH 7). They also fixed CO2 to yield
pyruvic acid (CH3C(QO)COOH) in acetyl-coenzyme A using
pyruvate dehydrogenase complexes as electrocatalysts.362 Solvent found that the addition of carbonic anhydrase could efficiently
molecules seemed to be involved in the reactions. Kuwabat accelerate CO2 reduction achieved by formate, aldehyde, and
et al.363 reported that at potentials between 0.7 and 0.9 V vs. alcohol dehydrogenase, although the process could be cata-
SCE, the electrolysis of CO2-saturated phosphate buffer solu- lyzed only by dehydrogenase. With respect to this, a com-
tions (pH 7) that contained formate dehydrogenase (FDH) and parison is provided in Fig. 25.364 Hansen et al.365 recently
either methyl viologen (MV2+) or pyrroloquinolinequinone as developed a model based on DFT calculations to describe
an electron mediator yielded HCOO with current efficiencies (1) trends in catalytic activity for CO2 reduction to CO at a metal
of 90%. The FDH demonstrated considerable durability. surface and (2) the active sites in CODH enzymes, in terms of the
They also found that the electrolysis of phosphate buffer adsorption energy of the reaction intermediates, CO and COOH.
solutions containing HCOONa in the presence of methanol Enzymes that were able to catalytically transform small mole-
dehydrogenase (MDH) and MV2+ at 0.7 V vs. SCE yielded cules (e.g., CO, formate, or protons) were a special category of
HCHO when the enzyme concentration was low, whereas it electrocatalysts. Due to the presence of enzymes (catalysts), those
produced both HCHO and CH3OH when the concentration previously irreversible processes could become electrochemically
became relatively high. The formation and accumulation of reversible, as shown in Fig. 26.366 In the active sites, some
formaldehyde promoted methanol production.363 Addo et al.364 elements (e.g., Ni, Fe, Cu, Se, Mo, and W) have been found.366–368

3. Product selectivity in the


electrocatalytic reduction of CO2
As discussed above, catalyst selectivity to produce desired products in
catalyzed CO2 electroreduction is very important for practical appli-
cations. Normally, this selectivity is closely related to the reduction
mechanism, with different reaction pathways or combinations of
different pathways leading to different products. In the initial
reduction step of a typical CO2 electroreduction mechanism, CO2
can obtain electrons either directly from the cathode surface (a bare
Fig. 25 Schematic of the enzyme cascade reaction from CO2 to metha-
nol on an electrode surface without carbonic anhydrase IV (left) and with
electrode surface or a surface coated catalyst) or indirectly from a
carbonic anhydrase IV (right). Reprinted with permission from ref. 364. medium, such as a soluble catalyst, to produce an intermediate,
Copyright r 2011 American Electrochemical Society. such as CO2 , which then absorbs on the cathode for product

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 661
View Article Online

Review Article Chem Soc Rev

Table 9 Proposed reaction pathways in CO2 electroreduction252,312,322,370

CO2 + e -  CO2(ads) 
CO2(ads) + H+ + e CO + 4H+ + 4e 
CH2(ads) + 2H+ + 2e - CH4
or - CO + OH -  CH2(ads) + H2O 2 CH2(ads) - C2H4
NR4+ + e -  NR4 2 CH2(ads) + 2H+ + 2e - C2H6
CO2 +  NR4 
CO2(ads) + H2O HCO2 (ads) + e - HCO2 HCO2 + CH3OH
-  CO2 + NR4+ (ref. 252) -  HCO2(ads) + OH - HCOOCH3 + OH (ref. 312)
2 CO2(ads) - C2O42 C2O42 + 2H+ + 2e HC(O)COO + 2e + 2H+
- HC(O)COO + OH - H2C(OH)COO (ref. 322)
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

formation. In general, which kind of pathways and how many Mo, Fe, Ru, Co, Ni, Pd, Cu, Zn, Cd, Ga, In, Ge, Sn, Pb), the current
pathways are required for the reduction process will be strongly efficiency of CO formation at CoPc and NiPc catalysts in the
affected by experimental conditions, such as the catalysts and/or reduction of CO2 alone was further demonstrated to be far higher
electrodes, electrode potential, electrolyte solution, buffer strength, than at pure metal catalysts; hence, these M/Pc catalysts were
pH, CO2 concentration and pressure, as well as temperature.369 expected to have a fairly high capacity for urea formation.373 The
The possible pathways are summarized in Table 9.252,312,322,370 effects of various aza-macrocyclic ligands on the production of
fuels from macrocyclic complexes were also investigated, including
Effects of single metal electrode type. Regarding product simple tetraazamacrocycles, porphyrin, phthalocyanine, and
selectivity, single metal electrodes seem to be the most popular biphenanthrolinic hexaazacyclophanes.374,375 The mechanism
type of electrocatalysts for CO2 reduction.241,290 Two groups can for producing a particular product is normally related to the
be roughly designated:176 (1) CO formation metals (Cu, Au, structural features of the azacyclam framework and its inter-
Ag, Zn, Pd, Ga, Ni, and Pt) and (2) formate formation metals action with the central metal and CO2 or CO molecules, while
(Pb, Hg, In, Sn, Cd, and Tl). There are other types of catalysts replacement of a –CH2 group in the ligand backbone by an amide
that also have both high selectivity and high current effici- residue, for example, does not disturb the catalytic process.
ency. In addition, some electrocatalysts have been specifically Effects of cations and anions in the electrolyte. As described
designed for and show unique catalytic activities toward previously, alkaline metals and alkaline earth metals cannot be
CO2 reduction to produce desired products with high current used as electrodes for CO2 catalytic reduction catalysts. However,
efficiencies. their salts have commonly been used as supporting electrolytes
Effects of metal complex, metal center, and ligand type. It is in electrochemical cells for CO2 electroreduction, and they have
well known that the catalytic performance of metal complex different effects on product selectivity. For example, in the case
catalysts for CO2 reduction strongly depends on the chemical of FeTPP catalyst, the addition of Lewis acid cations such as
properties of the metal center and ligand. Therefore, it is Mg2+, Ca2+, Ba2+, Li+, or Na+ decreased HCOOH formation as the
expected that the distribution of the electrolysis products, the acidity increased; the order of reactivity of these Lewis acid
current efficiencies, and the reaction mechanism of CO2 elec- synergists was Mg2+ = Ca2+ > Ba2+ > Li+ > Na+.33 Thorson
troreduction will also be strongly affected by the type of central et al.376 confirmed that the presence of large cations such as
metals and ligands in macrocyclic complexes. For example, CO2 cesium (Cs) and rubidium (Ru) in the electrolyte could enhance
electroreduction on a GCE modified with polymeric M-tetrakis the electrochemical conversion of CO2 to CO. This was explained
aminophthalocyanines (M = Co, Ni, Fe) indicated that different by the interplay between the level of cation hydration and the
metal centers produced different products.103 When M was Co, extent of cation adsorption on the metal electrodes. The effects
HCOOH was the only product; when M was Fe, a mixture of of anions in the electrolyte on the products of CO2 electro-
CH2O and H2 was produced, whereas when M was Ni, a mixture reduction were also investigated using a copper mesh electrode
of HCOOH and CH2O was observed.371 Furuya and Matsui372 in aqueous solutions containing 3 M KCl, KBr, and KI as the
investigated the electrocatalytic reduction of CO2 on GDEs respective electrolytes.377 The results showed that the bond
modified by 16 kinds of metal phthalocyanine (MPc) catalysts between the adsorbed halide anion (e.g. Br, Cl, or I) and
(where M = Co, Ni, Fe, Pd, Sn, Pb, In, Zn, Al, Cu, Ga, Ti, V, Mn, carbon helped the electron transfer from the adsorbed halide
Mg, Pt) in 0.5 M KHCO3. They found that the distribution and anion to the vacant orbital of CO2, promoting CO2 conver-
current efficiencies of the electrolysis products were strongly sion.377 The stronger the adsorption of the halide anion to the
dependent on the nature of the central metal coordinated to the electrode was, the more strongly CO2 was restrained, resulting
phthalocyanines. With transition metals of Co- and Ni-phthalo- in a higher CO2 reduction current. Furthermore, the specifically
cyanines, the main electrolysis product was CO, with a current adsorbed halide anions suppressed the adsorption of protons,
efficiency of B100%. On the other hand, HCOOH was the main leading to a higher hydrogen overvoltage. This reaction mecha-
product on phthalocyanines with Sn, Pb, or In metal centers. nism was also confirmed by Schizodimou and Kyriacou,311 who
The highest current efficiency, B70%, was observed on SnPc showed that the rate of electrochemical reduction of CO2
around 1.6 V. In the case of Cu-, Ga-, and Ti-phthalocyanines, increased in the order Cl o Br o I.
CH4 was the main product, with the highest current efficiencies Effects of supporting electrolytes. In fact, even for the same
being 30–40%. Analogously, in the simultaneous reduction of metal electrode with the same purity, different supporting
CO2 and NO3 with various MPc catalysts (where M = Ti, V, Cr, electrolytes have a great effect on the final products. For

662 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

example, in the electroreduction of CO2 at a copper electrode the large-scale electrochemical reduction of CO2 to HCOOH
(99.999% purity) in methanol, Saeki et al.252 observed two and HCOO has been discussed by Agarwal et al.8
different main products—CO (current efficiency 48.1–86.8%) Electrocatalytic reduction of CO2 by an enzyme catalyst,
and CH3COOH (54.5 and 46.7%)—when the supporting electro- namely, formate dehydrogenase enzyme (FDH1), which was
lyte was either tetrabutylammonium (TBA) salts (i.e., TBABF4 and isolated from Syntrophobacter fumaroxidans, was found (as
TBAClO4) or lithium salts (LiBF4 and LiClO4). They proposed that either a homogeneous or a heterogeneous catalyst) to produce
the intermediate, CO2*, was stabilized by forming a TBA+–CO2* formate exclusively.58 Two acetogenic bacteria, Moorella
ion pair or by being adsorbed on the electrode surface. By thermoacetica (Mt, formerly Clostridium thermoaceticum, ATCC
contrast, NH4ClO4 as a supporting electrolyte in the same 35608) and Clostridium formicoaceticum (Cf, DSM 92), were
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

electrolysis only led to hydrogen evolution (84.6%). Other explored as catalysts for CO2 electroreduction in 1.0 atm CO2-
electrolytes have also been explored, such as tetraethylammonium saturated 0.1 M phosphate buffer solution (pH 7.0) at 0.58 V vs.
perchlorate (TEA).339 NHE; the results showed that these catalysts could efficiently
Effects of solvent–CO2 concentration. It should be mentioned convert CO2 to formate with current efficiencies of 80% for Mt
that CO2 utilization in aqueous solution can be limited by its low and 100% for Cf.382
solubility in water at standard temperature and pressure.378 This
is because there are relatively small amounts of CO2 available for 3.2 Selective production of carbon monoxide
the reaction to proceed at the electrode surface. To speed up the CO is one of the important products generated in CO2 electro-
reaction process for industrial purposes, pressurized CO2 is reduction. Some metal cathodes, such as Ag, Au, and Zn, are
normally required,79,379 which often causes a certain degree of highly selective for the electrocatalytic reduction of CO2 to CO
change in product selectivity. Normally, in aqueous solutions, in KHCO3 aqueous solutions.38 [NiII(cyclam)]2+ is a well-known
metallic catalysts (or electrodes), such as sp group metals (e.g., In catalyst for CO2 electroreduction to CO on a mercury cathode
or Pb), tend to give higher CO production at pH levels higher than 4, at 0.9 V in aqueous solutions.111 Two other electrocatalysts
while d group metals (such as Pd and Cu) can promote HCOOH with a similar structure to [Ni(cyclam)]2+ were also found to be
production.369 It was concluded that the main products obtained in selective for CO production.120 In addition, metal polyphosphine
aqueous media under ambient conditions were strongly dependent complexes, such as Pd(triphosphine)L2+ (L = CH3CN, P(OMe)3,
on the type of cathode: Cu electrodes mainly yielded mixtures of PEt3, P(CH2OH)3, and PPh3), exhibited high catalytic activity for
hydrocarbons (mostly methane and ethylene) and alcohols; Au, Ag, the reduction of CO2 to CO in acidic CH3CN solutions.216
and Zn mainly produced CO, whereas other metals, such as In, Sn, As an enzyme-based catalyst, a Ni- and Fe-containing metallo-
Hg, and Pb, were selective for the production of formic acid/ enzyme isolated from Moorella thermoacetica showed highly
formate.365 Compared to water solutions, the application of selective activity towards the conversion of CO2 to CO, with a
nonaqueous solvents is relatively popular due to their high current efficiency as high as B100% at 0.57 V vs. NHE in a
solubility for CO2. For example, DMF, PC, and CH3OH may 0.1 M phosphate buffer solution (pH 6.3).173
contain, respectively, up to 20, 8, and 5 times more CO2 than
corresponding amounts of aqueous solutions. Among them, 3.3 Selective production of formaldehyde
solvents having low proton availability, such as DMF, favor the Senda et al.57 reported that [M(4-v-tpy)2]2+ and [M(6-v-tpy)2]2+
formation of oxalate and CO, while aqueous solution favors (M = Cr, Ni, Co, Fe, Ru, or Os), after being electropolymerized onto
formate.380 Strategies that have been demonstrated experimentally GCEs to form films, exhibited electrocatalytic activity toward CO2
are further described below according to the selective generation of reduction, with formaldehyde as virtually the only product. The
desired products. current efficiency for films of Cr[(4-v-tpy)2]2+ was as high as 87%.

3.4 Selective production of methanol


3.1 Selective production of formic acid (and formate) Some electrocatalysts have been found to be selective for
Electrochemical reduction of CO2 in aqueous solution to formic methanol production in CO 2 electroreduction, as listed in
acid and formate was reported as early as 1870.381 Noda et al.38 Table 10. In addition, the photoelectrocatalytic reduction
reported that Zn, Cd, Hg (Group 12), In (Group 13), Sn, and Pb of CO 2 by semiconductor materials also showed selectivity
(Group 14) metal cathodes in 0.1 M KHCO3 aqueous solution for methanol production. The pyridinium cation on Pt also
exhibited high production selectivity for HCOO formation. reduced CO 2 to methanol, with a Faradaic efficiency of
Chen et al.232 reported that HCOOH was the main product 30%. 353,354
of CO2 electroreduction by both [Ir2(dimen)42+](PF6)2 and
[Ir2(dimen)42+](B(C6H5)4)2 (dimen = 1,8-diisocyanomenthane) 3.5 Selective production of oxalic acid (oxalate)
electrocatalysts. As previously mentioned, to optimize the large- Electroreduction of CO2 in DMF solution showed a current
scale (even industrial-scale) electrocatalytic reduction of CO2 to efficiency of 73% for oxalic acid, with a little formate and
formate, Li and Oloman331 investigated a series of condition CO production.140 The macrocyclic nickel complex Ni-Etn-
variables that might affect the performance of a reactor. The (Me/COOEt)-Etn was found to be one of the most active and
current efficiency of formate formation was reportedly as high persistent homogeneous catalysts for CO2 electroreduction
as 91%. Recently, the engineering and economic feasibility of selectively to oxalate.128 Dinuclear copper(I) complexes

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 663
View Article Online

Review Article Chem Soc Rev

Table 10 Summary of electrocatalytic reduction of CO2 to selectively produce several important low-carbon fuels

Temperature/
Electrode pressure and Faradaic
Product Electrode/electrocatalysts potential (V) Electrolyte other conditions efficiency (%) Ref.
CO Ag(99.98%) electrode 1.6 V vs. 64.7 38
Au(99.95%) electrode Ag/AgCl saturated 81.5
with KCl
Glassy carbon electrode (WE)/Re(I)(bpy)(CO)3Cl 1.25 V vs. NHE 0.1 M Et4NCl DMF-H2O 25 1C 98 61
(10%) solutions
Platinum gauze electrode (WE)/Re(CO)3(vbpy)Cl 1.55 V vs. CH3CN-Bu4NPF6 satu- 92.3 63
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

sodium SCE rated with CO2


Platinum (WE), Ag/10 mM Ag+ (RE)/(1 mM) fac- 1.85 V CH3CN + 0.1 M TBAP 92 65,
Re(L)(CO)3Cl (L = pyrrole-substituted 2,2 0 - 66
bipyridine)
Carbon felt (WE)/fac-Re(L)(CO)3Cl (L = pyrrole- Max. 98.5 66
substituted 2,2 0 -bipyridine)
Polymeric films formed by coelectropolymerization 1.55 V vs. SSCE 0.1 M TBAH in CH3CN 90–98 67
of cis-[(bpy)2Ru(vpy)2]2+ with fac-[Re(CO)3(vbpy)Cl] saturated with CO2
or fac-[Re(CO)3(vbpy)CH3CN]+
Re film electrodeposited onto a polycrystalline gold 1.35 V 0.1 M LiClO4 CH3OH 1 atm CO2, stirred 87 68
support solution conditions
Quiescent 57
conditions
fac-(5,5 0 -Bisphenylethynyl-2,2 0 -bipyridyl)Re(CO)3Cl 1.750 V vs. NHE CH3CN with 0.1 M 45 72
(n-CH3(CH2)3)4NPF6
[Mn(bpy)(CO)3]+ 1.70 V 0.1 M TBAP in MeCN At room 85 77
[Mn(dmbpy)(CO)3]+ temperature 100
Glassy carbon (WE), Pt wire (CE), Ag/AgCl (RE)/ 2.2 V vs. SCE DMF/2 M H2O 100 78
[Mn(bpy-t-Bu)(CO)3]+
Glassy carbon (WE); aqueous SCE electrode (RE); Pt DMF + 0.1 M n-Bu4NPF6 21 1C/? >90 86
wire (CE) + 2 M H2O
(1) Iron 5,10,15,20-tetrakis (2 0 ,6 0 -dihydroxyphenyl)- 1.333 V vs. NHE
porphyrin (Fe0TDHPP)
(2) Iron 5,10,15,20-tetrakis (2 0 ,6 0 -dimethoxyphenyl)- 1.69 V vs. NHE
porphyrin (Fe0TDMPP)
Carbon monoxide dehydrogenase (CODH) from 173
Moorella thermoacetica
Pd metal 1.8 V vs. Ag/AgCl 0.1 M KHCO3 aqueous 50 atm 57.9 184
solution
cis(Cl)-[Ru(bpy)-(CO)2Cl2], cis(CO)- 1.5 V Ag/Ag+ TBAP (0.1 M) + CH3CN 95–97 222
[Ru(bpy)(CO)2(C(O)OMe)Cl]/ (0.01 M) saturated with CO2
3.5 CH3OH with 500 mM 30 1C, CO2 84 255
CsOH supporting salt pressure of 10 atm
Ag(99.98%) electrode 1.6 V vs. 0.1 M KHCO3 aqueous 25 1C 64.7 38
Au(99.95%) electrode Ag/AgCl saturated solution saturated with 81.5
with KCl KCl
Metallic Zn electrode 0.1 M KHCO3 aqueous 39.6 38
Metallic Cd electrode solution 14.4
Metallic Hg electrode ND

HCOOH/ Working electrode: metallic electrodes 0.1 M KHCO3 aqueous 38


HCOO Zn solution 20
Cd 39
Hg 94
Polished pyrolytic graphite edge electrode (WE)/ 20 mM Na2CO3 (pH 6.5) 37 1C B100 58
tungsten-containing formate dehydrogenase solution
enzyme (FDH1) from Syntrophobacter fumaroxidans
Fe wire (99.5%, 0.16–0.63 cm2) electropolished in 1.53 to 1.61 0.1 M KClO4 25 1C; 30 atm, 59.5–59.6 79
HClO4-(CH3CO)O-H2O vs. Ag/AgCl 120 mA cm2
Ni(cyclam)2+ 75 118
[Ru(bpy)2(CO)2]2+ or [Ru(bpy)2(CO)Cl]+ (homo- 1.3 V vs. SCE Saturated H2O/DMF 84.3 218
geneous catalysts)/Hg pool (9 : 1, v/v) solution in
the presence of Me2NH
HCl, pH 9.5
[(bpy)2Ru(dmbbbpy)](PF6)2 (dmbbbpy = 2,2 0 -bis(1- 1.65 V vs. CO2 saturated MeCN + 89 224
methylbenzimidazol-2-yl)-4,4 0 -bipyridine) Ag/AgCl 2.5% H2O
[(bpy)2Ru(dmbbbpy)Ru(bpy)2](PF6)4 1.55 V vs. CO2 saturated MeCN + 90 224
Ag/AgCl 2.5% H2O
Tinned-copper sheet 0.45 M KHCO3 0.22 kA m2, ambient 86 264
conditions
Metallic Pb electrode 2.4 V Ag/AgCl 0.1 M TEAP/H2O 100 1C, normal 78.9 322
pressure

664 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

Table 10 (continued)

Temperature/
Electrode pressure and Faradaic
Product Electrode/electrocatalysts potential (V) Electrolyte other conditions efficiency (%) Ref.
Sn-powder-decorated gas diffusion layer (SnGDL) 1.6 V vs. NHE Aqueous NaHCO3 27 mA cm2 70 334
electrode solution
Sn foil (Alfa Aesar, 99.998%) with an active surface 1.7 V vs. SCE 0.1 M Na2SO4 B95 335
area of 1 cm2
Pb granule electrodes 1.8 V vs. SCE 0.2 M K2CO3 aqueous B80 1C, B50 atm 94 337
solution
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

CH3OH RuO2–TiO2 nanoparticle (NP) composite electrodes 0.8 V vs. SCE 0.5 M NaHCO3 solution 40.2 46
(WE); SCE (RE); Pt plate (CE)–RuO2 saturated with CO2
RuO2–TiO2 nanotube (NT) composite electrodes 60.5
(WE); SCE (RE); Pt plate (CE)–RuO2
Molybdenum metal 0.7 to 0.8 V vs. 0.2 M Na2SO4 solution 20 1C/?/pH 4.2 >50 47
SCE saturated with CO2
Electrodeposited cuprous oxide film 1.1 V vs. SCE 0.5 M KHCO3 solution 38 142
saturated with CO2
II II
Platinum plate electrode/KFe [Fe (CN)6] >80 154
Ru 0.54 V vs. SCE 42 178
Ru/Cu 0.8 V vs. SCE 0.5 M NaHCO3 solution 41.3 177
saturated with CO2
Cu 1.1 V vs. SCE 40 370
RuO2–TiO2 0.95 V vs. SCE 30 209
p-GaP 1.4 V vs. SCE 60 317
Illuminated p-GaP photoelectrodes 0.52 V vs. SCE B100 320
p-GaAs 1.3 V vs. SCE 55 326
p-InP 70
n-GaAs 100
Pyridinium cation and substituted derivatives 30 353
(homogeneous electrocatalysts)

H2C2O4/ Ni-Etn(Me/COOEt)Etn ACN + 0.25 M Bu4NClO4 98 128


C2O42 saturated with CO2
[(bpy)2Ru(dmbbbpy)](PF6)2 (dmbbbpy = 2,2 0 -bis(1- 1.65 V vs. CO2-saturated MeCN + 64 224
methylbenzimidazol-2-yl)-4,4 0 -bipyridine) Ag/AgCl 2.5% H2O
[(bpy)2Ru(dmbbbpy)Ru(bpy)2](PF6)4 1.55 V vs. CO2-saturated MeCN + 70 224
Ag/AgCl 2.5% H2O
Metallic Pb electrode 2.6 V Ag/AgCl 0.1 M TEAP/propylene 100 1C under 73.3 322
carbonate (PC) normal pressure

CH4 Au electrodes 1.35 V 0.1 M LiClO4 in CH3OH Atmospheric 68


PpyRe microalloy polypyrrole solution pressure of CO2 34
PpyCu–Re microalloy polypyrrole 31
99.999% Cu sheet cathode 0.5 M KHCO3 aqueous 0 1C 65 247
solution saturated with
CO2

C2H4 99.999% Cu sheet cathode 0.5 M KHCO3 aqueous 40 1C 20 247


solution saturated with
CO2
99.999% Cu electrode 1.40 V vs. NHE, 0.1 M KClLO4, aqueous 19 1C/pH 5.9 48.1 248
2
5 mA cm , solution saturated with
CO2
99.98% Cu electrode 3.5 V vs. CH3OH with 80 mM 30 1C 32.3 254
Ag/AgCl CsOH supporting salt,
Cu single-crystal electrodes 0.1 M KClO4 18 1C/pH 6.8 31.7 276
Polycrystal 1.44 V
(100) 1.42 V
(110) 1.55 V
(111) 1.56 V vs. NHE

HCHO GCEs (WE); Ag/AgCl (saturated with sodium chlor- 1.100 V vs. 0.10 M NaClO4 solution 87 57
ide) (+0.222 vs. NHE) electrode (RE); Pt plate (CE)/ Ag/AgCl saturated with CO2
electropolymerized [Cr(4-v-tpy)2]2+ film

electrocatalyzed CO2 conversion selectively to oxalate in precipitation of lithium oxalate.172 In addition, anion radicals of
CH3CN with a soluble lithium salt, resulting in quantitative aromatic esters such as phenyl benzoate and methyl benzoate, and

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 665
View Article Online

Review Article Chem Soc Rev

of nitriles such as benzonitrile, in DMF at an inert electrode (e.g., catalysts, in the electrochemical reduction of CO2 (ERC) to
mercury) were able to reduce CO2 exclusively to oxalate.358,359 HCOO/HCOOH at a gas/solid/liquid interface, using a flow-
through reactor. Although better durability was observed in Sn
3.6 Selective production of lactic acid than in Cu, a color change appeared on the electrode surface, as
Lactic acid (CH3CH(OH)COOH) is an organic C3 compound that well as slight deactivation. Wu et al.335 observed the effects of
plays an important role in numerous industries, including food, the electrolyte on selectivity and activity with a Sn electrode. For
medicine, and cosmetics. Ogura et al.159 found selective a pure Sn electrocatalyst, a decrease in performance could be
reduction of CO2 to lactic acid, catalyzed by Fe(II)-4,5-dihydroxy- caused by several factors:388 (1) cathodic degradation of the
benzene-1,3-disulfonate immobilized on a PAn/PB-modified catalyst surface, (2) deposition of non-catalytic species from
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Pt electrode in 0.5 M KCl solution. However, if Co-4,5- reaction intermediates in the reduction of the pollutant species,
dihydroxybenzene-1,3-disulfonate was used, acetaldehyde was (3) deposition of non-catalytic metallic species from contami-
also produced. nants in the electrolyte,335 and (4) anodic degradation of the
To summarize the selective generation of desired products catalyst at sites where gas bubbles formed, preventing the
in electrocatalytic CO2 reduction, Table 10 lists typical examples cathodic polarization of the catalyst. Bujno et al.389 conducted
of several important low-carbon fuels, together with their gen- experiments in diluted solutions and confirmed that the Ni(I)
eration conditions. complex catalysts present at the electrode surface were trans-
formed into a catalytically inactive Ni(0) carbonyl deposit,
blocking the electrode surface against further catalysis.
4. Catalyst stability, activity Benson and Kubiak390 investigated the deactivation pathway
degradation, and mitigation of the Lehn catalyst. One pathway was concluded to be the
formation of thermodynamically stable and often catalytically
Catalytic activity is normally evaluated by considering both the inactive dimers.73,216,391 Pugh et al.220 found that the electro-
onset potential of reduction and the Faradaic efficiency, while catalytic activity of cis-[Ru(bpy)2(CO)H]+ decreased slowly over
catalyst stability (or durability) is assessed according to varia- an extended period. Normally, active species (sites) in catalysts
tions in catalyst behavior with electrolysis time.331 are always responsible for the catalytic activity and are indis-
With respect to catalyst stability, the issue of deactivation pensable for electrocatalytic CO2 reduction.61,216,297,392 Loss in
has often been reported; the formation of poisonous inter- catalytic activity is always associated with the disappearance of
mediates and the deposition of inactive compositions on active sites. For example, during CO2 reduction, Ru-based com-
electrode surfaces are the main causes.233,245,259,268,383,384 Hori plex catalysts gradually lost their carbonyl-containing complexes,
et al.268 put forth several possible factors, which can be sum- and inactive species, such as [Ru(bpy)2(CO3)], were formed. The
marized as (1) heavy metal impurities contained in reagent instability of the [Cl(CO)2-(bpy*)Ru–Ru(bpy)(CO)2Cl] species
chemicals and introduced to the electrolyte solution; (2) very was confirmed by a voluminous black precipitate, produced by
small amounts of organic substances possibly contained in exhaustive electrolysis at 2.00 V.222 In a recent study of CO2
water, and (3) intermediate poisoning species or products formed reduction to CO at low overpotential in neutral aqueous
during CO2 reduction and adsorbed on electrodes. Besides these, solution by a Ni(cyclam) complex attached to poly(allylamine),
electrolysis mode and condition can also affect catalyst stabi- Saravanakumar et al.124 achieved a current efficiency of 92%
lity.142,297,385,386 For example, deactivation of a Cu cathode was during the initial 6 hours of electrolysis, but this dropped to
observed after only 3 hours of electrolysis in a constant potential 88% at 12 hours and then to 79% at 24 hours.
mode, while the electrocatalytic activity of Cu remained constant Hence, to mitigate the degradation of catalyst activity and
for 7 hours if a superimposed potential method was applied.269 stability, two major factors should be considered: (1) the effect
Using the latter, the surface structure of the copper electrode was
changed with the formation of cuprous oxide (Cu2O), then the
adsorption of amorphous graphite was prevented, leading to
stable long-term electrolysis for CH4 production.269 The pulse
electrolysis mode was also found to have a mitigating effect on
electrode deactivation.387 Changing the electrolytic conditions
led to the deposition of poisoning species on a Cu electrode
being highly suppressed, while the selectivity for C2H4 formation
was enhanced. Details of the reduction mechanism on Cu2O
remained unclear, and further study of the electrodeposition of
CO2 on the Cu2O cathode should be considered.
An early study found that the deactivation of a Sn electrode
was related to the formation of organometallic complexes on
the electrode surface, which could accelerate the rate of hydro- Fig. 27 Schematic representation of the electrochemical process to
gen evolution.321 Recently, Agarwal et al.8 investigated the long- convert CO2 into formate/formic acid. Reprinted with permission from
term performance of Sn, together with other proprietary ref. 8. Copyright r 2011 WILEY-VCH Verlag GmbH & Co. KGaA.

666 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

of catalyst type, structure, and composite, and (2) the effect of and Agarwal et al.8 on the engineering and economic feasibility
catalyst operating conditions. of CO2 electroreduction to produce formate/formic acid with a
planned conversion of 100 tonnes of CO2 per day---optimization
of the electrode/reactor and system design seems to be the
5. Technological challenges in CO2 second biggest challenge, next to low catalyst stability. There-
electrocatalytic reduction fore, efforts to optimize system designs and at the same time
develop durable catalysts should be carried out.
Regarding CO2 electrochemical reduction technology, several
technical challenges remain, mainly (1) low catalyst activity, (2) low
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

product selectivity, and (3) insufficient stability. In terms of the 6. Summary and proposed research
practical application of CO2 reduction to produce usable low-carbon directions
fuels, our technology still seems to be far from adequate. Particularly
for industrial-scale implementation (Fig. 27), low catalyst stability To facilitate the research and development of CO2 electro-
seems to be the major limitation at present. Therefore, developing reduction, this paper gives a comprehensive overview of several
highly active, selective, and stable electrocatalysts for the reduction decades of development and recent trends in the catalysis of
of CO2 is still the major focus of activity in this area. The several CO2 electroreduction. Various electrocatalysts explored and
challenges can be summarized as follows: reported in the literature are summarized and classified into
(1) Low catalyst activity. It can be seen from the above metals, metal complexes, and organic/bio-organic compounds.
discussion that for every single type of catalyst developed in The composition of electrocatalysts, catalytic mechanisms of CO2
the literature, the overpotential for CO2 electroreduction is reduction, conditions of use (medium, electrolyte, concentration,
normally too high, indicating that these catalysts’ activities electrolysis potential, temperature, and pressure), current or
are still not good enough for practical applications in terms Faradaic efficiency, and product selectivity are reviewed in depth
of energy efficiency. with respect to catalyst activity and stability. Some typical cata-
(2) Low product selectivity. Although some of the catalysts lysts and their associated data for catalytic activity, product
discussed above gave desirable product selectivity and stable selectivity, and catalytic stability are summarized and presented
yields under continuous operation, so far only a few attempts in tables to help readers quickly locate the information they are
have resulted in selective production—for instance, the use of looking for.
Sn, Pb, and Hg metal electrodes to produce formate/formic Furthermore, industrial attempts to scale up the technology of
acid.8,18,264,330,331 Unfortunately, the stability of the catalysis CO2 electroreduction for practical applications, and achievements
processes on these electrodes/catalysts is too low to be practical. in this area over the past several decades, are also discussed to
Although other types of catalysts, such as metal complexes, give a clear picture of the current state of technology.
have been explored, and some high catalytic selectivity has been It seems that the maturity of CO2 electroreduction technol-
achieved, they suffer from the same low stability. For the ogy to produce low-carbon fuels is still far from reaching the
majority of the catalysts explored, even if they show high requirements for commercialization, due to several major
activity, the product selectivities are low, with some undesired technological challenges, including low catalyst activity, low
products also resulting. product selectivity, and insufficient catalyst stability. To overcome
(3) Insufficient catalyst stability/durability. This is probably these challenges, we propose several future research directions:
the single biggest challenge. In the literature, the normally (1) Enhancement of catalytic activity and stability by exploring
reported stability tests are in the region of under 100 hours, innovative electrocatalysts. Generally speaking, almost all of the
while long-term tests have yet to be done. As discussed above, possible pure metals and their associated compounds that
as the CO2 reduction reaction proceeds, the active electrode/ show some catalytic activity toward CO2 electroreduction have
catalyst surface can gradually become covered by reaction been tried as electrocatalysts, and some progress has been
intermediates and by-products (such as carbon films and achieved in terms of catalytic activity, product selectivity, and
poisonous species), blocking and poisoning the catalyst’s active catalytic stability. However, these technological advances are
sites and leading to rapid catalytic activity degradation.330 still not sufficient for practical applications. Breakthroughs in
(4) Insufficient fundamental understanding. The literature catalysis are definitely desirable and, indeed, necessary. With
contains attempts to fundamentally understand the CO2 reduction respect to these, developing new material synthesis technology
process on catalysts through both experimental and theoretical to give innovative new catalysts with optimal performance is
modeling approaches to catalyst downselection with respect to the priority. Two important types of catalyst materials should
catalyst activity, new catalyst design, and catalyst operation opti- be emphasized here: (i) composite catalyst materials, which are
mization. However, the work in this area seems to be insufficient. synthesized by combining several different materials together,
(5) Non-optimized electrode/reactor and system design for and (ii) nanostructured catalyst materials. Composite catalyst
practical applications. The scale-up of the electroreduction of materials should have different properties and catalytic perfor-
CO2 for practical applications is a necessary step toward the mance than their individual components because the indivi-
success of this technology.263 Although there have been some dual substances in the composites experience a synergistic
attempts in this respect—such as the work by Li and Oloman18 effect; this comes about through optimizing particle size,

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 667
View Article Online

Review Article Chem Soc Rev

specific surface area, porosity, and active sites, preventing undesirable CO2 to solve our energy shortage issues. The
particles from agglomerating, facilitating electron and proton authors of this article deeply believe that with continued and
conduction, and protecting active materials from chemical and extensive efforts focused on developing innovative composite
mechanical degradation. As a result, the obtained composites and nanostructured catalyst materials to overcome the chal-
may have high catalytic activity, high product selectivity, and lenges of insufficient catalytic activity, product selectivity, and
high catalytic stability in CO2 electroreduction. Regarding catalytic stability, the technology of CO2 electroreduction will
nanostructured catalyst materials, such as nanoaerogels, nano- become practical in the near future.
tubes/rods, nanoplates/sheets, nanospheres, and so on, their
unique properties—such as high specific surface area and 1-,
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

2-, and 3-dimensional structures—can provide easy transport/ Acknowledgements


diffusion pathways for substrates to access, leading to faster
This investigation was financially supported by the National
kinetics, more efficient contact for electrolyte ions, and more
Natural Science Foundation of China (grant no. 21173039), the
active sites for the catalytic process. Furthermore, these nanos-
Specialized Research Fund for the Doctoral Program of Higher
tructured materials can be used as catalyst supports to provide
Education, SRFD (20110075110001); the Innovation Program of
a synergistic effect between the catalyst and the support parti-
the Shanghai Municipal Education Commission, the Program
cles, resulting in highly active and stable electrocatalysts for
for New Century Excellent Talents in University (NCET-12-
CO2 electroreduction.
0828), the Science and Technology Commission of Shanghai
(2) Further fundamental understanding through both experi-
Municipality (11230700600), the Fundamental Research Funds
ments and theoretical modeling. For down-selecting catalysts,
for the Central Universities, the State Environmental Protection
designing and optimizing new catalyst structures with respect
Engineering Center for Pollution Treatment and Control in the
to improving catalytic activity, product selectivity, and catalytic
Textile Industry of China and Sendai Kankyo Kaihatsu Corpora-
stability, better fundamental understanding through both
tion, Japan. All of the above financial support is gratefully
experiments and theoretical modeling is necessary. For exam-
acknowledged.
ple, we need to fundamentally understand the mechanisms of
CO2 electroreduction and their relationship to catalyst active
site structures and composition, using both theoretical calcula- References
tions (molecular/electronic-level modeling) and experimental
approaches, to guide new catalyst development. To mitigate 1 J. Sawyer, Nature, 1972, 239, 2.
catalyst degradation, it is necessary to understand the degrada- 2 T. Volk, CO2 rising: the world’s greatest environmental
tion mechanisms and failure modes, which can be done using challenge, MIT Press, 2008.
both experimental and theoretical modeling approaches. For 3 G. A. Olah, A. Goeppert and G. K. S. Prakash, J. Org. Chem.,
instance, a variety of instrumental analysis methods (e.g., SEM, 2009, 74, 487–498.
TEM, XRD, XPS, NMR, HPLC, GC, and so on) and electro- 4 N. S. Spinner, J. A. Vega and W. E. Mustain, Catal. Sci.
chemical methods (CV, RDE/RRDE, and EIS) can be used to Technol., 2012, 2, 19–28.
characterize catalysts before and after lifetime tests. With 5 D. P. Schrag, Science, 2007, 315, 812–813.
greater understanding, it should be possible to develop new 6 D. D. Yuan, C. H. Yan, B. Lu, H. X. Wang, C. M. Zhong and
mitigation strategies. Q. H. Cai, Electrochim. Acta, 2009, 54, 2912–2915.
(3) Optimizing electrodes, reactors, and system designs for 7 D. T. Whipple and P. J. A. Kenis, J. Phys. Chem. Lett., 2010,
practical applications. In the current state of technology for 1, 3451–3458.
industrial-scale CO2 electroreduction to produce low-carbon 8 A. S. Agarwal, Y. M. Zhai, D. Hill and N. Sridhar, ChemSusChem,
fuels, the major limitations seemed to be rapid catalyst degra- 2011, 4, 1301–1310.
dation and slow CO2 transfer to the electrode surface, both of 9 T. J. Meyer, Acc. Chem. Res., 1989, 22, 163–170.
which should be partially related to the electrode/reactor 10 W. Leitner, Coord. Chem. Rev., 1996, 153, 257–284.
design. It is believed that catalyst degradation is mainly related 11 M. Cheng, E. B. Lobkovsky and G. W. Coates, J. Am. Chem.
to the material itself but also to its operating environment. Soc., 1998, 120, 11018–11019.
Besides improving the catalyst material’s stability, improving 12 J. H. Alstrum-Acevedo, M. K. Brennaman and T. J. Meyer,
the electrode/reactor design to optimize the operating condi- Inorg. Chem., 2005, 44, 6802–6827.
tions is also important for performance. For example, certain 13 M. Aresta and A. Dibenedetto, J. Chem. Soc., Dalton Trans.,
innovative designs for GDEs and electrolyte membrane-based 2007, 2975–2992.
electrochemical cells developed in recent years seem to be the 14 G. Centi and S. Perathoner, Catal. Today, 2009, 148, 191–205.
right approaches for reducing internal resistance and improv- 15 I. Omae, Coord. Chem. Rev., 2012, 256, 1384–1405.
ing the reactant mass transfer process. 16 J. J. Concepcion, R. L. House, J. M. Papanikolas and T. J.
In summary, CO2 electroreduction to produce low-carbon Meyer, Proc. Natl. Acad. Sci. U. S. A., 2012, 109, 15560–15564.
fuels is an important research and development subject for 17 T. Sakakura, J. C. Choi and H. Yasuda, Chem. Rev., 2007,
overcoming excessive, environmentally harmful CO2 emissions; 107, 2365–2387.
it also presents the option of producing useful fuels using 18 C. Oloman and H. Li, ChemSusChem, 2008, 1, 385–391.

668 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

19 E. E. Benson, C. P. Kubiak, A. J. Sathrum and J. M. Smieja, 48 B. X. Wang and S. J. Dong, J. Electroanal. Chem., 1994, 379,
Chem. Soc. Rev., 2009, 38, 89–99. 207–214.
20 J. Lee, Y. Kwon, R. L. Machunda and H. J. Lee, Chem.–Asian 49 M. Bertotti and D. Pletcher, Electroanalysis, 1996, 8,
J., 2009, 4, 1516–1523. 1105–1111.
21 C. D. Windle and R. N. Perutz, Coord. Chem. Rev., 2012, 50 M. Bertotti and D. Pletcher, Quim. Nova, 1998, 21, 167–171.
256, 2562–2570. 51 L. Kosminsky and M. Bertotti, J. Electroanal. Chem., 1999,
22 C. Finn, S. Schnittger, L. J. Yellowlees and J. B. Love, Chem. 471, 37–41.
Commun., 2012, 48, 1392–1399. 52 L. Kosminsky and M. Bertotti, Electroanalysis, 1999, 11,
23 J. L. Inglis, B. J. MacLean, M. T. Pryce and J. G. Vos, Coord. 623–626.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

Chem. Rev., 2012, 256, 2571–2600. 53 V. S. Saji and C. W. Lee, ChemSusChem, 2012, 5, 1146–1161.
24 J. Schneider, H. F. Jia, J. T. Muckerman and E. Fujita, 54 M. Nakazawa, Y. Mizobe, Y. Matsumoto, Y. Uchida,
Chem. Soc. Rev., 2012, 41, 2036–2051. M. Tezuka and M. Hidai, Bull. Chem. Soc. Jpn., 1986, 59,
25 K. Mori, H. Yamashita and M. Anpo, Rsc. Adv., 2012, 2, 809–814.
3165–3172. 55 K. Ogura and I. Yoshida, J. Mol. Catal., 1988, 47,
26 M. Aresta, Carbon dioxide recovery and utilization, Springer, 51–57.
2003. 56 K. T. Potts, D. A. Usifer, A. Guadalupe and H. D. Abruna,
27 M. M. Halmann, Chemical Fixation of Carbon Dioxide- J. Am. Chem. Soc., 1987, 109, 3961–3967.
Methods for Recycling CO2 into Useful Products, CRC, 1993. 57 J. A. R. Sende, C. R. Arana, L. Hernandez, K. T. Potts,
28 B. P. Sullivan, K. Krist and H. Guard, Electrochemical M. Keshevarzk and H. D. Abruna, Inorg. Chem., 1995, 34,
and electrocatalytic reactions of carbon dioxide, Elsevier, 3339–3348.
Amsterdam etc., 1993. 58 T. Reda, C. M. Plugge, N. J. Abram and J. Hirst, Proc. Natl.
29 C. S. Song, Catal. Today, 2006, 115, 2–32. Acad. Sci. U. S. A., 2008, 105, 10654–10658.
30 A. J. Bard, R. Parsons and J. Jordan, Standard potentials in 59 M. Tiberti, E. Papaleo, N. Russo, L. De Gioia and
aqueous solutions, CRC press, 1985. G. Zampella, Inorg. Chem., 2012, 51, 8331–8339.
31 E. Lamy, L. Nadjo and J. M. Saveant, J. Electroanal. Chem., 60 J. Hawecker, J. M. Lehn and R. Ziessel, J. Chem. Soc., Chem.
1977, 78, 403–407. Commun., 1983, 536–538.
32 M. Hammouche, D. Lexa, M. Momenteau and J. M. Saveant, 61 J. Hawecker, J. M. Lehn and R. Ziessel, J. Chem. Soc., Chem.
J. Am. Chem. Soc., 1991, 113, 8455–8466. Commun., 1984, 328–330.
33 I. Bhugun, D. Lexa and J. M. Saveant, J. Phys. Chem., 1996, 62 J. Hawecker, J. M. Lehn and R. Ziessel, Helv. Chim. Acta,
100, 19981–19985. 1986, 69, 1990–2012.
34 K. Y. Wong, W. H. Chung and C. P. Lau, J. Electroanal. 63 T. R. O’Toole, L. D. Margerum, T. D. Westmoreland,
Chem., 1998, 453, 161–169. W. J. Vining, R. W. Murray and T. J. Meyer, J. Chem. Soc.,
35 J. M. Saveant, Chem. Rev., 2008, 108, 2348–2378. Chem. Commun., 1985, 1416–1417.
36 M. R. DuBois and D. L. DuBois, Acc. Chem. Res., 2009, 42, 64 C. R. Cabrera and H. D. Abruna, J. Electroanal. Chem., 1986,
1974–1982. 209, 101–107.
37 C. Costentin, M. Robert and J. M. Saveant, Chem. Soc. Rev., 65 S. Cosnier, A. Deronzier and J. C. Moutet, J. Electroanal.
2013, 42, 2423–2436. Chem., 1986, 207, 315–321.
38 H. Noda, S. Ikeda, Y. Oda, K. Imai, M. Maeda and K. Ito, 66 S. Cosnier, A. Deronzier and J. C. Moutet, J. Mol. Catal.,
Bull. Chem. Soc. Jpn., 1990, 63, 2459–2462. 1988, 45, 381–391.
39 A. Fujishima, X. T. Zhang and D. A. Tryk, Surf. Sci. Rep., 67 T. R. O’Toole, B. P. Sullivan, M. R. M. Bruce, L. D.
2008, 63, 515–582. Margerum, R. W. Murray and T. J. Meyer, J. Electroanal.
40 A. Monnier, J. Augustynski and C. Stalder, J. Electroanal. Chem., 1989, 259, 217–239.
Chem. Interfacial Electrochem., 1980, 112, 383–385. 68 R. Schrebler, P. Cury, F. Herrera, H. Gomez and
41 M. Koudelka, A. Monnier and J. Augustynski, J. Electrochem. R. Cordova, J. Electroanal. Chem., 2001, 516, 23–30.
Soc., 1984, 131, 745–750. 69 F. Cecchet, M. Alebbi, C. A. Bignozzi and F. Paolucci, Inorg.
42 A. Bandi, J. Electrochem. Soc., 1990, 137, 2157–2160. Chim. Acta, 2006, 359, 3871–3874.
43 L. F. Cueto, G. A. Hirata and E. M. Sanchez, J. Sol–Gel Sci. 70 K. C. Cheung, P. Guo, M. H. So, L. Y. S. Lee, K. P. Ho,
Technol., 2006, 37, 105–109. W. L. Wong, K. H. Lee, W. T. Wong, Z. Y. Zhou and
44 D. Chu, G. X. Qin, X. M. Yuan, M. Xu, P. Zheng and J. Lu, K. Y. Wong, J. Organomet. Chem., 2009, 694, 2842–2845.
ChemSusChem, 2008, 1, 205–209. 71 J. M. Smieja and C. P. Kubiak, Inorg. Chem., 2010, 49,
45 T. Mizuno, A. Naitoh and K. Ohta, J. Electroanal. Chem., 9283–9289.
1995, 391, 199–201. 72 E. Portenkirchner, K. Oppelt, C. Ulbricht, D. A. M. Egbe,
46 J. P. Qu, X. G. Zhang, Y. G. Wang and C. X. Xie, Electrochim. H. Neugebauer, G. Knor and N. S. Sariciftci, J. Organomet.
Acta, 2005, 50, 3576–3580. Chem., 2012, 716, 19–25.
47 D. P. Summers, S. Leach and K. W. Frese, J. Electroanal. 73 B. P. Sullivan, C. M. Bolinger, D. Conrad, W. J. Vining and
Chem., 1986, 205, 219–232. T. J. Meyer, J. Chem. Soc., Chem. Commun., 1985, 1414–1415.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 669
View Article Online

Review Article Chem Soc Rev

74 F. P. A. Johnson, M. W. George, F. Hartl and J. J. Turner, 101 C. M. Lieber and N. S. Lewis, J. Am. Chem. Soc., 1984, 106,
Organometallics, 1996, 15, 3374–3387. 5033–5034.
75 T. Scheiring, A. Klein and W. Kaim, J. Chem. Soc., Perkin 102 S. Kapusta and N. Hackerman, J. Electrochem. Soc., 1984,
Trans. 2, 1997, 2569–2571. 131, 1511–1514.
76 C. L. Anfuso, R. C. Snoeberger, A. M. Ricks, W. M. Liu, 103 T. Yoshida, K. Kamato, M. Tsukamoto, T. Iida,
D. Q. Xiao, V. S. Batista and T. Q. Lian, J. Am. Chem. Soc., D. Schlettwein, D. Wohrle and M. Kaneko, J. Electroanal.
2011, 133, 6922–6925. Chem., 1995, 385, 209–225.
77 M. Bourrez, F. Molton, S. Chardonnoblat and A. Deronzier, 104 T. Abe, T. Yoshida, S. Tokita, F. Taguchi, H. Imaya and
Angew. Chem., Int. Ed., 2011, 50, 9903–9906. M. Kaneko, J. Electroanal. Chem., 1996, 412, 125–132.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

78 J. M. Smieja, M. D. Sampson, K. A. Grice, E. E. Benson, J. D. 105 T. Abe, F. Taguchi, T. Yoshida, S. Tokita, G. Schnurpfeil,
Froehlich and C. P. Kubiak, Inorg. Chem., 2013, 52, 2484–2491. D. Wohrle and M. Kaneko, J. Mol. Catal. A: Chem., 1996,
79 K. Hara, A. Kudo and T. Sakata, J. Electroanal. Chem., 1995, 112, 55–61.
386, 257–260. 106 J. Zhang, W. J. Pietro and A. B. P. Lever, J. Electroanal.
80 O. Koga and Y. Hori, Electrochim. Acta, 1993, 38, 1391–1394. Chem., 1996, 403, 93–100.
81 A. Kudo, S. Nakagawa, A. Tsuneto and T. Sakata, 107 H. Z. Zhao, Y. Zhang, B. Zhao, Y. Y. Chang and Z. S. Li,
J. Electrochem. Soc., 1993, 140, 1541–1545. Environ. Sci. Technol., 2012, 46, 5198–5204.
82 O. Koga, T. Matsuo, H. Yamazaki and Y. Hori, Bull. Chem. 108 M. Isaacs, J. C. Canales, M. J. Aguirre, G. Estiu, F. Caruso,
Soc. Jpn., 1998, 71, 315–320. G. Ferraudi and J. Costamagna, Inorg. Chim. Acta, 2002,
83 K. Takahashi, K. Hiratsuka, H. Sasaki and S. Toshima, 339, 224–232.
Chem. Lett., 1979, 305–308. 109 M. Isaacs, J. C. Canales, A. Riquelme, M. Lucero,
84 I. Bhugun, D. Lexa and J. M. Saveant, J. Am. Chem. Soc., M. J. Aguirre and J. Costamagna, J. Coord. Chem., 2003,
1994, 116, 5015–5016. 56, 1193–1201.
85 I. Bhugun, D. Lexa and J. M. Saveant, J. Am. Chem. Soc., 110 J. Grodkowski, P. Neta, E. Fujita, A. Mahammed, L. Simkhovich
1996, 118, 1769–1776. and Z. Gross, J. Phys. Chem. A, 2002, 106, 4772–4778.
86 C. Costentin, S. Drouet, M. Robert and J. M. Saveant, 111 M. Beley, J. P. Collin, R. Ruppert and J. P. Sauvage, J. Chem.
Science, 2012, 338, 90–94. Soc., Chem. Commun., 1984, 1315–1316.
87 H. Z. Zhao, Y. Y. Chang and C. Liu, J. Solid State Electro- 112 M. Beley, J. P. Collin, R. Ruppert and J. P. Sauvage, J. Am.
chem., 2013, 17, 1657–1664. Chem. Soc., 1986, 108, 7461–7467.
88 K. Leung, I. M. B. Nielsen, N. Sai, C. Medforth and 113 M. Fujihira, Y. Hirata and K. Suga, J. Electroanal. Chem.,
J. A. Shelnutt, J. Phys. Chem. A, 2010, 114, 10174–10184. 1990, 292, 199–215.
89 T. Atoguchi, A. Aramata, A. Kazusaka and M. Enyo, 114 S. Sakaki, J. Am. Chem. Soc., 1992, 114, 2055–2062.
J. Electroanal. Chem., 1991, 318, 309–320. 115 G. B. Balazs and F. C. Anson, J. Electroanal. Chem., 1992,
90 H. Tanaka and A. Aramata, J. Electroanal. Chem., 1997, 437, 322, 325–345.
29–35. 116 G. B. Balazs and F. C. Anson, J. Electroanal. Chem., 1993,
91 T. V. Magdesieva, T. Yamamoto, D. A. Tryk and 361, 149–157.
A. Fujishima, J. Electrochem. Soc., 2002, 149, D89–D95. 117 Y. Hirata, K. Suga and M. Fujihira, Chem. Lett., 1990, 1155–1158.
92 O. Enoki, T. Imaoka and K. Yamamoto, Macromol. Symp., 118 J. P. Collin, A. Jouaiti and J. P. Sauvage, Inorg. Chem., 1988,
2003, 204, 151–158. 27, 1986–1990.
93 T. Imaoka, R. Tanaka and K. Yamamoto, J. Polym. Sci., 119 R. W. Hay, J. A. Crayston, T. J. Cromie, P. Lightfoot and
Part A: Polym. Chem., 2006, 44, 5229–5236. D. C. L. deAlwis, Polyhedron, 1997, 16, 3557–3563.
94 S. A. Yao, R. E. Ruther, L. H. Zhang, R. A. Franking, R. J. Hamers 120 J. Schneider, H. F. Jia, K. Kobiro, D. E. Cabelli, J. T.
and J. F. Berry, J. Am. Chem. Soc., 2012, 134, 15632–15635. Muckerman and E. Fujita, Energy Environ. Sci., 2012, 5,
95 D. Behar, T. Dhanasekaran, P. Neta, C. M. Hosten, D. Ejeh, 9502–9510.
P. Hambright and E. Fujita, J. Phys. Chem. A, 1998, 102, 121 J. D. Froehlich and C. P. Kubiak, Inorg. Chem., 2012, 51,
2870–2877. 3932–3934.
96 J. Y. Becker, B. Vainas, R. Eger and L. Kaufman, J. Chem. 122 F. Abba, G. Desantis, L. Fabbrizzi, M. Licchelli, A. M. M.
Soc., Chem. Commun., 1985, 1471–1472. Lanfredi, P. Pallavicini, A. Poggi and F. Ugozzoli, Inorg.
97 N. Sonoyama, M. Kirii and T. Sakata, Electrochem. Commun., Chem., 1994, 33, 1366–1375.
1999, 1, 213–216. 123 A. Jarzebinska, P. Rowinski, I. Zawisza, R. Bilewicz,
98 W. Hieringer, K. Flechtner, A. Kretschmann, K. Seufert, L. Siegfried and T. Kaden, Anal. Chim. Acta, 1999, 396, 1–12.
W. Auwarter, J. V. Barth, A. Gorling, H. P. Steinruck and 124 D. Saravanakumar, J. Song, N. Jung, H. Jirimali and
J. M. Gottfried, J. Am. Chem. Soc., 2011, 133, 6206–6222. W. Shin, ChemSusChem, 2012, 5, 634–636.
99 K. Hiratsuka, K. Takahashi, H. Sasaki and S. Toshima, 125 B. Fisher and R. Eisenberg, J. Am. Chem. Soc., 1980, 102,
Chem. Lett., 1977, 1137–1140. 7361–7363.
100 S. Meshitsu, M. Ichikawa and K. Tamaru, J. Chem. Soc., 126 E. Fujita, J. Haff, R. Sanzenbacher and H. Elias, Inorg.
Chem. Commun., 1974, 158–159. Chem., 1994, 33, 4627–4628.

670 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

127 K. Bujno, R. Bilewicz, L. Siegfried and T. Kaden, 154 K. Ogura and H. Uchida, J. Electroanal. Chem., 1987, 220,
J. Electroanal. Chem., 1996, 407, 131–140. 333–337.
128 M. Rudolph, S. Dautz and E. G. Jager, J. Am. Chem. Soc., 155 K. Ogura, J. Electrochem. Soc., 1987, 134, 2749–2754.
2000, 122, 10821–10830. 156 K. Ogura, C. T. Migita and H. Imura, J. Electrochem. Soc.,
129 A. R. Guadalupe, D. A. Usifer, K. T. Potts, H. C. Hurrell, 1990, 137, 1730–1732.
A. E. Mogstad and H. D. Abruna, J. Am. Chem. Soc., 1988, 157 K. Ogura, C. T. Migita and K. Wadaka, J. Mol. Catal., 1991,
110, 3462–3466. 67, 161–173.
130 H. C. Hurrell, A. L. Mogstad, D. A. Usifer, K. T. Potts and 158 K. Ogura, K. Mine, J. Yano and H. Sugihara, J. Chem. Soc.,
H. D. Abruna, Inorg. Chem., 1989, 28, 1080–1084. Chem. Commun., 1993, 20–21.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

131 C. Arana, S. Yan, M. Keshavarzk, K. T. Potts and 159 K. Ogura, H. Sugihara, J. Yano and M. Higasa,
H. D. Abruna, Inorg. Chem., 1992, 31, 3680–3682. J. Electrochem. Soc., 1994, 141, 419–424.
132 C. Arana, M. Keshavarz, K. T. Potts and H. D. Abruna, 160 K. Ogura, M. Nakayama and C. Kusumoto, J. Electrochem.
Inorg. Chim. Acta, 1994, 225, 285–295. Soc., 1996, 143, 3606–3615.
133 H. Aga, A. Aramata and Y. Hisaeda, J. Electroanal. Chem., 161 K. Ogura, N. Endo, M. Nakayama and H. Ootsuka,
1997, 437, 111–118. J. Electrochem. Soc., 1995, 142, 4026–4032.
134 G. Chiericato, C. R. Arana, C. Casado, I. Cuadrado and 162 M. Nakayama, M. Iino and K. Ogura, J. Electroanal. Chem.,
H. D. Abruna, Inorg. Chim. Acta, 2000, 300, 32–42. 1997, 440, 251–257.
135 S.-N. Pun, W.-H. Chung, K.-M. Lam, P. Guo, P.-H. Chan, 163 P. Braunstein, D. Matt and D. Nobel, Chem. Rev., 1988, 88,
K.-Y. Wong, C.-M. Che, T.-Y. Chen and S.-M. Peng, J. Chem. 747–764.
Soc., Dalton. Trans., 2002, 575–583. 164 K. S. Ratliff, R. E. Lentz and C. P. Kubiak, Organometallics,
136 D. J. Pearce and D. Pletcher, J. Electroanal. Chem., 1986, 1992, 11, 1986–1988.
197, 317–330. 165 C. de Alwis, J. A. Crayston, T. Cromie, T. Eisenblatter,
137 A. A. Isse, A. Gennaro, E. Vianello and C. Floriani, J. Mol. R. W. Hay, Y. D. Lampeka and L. V. Tsymbal, Electrochim.
Catal., 1991, 70, 197–208. Acta, 2000, 45, 2061–2074.
138 J. Losada, I. Delpeso, L. Beyer, J. Hartung, V. Fernandez 166 E. Simón-Manso and C. P. Kubiak, Organometallics, 2005,
and M. Mobius, J. Electroanal. Chem., 1995, 398, 89–93. 24, 96–102.
139 B. A. Averill, T. Herskovi, R. H. Holm and J. A. Ibers, J. Am. 167 E. Y. Lee, D. W. Hong, H. W. Park and M. P. Suh, Eur. J.
Chem. Soc., 1973, 95, 3523–3534. Inorg. Chem., 2003, 3242–3249.
140 M. Tezuka, T. Yajima, A. Tsuchiya, Y. Matsumoto, 168 J. S. Field, R. J. Haines, C. J. Parry and S. H. Sookraj,
Y. Uchida and M. Hidai, J. Am. Chem. Soc., 1982, 104, Polyhedron, 1993, 12, 2425–2428.
6834–6836. 169 R. J. Haines, R. E. Wittrig and C. P. Kubiak, Inorg. Chem.,
141 B. D. Yuhas, C. Prasittichai, J. T. Hupp and M. G. Kanatzidis, 1994, 33, 4723–4728.
J. Am. Chem. Soc., 2011, 133, 15854–15857. 170 G. N. A. Nallas and K. J. Brewer, Inorg. Chim. Acta, 1996,
142 M. Le, M. Ren, Z. Zhang, P. T. Sprunger, R. L. Kurtz and 253, 7–13.
J. C. Flake, J. Electrochem. Soc., 2011, 158, E45–E49. 171 J. W. Raebiger, J. W. Turner, B. C. Noll, C. J. Curtis,
143 M. Gangeri, S. Perathoner, S. Caudo, G. Centi, J. Amadou, A. Miedaner, B. Cox and D. L. DuBois, Organometallics,
D. Begin, C. Pham-Huu, M. J. Ledoux, J. P. Tessonnier, 2006, 25, 3345–3351.
D. S. Su and R. Schlogi, Catal. Today, 2009, 143, 57–63. 172 R. Angamuthu, P. Byers, M. Lutz, A. L. Spek and
144 C. Ampelli, G. Centi, R. Passalacqua and S. Perathoner, E. Bouwman, Science, 2010, 327, 313–315.
Energy Environ. Sci., 2010, 3, 292–301. 173 W. Shin, S. Lee, J. Shin, S. Lee and Y. Kim, J. Am. Chem.
145 R. Arrigo, M. E. Schuster, S. Wrabetz, F. Girgsdies, Soc., 2003, 125, 14688–14689.
J. P. Tessonnier, G. Centi, S. Perathoner, D. S. Su and 174 S. W. Ragsdale, Crit. Rev. Biochem. Mol. Biol., 2004, 39,
R. Schlogl, ChemSusChem, 2012, 5, 577–586. 165–195.
146 G. Centi and S. Perathoner, ChemSusChem, 2010, 3, 195–208. 175 A. Parkin, J. Seravalli, K. A. Vincent, S. W. Ragsdale and
147 A. B. Bocarsly, Q. D. Gibson, A. J. Morris, R. P. L’Esperance, F. A. Armstrong, J. Am. Chem. Soc., 2007, 129, 10328–10329.
Z. M. Detweiler, P. S. Lakkaraju, E. L. Zeitler and 176 Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Electro-
T. W. Shaw, ACS Catal., 2012, 2, 1684–1692. chim. Acta, 1994, 39, 1833–1839.
148 M. D. Rail and L. A. Berben, J. Am. Chem. Soc., 2011, 133, 177 J. P. Popic, M. L. AvramovIvic and N. B. Vukovic,
18577–18579. J. Electroanal. Chem., 1997, 421, 105–110.
149 K. Ogura and S. Yamasaki, J. Chem. Soc., Faraday Trans. 1, 178 K. W. Frese and S. Leach, J. Electrochem. Soc., 1985, 132,
1985, 81, 267–271. 259–260.
150 K. Ogura and M. Kaneko, J. Mol. Catal., 1985, 31, 49–56. 179 M. Lukaszewski and A. Czerwinski, J. Solid State Electrochem.,
151 K. Ogura and K. Takamagari, J. Chem. Soc., Dalton. Trans., 2007, 11, 339–349.
1986, 1519–1523. 180 M. Lukaszewski, H. Siwek and A. Czerwinski, J. Solid State
152 K. Ogura and I. Yoshida, J. Mol. Catal., 1986, 34, 67–72. Electrochem., 2009, 13, 813–827.
153 K. Ogura and M. Fujita, J. Mol. Catal., 1987, 41, 303–311. 181 D. W. McKee, J. Catal., 1967, 8, 240–249.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 671
View Article Online

Review Article Chem Soc Rev

182 M. Spichigerulmann and J. Augustynski, J. Chem. Soc., 209 A. Bandi and H. M. Kühne, J. Electrochem. Soc., 1992, 139,
Faraday Trans. 1, 1985, 81, 713–716. 1605–1610.
183 M. Azuma, K. Hashimoto, M. Watanabe and T. Sakata, 210 N. Spataru, K. Tokuhiro, C. Terashima, T. N. Rao and
J. Electroanal. Chem., 1990, 294, 299–303. A. Fujishima, J. Appl. Electrochem., 2003, 33, 1205–1210.
184 S. Nakagawa, A. Kudo, M. Azuma and T. Sakata, 211 C. M. Bolinger, B. P. Sullivan, D. Conrad, J. A. Gilbert,
J. Electroanal. Chem., 1991, 308, 339–343. N. Story and T. J. Meyer, J. Chem. Soc., Chem. Commun.,
185 K. Ohkawa, K. Hashimoto, A. Fujishima, Y. Noguchi and 1985, 796–797.
S. Nakayama, J. Electroanal. Chem., 1993, 345, 445–456. 212 Z. F. Chen, C. C. Chen, D. R. Weinberg, P. Kang,
186 K. Ohkawa, Y. Noguchi, S. Nakayama, K. Hashimoto and J. J. Concepcion, D. P. Harrison, M. S. Brookhart and
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

A. Fujishima, J. Electroanal. Chem., 1993, 348, 459–464. T. J. Meyer, Chem. Commun., 2011, 47, 12607–12609.
187 K. Ohkawa, Y. Noguchi, S. Nakayama, K. Hashimoto and 213 P. Paul, Proc. Indian Acad.-Chem. Sci., 2002, 114,
A. Fujishima, J. Electroanal. Chem., 1994, 367, 165–173. 269–276.
188 B. I. Podlovchenko, E. A. Kolyadko and S. G. Lu, 214 S. Slater and J. H. Wagenknecht, J. Am. Chem. Soc., 1984,
J. Electroanal. Chem., 1994, 373, 185–187. 106, 5367–5368.
189 K. Ohkawa, Y. Noguchi, S. Nakayama, K. Hashimoto and 215 D. L. DuBois and A. Miedaner, J. Am. Chem. Soc., 1987, 109,
A. Fujishima, J. Electroanal. Chem., 1994, 369, 247–250. 113–117.
190 B. R. Eggins and J. McNeill, J. Electroanal. Chem., 1983, 148, 216 D. L. DuBois, A. Miedaner and R. C. Haltiwanger, J. Am.
17–24. Chem. Soc., 1991, 113, 8753–8764.
191 G. M. Brisard, A. P. M. Camargo, F. C. Nart and T. Iwasita, 217 B. D. Steffey, A. Miedaner, M. L. Maciejewskifarmer,
Electrochem. Commun., 2001, 3, 603–607. P. R. Bernatis, A. M. Herring, V. S. Allured, V. Carperos
192 K. Hara, A. Kudo, T. Sakata and M. Watanabe, and D. L. DuBois, Organometallics, 1994, 13, 4844–4855.
J. Electrochem. Soc., 1995, 142, L57–L59. 218 H. Ishida, H. Tanaka, K. Tanaka and T. Tanaka, J. Chem.
193 Y. Tomita and Y. Hori, Stud. Surf. Sci. Catal., 1998, 114, Soc., Chem. Commun., 1987, 131–132.
581–584. 219 H. Ishida, K. Tanaka and T. Tanaka, Organometallics, 1987,
194 Y. Hori and Y. Tomita, Abstr. Pap. Am. Chem. Soc., 1998, 6, 181–186.
215, U405–U406. 220 J. R. Pugh, M. R. M. Bruce, B. P. Sullivan and T. J. Meyer,
195 Y. Tomita, S. Teruya, O. Koga and Y. Hori, J. Electrochem. Inorg. Chem., 1991, 30, 86–91.
Soc., 2000, 147, 4164–4167. 221 S. Chardonnoblat, M. N. Collombdunandsauthier,
196 G. Centi, S. Perathoner, G. Wine and M. Gangeri, Green A. Deronzier, R. Ziessel and D. Zsoldos, Inorg. Chem.,
Chem., 2007, 9, 671–678. 1994, 33, 4410–4412.
197 Q. J. Feng, S. Q. Liu, X. Y. Wang and G. H. Jin, Appl. Surf. 222 S. Chardonnoblat, A. Deronzier, R. Ziessel and D. Zsoldos,
Sci., 2012, 258, 5005–5009. Inorg. Chem., 1997, 36, 5384–5389.
198 B. X. Hu, V. Stancovski, M. Morton and S. L. Suib, Appl. 223 M. N. Collombdunandsauthier, A. Deronzier and
Catal., A, 2010, 382, 277–283. R. Ziessel, Inorg. Chem., 1994, 33, 2961–2967.
199 N. R. de Tacconi, W. Chanmanee, B. H. Dennis, F. M. 224 Md. M. Ali, H. Sato, T. Mizukawa, K. Tsuge, M. Haga and
MacDonnell, D. J. Boston and K. Rajeshwar, Electrochem. K. Tanaka, Chem. Commun., 1998, 249–250.
Solid-State Lett., 2012, 15, B5–B8. 225 K. Tanaka and T. Mizukawa, Appl. Organomet. Chem., 2000,
200 X. Yuan, B. Lu, J. Z. Liu, X. L. You, J. X. Zhao and Q. H. Cai, 14, 863–866.
J. Electrochem. Soc., 2012, 159, E183–E186. 226 D. Ooyama, T. Tomon, K. Tsuge and K. Tanaka,
201 B. Z. Nikolic, H. Huang, D. Gervasio, A. Lin, C. Fierro, J. Organomet. Chem., 2001, 619, 299–304.
R. R. Adzic and E. B. Yeager, J. Electroanal. Chem., 1990, 227 K. Tanaka and D. Ooyama, Coord. Chem. Rev., 2002, 226,
295, 415–423. 211–218.
202 N. Hoshi, T. Mizumura and Y. Hori, Electrochim. Acta, 228 A. Begum and P. G. Pickup, Electrochem. Commun., 2007, 9,
1995, 40, 883–887. 2525–2528.
203 N. Hoshi, T. Suzuki and Y. Hori, J. Phys. Chem. B, 1997, 101, 229 C. M. Bolinger, N. Story, B. P. Sullivan and T. J. Meyer,
8520–8524. Inorg. Chem., 1988, 27, 4582–4587.
204 N. Hoshi, S. Kawatani, M. Kudo and Y. Hori, J. Electroanal. 230 M. R. M. Bruce, E. Megehee, B. P. Sullivan, H. Thorp,
Chem., 1999, 467, 67–73. T. R. O’Toole, A. Downard and T. J. Meyer, Organometallics,
205 N. Hoshi and Y. Hori, Electrochim. Acta, 2000, 45, 1988, 7, 238–240.
4263–4270. 231 M. R. M. Bruce, E. Megehee, B. P. Sullivan, H. H. Thorp,
206 F. ChunJie, F. YbuJun, Z. ChunHua, Z. QingWei and T. R. O’Toole, A. Downard, J. R. Pugh and T. J. Meyer, Inorg.
S. ShiGang, Sci. China, Ser. B: Chem., 2007, 50, 593–598. Chem., 1992, 31, 4864–4873.
207 C. M. Sanchez-Sanchez, J. Souza-Garcia, E. Herrero and 232 S. C. Cheng, C. A. Blaine, M. G. Hill and K. R. Mann, Inorg.
A. Aldaz, J. Electroanal. Chem., 2012, 668, 51–59. Chem., 1996, 35, 7704–7708.
208 N. Hoshi, M. Noma, T. Suzuki and Y. Hori, J. Electroanal. 233 A. G. M. M. Hossain, T. Nagaoka and K. Ogura, Electrochim.
Chem., 1997, 421, 15–18. Acta, 1996, 41, 2773–2780.

672 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

234 A. Szymaszek and F. P. Pruchnik, J. Organomet. Chem., 262 W. Tang, A. A. Peterson, A. S. Varela, Z. P. Jovanov, L. Bech,
1989, 376, 133–140. W. J. Durand, S. Dahl, J. K. Norskov and I. Chorkendorff,
235 A. Miedaner, C. J. Curtis, R. M. Barkley and D. L. DuBois, Phys. Chem. Chem. Phys., 2012, 14, 76–81.
Inorg. Chem., 1994, 33, 5482–5490. 263 A. H. Zhou, D. L. He, N. X. Xie, Q. J. Xie, L. H. Nie and
236 S. C. Cheng, C. A. Blaine, M. G. Hill and K. R. Mann, Inorg. S. Z. Yao, Electrochim. Acta, 2000, 45, 3943–3950.
Chem., 1996, 35, 7704–7708. 264 H. Li and C. Oloman, J. Appl. Electrochem., 2005, 35,
237 K. Tanaka, Y. Kushi, K. Tsuge, K. Toyohara, T. Nishioka 955–965.
and K. Isobe, Inorg. Chem., 1998, 37, 120–126. 265 K. P. Kuhl, E. R. Cave, D. N. Abram and T. F. Jaramillo,
238 P. G. Jessop, T. Ikariya and R. Noyori, Nature, 1994, 368, Energy Environ. Sci., 2012, 5, 7050–7059.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

231–233. 266 H. De Jesus-Cardona, C. del Moral and C. R. Cabrera,


239 P. G. Jessop, F. Joo and C. C. Tai, Coord. Chem. Rev., 2004, J. Electroanal. Chem., 2001, 513, 45–51.
248, 2425–2442. 267 D. W. Dewulf, T. Jin and A. J. Bard, J. Electrochem. Soc.,
240 P. Kang, C. Cheng, Z. F. Chen, C. K. Schauer, T. J. Meyer 1989, 136, 1686–1691.
and M. Brookhart, J. Am. Chem. Soc., 2012, 134, 5500–5503. 268 Y. Hori, H. Konishi, T. Futamura, A. Murata, O. Koga,
241 Y. Hori, K. Kikuchi and S. Suzuki, Chem. Lett., 1985, 1695–1698. H. Sakurai and K. Oguma, Electrochim. Acta, 2005, 50,
242 M. Jitaru, D. A. Lowy, M. Toma, B. C. Toma and L. Oniciu, 5354–5369.
J. Appl. Electrochem., 1997, 27, 875–889. 269 J. Lee and Y. Tak, Electrochim. Acta, 2001, 46, 3015–3022.
243 M. Gattrell, N. Gupta and A. Co, J. Electroanal. Chem., 2006, 270 H. Yano, T. Tanaka, M. Nakayama and K. Ogura,
594, 1–19. J. Electroanal. Chem., 2004, 565, 287–293.
244 R. L. Cook, R. C. Macduff and A. F. Sammells, J. Electrochem. 271 K. Ogura, H. Yano and T. Tanaka, Catal. Today, 2004, 98,
Soc., 1987, 134, 2375–2376. 515–521.
245 S. Wasmus, E. Cattaneo and W. Vielstich, Electrochim. Acta, 272 N. Hoshi, H. Ito, T. Suzuki and Y. Hori, J. Electroanal.
1990, 35, 771–775. Chem., 1995, 395, 309–312.
246 R. L. Cook, R. C. Macduff and A. F. Sammells, J. Electrochem. 273 N. Hoshi, M. Kuroda and Y. Hori, J. Electroanal. Chem.,
Soc., 1990, 137, 607–608. 2002, 521, 155–160.
247 Y. Hori, K. Kikuchi, A. Murata and S. Suzuki, Chem. Lett., 274 N. Hoshi, T. Uchida, T. Mizumura and Y. Hori,
1986, 897–898. J. Electroanal. Chem., 1995, 381, 261–264.
248 Y. Hori, A. Murata, R. Takahashi and S. Suzuki, J. Chem. 275 N. Hoshi, M. Kato and Y. Hori, J. Electroanal. Chem., 1997,
Soc., Chem. Commun., 1988, 17–19. 440, 283–286.
249 Y. Hori, A. Murata and R. Takahashi, J. Chem. Soc., Faraday 276 Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Surf. Sci.,
Trans. 1, 1989, 85, 2309–2326. 1995, 335, 258–263.
250 H. Shibata, J. A. Moulijn and G. Mul, Catal. Lett., 2008, 123, 277 I. Takahashi, O. Koga, N. Hoshi and Y. Hori, J. Electroanal.
186–192. Chem., 2002, 533, 135–143.
251 X. Y. Wang, S. Q. Liu, K. L. Huang, Q. J. Feng, D. L. Ye, 278 Y. Hori, I. Takahashi, O. Koga and N. Hoshi, J. Mol. Catal.
B. Liu, J. L. Liu and G. H. Jin, Chin. Chem. Lett., 2010, 21, A: Chem., 2003, 199, 39–47.
987–990. 279 M. R. Goncalves, A. Gomes, J. Condeco, R. Fernandes,
252 T. Saeki, K. Hashimoto, N. Kimura, K. Omata and T. Pardal, C. A. C. Sequeira and J. B. Branco, Energy
A. Fujishima, J. Electroanal. Chem., 1995, 390, 77–82. Convers. Manage., 2010, 51, 30–32.
253 S. Kaneco, K. Iiba, S. K. Suzuki, K. Ohta and T. Mizuno, 280 K. J. P. Schouten, Z. S. Qin, E. P. Gallent and M. T. M.
J. Phys. Chem. B, 1999, 103, 7456–7460. Koper, J. Am. Chem. Soc., 2012, 134, 9864–9867.
254 S. Kaneco, K. Iiba, N. Hiei, K. Ohta, T. Mizuno and 281 K. J. P. Schouten, Y. Kwon, C. J. M. van der Ham, Z. Qin
T. Suzuki, Electrochim. Acta, 1999, 44, 4701–4706. and M. T. M. Koper, Chem. Sci., 2011, 2, 1902–1909.
255 S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, 282 W. J. Durand, A. A. Peterson, F. Studt, F. Abild-Pedersen
Chem. Eng. J., 2007, 128, 47–50. and J. K. Norskov, Surf. Sci., 2011, 605, 1354–1359.
256 S. Kaneco, Y. Ueno, H. Katsumata, T. Suzuki and K. Ohta, 283 C. Liu, T. R. Cundari and A. K. Wilson, J. Phys. Chem. C,
Chem. Eng. J., 2006, 119, 107–112. 2012, 116, 5681–5688.
257 R. L. Cook, R. C. Macduff and A. F. Sammells, 284 M. Maeda, Y. Kitaguchi, S. Ikeda and K. Ito, J. Electroanal.
J. Electrochem. Soc., 1987, 134, 1873–1874. Chem., 1987, 238, 247–258.
258 J. J. Kim, D. P. Summers and K. W. Frese, J. Electroanal. 285 T. Ohmori, A. Nakayama, H. Mametsuka and E. Suzuki,
Chem., 1988, 245, 223–244. J. Electroanal. Chem., 2001, 514, 51–55.
259 G. Kyriacou and A. Anagnostopoulos, J. Electroanal. Chem., 286 G. B. Stevens, T. Reda and B. Raguse, J. Electroanal. Chem.,
1992, 322, 233–246. 2002, 526, 125–133.
260 K. Ohta, K. Suda, S. Kaneco and T. Mizuno, J. Electrochem. 287 D. R. Kauffman, D. Alfonso, C. Matranga, H. F. Qian and
Soc., 2000, 147, 233–237. R. C. Jin, J. Am. Chem. Soc., 2012, 134, 10237–10243.
261 R. L. Cook, R. C. Macduff and A. F. Sammells, 288 M. Shibata, K. Yoshida and N. Furuya, J. Electrochem. Soc.,
J. Electrochem. Soc., 1988, 135, 1320–1326. 1998, 145, 2348–2353.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 673
View Article Online

Review Article Chem Soc Rev

289 L. V. Haynes and D. T. Sawyer, Anal. Chem., 1967, 39, 332–338. 316 R. J. Haines, R. E. Wittrig and C. P. Kubiak, Inorg. Chem.,
290 M. Azuma, K. Hashimoto, M. Hiramoto, M. Watanabe and 1994, 33, 4723–4728.
T. Sakata, J. Electrochem. Soc., 1990, 137, 1772–1778. 317 M. Halmann, Nature, 1978, 275, 115–116.
291 W. Paik, T. N. Andersen and H. Eyring, Electrochim. Acta, 318 B. Aurianblajeni, M. Halmann and J. Manassen, Sol. Energy
1969, 14, 1217–1232. Mater., 1983, 8, 425–440.
292 A. Bewick and G. P. Greener, Tetrahedron Lett., 1969, 4623–4626. 319 D. Canfield and K. W. Frese, J. Electrochem. Soc., 1983, 130,
293 Y. Hori and S. Suzuki, Bull. Chem. Soc. Jpn., 1982, 55, 660–665. 1772–1773.
294 M. Todoroki, K. Hara, A. Kudo and T. Sakata, J. Electroanal. 320 E. E. Barton, D. M. Rampulla and A. B. Bocarsly, J. Am.
Chem., 1995, 394, 199–203. Chem. Soc., 2008, 130, 6342–6344.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

295 T. Y. Chang, R. M. Liang, P. W. Wu, J. Y. Chen and 321 S. Kapusta and N. Hackerman, J. Electrochem. Soc., 1983,
Y. C. Hsieh, Mater. Lett., 2009, 63, 1001–1003. 130, 607–613.
296 S. Ohya, S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, 322 S. Ikeda, T. Takagi and K. Ito, Bull. Chem. Soc. Jpn., 1987,
Catal. Today, 2009, 148, 329–334. 60, 2517–2522.
297 C. W. Li and M. W. Kanan, J. Am. Chem. Soc., 2012, 134, 323 K. Ito, S. Ikeda, N. Yamauchi, T. Iida and T. Takagi, Bull.
7231–7234. Chem. Soc. Jpn., 1985, 58, 3027–3028.
298 H. W. Wu, N. Zhang, Z. J. Cao, H. M. Wang and S. U. Hong, 324 T. Mizuno, K. Ohta, A. Sasaki, T. Akai, M. Hirano and
Int. J. Quantum Chem., 2012, 112, 2532–2540. A. Kawabe, Energy Sources, 1995, 17, 503–508.
299 H. Li, M. Eddaoudi, M. O’Keeffe and O. M. Yaghi, Nature, 325 S. R. Narayanan, B. Haines, J. Soler and T. I. Valdez,
1999, 402, 276–279. J. Electrochem. Soc., 2011, 158, A167–A173.
300 S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen 326 B. A. Parkinson and P. F. Weaver, Nature, 1984, 309, 148–149.
and I. D. Williams, Science, 1999, 283, 1148–1150. 327 S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Chem.
301 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, Eng. J., 2006, 116, 227–231.
M. O’Keeffe and O. M. Yaghi, Science, 2002, 295, 469–472. 328 S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Appl.
302 H. K. Chae, D. Y. Siberio-Perez, J. Kim, Y. Go, M. Eddaoudi, Catal., B, 2006, 64, 139–145.
A. J. Matzger, M. O’Keeffe and O. M. Yaghi, Nature, 2004, 329 S. Kaneco, Y. Ueno, H. Katsumata, K. T. Suzuki and
427, 523–527. K. Ohta, Chem. Eng. J., 2009, 148, 57–62.
303 J.-y. Ye and C.-j. Liu, Chem. Commun., 2011, 47, 2167–2169. 330 H. Li and C. Oloman, J. Appl. Electrochem., 2006, 36,
304 R. Senthil Kumar, S. Senthil Kumar and M. Anbu 1105–1115.
Kulandainathan, Microporous Mesoporous Mater., 2013, 331 H. Li and C. Oloman, J. Appl. Electrochem., 2007, 37,
168, 57–64. 1107–1117.
305 R. S. Kumar, S. S. Kumar and M. A. Kulandainathan, 332 R. L. Machunda, H. Ju and J. Lee, Curr. Appl. Phys., 2011,
Electrochem. Commun., 2012, 25, 70–73. 11, 986–988.
306 M. Schwartz, R. L. Cook, V. M. Kehoe, R. C. Macduff, 333 Y. H. Chen and M. W. Kanan, J. Am. Chem. Soc., 2012, 134,
J. Patel and A. F. Sammells, J. Electrochem. Soc., 1993, 1986–1989.
140, 614–618. 334 G. K. S. Prakash, F. A. Viva and G. A. Olah, J. Power Sources,
307 Y. H. Chen, C. W. Li and M. W. Kanan, J. Am. Chem. Soc., 2013, 223, 68–73.
2012, 134, 19969–19972. 335 J. J. Wu, F. G. Risalvato, F. S. Ke, P. J. Pellechia and
308 M. Watanabe, M. Shibata, A. Katoh, T. Sakata and X. D. Zhou, J. Electrochem. Soc., 2012, 159, F353–F359.
M. Azuma, J. Electroanal. Chem., 1991, 305, 319–328. 336 A. Bewick and G. P. Greener, Tetrahedron Lett., 1970, 391–394.
309 M. Watanabe, M. Shibata, A. Kato, M. Azuma and 337 F. Koleli and D. Balun, Appl. Catal., A, 2004, 274, 237–242.
T. Sakata, J. Electrochem. Soc., 1991, 138, 3382–3389. 338 S. Kaneco, R. Iwao, K. Iiba, K. Ohta and T. Mizuno, Energy,
310 J. Christophe, T. Doneux and C. Buess-Herman, Electro- 1998, 23, 1107–1112.
catalysis, 2012, 3, 139–146. 339 B. Eneau-Innocent, D. Pasquier, F. Ropital, J. M. Leger and
311 A. Schizodimou and G. Kyriacou, Electrochim. Acta, 2012, K. B. Kokoh, Appl. Catal., B, 2010, 98, 65–71.
78, 171–176. 340 B. Innocent, D. Liaigre, D. Pasquier, F. Ropital, J. M. Leger
312 T. V. Magdesieva, I. V. Zhukov, D. N. Kravchuk, O. A. and K. B. Kokoh, J. Appl. Electrochem., 2009, 39, 227–232.
Semenikhin, L. G. Tomilova and K. P. Butin, Russ. Chem. 341 K. Subramanian, K. Asokan, D. Jeevarathinam and
Bull., 2002, 51, 805–812. M. Chandrasekaran, J. Appl. Electrochem., 2007, 37, 255–260.
313 B. Verdejo, J. Aguilar, E. Garcia-Espana, P. Gavina, 342 A. Murata and Y. Hori, Bull. Chem. Soc. Jpn., 1991, 64,
J. Latorre, C. Soriano, J. M. Llinares and A. Domenech, 123–127.
Inorg. Chem., 2006, 45, 3803–3815. 343 G. Z. Kyriacou and A. K. Anagnostopoulos, J. Appl. Electro-
314 J. P. Muena, M. Villagran, J. Costamagna and M. J. Aguirre, chem., 1993, 23, 483–486.
J. Coord. Chem., 2008, 61, 479–489. 344 S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Energy
315 B. Verdejo, S. Blasco, J. Gonzalez, E. Garcia-Espana, P. Gavina, Fuels, 2006, 20, 409–414.
S. Tatay, A. Domenech, M. T. Domenech-Carbo, H. R. Jimenez 345 S. Kaneco, K. Iiba, K. Ohta and T. Mizuno, J. Solid State
and C. Soriano, Eur. J. Inorg. Chem., 2008, 84–97. Electrochem., 1999, 3, 424–428.

674 | Chem. Soc. Rev., 2014, 43, 631--675 This journal is © The Royal Society of Chemistry 2014
View Article Online

Chem Soc Rev Review Article

346 S. Kaneco, K. Iiba, K. Ohta and T. Mizuno, Energy Sources, 369 R. P. S. Chaplin and A. A. Wragg, J. Appl. Electrochem., 2003,
2000, 22, 127–135. 33, 1107–1123.
347 S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, 370 J. W. Li and G. Prentice, J. Electrochem. Soc., 1997, 144,
J. Solid State Electrochem., 2007, 11, 490–495. 4284–4288.
348 S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electro- 371 M. Isaacs, F. Armijo, G. Ramirez, E. Trollund, S. R. Biaggio,
chim. Acta, 2006, 51, 3316–3321. J. Costamagna and M. J. Aguirre, J. Mol. Catal. A: Chem.,
349 Y. Oh and X. L. Hu, Chem. Soc. Rev., 2013, 42, 2005, 229, 249–257.
2253–2261. 372 N. Furuya and K. Matsui, J. Electroanal. Chem., 1989, 271,
350 F. Koleli, T. Ropke and C. H. Hamann, Synth. Met., 2004, 181–191.
Published on 01 November 2013. Downloaded by University of Rostock on 9/24/2018 3:46:26 PM.

140, 65–68. 373 M. Shibata and N. Furuya, Electrochim. Acta, 2003, 48,
351 R. Aydin and F. Koleli, Synth. Met., 2004, 144, 75–80. 3953–3958.
352 R. D. L. Smith and P. G. Pickup, Electrochem. Commun., 374 J. Costamagna, J. Canales, J. Vargas and G. Ferraudi, Pure
2010, 12, 1749–1751. Appl. Chem., 1995, 67, 1045–1052.
353 G. Seshadri, C. Lin and A. B. Bocarsly, J. Electroanal. Chem., 375 J. Costamagna, G. Ferraudi, J. Canales and J. Vargas,
1994, 372, 145–150. Coord. Chem. Rev., 1996, 148, 221–248.
354 E. B. Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, 376 M. R. Thorson, K. I. Siil and P. J. A. Kenis, J. Electrochem.
E. Abelev and A. B. Bocarsly, J. Am. Chem. Soc., 2010, 132, Soc., 2013, 160, F69–F74.
11539–11551. 377 K. Ogura, J. R. Ferrell, A. V. Cugini, E. S. Smotkin and M. D.
355 M. Z. Ertem, S. J. Konezny, C. M. Araujo and V. S. Batista, Salazar-Villalpando, Electrochim. Acta, 2010, 56, 381–386.
J. Phys. Chem. Lett., 2013, 4, 745–748. 378 J. J. Carroll, J. D. Slupsky and A. E. Mather, J. Phys. Chem.
356 J. A. Keith and E. A. Carter, Chem. Sci., 2013, 4, 1490–1496. Ref. Data, 1991, 20, 1201–1209.
357 C. H. Lim, A. M. Holder and C. B. Musgrave, J. Am. Chem. 379 K. Hara, A. Kudo and T. Sakata, J. Electroanal. Chem., 1997,
Soc., 2013, 135, 142–154. 421, 1–4.
358 A. Gennaro, A. A. Isse, M. G. Severin, E. Vianello, I. Bhugun 380 C. Amatore and J. M. Saveant, J. Am. Chem. Soc., 1981, 103,
and J. M. Saveant, J. Chem. Soc., Faraday Trans., 1996, 92, 5021–5023.
3963–3968. 381 M. Royer, Compt. Rend., 1870, 731–732.
359 A. Gennaro, A. A. Isse, J. M. Saveant, M. G. Severin and 382 J. Song, Y. Kim, M. Lim, H. Lee, J. I. Lee and W. Shin,
E. Vianello, J. Am. Chem. Soc., 1996, 118, 7190–7196. ChemSusChem, 2011, 4, 587–590.
360 L. L. Snuffin, L. W. Whaley and L. Yu, J. Electrochem. Soc., 383 B. D. Smith, D. E. Irish, P. Kedzierzawski and J. Augustynski,
2011, 158, F155–F158. J. Electrochem. Soc., 1997, 144, 4288–4296.
361 K. Sugimura, S. Kuwabata and H. Yoneyama, J. Am. Chem. 384 P. Friebe, P. Bogdanoff, N. AlonsoVante and H. Tributsch,
Soc., 1989, 111, 2361–2362. J. Catal., 1997, 168, 374–385.
362 S. Kuwabata, N. Morishita and H. Yoneyama, Chem. Lett., 385 L. D. Burke and J. A. Collins, J. Appl. Electrochem., 1999, 29,
1990, 1151–1154. 1427–1438.
363 S. Kuwabata, R. Tsuda and H. Yoneyama, J. Am. Chem. Soc., 386 R. Senthil Kumar, S. Senthil Kumar and M. Anbu
1994, 116, 5437–5443. Kulandainathan, Electrochem. Commun., 2012, 25, 70–73.
364 P. K. Addo, R. L. Arechederra, A. Waheed, J. D. Shoemaker, 387 J. Yano, T. Morita, K. Shimano, Y. Nagami and S. Yamasaki,
W. S. Sly and S. D. Minteer, Electrochem. Solid-State Lett., J. Solid State Electrochem., 2007, 11, 554–557.
2011, 14, E9–E13. 388 L. M. Chiacchiarelli, Y. Zhai, G. S. Frankel, A. S. Agarwal
365 H. A. Hansen, J. B. Varley, A. A. Peterson and J. K. Norskov, and N. Sridhar, J. Appl. Electrochem., 2012, 42, 21–29.
J. Phys. Chem. Lett., 2013, 4, 388–392. 389 K. Bujno, R. Bilewicz, L. Siegfried and T. Kaden, Electro-
366 F. A. Armstrong and J. Hirst, Proc. Natl. Acad. Sci. U. S. A., chim. Acta, 1997, 42, 1201–1206.
2011, 108, 14049–14054. 390 E. E. Benson and C. P. Kubiak, Chem. Commun., 2012, 48,
367 J. J. G. Moura, C. D. Brondino, J. Trincao and M. J. Romao, 7374–7376.
JBIC, J. Biol. Inorg. Chem., 2004, 9, 791–799. 391 V. S. Thoi and C. J. Chang, Chem. Commun., 2011, 47,
368 F. A. M. de Bok, P. L. Hagedoorn, P. J. Silva, W. R. Hagen, 6578–6580.
E. Schiltz, K. Fritsche and A. J. M. Stams, Eur. J. Biochem., 392 M. Beley, J. P. Collin, R. Ruppert and J. P. Sauvage, J. Chem.
2003, 270, 2476–2485. Soc., Chem. Commun., 1984, 1315–1316.

This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev., 2014, 43, 631--675 | 675

You might also like