You are on page 1of 7

Article

pubs.acs.org/Langmuir

Switching the Reaction Course of Electrochemical CO2 Reduction


with Ionic Liquids
Liyuan Sun,† Ganganahalli K. Ramesha,‡ Prashant V. Kamat,*,†,‡,§ and Joan F. Brennecke*,†

Department of Chemical and Biomolecular Engineering, ‡Radiation Laboratory, and §Department of Chemistry and Biochemistry,
University of Notre Dame, Notre Dame, Indiana 46556, United States
*
S Supporting Information

ABSTRACT: The ionic liquid 1-ethyl-3-methylimidazolium


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

bis(trifluoromethylsulfonyl)imide ([emim][Tf2N]) offers new


ways to modulate the electrochemical reduction of carbon
Downloaded via UNIV OF WOLLONGONG on February 17, 2019 at 23:55:20 (UTC).

dioxide. [emim][Tf2N], when present as the supporting


electrolyte in acetonitrile, decreases the reduction over-
potential at a Pb electrode by 0.18 V as compared to
tetraethylammonium perchlorate as the supporting electrolyte.
More interestingly, the ionic liquid shifts the reaction course
during the electrochemical reduction of carbon dioxide by
promoting the formation of carbon monoxide instead of
oxalate anion. With increasing concentration of [emim]-
[Tf2N], a carboxylate species with reduced CO2 covalently
bonded to the imidazolium ring is formed along with carbon monoxide. The results highlight the catalytic effects of the medium
in modulating the CO2 reduction products.

1. INTRODUCTION tional toxic and volatile organic solvents in many chemical


The steadily increasing atmospheric concentration of carbon processes. Additionally, the high solubility of CO2 in ILs,19−21
dioxide (CO2), a major greenhouse gas, mainly generated from intrinsic ionic conductivity, and wide potential windows make
burning of fossil fuels, and its potential impact on climate has ILs attractive media for electrochemical reduction of CO2.
been the topic of much discussion.1−4 On the other hand, CO2 Recently, Rosen et al. reported the effective lowering of the
is recognized as a naturally abundant and inexpensive carbon reduction overpotential for CO2 conversion into CO at Pt and
source for the production of fuels and feed stocks for a variety Ag electrodes in an aqueous solution of the IL 1-ethyl-3-
of chemical syntheses in industry.2,4 Therefore, an attractive methylimidazolium tetrafluoroborate ([emim][BF4]). They
way to mitigate increasing atmospheric CO2 concentrations is proposed that the effectiveness of this newly observed catalytic
through conversion of CO2 into fuels and useful chemicals reduction relies on the ability of the IL to lower the energy
using carbon-neutral energy resources.4 Approaches such as barrier for the formation of CO2−• through the formation of a
chemical, thermochemical, photochemical, biochemical, and complex.11 Wipple et al. also proposed that the overpotential of
electrochemical reduction methods have been investigated to CO2 reduction could be lowered by stabilizing the intermediate
achieve CO2 conversion.1,5−11 Of these methods, electro- CO2−• using a suitable catalyst.4
chemical reduction of CO2 is considered to be attractive Electrochemical reduction is a convenient approach to carry
because of its higher achievable conversion efficiency, its out reduction of CO2 in a controlled way. The product
product selectivity, and its potential to store electrical energy distribution varies depending on the type of metal electrodes
that is extracted from renewable energy sources like the employed in a certain electrolyte. For example, oxalate has been
sun.4,12,13 CO2 is thermodynamically stable and kinetically widely reported as the major product of CO2 reduction at a Pb
inert; the main hurdle in the electrochemical CO2 reduction lies electrode in nonaqueous media.12,22−24 Also, Pb is considered
in the first step, the one-electron reduction of CO2. This to be an inert electrode material that does not interfere with the
activation of CO2 to form an anion radical (CO2−•) requires an reduction of CO2 and the competing pathways for product
unusually high reduction potential of −1.9 V vs NHE,14 formation.17 Here we investigate the role that the IL 1-ethyl-3-
associated with an overpotential that is dependent on the methylimidazolium bis(trifluoromethylsulfonyl)imide ([emim]-
electrode and medium.11,13,15−17 Efforts have been made to [Tf2N], shown in Figure 1) plays as a homogeneous catalyst for
achieve CO2 reduction at lower electrochemical potentials, e.g., CO2 reduction using electrochemical reduction at a Pb cathode.
using pyridinium-based homogeneous catalysts.18
Due to their low volatility, nonflammability, good solvating Received: March 7, 2014
ability, and good thermal and chemical stability, ionic liquids Revised: May 9, 2014
(ILs) have been regarded as “green” alternatives for conven- Published: May 13, 2014

© 2014 American Chemical Society 6302 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308
Langmuir Article

2.2. Electrolysis at Controlled Potentials and Product


Analysis. We carried out electrolysis at controlled potentials to
monitor the CO2 reduction products. When the electrolysis was
carried out for 60 min in 0.1 M TEAP/AcN we observed the
Figure 1. Structure of the ionic liquid [emim][Tf2N]. formation of a white precipitate. Analysis of the precipitate
using FTIR spectra (see Supporting Information, Figure S3)
confirmed the formation of oxalate as the primary product of
Specifically, we seek answers to the following questions: (1) CO2 reduction. Oxalate generated from CO2 reduction in 0.1
how does [emim][Tf2N] mediate the product selectivity during M TEAP/AcN was quantified using HPLC (Supporting
CO2 reduction and (2) how does the mechanism of CO2 Information, Figure S4). The formation of oxalate anion is
reduction switch from oxalate anion to CO in an IL-based the result of dimerization of CO2−•. This selective reduction
nonaqueous medium? product formed at Pb electrodes agrees with earlier
findings.22,23
2. RESULTS AND DISCUSSION The oxalate formation was significantly decreased when 0.1
2.1. Voltammetry Study of CO2 Reduction in TEAP/ M [emim][Tf2N]/AcN was used as an electrolyte. Efforts were
AcN and [emim][Tf2N]/AcN at a Pb Working Electrode. made to isolate and identify reduction products using 1H and
13
The cyclic voltammograms (CVs) measured using a Pb cathode C NMR. In addition, the gaseous reduction product, carbon
in 0.1 M tetraethylammonium perchlorate/acetonitrile (TEAP/ monoxide (CO), was analyzed using GC analysis of the head
AcN) and 0.1 M [emim][Tf2N]/AcN are shown in Figure 2. space samples. The resonance lines of [emim][Tf2N] in the 1H
These voltammograms were recorded at a potential sweep rate NMR spectrum were assigned from their multiplet splitting
of 50 mV/s at room temperature after the electrolyte solution patterns, as well as comparison with peak assignments for other
was purged with N2 for 30 min to remove dissolved oxygen, [emim]+ ILs whose NMR spectra are available in the
followed by purging with CO2 for another 30 min. Parts a and b literature.26 The chemical shifts of the protons in [emim]-
show the reductive scan of N2-saturated and CO2-saturated [Tf2N] are summarized in Table 1, according to the labeling
electrolyte solutions, respectively. The increased current seen in
the CO2-saturated solutions confirms the ability of Pb electrode Table 1. 1H NMR Chemical Shifts (DMSO, δ, ppm relative
to reduce CO2 to CO2−•. to TMS) of the Pure 1-ethyl-3-methylimidazolium
The onset potential corresponding to the reduction of CO2 bis(trifluoromethylsulfonyl)imide ([emim][Tf2N]) and of
in the presence of 0.1 M [emim][Tf2N]/AcN is decreased by the Solution Obtained by CO2 Electrochemical Reduction at
0.18 V compared to that of 0.1 M TEAP/AcN. The onset −2.25 V vs Ag/AgNO3 in 0.1 M [emim][Tf2N]/AcN at
potential is −2.30 V in 0.1 M TEAP/AcN and −2.12 V in 0.1 Room Temperature, under Atmospheric Pressure for 1 h
M [emim][Tf2N]/AcN with respect to the reference electrode
no. of H
(Ag/AgNO3), as seen in Figure 2. Since it is difficult to atom pure [emim][Tf2N] COO−[emim][Tf2N] [emim][Gly]26
establish the exact reduction peak, we selected potentials that 7 1.42 1.35 1.41
gave a current of 0.6 mA/cm2 for performing reduction of CO2 8 3.85 3.96 3.87
in acetonitrile. Potentials of −2.40 and −2.25 V (vs Ag/ 6 4.20 4.48 4.22
AgNO3) were employed for CO2 reduction in 0.1 M TEAP/ 5 7.69 7.54 7.75
AcN and 0.1 M [emim][Tf2N]/AcN, respectively, in our 4 7.77 7.61 7.84
experiments, unless otherwise indicated. This approach of 2 9.12 9.68
monitoring the apparent current density at a fixed potential is
consistent with earlier reports.25 The difference in the two
reduction potentials represents the decrease in overpotential in shown in Figure 3. As is shown in the spectra in Figure 3a−c,
the [emim][Tf2N]/AcN system compared to TEAP/AcN. the original resonance lines of pure [emim][Tf2N], with the

Figure 2. Cyclic voltammograms (CVs) recorded using a Pb working electrode in an electrolyte containing (A) 0.1 M TEAP/AcN and (B) 0.1 M
[emim][Tf2N]/AcN. The CVs were recorded in (a) N2-saturated and (b) CO2-saturated solutions. (Scan rate 50 mv/s. The vertical line in each plot
is the potential employed for electrolysis experiments.)

6303 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308


Langmuir Article

Figure 3. 1H NMR spectra (in DMSO, reference to TMS) of (a) pure [emim][Tf2N], (b) 0.1 M [emim][Tf2N]/AcN after purging CO2 for 30 min,
(c) 0.1 M [emim][Tf2N]/AcN electrolysis solution after 1 h of CO2 reduction, and (d) separated precipitate generated in CO2 reduction. The
schematic representation of the [emim][Tf2N] ion pair with the labeling of the carbon atom used in the assignment is displayed.

exception of the proton on C2, become accompanied by a solution after electrolysis (Supporting Information, Figure S6),
corresponding secondary line of lower intensity after CO2 which can be attributed to the carboxylate group covalently
electrochemical reduction is carried out in [emim][Tf2N]/ bound to C2 of the imidazolium ring,27 providing further
AcN. The composite spectrum of the electrolysis solution confirmation that a carboxylate species was formed. It needs to
mixture indicates a substitution reaction at C2 of the be pointed out that electrospray ionization (ESI) MS provides
imidazolium ring to form a reacted imidazolium species. further evidence of the formation of the carboxylate species by
Figure 3d shows the 1H NMR spectrum of a precipitate showing a peak with an m/z value of 155.0822 (Supporting
generated during the electrolysis of CO2-saturated 0.1 M Information, Figure S7). The 1-ethyl-3-methylimidazolium-2-
[emim][Tf2N]/AcN after it was isolated from the electrolysis carboxylate species was quantified on the basis of integration of
solution mixture, purified, and dissolved in DMSO. The main peak intensities in the 1H NMR spectra, and CO was quantified
resonance lines in this spectrum correspond to the protons of by GC (Supporting Information, Figure S5).
the new reacted imidazolium species, which we believe to be a Rosen et al. 10 suggested that an adsorbed neutral
carboxylate complex. The less intense lines at the same [emimCO2] complex was responsible for lowering the
chemical shifts as the pure IL can be attributed to residual overpotential for the electrochemical conversion of CO2 to
unreacted IL in the precipitate. There have been abundant CO, lending support to the discovery presented here of a 1-
literature reports describing carboxylation at the C2 position of ethyl-3-methylimidazolium-2-carboxylate species in solution
imidazolium-based ILs, triggered by reaction with neutral CO2 and as a bulk precipitate. From the 1H NMR spectra we
molecules.3,27−29 The reaction scheme is shown in Figure 4, cannot determine if the complex is neutral or negatively
although there is some debate regarding the reversibility of the charged. We envision that the [emim]+ cation is deprotonated
reaction. We reason from the spectra in Figure 3 that to form the neutral carbene (Figure 4). The carbene then reacts
electrochemically reduced CO2 is also capable of deprotonating with the reduced CO2−•, which would form a negatively
the imidazolium ring to form a carboxylate species. In addition, charged [emimCO 2]−. However, this complex may be
we observe a resonance line at 155.1 ppm in the 13C NMR reprotonated to form the overall neutral molecule shown in
spectrum of the CO2-saturated 0.1 M [emim][Tf2N]/AcN eq 1. The proton removed from the C2 position is not seen in
the 1H NMR of the solution after electrolysis (c) or in the
precipitate (d); i.e., there is no 2′ in either spectrum. However,
if the H+ is attached to the complex, we would anticipate fast
exchange of that proton between the carboxylate site and the
polar solvent, so that it would not show up in the 1H NMR.27,30

[emim] + CO2 + H+ + e− → ([emimCO2 ]− ···H+) (1)

Figure 5 shows the products formed and the Faradaic


Figure 4. A schematic diagram of the carboxylation of C2 of the efficiency as a function of time for electrolysis of CO2-saturated
imidazolium ring by CO2. Redrawn according to refs 16−18. 0.1 M TEAP/AcN solution (A and B) and CO2-saturated 0.1
6304 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308
Langmuir Article

Figure 5. Formation of the products (a) oxalate anion, (b) CO, and (c) carboxylate anion during the course of electrolysis and the corresponding
Faradaic efficiency using two different electrolytes: (A and B) 0.1 M TEAP/AcN (applied potential of −2.40 V vs Ag/AgNO3) and (C and D) 0.1 M
[emim][Tf2N]/AcN (applied potential of −2.25 V vs Ag/Ag NO3). The electrolyte was kept saturated with CO2. The Faradaic efficiency was
monitored from the charges flowed during the reduction period.

Figure 6. Effect of [emim][Tf2N] concentration on the distribution of CO2 reduction products: (A) (a) oxalate and (b) CO, and (B) (b) CO and
(c) carboxylate. The electrolysis was carried out at −2.34 V (vs Ag/AgNO3) in CO2-saturated AcN containing 0.1 M TEAP and varying
concentrations of [emim][Tf2N]. The products were analyzed after 1 h of electrolysis.

M [emim][Tf2N]/AcN solution (C and D). In agreement with Obviously, the presence of [emim][Tf2N] changes the
previous studies, the product of electrochemical reduction of course of the CO2 reduction reaction so that it produces CO
CO2 in nonaqueous media using Pb as the working electrode and an imidazolium−carboxylate complex and suppresses the
and TEAP as the supporting electrolyte is predominantly dimerization of CO2−• to form oxalate. We further explored the
oxalate anion.12,22−24,31 When we used 0.1 M [emim][Tf2N]/ role of [emim][Tf2N] in switching the course of the reaction by
AcN as the supporting electrolyte, CO and an imidazolium− varying its concentration in a 0.1 M TEAP/AcN solution and
carboxylate species become the prominent products. In the 0.1 monitoring the product distribution of CO2 electrochemical
M [emim][Tf2N]/AcN system, the Faradaic efficiency of the reduction.
imidazolium carboxylate decreased with time and that of CO 2.3. Shifting the Course of the Electrochemical
increased with time, until equilibrium values were reached after Reduction of CO2 with [emim][Tf2N]. Cyclic voltammo-
an hour or two. grams were recorded using a Pb working electrode (Supporting
6305 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308
Langmuir Article

Scheme 1. Reaction Pathways for the Electrochemical Reduction of CO2 in the (A) Absence and (B, C) Presence of
[emim][Tf2N] at a Pb Electrode in AcN

Information, Figure S8) in N2- and CO2-saturated 0.1 M hydroxide. This produced no change in the amount of CO
TEAP/AcN solutions with different concentrations of [emim]- produced, thus eliminating reaction 3 as the mechanism.
[Tf2N]. We applied a constant potential of −2.34 V (vs Ag/ Gu et al. reported a mechanism where N-heterocyclic
AgNO3) to carry out CO2 electrolysis for a period of 1 h. The carbenes (NHC) catalyzed CO2 splitting to form CO when
concentration of products and current that flowed through the aromatic aldehydes were used as oxygen acceptors. They
circuit were monitored to estimate the Faradaic efficiency. As proposed a reaction scheme where CO2 reacted with the NHC,
shown in Figure 6A, CO increases and oxalate anion decreases resulting in the formation of an imidazolium carboxylate, which
with increasing concentration of the IL [emim][Tf2N]. Another subsequently reacted with added aromatic aldehyde to generate
reduction product is imidazolium carboxylate, which is formed an intermediate species that was likely stabilized by
in significant amounts during the electrolysis (Figure 6B). intermolecular hydrogen bonding. The H+ on the imidazolium
It is evident from the product analysis that [emim][Tf2N] ring of the intermediate species was trapped by the added base,
can switch the course of the electrochemical reduction reaction, leading to the splitting of CO2 and formation of the
leading to the formation of CO and suppressing the oxalate corresponding aromatic salt. Finally, CO was released from
anion formation. It has been reported that CO is produced in the NHC−CO complex, which regenerated the NHC, which
the electrochemical reduction of CO2 in nonaqueous media at could proceed to the next catalytic cycle.8 The detection of a 1-
metal working electrodes, such as In, Zn, Sn, Au, Hg, Pb, and ethyl-3-methylimidazolium-2-carboxylate complex in our ex-
Pt, followed by the disproportionation of CO2−• (reaction periments and the leveling off of the Faradaic efficiency of both
2).22,23,31−33 On the other hand, in aqueous media protons can the carboxylate and CO as a function of electrolysis time
interact with CO2−• to produce CO (reaction 3).1,7,10,11,23,34 suggest that CO could be generated from the imidazolium
carboxylate in our system, as well. To test this hypothesis, we
2CO2−• → CO + CO32 − (2) used trimethylimidazolium bis(trifluoromethylsulfonyl)imide
([mmmim][Tf2N]) as the supporting electrolyte. In this IL
CO2−• + 2H+ → CO + H 2O (3) the proton at the C2 position is replaced with a methyl group,
making the formation of a carbene impossible. To our surprise,
We further probed the involvement of reactions 2 and 3 in CO was produced with an Faradaic efficiency of 57.2% for 1 h
our experiments. The 13C NMR spectrum of the electrolysis electrolysis at −2.25 V (vs Ag/Ag NO3), even higher than that
sample (Supporting Information, Figure S6) did not show any of 39.2% with [emim][Tf2N] as the electrolyte, which suggests
carbonate species, which would have to be produced in an that CO does not originate from the 1-ethyl-3-methylimidazo-
amount equal to the amount of CO formed if reaction 2 were lium-2-carboxylate; rather, the carboxylate may be a competing
taking place. This rules out the disproportionation of CO2−• species formed concomitantly with CO.
(reaction 2) in the electrolysis of CO2-saturated 0.1 M TEAP/ Finally, our observation that CO is formed preferentially over
AcN plus [emim][Tf2N] mixtures. We also explored the oxalate when [emim][Tf2N] is added to TEAP/AcN is similar
possibility of CO being formed by reaction of CO2−• with H+ to what occurs with the addition of a surfactant.35 This
ions (reaction 3) by adjusting the proton ion concentration indicates that ILs ([emim][Tf2N] and [mmmim][Tf2N]) may
through addition of sulfuric acid or tetraethylammonium play a role similar to that of surfactants, thus helping to stabilize
6306 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308
Langmuir Article

CO2−•, which is responsible for the reduced overpotential for electrode was manufactured from a 25 × 25 mm platinum gauze (Alfa,
the CO2 reduction reaction. Moreover, the presence of ILs Aesar, 99.9%, 52 mesh). The working electrode and the reference
modifies the course of the electrochemical reduction of CO2 in electrode were immersed in 10 mL of electrolyte solution in the
nonaqueous solution. Analogous to Rosen et al.’s study,10 the cathodic compartment of the cell, and the counter electrode was
immersed in 5 mL of electrolyte solution in the anodic compartment.
formation of CO in our system can be attributed to the The applied potential between the reference and working electrodes
interaction between CO2 and the cation of the IL at the was maintained by a Gamry potentiostat. Before each set of bulk
electrode surface. When only TEAP is used as the supporting electrolysis experiments, the cathodic compartment was purged with
electrolyte, CO2−• dimerizes on the surface of the Pb electrode N2 for 30 min to remove any moisture and dissolved oxygen, followed
to yield oxalate (shown in Scheme 1A). When an imidazolium by purging with CO2 for 30 min to saturate the electrolyte. Cyclic
IL is added, the cathode surface becomes covered with both voltammograms of the Pb electrode in N2- and CO2-saturated
CO2−• and the cations of the IL. The imidazolium cation electrolyte were taken between −1.50 and −2.50 V (vs Ag/Ag NO3) at
stabilizes the CO2−•, prevents close approach of two CO2−•, a sweep rate of 50 mV/s right after purging the gas. Generally, the
and inhibits dimerization to form oxalate (Scheme 1B). In surface of the working electrode was electrochemically conditioned,
and equilibrium was reached after 10 CV cycles with the electrolyte
addition, CO2−• reacts with the [emim]+ cation to form kept under the N2 or CO2 atmosphere. The data reported corresponds
carboxylate (Scheme 1C). Therefore, when [emim][Tf2N] is to the last cycle. Chronocoulometry was carried out in the closed cell
added, two competing pathways replace dimerization of CO2−• with an applied potential of −2.40 and −2.25 V for 0.1 M TEAP/AcN
to form oxalate (Scheme 1A), with one leading to the and 0.1 M [emim][Tf2N]/AcN, respectively.
formation of CO (Scheme 1B) and the other to the formation 4.3. Product Characterization. Fourier transform infrared
of the carboxylate (Scheme 1C). spectroscopy (FTIR) (Bruker Tensor 27) was used for the IR
measurements of the precipitate generated from CO2 reduction with
3. CONCLUSIONS TEAP as the electrolyte. Sodium oxalate powder was used as a
standard. The precipitate was separated and purified from the
The electrochemical reduction of CO2 at a Pb electrode in the electrolysis solution in a centrifuge at 6000 rpm for 15 min, followed
presence of ionic liquid [emim][Tf2N] shows the catalytic role by rinsing with pure AcN. The separation and purification were
of ionic liquid in shifting the course of the reduction pathway. repeated three times. Figure S3 (Supporting Information) shows the
The cyclic voltammograms confirm that imidazolium IL lowers IR spectra of the blank, standard, and unknown samples.
the overpotential for CO2 reduction by about 0.18 V. Most The NMR measurements were performed on a Varian spectrometer
operating at 600 MHz Larmor frequency. The samples were dissolved
importantly, our study shows that the course of the in DMSO, wherein the electrolysis solution was dried on a rotary
electrochemical reduction of CO2 can be altered dramatically evaporator at 40 °C under vacuum. The precipitate generated in
by the presence of [emim][Tf2N]. Specifically, increasing the [emim][Tf2N]/AcN was separated and purified similarly.
concentration of [emim][Tf2N] in 0.1 M TEAP/AcN switches High-performance liquid chromatography (HPLC) was performed
the product from oxalate, the widely reported principle product on an Alliance-Waters HPLC instrument with a C18 reversed phase
of CO2 reduction on Pb in AcN, to CO and an imidazolium (RP) column (250 × 4.6 mm i.d., OOG-4299-E0, Phenomenex). The
carboxylate complex. The results discussed in this investigation mobile phase was AcN/10 mM KH2PO4 aqueous solution of pH 2.5
further ascertain the role of imidazolium-based ionic liquid as (5:95) at a flow rate of 0.5 mL/min. A standard calibration curve was
homogeneous catalyst for lowering the CO2 reduction potential developed to determine the concentration of oxalate generated in the
unknown samples (Supporting Information, Figure S4).
as well as modulating the course of the reaction.
The gaseous product of CO2 reduction was analyzed with a gas
chromatography (GC) instrument (Thermo Scientific) equipped with
4. EXPERIMENTAL SECTION a thermal conductivity detector (TCD). A molecular sieve 5A column
4.1. Materials. Acetonitrile (AcN, Fisher Scientific, 99.9%), was used with helium as the carrier gas at a constant pressure of 3 psi.
sodium oxalate powder (Na2C2O4, Alfa Aesar, A.C.S, 99.5+%), The temperature of the oven was set at 65 °C for 4.5 min. The
potassium phosphate monobasic (KH2PO4, Fisher Scientific, A.C.S, standard calibration curve developed to quantify CO in unknown
99.9%), and dimethyl sulfoxide (DMSO, D, Cambridge Isotope samples is shown in Figure S5 (Supporting Information).


Laboratories, 99.9%, +1% v/v TMS,) were used as purchased.
Tetraethylammonium perchlorate (TEAP) (Alfa Aeser, 98%) was ASSOCIATED CONTENT
desiccated at 40 °C overnight in a vacuum oven before use.
[emim][Tf2N] and [mmmim][Tf2N] were synthesized by standard *
S Supporting Information
techniques. N2 and CO2 were supplied from Airgas (purity 99.995%). Construction of the electrochemical cell, IR spectra, and
The Nafion 117 membrane (manufactured by PuPont Fuel Cells) was calibration curves are presented. This material is available free
treated as described in the literature before use:36,37 boiling in 3% of charge via the Internet at http://pubs.acs.org/.


H2O2 aqueous solution, deionized water, 0.5 M H2SO4 aqueous
solution, and finally deionized water, successively, for 1 h for each step.
4.2. Electrochemical Measurements. Cyclic voltammetry and AUTHOR INFORMATION
chronocoulometry were performed in a two-compartment, three-
electrode glass cell. Figure S1a,b (Supporting Information) shows the Corresponding Author
construction of the cell. The working electrode was a lead sheet (Alfa, *E-mail: jfb@nd.edu (J.F.B.), pkamat@nd.edu (P.V.K.).
Aesar, 99.998%), which was polished with sand paper with a grit Notes
designation of 320, followed by rinsing with AcN before use to remove
any oxide impurities on the surface. A geometrical surface area of 2.0 The authors declare no competing financial interest.
cm2 of the lead working electrode was maintained in the solution in all
electrochemical experiments. The Ag/AgNO3 electrode, made of a
silver wire immersed in 0.01 M silver nitrate dissolved in 0.1 M TEAP/
AcN, was used as the reference electrode. The reference electrode was
■ ACKNOWLEDGMENTS
The authors acknowledge the Notre Dame Sustainable Energy
calibrated against a ferrocene/ferrocenium (Fc+/Fc) redox couple to Initiative for financial support during this project. We also
make sure it gave the right potential in reference to SHE.38 Figure S2 thank the Notre Dame Center for Environmental Science and
(Supporting Information) shows the calibration curves. The counter Technology (CEST) for free access to the facilities.
6307 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308
Langmuir


Article

REFERENCES (21) Goodrich, B. F.; de la Fuente, J. C.; Gurkan, B. E.; Zadigian, D.


J.; Price, E. A.; Huang, Y.; Brennecke, J. F. Experimental measurements
(1) Chen, Y.; Li, C. W.; Kanan, M. W. Aqueous CO2 reduction at of amine-functionalized anion-tethered ionic liquids with carbon
very low overpotential on oxide-derived Au nanoparticles. J. Am. Chem. dioxide. Ind. Eng. Chem. Res. 2011, 50, 111−118.
Soc. 2012, 134, 19969−72. (22) Amatore, C.; Saveant, J. M. Mechanism and kinetic character-
(2) Chu, D.; Qin, G. X.; Yuan, X. M.; Xu, M.; Zheng, P.; Lu, J. istics of the electrochemical reduction of carbon dioxide in media of
Fixation of CO2 by electrocatalytic reduction and electropolymeriza- low proton availability. J. Am. Chem. Soc. 1981, 103, 5021−5023.
tion in ionic liquid-H2O solution. ChemSusChem 2008, 1, 205−209. (23) Jitaru, M.; Lowy, D. A.; Toma, M.; Toma, B. C.; Oniciu, L.
(3) Gurkan, B.; Goodrich, B. F.; Mindrup, E. M.; Ficke, L. E.; Massel, Electrochemical reduction of carbon dioxide on flat metallic cathodes.
M.; Seo, S.; Senftle, T. P.; Wu, H.; Glaser, M. F.; Shah, J. K.; Maginn, J. Appl. Electrochem. 1997, 27, 875−889.
E. J.; Brennecke, J. F.; Schneider, W. F. Molecular design of high (24) Tomita, Y.; Teruya, S.; Koga, O.; Hori, Y. Electrochemical
capacity, low viscosity, chemically tunable ionic liquids for CO2 reduction of carbon dioxide at a platinum electrode in acetonitrile−
capture. J. Phys. Chem. Lett. 2010, 1, 3494−3499. water mixtures. J. Electrochem. Soc. 2000, 147, 4164−4167.
(4) Whipple, D. T.; Kenis, P. J. A. Prospects of CO2 utilization via (25) Nakagawa, T.; Beasley, C. A.; Murray, R. W. Efficient electro-
direct heterogeneous electrochemical reduction. J. Phys. Chem. Lett. oxidation of water near its reversible potential by a mesoporous IrOx
2010, 1, 3451−3458. nanoparticle film. J. Phys. Chem. C 2009, 113, 12958−12961.
(5) Barton, E. E.; Rampulla, D. M.; Bocarsly, A. B. Selective solar- (26) Kasahara, S.; Kamio, E.; Ishigami, T.; Matsuyama, H. Amino
driven reduction of CO2 to methanol using a catalyzed p-GaP based acid ionic liquid-based facilitated transport membranes for CO2
photoelectrochemical cell. J. Am. Chem. Soc. 2008, 130, 6342−6344. separation. Chem. Commun. 2012, 48, 6903−6905.
(6) Chueh, W. C.; Falter, C.; Abbott, M.; Scipio, D.; Furler, P.; Haile, (27) Besnard, M.; Cabaco, M. I.; Vaca Chavez, F.; Pinaud, N.;
S. M.; Steinfeld, A. High-flux solar-driven thermochemical dissociation Sebastiao, P. J.; Coutinho, J. A.; Mascetti, J.; Danten, Y. CO2 in 1-
of CO2 and H2O using nonstoichiometric ceria. Science 2010, 330, butyl-3-methylimidazolium acetate. 2. NMR investigation of chemical
1797−1801. reactions. J. Phys. Chem. A 2012, 116, 4890−4901.
(7) Costentin, C.; Drouet, S.; Robert, M.; Saveant, J. M. A local (28) Duong, H. A.; Tekavec, T. N.; Arif, A. M.; Louie, J. Reversible
proton source enhances CO2 electroreduction to CO by a molecular carboxylation of N-heterocyclic carbenes. Chem. Commun. 2004, 112−
Fe catalyst. Science 2012, 338, 90−94. 113.
(8) Gu, L. Q.; Zhang, Y. G. Unexpected CO2 splitting reactions to (29) Gurau, G.; Rodriguez, H.; Kelley, S. P.; Janiczek, P.; Kalb, R. S.;
form CO with N-heterocyclic carbenes as organocatalysts and Rogers, R. D. Demonstration of chemisorption of carbon dioxide in
aromatic aldehydes as oxygen acceptors. J. Am. Chem. Soc. 2010, 1,3-dialkylimidazolium acetate ionic liquids. Angew. Chem. Int. Ed.
132, 914−915. 2011, 50, 12024−12026.
(9) Parkinson, B. A.; Weaver, P. F. Photoelectrochemical pumping of (30) Hore, P. J. Nuclear magnetic resonance; Oxford University Press:
enzymatic CO2 reduction. Nature 1984, 309, 148−149. New York, 1995.
(10) Rosen, B. A.; Haan, J. L.; Mukherjee, P.; Braunschweig, B.; Zhu, (31) Ikeda, S.; Takagi, T.; Ito, K. Selective formation of formic-acid,
W.; Salehi-Khojin, A.; Dlott, D. D.; Masel, R. I. In situ spectroscopic oxalic-acid, and carbon-monoxide by electrochemical reduction of
examination of a low overpotential pathway for carbon dioxide carbon-dioxide. Bull. Chem. Soc. Jpn. 1987, 60, 2517−2522.
conversion to carbon monoxide. J. Phys. Chem. C 2012, 116, 15307− (32) Gressin, J. C.; Michelet, D.; Nadjo, L.; Saveant, J. M.
15312. Electrochemical reduction of carbon-dioxide in low proton media.
(11) Rosen, B. A.; Salehi-Khojin, A.; Thorson, M. R.; Zhu, W.; Nouv. J. Chem.New J. Chem. 1979, 3, 545−554.
Whipple, D. T.; Kenis, P. J.; Masel, R. I. Ionic liquid-mediated selective (33) Kaiser, U.; Heitz, E. Mechanism for electrochemical
conversion of CO2 to CO at low overpotentials. Science 2011, 334, dimerization of CO2 into oxalic-acid. Ber. Bunsen-Ges. Phys. Chem.
643−644. Chem. Phys. 1973, 77, 818−823.
(12) Eneau-Innocent, B.; Pasquier, D.; Ropital, F.; Léger, J. M.; (34) Rosen, B. A.; Zhu, W.; Kaul, G.; Salehi-Khojin, A.; Masel, R. I.
Kokoh, K. B. Electroreduction of carbon dioxide at a lead electrode in Water enhancement of CO2 conversion on silver in 1-ethyl-3-
propylene carbonate: A spectroscopic study. Appl. Catal., B: Environ. methylimidazolium tetrafluoroborate. J. Electrochem. Soc. 2012, 160,
2010, 98, 65−71. H138−H141.
(13) Spinner, N. S.; Vega, J. A.; Mustain, W. E. Recent progress in the (35) Tezuka, M.; Yajima, T.; Tsuchiya, A.; Matsumoto, Y.; Uchida,
electrochemical conversion and utilization of CO2. Catal. Sci. Technol. Y.; Hidai, M. Electroreduction of carbon-dioxide catalyzed by iron-
2012, 2, 19. sulfur clusters [Fe4S4(SR)4]2‑. J. Am. Chem. Soc. 1982, 104, 6834−
(14) Schneider, J.; Jia, H. F.; Muckerman, J. T.; Fujita, E. 6836.
Thermodynamics and kinetics of CO2, CO, and H+ binding to the (36) Chae, K. J.; Choi, M.; Ajayi, F. F.; Park, W.; Chang, I. S.; Kim, I.
metal centre of CO2 reduction catalysts. Chem. Soc. Rev. 2012, 41, S. Mass transport through a proton exchange membrane (Nafion) in
2036−2051. microbial fuel cells. Energy Fuels 2008, 22, 169−176.
(15) Bockris, J. O.; Wass, J. C. The photoelectrocatalytic reduction of (37) Lu, Z. J.; Polizos, G.; Macdonald, D. D.; Manias, E. State of
carbon-dioxide. J. Electrochem. Soc. 1989, 136, 2521−2528. water in perfluorosulfonic ionomer (Nafion 117) proton exchange
(16) Chandrasekaran, K.; Bockris, J. O. In situ spectroscopic membranes. J. Electrochem. Soc. 2008, 155, B163−B171.
investigation of adsorbed intetmediate radicals in electrochemical (38) Pavlishchuk, V. V.; Addison, A. W. Conversion constants for
reactionsCO2− on platinum. Surf. Sci. 1987, 185, 495−514. redox potentials measured versus different reference electrodes in
(17) Costentin, C.; Robert, M.; Saveant, J. M. Catalysis of the acetonitrile solutions at 25 degrees C. Inorg. Chim. Acta 2000, 298,
electrochemical reduction of carbon dioxide. Chem. Soc. Rev. 2013, 42, 97−102.
2423−2436.
(18) Yan, Y.; Zeitler, E. L.; Gu, J.; Hu, Y.; Bocarsly, A. B.
Electrochemistry of aqueous pyridinium: Exploration of a key aspect of
electrocatalytic reduction of CO2 to methanol. J. Am. Chem. Soc. 2013,
135, 14020−14023.
(19) Blanchard, L. A.; Hancu, D.; Beckman, E. J.; Brennecke, J. F.
Green processing using ionic liquids and CO2. Nature 1999, 399, 28−
29.
(20) Brennecke, J. E.; Gurkan, B. E. Ionic liquids for CO2 capture and
emission reduction. J. Phys. Chem. Lett. 2010, 1, 3459−3464.

6308 dx.doi.org/10.1021/la5009076 | Langmuir 2014, 30, 6302−6308

You might also like