Controllable Modification of the Electronic Structure of Carbon

May 13, 2014 - Nitrogen-doped carbon-TiO2 composite as support of Pd electrocatalyst for formic acid oxidation. Yuan-Hang Qin , Yunfeng Li , Thomas La...
0 downloads 6 Views 2MB Size
Subscriber access provided by UNIV OF ALABAMA BIRMINGHAM

Article

Controllable Modification of the Electronic Structure of CarbonSupported Core–Shell Cu@Pd Catalysts for Formic-Acid Oxidation Mingjun Ren, Yi Zhou, Feifei Tao, Zhiqing Zou, Daniel L. Akins, and Hui Yang J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/jp5033417 • Publication Date (Web): 13 May 2014 Downloaded from http://pubs.acs.org on May 26, 2014

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 26

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

Controllable Modification of the Electronic Structure of Carbon-Supported

2

Core–Shell Cu@Pd Catalysts for Formic Acid Oxidation

3

Mingjun Ren,†,‡ Yi Zhou,† Feifei Tao,† Zhiqing Zou,† Daniel L. Akins,§and Hui Yang,†,*

4

† Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China

5

‡ University of Chinese Academy of Sciences (CAS), Beijing 100039, China

6

§CASI, The City College of The City University of New York, NY 10031, USA

7

ABSTRACT: This study analyzes the synthesis of carbon-supported core–shell structured Cu@Pd

8

catalysts (Cu@Pd/C) through a galvanic replacement reaction to be utilized in the electrocatalytic

9

oxidation of formic acid. The strategy used in this study explores the relationship among lattice

10

strain, electronic structure, and catalytic performance. X-ray diffraction and X-ray photoelectron

11

spectroscopy indicate that the inclusion of Cu in the nanocatalyst increases lattice strain and results

12

in a downshift of the d-band of palladium. Electrochemical tests show that Cu@Pd/C catalysts

13

exhibit weaker adsorption strength for CO with increased Cu content, which can be attributed to

14

the downshift of the electronic d-band. For the synthesized materials, the Cu@Pd/C catalyst with a

15

Cu:Pd atomic ratio of 27:73 is found to have the highest activity for formic acid oxidation. A peak-

16

like plot between activity and atomic composition is acquired and reveals the relationship among

17

lattice strain, electronic structure, and catalytic performance.

18

Keywords: Electrocatalysis, d-band center, lattice strain, galvanic replacement reaction

19

ACS Paragon Plus Environment

1

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

20

Page 2 of 26

1. Introduction

21

Research aimed at creating high-performance heterogeneous bimetallic/trimetallic catalysts to

22

overcome commercial viability issues regarding the wide utility of fuel cells has been conducted

23

for several decades.1–5 The main reason for the focus on bimetallic/trimetallic catalysts is the

24

compositional diversity inherent in using such catalysts, which provides a number of opportunities

25

to control the properties of a material with resultant performance enhancements, e.g., catalytic

26

activity and stability through geometrical or ligand effects.5,6 Progress has been made in the

27

preparation of core–shell Cu@Pt and hollow-structured Pt nanoscale materials, which have

28

suggested that lattice contraction at the surface plays a role in performance enhancement for the

29

oxygen reduction reaction.7,8 Kibler and coworkers deposited Pd monolayer atoms onto a series of

30

metallic substrates and clarified the relationship among lattice strain (geometric effect), ligand

31

effect, and catalytic performance for formic acid oxidation.9 Generally, such properties are

32

influenced by the d-band of the host metal, which is known to affect the performance of catalysts

33

significantly.9–11 However, significantly more mechanistic studies of related properties should be

34

conducted, such as the characterization of how lattice strain affects the electronic structure of a

35

metal, the catalytic performance, and the design of high-performance catalysts.

36

In the last decade, direct formic acid fuel cells (DFAFCs) have drawn increasing attention for

37

transportation and portable electronic device applications. Compared with methanol, formic acid

38

has numerous advantages as a fuel. These advantages include higher kinetic activity, lower

39

crossover through the Nafion membrane, high practical energy density, ease of integration into a

40

power system, nontoxicity, and others.1,2,12,13 However, the performances of DFAFCs are still

41

negatively affected by several problems.14–18 In particular, Pd has been shown to perform better

42

than Pt for formic acid oxidation,1,18 and the high initial catalytic activity of Pd compared with Pt

ACS Paragon Plus Environment

2

Page 3 of 26

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

43

results in its preferred use for formic acid oxidation.1,18 However, state-of-the-art Pd-based

44

catalysts suffer from lower stability and activity than Pt-based catalysts because of poisoning of

45

the electrode by CO adsorbate (COad).14,15,17 As a result, the development of high-performance Pd-

46

based catalyst for use in DFAFCs continues to be an active research focus area. For example,

47

porous Cu–Pd alloy and Pd-enriched Cu–Pd nanoparticles have been recently prepared, and the

48

authors have intimated that the enhancement might be the result of variation of electronic structure

49

through a synergistic effect that involves Cu.19–21 Furthermore, recent investigations that use X-ray

50

photoelectron spectroscopy (XPS) have detected electronic structure variation in Cu–Pd catalysts

51

and have indicated that enhanced performance for formic acid correlates with the downshift of the

52

d-band center.22,23 However, such findings are insufficient to account for the level of improved

53

performance attained.

54

In the present study, Cu@Pd/C catalysts are prepared with different atomic compositions

55

through galvanic replacement reactions. The lattice strain increase associated with the lattice

56

contraction in the crystal and its relationship to the electronic structure and catalytic performance

57

for formic acid oxidation are discussed.

58

2. Experimental Section

59

2.1 Preparation of Catalysts

60

Palladium chloride (PdCl2 AR), copper sulfate (CuSO4·5H2O, AR), ethylene glycol (EG, AR),

61

sodium chloride (AR), formic acid (AR), sodium hydroxide (AR), and sulfuric acid (AR) were

62

obtained from Sinopharm Chemical Reagent Co. Ltd (SCRC). VXC-72 carbon and 5 wt.% Nafion

63

solution were purchased from Cabot Co. and Sigma–Aldrich, respectively.

64

The Cu@Pd/C catalysts were synthesized by a two-step process. Briefly, a selected amount of

65

copper sulfate was added to a flask that contained 20 mL of 0.25 M sodium hydroxide in EG. The

ACS Paragon Plus Environment

3

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 26

66

mixture was continuously stirred at 80 °C under high-purity N2 flow for 1 h. Then, the mixture

67

was refluxed at 160 °C under N2 gas atmosphere for 12 h. The resultant reddish brown mixture

68

was cooled in ice-cold water. Next, 0.7 mL of concentrated sulfuric acid and 12.5 mL of 40 mM

69

Na2PdCl4 were added to the mixture under ultrasonic dispersion for 1 h. Finally, 80 mg VXC-72

70

carbon dispersed in 200 mL EG was introduced to support the metallic nanoparticles. Precipitates

71

were collected by filtration and washed three times with ultrapure water.

72

Pd/C catalyst was prepared using ethylene glycol at 140 °C as the dispersant and reductant.

73

Typically, a selected amount of Na2PdCl4 in EG was mixed with 20 ml of 0.25 M sodium

74

hydroxide in EG, followed by the addition of 80 mg VXC-72 carbon in EG. The resultant 200 mL

75

mixture was heated at 140 °C for 12 h under a N2 gas atmosphere. Lastly, the precipitates were

76

collected by filtration and washed three times with ultrapure water.

77

2.2 Physical Characterization

78

X-ray diffraction (XRD) measurements were conducted using a Bruke D8 Advanced XRD, with

79

Cu Kα1 radiation (λ = 1.54 Å). The voltage of the tube was maintained at 40 kV and the current at

80

100 mA. Diffraction patterns in the range between 20° and 90° were collected at a scanning rate of

81

2° min-1 with a step size of 0.02°; in the ranges of 30° to 45° and 61° to 76°, the scanning rate was

82

1° min-1.

83

Evaluation of the electronic structure of the catalysts was conducted with an XPS (Kratos Axis

84

Ultra DLD, Britain) using Al Kα radiation. The binding energy (BE) was calibrated using the C1s at

85

284.45 eV.

86

Transmission electron microscopy (TEM) characterizations were conducted on a JEOL JEM-

87

2100F TEM. The samples were prepared by ultrasonically suspending catalyst powder in ethanol.

88

A drop of suspension was placed onto a holey copper grid and dried under air. The atomic

ACS Paragon Plus Environment

4

Page 5 of 26

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

89

composition of the catalysts was determined with an IRIS advantage inductively coupled plasma

90

atomic emission spectroscopy (ICP-AES).

91

2.3 Electrochemical Evaluation of Catalysts

92

Electrochemical experiments were performed using a Solartron Electrochemical Interface

93

SI1287 with a standard three-electrode cell. Exactly, 10 mg of the catalyst, 0.5 mL of 5 wt.%

94

Nafion solution, and 2.5 mL of ultrapure water were ultrasonically mixed to form the catalyst ink.

95

Subsequently, 3 µL of such a mixture was transferred onto a glassy carbon (GC; 3 mm in diameter)

96

electrode using a microsyringe. The electrolyte used was 0.5 M H2SO4 or 0.5 M H2SO4 + 0.5 M

97

HCOOH. High-purity N2 was used to deaerate the solutions. To determine the real electrochemical

98

surface area (ECSA) of the catalyst, cyclic voltammograms (CVs) and CO-stripping

99

voltammograms were recorded. For the CO-stripping voltammograms, CO was pre-adsorbed at an

100

open potential for 30 min by bubbling CO into 0.5 M H2SO4 solution. Then, dissolved CO was

101

subsequently removed by purging for 30 min with high-purity N2. In all cases, electrochemical

102

measurements were conducted at a temperature of 25 ± 1 °C.

103 104 105

The ECSA of catalysts can be obtained by calculating the charge accumulated during the CO desorption and/or H desorption: ECSA = 100×Q/(m×C)

106

where Q(µC) is the charge caused by the H desorption or CO desorption; C is the electrical charge

107

associated with monolayer adsorption of H (210 µC cm-2) or CO (420 µC cm-2); and m is the

108

catalyst loading on the GC electrode (10 µg).

109

3. Results and Discussion

110

3.1 Composition and Morphology

ACS Paragon Plus Environment

5

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 26

111

The average bulk compositions of the catalysts were determined using ICP-AES analysis by

112

atomic compositions of the Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and

113

Cu44@Pd56/C catalysts. The Pd loadings were 38.5, 36.9, 36.73, 36.04, 35.63, and 34.5 wt.% for

114

the

115

Cu44@Pd56/C, respectively; whereas the Cu loadings exhibited a reverse ordering of 0, 4.0, 5.92,

116

7.77, 10.53, and 15.74 wt.%. For clear comparisons, the as-prepared Pd/C, Cu15@Pd85/C,

117

Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and Cu44@Pd56/C catalysts will be correspondingly

118

labelled using letters a to f in some of the following figures. Given that the Cu is inactive in the

119

formic acid oxidation, comparing only the Pd loading levels to analyze the performance of the

120

various catalyst compositions would seem valid.21,22

as-prepared

Pd/C,

Cu15@Pd85/C,

Cu21@Pd79/C,

Cu27@Pd73/C,

Cu35@Pd65/C,

and

121 122

Figure 1. TEM images (scale bar: 20 nm) of the as-prepared Pd/C (A), Cu15@Pd85/C (B),

123

Cu21@Pd79/C (C), Cu27@Pd73/C (D), Cu35@Pd65/C (E), and Cu44@Pd56/C (F).

ACS Paragon Plus Environment

6

Page 7 of 26

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

124

The morphologies of the as-prepared Pd/C and Cu@Pd/C were determined using TEM. Figures

125

1A–1F shows that all the metallic NPs are well dispersed on the surface of VXC-72 carbon. The

126

corresponding particle sizes of the catalysts have been determined by measurement of more than

127

100 NPs. The measured mean particle diameters obtained are ca. 3.9 to 4.0 nm for the Cu@Pd/C

128

catalysts and 3.8 nm for the as-prepared Pd/C catalyst. Hence, a comparison of catalytic

129

performance in this work is appropriate.

130

3.2 Crystalline and Electronic Structure

131

Figure 2 presents the XRD patterns of the Cu@Pd/C and Pd/C catalysts. Figure 2A shows that

132

the first peak located at ca. 25° is attributable to VXC-72 carbon .24 All the catalysts also exhibit

133

five characteristic peaks indexed with planes (111), (200), (220), (311), and (222), demonstrating

134

that all the bimetallic catalysts are single-phase disordered Pd face-centered-cubic structures

135

(JCPDF #46-1043). Compared with the reflections in the as-prepared Pd/C, the diffraction peaks

136

of the Cu@Pd/C catalysts shift to higher 2θ with increase in Cu content. At a slower scan rate, the

137

Pd (111) and (220) facets exhibit the same trends, as shown in Figures 2B and 2C. The lattice

138

spacings for the catalysts, determined by fitting the XRD intensity profiles, are provided in Table

139

1 .24,25 The lattice spacings obtained for the (111) and (220) planes decrease with increased Cu

140

content relative to that of bulk Pd (JCPDF #46-1043). The observed lattice contraction can be

141

ascribed to Vegard’s law, which involves metals with different lattice spacings.25 Specifically, the

142

addition of Cu, whose lattice spacings for the (111) and (220) facets are 2.09 and 1.28 Å (JCPDF

143

#04-0836), which are lower than that of the corresponding Pd facets 2.25 and 1.38 Å (JCPDF #46-

144

1043), respectively, would markedly decrease the mean lattice distance of Pd.

145

To further understand the lattice contraction, lattice strain was calculated using the following

146

formula: ε = (a-a0)/a0. For the (111) and (220) facets, with a0 = 2.25 and 1.38 Å, respectively,7,8

ACS Paragon Plus Environment

7

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 26

147

the lattice strains are provided in Table 1. The average lattice strains for the (111) and (220) facets

148

of the Cu@Pd/C catalysts have values ranging from 0% to –2.67% and –0.72% to –3.62%,

149

respectively. Obviously, the addition of Cu accounts for the lattice strain in the Cu@Pd/C. This

150

finding is similar to that reported in the investigation of Cu@Pt catalysts by Strasser et al. .7

151 152

Figure 2. XRD patterns of the catalysts at a scan rate of 2° min-1 (A) and at a lower scan rate of 1°

153

min-1 of the Pd (111) (B) and Pd (220) facets (C). Traces a, b, c, d, e, and f are corresponding to

154

the Pd/C, Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and Cu44@Pd56/C catalysts.

155

ACS Paragon Plus Environment

8

Page 9 of 26

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

156

Table 1. Lattice distance, lattice strain (ε1 and ε2), and εd results of the catalysts. ε2 / % b

εd / eV

0

–0.72

2.79

1.37

–0.89

–1.72

2.92

2.22

1.36

–1.33

–1.45

3.10

Cu27@Pd73/C

2.20

1.35

–2.22

–2.17

3.26

Cu35@Pd65/C

2.20

1.34

–2.22

–2.90

3.53

Cu44@Pd56/C

2.19

1.33

–2.67

–3.62

3.84

Sample

Pd (111)/Å

Pd (220)/Å

Pd/C

2.25

1.37

Cu15@Pd85/C

2.23

Cu21@Pd79/C

ε1 / % a

157

a

a0 = 2.25 Å for the Pd (111) planes (JCPDF #46-1043).

158

b

a0 = 1.38 Å for the Pd (220) planes (JCPDF #46-1043).

159 160

Historically, the variation of the lattice distance has been considered to be related to the

161

electronic structure of the nanocrystal.6,7,9,26,27 Such a determination, generally from XRD

162

measurements, necessitates the exploration of the electronic properties of the Cu@Pd/C catalysts.

163

Hence, the samples were analyzed using XPS. Figure 3A shows XPS Pd 3d spectra for the Pd 3d5/2

164

peak with BE of 335.29, 335.37, 335.44, 335.60, 335.68, and 335.80 eV, as well as for the Pd 3d3/2

165

peak with BE of 340.59, 340.67, 340.83, 340.91, 341.02, and 341.10 eV for as-prepared Pd/C,

166

Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and Cu44@Pd56/C, respectively.18,23

167

The BEs of the Pd 3d peaks positively shift with increase in Cu content, indicating electron lose

168

for surface Pd atoms. Such a result is consistent with the finding by Hu et al..23 No oxide species is

169

formed in the catalysts, and no broad trend of the reflection peaks can be observed. Thus, the

170

positive surface core-level shift of the Pd 3d can be attributed to the electron transfer from the

171

surface Pd atoms to their adjacent sub-layer atoms. This latter finding suggests an enhanced Pd–Pd

172

strength and a decreased bond length, which might explain the lattice contraction.28

ACS Paragon Plus Environment

9

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 26

173

174 175

Figure 3. XPS spectra in the binding energy ranges for Pd 3d (A), and XPS spectra in the valence

176

band region for the catalysts (B). Traces a, b, c, d, e, and f are corresponding to the Pd/C,

177

Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and Cu44@Pd56/C catalysts.

178

Pd valence band spectra were also acquired in the present study, as shown in Figure 3B. The

179

spectra of the Cu@Pd/C catalysts show broader d-bands with increase in Cu content. The d-band

180

center (εd) is found to shift downward relative to the Fermi level, as shown in Table 1. Obviously,

181

the broadening of the d-band may be related to the compressive strain associated with lattice

182

contraction, which means increased electronic state overlap between the metal atoms. To keep the

183

d-occupancy constant, a downward shift of the d-band center occurs.9,29 Such an observation is

ACS Paragon Plus Environment

10

Page 11 of 26

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

184

also reported by Hu.23 The positive core-level shift of the 3d spectrum and the downshift of the d-

185

band center occur together owing to electron loss.23,30,31 Generally, the downshift of the d-band

186

center is significant because it suggests a weaker adsorption strength between the catalysis surface

187

and the adsorption species, which is a result of decreased electron back-donation from the metallic

188

surface to the anti-bonding level of adsorbed molecule.7,10,32,33 The increase in lattice strain and the

189

downshift of the d-band center explain the performance improvement for formic acid oxidation.

190

The deduction in this study is based on substantial syntheses efforts to form the Pd-based catalysts,

191

as well as important characterization studies using XRD and/or XPS.3,9,22,34–38

192

3.3 Electrocatalytic activity for formic acid oxidation

193

To verify the conclusions in this study, the electrochemical performance of the Cu@Pd/C was

194

tested by depositing the catalysts onto a GC electrode. Figure 4 shows the CVs of the as-prepared

195

Pd/C and Cu@Pd/C catalysts at a scan rate of 20 mV s-1 in 0.5 M H2SO4 solution at room

196

temperature. All the catalysts possess typical H adsorption/desorption regions at the lower

197

potential, which can be assigned to the behavior of Pd. The ECSA, acquired by integration of the

198

H desorption region, was obtained and is provided in Table 2. The ECSAs of the catalysts exhibit

199

an optimized maximum value of 12.33 m2 g-1 for the Cu27@Pd73/C. A similar trend is also

200

observed for the oxide reduction peaks at ca. 0.5 V. Identical peaks of the Cu redox are indicated

201

by arrows in the figure, revealing the presence of Cu atoms in the catalysts.19,22

202

ACS Paragon Plus Environment

11

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 26

203 204

Figure 4. CVs of the catalysts in 0.5 M H2SO4 solution at a scan rate of 20 mV s-1. Traces a, b, c,

205

d, e, and f are corresponding to the Pd/C, Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C,

206

Cu35@Pd65/C, and Cu44@Pd56/C catalysts.

207 208

Table 2. Electrochemical characteristics of the catalysts. Sample

ECSA / m2g-1 a

ECSA / m2g-1 b

[email protected] / mA cm-2 c

Pd/C

7.16

6.31

49.85

Cu15@Pd85/C

8.28

7.98

53.27

Cu21@Pd79/C

12.04

12.62

61.26

Cu27@Pd73/C

12.33

13.92

82.56

Cu35@Pd65/C

7.73

7.40

39.96

Cu44@Pd56/C

4.34

6.99

22.09

209

a

Results obtained from CVs.

210

b

Results obtained from CO-stripping voltammograms.

211

c

Results obtained with the CVs measured in electrolyte that contains formic acid.

212

ACS Paragon Plus Environment

12

Page 13 of 26

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

213 214

Figure 5. CO-stripping voltammograms of the catalysts in 0.5 M H2SO4 solution at a scan rate of

215

20 mV s-1.

ACS Paragon Plus Environment

13

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 26

216

To gain more insight into the electrochemical performance of the catalysts, CO-stripping

217

voltammograms of the catalysts were obtained, as shown in Figure 5. The figure shows that the

218

oxidation peaks of the adsorbed CO on the Cu21@Pd79/C and Cu27@Pd73/C have the maximum

219

current density, indicating that the two catalysts might have the most catalytic activity sites for

220

adsorbing CO. The ECSAs of the catalysts are also calculated using the CO oxidation peaks and

221

are provided in Table 2. The results clearly demonstrate that Cu27@Pd73/C has the highest ECSA

222

among the catalysts, which qualitatively agree with the data obtained from CVs. More

223

interestingly, the Cu@Pd/C catalysts display improved CO tolerance, as indicated by the negative

224

shift of the onset oxidation potential associated with the vertical line in the figure. The downshift

225

of the d-band center would decrease the back-donation of d electrons to the 2π* level of the COad.

226

Hence, the adsorption strength between the metallic surface and COad onto Cu@Pd/C would be

227

much weaker than that for the as-prepared Pd/C.33,39 As a result, the COad on Cu@Pd/C with

228

higher Cu content can be more easily removed. Given that COad is inevitably formed during the

229

electrooxidation of formic acid, and certainly would poison the Pd-based catalysts, the increased

230

resistance to the COad poisoning effect could be a meaningful avenue to improve

231

performance.14,15,17

232

Figure 6 is a comparison of formic acid oxidation on the Pd/C and Cu@Pd/C catalysts. All

233

measurements were conducted at 50 mV s-1 in 0.5 M H2SO4 + 0.5 M formic acid. Compared with

234

the as-prepared Pd/C, only one formic acid oxidation peak exists for each of the Cu@Pd/C

235

catalysts, suggesting different oxidation mechanisms, as shown in Figure 6A. The current density

236

for formic acid oxidation on the catalysts varies with atomic compositions of the catalysts in the

237

order Cu27@Pd73/C > Cu21@Pd79/C > Cu15@Pd85/C > Pd/C > Cu35@Pd65/C > Cu44@Pd56/C,

238

demonstrating enhanced formic acid catalytic activity attributable to a synergistic effect of copper

239

atoms.

ACS Paragon Plus Environment

14

Page 15 of 26

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

240 241

Figure 6. Cyclic voltammograms of the catalysts in 0.5 M H2SO4 + 0.5 M formic acid solution at

242

a scan rate of 50 mV s-1 (A) and a relationship between current density at 0.16 V (vs. SCE) and the

243

composition of the catalysts (B). Traces a, b, c, d, e, and f are corresponding to the Pd/C,

244

Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and Cu44@Pd56/C catalysts.

245

To further understand the formic acid oxidation mechanism on the Cu@Pd/C catalysts, the

246

current density of formic acid oxidation at 0.16 V (vs. SCE) is plotted as a function of the

247

compositions of the catalysts, as shown in Figure 6B. A peak-like profile is revealed, as frequently

248

occurs in catalysis sciences. Generally, the shape of this plot could be a result of several

249

phenomena, such as increased particle size and electronic structure change.30,40,41 For the

250

Cu@Pd/C system, the performance variation is clearly dependent on the lattice strain associated

251

with lattice contraction and also the formation of core–shell structure. The key ingredient to the

252

peak–like profile could be the tuning of the reactive species adsorption change. In another study

253

involving chronoamperometry in 0.5 M H2SO4 + 0.5 M HCOOH, the electrodes were held at 0.16

254

V (vs. SCE) for 1 h to obtain i–t curves, as shown in Figure 7. The Cu27@Pd73/C catalyst exhibits

255

the highest initial current density compared with the other catalysts: the initial current densities

256

decreased in the order Cu27@Pd73/C > Cu21@Pd79/C > Cu15@Pd85/C > Cu35@Pd65/C > Pd/C >

ACS Paragon Plus Environment

15

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 16 of 26

257

Cu44@Pd56/C, a slight difference from that shown in Figure 6A at 0.16 V. Despite the enhanced

258

catalytic activity for formic acid oxidation of the Cu15@Pd85/C, Cu21@Pd79/C, and Cu27@Pd73/C

259

catalysts, compared with that of the as-prepared Pd/C, no improved stability toward formic acid

260

oxidation was observed. By contrast, the decreased catalytic activity for the Cu35@Pd65/C and

261

Cu44@Pd56/C catalysts indicates the greater stability of the latter than that of the as-prepared Pd/C.

262 263

Figure 7. I–t curves of the catalysts in 0.5 M H2SO4 + 0.5 M formic acid solution at a given

264

potential of 0.16 V (vs. SCE). Traces a, b, c, d, e, and f are corresponding to the Pd/C,

265

Cu15@Pd85/C, Cu21@Pd79/C, Cu27@Pd73/C, Cu35@Pd65/C, and Cu44@Pd56/C catalysts.

266

The downshift of the d-band center promotes improved performance of the Pd-based catalysts

267

for formic acid oxidation.9 Although Cu27@Pd73/C can produce the highest catalytic activity, a

268

compromise between the oxidation activity of formic acid and the poisoning effect of the catalysts

269

(by COad and/or (COOH)ad) is important.23,30,40,42 By contrast, the downshift of the d-band relative

270

to the Fermi level implies weaker adsorption strength of the formic acid molecule on the metal

271

surface, which makes breaking bonds harder, resulting in a diminution of the catalytic activity

272

toward the formic acid.10,41 The downshift of the d-band center would also lead to weaker bonds

ACS Paragon Plus Environment

16

Page 17 of 26

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

273

between the metal surface and the poisoning species that is produced during the oxidation of

274

formic acid, indicating that the poisoning species can be more easily removed. Surface active sites

275

occupied by the poisoning species may be freed more quickly in this latter case.

276

As reported, the galvanic replacement reaction is an efficient method to synthesize core-shell

277

metallic nanoparticles with improved catalytic activity.8 The gradually enhanced electrochemical

278

activity on the Cu15@Pd85/C, Cu21@Pd79/C and Cu35@Pd65/C catalysts may be due to the

279

formation of core-shell structure, which would result in an increased ratio of coordination

280

unsaturated Pd. However, with the further increase in Cu content within the catalysts, the thickness

281

of the Pd shell will gradually decrease, and the lattice strain in the nanocrystals is raised, thus

282

leading to a down-shift of the d-band center. Hence, the adsorption strength between the adsorbate

283

and the active site on the catalysts would be decreased with the increase in Cu content. Therefore,

284

an excessive increase of Cu would lead to a decrease of the catalytic activity and to an

285

improvement in durability, as verified by using the Cu35@Pd65/C, and Cu44@Pd56/C catalysts.

286

4. Conclusions

287

The Cu@Pd/C catalysts with different atomic compositions have been successfully prepared.

288

The lattice strain–induced electronic structure variation and modulation of catalytic performance

289

toward the formic acid oxidation have been clarified. Lattice contraction increases with Cu content

290

and leads to lattice strain associated with surface Pd atoms, which, in turn, directly affects the

291

electronic structure of the catalysts. Specifically, increased lattice strain leads to a downshift of the

292

d-band center. The catalytic performance of the catalysts exhibits a peak-like profile when plotted

293

versus the composition of the catalysts, with the maximum value found for Cu27@Pd73/C.

294

Catalytic stabilities are improved for the Cu35@Pd65/C and Cu44@Pd56/C catalysts.

ACS Paragon Plus Environment

17

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

295

AUTHOR INFORMATION

296

Corresponding Author

297

Tel. & Fax: +86 21 20321112

298

E-mail: [email protected]

299

Notes

300

The authors declare no competing financial interest.

301

ACKNOWLEDGMENT

Page 18 of 26

302

We are grateful for the financial support from the National Basic Research Program of China

303

(973 Program) (2012CB932800), the National Natural Science Foundation of China (21073219),

304

the Shanghai Science and Technology Committee (11DZ1200400), the Knowledge Innovation

305

Engineering of the CAS (12406, 124091231), and the Scientific and Technological Innovation

306

Fund for Graduate Students of the CAS. DLA also thanks the U.S. NSF under Cooperative

307

Agreement number HRD-0833180 for supporting this work.

308 309

REFERENCES

310

(1) Zhang, S.; Shao, Y.; Yin, G.; Lin, Y. Electrostatic Self–Assembly of a Pt–Around–Au

311

Nanocomposite with High Activity Towards Formic Acid Oxidation. Angew. Chem. Int. Ed.

312

2010, 49, 2211-2214.

313 314

(2) Yu, X.; Pickup, P.G. Recent Advances in Direct Formic Acid Fuel Cells (DFAFC). J Power Sources, 2008, 182, 124-132.

ACS Paragon Plus Environment

18

Page 19 of 26

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

315

(3) Zhang, L.; Wan, L.; Ma, Y.; Chen, Y.; Zhou, Y.; Tang, Y.; Lu, T. Crystalline Palladium–

316

Cobalt Alloy Nanoassemblies with Enhanced Activity and Stability for the Formic Acid

317

Oxidation Reaction. Appl. Catal. B, 2013, 138-139, 229-235.

318

(4) Du, C.; Chen, M.; Wang, W.; Yin, G.; Shi, P. Electrodeposited PdNi2 Alloy with Novelly

319

Enhanced Catalytic Activity for Electrooxidation of Formic Acid. Electrochem. Commun.,

320

2010, 12, 843-846.

321 322

(5) Tao, F. Synthesis, Catalysis, Surface Chemistry and Structure of Bimetallic Nanocatalysts. Chem. Soc. Rev., 2012, 41, 7977-7979.

323

(6) Kitchin, J. R.; Nørskov, J. K.; Barteau, M. A.; Chen, J. G. Role of Strain and Ligand Effects in

324

the Modification of the Electronic and Chemical Properties of Bimetallic Surfaces. Phys. Rev.

325

Lett. 2004, 93, 156801.

326

(7) Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.; Liu, Z.; Kaya, S.; Nordlund,

327

D.; Ogasawara, H.; et al. Lattice–Strain Control of the Activity in Dealloyed Core–Shell Fuel

328

Cell Catalysts. Nat. Chem., 2010, 2, 454-460.

329

(8) Wang, J. X.; Ma, C.; Choi, Y.; Su, D.; Zhu, Y.; Liu, P.; Si, R.; Vukmirovic, M. B.; Zhang, Y.;

330

Adzic, R. R. Kirkendall Effect and Lattice Contraction in Nanocatalysts: A New Strategy to

331

Enhance Sustainable Activity. J. Am. Chem. Soc., 2011, 133, 13551-13557.

332 333 334 335 336 337

(9) Kibler, L. A.; El-Aziz, A. M.; Hoyer, R.; Kolb, D. M. Tuning Reaction Rates by Lateral Strain in a Palladium Monolayer. Angew. Chem. Int. Ed., 2005 44, 2080-2084. (10)

Norskov, J. K.; Bligaard, T.; Rossmeisl, J.; Christensen, C. H. Towards the Computational

Design of Solid Catalysts. Nat. Chem., 2009, 1, 37-46. (11)

Kibler, L. A.; El-Aziz, A. M.; Kolb, D. M. Electrochemical Behaviour of Pseudomorphic

Overlayers: Pd on Au(111). J. Mol. Catal. A, 2003, 199, 57-63.

ACS Paragon Plus Environment

19

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

338

(12)

Page 20 of 26

Wang, S.; Wang, X.; Jiang, S. P. Controllable Self–Assembly of Pd Nanowire Networks as

339

Highly Active Electrocatalysts for Direct Formic Acid Fuel Cells. Nanotechnology, 2008, 19,

340

455602.

341 342 343 344 345

(13)

Rees, N. V.; Compton, R. G. Sustainable Energy: A Review of Formic Acid

Electrochemical Fuel Cells. J Solid State Electrochem, 2011, 15, 2095-2100. (14)

Wang, J.-Y.; Zhang, H.-X.; Jiang, K.; Cai, W.-B. From HCOOH to CO at Pd Electrodes: A

Surface–Enhanced Infrared Spectroscopy Study. J. Am. Chem. Soc., 2011, 133, 14876-14879. (15)

Zhang, H.-X.; Wang, S.-H.; Jiang, K.; André, T.; Cai, W.-B. In Situ Spectroscopic

346

Investigation of CO Accumulation and Poisoning on Pd Black Surfaces in Concentrated

347

HCOOH. J Power Sources, 2012, 199, 165-169.

348

(16)

Yu, X.; Pickup, P. G. Deactivation/Reactivation of A Pd/C Catalyst in A Direct Formic

349

Acid Fuel Cell (DFAFC): Use of Array Membrane Electrode Assemblies. J Power Sources,

350

2009, 187, 493-499.

351 352 353

(17)

Yu, X.; Pickup, P.G. Mechanism Study of the Deactivation of Caron Supported Pd During

Formic Acid Oxidation. Electrochem. Commun. , 2009, 11, 2012-2014. (18)

Ren, M.; Kang, Y.; He, W.; Zou, Z.; Xue, X.; Akins, D. L.; Yang, H.; Feng, S. Origin of

354

Performance Degradation of Palladium–Based Direct Formic Acid Fuel Cells. Appl. Catal. B,

355

2011, 104, 49-53.

356 357 358 359 360 361

(19)

Xu, C.; Liu, A.; Qiu, H.; Liu, Y. Nanoporous PdCu Alloy with Enhanced Electrocatalytic

Performance. Electrochem. Commun. , 2011, 13, 766-769. (20)

Xu, C.; Liu, Y.; Wang, J.; Geng, H.; Qiu, H. Nanoporous PdCu Alloy for Formic Acid

Electro–Oxidation. J Power Sources, 2012, 199, 124-131. (21)

Dai, L.; Zou, S. Enhanced Formic Acid Oxidation on Cu–Pd Nanoparticles. J Power

Sources, 2011, 196, 9369-9372.

ACS Paragon Plus Environment

20

Page 21 of 26

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

362

The Journal of Physical Chemistry

(22)

Lu, L.; Shen, L.; Shi, Y.; Chen, T.; Jiang, G.; Ge, C.; Tang, Y.; Chen, Y.; Lu, T. New

363

Insights into Enhanced Electrocatalytic Performance of Carbon Supported Pd–Cu Catalyst for

364

Formic Acid Oxidation. Electrochim. Acta, 2012, 85, 187-194.

365 366 367

(23)

Hu, S.; Scudiero, L.; Ha, S. Electronic Effect on Oxidation of Formic Acid on Supported

Pd–Cu Bimetallic Surface. Electrochim. Acta, 2012, 83, 354-358. (24)

Wang, W.; Huang, Q.; Liu, J.; Zou, Z.; Zhao, M.; Vogel, W.; Yang, H. Surface and

368

Structure Characteristics of Caron–Supported Pd3Pt1 Bimetallic Nanoparticles for Methanol–

369

Tolerant Oxygen Reduction Reaction. J. Catal., 2009, 266, 156-163.

370

(25)

Liu, J.; Cao, J.; Huang, Q.; Li, X.; Zou, Z.; Yang, H. Methanol Oxidation on Carbon–

371

Supported Pt–Ru–Ni Ternary Nanoparticle Electrocatalysts. J Power Sources, 2008, 175, 159-

372

165.

373 374 375

(26)

Rodriguez, J. A.; Goodman, D. W. The Nature of the Metal–Metal Bond in Bimetallic

Surfaces. Science, 1992, 257, 897-903. (27)

Yoo, S. J.; Kim, S. K.; Jeon, T.Y.; Hwang, S. J.; Lee, J. G.; Lee, S. C.; Lee, K.S.; Cho, Y.

376

H.; Sung, Y. E.; Lim, T. H. Enhanced Stability and Activity of Pt–Y Alloy Catalysts for

377

Electrocatalytic Oxygen Reduction. Chem. Commun. (Camb), 2011, 47, 11414-11416.

378 379 380 381 382

(28)

Richter, B.; Kuhlenbeck, H.; Freund, H. J.; Bagus, P. Cluster Core–Level Binding–Energy

Shifts: The Role of Lattice Strain. Phys. Rev. Lett. , 2004, 93, 026805. (29)

Bligaard, T.; Nørskov, J. K. Ligand Effects in Heterogeneous Catalysis and

Electrochemistry. Electrochim. Acta, 2007, 52, 5512-5516. (30)

Zhou, W. P.; Lewera, A.; Larsen, R.; Masel, R. I.; Bagus, P. S.; Wieckowski, A. Size

383

Effects in Electronic and Catalytic Properties of Unsupported Palladium Nanoparticls in

384

Electrooxidation of Formic Acid. J. Phys. Chem. B, 2006, 110, 13393-13398.

ACS Paragon Plus Environment

21

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

385 386 387 388 389

(31)

Page 22 of 26

Corcoran, C. J.; Tavassol, H.; Rigsby, M. A.; Bagus, P. S.; Wieckowski, A. Application of

XPS to Study Electrocatalysts for Fuel Cell. J Power Sources, 2010, 195, 7856-7879. (32)

Suo, Y.; Zhuang, L.; Lu, J. First–Principles Considerations in the Design of Pd–Alloy

Catalysts for Oxygen Reduction. Angew. Chem. Int. Ed. , 2007, 46, 2862-2864. (33)

Tong, Y. Y.; Kim, H. S.; Babu, P. K.; Waszczuk, P.; Wieckowski, A.; Oldfield, E. An

390

NMR Investigation of CO Tolerance in A Pt/Ru Fuel Cell Catalyst. J. Am. Chem. Soc., 2002,

391

124, 468-473.

392 393 394 395 396

(34)

Gao, Y.; Wang, G.; Wu, B.; Deng, C.; Gao, Y. Highly Active Carbon–Supported PdNi

Catalyst for Formic Acid Electrooxidation. J. Appl. Electrochem. , 2011, 41, 1-6. (35)

Wang, R.; Liao, S.; Ji, S. High Performance Pd–Based Catalysts for Oxidation of Formic

Acid. J Power Sources, 2008, 180, 205-208. (36)

Feng, L.G.; Cui, Z. M.; Yan, L. A.; Xing, W.; Liu, C. P. The Enhancement Effect of MnOx

397

on Pd/C Catalyst for the Electrooxidation of Formic Acid. Electrochim. Acta, 2011, 56, 2051-

398

2056.

399

(37)

Yin, M.; Li, Q.; Jensen, J. O.; Huang, Y.; Cleemann, L. N.; Bjerrum, N. J.; Xing, W.

400

Tungsten Carbide Promoted Pd and Pd–Co Electrocatalysts for Formic Acid Electrooxidation.

401

J Power Sources, 2012, 219, 106-111.

402

(38)

Sun, H.; Xu, J.; Fu, G.; Mao, X.; Zhang, L.; Chen, Y.; Zhou, Y.; Lu, T.; Tang, Y.

403

Preparation of Highly Dispersed Palladium–Phosphorus Nanoparticles and Its Electrocatalytic

404

Performance for Formic Acid Electrooxdation. Electrochim. Acta, 2012, 59, 279-283.

405 406 407 408

(39)

Hammer, B.; Morikawa, Y.; Nørskov, J. K. CO Chemisorption at Metal Surfaces and

Overlayers. Phys. Rev. Lett. , 1996, 76, 2141-2144. (40)

Zhou, W.; Lee, J.Y. Particle Size Effects in Pd–Catalyzed Electrooxidation of Formic Acid.

J. Phys. Chem. C, 2008, 112, 3789-3793.

ACS Paragon Plus Environment

22

Page 23 of 26

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

409

The Journal of Physical Chemistry

(41)

Tedsree, K.; Li, T.; Jones, S.; Chan, C.W.A.; Yu, K.M.K.; Bagot, P.A.J.; Marquis, E.A.;

410

Smith, G.D.W.; Tsang, S.C.E. Hydrogen Production From Formic Acid Decomposition at

411

Room Temperature Using A Ag–Pd Core–Shell Nanocatalyst. Nat. Nano. , 2011, 6, 302-307.

412

(42)

Yang, G.; Chen, Y.; Zhou, Y.; Tang, Y.; Lu, T. Preparation of Carbon Supported Pd–P

413

Catalyst with High Content of Element Phosphorus and Its Electrocatalytic Performance for

414

Formic Acid Oxidation. Electrochem. Commun., 2010, 12, 492-495.

415 416 417 418 419 420 421 422 423 424 425 426

ACS Paragon Plus Environment

23

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

427

Page 24 of 26

Table of Content

428

ACS Paragon Plus Environment

24

Page 25 of 26

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

Revised Figures Figure 4

Figure 4

Figure 6B

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

TOC

ACS Paragon Plus Environment

Page 26 of 26