Is the Imidazolium Cation a Unique Promoter for Electrocatalytic

Sep 21, 2016 - *E-mail: [email protected]. Tel: +61 3 9545 8866. Fax: +61 3 9562 8919., *E-mail: [email protected]. Tel: +61 3 99056289. Fax: +61...
0 downloads 0 Views 989KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Is the Imidazolium Cation a Unique Promoter for Electrocatalytic Reduction of Carbon Dioxide? Shu-Feng Zhao, Mike Horne, Alan M Bond, and Jie Zhang J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.6b08182 • Publication Date (Web): 21 Sep 2016 Downloaded from http://pubs.acs.org on September 25, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Is the Imidazolium Cation a Unique Promoter for Electrocatalytic Reduction of Carbon Dioxide?



∗†

Shu-Feng Zhao,†‡ Mike Horne, ‡ Alan M. Bond† and Jie Zhang



School of Chemistry and Australian Research Council Centre of Excellence for Electromaterials Science, Monash University, Clayton, Victoria, 3800, Australia.



CSIRO Process Science and Engineering, Box 312, Clayton South, Victoria, 3168, Australia.



School of Chemistry and Australian Research Council Centre of Excellence for Electromaterials Science, Monash University, Clayton, Victoria, 3800, Australia. E-mail: [email protected]; Tel: +61 3 99056289; Fax: +61 3 99054597.



CSIRO Process Science and Engineering, Box 312, Clayton South, Victoria, 3168, Australia. Email: [email protected]; Tel: +61 3 9545 8866; Fax: +61 3 9562 8919.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 39

Abstract There has been considerable recent interest in the use of the imidazolium cation as a promoter in the heterogeneous and homogeneous electrocatalysis of CO2 reduction. However, despite its widespread use for this purpose, the mechanism by which imidazolium operates is not yet fully established. The present work reveals that enhanced catalytic activity is achieved by addition of many cations other than imidazolium. Under cyclic voltammetric conditions at a Ag electrode in acetonitrile solutions (0.1 M n-Bu4NPF6), 2.0 mM concentrations of imidazolium,

pyrrolidium,

ammonium,

phosphonium

and

(trimethylamine)-

(dimethylethylamine)-dihydroborate cations can all enhance the kinetics of catalytic CO2 reduction with imidazolium and pyrrolidium being the most active. Analysis of the voltammetric data suggests that imidazolium cations achieve their impact by directly acting as co-catalysts with Ag whereas the other cations affect the reaction rate by modifying the electrochemical double layer. The results also confirm that the active form of the co-catalyst is the reduced imidazolium radical which forms a complex with CO2 before being further reduced to CO or other products at Ag and not an imidazolium carboxylate formed between an imidazolium carbene and CO2. In fact, imidazolium is deactivated during CO2 reduction by the latter reaction. Addition of water inhibits this deactivation pathway allowing the imidazolium cation to act as in a long term for CO2 reduction. In contrast, the pyrrolidium cation, where enhanced catalysis is attributed to an electrochemical double layer effect, retains its catalytic activity for very long periods of time regardless of the presence or absence of water.

2 ACS Paragon Plus Environment

Page 3 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1. Introduction An efficient electrochemical route for the transformation of CO2 into valuable chemicals and fuels is a much sought after goal in chemical processing. If coupled with renewable energy sources, the electrochemical transformation of CO2 will not only translate a currently intractable waste material into a valuable chemical feedstock, but could help to address the persistent problem of energy storage from intermittent renewable energy sources. Unfortunately, CO2 is chemically inert. The first one-electron reduction process to form CO2.is kinetically slow and occurs at very negative potentials. Consequently, many potential catalysts including heterogeneous (such as metals1-2, metal oxides2 and non-metals,3-5) and homogeneous ones (such as metal complexes6) have been tested to see whether they lower the overpotential for CO2 reduction in protic and aprotic media. To improve the energy efficiency of the electrocatalytic reduction process, imidazoliumbased ionic liquids (ILs) also have been introduced as promoters7-8. They are promising for two reasons: they lower the reduction overpotential and they also suppress the competing H2 evolution reaction.9-20 Masel and co-workers20 reported that 18 mol% of 1-ethyl-3methylimidazolium tetrafluoroborate ([C2mim][BF4]) lowers the overpotential for CO formation on a silver (Ag) electrode in aqueous solutions by about 600 mV. Studies with in situ sum frequency generation spectroscopy indicate that a layer of [C2mim]+ adsorbed on the electrode surface promotes CO2 conversion to CO on Ag and platinum (Pt) electrodes14, 17. The mechanism proposed (Scheme 1A) involves the formation of an adsorbed C2mim-CO2 complex — a key intermediate that lowers the overpotential for the reduction17. In this context, imidazolium cations can be said to effectively promote the formation of CO. Rosenthal and co-workers9, 11, 16 also found that imidazolium facilitates for reduction of CO2 to CO on inexpensive and easily prepared post-transition metal electrodes in dry MeCN. A mechanism proposed in dry MeCN containing an imidazolium-based IL as an electrolyte 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 39

implies that the C-2 hydrogen on the imidazolium cation is an available source of protons and facilitates a thermodynamically more favourable proton coupled electron transfer reaction to form CO (Scheme 1B).21 The enhanced acidity of the C-2 hydrogen under a CO2 atmosphere was also reported in studies of the electroreduction of acetophenone.22 However, in all these studies, the concentration of the imidazolium-based ionic liquids was at least 0.1 M. Under these conditions, a significant proton concentration is available for CO formation, making it difficult to verify the role of imidazolium as a promoter during CO2 reduction. Furthermore, the presence of a high concentration of ionic liquid could enhance the solubility of CO2 in the molecule solvent and hence influence the mechanism for CO2 reduction.23

Scheme 1. Mechanisms proposed in references17, 21 for electrochemical reduction of CO2 in aqueous (A) and dry MeCN (B) solutions containing imidazolium-based ionic liquids as electrolytes. The lack of consensus concerning the mechanism by which imidazolium-based ionic liquids act during CO2 reduction poses several questions: (1) What is the role of the imidazolium cation other than a proton source? (2) Do substituents on the imidazolium cation affect its catalytic activity and stability? (3) Does the electrolyte anion influence the CO2 reduction reaction? (4) Is the catalytic effect specific to imidazolium cations? This work attempts to answer these questions. In order to test whether or not imidazolium cations do act

4 ACS Paragon Plus Environment

Page 5 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

as co-catalysts by the mechanism described in Scheme 1A, only small concentrations (2.0 mM) were added to the aprotic solvent MeCN containing 0.1 M tetrabutylammonium hexafluorophosphate (n-Bu4NPF6) as the supporting electrolyte. The small concentration of additive means that (1) the available proton concentration from the imidazolium cation is negligible compared with the amount of CO2 (0.27 M at 1 atm24) and (2) the presence of the additive will not modify the CO2 solubility appreciably, which is also confirmed by the fact that the CO2 reduction peak current remains essentially unchanged upon addition of the additives (Figure 3, vide infra). Ag was chosen as the electrode material since it catalyses the formation of CO from CO2 with high faradaic efficiency.1 To systematically investigate the potential role that imidazolium cations play as co-catalysts or promoters in this reaction, the catalytic activity of 20 compounds, including ionic liquids from several commonly-used classes as well as alkylammonium salts, was investigated. The structures of their constituent ions are shown in Scheme 2.

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 39

Scheme 2. Structures of the constituent ions of the additives used as potential co-catalysts for the electrochemical reduction of CO2. The abbreviations and their meanings are as follows: [BF4]-: tetrafluoroborate, [TFSI]-: bis(trifluoromethylsulfonyl)imide, [Cxmim]+: 1--3-methylimidazolium,

[C4mpyrd]+:

1-butyl-1-methylpyrrolidinium,

[N1114]+:

trimethylbutylammonium, [P2225]+: triethylpentylphosphonium, [NHHB]+: (trimethylamine)(dimethylethylamine)-dihydroborate, triethylsulfonium,

[C6DABCO]+:

[C3mpy]+:

N-propyl-2-methylpyridinium,

1-hexyl-1,4-diazabicyclo[2.2.2]octanium,

[S222]+: Me4NCl:

tetramethylammonium chloride, CTAB: cetyltrimethylammonium bromide, n-Hex4NBr: tetrahexylammonium bromide, n-Bu4NTfO: tetrabutylammonium triflate, n-Bu4NBF4: tetrahexylammonium tetrafluoroborate, n-Bu4NBr: tetrabutylammonium bromide.

6 ACS Paragon Plus Environment

Page 7 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2. Experimental section 2.1 Chemicals The ionic liquids [C4mim][BF4], [C10mim][BF4], [C10mim][TFSI], [C16mim][BF4], [C4mpyrd][BF4], [C4mpyrd][TFSI] and [C3mPy][TFSI] were purchased from IOLITEC (Germany). [N114][TFSI], [S222][TFSI], [C1mim-CO2] (≥ 80%), CTAB, Me4NCl, n-Hex4NBr, n-Bu4NTfO, n-Bu4NBF4, n-Bu4NBr and ferrocene (Fc) (≥ 98%) were purchased from Sigma Aldrich. [P2225][TFSI] was purchased from KANTO (Japan). MeCN and dimethyl sulfoxided6 (DMSO-d6) were purchased from Merck (Germany). The ionic liquids were dried over basic alumina for at least 24 h, then placed under vacuum at 80 oC for 24 h prior to use. [C6DABCO][TFSI] and [NHHB][TFSI] were synthesized by Dr. Thomas Ruether (CSIRO). n-Bu4NPF6 (GFS Chemicals, Inc, 98%) was recrystallised twice from ethanol prior to use. Other reagents were analytical grade and used as supplied by the manufacturer.

2.2 Voltammetric measurements Voltammetric measurements were undertaken at 20 ± 2

o

C using a CHI 700

electrochemical workstation (CH Instruments, Texas, USA) and a standard three-electrode cell. For transient cyclic voltammetric measurements, a 3.0 mm diameter Ag disk electrode or a 1.0 mm diameter glassy carbon (GC) electrode were used as the working electrodes. A Pt wire was used as a counter electrode. A Pt wire mounted into a capillary sealed with a sintered glass disk served as a quasi-reference electrode. The solution used in the reference compartment was MeCN (0.1 M n-Bu4NPF6). To convert the quasi-reference potential scale to a known scale, 0.5 mM Fc was added to each experiment and the voltammetric limits were extended to record the reversible Fc/Fc+ process. Analysis of the data allows potentials to be converted to the Fc/Fc+ scale. The working electrode was polished with an aqueous suspension of 0.3 µm alumina on a polishing cloth (Buehler), sonicated in deionized water,

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 39

rinsed sequentially with deionized water and acetone, and dried under a flow of nitrogen (N2) before use.

2.3 Bulk electrolysis After the solution was saturated with CO2, controlled potential bulk electrolysis experiments were undertaken at room temperature (20 ± 2 oC), under a CO2 atmosphere in a gas-tight two-compartment divided cell with a Ag wire as the cathode and Pt wire as the anode. The volume for each compartment was 29 mL. The reference electrode was the same as that used in the voltammetric measurements. Both compartments were filled with 10 mL MeCN solution (0.1 M n-Bu4NPF6) containing 2.0 mM of the chosen ionic liquid and a constant stream of CO2 was passed through the cell to achieve a saturated solution before the electrolysis. During electrolysis, the solutions in both compartments were stirred. The volatile reduction products were identified by withdrawing 200 µL samples from the head space of the cathode compartment using a gas-tight syringe and injecting these samples into a gas chromatograph (Agilent 7820A) equipped with an Agilent HP-5 column. Helium was used as the carrier gas for CO detection and N2 as the carrier gas for H2 detection. The liquid phase reduction products were quantified using 1H and 13C NMR (400 MHz, Bruker) spectroscopy. Following bulk electrolyses, 400 µl of the condensed solution was withdrawn and 50 µl pxylene was added as an internal standard. 400 µl DMSO-d6 was also added to provide a deuterated solvent. 1H NMR spectra were used to quantify the concentrations of formate and imidazolium-based species, whereas

13

C NMR spectra were used to quantify the oxalate

concentration.

8 ACS Paragon Plus Environment

Page 9 of 39

3. Results and discussion 3.1 Voltammetric measurements Voltammetric studies were undertaken to investigate the influence of the ions listed in Scheme 2 on the electrocatalytic reduction of CO2 at a Ag electrode. When needed, control experiments were undertaken using a GC electrode. 3.1.1 Catalysis with [C10mim]+ 2.0 mM [C10mim][BF4]-N2

N2 0.0

2.0 mM [C10mim][BF4]-CO2 -0.2

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CO2

-0.4

-0.6 -3.5

-3.0

-2.5

-2.0

-1.5

-1.0

+

E (V vs. (Fc/Fc ))

Figure 1. Cyclic voltammograms obtained with a 3.0 mm diameter Ag electrode at a scan rate of 0.1 V s-1 in the absence and presence of 2.0 mM [C10mim][BF4] under N2 and CO2 atmospheres in MeCN (0.1 M n-Bu4NPF6). In initial studies, [C10mim][BF4] was chosen to investigate the effect of the imidazolium cation on CO2 reduction at a Ag electrode in MeCN (0.1 M n-BuN4PF6). Voltammograms obtained under a N2 atmosphere in the presence of 2.0 mM [C10mim][BF4] display two 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 39

irreversible reduction processes prior to the cathodic potential limit with peak potentials (Ep1,N2 and Ep2,N2) of -2.60 V and -2.85 V vs. Fc/Fc+ (blue trace, Figure 1). These observations are consistent with other reports on the reduction of the [C4mim]+ cation which was shown to undergo a series of one electron reduction and dimerization processes to form a range of products including N-heterocyclic carbene, through a mechanism described in Scheme 325-28.

Scheme 3. Electrochemical reduction of [C4mim][BF4] at a Pt electrode25-27. Under a CO2 atmosphere, the voltammogram is significantly different to that under N2 with only one irreversible reduction process being evident. The a peak potential (Ep,CO2) value of -2.30 V vs. Fc/Fc+ (green trace, Figure 1) is about 0.3 V more positive than Ep1,N2 detected under a N2 atmosphere. As this reduction process is not observed in the absence of [C10mim][BF4], it is attributed to the reduction of either [C10mim]+ or CO2 followed by the formation of a reduced imidazolium-CO2 complex, which according to a previous study,10 is influenced by the mass-transport associated with imidazolium. Since the reduction peak current (Ip,CO2) magnitude is about five times that observed for the first reduction process under a N2 atmosphere (Ip1,N2), the enhancement cannot result solely from an increase in the number of electron transferred (e.g. 2 e- vs. 1 e- process). Moreover, the onset potential for 10 ACS Paragon Plus Environment

Page 11 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CO2 reduction (Eonset,CO2) in the presence of [C10mim][BF4] is also significantly more positive (-1.95 vs. -2.27 vs. Fc/Fc+). This onset potential is derived from the intercept of a linear extrapolation of the low current flat and steep large current regions of the voltammogram. A similar positive shift was also observed in the electrochemical reduction of CO2 at a Bimodified electrode in the presence of 20 mM [C2mim][BF4]16. These observations suggest a different reduction mechanism occurs in the prsence of CO2 than under N2 and this is most probably the catalytic process reported by Wang et. al. in which a reduced imidazolium-CO2 complex is further reduced to regenerate reduced imidazolium29 (Scheme 4). As will be confirmed later, the reaction ultimately leads to the reduction of CO2, so from now on this reduction process is referred to as “CO2 reduction”.

Scheme 4 Mechanism proposed by Wang et. al.29 for electrochemical activation of CO2 by ionic liquid [C2mim][BF4]. As highlighted in the Introduction, much higher concentrations of imidazolium-based ionic liquids (typically 0.1 M) have been used as supporting electrolytes in previous studies on the electrochemical reduction of CO29, 11, 16, 20, 30. Therefore, voltammetric experiments were also undertaken to establish the influence of the concentration of [C10mim][BF4] (Figure S1). The key parameters extracted from the voltammetric data are summarized in Table 1. 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 39

The onset potentials (Eonset,N2 or Eonset,CO2) are almost independent of the concentration of [C10mim][BF4] (CIL). The negative shift in Ep,CO2 as the concentration of [C10mim][BF4] increases can be attributed to an increase IRu drop (Ru is typically 250 Ω), associated with the concentration dependent increase in the peak current (Ip,CO2). Importantly, when the concentration of [C10mim][BF4] is doubled from 2.0 to 4.0 mM the Ip,CO2:Ip1,N2 ratio almost remains unchanged, implying the reaction occurs under pseudo first order conditions associated with a large concentration excess of CO2. However, when the IL concentration is increased to 50.0 and 100.0 mM the ratio decreases due to the mass transport limitation associated with partial depletion of CO2. These data are consistent with theoretical predictions for a reaction catalysed by a dissolved electron transfer mediator. Table 1. Key parameters extracted from the voltammetric data obtained at a scan rate of 0.1 V s-1 under N2 or CO2 atmospheres in the presence and absence of [C10mim][BF4]. CIL/mM Ep1,N2/V Eonset,N2/V Ep,CO2/V Eonset,CO2/V Ip,CO2/Ip1,N2 2.0 -2.60 -2.28 -2.30 -1.95 5.25 4.0 -2.62 -2.28 -2.34 -1.93 5.69 50.0 -2.85 -2.22 -2.74 -1.87 2.15 100.0 -2.96 -2.30 -2.84 -1.87 1.92

3.1.2 The influence of other imidazolium-based ionic liquids The imidazolium ionic liquids, [C4mim][BF4] and [C16mim][BF4] were chosen to ascertain if the alkyl chain length of the substituents on the imidazolium affects the catalytic activity for CO2 reduction at a Ag eletrode. Cyclic voltammetric measurements were again undertaken under either N2 or CO2 atmospheres and in the absence and presence of 2.0 mM ionic liquid. With the exception of the magnitude of Ip,CO2, the voltammetry was almost independent of the identity of the ionic liquid, as shown in Figure S2. Under a N2 atmosphere, two reduction processes were observed. The Ep1,N2 values are about -2.60 V vs. Fc/Fc+ in all ionic liquids and Ip,CO2/Ip1,N2 ratios are in the range 4.0~5.5 when a scan rate of 0.1 V s-1 is used. Under a CO2 atmosphere, both ionic liquids significantly promote the rate of

12 ACS Paragon Plus Environment

Page 13 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

CO2 reduction, as was the case with [C10mim][BF4]. As shown by inspection of data summarized in Table 2, the onset potential (Eonset,N2 or Eonset,CO2) for reduction of CO2 is not strongly dependent on the identity of the added ionic liquid. The Ip,CO2:Ip1,N2 ratio increases with chain length on the alkyl substituent when comparing C4 and C10 data, but the value for C10 and C16 are similar. Increasing the alkyl chain length increases the bulkiness of the imidazolium cation which diminishes the likelihood that the reduced imidazolium intermediate will dimerize (see Scheme 3). Since the [C10mim]+ and [C16mim]+ cations are more active in catalysing CO2 reduction, it is plausible that the monomers rather than dimers are the active forms of the co-catalysts or promoters. This hypothesis is consistent with that of Wang et, al.29 who proposed in Scheme 4 by quantum chemistry that a reduced monomeric form of the imidazolium cation containing a C2-hydrogen is the key intermediate in CO2 reduction. Table 2. Key parameters extracted from the voltammetric data obtained at a Ag electrode with a scan rate of 0.1 V s-1 in the presence of 2.0 mM [C4mim]+, [C10mim]+ and [C16mim]+based ionic liquids under N2 and CO2 atmospheres. IL Ep1,N2/V Eonset,N2/V Ep,CO2/V Eonset,CO2/V Ip,CO2/Ip1,N2 [C4mim][BF4] -2.65 -2.38 -2.30 -2.03 4.19 [C10mim][BF4] -2.60 -2.28 -2.30 -1.95 5.25 [C16mim][BF4] -2.60 -2.31 -2.32 -1.99 5.53

The influence of the anion on the catalytic activity of imidazolium-based ionic liquids for CO2 reduction also has been investigated. In this case, 0.1 M of either [C10mim][TFSI] or [C10mim][BF4] was used as both the supporting electrolytes and co-catalysts with n-Bu4NPF6 being absent. Cyclic voltammetric results obtained under N2 and CO2 atmospheres reveal that catalytic activity of [C10mim]+ is almost independent of either anion, as judged from the similarity of the onset potentials (Figure S3).

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 39

3.1.3 Electrochemical reduction of [C10mim]+ at a glassy carbon electrode The results presented above confirm that imidazolium cations accelerate the rate of CO2 reduction at a Ag electrode. However, to establish whether the cations act as the catalysts or co-catalyst with Ag, 2.0 mM [C10mim][BF4] in MeCN solutions were electrochemically reduced at a 1.0 mm diameter GC electrode under both N2 and CO2 atmospheres (GC has a low activity for CO2 reduction22). As shown in Figure 2, under a N2 atmosphere, two reduction processes are observed at -2.81 V and -3.05 V vs. Fc/Fc+, which is more negative than found at a Ag electrode (see Figures S1 and S2). Under a CO2 atmosphere, two reduction processes were again seen at -2.65 V and -2.78 V on the first negative potential sweep, but only one reduction process at -2.86 V on repetitive cycling of the potential. Presumably, reduction products generated on the first negative-going scan are irreversibly adsorbed onto the GC electrode and inhibit subsequent reduction reactions. Importantly, the positive shift in Eonset,CO2 and enhancement of Ip,CO2 is much less pronounced on GC leading to the conclusion that imidazolium cations act as a co-catalyst along with Ag for CO2 reduction at a Ag electrode. In additional to those reduction processes, an oxidation process was observed at -1.0 V. This process is attributed to the oxidation of an adduct of CO2.- and [C10mim.] radicals generated during co-reduction of both CO2 and [C10mim]+ according to a study by Hanc-Scherer et al..31

14 ACS Paragon Plus Environment

Page 15 of 39

0.00

-0.01

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1st cycle -0.02

-0.03 -3.5

-3.0

-2.5

-2.0

-1.5

-1.0

+

E (V vs. (Fc/Fc ))

Figure 2. Cyclic voltammograms of 2.0 mM [C10mim][BF4] obtained at a 1.0 mm diameter GC electrode with a scan rate of 0.1 V s-1 in MeCN (0.1 M n-Bu4NPF6) under N2 (─) and CO2 (─) atmospheres. 3.1.4 Catalysis with [C4mpyrd]+ and [C4mmim]+ Our data and other studies with imidazolium-based ionic liquids10, raise two important questions. Are other ionic liquid cations also capable of catalysing CO2 reduction? Does the C2-hydrogen of imidazolium play a critical role in CO2 reduction? To answer these questions, voltammetric studies were also undertaken using the pyrrolidinium ionic liquids [C4mpyrd][BF4] and [C4mpyrd][TFSI], and [C4mmim][BF4] as additives; these ionic liquids do not contain a reactive C2-hydrogen in the cation. Pyrrolidinium ionic liquids are well-known for their high electrochemical stability and low proton availability.32-33 15 ACS Paragon Plus Environment

The Journal of Physical Chemistry

Under a N2 atmosphere (black and blue curves, Figure 3), no significant reduction current associated with the presence of 2.0 mM [C4mpyrd]+ was detected within the potential window available in MeCN (0.1 M n-Bu4NPF6). Under a CO2 atmosphere, the onset potential for CO2 reduction shifted positively by 290 mV in the presence of 2.0 mM [C4mpyrd][BF4] or [C4mpyrd][TFSI], a similar magnitude to that observed with 2.0 mM imidazolium-based ionic liquids (320 mV). The catalytic activity of [C4mpyrd]+ for CO2 reduction is also independent of the BF4- or TFSI- these anions, as judged by the similarity of the Eonset,CO2. Even if mechanistic details differ, the degrees by which imidazolium and pyrrolidinium cations promote CO2 reduction are comparable and hence catalysis of CO2 reduction is not unique to imidazolium-based ionic liquids.

0.5

2.0 mM [C4mpyrd][BF4]-N2

0.0 -0.5

2.0 mM [C4mpyrd][TFSI]-N2

-1.0

2.0 mM [C10mim][BF4]-CO2

0.0

I (mA)

-0.2

-1.5

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 39

-2.0 -0.4

CO2

-2.5

-2.8

-2.4

-2.0 +

E (V vs. (Fc/Fc ))

-3.0 -3.5

2.0 mM [C4mpyrd][TFSI]-CO2

-4.0

2.0 mM [C4mpyrd][BF4]-CO2 -5

-4

-3

-2

-1

+

E (V vs. (Fc/Fc ))

Figure 3. Cyclic voltammograms obtained with a 3.0 mm diameter Ag electrode in MeCN (0.1 M n-Bu4NPF6) at a scan rate of 0.1 V s-1 with 2.0 mM [C4mpyrd][BF4] or [C4mpyrd][TFSI] and under either N2 or CO2 atmospheres. Cyclic voltammograms of CO2 in

16 ACS Paragon Plus Environment

Page 17 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

the absence and presence of 2.0 mM [C10mim][BF4] recorded under the same conditions were shown for comparison. When the concentration of [C4mpyrd]+ is increased up to 100.0 mM, Eonset,CO2 shifts positively (Figure S4), which differs from results with [C10mim]+ where Eonset,CO2 is almost independent of its concentration. Furthermore, the magnitude of Ip,CO2 is minimally dependent on [C4mpyrd]+ concentration, which suggests that CO2 reduction in the presence of this cation is mass transport limited, whereas it is kinetically limited in the presence of imidazolium. It is highly unlikely that [C4mpyrd]+ acts as a redox mediator as the rate would have to be improbably high for the reduction process to be mass transport controlled at low [C4mpyrd]+ concentrations, based on a semi-quantitative estimation using simulation of the voltammograms obtained with the DigiElch software package (results not shown). Since CO2 reduction occurs at a very negative potential which has to be much more negative than the potential of zero charge for a Ag electrode in MeCN, the electrochemical double layer will be dominated by cations. Since CO2 reduction is an inner sphere process, the cations within the electrochemical double layer are expected to play an important role in the reduction reaction. Because [C4mpyrd]+ is smaller than [n-Bu4N]+, it may displace the latter cation in the electrochemical double layer. This can allow CO2 to approach the electrode surface more closely and facilitate electron transfer. That is, the heterogeneous charge transfer constant for the CO20/.- process may increase, as has been reported for other reactions.34-35 In addition to inducing an electrochemical double layer effect, [C4mpyrd]+ also may promote the kinetics of CO2 reduction by providing a nitrogen site for binding CO2 and stabilizing the intermediate CO2.- radical anion. Electrochemical reduction of 2.0 mM [C4mmim][BF4] (Figure 4) at a Ag electrode under a N2 atmosphere exhibited a reduction process with a peak potential of -3.02 V vs. Fc/Fc+.

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry

The voltammogram obtained under a CO2 atmosphere had an Eonset,CO2 value of -2.02 V vs. Fc/Fc+, which is similar to those obtained in the presence of pyrrolidinium-based ionic liquids, suggesting that [C4mmim]+ also modifies the electrochemical double layer in a manner that favours CO2 reduction.

N2

2.0 mM [C4mmim][BF4]-N2

0

-1

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 39

CO2 2.0 mM [C4mmim][BF4]-CO2

-2

-3

-3.5

-3.0

-2.5

-2.0

-1.5

-1.0

+

E (V vs. (Fc/Fc ))

Figure 4. Cyclic voltammograms obtained with a 3.0 mm diameter Ag electrode at a scan rate of 0.1 V s-1 in MeCN (0.1 M n-Bu4NPF6) under N2 (─) and CO2 (─) atmospheres and in the presence of 2.0 mM [C4mmim][BF4] under N2 (─)and CO2 (─). 3.1.5 Catalysis with other ionic liquid cations To further explore the catalytic influence of other ionic liquid cations on CO2 reduction, the voltammetry in the presence of 2.0 mM concentrations of six other ionic liquids containing [TFSI]- as a common anion (structures are provided in Scheme 2) were studied in MeCN (0.1 M n-Bu4NPF6). Under a N2 atmosphere, no reduction was observed prior to the cathodic potential limit for MeCN (0.1 M n-Bu4NPF6) in the presence of [N1114][TFSI], 18 ACS Paragon Plus Environment

Page 19 of 39

([P2225][TFSI]), [NHHB][TFSI] and [C6DABCO][TFSI] (black curves, Figure 5). However, under a CO2 atmosphere, the onset potentials for CO2 reduction again shifted to significantly more positive values and the peak currents had mass transport limiting current characteristics (Figure 5). Thus the cations associated with all these ionic liquids also promote the electrochemical catalysis of CO2 reduction in a manner similar to [C4mpyrd]+. Interestingly, the magnitude of the potential shift is significantly larger for ionic liquids containing nitrogen heteroatoms than for those with phosphonium-based ones; implying CO2 interacts more strongly with the more electronegative nitrogen heteroatom.

2.0 mM [N1114][TFSI]-N2

2.0 mM [P2225][TFSI]-N2

0

0

-1

2.0 mM [N1114][TFSI]-CO2 I (mA)

I (mA)

-1

CO2

-2

-3

2.0 mM [P2225][TFSI]-CO2 -2

-3

-4

CO2

-4 -3.5

-3.0

-2.5

-2.0

-1.5

-1.0

-0.5

-3.5

2.0 mM [NHHB][TFSI]-N2 0

0

-1

-1

2.0 mM [NHHB][TFSI]-CO2 -2

CO2

-3

-3.0

-2.5

-2.0

-1.5

-1.0

-0.5

2.0 mM [C6DABCO][TFSI]-N2

I (mA)

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2.0 mM [C6DABCO][TFSI]-CO2 CO2

-2

-3

-4

-4 -3.5

-3.0

-2.5

-2.0

-1.5

-1.0

-0.5

-3.5

-3.0

+

E (V vs. (Fc/Fc ))

-2.5

-2.0

-1.5

-1.0

-0.5

+

E (V vs. (Fc/Fc ))

Figure 5. Cyclic voltammograms of 2.0 mM ionic liquids including [N1114][TFSI], [P2225][TFSI], [NHHB][TFSI] and [C6DABCO][TFSI] obtained at a 3.0 mm diameter Ag electrode with a scan rate of 0.1 V s-1 under N2 (─) and CO2 (─) atmospheres in MeCN (0.1 19 ACS Paragon Plus Environment

The Journal of Physical Chemistry

M n-Bu4NPF6). Cyclic voltammogram of CO2 (─) in MeCN (0.1 M n-Bu4NPF6) obtained at a 3.0 mm Ag electrode with a scan rate of 0.1 V s-1 is included in each figure for comparison. For solutions containing 2.0 mM [C3mpy][TFSI]) or [S222][TFSI] and under a N2 atmosphere, irreversible reduction processes associated with the cations were found at -1.76 V and at -2.21 V vs. Fc/Fc+, respectively (Figure 6). However, in these cases, the irreversible reduction process remains unaffected under a CO2 atmosphere (Figure 6), indicating the absence of interaction between the reduced species and CO2. Furthermore, if the positive charge on the cation is removed upon reduction, it cannot modify the electrochemical double layer.

2.0 mM [C3mpy][TFSI]-N2

I (mA)

0.0

2.0 mM [C3mpy][TFSI]-CO2 CO2

-0.2

-0.4

2.0 mM [S222][TFSI]-N2 0.0

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 39

2.0 mM [S222][TFSI]-CO2

-0.2

CO2 -0.4 -3.0

-2.5

-2.0

-1.5

+

E (V vs. (Fc/Fc ))

Figure 6. Cyclic voltammograms of 2.0 mM [C3mpy][TFSI] and [S222][TFSI] obtained at a 3.0 mm diameter Ag electrode with a scan rate of 0.1 V s-1 under N2 (─) and CO2 (─) atmospheres in MeCN (0.1 M n-Bu4NPF6). Cyclic voltammogram of CO2 (─) obtained under the same conditions is included in each figure for comparison. 20 ACS Paragon Plus Environment

Page 21 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

In summary, the Eonset,CO2 values obtained in the presence of 2.0 mM ionic liquid promoters, decrease in the order: [C10mim]+ (-1.95 V) > [C4mpyrd]+ (-1.98 V) > [C16mim]+ (1.99 V) > [N1114]+ (-2.0 V) > [C4mmim]+ (-2.02 V) > [C4mim]+ (-2.03 V) > [C6DABCO]+ (2.04 V) > [NHHB]+ (-2.11 V) > [P2225]+ (-2.17 V), suggesting that [C10mim]+ and [C4mpyrd]+ are superior for this application. 3.1.6 Catalysis with ammonium cations The results presented above show that catalysis of CO2 reduction by addition of ionic liquids is not restricted to imidazolium based ones. To test if the cation induced catalysis is general, three simple ammonium-based salts having variable alkyl lengths in their cations were also examined with respect to their influence for the electroreduction of CO2 at a Ag electrode. Voltammetric results from studies of 2.0 mM Me4NCl, CTAB and n-Hex4NBr in MeCN (0.1 M n-Bu4NPF6) under N2 and CO2 atmospheres are shown in Figure 7. Under a N2 atmosphere, Me4N+ form Me4NCl is reduced at -2.86 V vs. Fc/Fc+. By contrast, the other two additives n-Hex4NBr and CTAB are more difficult36 to reduce based on their reduction onset potentials of -2.95 and -3.15 V vs. Fc/Fc+.

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry

2.0 mM Me4NCl-N2

I (mA)

0 -1

CO2

-2

2.0 mM Me4NCl-CO2

I (mA)

-3 0 -1

2.0 mM CTAB-N2

CO2 2.0 mM CTAB-CO2

-2 -3 0

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 39

2.0 mM n-Hex4NBr-N2 CO2

-1

2.0 mM n-Hex4NBr-CO2

-2 -3 -3.5

-3.0

-2.5

-2.0

-1.5

-1.0

-0.5

+

E (V vs. (Fc/Fc ))

Figure 7. Cyclic voltammograms recorded on a 3.0 mm diameter Ag electrode with a scan rate of 0.1 V s-1 in MeCN (0.1 M n-Bu4NPF6) containing 2.0 mM Me4NCl, CTAB and nHex4NBr, respectively, under N2 (─) and CO2 (─) atmospheres. Cyclic voltammogram of CO2 (─) in MeCN (0.1 M n-Bu4NPF6) under the same conditions is included in each figure for comparison. Under a CO2 atmosphere, Eonset,CO2 shifts positively in the presence of Me4NCl or CTAB and the CO2 reduction process is mass transport controlled. Again, catalysis is attributed to a modification of the electrochemical double layer. Even though [Me4N]+ can be electrochemically reduced to a compound that could act as an electron transfer mediator, any contribution from this mechanism should be insignificant. No impact on CO2 reduction was 22 ACS Paragon Plus Environment

Page 23 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

observed in the presence of 2.0 mM n-Hex4NBr. However, this cation is larger and has a lower charge density than [n-Bu4N]+, so at a 2.0 mM concentration it is not expected to modify the double layer to any significant extent. To systematically vary the size of the cation and identity of the anion in ammoniumbased salts, voltammetric reduction of CO2 in MeCN was undertaken in 0.1 M solutions of six tetraalkylammonium salts. The cation size difference was probed by cyclic voltammograms obtained in the presence of n-Hex4NBr, n-Bu4NPF6 and Et4NI and the anion identity using n-Bu4NPF6, n-Bu4NBF4 and n-Bu4NBr. Results are summarized in Figure S5. Utilizing this tetraalkylammonium group, Eonset,CO2 was most positive when the smallest cation was used ([n-Et4N]+) and became more negative as the cation size increased. This trend is consistent with an electrochemical double layer effect since the smaller tetraalkylammonium cations also have a higher charge density and they can more effectively stabilize CO2.-, which facilitates CO2 reduction. The identity of the anions ([PF6]-, [BF4]-, [Br]-) used does not significantly affect the reduction reaction with the [n-Bu4N]+ series of salts (inset Figure in Figure S5). 3.1.7 The influence of water Proton coupling to electron transfer is known to lower the overpotential for CO2 reduction1. Adventitious water, a proton source, is always likely to present and may play a role. The impact of deliberately added water on the reduction of CO2 in the presence of 2.0 mM [C10mim][BF4] or [C4mpyrd][BF4] in MeCN (0.1 M n-Bu4NPF6) was therefore examined. As described above, the reduction of 2.0 mM [C10mim][BF4] with no deliberately added water and in the presence of CO2 exhibits an irreversible catalytic reduction process with an Ep,CO2 value of -2.30 V vs. Fc/Fc+ (black trace, Figure 8a). Upon addition of 0.5% water, the

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry

peak current increases significantly confirming that water promotes the catalytic activity of imidazolium. However, the potential for direct reduction of CO2 at a Ag electrode also becomes more positive (olive green trace, Figure 8a), so that the mediated and direct CO2 reduction processes overlap. Upon addition of more H2O, the extent of overlap increased and with 2.0% added H2O, the two processes fully merged giving rise to an overall mass transport controlled process with an Ep,CO2 value of -3.09 V. However, the addition of water does not significantly alter the onset potential for CO2 reduction.

(a) 2.0 mM [C10mim][BF4] 0

without H2O -1

0.5% H2O I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 39

-2

2.0% H2O -3

1.0% H2O without IL + 0.5% H2O -4

without IL + 1.0% H2O -4.0

-3.5

-3.0

-2.5

-2.0

-1.5

-1.0

+

E (V vs. (Fc/Fc ))

24 ACS Paragon Plus Environment

Page 25 of 39

(b) 2.0 mM [C4mpyrd][BF4] 0

-1

I (mA)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

-2

2.0% H2O 1.0% H2O without H2O

-3

0.5% H2O

-4 -3.5

-3.0

-2.5

-2.0

-1.5

-1.0

+

E (V vs. (Fc/Fc ))

Figure 8. The impact of water on cyclic voltammograms obtained with a 3.0 mm diameter Ag electrode at a scan rate of 0.1 V s-1 in MeCN (0.1 M n-Bu4NPF6) containing 2.0 mM of [C10mim][BF4] (a) or [C4mpyrd][BF4] (b) under a CO2 atmosphere: no added water (─), 0.5% added water (─), 1.0% added water (─), 2.0% added water (─) and without IL but with 0.5% (─) or 1.0% (─) added water. Results from experiment with 2.0 mM [C4mpyrd][BF4] under a CO2 atmosphere are shown in Figure 8b. Without added water, introduction of the pyrrolidinium cation shifts the onset potential for CO2 reduction in a positive direction and catalyses CO2 reduction. When water is now added to this solution, the peak current and the onset potential are almost unchanged. This different response to that with imidazolium reinforces the conclusion drawn previously that the mechanisms by which pyrrolidinium and imidazolium cations catalyse CO2 reduction are not the same. The slightly diminishes peak current found when 2.0% H2O was added may reflect a decrease in the solubility of CO2 as the water concentration increases37.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 39

3.2 Bulk electrolysis and mechanistic analysis Catalysis of CO2 reduction can affect not only voltametric characteristics, but may also influence the product distribution. To investigate the identity of the cation additive in this context, bulk electrolysis experiments were undertaken in a divided cell using a large surface area Ag wire as the cathode in MeCN (0.1 M n-Bu4NPF6) along with 2.0 mM [C10mim][BF4] or [C4mpyrd][BF4]. Water (1.0%) was deliberately added in some of these experiments. Unless otherwise stated, the applied potential (Eappied) was -2.30 V vs. Fc/Fc+, which corresponds to Ep,CO2 in the presence of 2.0 mM [C10mim]+ under voltammetric conditions with a scan rate of 0.1 V s-1. This potential was selected to avoid any significant contribution from the direct reduction of CO2 at a Ag electrode, which could be particularly significant in the presence of 1.0% H2O. After electrolysis, the gaseous products (CO and H2) were detected by gas chromatography and the liquid phase products (formate and oxalate) were concentrated before detection by NMR spectroscopy. The products obtained under different conditions and their faradaic efficiencies (FE) are summarized in Table 3. It should be noted that to calculate the FE, the charge used to reduce the co-catalyst, [C10mim]+ was ignored since its magnitude insignificant relative to the total charge consumed. Under all conditions examined, reduction of CO2 forms CO and C2O42- as the major products, with only a small amount of formate (FE < 1.0%). In the absence of either ionic liquid or added water, an Eappied value of -2.6 V was used instead of -2.3 V to obtain adequate current density. The major products under this condition are CO and C2O42- with FE values of 53.9% and 39.3%, respectively. The FE for H2 is negligible due to the aprotic environment. Upon addition of 2.0 mM ionic liquid, CO and C2O42- remain the major products although their FE values do depend on the identity of ionic liquid with a higher FE value for CO detected in the presence of [C4mpyrd][BF4].

26 ACS Paragon Plus Environment

Page 27 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 3 Products formed by bulk electrolysis of CO2 at a Ag cathode in MeCN (0.1 M nBu4NPF6). FE/% Eappied/V Charge Additives vs. Fc/Fc+ /C H2 CO C2O42- HCO2without ILa -2.6b 12.5 0.2 53.9 39.3 0.8 a 2.0 mM [C10mim][BF4] -2.3 28.5 0 45.1 53.4 0.4 a 2.0 mM [C4mpyrd][BF4] -2.3 50.1 0 66.4 28.4 0.8 without IL + 1.0% H2O -2.3 15.3 87.4 9.5 3.5 ---2.3 61.8 1.8 70.4 27.1 0.4 2.0 mM [C10mim][BF4] + 1.0% H2O 2.0 mM [C4mpyrd][BF4] + 1.0% H2O -2.3 50.2 0.4 58.4 42.6 0.8 a Water content: 5.0 mM. b This more negative potential was chosen to achieve a reduction rate that is comparable to that found in the presence of IL additive. With added water, but in the absence of ionic liquid, H2 is the main product (FE = 87.4%). The hydrogen evolution reaction (HER) is almost completely suppressed by addition of ionic liquids. Since HER is an inner-sphere process, the presence of any surface active species is expected to hinder its formation. In the presence of water and imidazolium-based ionic liquid, the FE for CO formation is also significantly enhanced. CO formation can occur regardless of proton availability via either homogenous or heterogeneous pathways. In the homogeneous pathway and in the absence of protons, CO2 is first reduced to CO2.- at the electrode surface. If this intermediate becomes detached, it can then react with dissolved CO2 to form CO via a multi-step pathway (Scheme 523). If CO2.remains attached, it can dimerise to form oxalate (Scheme 623). In the presence of H+, a proton coupled heterogeneous reaction occurs to form CO (Scheme 71, 23). In the presence of added water and ionic liquids, clearly, Scheme 6 is dominant.

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 39

Scheme 5. Homogeneous pathway for CO formation

Scheme 6. The dimerization of the carbon dioxide radical anion in the absence of H+

Scheme 7. Heterogeneous pathway for CO formation in the presence of H+. To investigate the stability of ionic liquids under bulk electrolysis conditions, the amount of CO formed and FE were measured as a function of time (or overall charge consumed) with Eapplied -2.3 V and in the presence of 2.0 mM ionic liquid (Figure 9). With 2.0 mM [C10mim][BF4] and no added H2O, the current initially remains relatively constant and the amount of CO formed is approximately proportional to the charge passed up to 20 C. However, when the electrolysis proceeds further, both the current passed and FE for CO production fall rapidly, most probably because [C10mim]+ is consumed during bulk electrolysis. Once 28.5 C of charge had been passed, the catalytic current is essentially zero. However, catalytic activity of imidazolium is recovered by addition of 1.0% water. In contrast, in the presence of [C4mpyrd]+, the FE for CO formation is almost independent of electrolysis time, with the small diminution in the reduction current at long times probably due to the consumption of CO2.

28 ACS Paragon Plus Environment

Page 29 of 39

2.0 mM [C10mim][BF4] 2.0 mM [C4mpyrd][BF4] 2.0 mM [C10mim][BF4]+ 1.0% H2O 2.0 mM [C4mpyrd][BF4]+ 1.0% H2O

n(CO) (µmol)

200

100

0 100

FE (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0

20

0

20

40

60

40

60

50

0

Charge(C)

Figure 9. The quantity of CO formed and FE for CO formation at a Ag wire cathode during bulk electrolysis of CO2 in the presence of 2.0 mM [C10mim][BF4] or [C4mpyrd][BF4] in MeCN (0.1 M n-Bu4NPF6). The addition of 1.0 % water to 2.0 mM [C10mim][BF4] solutions significantly increases the FE for CO formation. Moreover, the current now remains almost constant when up to 60 C of charge is passed. Clearly, the presence of water significantly extends the catalytic lifetime of imidazolium for CO2 reduction, indicating that a catalytic active form of the imidazolium cation is regenerated when water is present and that this is important in the catalytic reduction of CO2. By contrast, the addition of 1.0 % water to 2.0 mM solutions of [C4mpyrd][BF4] does not increase the FE for CO formation or the stability of [C4mpyrd]+ to any appreciable extent. The stability of the ionic liquid cations was assessed by examining the solutions before and after electrolysis by cyclic voltammetry. As seen in Figure S6a, in the absence of deliberately added water, the catalytic reduction process associated with [C10mim]+ has 29 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 39

essentially disappeared after bulk electrolysis, even though the unreacted CO2 is still ~ 79% of the original solution concentration, as judged from the magnitude of the mass-transport limited current for CO2 reduction. By contrast, in 2.0 mM [C4mpyrd][BF4] solutions (Figure S6b) the onset potential for reduction remains unchanged after electrolysis and the CO2 reduction peak current decreases by about 32 % due to the consumption of CO2. In the presence of 1.0 % added water, the onset potential and voltametric features were almost the same before and after bulk electrolysis when the charges listed in Table 3 were passed (Figures S6c, d). The magnitude of the peak reduction current after electrolysis decreases because the concentration of dissolved CO2 decreases. These observations again suggest that in the absence of water, imidazolium cations are consumed during bulk electrolysis, while the presence of water extends the catalytic lifetime for CO2 reduction. By contrast, the stability of [C4mpyrd]+ for CO2 reduction is unaffected by the presence of water.

3.3 NMR investigation of the deactivation mechanism for imidazolium cations To better understand the deactivation mechanism associated with the imidazolium cation under bulk electrolysis conditions, 1H NMR spectra were used to identify the products generated. The 1H NMR spectrum for a control solution containing 12.5 mM [C10mim][BF4] in MeCN is shown in Figure 10a. On the basis of the 1H proton chemical shifts reported in the literatures10,

25

, H1, H2 and H3 resonances are assigned to the hydrogens from the

imidazolium ring, and H4 and H5 resonances are the hydrogens from the alkyl groups.

30 ACS Paragon Plus Environment

Page 31 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

H4 C 9H 19

H4

H1

H5

N

N

H2

H3

H5 H5

H5 12.5 mM [C10mim][BF4]

H4

H2 H3

H1 a

After bulk electrolysis of 28.5 C H2' H3'

H2 H3 H1

HCO2-

b

H5' H4'

H5 H4

After bulk electrolysis of 28.5 C + 1.0% H2O H4 H1

H2 HCO2-

c 10

9

8

H3 H2' H3' 7

H5

H4' 6

5

H5'

4

f1 (ppm)

Figure 10. 1H NMR spectra of (a) 12.5 mM [C10mim][BF4] in MeCN, (b) a concentrated solution of 2.0 mM [C10mim][BF4] in MeCN (0.1 M n-Bu4NPF6) after electrolysis (28.5 C) under a CO2 atmosphere and (c) as for (b) but with 1.0% added water. After reductive electrolysis (28.5 C) of a 2.0 mM solution of [C10mim][BF4] in MeCN (0.1 M n-Bu4NPF6) new resonances were detected in the spectrum, labelled H'2, H'3, H'4 and H'5, while the resonance for H1 had shifted from 8.97 to 9.35 ppm (Figure 10b) due to the solution becoming more basic during the course of the bulk electrolysis experiment. The heights of all the resonances related to the imidazolium cation significantly decreased. Similar new resonances have been reported after the electrochemical reduction of CO2 in the presence of 100 mM [C2mim][TFSI] in MeCN10. This study identified the formation of a carboxylate complex formed between an imidazolium carbene and CO2. This species is also well documented as a product arising from a complexation reaction between an imidazolium 31 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

carbene and CO2

Page 32 of 39

10, 22, 38

. Thus, it is concluded that the C2-H of the imidazolium cation is

consumed during bulk electrolysis to form carbene, which reacts with CO2 to generate the carboxylate complex that gives rise to the H'2, H'3, H'4 and H'5 resonances, derived from H2, H3, H4 and H5, respectively. It is also notable in Figure 10b, that the imidazolium dimer is not detected. The bulky alkyl group on the imidazolium cation would hinder dimerization. Apparently, imidazolium dimers do not act as co-catalysts for CO2 reduction under our bulk electrolysis conditions. The integrated 1H NMR resonances (a) and (b) in Figure 10 show that the majority of the [C10mim]+ cations were converted to the carboxylate complex during bulk electrolysis. Upon addition of 1.0% water (Figure 10c), the integrated resonance for the 1H in imidazolium increases, as expected if [C10mim]+ is regenerated by protonation of the carbene in the carboxylate complex. This result is consistent with the fact that the catalytic activity of the deactivated [C10mim]+ is recovered upon addition of water. On the basis of the NMR data, it is concluded that the imidazolium carboxylate formed between carbene and CO2 does not catalyse reduction of CO2, and results a deactivated form of the imidazolium cations. These conclusions are consistent with the experimental observations of Sun et. al.10 and Watkins et. al.19 The former authors suggested that CO and imidazolium carboxylate are major products in the electrochemical reduction of CO2 at a Pb cathode in MeCN solution containing 0.1 M [C2mim][TFSI] as the supporting electrolyte.10 The latter authors reported the electrochemical reduction of 1,3-dimethylimidazolium-2-carboxylate ([C1mim-CO2]) in dry 1-ethyl-3-methylimidazolium trifluoroacetate.19 No catalytic activity for CO2 reduction was found under either argon or CO2 atmospheres. These results are also consistent with the fact that imidazolium carboxylates are formed by a complexation reaction between carbene and CO2, and that they are the precursors for generating of N-heterocyclic carbene catalysts in relevant area of organic syntheses39. 32 ACS Paragon Plus Environment

Page 33 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

In summary, a combination of data obtained by voltammetry and bulk electrolyses of solutions containing imidazolium ionic liquids with 1H NMR detection of products allows a mechanism for catalytic reduction of CO2 to be proposed (Scheme 8) using CO formation as an example. In this scheme, all imidazolium related species are assumed to be adsorbed on the silver electrode surface since clear evidence (Figures 1 and 2) suggests that the electrode material also plays a decisive role in the catalytic reduction of CO2. This scheme is analogous to that proposed by Wang et. al. on the basis of theoretical calculations29. According to this mechanism, an imidazolium cation is reduced to generate a neutral radical. This radical then forms a complex with CO2 that is more easily reduced at the Ag electrode leading to formation of CO through a multi-step reaction pathway. This catalytic mechanism also implies a source of proton is essential for imidazolium to act as a catalyst. Reduction of imidazolium also can form carbene, which then forms a catalytically inactive complex with CO2, which deactivates the catalytic action of imidazolium on CO2 reduction. This competing deactivation process is slowed down when proton availability increases. Furthermore, the imidazolium carboxylate can be reactivated, if formed, to regenerate the imidazolium cocatalyst if a proton source, such as water, is present, as suggested by the NMR data. This reaction with water explains why imidazolium ionic liquids are excellent catalysts for the electrochemical reduction of CO2 in aqueous media.12, 15, 20

33 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 39

Scheme 8. Mechanism associated with [C10mim]+ catalysis of the electrochemical reduction of CO2 to CO along with its deactivation pathway.

Conclusion The electrochemical reduction of CO2 at a Ag electrode has been investigated in MeCN (0.1 M n-Bu4NPF6) in the presence of 13 ionic liquids or salts that contain imidazolium, pyrrolidinium, ammonium, phosphonium, borate, pyridinium and sulfonium-based cations, and 7 ammonium cations of variable alkyl chain lengths. The results show many cations can significantly catalyse the electrochemical reduction CO2 along with a Ag electrode. Thus, catalytic CO2 reduction is not restricted to the widely reported imidazolium ionic liquids and is a property associated with many other cations, although the mechanism for catalysis varies with the nature of the cation. Imidazolium cations act as co-catalysts at a Ag electrode, and after reduction form a complex between the electrogenerated imidazolium radical and CO2. The imidazolium radical is the key species for catalysing CO2 reduction rather than the imidazolium carboxylate complex formed by a reaction between an imidazolium carbene and CO2. NMR results suggest that the formation of imidazolium carboxylate leads to deactivation of the co-catalyst, but addition of water leads to catalysis recovery. Other catalytically active cations are believed to enhance CO2 reduction by an electrochemical double layer effect. In terms of the stability for long-term applications, pyrrolidinium-based ionic liquids are recommended for catalytic reduction of CO2 in the absence of water, while both imidazolium and pyrrolidinium are both very effective in the presence of water.

Associated Content Supporting Information

34 ACS Paragon Plus Environment

Page 35 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Supporting data include voltammograms of ionic liquids and salts under N2 and CO2 atmospheres (Figures S1-S6). This material is available free of charge via the Internet at http://pubs.acs.org

Author Information Corresponding Authors *Email: [email protected] (J.Z.) *Email: [email protected] (M.H.) Notes The authors declare no competing financial interest.

Acknowledgements SFZ gratefully acknowledges the CSIRO Office of the Chief Executive for provision of a postdoctoral research fellowship. JZ and AMB would like to thank the Australian Research Council for financial support. The provision of the ionic liquids [NHHB][TFSI] and [C6DABCO][TFSI] by Dr. Thomas Ruether is also much appreciated.

35 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 39

References (1) Hori, Y., Electrochemical CO2 Reduction on Metal Electrodes. In Modern Aspects of Electrochemistry, Vayenas, C. G.; White, R. E.; Gamboa-Aldeco, M. E., Eds. Springer New York: New York, NY, 2008; pp 89-189. (2) Appel, A. M.; Bercaw, J. E.; Bocarsly, A. B.; Dobbek, H.; DuBois, D. L.; Dupuis, M.; Ferry, J. G.; Fujita, E.; Hille, R.; Kenis, P. J. A., et al., Frontiers, Opportunities, and Challenges in Biochemical and Chemical Catalysis of CO2 Fixation. Chem. Rev. 2013, 113, 6621-6658. (3) Sharma, P. P.; Wu, J.; Yadav, R. M.; Liu, M.; Wright, C. J.; Tiwary, C. S.; Yakobson, B. I.; Lou, J.; Ajayan, P. M.; Zhou, X. D., Nitrogen-Doped Carbon Nanotube Arrays for HighEfficiency Electrochemical Reduction of CO2 : On the Understanding of Defects, Defect Density, and Selectivity. Angew. Chem. Int. Ed. 2015, 54, 13701-13705. (4) Sreekanth, N.; Nazrulla, M. A.; Vineesh, T. V.; Sailaja, K.; Phani, K. L., Metal-Free Boron-Doped Graphene for Selective Electroreduction of Carbon Dioxide to Formic acid/Formate. Chem. Commun. 2015, 51, 16061-16064. (5) Mikkelsen, M.; Jørgensen, M.; Krebs, F. C., The Teraton Challenge. A Review of Fixation and Transformation of Carbon Dioxide. Energy Environ. Sci. 2010, 3, 43-81. (6) Benson, E. E.; Kubiak, C. P.; Sathrum, A. J.; Smieja, J. M., Electrocatalytic and Homogeneous Approaches to Conversion of CO2 to Liquid Fuels. Chem. Soc. Rev. 2009, 38, 89-99. (7) Alvarez-Guerra, M.; Albo, J.; Alvarez-Guerra, E.; Irabien, A., Ionic Liquids in the Electrochemical Valorisation of CO2. Energy Environ. Sci. 2015, 8, 2574-2599. (8) MacFarlane, D. R.; Forsyth, M.; Howlett, P. C.; Kar, M.; Passerini, S.; Pringle, J. M.; Ohno, H.; Watanabe, M.; Yan, F.; Zheng, W., et al., Ionic Liquids and Their Solid-State Analogues as Materials for Energy Generation and Storage. Nat. Rev. Mater. 2016, 1-15. (9) Medina-Ramos, J.; Pupillo, R. C.; Keane, T. P.; DiMeglio, J. L.; Rosenthal, J., Efficient Conversion of CO2 to CO Using Tin and Other Inexpensive and Easily Prepared PostTransition Metal Catalysts. J. Am. Chem. Soc. 2015, 137, 5021-5027. (10) Sun, L.; Ramesha, G. K.; Kamat, P. V.; Brennecke, J. F., Switching the Reaction Course of Electrochemical CO2 Reduction with Ionic Liquids. Langmuir 2014, 30, 6302-6308. (11) Medina-Ramos, J.; DiMeglio, J. L.; Rosenthal, J., Efficient Reduction of CO2 to CO with High Current Density Using in Situ or Ex Situ Prepared Bi-Based Materials. J. Am. Chem. Soc. 2014, 136, 8361-8367. (12) Asadi, M.; Kumar, B.; Behranginia, A.; Rosen, B. A.; Baskin, A.; Repnin, N.; Pisasale, D.; Phillips, P.; Zhu, W.; Haasch, R. et al., Robust Carbon Dioxide Reduction on Molybdenum Disulphide Edges. Nat. Commun. 2014, 5, 4470. (13) Salehi-Khojin, A.; Jhong, H.-R. M.; Rosen, B. A.; Zhu, W.; Ma, S.; Kenis, P. J.; Masel, R. I., Nanoparticle Silver Catalysts That Show Enhanced Activity for Carbon Dioxide Electrolysis. J. Phys. Chem. C 2013, 117, 1627-1632. (14) Rosen, B. A.; Zhu, W.; Kaul, G.; Salehi-Khojin, A.; Masel, R. I., Water Enhancement of CO2 Conversion on Silver in 1-Ethyl-3-Methylimidazolium Tetrafluoroborate. J. Electrochem. Soc. 2013, 160, H138-H141. (15) Kumar, B.; Asadi, M.; Pisasale, D.; Sinha-Ray, S.; Rosen, B. A.; Haasch, R.; Abiade, J.; Yarin, A. L.; Salehi-Khojin, A., Renewable and Metal-Free Carbon Nanofibre Catalysts for Carbon Dioxide Reduction. Nat. Commun. 2013, 4, 2819. (16) DiMeglio, J. L.; Rosenthal, J., Selective Conversion of CO2 to CO with High Efficiency Using an Inexpensive Bismuth-Based Electrocatalyst. J. Am. Chem. Soc. 2013, 135, 87988801. 36 ACS Paragon Plus Environment

Page 37 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(17) Rosen, B. A.; Haan, J. L.; Mukherjee, P.; Braunschweig, B.; Zhu, W.; Salehi-Khojin, A.; Dlott, D. D.; Masel, R. I., In Situ Spectroscopic Examination of a Low Overpotential Pathway for Carbon Dioxide Conversion to Carbon Monoxide. J. Phys. Chem. C 2012, 116, 15307-15312. (18) Bocarsly, A. B.; Gibson, Q. D.; Morris, A. J.; L’Esperance, R. P.; Detweiler, Z. M.; Lakkaraju, P. S.; Zeitler, E. L.; Shaw, T. W., Comparative Study of Imidazole and Pyridine Catalyzed Reduction of Carbon Dioxide at Illuminated Iron Pyrite Electrodes. ACS Catal. 2012, 2, 1684-1692. (19) Watkins, J. D.; Bocarsly, A. B., Direct Reduction of Carbon Dioxide to Formate in High-Gas-Capacity Ionic Liquids at Post-Transition-Metal Electrodes. ChemSusChem 2014, 7, 284-290. (20) Rosen, B. A.; Salehi-Khojin, A.; Thorson, M. R.; Zhu, W.; Whipple, D. T.; Kenis, P. J. A.; Masel, R. I., Ionic Liquid–Mediated Selective Conversion of CO2 to CO at Low Overpotentials. Science 2011, 334, 643-644. (21) Matsubara, Y.; Grills, D. C.; Kuwahara, Y., Thermodynamic Aspects of Electrocatalytic CO2 Reduction in Acetonitrile and with an Ionic Liquid as Solvent or Electrolyte. ACS Catal. 2015, 5, 6440-6452. (22) Zhao, S.-F.; Wu, L.-X.; Wang, H.; Lu, J.-X.; Bond, A. M.; Zhang, J., A Unique Proton Coupled Electron Transfer Pathway for Electrochemical Reduction of Acetophenone in the Ionic Liquid [Bmim][BF4] under a Carbon Dioxide Atmosphere. Green Chem. 2011, 13, 3461-3468. (23) Gennaro, A.; Isse, A. A.; Severin, M.-G.; Vianello, E.; Bhugun, I.; Saveant, J.-M., Mechanism of the Electrochemical Reduction of Carbon Dioxide at Inert Electrodes in Media of Low Proton Availability. Faraday Trans. 1996, 92, 3963-3968. (24) Tomita, Y.; Teruya, S.; Koga, O.; Hori, Y., Electrochemical Reduction of Carbon Dioxide at a Platinum Electrode in Acetonitrile‐Water Mixtures. J. Electrochem. Soc. 2000, 147, 4164-4167. (25) Gorodetsky, B.; Ramnial, T.; Branda, N. R.; Clyburne, J. A., Electrochemical Reduction of an Imidazolium Cation: A Convenient Preparation of Imidazol-2-Ylidenes and Their Observation in an Ionic Liquid. Chem. Commun. 2004, 1972-1973. (26) Feroci, M.; Chiarotto, I.; Forte, G.; Vecchio Ciprioti, S.; Inesi, A., Stability and CO2 Capture Ability of Electrogenerated N-Heterocyclic Carbene in Parent 1-Butyl-3methylimidazoliun Ionic Liquid (BMIm-X): The Role of X−. ChemElectroChem 2014, 1, 1407-1414. (27) Forte, G.; Chiarotto, I.; Inesi, A.; Loreto, M. A.; Feroci, M., Electrogenerated NHeterocyclic Carbene in Ionic Liquid: An Insight into the Mechanism of the Oxidative Esterification of Aromatic Aldehydes. Adv. Synth. Catal. 2014, 356, 1773-1781. (28) Xiao, L.; Johnson, K. E., Electrochemistry of 1-Butyl-3-methyl-1H-imidazolium Tetrafluoroborate Ionic Liquid. J. Electrochem. Soc. 2003, 150, E307. (29) Wang, Y.; Hatakeyama, M.; Ogata, K.; Wakabayashi, M.; Jin, F.; Nakamura, S., Activation of CO2 by Ionic Liquid EMIM-BF4 in the Electrochemical System: a Theoretical Study. Phys. Chem. Chem. Phys. 2015, 17, 23521-23531. (30) Oh, Y.; Hu, X., Ionic Liquids Enhance the Electrochemical CO2 Reduction Catalyzed by MoO2. Chem. Commun. 2015, 51, 13698-13701. (31) Hanc-Scherer, F. A.; Montiel, M. A.; Montiel, V.; Herrero, E.; Sanchez-Sanchez, C. M., Surface Structured Platinum Electrodes for the Electrochemical Reduction of Carbon Dioxide in Imidazolium Based Ionic Liquids. Phys. Chem. Chem. Phys. 2015, 17, 23909-23916. (32) Lewandowski, A.; Świderska-Mocek, A., Ionic Liquids as Electrolytes for Li-Ion Batteries—an Overview of Electrochemical Studies. J. Power Sources 2009, 194, 601-609.

37 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 39

(33) Zhao, S.-F.; Horne, M.; Bond, A. M.; Zhang, J., Electrocarboxylation of Acetophenone in Ionic Liquids: The Influence of Proton Availability on Product Distribution. Green Chem. 2014, 16, 2242-2251. (34) Corrigan, D. A.; Evans, D. H., Cyclic Voltammetric Study of Tert-Nitrobutane Reduction in Acetonitrile at Mercury and Platinum Electrodes: Observation of a Potential Dependent Electrochemical Transfer Coefficient and the Influence of the Electrolyte Cation on the Rate Constant. J. Electroanal. Chem. 1980, 106, 287-304. (35) Evans, D. H., One-Electron and Two-Electron Transfers in Electrochemistry and Homogeneous Solution Reactions. Chem. Rev. 2008, 108, 2113-2144. (36) Izutsu, K., Selection and Preparation of Supporting Electrolytes. In Electrochemistry in Nonaqueous Solutions, Wiley-VCH Verlag GmbH & Co. KGaA: 2003; pp 301-311. (37) Husson, P.; Pison, L.; Jacquemin, J.; Gomes, M. F. C., Influence of Water on the Carbon Dioxide Absorption by 1-Ethyl-3-Methylimidazolium Bis(trifluoromethylsulfonyl)amide. Fluid Phase Equilib. 2010, 294, 98-104. (38) de Robillard, G.; Devillers, C. H.; Kunz, D.; Cattey, H.; Digard, E.; Andrieu, J., Electrosynthesis of Imidazolium Carboxylates. Org. Lett. 2013, 15, 4410-4413. (39) Duong, H. A.; Tekavec, T. N.; Arif, A. M.; Louie, J., Reversible Carboxylation of NHeterocyclic Carbenes. Chem. Commun. 2004, 112-113.

38 ACS Paragon Plus Environment

Page 39 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table of Content

39 ACS Paragon Plus Environment