Sie sind auf Seite 1von 7

Journal of Applied Electrochemistry

https://doi.org/10.1007/s10800-018-1166-6

RESEARCH ARTICLE

Electrochemical reduction of ­CO2 at CuAu nanoparticles: size and alloy


effects
Evan Andrews1 · Yuxin Fang1 · John Flake1

Received: 13 October 2017 / Accepted: 5 February 2018


© Springer Science+Business Media B.V., part of Springer Nature 2018

Abstract
Reduction of C ­ O2 at Cu or Au electrodes typically yields methane or CO, respectively. Cu and Au nanoparticles and their
alloys offer unique advantages over foil electrodes in terms of reduced overpotentials and product selectivities. In this work,
we explore the electrochemical reduction of ­CO2 in aqueous electrolytes using alloys of Cu and Au including 2 and 6 nm
nanoparticles along with polycrystalline foils. These results show the CuAu alloys primarily produce CO; however, yields
are dramatically increased relative to Au. CuAu electrodes in the form of planar foils produce up to 3.4 times more CO
yields relative to Au foil. Most remarkably, nanoparticle electrodes provide up to 12.5-fold CO yield increases relative to
polycrystalline alloy foils and 175-fold CO yield increases relative to bulk Au foils. Voltammetry shows that onset potentials
for ­CO2 reduction are shifted anodically with smaller nanoparticle sizes and with greater Au content. The dramatic increase
in CO yields with nanoparticle alloys is attributed to the improved ­CO2 deoxygenation associated with Cu interfaces and
the relatively facile desorption of CO from low-coordination Au sites.
Graphical Abstract

Foil 2nm 6nm

O=C=O CO
O=

C C
Cu Au

Keywords  Copper · Gold · Alloy · CO2 reduction · Electrocatalysis · Carbon dioxide · Nanoparticle

1 Introduction

Transition metals such as Cu, Ni, Sn, and Fe have long been
Electronic supplementary material  The online version of this considered as potential catalysts for ­CO2 reduction [1–4].
article (https​://doi.org/10.1007/s1080​0-018-1166-6) contains
supplementary material, which is available to authorized users.
Of the transition metals, Cu [5–7] has been the most heav-
ily investigated as it produces hydrocarbons at relatively
* John Flake high current densities (~ 5 mA cm−2) and Faradaic efficien-
johnflake@lsu.edu cies > 60% [6, 8]. Likewise, other noble metals such as Au
1
Cain Department of Chemical Engineering, Louisiana State
and Ag are known to produce CO at relatively high current
University, Baton Rouge, LA 70803, USA densities and Faradaic efficiencies exceeding 90% [9, 10].

13
Vol.:(0123456789)
Journal of Applied Electrochemistry

In recent years, there has been significant progress in the follow the same linear relationship as pure metals [28]; so
synthesis of well-defined nanoparticles using wet proce- these may offer an opportunity to break the linear scaling in
dures such as Brust Schiffrin, Perrault [11], Martin [12], or binding energies and alter product selectivities.
Turkevich methods. These particles are particularly inter- Cu and Au are prime candidates for alloying, as Au has a
esting in their application as electrocatalysts as they offer high CO activity and Cu is known to be active in the hydro-
unique advantages over bulk analogs including remarkably genation of CO. Alloying Cu and Au causes the adsorption
greater activities. In terms of wet synthesis methods, the energies of *CO to shift by < 0.2 eV, while the energies of
most significant advances have been centered on Au nano- *OCOH shift by up to 0.5 eV [29]. A DFT simulation of a
particles including “magic number” nanoparticles that are ­Cu3Au1 surface by Hirunsit et al. suggests that the decrease
stabilized using ligands such as glutathione, dodecanethiol, in hydrocarbons observed with CuAu alloys relative to Cu is
triphenylphosphine, or phenylethanethiol [13]. These Au due to a weakened ability to bond CO [29]. Another mecha-
nanoparticles are relatively stable and do not agglomerate for nism to compensate for the weakened CO bonding is via
several months of low-temperature storage [13, 14]. The wet the use of nanoscale catalysts, which also has the benefit of
synthesis of Cu nanoparticles is somewhat more challenging increasing mass activity. CO preferentially adsorbs at edge
since these particles are relatively more susceptible to oxi- and corner sites, which are more numerous with nanoscale
dation and agglomeration; however, there have been several catalysts. The preference of CO for edge and corners is also
reports showing the synthesis of Cu nanoparticles [15–17]. apparent on high disorder single crystal catalysts [5]. Here,
In terms of the particle size effect on the electrochemical we explore the electrocatalytic behavior of CuAu alloys as
behavior of C ­ O2 reduction at Au and Cu electrodes, Zhu a function of alloy composition and particle size.
et al. and Kauffman et al. have shown that Au nanoparti-
cles are more active than bulk Au, producing at up to 97%
CO Faradaic efficiency at 8 nm sizes, and 100% using A ­ u25
[10, 18, 19]. Likewise, Reske et al. and Baturina et al. have 2 Experimental
shown that smaller Cu particles are nearly twice as active as
foil electrodes [15, 20]. The 2 nm nanoclusters were synthesized using the following
Previous reports with CuAu alloys have shown that sin- procedure modified from Hostetler et al. [30] and Yin et al.
gle crystal CuAu electrodes result in decreased methane, [31] ­HAuCl4 ­xH2O and ­CuCl2 were dissolved in DI water
ethylene, and ethanol yields compared to bulk Cu, but show along with KBr. Separate solutions with Cu/Au molar ratios
increases in CO yield, and at lower overpotentials [21, 22]. of 1:1, 3:1, and 9:1 were prepared. Tetraoctylammonium
Increasing the molar ratio of Au relative to Cu also shows bromide (TOAB) was dissolved in toluene. The solutions
increased CO production. For example, at a C ­ u50Au50 ratio, were stirred together until the aqueous solution was clear,
the CO partial current was ~ 1.5 times greater than bulk indicating the phase transfer of the ions. Dodecanethiol
Au [22]. Likewise, 10-nm CuAu nanoclusters have been (600 µL) was added and the solution was stirred for 1 h or
reported to produce CO at up to 65% Faradaic efficiency, until the toluene phase became clear. 0.5 g of ­NaBH4 in DI
with ­Cu1Au3 alloys outperforming pure Au nanoparticles water was slowly added to the solution, resulting in a brown
in both yield (~ 10% greater) and Faradaic efficiency [21]. color. The solution was stirred for 3 h and the aqueous phase
Density functional theory (DFT) studies have shown that removed by pipette. The toluene solvent was evaporated
alloy electrodes offer a potential means to create catalysts under vacuum and the nanoparticles washed and filtered
that outperform pure metal catalysts due to low coordina- with ethanol.
tion. DFT has been used to determine the binding energies The 6-nm nanoclusters were synthesized using a modified
of intermediates, allowing for the prediction of reaction procedure based on the literature [31, 32]. 2-nm nanoparti-
pathways based on thermodynamics and kinetics [23–25]. cles were synthesized following the previous procedure, but
Hansen et al. show that the binding energies of the *CO and after removing the aqueous phase, the toluene was evapo-
*COOH intermediates adsorbed on transition metal surfaces rated to give a nanoparticle solution of ~ 15× concentra-
follow linear scaling relationships [26]. On all transition tion. The resulting solution was sealed in a glass reactor
metals, the adsorption energies of the *CO, *COOH, and and heated at 150 °C for 2 h. The remaining toluene was
*CHO intermediates are proportionally in such a way that evaporated under vacuum and the nanoparticles washed and
the reaction energetics and pathways are not altered by the filtered with ethanol.
electronic properties of the catalyst [27]. This limitation sug- Bulk alloy foils were obtained commercially from ESPI
gests that the optimal pathways for ­CO2 reduction to hydro- metals. The alloys were 50, 25, and 10% Au by weight, cor-
carbons cannot be obtained simply by using pure transition responding to molar ratios of C­ u75.6Au24.4, ­Cu90.3Au9.7, and
metal electrodes, although Cu is nearest to the optimal spot ­Cu96.6Au3.4, respectively. Several 1 cm2 squares were cut
[26]. Other DFT simulations show that Cu alloys do not from the foils and used as working electrodes for electrolysis.

13
Journal of Applied Electrochemistry

Transmission electron microscopy (TEM) samples were purged with ­N2, and used for converting to the RHE scale.
prepared by dispersing dilute samples in toluene onto Cu Voltammetry was conducted at a scan rate of 10 mV s−1.
mesh grids (Fig. 1). The diameters were measured and aver- Onset potentials were calculated by plotting the natural log
aged. (Table S1). of current against voltage and determining where the curve
X-ray photoelectron spectroscopy (XPS) was used to departed from linear behavior (Figure S3). Reaction yields
determine the Cu:Au ratios of the nanoparticles after syn- were determined using potentiostatic experiments using gas
thesis (Figures S1, S2). chromatography (TCD and FID).
Electrodes were prepared by mixing the nanoparticles The yield data were normalized using the surface area of
with carbon black inks using Nafion™ as a conductive poly- the nanoparticles in order to compare nano and bulk scale
mer binder. 100-mg carbon black was mixed with 50 mg of catalysts. Surface area normalization was also necessary
Nafion™ resin and 10 mg of nanoparticles in 2 mL acetone, to compare nanoparticles of the same size, as the density
then sonicated for 30 min to ensure even dispersion. Ink was of the alloy changes with composition. The nanoparticle
applied to a glassy carbon electrode in a Teflon™ holder by weight loading was estimated from the ink composition and
paintbrush, giving a nanoparticle loading of 0.06 mg cm−2. the number and geometric surface area of the nanoparticles
Electrochemical experiments were performed in a sealed was calculated using the average sizes from the TEM images
two-compartment cell with a Nafion™ membrane separat- (Equation S1). 50% of the surface area was assumed inac-
ing the two compartments. One compartment contained the cessible due to the carbon black substrate and the Nafion™
Ag/AgCl reference electrode and working electrode, and the polymer.
other compartment contained the Pt counter electrode. 0.1M
­KHCO3 was used as the electrolyte. The catholyte in the
working compartment was continuously bubbled with ­CO2 3 Results and discussion
and effluent gas vented to a gas chromatograph injection
port (Shimadzu GC-2014). The pH of carbonate electrolyte XPS characterization of the alloys showed the 2 nm nan-
was measured at 6.8 when saturated with C ­ O2, and 8.9 when oparticles’ compositions to be ­Cu25Au75, ­Cu62Au38, and
­Cu70Au30 while the 6 nm nanoparticles’ compositions were
­Cu9Au91, ­Cu38Au62, and ­Cu59Au41. The noticeable difference
in 2 and 6 nm compositions is likely due to migration of Au
towards the surface during heat treatment due to its lower
surface energy [33]. While there is some level of surface
segregation, our XPS and that of literature results using the
same synthesis techniques show that an alloy is present on
the surface despite segregation [30, 31, 34].
Bulk foils were evaluated as catalysts in order to set a
baseline for comparison with the nanoparticles. Figure 2
shows the CO yields of the bulk alloy foils as well as indi-
vidual Cu and Au foils. All alloy foils show higher CO yields
compared to individual Cu and Au foils, with the ­Cu75Au25
foil yielding the most. At − 1.0 V vs. RHE, ­Cu75Au25 foils
yield 3.4 times more CO than bulk Au foil and 22.3 times
more than bulk Cu foil. However, Au foil showed a higher
Faradaic efficiency than ­Cu90Au10 and ­Cu96Au4 alloys at
lower overpotentials, likely due to the increased hydrogen
affinity of Cu (Fig. 3). This is consistent with the results
published previously by Christophe et al. [22].
The 2 nm alloy nanoparticles’ CO yields and Faradaic
efficiencies are shown in Figs. 4 and 5, respectively. The
­C u 25Au 75 alloy gave the highest CO yield, peaking at
1.47 mmol cm−2 h−1 and producing ~ 9 times the CO yield
as 2 nm ­Cu70Au30 at − 1.1 V vs. RHE. ­Cu25Au75 also had the
greatest peak CO Faradaic efficiency of 37%, while ­Cu70Au30
possessed only 22% Faradaic efficiency towards CO.
­Cu68Au32 showed a maximum yield of 0.65 mmol cm−2 h−1
Fig. 1  Cu68Au32 2 nm (top) and ­Cu38Au62 6 nm (bottom) TEM and a Faradaic efficiency of 28%.

13
Journal of Applied Electrochemistry

Fig. 2  CO yield on bulk CuAu, pure Au, and pure Cu foils. Yields are
Fig. 4  CO yield on 2 nm CuAu nanoparticle catalysts. Yields are nor-
normalized to foil surface area
malized to the surface areas of the nanoparticles

Fig. 3  Faradaic efficiency of CO on bulk CuAu, Au, and Cu foils


Fig. 5  Faradaic efficiency of 2 nm CuAu nanoparticle catalysts

The 6 nm alloy nanoparticles’ CO yields and Faradaic effi- and 2 nm ­Cu25Au75 showing ~ 29× the CO yield of bulk
ciencies are shown in Figs. 6 and 7, respectively. ­Cu38Au62 Au. At their maximums, the 6-nm nanoparticles (Fig. 6)
gave the highest CO yield, peaking at 8.8 mmol cm−2 h−1 yield 8.8 mmol cm−2 h−1 of CO; the 2-nm nanoparticles
with a Faradaic efficiency of 50%. Despite having a lower (Fig. 4) yield 1.47 mmol cm−2 h−1, and the foils (Fig. 2)
relative yield of 5.7 mmol cm−2 h−1, ­Cu9Au91 possessed the yield only 0.175 mmol cm−2 h−1. Previous reports involv-
greatest Faradaic efficiency towards CO with a maximum ing ­CO2 reduction on 1.9 and 4.8 nm Cu nanoparticles [15]
of 73%. ­Cu59Au41 showed a maximum yield and Faradaic and 2 and 6 nm Au nanoparticles [10] have described single
efficiency of 1.25 mmol cm−2 h−1 and 18%, respectively. order of magnitude increases (×9 and ×40 for Au 6 and 2 nm
When compared with foils of similar composition, the nanoparticles respectively at − 1.2 V vs. RHE).
CuAu nanoparticles exhibit increased CO yields. At − 1.1 V The ­CO2 reduction onset potentials (Table 1) are indic-
vs. RHE, 6 nm C ­ u59Au41 produces CO at ~ 10 times the ative of a size effect being responsible, as the size of the
yield of bulk C­ u75Au25. When compared with the pure Cu catalyst has a significant effect on the onset potential. The
and Au foils, the difference is even more distinct, with 6 nm 6-nm nanoparticle electrodes show at least a 200 mV anodic
­Cu38Au62 showing ~ 175 times the CO yield of bulk Au shift compared to bulk alloys, and the 2-nm nanoparticles

13
Journal of Applied Electrochemistry

Table 2  HER formation onset potentials in N


­ 2 purged 0.1 M ­KHCO3
HER formation onset potentials/V vs. RHE
Bulk 2 nm 6 nm

Au − 0.36 – – Cu9Au91 − 0.41


Cu75Au25 − 0.38 Cu25Au75 − 0.45 Cu38Au62 − 0.47
Cu90Au10 − 0.37 Cu62Au38 − 0.48 Cu59Au41 − 0.54
Cu96Au4 − 0.38 Cu70Au30 − 0.50 – –

show at least a 150 mV anodic shift. The onset potential is


not heavily dependent on alloy composition, with catalysts
of the same size having potentials within 50 mV windows.
The onset potentials for the hydrogen evolution reaction
are given in Table 2. While there was some cathodic shift
in hydrogen evolution reaction onset, it was not significant
relative to that observed with C ­ O2 reduction, indicating that
Fig. 6  CO yield on 6 nm CuAu nanoparticle catalysts. Yields are nor-
malized to the surface areas of the nanoparticles
the increases in CO yield are not simply from changes in
surface area.
The relatively high activity of the 6-nm nanoparticles
compared to the 2 nm nanoparticles is somewhat surpris-
ing. The difference may be due to the size effects observed
on small nanoparticles where there are optimal particle
sizes for reaction activity [35, 36]. The electrode morphol-
ogy and binder may also play a role in the activity of small
particle electrodes; smaller nanoparticles are more likely to
be trapped in internal pores, limiting their access to C ­ O2
[37]. However, the most likely explanation is that the ­CO2
reduction active sites on a 6 nm are denser than on a 2-nm
nanoparticle. The edge site to corner site ratio is higher on
a larger diameter spherical nanoparticle [38], and as shown
by Zhu et al., metal catalysts with high edge to corner ratios
such as Au nanowires are very active for ­CO2 reduction [9].
Mistry et al. show that Au nanoparticles in the 2–8 nm size
range have significant size effects, where hydrogen selectiv-
ity increases as the particle size decreases, a result attributed
to the higher ratio of corner sites where hydrogen evolution
Fig. 7  Faradaic efficiency of CO on 6  nm CuAu nanoparticle cata-
lysts is preferred [39]. The relatively low CO Faradaic efficiency
of the 2-nm CuAu nanoparticles compared to CuAu foils
matches this observation. Despite the improved yields, the
Table 1  CO2 reduction onset potentials in C
­ O2 saturated 0.1  M overall selectivity shifts towards hydrogen.
­KHCO3 Although lesser than the size effects, alloy composition
CO2 reduction onset potentials/V vs. RHE effects were also readily observed in both the foil and the
2 and 6 nm nanoparticle electrodes, with a larger Au ratio
Bulk 2 nm 6 nm
in the alloy generally correlating to a greater CO yield. At
Au − 0.42 – – Cu9Au91 − 0.14 − 1.0 V vs. RHE, the ­Cu75Au25 alloy foil yields ~ 1.2 times
Cu75Au25 − 0.44 Cu25Au75 − 0.22 Cu38Au62 − 0.12 more CO than the ­Cu90Au10 and ­Cu96Au4 foils. Similarly,
Cu90Au10 − 0.43 Cu62Au38 − 0.25 Cu59Au41 − 0.13 at − 1.1 V vs. RHE, the 2 nm C ­ u25Au75 produces ~ 9 times
Cu96Au4 − 0.37 Cu70Au30 − 0.26 – – the CO yield as the 2 nm C ­ u70Au30 and the 6 nm C ­ u38Au62
produces ~ 11 times the CO yield as the ­Cu59Au41. How-
ever, the 6-nm ­C u 9Au 91 nanoparticles yield ~ 50% less
CO than the 6-nm C ­ u38Au62 nanoparticles, despite having
a Faradaic efficiency of 77% towards CO. An excess of

13
Journal of Applied Electrochemistry

Au content may result in higher Faradaic efficiency, up The results from this work suggest that the dramatic yield
to the > 95% seen from the pure Au nanoparticles from increases originate from both improved deoxygenation of
Kauffman et al. [10], but decrease the net catalytic activity. ­CO2 at Cu and the reduced adsorption energy of CO at Au.
There may be an optimum CuAu alloy ratio for CO yield Likewise, the remarkable increases in yield associated with
between the ­Cu38Au62 and ­Cu9Au91 compositions. nanoscale CuAu alloys suggest that the relative impact of
The increased CO yields have been attributed to these effects is significantly greater at low-coordinated
improved CO desorption on alloys due to weakened CO Cu–Au interfaces, as the 6-nm nanoparticles which have
bond energy on Cu [22]. Alloying the metals results in the highest ratio of edge sites give the greatest CO yield
a change in electronegativity, showing a shift in elec- [38]. It was also found that the nanoparticles with higher
tron charge of ~ 0.2 eV towards Au [40]. This results in Au ratios were more favorable for CO formation which is
improved Au binding sites for CO, which is an electron in agreement with DFT by Lysgaard et al. [42]. The DFT
acceptor, primarily bonding to the surface via pi orbitals, suggests that Au heavy surfaces with Cu cores are the most
and prefers to bind at metal sites with electron rich d orbit- stable structurally and have similar adsorption strength to
als [41]. Likewise, *COOH is expected to be stabilized on pure Au nanoparticles, though CuAu bimetallic sites are the
CuAu alloy corner sites [42]. As such, *CO and *COOH most preferred for CO adsorption.
bonded to CuAu will be less stable relative to Cu surfaces,
but more stable than Au.
There are several explanations for the underlying nature 4 Conclusion
of the dramatic improvements in CO yield observed with
nanoparticle CuAu alloys. Christophe et al. proposed that CuAu alloys significantly increase CO yield per area com-
electronic effects can alter the bond strength of CO, causing pared to pure Au or Cu electrocatalysts, and as nanoscale
an increase in kinetics [22]. Hirunsit et al. have used DFT catalysts enhance the CO yield even further. The use of alloy
calculations to reach a similar conclusion, showing that the foils leads to a ~ 3× increase in CO yields compared to pure
adsorption energy of *CO is 0.19 eV weaker on C ­ u75Au25 metals, while the use of CuAu nanoparticles leads to an
than on Cu and that the change in bond energy plays a increase in CO yield by > 175×. In addition to the increased
role on several alloys [28, 29]. The most significant yield yields, voltammetry shows reduction C ­ O2 reduction onset
increases are associated with nanoparticle alloys. Kim et al. potentials are shifted anodically by > 150 mV for 2 nm
proposed that the presence of Cu adjacent to Au surface alloys and > 200 mV for 6 nm alloy electrodes. Likewise, the
atoms also increases the stability of the *COOH intermedi- increased yields and anodically shifted onset potentials are
ate by allowing a single *COOH to bond to both Au and Cu associated with an increase in edge sites and the synergistic
atoms [21]. The Cu atom is the more oxyphilic of the two electronic effects of CuAu alloys on the nanoparticle alloy
metals, so the C atom bonds to the Au, and an O to the Cu. electrodes. These results suggest that CuAu alloys, particu-
As the alloys increase in Au content, the number of Cu sites larly low-coordinated interfaces thereof, improve both the
with neighboring Au atoms increases, which contributes deoxygenation ­CO2 and desorption of CO.
towards the increase in CO yield. The aforementioned reac-
tions are most likely to occur at low-coordinated sites and Acknowledgements  This work was supported by the National Science
thus the alloy effect is more pronounced at the nanoscale. Foundation under Grant CBET-1438385.
Other groups utilizing alloys such as CuIn [43], ­Cu3Ni [44],
and AuCd [45] have also proposed bimetallic active sites as
the key part of the electrocatalysts; however, DFT calcula- References
tions of CuAu alloy nanoclusters performed by Lysgaard
et al. suggest no O–Cu bonds were formed by *COOH in 1. Hori Y, Kikuchi K, Suzuki S (1985) Chem Lett 14(11):1695–1698
2. Hori Y (2008) Modern aspects of electrochemistry. Springer, New
their simulation, although Au edge and apex sites adjacent to
York, pp 89–189
Cu were still preferred for reactivity [42]. Other works pro- 3. Hori Y, Wakebe H, Tsukamoto T, Koga O (1994) Electrochim
pose the metal with lower surface energy (Au in this case) Acta 39(11):1833–1839
is observed to migrate to the surface, causing the alloy to act 4. Jitaru M, Lowy DA, Toma M, Toma BC, Oniciu L (1997) J Appl
Electrochem 27(8):875–889
similarly to the segregated surface metal [33, 46]. Likewise,
5. Hori Y, Takahashi I, Koga O, Hoshi N (2002) J Phys Chem B
works considering intentionally segregated electrocatalysts 106(1):15–17
such as core–shell nanoparticles [47] and thin overlayers 6. Hori Y, Kikuchi K, Murata A, Suzuki S (1986) Chem Lett
have noted the lattice strain as another factor affecting the 15(6):897–898
7. DeWulf DW, Jin T, Bard AJ (1989) J Electrochem Soc
reaction [33, 48–50]. In all cases, edge sites [9] and other
136(6):1686–1691
low-coordination sites such as grain boundaries [51] are 8. Kuhl KP, Cave ER, Abram DN, Jaramillo TF (2012) Energy Envi-
identified as the key active sites for ­CO2 reduction. ron Sci 5(5):7050–7059

13
Journal of Applied Electrochemistry

9. Zhu W, Zhang Y-J, Zhang H, Lv H, Li Q, Michalsky R, Peterson 32. Maye MM, Zheng W, Leibowitz FL, Ly NK, Zhong C-J (2000)
AA, Sun S (2014) J Am Chem Soc 136(46):16132–16135 Langmuir 16(2):490–497
10. Kauffman DR, Alfonso D, Matranga C, Qian H, Jin R (2012) J 33. Friebel D, Mbuga F, Rajasekaran S, Miller DJ, Ogasawara H,
Am Chem Soc 134(24):10237–10243 Alonso-Mori R, Sokaras D, Nordlund D, Weng T-C, Nilsson A
11. Per rault SD, Chan WCW (2009) J Am Chem Soc
(2014) J Phys Chem C 118(15):7954–7961
131(47):17042–17043 34. Llorca J, Domínguez M, Ledesma C, Chimentão RJ, Medina
12. Martin MN, Basham JI, Chando P, Eah S-K (2010) Langmuir F, Sueiras J, Angurell I, Seco M, Rossell O (2008) J Catal
26(10):7410–7417 258(1):187–198
13. Negishi Y, Nobusada K, Tsukuda T (2005) J Am Chem Soc 35. Kinoshita K (1990) J Electrochem Soc 137(3):845–848
127(14):5261–5270 36. Shao M, Peles A, Shoemaker K (2011) Nano Lett

14. Boyen HG, Kästle G, Weigl F, Koslowski B, Dietrich C, Ziemann 11(9):3714–3719
P, Spatz JP, Riethmüller S, Hartmann C, Möller M (2002) Science 37. Uchida M, Park Y-C, Kakinuma K, Yano H, Tryk DA, Kamino
297(5586):1533–1536 T, Uchida H, Watanabe M (2013) Phys Chem Chem Phys
15. Reske R, Mistry H, Behafarid F, Roldan B, Cuenya, Strasser P 15(27):11236–11247
(2014) J Am Chem Soc 136(19):6978–6986 38. Back S, Yeom MS, Jung Y (2015) ACS Catal 5(9):5089–5096
16. Wu S-H, Chen D-H (2004) J Colloid Interface Sci 273(1):165–169 39. Mistry H, Reske R, Zeng Z, Zhao Z-J, Greeley J, Strasser P,
17. Lisiecki I, Billoudet F, Pileni MP (1996) J Phys Chem Cuenya BR (2014) J Am Chem Soc 136(47):16473–16476
100(10):4160–4166 40. Sham TK, Hiraya A, Watanabe M (1997) Phys Rev B 55(12):7585
18. Zhu W, Michalsky R, Metin Ö, Lv H, Guo S, Wright CJ, Sun X, 41. Bagus PS, Nelin CJ, Bauschlicher CW (1984) J Vac Sci Technol
Peterson AA, Sun S (2013) J Am Chem Soc 135(45):16833–16836 A 2(2):905–909
19. Chen Y, Li CW, Kanan MW (2012) J Am Chem Soc 42. Lysgaard S, Myrdal JSG, Hansen HA, Vegge T (2015) Phys Chem
134(49):19969–19972 Chem Phys 17(42):28270–28276
20. Baturina OA, Lu Q, Padilla MA, Xin L, Li W, Serov A, Artyush- 43. Rasul S, Anjum DH, Jedidi A, Minenkov Y, Cavallo L, Takanabe
kova K, Atanassov P, Xu F, Epshteyn A, Brintlinger T, Schuette K (2015) Angew Chem 127(7):2174–2178
M, Collins GE (2014) ACS Catal 4(10):3682–3695 44. Adit Maark T, Nanda BRK (2016) J Phys Chem C

21. Kim D, Resasco J, Yu Y, Asiri AM, Yang P (2014) Nat Commun 120(16):8781–8789
5:4948 45. Jovanov ZP, Hansen HA, Varela AS, Malacrida P, Peterson
22. Christophe J, Doneux T, Buess-Herman C (2012) Electrocatalysis AA, Nørskov JK, Stephens IEL, Chorkendorff I (2016) J Catal
3(2):139–146 343:215–231
23. Hansen HA, Montoya JH, Zhang Y-J, Shi C, Peterson AA, Nør- 46. Choi J, Kim MJ, Ahn SH, Choi I, Jang JH, Ham YS, Kim JJ, Kim
skov JK (2013) Catal Lett 143(7):631–635 S-K (2016) Chem Eng J 299:37–44
24. Nie X, Esopi MR, Janik MJ, Asthagiri A (2013) Angew Chem Int 47. Monzo J, Malewski Y, Kortlever R, Vidal-Iglesias FJ, Solla-
Ed 52(9):2459–2462 Gullon J, Koper MTM, Rodriguez P (2015) J Mater Chem A
25. Schouten KJP, Kwon Y, van der Ham CJM, Qin Z, Koper MTM 3(47):23690–23698
(2011) Chem Sci 2(10):1902–1909 48. Reske R, Duca M, Oezaslan M, Schouten KJP, Koper MTM,
26. Hansen HA, Varley JB, Peterson AA, Nørskov JK (2013) J Phys Strasser P (2013) J Phys Chem Lett 4(15):2410–2413
Chem Lett 4(3):388–392 49. Todoroki N, Yokota N, Nakahata S, Nakamura H, Wadayama T
27. Peterson AA, Nørskov JK (2012) J Phys Chem Lett 3(2):251–258 (2016) Electrocatalysis 7(1):97–103
28. Hirunsit P, Soodsawang W, Limtrakul J (2015) J Phys Chem C 50. Varela AS, Schlaup C, Jovanov ZP, Malacrida P, Horch
119(15):8238–8249 S, Stephens IEL, Chorkendorff I (2013) J Phys Chem C
29. Hirunsit P (2013) J Phys Chem C 117(16):8262–8268 117(40):20500–20508
30. Hostetler MJ, Zhong C-J, Yen BKH, Anderegg J, Gross SM, 51. Feng X, Jiang K, Fan S, Kanan MW (2015) J Am Chem Soc
Evans ND, Porter M, Murray RW (1998) J Am Chem Soc 137(14):4606–4609
120(36):9396–9397
31. Yin J, Shan S, Yang L, Mott D, Malis O, Petkov V, Cai F, Shan
Ng M, Luo J, Chen BH (2012) Chem Mater 24(24):4662–4674

13

Das könnte Ihnen auch gefallen