Nickel-Gallium-Catalyzed Electrochemical Reduction of CO2 To Highly Reduced Products at Low Overpotentials

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Subscriber access provided by NEW YORK UNIV

Letter
Nickel-Gallium-Catalyzed Electrochemical Reduction of
CO2 to Highly Reduced Products at Low Overpotentials
Daniel A. Torelli, Sonja A Francis, J. Chance Crompton, Alnald Javier, Jonathan
R Thompson, Bruce S. Brunschwig, Manuel P. Soriaga, and Nathan S. Lewis
ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.5b02888 • Publication Date (Web): 17 Feb 2016
Downloaded from http://pubs.acs.org on February 21, 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.

ACS Catalysis 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 16 ACS Catalysis

1
2
3
4 Nickel-Gallium-Catalyzed Electrochemical Reduction of CO2 to
5
6
Highly Reduced Products at Low Overpotentials
7
8 Daniel A. Torelli¶†‡, Sonja A. Francis¶†‡£, J. Chance Crompton†‡, Alnald Javier‡, Jonathan R.
9 Thompsonǁ, Bruce S. Brunschwig*§, Manuel P. Soriaga*‡, and Nathan S. Lewis*†‡§
10
11 †Division of Chemistry and Chemical Engineering, ‡Joint Center for Artificial Photosynthesis,
£Resnick Sustainability Institute, ǁ Division of Engineering and Applied Sciences, and §Beckman
12
13 Institute Molecular Materials Research Center, California Institute of Technology, Pasadena,
14 California 91125, United States
15
16 ¶These authors contributed equally
17
18
19
Abstract: We report the electrocatalytic reduction of CO2 to the highly reduced C2 products, ethylene and ethane, as
20
21
well as to the fully reduced C1 product methane, on three different phases of nickel-gallium (NiGa, Ni3Ga, and
22
23 Ni5Ga3) films prepared by drop-casting. In aqueous bicarbonate electrolytes at neutral pH, the onset potential for
24
25 methane, ethylene, and ethane production on all three phases was found to be -0.48 V vs. the reversible hydrogen
26
27 electrode (RHE), among the lowest onset potentials reported to date for the production of C2 products from CO2.
28
29 Similar product distributions and onset potentials were observed for all three nickel-gallium stoichiometries tested.
30
31 The onset potential for the reduction of CO2 to C2 products at low current densities catalyzed by nickel-gallium was
32
33 > 250 mV more positive than polycrystalline copper, and approximately equal to single crystals of copper, which
34
35 have some of the lowest overpotentials to date for the reduction of CO2 to C2 products and methane. The nickel-
36
37 gallium films also reduced CO to ethylene, ethane, and methane, consistent with a CO2 reduction mechanism that
38
39 first involves the reduction of CO2 to CO. Isotopic labeling experiments with 13
CO2 confirmed that the detected
40
41 products were produced exclusively by the reduction of CO2.
42
43
44
45 Keywords: CO2 reduction, electrocatalysis, nickel gallium, NiGa, ethane, methane, low overpotential, C2
46
47 production
48
49
50 Introduction
51
52
53 Pioneering work in the 1980s showed that Cu is the only pure metal capable of catalyzing the electrochemical
54
55 reduction of CO2 to highly reduced products (any product requiring the transfer of > 6 electrons, with needed
56
57 protons, to one molecule of CO2) in an aqueous solution at near-neutral pH in high yields.1-3 Under 1 atm of CO2,
58
59
60
1
ACS Paragon Plus Environment
ACS Catalysis Page 2 of 16

1
2
3 polycrystalline Cu produces a mixture of gaseous and liquid hydrocarbons (CH4, C2H4, C2H5OH, and C3H7OH) at
4
5 >70% Faradaic efficiency. However, potentials more negative than -1 V vs. the reversible hydrogen electrode
6
7 (RHE) are required to reduce CO2 to these products at cathodic current densities > 1 mA cm-2. Detailed
8
9 investigations indicate that as many as 14 different highly reduced products, at a variety of yields, can be produced
10
11 by Cu under aqueous, neutral pH conditions.4
12
13
14 The electrocatalytic activity of Cu has inspired efforts to improve the selectivity, and to lower the overpotentials
15
16 required to produce these more reduced products.5-13 Although these efforts have resulted in an improvement in
17
18 selectivity for both CH48 and various C2 products,5,7,10-12 overpotentials of > 500 mV are required to reach current
19
20 densities > 1 mA cm-2 towards these products. The discovery of new families of electrocatalysts that do not contain
21
copper, but produce highly reduced products from CO2 could be advantageous because CO2 reduction at non-copper
22
23
surfaces might proceed via lower-energy pathways than those accessible on copper-containing surfaces. Therefore,
24
25
such a discovery would offer not only alternatives to copper for the electrocatalytic reduction of CO2, but also the
26
27
possibility that different constraints than those which govern the highly studied copper surface would govern the
28
29
performance of the new catalysts. Such a system could provide a route to low-overpotential, tunable, and selective
30
31 catalysts. Alloys and intermetallics are an attractive set of materials for this purpose as they not only provide
32
33 multiple binding sites for reaction intermediates, but also provide materials with variable composition that may
34
35 exhibit altered reactivity relative to their bulk counterparts.
36
37
38 Theoretical14 and experimental13,15-24 investigations have recently focused on the development of new or improved
39
40 electrocatalysts for CO2 (or CO) reduction based on the use of metal alloys. Much of the experimental work has
41
42 focused on improving the selectivity and/or overpotential of Cu by forming alloys of Cu with other metals.15,17-24
43
44 Non-copper-containing alloys such as AuPd16 or PdPt25 have been used to study the effects of composition on
45
46 reactivity. Other non-copper-containing catalysts such as the III-V semiconductors GaP, GaAs, and InP have been
47
48 reported to produce CH3OH,26-28 although the products, yields, and reproducibility of these reports have not been
49
50 well established.29,30 Modified carbon supports such as metal-doped nitrogenated carbon,31 or nitrogen-doped
51
52 nanodiamond32, as well as modified gold electrodes with self-assembled monolayers33 have also been shown to
53
54 provide copper-free materials that can produce highly reduced products at a range of efficiencies and overpotentials.
55
56
57
58
59
60
2
ACS Paragon Plus Environment
Page 3 of 16 ACS Catalysis

1
2
3 Ni-based alloys and intermetallics are a particularly promising route to explore alternatives to Cu-based materials,
4
5 because Ni is the only other single metal reported to reduce CO2 to C2 products at potentials more negative than -1
6
7 V vs. RHE, although Ag has also been shown to make small amounts of CH3CH2OH at -1.35 V vs. RHE.3,34
8
9 Furthermore, Ni has been shown to have the second highest yields of CH4, albeit only at 0.5% Faradaic efficiency at
10
11 highly reducing potentials, i.e. more negative than -1 V vs RHE.34 The synthesis of Ni-based intermetallics or alloys
12
13 should offer the opportunity to modify the reactivity of Ni, possibly improving the activity of Ni for CO2 reduction.
14
15
16 The work described herein follows, and was inspired by, theoretical and experimental literature on CO2
17
18 hydrogenation. Specifically, SiO2-supported nanoparticles of Ni5Ga3, Ni3Ga, and NiGa are highly active for
19
20 thermochemical, gas-phase CO2 hydrogenation to CH3OH at temperatures above ambient.35 These materials are
21
stable intermetallic compounds with large ordering energies, and have calculated oxygen-adsorption energies close
22
23
to the optimum benchmark methanol-producing catalyst (Cu/ZnO/Al2O3). A volcano curve has been developed to
24
25
describe the theoretically predicted activity for some metals, including Ni and Ni intermetallics, for the
26
27
thermochemical reduction of CO2. The volcano curve indicated that the oxygen-adsorption energy is a relevant
28
29
parameter for predicting the catalytic activity. Experimental data were in accord with this suggestion, and the
30
31 highest yields of ~ 0.25 mol h-1 were obtained for the production of CH3OH from the Ni5Ga3 phase. Based on these
32
33 findings and conclusions, we have evaluated these NixGay materials for their electrocatalytic CO2 reduction activity.
34
35
36 Experimental Section
37
38 NiGa, Ni5Ga3, and Ni3Ga films were synthesized by a temperature-programmed reduction method adapted from
39
40 Studt et. al.35 (See Supporting Information, SI). After annealing, scanning-electron microscopy (SEM) showed that
41
42 the synthesized films consisted of aggregates of particles in the size range of 1-5 µm (Figure S1). X-ray diffraction
43
44 (XRD) indicated that the materials were crystalline and of the desired phase (Figure 1), with all peaks assignable
45
46 and consistent with the proposed composition (Table S1). X-ray photoelectron spectroscopy (XPS) showed that the
47
48 films had Ni:Ga ratios of approximately 1:1, 3:1, and 5:3 for NiGa, Ni3Ga, and Ni5Ga3 respectively (Figure S2 and
49
50 Table S2), with some surface Ni and Ga oxides resulting from exposure of the films to air during transfer to the
51
52 ultra-high vacuum instrumentation.36
53
54 The electrocatalytic activity of the films for CO2 reduction was evaluated in a modified two-compartment
55
56 electrochemical cell, Figure S3. The cell was sealed to facilitate bulk electrolysis, and gas chromatography (GC)
57
58
59
60
3
ACS Paragon Plus Environment
ACS Catalysis Page 4 of 16

1
2
3 and nuclear magnetic resonance (NMR) spectroscopy were used to analyze the gaseous and liquid products,
4
5 respectively, (see SI). Electrochemical surface area measurements were carried out similar to literature reports, and
6
7 all electrodes were found to be only slightly roughened, Figure S4.37
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 1. XRD data for the 3 different phases of nickel-gallium. The asterisk denotes signals from the quartz
27 substrate.
28
29 Results and Discussion
30
31 Analysis of the electrolysis products using the synthesized NixGay phases as the catalysts indicated the formation
32
33 of highly reduced products such as CH4, C2H4, C2H6, representing reductions that require the transfer of 8, 12, or 14
34
35 electrons per molecule of product, respectively. Figure 2 and Figure S5 show the product distribution and rate of
36
37 production of hydrocarbon products as a function of the electrode potential. The partial cathodic current density for
38
39 CH4 production exceeded 140 µA cm-2 at -1.18 V vs. RHE, with Faradaic efficiencies > 2 % at potentials more
40
41 negative than -0.48 V vs. RHE for NixGay films of varying stoichiometries. Partial cathodic current densities for the
42
43 production of C2H6 reached 100 µA cm-2 at -1.18 V vs. RHE, while Faradaic efficiencies were 1.3% at -0.48 V for
44
45 C2H6. The sum of the partial cathodic current densities towards all three hydrocarbons reached > 200 µA cm-2
46
47 at -1.18 V. The balance of the current yielded H2 along with trace amounts of CO and CH3OH.
48
49
50
51
52
53
54
55
56
57
58
59
60
4
ACS Paragon Plus Environment
Page 5 of 16 ACS Catalysis

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 Figure 2. Potential-dependent Faradaic efficiencies (solid lines) and current densities (dotted line) for CO2 reduction
36 in CO2-saturated 0.1M NaHCO3 (aq) to methane (triangles), ethane (exes) and ethylene (squares). (a) Hydrocarbon
37 production for the Ni5Ga3 phase of NixGay and (b) for Cu where data were taken from ref.4 in 0.1 M KHCO3.
38
39 No electrocatalysts, other than Cu and to some extent Ni, have been reported to produce gaseous C2 products in
40
41 any detectable yields from the electrochemical reduction of CO2 in an aqueous electrolyte solution.3,38,39 Other Cu-
42
43 free materials, such as organic self-assembled monolayers,33 or nonmetallic electrocatalysts,32 have produced liquid
44
45 C2 products at low overpotentials, while Ag40 and carbon nanotube confined metallic nanoparticles41 have been
46
47 shown to produce liquid C2 products at large overpotentials. The production of gaseous C2 products allows for
48
49 simpler product separation and collection. Moreover, the onset potential of -0.48 V vs. RHE for C2 and CH4
50
51 production by any of the NixGay phases studied is approximately equal to that of single crystal Cu(100), to date the
52
53 lowest overpotential catalyst for C2 production,42 while the onset potential for C2 and CH4 production are 250 mV
54
55 and 300 mV, respectively, more positive than that reported for polycrystalline Cu, Figure 2.4 The NixGay films
56
57 prepared in this work were not single crystals, nor were they optimized to maximize activity. Hence an appropriate
58
59
60
5
ACS Paragon Plus Environment
ACS Catalysis Page 6 of 16

1
2
3 comparison is with the behavior of polycrystalline copper catalysts, noting that opportunities remain for improving
4
5 the catalytic activity of NixGay. For example, single crystals of NiGa could possibly allow increases in Faradaic
6
7 efficiencies, partial cathodic current densities, and overpotentials, similar to improvements over polycrystalline
8
9 copper realized using Cu single crystals as electrocatalysts. This optimization could yield higher current densities,
10
11 making NixGay a potential candidate for CO2 reduction applications. At present, the partial current densities towards
12
13 highly reduced products at large overpotentials are substantially higher on Cu electrodes than on NixGay.
14
15 Polycrystalline copper and the nickel-gallium films evaluated in this work both produce CH4 and C2H4 from CO2.
16
17 Because the two materials produce common products, the reactions at the two surfaces likely possess some
18
19 mechanistic commonalities. Both theory43 and experiments42,44 have shown the reduction of CO2 to both CH4 and
20
21 C2H4 on Cu involves the reduction of CO2 to CO as the first step in the formation of products. In the electrocatalytic
22
23 reduction of CO2 by NixGay films, CO was detected at <0.01 % Faradaic efficiency by GC (Figure S6), suggesting
24
25 that CO may be an intermediate on the three different NixGay phases. Hence, electrolyses were performed in
26
27 phosphate buffer at pH = 7 under N2, CO2, or CO. Both CO2 and CO yielded CH4, C2H4, and C2H6 (Figure S7),
28
29 whereas no significant amounts of CO2 reduction products, and only H2 production, were observed under N2. These
30
31 observations suggest that CO is an intermediate in at least some pathways for conversion of CO2 into these highly
32
33 reduced products on both the NixGay and Cu surfaces, and that related reaction pathways for CO2 reduction may be
34
35 accessible even at surfaces of two different materials (NixGay and Cu).
36
37 Cyclic voltammograms (CVs) were obtained in aqueous solution by sweeping the potential from 0 V vs. the open-
38
39 circuit potential (OCP) to -0.68 V vs. RHE then back to +1.12 V vs. RHE, Figure 3. Under 1 atm of CO2, on the
40
41 return scan the nickel-gallium films, Figure 3a, showed a small irreversible anodic wave at +0.54 V vs. RHE. For
42
43 CVs performed on NixGay after CO2 reduction electrolysis, on the return scan the anodic wave, previously observed
44
45 at +0.54 V vs. RHE, was ~3x larger than observed on pristine samples. After a second cycle, the anodic wave
46
47 disappeared almost completely, Figure 3. To determine the nature of this wave, CVs were conducted under an
48
49 atmosphere of CO in phosphate buffer at pH 7. Under CO, a large wave was observed at +0.55 V, close to the
50
51 position of the peak under CO2 suggesting the wave corresponds to CO oxidation or another chemisorbed CO2 (or
52
53 CO) reduction intermediate. The more negative onset potential and lower current densities for catalysis under CO is
54
55 likely caused by chemisorbed CO binding to H2O or proton reduction sites, suppressing H2 evolution, and shifting
56
57 the onset to more negative electrode potentials. Upon further cycling, the peak at + 0.55 V almost completely
58
59
60
6
ACS Paragon Plus Environment
Page 7 of 16 ACS Catalysis

1
2
3 disappeared suggesting that the rebinding of CO is a relatively slow process and does not occur on the time scale of
4
5 the CV. When the electrode was immersed in solution for 3-5 min, and another CV was performed, the anodic wave
6
7 reappeared implying that CO eventually rebinds. This behavior contrasts with that of Ni under the same conditions
8
9 (Figure 3b). Upon cycling a Ni electrode under CO, synthesized using the same method as the NiGa but excluding
10
11 Ga, an anodic wave that did not disappear upon subsequent cycling was seen at +0.59 V (Figure 3b). This behavior
12
13 indicates that CO more rapidly rebinds on Ni compared to NiGa. Literature reports have found CO is tightly bound
14
15 on metallic Ni and its electrochemical oxidation is seen at similar potentials.45 The introduction of Ga into the Ni
16
17 films appears to destabilize the Ni-CO interaction, slowing the poisoning of the surface and greatly improving yields
18
19 of highly reduced products.
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 Figure 3. Cyclic voltammetry for (a) Ni5Ga3 films and (b) Ni in 0.1 M K2HPO4 buffered to pH = 6.8 with KH2PO4
54 under CO2 and CO at 50 mV s-1 scan rate. The anodic wave peak potentials (grey vertical lines) are shown in the
55 inset, of (a) Ni5Ga3 peak at +0.55 V and in (b) Ni peak at +0.59 V.
56
57
58
59
60
7
ACS Paragon Plus Environment
ACS Catalysis Page 8 of 16

1
2
3 Isotopic labeling experiments were performed to confirm that the carbon in the products was derived from CO2.
4
5 13
CO2 was used as the reactant gas, and gas chromatography mass/spectrometry (GCMS) rather than a thermal
6
7 conductivity detector (TCD) was used to analyze the headspace. In this experiment, 100 µL headspace aliquots were
8
9 taken every few hours and analyzed for isotopically labeled products. Figure S8 shows the resulting chromatograms
10
11 for several important mass fragments. Mass fragment m/z=17 corresponds to 13
CH4+, m/z=29 corresponds to
12
13 13
CH213CH+ fragments from both C2H4 and C2H6 that eluted at different times, and m/z=31 corresponds to 13C2H5+
14
15 from 13C2H6. Other mass fragments corresponding to 13CH4, 13C2H4, and 13C2H6 were observed, and the estimated
16
17 Faradaic efficiencies for the products were similar to when 12
CO2 was used. Additionally, over the >10 h
18
19 electrolysis, no decrease in the rate of 13CH4, 13C2H4, or 13C2H6 production was observed.
20
21
Several important differences are apparent between the reactivity of polycrystalline Cu and the NixGay films. The
22
23
potential at which production of CH4 and C2H4 begins on the polycrystalline NixGay films is several hundred mV
24
25
more positive than on polycrystalline Cu. In the potential range from -0.5 V to -0.85 V vs. RHE, polycrystalline Cu
26
27
largely makes only the two-electron reduction products, CO and HCOO-, whereas the NixGay films form the 8, 12,
28
29
and 14-electron reduction products, CH4, C2H4, and C2H6 respectively, in this potential range. Additionally, the
30
31 onset potentials are nearly equivalent to the best-reported catalysts to date, single crystals of Cu and oxide derived
32
33 Cu.6,12,46 The oxidation of chemisorbed CO is not observed on Cu surfaces, most likely due to its weaker interaction
34
35 with CO compared to that of Ni. This stronger interaction with CO, along with the possibility of tuning its strength
36
37 by modifying the composition of the intermetallic, make NixGay an interesting catalyst for electrochemical CO2
38
39 reduction.
40
41
42 When large overpotentials are applied (e.g. at -1.2 V vs RHE), many catalysts exhibit higher current densities
43
44 than NiGa for the production of hydrocarbons. For example, at these potentials, Cu exhibits cathodic current
45
46 densities of 6 mA/cm2 as compared to the 100 µA/cm2 cathodic current densities observed for the NixGay alloys
47
48 reported herein.4,11,12,46 However, the onset potential of Ni-Ga is -0.48 V for highly reduced products, with cathodic
49
50 current densities >100 µA/cm2 at -0.88 V vs. RHE. These materials improve on the state-of-the-art catalysts by
51
52 lowering the overpotential required to generate highly reduced products with a polycrystalline electrode. Although
53
54 current densities reported here are low, high surface area electrodes, or single crystal electrodes, could generate
55
56 significantly higher rates of hydrocarbon production. Additionally, these materials provide a key link between CO2
57
reduction electrocatalysts and theoretically predicted high temperature CO2 hydrogenation catalysts. This link could
58
59
60
8
ACS Paragon Plus Environment
Page 9 of 16 ACS Catalysis

1
2
3 be exploited in future CO2 reduction electrocatalysts discovery studies. Because neither Ni nor Ga alone exhibit low
4
5 overpotentials for CO2 reduction, this work strikingly shows that unique and tunable reactivity can be obtained for
6
7 CO2 reduction by use of metal alloys rather than their pure phases, and thus opens up new lines of investigation for
8
9 the discovery of CO2 reduction electrocatalysts.
10
11
12 Supporting Information Available:
13
14 Experimental procedures, catalyst characterization, detailed product analysis, and isotopic labeling experiments are
15
16 included in SI. This material is available free of charge via the Internet at http://pubs.acs.org.
17
18
19 Author Information
20
Corresponding Author
21
22
*Email: nslewis@caltech.edu
23
24
25
26 Acknowledgements
27
28
The authors would like to thank Prof. Matthew McDowell for his help with TEM and insightful discussions, Dr.
29
30
Ivonne Ferrer for her help with GC setup and calibration, and Dr. Kimberly Papadantonakis for help editing. This
31
32
material is based upon work performed by the Joint Center for Artificial Photosynthesis, a DOE Energy Innovation
33
34
Hub, supported through the Office of Science of the U.S. Department of Energy under Award Number DE-
35
36 SC0004993. DAT recognizes a Graduate Research Fellowship from the National Science Foundation for support
37
38 and the NWA. SAF acknowledges the Resnick Sustainability Institute at Caltech for a Postdoctoral Fellowship. JCC
39
40 acknowledges support from the Department of Defense through the National Defense Science & Engineering
41
42 Graduate Fellowship Program. NSL acknowledges support through the Multidisciplinary University Research
43
44 Initiative (MURI) under AFOSR Award No. FA9550-10-1-0572. BSB acknowledges support through the Beckman
45
46 Institute of the California Institute of Technology.
47
48
49
50 References
51
52 (1)Hori, Y.; Kikuchi, K..Suzuki, S. Chem. Lett. 1985, 14, 1695-1698.
53 (2)Hori, Y.; Murata, A..Takahashi, R. J. Chem. Soc., Faraday Trans. 1989, 85, 2309-2326.
54 (3)Hori, Y. In Modern Aspects of Electrochemistry; Vayenas, C., Ed.; Springer: New York, 2008; Vol. 42, p 89-189.
55 (4)Kuhl, K. P.; Cave, E. R.; Abram, D. N..Jaramillo, T. F. Energy Environ. Sci. 2012, 5, 7050-7059.
56 (5)Li, C. W.; Ciston, J..Kanan, M. W. Nature 2014, 508, 504-507.
57 (6)Li, C. W..Kanan, M. W. J. Am. Chem. Soc. 2012, 134, 7231-7234.
58 (7)Roberts, F. S.; Kuhl, K. P..Nilsson, A. Angew. Chem. Int. Ed. 2015, 54, 5179-5182.
59
60
9
ACS Paragon Plus Environment
ACS Catalysis Page 10 of 16

1
2
3 (8)Manthiram, K.; Beberwyck, B. J..Aivisatos, A. P. J. Am. Chem. Soc. 2014, 136, 13319-13325.
4 (9)Tang, W.; Peterson, A. A.; Varela, A. S.; Jovanov, Z. P.; Bech, L.; Durand, W. J.; Dahl, S.; Norskov, J.
5 K..Chorkendorff, I. PCCP 2012, 14, 76-81.
6 (10)Yano, H.; Tanaka, T.; Nakayama, M..Ogura, K. J. Electroanal. Chem. 2004, 565, 287-293.
7 (11)Kas, R.; Kortlever, R.; Milbrat, A.; Koper, M. T. M.; Mul, G..Baltrusaitis, J. PCCP 2014, 16, 12194-12201.
8 (12)Ren, D.; Deng, Y.; Handoko, A. D.; Chen, C. S.; Malkhandi, S..Yeo, B. S. ACS Catal. 2015, 5, 2814-2821.
9 (13)Neatu, S.; Antonio Macia-Agullo, J.; Concepcion, P..Garcia, H. J. Am. Chem. Soc. 2014, 136, 15969-15976.
10 (14)Karamad, M.; Tripkovic, V..Rossmeisl, J. ACS Catal. 2014, 4, 2268-2273.
11 (15)Watanabe, M.; Shibata, M.; Katoh, A.; Sakata, T..Azuma, M. J. Electroanal. Chem. 1991, 305, 319-328.
12 (16)Hahn, C., Abram, D.N., Hansen, H.A., Hatsukade, T., Jackson, A., Johnson, N.C., Hellstern, T.R., Kuhl, K.P.,
13 Cave, E.R., Feaster, J.T., Jaramillo, T.F. J. Mater. Chem. A 2015.
14 (17)Christophe, J.; Doneux, T..Buess-Herman, C. Electrocatalysis 2012, 3, 139-146.
15 (18)Jia, F.; Yu, X..Zhang, L. J. Power Sources 2014, 252, 85-89.
16 (19)Nakato, Y.; Yano, S.; Yamaguchi, T..Tsubomura, H. Denki Kagaku 1991, 59, 491-498.
17 (20)Rasul, S.; Anjum, D. H.; Jedidi, A.; Minenkov, Y.; Cavallo, L..Takanabe, K. Angew. Chem. Int. Ed. 2015, 54,
18 2146-2150.
19 (21)Watanabe, M.; Shibata, M.; Katoh, A.; Azuma, M..Sakata, T. Denki Kagaku 1991, 59, 508-516.
20 (22)Ishimaru, S.; Shiratsuchi, R..Nogami, G. J. Electrochem. Soc. 2000, 147, 1864-1867.
21 (23)Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M..Yang, P. Nature Communications 2014, 5:4948.
22 (24)Zhao, X.; Luo, B.; Long, R.; Wang, C..Xiong, Y. J. Mater. Chem. A 2015, 3, 4134-4138.
23 (25)Kortlever, R.; Peters, I.; Koper, S..Koper, M. T. M. ACS Catal. 2015, 5, 3916-3923.
24 (26)Canfield, D..Frese, K. W. J. Electrochem. Soc. 1983, 130, 1772-1773.
25 (27)Frese, K. W..Canfield, D. J. Electrochem. Soc. 1984, 131, 2518-2522.
26 (28)Barton, E. E.; Rampulla, D. M..Bocarsly, A. B. J. Am. Chem. Soc. 2008, 130, 6342-6344.
27 (29)Costentin, C.; Canales, J. C.; Haddou, B..Saveant, J.-M. J. Am. Chem. Soc. 2013, 135, 17671-17674.
28 (30)Sears, W. M..Morrison, S. R. J. Phys. Chem. 1985, 89, 3295-3298.
29 (31)Varela, A. S.; Sahraie, N. R.; Steinberg, J.; Ju, W.; Oh, H. S..Strasser, P. Angew. Chem. Int. Ed. 2015, 54,
30 10758-10762.
31 (32)Liu; Liu, Y.; Chen, S.; Quan, X..Yu, H. J. Am. Chem. Soc. 2015, 137, 11631-11636.
32 (33)Tamura, J., Ono, A., Sugano, Y., Huang, C., Nishizawa, H., Mikoshiba, S. PCCP 2012.
33 (34)Kuhl, K. P.; Hatsukade, T.; Cave, E. R.; Abram, D. N.; Kibsgaard, J..Jaramillo, T. F. J. Am. Chem. Soc. 2014.
34 (35)Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjaer, C. F.; Hummelshoj, J. S.; Dahl, S.; Chorkendorff,
35 I..Norskov, J. K. Nature Chemistry 2014, 6, 320-324.
36 (36)Hsu, L. S..Williams, R. S. J. Phys. Chem. Solids 1994, 55, 305-312.
37 (37)McCrory, C. C. L.; Jung, S. H.; Peters, J. C..Jaramillo, T. F. J. Am. Chem. Soc. 2013, 135, 16977-16987.
38 (38)Qiao, J.; Liu, Y.; Hong, F..Zhang, J. Chem. Soc. Rev. 2014, 43, 631-675.
39 (39)Qiao, J.; Liu, Y.; Hong, F..Zhang, J. Chem. Soc. Rev. 2014, 43, 631-675.
40 (40)Hatsukade, T.; Kuhl, K. P.; Cave, E. R.; Abram, D. N..Jaramillo, T. F. PCCP 2014, 16, 13814-13819.
41 (41)Gangeri, M.; Perathoner, S.; Caudo, S.; Centi, G.; Amadou, J.; Begin, D.; Pham-Huu, C.; Ledoux, M. J.;
42 Tessonnier, J. P.; Su, D. S..Schloegi, R. Catal. Today 2009, 143, 57-63.
43 (42)Schouten, K. J. P.; Gallent, E. P..Koper, M. T. M. ACS Catal. 2013, 3, 1292-1295.
44 (43)Peterson, A. A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J..Norskov, J. K. Energy Environ. Sci. 2010, 3, 1311-
45 1315.
46 (44)Hori, Y.; Takahashi, R.; Yoshinami, Y..Murata, A. J. Phys. Chem. B 1997, 101, 7075-7081.
47 (45)Hori, Y..Murata, A. Electrochim. Acta 1990, 35, 1777-1780.
48 (46)Schouten, K. J. P.; Gallent, E. P..Koper, M. T. M. J. Electroanal. Chem. 2014, 716, 53-57.
49
50
51
52
53
54
55
56
57
58
59
60
10
ACS Paragon Plus Environment
Page 11 of 16 ACS Catalysis

1
2
3
Table of Contents Graphic
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
S2
ACS Paragon Plus Environment
ACS Catalysis Page 12 of 16

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
ACS Paragon Plus Environment
45
46
47
Page 13 of 16 ACS Catalysis

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
ACS Paragon Plus Environment
45
46
47
ACS Catalysis Page 14 of 16

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
ACS Paragon Plus Environment
45
46
47
Page 15 of 16 ACS Catalysis

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
ACS Paragon Plus Environment
45
46
47
ACS Catalysis Page 16 of 16

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
ACS Paragon Plus Environment
45
46
47

You might also like