47
1 Submitted, accepted and published in 1 Applied Catalysis B: environmental 163 83-95 (2015); ISSN 0926-3373, 2 http://dx.doi.org/10.1016/j.apcatb.2014.07.031 3 4 Pd catalysts supported onto nanostructured carbon materials for CO 2 5 valorization by electrochemical reduction 6 7 S. Pérez-Rodríguez 1,2 , N. Rillo 1 , M.J. Lázaro 1 , E. Pastor 2 8 9 10 1 Instituto de Carboquímica (CSIC), Miguel Luesma Castán 4, 50018 Zaragoza, Spain. 11 2 Universidad de La Laguna, Dpto. Química-Física, Avda. Astrofísico Francisco 12 Sánchez s/n, 38071 La Laguna (Tenerife), Spain. 13 14 * Corresponding author: Tel.: +34 976733977; fax: +34 976733318. E-mail address: 15 [email protected] (M.J. Lázaro). 16

Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

1

Submitted, accepted and published in 1

Applied Catalysis B: environmental 163 83-95 (2015); ISSN 0926-3373, 2

http://dx.doi.org/10.1016/j.apcatb.2014.07.031 3

4

Pd catalysts supported onto nanostructured carbon materials for CO2 5

valorization by electrochemical reduction 6

7

S. Pérez-Rodríguez1,2

, N. Rillo1, M.J. Lázaro

1, E. Pastor

2 8

9

10

1Instituto de Carboquímica (CSIC), Miguel Luesma Castán 4, 50018 Zaragoza, Spain. 11

2Universidad de La Laguna, Dpto. Química-Física, Avda. Astrofísico Francisco 12

Sánchez s/n, 38071 La Laguna (Tenerife), Spain. 13

14

* Corresponding author: Tel.: +34 976733977; fax: +34 976733318. E-mail address: 15

[email protected] (M.J. Lázaro). 16

Page 2: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

2

Abstract 17

Pd nanoparticles have been supported onto different novel carbon materials to be 18

used in the electroreduction of CO2 to high-added value products. Carbon nanocoils 19

(CNCs), carbon nanofibers (CNFs) and ordered mesoporous carbon materials (OMCs) 20

have been studied as support of the catalysts. In addition, Pd catalyst supported onto 21

Vulcan XC-72R has been synthesized in order to establish a comparison with the 22

commercial support. In this way, the influence of the carbon material on the 23

physicochemical and electrochemical properties of the catalysts for the CO2 24

electroreduction process can be analyzed. Supports presented different morphologies 25

and structures. CNFs and CNCs exhibited a crystalline structure with well-aligned 26

graphitic layers and OMCs a hexagonal ordered structure composed of not crystalline 27

carbon. Finally, Vulcan presented an intermediate structure between amorphous and 28

graphitic. These differences do not affect significantly the average Pd crystallite size, 29

although a different metal dispersion was found depending on the carbon material. On 30

the other hand, cyclic voltammetry studies showed that CO2 was effectively reduced to 31

other species at the surface of Pd/C catalysts. Additionally, it was proved that these 32

species were adsorbed onto Pd at -1.0 V vs Ag/AgCl. Finally, a different catalytic 33

activity towards the CO2 reduction reaction was observed for the different 34

electrocatalysts, indicating an influence of the carbon support. 35

36

Keywords: Carbon nanofibers, carbon nanocoils, ordered mesoporous carbon materials, 37

CO2 electroreduction, palladium. 38

39

Page 3: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

3

1. Introduction 40

Nowadays, CO2 is the main responsible of global warming, and therefore, 41

several CO2 mitigation strategies are being studying. Among them, the CO2 conversion 42

represents a promising alternative for the reutilization of this gas, with the aim of 43

producing useful products, with important applications in chemical industry and for 44

power generation. Different methods have been carried out to convert CO2, such as 45

chemical, thermochemical, photochemical, electrochemical and biochemical 46

procedures. The electrochemical route shows a wide selectivity towards valuable 47

products (methane, methanol, monoxide carbon, acid formic, etc) [1]. 48

Various metallic electrodes have been studied for the CO2 electrochemical 49

reduction in aqueous solutions, since the catalytic activity and the product distribution 50

depends strongly on the used material and the surface metallic arrangement [2-6]. 51

Copper has attracted special attention [1, 2, 7-11], due to its exceptional selectivity for 52

CO2 conversion to hydrocarbons and alcohols at significant current densities. On the 53

other hand, metals of the group VIII elements, such as Pt and Pd, have been also studied 54

in detail by several authors [12-24], since these electrodes present low hydrogen 55

overpotentials and easily adsorb hydrogen. Adsorbed hydrogen favor the adsorption of 56

species derived from CO2 on the electrode surface, since the formation of "reduced 57

CO2" adsorbates requires interaction between CO2 molecules and adsorbed hydrogen 58

[12, 13, 16-18]. These adsorbates are expected to play an important role in the reduction 59

pathway of the process, since adsorbed species may behave as both reactive 60

intermediate and poison. Beden et al. revealed by means of infrared spectroscopy, that 61

CO2 is reduced to CO, which is strongly adsorbed onto Pt, inhibiting further reduction 62

of CO2 and, therefore, obtaining H2 (formed by water reduction) as main product [20]. 63

Same behavior was found for the rest of metals of the group VIII elements (Ni, Fe, etc.), 64

Page 4: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

4

excepting Pd [3, 25, 26]. On this last metal, CO and formic acid have been reported as 65

main products, and even hydrocarbons have been obtained at low current efficiencies 66

[12, 15]. Taguchi et al. attributed this different behavior to the higher adsorption affinity 67

of "reduced CO2" on Pt that on Pd and suggested that "reduced CO2" would be behaving 68

as a catalytic self-poisoning for the electrochemical reduction of CO2 on Pt, but it partly 69

becoming an intermediate on Pd [14]. On the other hand, this different activity can be 70

also explained by the differences in the process of hydrogen sorption (adsorption and 71

absorption)/desorption. It is widely known that the bulk absorption of hydrogen in Pt is 72

insignificant in comparison to Pd, under the same conditions [27]. Therefore, a 73

substantial amount of the hydrogen obtained on the Pd electrodes would be bulk 74

absorbed by the metal, lowering the rate of detected hydrogen and changing the 75

adsorption affinity of the species derived from the CO2 reduction reaction. 76

Despite all this extended research of the electrochemical reduction of CO2 on 77

different metallic electrodes, the involved mechanism and the implied intermediates of 78

the reaction are not fully clear, however. Moreover, many limitations must be addressed 79

to make the process viable. One of the main problems is the low solubility of CO2 in 80

aqueous systems, at atmospheric pressure and room temperature. For this reason, the 81

electroreduction of CO2 has been also studied onto different metals at high pressure (30 82

– 60 atm) [8, 28, 29], low temperatures (0 – (-30) ºC) [4, 8-10, 30] and/or in aprotic 83

solvents (methanol, acetonitrile, dimethyl sulphoxide, propylene carbonate, etc.) [21, 84

31, 32], in order to avoid these mass transport limitations. Despite of obtaining 85

hydrocarbons and more complexes products, the current efficiencies remained being 86

low. In addition, a quick deactivation of the electrodes was observed due to the higher 87

current densities under these conditions [7, 11]. In order to partly resolve these issues, 88

the use of gas diffusion electrodes (GDEs) and/or catalysts supported on nanostructured 89

Page 5: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

5

carbon materials has been proposed over the last few decades [21, 33-40]. These porous 90

electrodes allow a good distribution of the reactives over the catalyst surface, obtaining 91

a large reaction area while provide low current density with high current efficiency for 92

the formation of products. Additionally, support nanopores help to confine CO2, 93

resulting in higher pressures at the electrocatalyst surface, which favor the reaction 94

kinetics. It has been found that using electrodes based on carbon materials the product 95

partial current densities are considerably higher than those obtained for the 96

corresponding bulk electrodes, and even the selectivity of the process can be modified. 97

Surprising a mixture of C1-C9 hydrocarbons and alcohols were found in different GDEs 98

based on noble metals supported on modified carbon cloths [36, 39]. Pérez-Rodríguez et 99

al. reported the formation of formic acid below -0.7 V vs. Ag/AgCl, at atmospheric 100

pressure and room temperature using iron-oxide electrodes supported on treated Vulcan 101

XC-72R [35], while bulk Fe electrodes produce H2 as the main product under the same 102

conditions [41]. Mahmood et al. employed a lead-impregnated GDE to reduce CO2 to 103

formic acid with a current efficiency of nearly 100 %, operating at -1.8 V vs. SCE [34]. 104

Most of these studies have been carried out using carbon blacks as support, 105

specifically Vulcan XC-72R, which is the most commonly carbon material used for 106

electrochemical applications, due to its low cost and its good compromise between 107

electrical conductivity and high specific surface area [42]. However, there are not many 108

studies about the electroreduction of CO2 on catalysts supported on different 109

nanostructured carbon materials, such as carbon nanofibers (CNFs), carbon nanocoils 110

(CNCs), ordered mesoporous carbon materials (OMCs) or carbon nanotubes (CNTs) 111

[33, 37, 38, 40]. These carbon materials present different properties, such as 112

crystallinity, morphology, porosity and surface chemistry than those of the commercial 113

material, Vulcan. Therefore, the mass transport, the conductivity and the 114

Page 6: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

6

electrochemical surface area of the electrocatalysts, as well as size, dispersion and 115

stability of the metallic nanoparticles could be affected. Accordingly the activity and 116

selectivity towards the production of hydrocarbonated compounds from CO2 would be 117

modified. Hence, the optimization of the carbon materials plays an important role on the 118

CO2 electroreduction research. 119

In this work, various Pd catalysts have been synthesized to be tested in the CO2 120

electroreduction process. Different novel nanostructured carbon materials, such as 121

CNCs, CNFs and OMCs, have been used as support of the catalysts. In addition, Vulcan 122

XC-72R was also tested in order to establish a comparison with the commercial support. 123

In this way, the influence of the carbon material on the physicochemical and 124

electrochemical properties of the catalysts for the CO2 electroreduction process can be 125

analyzed. In this context, to the best of our knowledge, this is the first time that CO2 126

reduction is carried out using different nanostructured carbon materials as supports, in 127

order to obtain a comparison of them and establish their influence on the activity of the 128

process. 129

2. Experimental 130

2.1 Synthesis of the nanostructured carbon materials 131

2.1.1 Preparation of Carbon Nanofibers (CNFs) 132

CNFs were grown by methane decomposition over a NiCuAl2O3 (molar ratio 133

78:6:16) catalyst. 134

The NiCuAl2O3 catalyst was prepared by coprecipitation of metal salts, 135

Ni(NO3)2·6H2O (98 wt. %, Alfa Aesar), Cu(NO3)2·3H2O (99 wt. %, Panreac) and 136

Al(NO3)3·9H2O (98 wt. %, Alfa Aesar), and calcination at 450 ºC for 4 h. The resultant 137

catalyst precursor was reduced under a H2 flow of 20 mLmin-1

at 550 ºC during 1h, 138

obtaining the active phase for the CNFs growing. 139

Page 7: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

7

The methane decomposition reaction was carried out in a fixed-bed vertical 140

reactor under a pure CH4 flow (20 mLmin-1

) at 700 ºC during 10 h. Further details can 141

be found elsewhere [43]. Other studies have shown that these conditions result in a high 142

carbon graphitization without catalyst deactivation [44, 45]. 143

Finally, CNFs were treated with concentrated nitric acid (65 wt. %, Panreac) at 144

room temperature for 2 h to remove the metal particles. Obtained CNFs were filtered, 145

washed with abundant deionized water and then dried overnight at 80 ºC. 146

2.1.2 Preparation of Carbon Nanocoils (CNCs) 147

CNCs were synthesized by the catalytic graphitization method, using 148

Ni(NO3)2·6H2O (98 wt. %, Alfa Aesar) and Co(NO3)2·6H2O (98 wt. %, Sigma–Aldrich) 149

metal nitrates as graphitization catalysts, and a mixture of resorcinol (> 98 wt. %, Safc) 150

and formaldehyde (37 wt. %, Sigma-Aldrich) as the carbon precursor. Moreover, silica 151

sol (Supelco) was used to generate mesoporosity. The synthesis involved the following 152

steps: 153

First, formaldehyde and silica sol were dissolved in deionized water. Then, the 154

nickel-cobalt salts and resorcinol were added under mechanical stirring. The molar ratio 155

H2O/Co nitrate/Ni nitrate/resorcinol/formaldehyde/silica used for the synthesis was 156

100:0.2:0.2:1:2:0.6 [46]. The obtained mixture was heated, first at 85 ºC for 3 h in a 157

closed system, and then at 108 ºC. Subsequently, it was carbonized under a N2 158

atmosphere at 900 ºC for 3 h, and the resultant material was washed with 1.1 M NaOH 159

(98 wt. %, Panreac) solution (an EtOH (96 % v/v, Panreac) - water mixture of 1:0.3 in 160

volume), in order to remove the silica particles. 161

Finally, CNCs were treated with concentrated nitric acid (65 wt. %, Panreac) at 162

room temperature for 2 h to remove the metal particles. Obtained CNCs were filtered, 163

thoroughly washed with deionized water and then dried overnight at 80 ºC [47]. 164

Page 8: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

8

2.1.3 Preparation of Ordered Mesoporous Carbons (OMCs) 165

OMCs were prepared by the nanocasting technique using a mesoporous silica 166

(SBA-15) as template and a furan resin/acetone as carbon precursor. The synthesis 167

involved the following steps: 168

First, the template was synthesized using Tetraethyl Orthosilicate (TEOS, 98 wt. 169

%, Aldrich) as silica precursor and Pluronic P123 copolymer (EO20PO70EO20, Aldrich) 170

as mesoporosity creator. With this purpose, the P123 surfactant was dissolved in a 1.8 171

M HCl (37 wt. %, Sigma-Aldrich) aqueous solution. The resultant mixture was 172

subjected to heat treatment at 50 ºC for 2 h under mechanical stirring. Then, TEOS 173

(with a mass relation TEOS/P123 = 20:4) was dropwise added and the solution was 174

heated 2 h more. The obtained gel was heated at 85 ºC during 24 h, followed by a 175

filtration and washing process. The sample was heated overnight at room temperature 176

and then at 108 ºC. Finally, the silica was calcinated under a N2 atmosphere at 500 ºC 177

for 5 h in order to eliminate the P123 surfactant and to obtain an ordered mesoporous 178

structure [48]. 179

After, OMCs were prepared from the silica templates. With this finality, the 180

calcinated silica was subjected to incipient-wetness impregnation with the carbon 181

precursor: a mixture of furan resin (furfuryl alcohol and formaldehyde, Huttenes-182

Albertus) and acetone (99.5 wt. %, Panreac), with a mass relation 5:1. The impregnated 183

silica was heated at 108 ºC for 24 h, and then was carbonizated under a N2 atmosphere 184

at 700 ºC for 3 h. Subsequently, the resulting material was washed with 1.1 M NaOH 185

(98 wt. %, Panreac) solution (an EtOH (96 % v/v, Panreac) - water mixture of 1:0.3 in 186

volume), in order to remove the silica particles. Obtained OMCs were filtered, 187

thoroughly washed with deionized water and then dried overnight at 80 ºC [49]. 188

2.2 Synthesis of the carbon-supported Pd electrocatalysts 189

Page 9: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

9

The carbon-supported Pd electrocatalysts were prepared by the method of 190

impregnation and reduction with sodium borohydride. The synthesized nanostructured 191

carbon materials (CNFs, CNCs and OMCs) were used as supports. Moreover, Pd 192

catalyst supported onto Vulcan XC-72R (Cabot Corporation), was also prepared in 193

order to establish a comparison with the commercial support. 194

The synthesis involved the following steps: first, a 3.0 mM metal precursor 195

Na2PdCl4, (98 wt. %, Sigma–Aldrich) solution was slowly added to a dispersion of the 196

different carbon materials in ultrapure water under sonication. Afterwards the pH of the 197

dispersion was adjusted to 5.0 using NaOH (98 wt. %, Panreac). Then, the metal was 198

reduced with a 48.0 mM sodium borohydride (99 wt. %, Sigma–Aldrich) solution, 199

keeping the temperature below 18 ºC and in the presence of sonication. Finally, 200

catalysts were filtered, washed with ultrapure water (Millipore Milli-Q system), and 201

then dried overnight at 70 ºC. 202

Appropriate amounts of the metal precursor were employed to obtain a 203

theoretical Pd loading of 20 wt. % on the different carbon materials. 204

2.3 Physicochemical characterization 205

Carbon materials obtained in the previously described procedures were 206

characterized by different physicochemical techniques like transmission electron 207

microscopy (TEM), X-ray diffraction (XRD) and N2 physisorption. On the other hand, 208

carbon-supported catalysts (Pd/C) were characterized by energy dispersive Xray 209

analysis (EDX), termogravimetric analysis (TGA), TEM and XRD. 210

Transmission electron microscope (TEM) micrographs for the carbon supports 211

and the Pd electrocatalysts were obtained using a JEOL-2000 FXII microscope 212

equipped with a LaB6 gun. The samples were grounded, dispersed in ethanol and a drop 213

of solution was then deposited on a copper grid. 214

Page 10: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

10

X-Ray diffraction (XRD) patterns of the materials were recorded using a Bruker 215

AXS D8 Advance diffractometer with a θ–θ configuration and using Cu Kα radiation ( 216

= 0.15406 nm). Scans were done at 1 ºmin-1

for 2Ɵ values from 10º to 80º, for Pd 217

catalysts or graphitic carbon materials (Vulcan, CNFs and CNCs) or at 0.24 ºmin-1

for 218

2Ɵ values from 0.5º to 4º for OMCs. The graphite interlayer spacing (d002) for 219

crystalline carbons and interplanar distance (d100) for OMCs were calculated from the 220

position of the (0 0 2) and (1 0 0) peaks, respectively, applying the Bragg’s equation 221

[50, 51]. Crystallite sizes were calculated from Scherrer’s equation on the (0 0 2) peak 222

for carbon supports (Lc) and on the (2 2 0) peak for Pd (with a K value of 0.89) [51, 223

52].The graphitization degree of supports (g) was calculated from Mering and Maire 224

equation (validity range: 0.3354 ≤ d002 ≤ 0.3440 nm) [53, 54]. Taking into account the 225

hexagonal structure of OMCs, the unit cell parameter or the distance between the carbon 226

cylinder centers (a) was obtained by using the formula 3

2 100da

[50]. 227

N2 adsorption-desorption isotherms of the carbon supports were measured at -228

196 ºC using a Micromeritics ASAP 2020. The total surface area was calculated from 229

BET (Brunauer, Emmett and Teller) equation and the total pore volume was determined 230

using the single point method at P/P0 = 0.99. Pore size distribution (PSD) curves were 231

obtained from the analysis of the desorption branch of the N2 isotherm using the BJH 232

(Barrett, Joyner and Halenda) method. 233

Metal content of the electrocatalysts were obtained by energy dispersive Xray 234

analysis and thermogravimetric analysis (TGA) in air. An EDX analyzer Röntec XFlash 235

Si(Li) coupled to a Hitachi S-3400N scanning electron microscope (SEM) was used. 236

TGA experiments were carried out on a thermogravimetric SETARAM Setsys 237

Evolution under air atmosphere. The temperature was varied between room temperature 238

and 950 ºC with a heating rate of 5 ºC min-1

. In order to obtain the metal content of the 239

Page 11: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

11

electrocatalysts, TGA analysis were also carried out for the carbon supports. In this way 240

the Pd load was calculated by the difference between the residual value of the catalysts 241

and the carbon materials, both results obtained by TGA under the same conditions 242

described above. 243

2.4. DEMS set-up 244

A scheme of the DEMS setup used for the in situ spectrometric and 245

electrochemical measurements is described in [55]. Briefly, the spectro-electrochemical 246

experiments were carried out in a conventional three electrodes electrochemical cell, 247

directly attached to the vacuum chamber of a mass spectrometer (Balzers Omnistar 248

quadrupole mass spectrometer), with a Faraday cup detector. The working electrode 249

consists of a glassy carbon disk (7.0 mm of exterior diameter, SIGRADUR ® G) with 250

an inner cavity of around 1.5 mm of diameter, where a PTFE membrane (Scimat) is 251

accurately placed allowing the detection of volatile and gaseous products and 252

intermediates generated in the electrochemical reactions with excellent sensibility. 253

Further details are given in [55]. 254

In order to carry out spectro-electrochemical measurements properly, the Pd 255

electrocatalysts were dispersed onto the total surface of the carbon disk. Therefore, the 256

designed cell allows to follow “in-situ” the electrochemical properties of electrodes and, 257

simultaneously, the gaseous species produced on the electroactive surface, providing 258

more information than a conventional electrochemical technique (i.e., cyclic 259

voltammetry or chronoamperometry). 260

2.5. Spectro-electrochemical measurements 261

Electrochemical measurements were carried out in a three electrodes 262

electrochemical cell at room temperature and atmospheric pressure. The potenciostat-263

galvanostat was an Autolab PGSTAT302 (Ecochemie). In the case of the spectro-264

Page 12: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

12

electrochemical measurements, the same cell was directly attached to a mass 265

spectrometer (section 2.4). In this way, the experimental set-up allows the simultaneous 266

acquisition of mass spectrometric cyclic voltammograms (MSCVs) for selected m/z 267

(mass to charge) ratios and conventional cyclic voltammograms (CVs). In this work, 268

DEMS experiments were carried out to follow the formation of molecular H2 (m/z = 2), 269

simultaneously produced by water reduction during the CO2 electrochemical reduction. 270

A high surface area carbon rod was used as a counter electrode, whereas an 271

Ag/AgCl/3M KCl electrode placed inside a Luggin capillary was employed as 272

reference. All potentials in the text are referred to this electrode. The working electrodes 273

were prepared depositing a thin-layer of the different electrocatalysts over the glassy 274

carbon disk (7.0 mm of diameter, SIGRADUR ® G) for the electrochemical 275

experiments. In the case of spectro-electrochemical measurements, it was used the 276

carbon disk with the Teflon membrane (7.0 mm of exterior diameter, SIGRADUR ® 277

G). Catalyst inks were prepared, in all the cases, by mixing 2 mg of catalyst, 15µL of 278

Nafion dispersion (5 wt. %, Aldrich) and 500 μL of ultrapure water (Millipore Milli-Q 279

system). The suspension was submitted under sonication for 40 min. An aliquot of 10 280

μL of the suspension was pipetted onto the carbon disk and dried at room temperature. 281

The working electrode was immersed into 0.1 M NaHCO3 (Merck, p.a.) 282

electrolyte solution. Ar (N50) was used to deaerate all solutions and CO (N47) was 283

employed for the adsorption experiences. In order to analyze the activity of catalysts 284

towards the reduction of CO2, previously to each measure, CO2 (99.99 %, Air Liquide) 285

was bubbled during 30 min, decreasing the pH value of the solution (8.4 in the presence 286

of Ar and 6.8 in the presence of CO2). 287

3. Results and discussion 288

3.1 Physicochemical characterization of the supports 289

Page 13: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

13

The different morphology of carbon materials was studied by TEM. Vulcan 290

images (Fig. 1.A and Fig. 1.B) showed that the commercial support is formed by an 291

aggregation of spherical carbon nanoparticles (primary particles) with range diameter 292

sizes of 15–80 nm. In addition, primary particles consist of small crystallites, which are 293

composed of several parallel graphene layers, arranged randomly around the particle 294

with a turbostratic structure (Fig. 1.B). These crystallites confer Vulcan a certain 295

crystallinity degree. On the other hand, OMCs exhibit a hexagonal ordered structure, 296

which consists of periodic amorphous carbon cylinders, of around 6 nm of diameter 297

(seen by TEM), with uniform mesopores between them (near 3-4 nm). Fig. 1.C displays 298

the parallel disposition of the pores and their hexagonal structure can be seen in Fig. 299

1.D. Finally, CNFs and CNCs show well-aligned graphitic layers (Fig. 1.F and Fig. 1.H, 300

respectively). CNFs appear as long carbon nanofilaments, with graphene layers oblique 301

to the growth axis or "fishbone" morphology (Fig. 1.E) and CNCs as spherical particles 302

composed of several long curved ribbons with a wide diameter distribution (Fig. 1.G). 303

Fig. 2.A shows the XRD patterns of all the carbon supports. All the samples, 304

with the exception of OMCs, show a peak around 2θ = 25º, which is associated to the 305

graphite (0 0 2) diffraction line. However, some differences between the carbon 306

materials are observed. An increase of the graphitization degree of the carbon supports 307

leads to an increase of the relative intensity and a narrower of the graphite (0 0 2) peak, 308

as well as a shift to more positive degrees. Therefore, the graphitization degree of the 309

materials increases in the order: Vulcan < CNC < CNF, being the commercial support 310

the less crystalline. These differences can be explained by the different structure of the 311

materials, as was previously seen by TEM. Vulcan has a structure mainly amorphous, 312

although with a certain crystallinity degree attributed to the turbostratic structure of the 313

primary particles, whereas CNFs and CNCs are more graphitic materials due to the 314

Page 14: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

14

regular-aligned graphitic layers. This observation is confirmed by calculating the 315

graphite interlayer distances (d002), the crystal domain sizes (Lc) and the graphitization 316

degree (g) from the position (2Ɵ) and full width at half maximum (FWHM) of (0 0 2) 317

peak for graphite through Bragg, Scherrer and Maire-Mering formulas, respectively. 318

High ordered structures present high values of Lc and g and low values of d002, close to 319

that of perfect graphite (0.3354 nm). According to the values in Tab. 1, CNFs are the 320

most graphitic material. 321

In addition, (1 0 0) and (1 0 1) reflections of the graphite, at 2θ = 42.4º and 2θ = 322

44.36º, are apparent for all the samples. However, in the case of Vulcan, these peaks are 323

hardly discerned, indicative of their more amorphous character. 324

In the case of the OMC sample, the graphite (0 0 2) diffraction line around 2θ = 325

25º is not perceived, confirming that these materials present an amorphous structure. 326

However, as can be seen in Fig. 2.B, OMCs exhibit a diffraction peak (1 0 0), indicating 327

the formation of a highly ordered hexagonal (p6m) mesoestructure [50, 56]. Through 328

Bragg equation, the interplanar distance (d100) was calculated from the position of the (1 329

0 0) diffraction peak. Taking into account the hexagonal structure of OMCs, the unit 330

cell parameter (a) was obtained. OMCs presented an interplanar distance of 8.02 nm and 331

a distance between carbon cylinders of 9.27 nm, in agreement with values of carbon 332

nanorods nm and mesoporous 3-4 nm) seen by TEM. Similar results have been 333

obtained by other authors [50, 56]. 334

Finally, it is remarkable that after the purification treatments of CNFs and 335

CNCs, the catalyst used in their preparation were practically removed as diffraction 336

peaks corresponding to Ni, Co or Cu were almost negligible. In accordance with these 337

results, TGA analysis showed that the ash content, composed of metal and mineral 338

residues was in the range of 4-5 % for both materials (Tab. 2.). It is important to note 339

Page 15: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

15

that this residual metal load is encapsulated in the carbon filaments or the nanocoils, and 340

therefore is not accessible for the electrochemical reactions. 341

Textural properties of the carbon materials obtained by N2-physisorption are 342

summarized in Tab. 2. As can be seen, the specific surface area increases in the order: 343

CNFs < CNCs < Vulcan < OMCs, while CNCs show the lowest total pore volume. It is 344

noticeable that OMCs present the highest pore area and volume, which is attributed to 345

their particular preparation method (nanocasting). N2 adsorption–desorption isotherms 346

of the carbons are given in Fig. 3.A. In all the samples, the obtained isotherms can be 347

classified as type IV (according to the IUPAC classification), as all isotherms present 348

hysteresis loops [57]. The hysteresis loops are associated with capillary condensation in 349

the mesopores, which can be caused by structural pores or spaces between carbon 350

filaments or particles. Different hysteresis loops were obtained for the carbon materials. 351

Vulcan and CNFs present H3 hysteresis (IUPAC), which is characteristic of slit-shaped 352

pores. On the other hand, CNCs show H2 hysteresis, which is distinctive of 353

interconnected pore networks with different shape and size, as the silica (Supelco) used 354

in the synthesis process of this material. Finally, the OMCs hysteresis loop can be 355

classified as type H4, which is characteristic of carbon materials with a size distribution 356

in the range of microporosity or narrow mesoporosity [57, 58]. Fig. 3.B shows the pore 357

size distribution (PSD) of the carbon materials. Vulcan and CNFs present wide PSD 358

with two peaks; the first one centred around 2.5 nm for the commercial support and 3.5 359

nm for the carbon nanofilaments, and followed by a second peak around 30 nm, for both 360

materials. Smaller mesopores are associated with the carbon surface roughness, whereas 361

larger ones (> 10 nm) are originated from the interstices between interlaced filaments or 362

carbon particles [59, 60]. On the other hand, CNCs and OMCs exhibit narrow PSD, 363

with only one peak around 3.0 nm. However, in the case of CNCs, when the PSD was 364

Page 16: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

16

obtained from the adsorption branch, two peaks centred at 2.5 and 15.0 nm were 365

developed. It could be associated to the presence of bottleneck pores, which would limit 366

N2 desorption and the available surface area. Additionally, Vulcan and CNFs presented 367

similar average pore sizes, being of 10 nm for Vulcan and 11 for CNF, while CNCs and 368

OMCs exhibited lower pore sizes. 369

3.2 Physicochemical characterization of the electrocatalysts 370

The catalysts metal loading was determined by EDX and TGA. Both 371

measurements confirmed that Pd content was similar to the nominal value of 20%, in all 372

the catalysts (Tab. 3). 373

TEM micrographs together with particle size distributions can be observed in 374

Fig. 4. In general, a regular distribution of the Pd particles over the carbonaceous 375

supports was observed, with the exception of CNC, where a not uniform dispersion was 376

obtained. Additionally, some particle aggregates can be observed in Fig. 4.D, although 377

this agglomeration is not extended over the whole sample. In the case of Pd/CNF, a 378

higher density of particles was detected on the edge of the carbon nanofilaments, as can 379

be seen in Fig. 4.C (enclosed figure in the particle size distribution). Narrow size 380

distributions were obtained for Pd catalysts supported on Vulcan, CNF and CNC, 381

whereas Pd/OMC sample exhibited the widest distribution. The average diameter was 382

close to 3-4 nm for all the samples, although when OMCs were used as support, 383

particles with a diameter in the range 6–10 nm were also observed. 384

Pd/C catalysts were analyzed by XRD technique and their patterns are given in 385

Fig. 5. In all the electrocatalysts three characteristic peaks appeared at 2θ = 39.4, 45.6, 386

67.4, which can be assigned to (1 1 1), (2 0 0) and (2 2 0) planes of Pd, respectively, 387

indicating that Pd was deposited with a face centred cubic structure. Pd crystallite size 388

was calculated from Scherrer equation and reported in Tab. 3. Similar sizes were 389

Page 17: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

17

obtained, with a value around 3-4 nm for all the catalysts, in agreement with the results 390

obtained from TEM images. Additionally, it is noticeable that the diffraction peak at 391

around 2θ = 25º, which is associated to the graphite (002) diffraction line of the support, 392

is still observed after the Pd deposition. 393

3.3 Electrochemical measurements. 394

The electrochemical behaviour of the Pd catalysts was studied in 0.1M NaHCO3, 395

at room temperature and atmospheric pressure by CV. In addition, DEMS experiments 396

were carried out in order to register simultaneously and "in-situ" the formation of 397

molecular hydrogen, which is generated during the CO2 electrochemical reduction. 398

First, several potential cycles (not shown) between -0.8 and 0.9 V in 0.1 M 399

NaHCO3 solution, previously deaerated with Ar, were recorded at 20 mVs-1

, to clean 400

and activate the catalyst surface. The voltammetric profile remains similar during the 401

activation step, confirming the stability of the catalysts in the electrolyte. After that, 402

spectro-electrochemical experiments were carried out. 403

3.3.1 Studies in the -0.8 V – 0.9 V potential range 404

At the beginning of the experiments, CVs were carried out, at 0.01 Vs-1

in 0.1 M 405

NaHCO3, cycling the potential between -0.8 and 0.9 V in absence (black curves) or in 406

presence (blue curves) of dissolved CO2. In this potential region, CO2 electrochemical 407

reduction doesn't take place. The curves for each electrode are recorded in Fig. 6. 408

As can be seen in the black curves (Ar saturated solution), during the cathodic 409

sweep, adsorption/absorption of hydrogen occurs at potentials below -0.3 V, followed 410

by the hydrogen evolution below -0.6 V. In the positive-going potential scan, hydrogen 411

is desorbed and palladium oxides are formed at E > 0.2 V. The reduction of these oxides 412

is observed in the negative scan at around -0.2 V. A small shift to more positive 413

potentials (about 50 mV) is observed in presence of CO2, (blue curves) which can be 414

Page 18: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

18

attributed to the different pHs of the electrolytes (8.4 in the presence of Ar and 6.8 in 415

the presence of CO2). 416

Comparing the CVs registered in the absence of CO2 (black curves), it is 417

remarkable the dependence of the shape of hydrogen desorption region (between -0.8 418

and 0 V in the positive scan) on the nature of the carbon support. The currents recorded 419

in this potential range are related to hydrogen adsorption/absorption (at the surface and 420

bulk of the Pd particles). This effect is also observed for CVs recorded in a CO2 421

saturated solution (blue curves). It is important to note that Pd/OMC catalyst exhibits 422

the highest reduction currents at -0.8 V, indicating that H2 evolution is favoured onto 423

this sample. 424

Different reactivity towards the hydrogen evolution reaction of the Pd/C 425

catalysts is evident from comparison of the faradic currents at -0.8 V in the presence and 426

in absence of CO2 (ICO2/IAr). All catalysts showed a ratio of ICO2/IAr around 2, with the 427

exception of Pd/CNF sample, in which the value was 3. Thus, the later catalyst presents 428

the highest reduction currents at this potential which can be ascribed to the presence of 429

high reactive Pd nanoparticles located on the edge of carbon nanofilaments, which 430

could favours hydrogen production in this potential region with CO2 disolved in the 431

solution. 432

Finally, it should be addressed that these Pd/C catalysts display differences in 433

the double layer currents, indicating that samples presented different resistance to 434

charge transfer and capacitance properties, properties that are directly related to the 435

porosity of the carbon materials. 436

3.3.2 Studies in the -1.4 V – 0.9 V potential range 437

After the experiments described in section 3.3.1, DEMS curves were recorded at 438

more negative potentials (between -1.4 V and 0.9 V), with the purpose of studying the 439

Page 19: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

19

electrochemical behaviour of the catalysts towards the CO2 electroreduction process in 440

parallel to the hydrogen evolution, both reactions occurring at these cathodic potentials. 441

Fig.7 shows the CVs (upper panel) and the corresponding MSCVs for hydrogen 442

evolution (bottom panel) recorded at 0.01 Vs-1

in the base electrolyte for the Pd/Vulcan 443

catalyst. The black curve is recorded in the base electrolyte saturated with Ar, while the 444

CO2 dissolved solution is represented by the blue curve. In the upper panel, it is 445

observed that after saturation with CO2 (blue curve), a peak around -1.0 V appears 446

during the cathodic scan while the hydrogen region is partly blocked in the anodic 447

sweep. Additionally, two oxidation peaks are developed around 0.1 and 0.3 V. These 448

results suggest that below - 0.8 V, CO2 is being reduced to other species (CO2)red, which 449

are adsorbed onto the active phase of Pd/C catalyst and oxidized during the anodic 450

scan. According to the bibliography [4, 12, 15, 23] these species are mainly COad, which 451

is adsorbed at Pd surface by the reaction: 452

OHCOHCO adad 22 2 (1) 453

although other adsorbates, such as COOHad, COHad or CHx could also be formed onto 454

Pd (see later). Consistent with these results, molecular hydrogen production is 455

significantly decreased by the presence of CO2, as it can be observed in the shift of H2 456

evolution to more negative potentials, observed in the MSCVs for m/z = 2 (Fig. 7, 457

bottom panel blue curve). 458

Similar results were obtained for the other electrocatalysts and the corresponding 459

CVs are presented in Fig. 8. All samples develop a peak around -1.0 V, which as was 460

mentioned before, is associated to the adsorption of species derived from the CO2 461

reduction, and the corresponding oxidation peaks at positive potentials. It is remarkable 462

the high reduction peak observed in the case of Pd/OMC sample (see Fig. 8 blue curve 463

for this catalyst). As mentioned above, this material presented the most significant 464

Page 20: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

20

hydrogen evolution current. Hydrogen plays a very important role in the adsorption of 465

species derived from CO2 reduction, since a certain Had coverage onto the electrode is 466

necessary for the interaction and final formation of "reduced CO2" species. Therefore, 467

higher Had coverage could lead to more significant CO2 reduction onto Pd. 468

On the other hand, with the exception of Pd/Vulcan, hydrogen evolution is 469

completely blocked due to the presence of (CO2)red species, and only for the former 470

material, the cathodic current related to this reaction is still apparent in the CVs. This 471

fact can be explained by a lower coverage of active Pd sites by adspecies at this 472

material. 473

Finally, comparing the CVs registered in the absence of CO2, it is remarkable the 474

dependence of the shape of hydrogen oxidation region (between -0.8 and 0.1 V in the 475

positive scan) on the nature of the carbon support. The currents recorded in this 476

potential range are mainly related to the oxidation of H2 retained into the carbon 477

materials, since the oxidation charges are significantly larger than a Had monolayer 478

charge, and consequently, it could not be only produced by adsorbed hydrogen. This H2 479

was formed during the cathodic scan and, due to diffusional hindrance it was not 480

released to the electrolyte, remaining into the carbon structure. Then, during the anodic 481

scan, retained hydrogen will be oxidized to protons. As can be observed (see Fig. 8 482

black curves and see later Fig. 11, black curves), Pd/Vulcan, Pd/CNF and Pd/CNC 483

catalysts presented a H2 oxidation peak, centred in the (-0.4) - (-0.25) potential range. It 484

is important to remark that the hydrogen oxidation peak is placed at more positive 485

potentials for Pd/CNC (≈ -0.2 V), in comparison with Pd/Vulcan and Pd/CNF (≈ -0.35 486

V). On the other hand, Pd/OMC catalyst shows a large oxidation range from -0.8 V to 487

0.0 V. 488

Page 21: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

21

Therefore, it seems that the diffusion and oxidation of this retained hydrogen is 489

more restricted when CNCs or OMCs are used as supports. In this context, the pore 490

structure of the materials allows to explain the results: Vulcan and CNFs present higher 491

average pore sizes, being of 10 and 11 nm respectively, while a value around 4 nm was 492

obtained for CNCs and OMCs. Pore size is a really important parameter to consider in 493

the synthesis of electrocatalysts: on one side, low diameter sizes can produce a pore 494

blockage when the metal nanoparticles are deposited onto the carbon supports; on the 495

other part, diffusion of reactives and products to or from the active sites of the 496

electrocatalysts can be hindered. In this case, textural properties of the support play an 497

important role in the release of produced H2 out from the carbon structure. Finally, 498

diffusional problems are more evident for Pd/OMC in comparison to Pd/CNC, which 499

could be attributed to the higher hydrogen evolution currents registered for the former 500

catalyst and/or their significant larger surface area (≈ 3.5 times higher for OMC than 501

CNC). As a consequence, larger amounts of hydrogen can be retained inside the carbon 502

structure. 503

3.3.3 CO and “reduced CO2” strippings 504

In order to verify if the species formed during the CO2 electroreduction were 505

irreversibly adsorbed onto the Pd/C electrocatalysts, as well as further investigate onto 506

the nature of these species, stripping voltammograms of CO and “reduced CO2” were 507

performed for all the catalysts. 508

In the case of CO stripping, CO was adsorbed on the catalyst surface by 509

bubbling this gas through the electrolyte for 30 min at - 0.5 V. The excess of CO was 510

then removed from the electrolyte with Ar during 30 min and the potential was cycled 511

between - 0.8 and 0.9 V at 0.01 Vs-1

for three complete cycles. Electrochemical active 512

areas were estimated by the integration of the oxidation of a bridge adsorbed CO 513

Page 22: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

22

monolayer (COad), assuming a charge of 210 mC cm-2

involved in the process [61, 62] . 514

The current densities (Acm-2

) given in the text were calculated from these electroactive 515

areas. 516

CO stripping curves can be observed in Fig. 9 whereas the comparison of the 517

first scan for all the catalysts is shown in Fig. 10. The hydrogen adsorption/absorption 518

region is completely blocked and two peaks associated to CO oxidation, which are 519

centred at approximately 0.15 and 0.40 V respectively, are depicted in all the CVs. 520

Other authors have obtained a bimodal distribution for CO oxidation on platinum metal 521

nanoparticles supported onto carbon materials [63-65]. It is widely known that the 522

catalytic activity towards this reaction strongly depends on the size and structure of the 523

metal nanoparticles [63, 64, 66]. Hara and co-workers showed that CO oxidation 524

reaction exhibits distinct structure sensitivity and that the onset potentials of the CO 525

oxidation reaction follow the sequence Pd (100) < Pd (110) < Pd (111) [66]. On the 526

other hand, Maillard et al. reported that Pt nanoparticles with particle size below 3 nm 527

exhibited an increase in the CO oxidation overpotential, whereas larger particles showed 528

a similar behaviour than that obtained on a polycrystalline Pt. Additionally, Pt 529

"agglomerates" with high defect sites density exhibited a remarkable activity in 530

comparison to isolated nanoparticles [64]. However, Guerin and co-workers explained 531

this bimodal distribution for CO oxidation by differences in the proportion of terrace 532

sites, which is affected by the metal particle size [63]. An increase in the particle 533

diameter results in a significant decrease in the number of edge sites and an increase of 534

the fraction of terrace sites. Accordingly, they attributed the peak at more negative 535

potentials to CO oxidation on the terraces of larger Pt nanoparticles, whereas the most 536

positive one was associated to CO oxidation on the small-particles edges. 537

Page 23: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

23

It is remarkable that the intensity of the second peak is higher for the samples 538

Pd/CNC and Pd/Vulcan, while onto catalysts supported on CNFs and OMCs the first 539

contribution becomes more important. Additionally, the onset for CO oxidation shifts to 540

more negative potential for Pd/CNF than for the rest, demonstrating an improved 541

condition for CO stripping from this catalyst surface. This different behaviour for Pd 542

catalysts could be attributed to the different nanoparticle size or dispersion onto the 543

carbon supports. As previously observed in TEM images, Pd/OMC displayed larger 544

metal nanoparticles, which could favour the growth of the appropriate Pd 545

crystallographic facets leading to a decrease in the onset potential for CO 546

electrooxidation [66]. On the other hand, Pd/CNF showed a higher density of particles 547

on the edge of the carbon nanofilaments, and consequently, an increase in the amount of 548

border sites contacting two nanoparticles. Therefore, larger particles or high density of 549

Pd may leads to a CO oxidation at more negative potentials, in agreement with the 550

bibliography [63, 64, 66]. However, Pd/CNC presented a bad dispersion of small size 551

nanoparticles, and in this case, the improvement condition towards CO oxidation was 552

not observed. 553

Afterthat, “reduced CO2” strippings were carried out for all the samples as 554

follows: first, CO2 was bubbled into the NaHCO3 electrolyte until saturation. Then, the 555

potential was set at -1.0 V for 30 min where CO2 is reduced forming reduced species 556

that adsorb onto the Pd/C catalysts. Subsequently, the solution was saturated with Ar 557

during 40 min. Finally, the potential was cycled between - 1.4 and 0.9 V at 0.01 Vs-1

for 558

three complete cycles. Fig. 11 shows the “reduced CO2” stripping voltammograms of 559

the Pd/C catalysts and Fig. 12 a comparison of the first scan of them. In all the samples, 560

two peaks associated to adsorbates derived from the CO2 electroreduction reaction were 561

observed. The first contribution appears centered in the 0.0 - 0.2 potential range, 562

Page 24: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

24

followed by a second one centered at about 0.3 – 0.4 V. The intensity of the first peak is 563

higher for all the catalysts, except for the sample Pd/OMC, on which the second 564

contribution becomes more important. It should be mentioned that the hydrogen 565

adsorption/absorption region is completely blocked for all catalysts including 566

Pd/Vulcan, opposite to the results obtained when potential was not set at -1 V for 30 567

min (Fig. 8.A). 568

A comparison of the CO and “reduced CO2” strippings for all the Pd/C 569

electrocatalysts can be seen in Fig. 13. It is observed that the oxidation of the reduced 570

species from the CO2 conversion takes place at lower potentials than for CO oxidation. 571

In addition, the enclosed area below the “reduced CO2” stripping voltammogram is 572

much greater, for all the samples. These results suggest that, at potentials below -0.8 V, 573

CO2 is reduced not only to adsorbed CO, but also other adsorbed species, such as 574

COOHad, COHad or CHx could be obtained. However, the presence of these adsorbates 575

on the Pd surface produces the blocking of the hydrogen region, and consequently, a 576

shift of the hydrogen oxidation features towards more positive values (some kind of 577

adduct may possible be formed between Hab and the adspecies [13, 22, 23]), partially 578

overlapping with the current of adsorbed species derived from CO2 reduction. 579

Therefore, part of this higher area observed in the “reduced CO2” strippings in 580

comparison to the CO ones, could be explained by the oxidation of absorbed hydrogen. 581

It is important to note that hydrogen atoms absorbed into Pd could take part in the 582

electrochemical reaction, exhibiting a higher chemical reactivity than gaseous hydrogen. 583

K. Ohkawa and coworkers reported an increase in the selectivity for CO2 reduction after 584

hydrogen was absorbed into a bulk-Pd electrode. This increase was a result from the 585

suppression of H2 evolution and/or the increase of the reaction rates of CO and HCOOH 586

Page 25: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

25

production [23]. They suggested the participation of absorbed hydrogen in the reduction 587

of CO2. 588

The ratio between the oxidation charges associated to “reduced CO2” and CO (589

CO

redCO

Q

Qk

,2 ) was calculated by the integration of the anodic curves obtained from the 590

stripping voltammograms (Figs. 10 and 12). Higher values of k were obtained for Pd 591

catalysts supported onto CNFs (3.21) followed by the commercial material Vulcan 592

(3.02). Smaller values were calculated for CNC and OMC supported catalysts (2.41 and 593

2.40, respectively). From these data, it can be concluded that COad cannot be the only 594

species on the surface. 595

If it is assumed that the amount of Hab is similar for all the samples, since the Pd 596

load and the average metal sizes were similar for them, the explanation for these values 597

should imply the presence of different adsorbates onto the electrocatalysts. As was said 598

above (section 3.3.2) COad can be formed at the Pd surface by reaction (1), which 599

probably cannot take place in one step but in a sequence of them: 600

adad COOHHCO 2 (2) 601

OHCOHCOOH adadad 2 (3) 602

Additionally, COad could interact with adsorbed hydrogen forming COHad, by 603

the reaction: 604

adadad COHHCO 605

(4) 606

These reactions do not represent the adsorption mechanism for CO2, only try to show 607

that the presence of other adspecies can be easily justified. Also absorbed hydrogen 608

could be involved in reactions (2)-(4). These adsorbed species when oxidized to CO2 609

involved different oxidation charges (2 e- for COad, 3 for COHad and 1 for COOHad), and 610

Page 26: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

26

accordingly, k values can vary assuming that reaction (2), (3) or (4) are favoured for 611

each electrocatalyst. This would suggest differences in the strength of Pd-Had and/or Pd-612

Hab depending on the support, which could have an influence on the CO2 613

electrochemical reduction process. 614

Nevertheless, other factors such as the conductivity, textural properties, 615

morphology or structure of the support, as well as the size and distribution of Pd 616

nanoparticles or a good compromise between them, could be also affecting the 617

performance of the Pd-based electrocatalysts towards the electrochemical reduction of 618

CO2. Additonally, the formation of other adsorbed species or even of volatiles products 619

is not discarded. Other authors have reported the formation of volatile CO and HCOOH 620

at potentials below -1.4 vs. Ag/AgCl [4, 15]. Azuma et al. reported also the formation of 621

hydrocarbons traces at higher overpotentials [15]. 622

In order to further investigate about the nature of these species spectro-623

electrochemical measurements should be carried out. However, this study is out of the 624

scope of the present work: the main objective of the present study is to know if it is 625

possible to electrochemically reduce CO2 on carbon-supported Pd catalysts and to 626

establish if the carbon support influences on this particular reaction. The product 627

distribution on Pd/C catalysts will be further discussed in a future paper. DEMS 628

analysis of the reduction products is currently underway. Additionally, in order to 629

analyse the influence of the absorbed hydrogen, a study in detail of CO strippings at 630

different potentials is being carried out. 631

4. Conclusions 632

Pd catalysts were synthesized for being used in the electroreduction of CO2 to 633

high-added value products. Different novel nanostructured carbon materials, such as 634

CNCs, CNFs and OMCs, as well as a commercial material, Vulcan XC-72R, were used 635

Page 27: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

27

as support of the catalysts. In this way, the influence of the carbon support on the 636

physicochemical and electrochemical properties of the catalysts for the CO2 637

electroreduction process was studied. On the basis of the results obtained, we concluded 638

as follows: 639

Supports presented different morphology and structure. CNCs and CNFs 640

exhibited a crystalline structure with well-aligned graphitic layers. On the other hand, 641

OMCs showed a hexagonal ordered structure, composed of not crystalline carbon. 642

Finally, Vulcan presented an intermediate structure between amorphous and graphitic. 643

Additionally, supports exhibited different textural properties, being OMCs the material 644

that presented the highest specific surface area and pore volume, due to their preparation 645

method (nanocasting technique). CNFs and Vulcan presented wide pore size 646

distribution, while OMCs and CNCs showed narrow distribution, with lower pore sizes 647

(3 – 4 nm). 648

These differences in the physicochemical properties of the supports do not affect 649

significantly the average Pd particle size of catalysts. All catalysts presented a Pd 650

average particle size around 3 – 4 nm, although when OMCs were used as support, 651

particles with a diameter in the range 6 – 10 nm were also observed. In general, a good 652

dispersion of the metal phase onto the carbon materials was obtained, with the 653

exception of Pd/CNC sample, where the Pd nanoparticles were not homogenously 654

distributed. Additionally, Pd/CNF electrocatalyst presented a a higher density of 655

particles on the edge of the carbon nanofilaments. 656

Catalysts displayed different responses to hydrogen adsorption/absorption, 657

evolution and oxidation processes, which have influence on the CO2 electroreduction 658

reaction. Additionally, the textural properties of the supports conditioned the diffusion 659

hindrance of H2 which is partially retained into the carbon structure. 660

Page 28: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

28

Cyclic voltammetry studies showed that CO2 was effectively reduced to other 661

species, which were adsorbed onto Pd, at -1.0 V vs. Ag/AgCl. By comparison of the 662

oxidation charges obtained from CO and “reduced CO2” strippings, it was concluded 663

that these species were not only COad, but also other adsorbates were formed. However, 664

Hab could be also affecting the results, forming some kind of adduct with the adspecies. 665

Finally, these electrochemical studies demonstrated that the catalytic activity of 666

Pd/C materials toward the CO2 electroreduction reaction are influenced by the carbon 667

support used. 668

669

Acknowledgments 670

The authors gratefully acknowledge financial support given by Spanish MINECO 671

(CTQ2011‐28913‐C02‐01 and 02). S. Pérez-Rodríguez acknowledges Gobierno de 672

Aragón for the DGA grant. 673

674

Page 29: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

29

References 675

[1] E.A. Batista, M.L.A. Temperini, Journal of Electroanalytical Chemistry 629 676

(2009) 158-163. 677

[2] M. Gattrell, N. Gupta, A. Co, Journal of Electroanalytical Chemistry 594 (2006) 678

1-19. 679

[3] Y. Hori, H. Wakebe, T. Tsukamoto, O. Koga, Electrochimica Acta 39 (1994) 680

1833-1839. 681

[4] M. Azuma, K. Hashimoto, M. Hiramoto, M. Watanabe, T. Sakata, Journal of the 682

Electrochemical Society 137 (1990) 1772-1778. 683

[5] Y. Hori, Handbook of Fuel Cells, John Wiley & Sons, Ltd, 2010. 684

[6] B. Innocent, D. Pasquier, F. Ropital, F. Hahn, J.M. Léger, K.B. Kokoh, Applied 685

Catalysis B: Environmental 94 (2010) 219-224. 686

[7] Y. Hori, H. Konishi, T. Futamura, A. Murata, O. Koga, H. Sakurai, K. Oguma, 687

Electrochimica Acta 50 (2005) 5354-5369. 688

[8] H. De Jesús-Cardona, C. Del Moral, C.R. Cabrera, Journal of Electroanalytical 689

Chemistry 513 (2001) 45-51. 690

[9] S. Kaneco, N.H. Hiei, Y. Xing, H. Katsumata, H. Ohnishi, T. Suzuki, K. Ohta, 691

Electrochimica Acta 48 (2002) 51-55. 692

[10] S. Kaneco, N.H. Hiei, Y. Xing, H. Katsumata, H. Ohnishi, T. Suzuki, K. Ohta, 693

Journal of Solid State Electrochemistry 7 (2003) 152-156. 694

[11] B. Jermann, J. Augustynski, Electrochimica Acta 39 (1994) 1891-1896. 695

[12] D. Kolbe, W. Vielstich, Electrochimica Acta 41 (1996) 2457-2460. 696

[13] M. Grdeń, A. Paruszewska, A. Czerwiński, Journal of Electroanalytical 697

Chemistry 502 (2001) 91-99. 698

[14] S. Taguchi, A. Aramata, M. Enyo, Journal of Electroanalytical Chemistry 372 699

(1994) 161-169. 700

[15] M. Azuma, K. Hashimoto, M. Watanabe, T. Sakata, Journal of Electroanalytical 701

Chemistry 294 (1990) 299-303. 702

[16] M.C. Arévalo, C. Gomis-Bas, F. Hahn, B. Beden, A. Arévalo, A.J. Arvia, 703

Electrochimica Acta 39 (1994) 793-799. 704

[17] A. Rodes, E. Pastor, T. Iwasita, Journal of Electroanalytical Chemistry 369 705

(1994) 183-191. 706

[18] A. Rodes, E. Pastor, T. Iwasita, Journal of Electroanalytical Chemistry 377 707

(1994) 215-225. 708

[19] Y.B. Vassiliev, V.S. Bagotzky, N.V. Osetrova, A.A. Mikhailova, Journal of 709

Electroanalytical Chemistry 189 (1985) 311-324. 710

[20] B. Beden, A. Bewick, M. Razaq, J. Weber, Journal of Electroanalytical 711

Chemistry 139 (1982) 203-206. 712

[21] P.A. Christensen, A. Hamnett, A.V.G. Muir, N.A. Freeman, Journal of 713

Electroanalytical Chemistry 288 (1990) 197-215. 714

[22] A. Czerwiński, S. Zamponi, R. Marassi, Journal of Electroanalytical Chemistry 715

and Interfacial Electrochemistry 304 (1991) 233-239. 716

[23] K. Ohkawa, K. Hashimoto, A. Fujishima, Journal of Electroanalytical Chemistry 717

345 (1993) 445-456. 718

[24] A. Rodes, E. Pastor, T. Iwasita, Journal of Electroanalytical Chemistry 373 719

(1994) 167-175. 720

[25] R.P.S. Chaplin, A.A. Wragg, Journal of Applied Electrochemistry 33 (2003) 721

1107-1123. 722

Page 30: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

30

[26] E.V. Kondratenko, G. Mul, J. Baltrusaitis, G.O. Larrazábal, J. Pérez-Ramírez, 723

Energy and Environmental Science 6 (2013) 3112-3135. 724

[27] F.A. Lewis, The Palladium-Hydrogen Systems, Academic Press, New York, 725

1967. 726

[28] K. Hara, A. Kudo, T. Sakata, Journal of Electroanalytical Chemistry 391 (1995) 727

141-147. 728

[29] S. Nakagawa, A. Kudo, M. Azuma, T. Sakata, Journal of Electroanalytical 729

Chemistry 308 (1991) 339-343. 730

[30] B. Eneau-Innocent, D. Pasquier, F. Ropital, J.M. Léger, K.B. Kokoh, Applied 731

Catalysis B: Environmental 98 (2010) 65-71. 732

[31] K. Ohta, M. Kawamoto, T. Mizuno, D.A. Lowy, Journal of Applied 733

Electrochemistry 28 (1998) 717-724. 734

[32] Y.B. Vassiliev, V.S. Bagotzky, O.A. Khazova, N.A. Mayorova, Journal of 735

Electroanalytical Chemistry 189 (1985) 295-309. 736

[33] C. Genovese, C. Ampelli, S. Perathoner, G. Centi, Journal of Energy Chemistry 737

22 (2013) 202-213. 738

[34] M.N. Mahmood, D. Masheder, C.J. Harty, Journal of Applied Electrochemistry 739

17 (1987) 1159-1170. 740

[35] S. Pérez-Rodríguez, G. García, L. Calvillo, V. Celorrio, E. Pastor, M.J. Lázaro, 741

International Journal of Electrochemistry 2011 (2011). 742

[36] G. Centi, S. Perathoner, Z.S. Rak, Applied Catalysis B: Environmental 41 743

(2003) 143-155. 744

[37] M. Gangeri, S. Perathoner, S. Caudo, G. Centi, J. Amadou, D. Bégin, C. Pham-745

Huu, M.J. Ledoux, J.P. Tessonnier, D.S. Su, R. Schlögl, Catalysis Today 143 (2009) 57-746

63. 747

[38] S. Perathoner, M. Gangeri, P. Lanzafame, G. Centi, Kinetics and Catalysis 48 748

(2007) 877-883. 749

[39] G. Centi, S. Perathoner, G. Winè, M. Gangeri, Green Chemistry 9 (2007) 671-750

678. 751

[40] G. Centi, S. Perathoner, Topics in Catalysis 52 (2009) 948-961. 752

[41] Y. Hori, K. Kikuchi, S. Suzuki, Chem. Lett. (1985) 1695-1698. 753

[42] C.W.B. Bezerra, L. Zhang, H. Liu, K. Lee, A.L.B. Marques, E.P. Marques, H. 754

Wang, J. Zhang, Journal of Power Sources 173 (2007) 891-908. 755

[43] D. Sebastián, I. Suelves, R. Moliner, M.J. Lázaro, Carbon 48 (2010) 4421-4431. 756

[44] D. Sebastián, I. Suelves, R. Moliner, M.J. Lázaro, A. Stassi, V. Baglio, A.S. 757

Aricò, Applied Catalysis B: Environmental 132-133 (2013) 22-27. 758

[45] D. Sebastián, I. Suelves, E. Pastor, R. Moliner, M.J. Lázaro, Applied Catalysis 759

B: Environmental 132-133 (2013) 13-21. 760

[46] V. Celorrio, L. Calvillo, M.V. Martínez-Huerta, R. Moliner, M.J. Lázaro, 761

Energy and Fuels 24 (2010) 3361-3365. 762

[47] V. Celorrio, L. Calvillo, S. Pérez-Rodríguez, M.J. Lázaro, R. Moliner, 763

Microporous and Mesoporous Materials 142 (2011) 55-61. 764

[48] L. Calvillo, V. Celorrio, R. Moliner, P.L. Cabot, I. Esparbé, M.J. Lázaro, 765

Microporous and Mesoporous Materials 116 (2008) 292-298. 766

[49] M.J. Lázaro, L. Calvillo, E.G. Bordejé, R. Moliner, R. Juan, C.R. Ruiz, 767

Microporous and Mesoporous Materials 103 (2007) 158-165. 768

[50] Y. Meng, D. Gu, F. Zhang, Y. Shi, H. Yang, Z. Li, C. Yu, B. Tu, D. Zhao, 769

Angewandte Chemie - International Edition 44 (2005) 7053-7059. 770

[51] J. Biscoe, B.E. Warren, Journal of Applied Physics 13 (1942) 364-371. 771

Page 31: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

31

[52] J.I. Langford, A.J.C. Wilson, Journal of Applied Crystallography 11 (1978) 102-772

113. 773

[53] D. Torres, J.L. Pinilla, M.J. Lázaro, R. Moliner, I. Suelves, International Journal 774

of Hydrogen Energy. 775

[54] J. Maire, J. Mering, Chem. Phys. Carbon 6 (1970) 125-189. 776

[55] G. García, J. Flórez, O. Guillén-Villafuerte, J.L. Rodríguez, E. Pastor, In 777

preparation. 778

[56] S. Song, K. Wang, Y. Liu, C. He, Y. Liang, R. Fu, D. Wu, Y. Wang, 779

International Journal of Hydrogen Energy 38 (2013) 1405-1412. 780

[57] H. Marsh, F. Rodríguez-Reinosos, Elsevier Science & Technology Books, 2006. 781

[58] G. Leofanti, M. Padovan, G. Tozzola, B. Venturelli, Catalysis Today 41 (1998) 782

207-219. 783

[59] T.V. Reshetenko, L.B. Avdeeva, Z.R. Ismagilov, V.V. Pushkarev, S.V. 784

Cherepanova, A.L. Chuvilin, V.A. Likholobov, Carbon 41 (2003) 1605-1615. 785

[60] S.K. Shaikhutdinov, V.I. Zaikovskii, L.B. Avdeeva, Applied Catalysis A: 786

General 148 (1996) 123-133. 787

[61] M.C. Ar valo, J.L. Rodr guez, E. Pastor, Journal of Electroanalytical Chemistry 788

505 (2001) 62-71. 789

[62] M.C. Ar valo, J.L. Rodr guez, E. Pastor, Journal of Electroanalytical Chemistry 790

472 (1999) 71-82. 791

[63] S. Guerin, B.E. Hayden, C.E. Lee, C. Mormiche, J.R. Owen, A.E. Russell, B. 792

Theobald, D. Thompsett, Journal of Combinatorial Chemistry 6 (2004) 149-158. 793

[64] F. Maillard, S. Schreier, M. Hanzlik, E.R. Savinova, S. Weinkauf, U. Stimming, 794

Physical Chemistry Chemical Physics 7 (2005) 385-393. 795

[65] S. Sun, Z. Jusys, R.J. Behm, Journal of Power Sources 231 (2013) 122-133. 796

[66] M. Hara, U. Linke, T. Wandlowski, Electrochimica Acta 52 (2007) 5733-5748. 797

798

799

800

Page 32: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

32

FIGURE CAPTIONS 801

Figure. 1. TEM images of carbon supports: Vulcan (A,B), OMCs (C,D), CNFs (E,F) 802

and CNCs (G,H). 803

Figure. 2. XRD patterns of carbon supports (A) and XRD pattern at low angles for the 804

OMC sample (B). 805

Figure. 3. Nitrogen adsorption–desorption isotherms (A) and PSD of carbon supports 806

(B). 807

Figure. 4. TEM images and PSD of 20 % Pd/C electrocatalysts: Vulcan (A), OMCs 808

(B), CNFs (C) and CNCs (D). 809

Figure. 5. XRD patterns of 20 % Pd/C electrocatalysts. 810

Figure. 6. CVs of Pd/C catalysts in 0.1 M NaHCO3 (v = 10 mVs-1

, Ei = 0.2 V vs. 811

Ag/AgCl). Black curves: Ar saturated solution. Blue curves: CO2 saturated solution. 812

Figure. 7. CVs (upper panel) and MSCV for H2, m/z = 2, (bottom panel) for Pd/Vulcan 813

catalyst in 0.1 M NaHCO3 (v = 10 mVs-1

, Ei = 0.2 V vs. Ag/AgCl). Black curves: Ar 814

saturated solution. Blue curves: CO2 saturated solution. Inset shows a zoom of the 815

oxidation region. 816

Figure. 8. CVs of Pd/C catalysts in 0.1 M NaHCO3 (v = 10 mVs-1

, Ei = 0.2 V vs. 817

Ag/AgCl). Black curves: Ar saturated solution. Blue curves: CO2 saturated solution. 818

Figure. 9. CO stripping voltammograms (blue line) of Pd/ C cataslysts in 0.1 M 819

NaHCO3 (v = 10 mVs-1

, Ead = -0.5 V vs. Ag/AgCl). 820

Figure. 10. Comparison of the first scan for CO stripping voltammograms of Pd/ C 821

cataslysts in 0.1 M NaHCO3 (v = 10 mVs-1

, Ead = -0.5 V vs. Ag/AgCl). 822

Figure. 11. “Reduced CO2” stripping voltammograms of Pd/ C cataslysts in 0.1 M 823

NaHCO3 (v = 10 mVs-1

, Ead = -1.0 V vs. Ag/AgCl). 824

Page 33: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

33

Figure. 12. Comparison of the first scan for “reduced CO2” stripping voltammograms 825

of Pd/ C cataslysts in 0.1 M NaHCO3 (v = 10 mVs-1

, Ead = -1.0 V vs. Ag/AgCl). 826

Figure. 13. Comparison of CO and “reduced CO2” stripping voltammograms for Pd/C 827

catalysts in 0.1 M NaHCO3 (v = 10 mVs-1

). 828

829

Page 34: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

34

Tables 830

Table. 1. XRD parameters of carbon supports 831

Sample 2θ (º) FWHM (º) d002 (Å) Lc (Å) g (%)

Vulcan 25.00 4.82 3.559 1.7 -

CNC 26.23 0.97 3.394 8.4 53

CNF 26.54 0.78 3.356 10.5 97

832

Table. 2. Textural and TGA parameters of carbon supports 833

Sample

SBET (m2g

-1) VT (cm

3g

-1) Dp (nm) % ash content -

TGA

Vulcan 237.9 0.38 10.4 2.29

CNF 75.9 0.24 11.5 4.00

CNC 233.9 0.17 4.1 5.37

OMC 812.3 0.55 3.8 3.25

834

Table. 3. EDX, XRD and TGA properties of 20 % Pd/C electrocatalysts 835

Sample

% Pd wt. -

EDX

% Pd wt. -

TGA

d (nm)

Pd/Vulcan 17.73 19.12 3.5

Pd/CNF 18.81 18.89 2.9

Pd/CNC 20.17 18.70 3.1

Pd/OMC 19.89 18.93 3.5

836

837

838

Page 35: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

35

Figure. 1. 839

840

841

842

843

844

A B

C D

E F

G H

Page 36: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

36

Figure. 2. 845

20 30 40 50 60 70

OMC

Vulcan

CNCInte

nsity (

a.u

.)

2Theta (deg)

C (101)C (002) C (100)

CNF

A

846

1.0 1.5 2.0 2.5 3.0 3.5 4.0

(100)

B

2Theta (deg)

Inte

nsity (

u.a

.)

OMC

847

848

849

850

Page 37: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

37

851

Figure. 3. 852

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.00

50

100

150

200

250

300

350

400

Vo

lum

e a

dso

rbe

d (

cm

3g

-1)

Relative pressure (P/P0)

Vulcan

CNF

CNC

OMC

A

853

10 100

0.0

0.2

0.4

0.6

0.8

1.0

Dp (nm)

Po

re v

olu

me

(cm

3g

-1) Vulcan

CNF

CNC

OMC

B

854

855

856

857

858

Page 38: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

38

Figure. 4. 859

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 150

3

6

9

12

15

18

21

24

27

30

33

(N/N

TO

TA

L)

x 1

00

(%

)

Particle size (nm)

Pd/Vulcan

860

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 150

3

6

9

12

15

18

21

24

27

30

33

(N/N

TO

TA

L)

x 1

00 (

%)

Particle size (nm)

Pd/OMC

861

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 150

3

6

9

12

15

18

21

24

27

30

33

(N/N

TO

TA

L)

x 1

00 (

%)

Particle size (nm)

Pd/CNF

862

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 150

3

6

9

12

15

18

21

24

27

30

33

(N/N

TO

TA

L)

x 1

00

(%

)

Particle size (nm)

Pd/CNC

863

A

B

C

D

Page 39: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

39

Figure. 5. 864

20 30 40 50 60 70

Pd/OMC

Pd/Vulcan

Pd/CNFInte

nsity (

a.u

.)

2Theta (deg)

Pd/CNC

Pd (111) Pd (200) Pd (220)C (002)

865

866

867

868

869

870

871

872

873

874

875

876

Page 40: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

40

Figure. 6. 877

-0,06

-0,04

-0,02

0,00

0,02

-0,8 -0,4 0,0 0,4 0,8

-0,06

-0,04

-0,02

0,00

0,02

-0,8 -0,4 0,0 0,4 0,8

Pd/Vulcan

Ar

CO2

Ar

CO2

Pd/CNF

Potential (V vs Ag/AgCl)

Curr

ent

Density (

mA

cm

-2)

Ar

CO2

Pd/CNC

Ar

CO2

Pd/OMC

878 879

880

881

882

883

Page 41: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

41

Figure. 7. 884

-0,2

-0,1

0,0

-1,6 -1,2 -0,8 -0,4 0,0 0,4 0,8

-0,3 0,0 0,3 0,6 0,9

-0,02

0,00

0,02

Potential (V vs Ag/AgCl)

Curr

ent D

ensity (

mA

cm

-2)

Ar

CO2

Pd/VulcanC

urr

ent

Density (

mA

cm

-2)

1 x 10-5

Potential (V vs Ag/AgCl)

Ionic

Curr

ent

(a.u

.)

m/z = 2

885

886

887

888

Page 42: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

42

Figure. 8. 889

-0,3

-0,2

-0,1

0,0

-1,5 -1,0 -0,5 0,0 0,5 1,0-0,3

-0,2

-0,1

0,0

-1,5 -1,0 -0,5 0,0 0,5 1,0

Curr

ent

Density (

mA

cm

-2)

Ar

CO2

Pd/Vulcan

Ar

CO2

Pd/CNF

Potential (V vs Ag/AgCl)

Ar

CO2

Pd/CNC

Ar

CO2

Pd/OMC

890 891

892

893

894

895

Page 43: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

43

Figure. 9. 896

-0,030

-0,015

0,000

0,015

0,030

-0,8 -0,4 0,0 0,4 0,8

-0,030

-0,015

0,000

0,015

0,030

-0,8 -0,4 0,0 0,4 0,8

Pd/Vulcan Pd/CNF

Cu

rren

t D

en

sity (

mA

cm

-2)

Pd/CNC

Potential (V vs Ag/AgCl)

Pd/OMC

897 898

899

900

901

902

Page 44: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

44

Figure. 10. 903

0,0 0,2 0,4 0,6 0,8 1,0-0,005

0,000

0,005

0,010

0,015

0,020

0,025 Pd/Vulcan

Pd/CNF

Pd/CNC

Pd/OMC

Cu

rre

nt

De

nsity (

mA

cm

-2)

Potential (V vs Ag/AgCl) 904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

Page 45: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

45

Figure. 11. 919

-0,3

-0,2

-0,1

0,0

-1,5 -1,0 -0,5 0,0 0,5 1,0-1,5 -1,0 -0,5 0,0 0,5 1,0-0,3

-0,2

-0,1

0,0

Pd/Vulcan Pd/CNF

Pd/OMC

Curr

ent

Density (

mA

cm

-2)

Potential (V vs Ag/AgCl)

Pd/CNC

920 921

922

923

924

Page 46: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

46

Figure. 12. 925

-0,4 -0,2 0,0 0,2 0,4 0,6 0,8 1,0

-0,02

-0,01

0,00

0,01

0,02

Pd/Vulcan

Pd/CNF

Pd/CNC

Pd/OMC

Cu

rre

nt

De

nsity (

mA

cm

-2)

Potential (V vs Ag/AgCl) 926

927

928

929

930

931

932

933

934

Page 47: Pd catalysts supported onto nanostructured carbon ...digital.csic.es/bitstream/10261/113454/1/Pd catalysts supported ont… · 127 reduction is carried out using different nanostructured

47

Figure. 13. 935

-0,02

0,00

0,02

-0,3 0,0 0,3 0,6 0,9-0,3 0,0 0,3 0,6 0,9

-0,02

0,00

0,02

CO stripping

CO2,red

stripping

Ar blank

CO stripping

CO2,red

stripping

Ar blank

CO stripping

CO2,red

stripping

Ar blank

CO stripping

CO2,red

stripping

Ar blank

Pd/Vulcan Pd/CNF

Potential (V vs Ag/AgCl)

Curr

ent

Density (

mA

cm

-2)

Pd/OMC

Pd/CNC

936

937

938

939

940

941

942

943

944