Hydrodeoxygenation of Acetophenone over Supported Precious Metal

Nov 23, 2015 - advantage of precious metal catalysts is the high activity, even at .... carried out running the reaction for long periods of time (60 ...
1 downloads 0 Views 675KB Size
Subscriber access provided by UNIV LAVAL

Article

Hydrodeoxygenation of acetophenone over supported precious metal catalysts at mild conditions: process optimization and reaction kinetics Celeste Gonzalez, Pablo Marin, Fernando V. Díez, and Salvador Ordonez Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.5b02112 • Publication Date (Web): 23 Nov 2015 Downloaded from http://pubs.acs.org on November 24, 2015

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.

Energy & Fuels 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 30

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

Energy & Fuels

1 2 3

Hydrodeoxygenation of acetophenone over supported precious metal catalysts at mild conditions: process optimization and reaction kinetics

4 5

Celeste González, Pablo Marín, Fernando V. Díez, Salvador Ordóñez*

6

Department of Chemical and Environmental Engineering, University of Oviedo, Facultad de Química,

7

Julián Clavería 8, Oviedo 33006, SPAIN

8

* Phone: 34-985 103 437, FAX: 34-985 103 434, e-mail: [email protected]

9 10

Abstract

11

Bio-oils obtained by pyrolysis of lignocellulose feedstocks must be upgraded to reduce the oxygen

12

content, improving their quality as bio-fuels. Catalytic hydrotreatment has been proposed to reduce

13

the oxygen content of bio-fuels and meet the standard requirements. Acetophenone is interesting as

14

a model compound for the study of hydrodeoxygenation of pyrolysis bio-oils, which contain aromatic

15

ketones. In this work, acetophenone gas phase hydrodeoxygenation over precious metal (Pt, Pd, Ru

16

and Rh) supported catalysts has been studied in a fixed-bed reactor (space time, W/F = 0.75-1.0 kgcat

17

s/moltot). The influence of catalyst active phase and operating conditions (pressure 0.5-1.5 MPa, and

18

temperature 275-375ºC) on the catalyst stability and activity, and product distribution was studied.

19

For the optimum pressure (1.0 MPa), and 275 and 375ºC, a reaction kinetic model based on the

20

reaction scheme has been proposed and fitted to the experimental data obtained at different space

21

velocities (W/F = 0-1.5 kgcat s/moltot).

22 23

Keywords: bio-oil; hydrotreating; palladium catalyst; fixed-bed reactor; kinetic modelling.

1 ACS Paragon Plus Environment

Energy & Fuels

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

Page 2 of 30

24

1. Introduction

25

The development of cleaner and renewable fuels is growing due to the environmental hazards and

26

possible shortage of traditional fossil fuels. In this context, biomass constitutes an energy resource

27

widely available around the world. The use of lignocellulose materials as feedstock to produce bio-

28

fuels increases the yield of the biomass conversion, in comparison to traditional bio-fuels (bioethanol

29

and lipid biodiesel), whose manufacture requires the use of a very small, and edible (starches, lipids),

30

fraction of this biomass.1

31

One of the proposed schemes for upgrading lignocellulosic feedstocks is based on biomass pyrolysis

32

followed by a chemical upgrading of the resulting oil. Pyrolysis is a thermal decomposition of the

33

biomass that produces a pyrolysis oil (bio-oil), formed by a complex mixture of hydrocarbons, with

34

high oxygen content. Some oxygenated functional groups appearing in bio-oils are carboxyl, carbonyl

35

(acetone and aldehyde), alcohol and ester.2-4 Oxygen content in bio-oils is usually 35-40%,5 while in

36

heavy petroleum fuel oil is only 1%.6 The presence of oxygenated compounds deteriorates the

37

properties of bio-oils as bio-fuels, increasing viscosity, acidity and instability, decreasing volatility and

38

energy density, etc.7 Upgrading of bio-oils is aimed to decrease their oxygen content, improving the

39

properties of the resulting bio-fuel. Two types of upgrading catalytic processes have been proposed:

40

cracking and hydrogenation.

41

Catalytic hydrogenation, and most specifically hydrodeoxygenation (HDO), is based on the reaction

42

with hydrogen at high temperature and pressure in presence of a catalyst, leading to bio-oil

43

constituents.8-10 Conventional hydrotreating catalysts, such as CoMo/Al2O3 and NiMo/Al2O3, are used

44

at industrial scale in hydrotreating petroleum fractions, for the simultaneous elimination of sulphur,

45

oxygen and nitrogen.10 The use of these catalysts for hydrotreating of bio-oils has been extensively

46

studied: optimization of catalyst formulation and operating conditions (typically 250-400ºC and 3.0-

47

20.0 MPa)8, 11, 12 and also kinetic studies.13-17 These catalysts are used in sulphided form, which

48

considerably outperforms the oxidic form in HDO of bio-oils, and hence the reaction must be carried

2 ACS Paragon Plus Environment

Page 3 of 30

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

Energy & Fuels

49

out in presence of sulphur. Sulphur content of bio-oil is too small to produce enough H2S during

50

hydrotreating, and for this reason an external source of H2S is required to maintain the catalyst in a

51

stable sulphided form.8 This is an important drawback of hydrotreating catalysts for bio-oil HDO.

52

Recently, precious metal-based catalysts, e.g. Pt18 and Ru,19-21 have been proposed and studied as

53

alternative to conventional hydrotreating catalysts.22, 23 The most important advantage of precious

54

metal catalysts is the high activity, even at low pressure and temperature, which would result in

55

smaller reactors and economic advantages.

56

Studies with model compounds are useful for determining the activity and stability of catalysts used

57

in hydrodeoxygenation of bio-oils, and also for understanding the reaction kinetics. The composition

58

of pyrolysis bio-oils is rather complex, formed by different families of compounds with concentration

59

varying according to the biomass feedstock or the processing conditions. Compounds often used as

60

models for bio-oils are guaiacol (2-methoxyphenol), ethyl phenol, or anisole (methoxybenzene),

61

among others. These compounds are representative of lignin precursor molecules.8 The composition

62

of bio-oils also accounts for acids, esters, ketones, etc. These compounds, with a higher oxidation

63

state, require a higher degree of hydrodeoxygenation.

64

In this work, the focus is set on ketones (around 27 wt% of the crude bio-oil24), and particularly

65

acetophenone, which is selected as model compound.25, 26 Acetophenone is a simple aromatic

66

ketone, representative of other more complex ketones and very interesting as model compound,

67

because contains two functional groups frequently present in bio-oils:4, 27 carbonyl and aromatic ring.

68

Hydrogenation of these functional groups leads to a complex reaction network formed by

69

competitive and consecutive steps. In bio-oils HDO, hydrogenation of the aromatic ring is not

70

desired, because the saturated product has a lower octane number, and the reaction consumes

71

valuable hydrogen.28 Hence, the optimization of catalyst formulation and reaction conditions is

72

critical to achieve high de-oxygenation with low hydrogen consumption. Although acetophenone

73

hydrogenation has been widely studied over different catalysts, such as Pt,29-34 Pd,28, 34-38 Rh,39 Ni,40, 41

3 ACS Paragon Plus Environment

Energy & Fuels

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 30

74

Co42 and Cu,43 most studies were carried out in liquid phase and focused on catalyst screening and

75

optimization to maximize the yield of 1-phenylethanol, an alcohol extensively used in perfumery and

76

pharmaceuticals.33 But for bio-oil upgrading, the catalyst and operating conditions should be

77

optimized to maximize hydrodeoxygenation of the carbonyl group in detriment of hydrogenation of

78

the aromatic ring.

79

The stability and activity of different commercial precious metal (Pt, Pd, Ru and Rh) supported

80

catalysts was firstly studied in a laboratory fixed-bed reactor. Then, the operating conditions were

81

optimized for the most promising catalyst. Finally, reaction kinetic models, based on a reaction

82

scheme, were proposed and fit to the experimental data. Most published studies on

83

hydrodeoxygenation are only focused on optimizing the catalyst formulation. However, kinetic

84

models for readily available commercial catalysts are very interesting, particularly in the scale-up of

85

the process to industrial scale.

86

87

2. Experimental section

88

2.1. Materials

89

The organic compounds used were acetophenone (99%, Merck), n-heptane (99%, Sigma–Aldrich),

90

used as solvent, and n-decane (>98%, Panreac), used as internal standard in the analysis. Hydrogen

91

(99.9%, Praxair) and nitrogen (99.9%, Praxair) were provided in gas cylinders.

92

Catalysts used in this work were precious metals (Pd, Pt, Ru and Rh) supported on γ-Al2O3, provided

93

by BASF (formerly Engelhard). For all the catalysts, the metal loading is 0.5% (wt.) with an egg-shell

94

impregnation pattern.

95

The textural properties of the fresh and aged catalysts were measured by nitrogen adsorption at 77 K

96

in a MICROMERITICS ASAP 2020 surface area and porosity analyzer.

4 ACS Paragon Plus Environment

Page 5 of 30

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

Energy & Fuels

97

The aged catalysts were also characterized by temperature programed oxidization (TPO) in a

98

MICROMERITICS TPD/TPR-2900 device, according to the procedure outlined in previous works44. The

99

sample (10·10-6 kg) was introduced in a U-shaped quartz tube with a flow rate of 1.5·10-6 m3/s of 2%

100

O2 (He balance). The temperature was increased at 5ºC/min (up to 900ºC) and effluent was analyzed

101

on-line in a PFEIFFER VACCUM 300 mass spectrometer.

102

103

2.2. Experimental device

104

The experimental device for the reaction experiments is depicted in Figure 1. Hydrogen and nitrogen

105

flow rates were set by BRONKHORST mass flow controllers (respectively, FIC-01 and 02). Nitrogen

106

was used for inertization and for leak testing in the apparatus. During reaction experiments, pure

107

hydrogen (flow rate 0-5·10-5 m3/s n.t.p) was used as gas feed. The organic feed (acetophenone

108

dissolved in n-heptane) was pumped by a piston pump (ALLTECH model 452, flow rate 0-5·10-8 m3/s).

109

Both streams were mixed in a “T-piece” in the reactor feed line, as shown in Figure 1. The reactor

110

consisted of a stainless steel tube with 12.7·10-3 m internal diameter and 0.600 m length, surrounded

111

by a temperature-controlled electrical oven. The catalyst sample (0.188-0.25·10-3 kg), ground and

112

sieved to 100-250 µm, was mixed with 1·10-3 kg of glass particles of 350-710 µm diameter and placed

113

inside the reactor tube. The catalyst bed height (8.5-9.3·10-3 m) was high enough to assume plug flow

114

behavior (bed height/particle size > 50). The resulting packed bed was hold in position with a

115

stainless steel mesh of 60 µm. Upstream the packed bed, the reactor tube was filled with glass

116

spheres with 10-3 m diameter, in order to ensure a flat radial velocity profile (plug flow) and pre-heat

117

the feed. In all the experiments, the reaction mixture was in gas phase at reaction conditions.

118

The reactor effluent was cooled down and conducted to the sampling section, formed by two

119

cylinders connected in parallel that were used alternatively to collect the samples. Thus, the

120

condensate organic liquid was separated from the gas and accumulated in one of the cylinders; when

5 ACS Paragon Plus Environment

Energy & Fuels

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 30

121

the amount of liquid accumulated was adequate, the valves were switched in order to accumulate

122

liquid in the other cylinder. Then, the first cylinder (now off-line) was emptied through an

123

AUTOCLAVE ENGINEERS valve situated in the bottom. When the second cylinder is ready to be

124

sampled, the valves are switched another time. Hence, samples are collected alternatively from the

125

cylinders. This way, the decrease in pressure during the sampling does not affect the reactor,

126

because it is local to the cylinder. The reactor outlet is always connected to a backpressure regulator

127

through the gas outlet of the corresponding cylinder. Pressure in the reactor is maintained constant

128

at the set value (in the range 0.5-1.5 MPa).

129

Liquid samples were analyzed by gas chromatography. A Shimadzu GCMS-QP2010 Ultra gas

130

chromatograph equipped with HP-5MS capillary column and a mass spectrometer detector was used

131

to identify the components in the sample. Quantitative concentration measurements were done with

132

a Shimadzu GC-17A chromatograph, equipped with a HP-5 capillary column and a flame ionization

133

detector (FID), using n-decane as internal standard.

134

135

2.3. Reaction experiments

136

Catalyst stability and reaction kinetics were studied in this work. Catalyst stability studies were

137

carried out running the reaction for long periods of time (60 h) at constant operating conditions,

138

taking effluent samples every 1-2 h. Kinetic experiments were performed running the reaction at

139

different space times (W/F in the range 0-1.5 kgcat s/moltot) and constant pressure and temperature

140

during the period of constant catalytic activity. For this purpose, the kinetic experiments were done

141

with a new catalyst sample, which was initially stabilized, following the same procedure as in the

142

stability study, until constant conversion was obtained (40 h). Then, flow rates were varied randomly

143

(and hence space time), repeating the reference space time (0.75 kgcat s/moltot) every three flow

144

rates, in order to monitor any possible change in the catalyst activity during the kinetic test.

6 ACS Paragon Plus Environment

Page 7 of 30

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

Energy & Fuels

145

In both types of studies, the liquid feed consisted of a solution of 1000 mol/m3 acetophenone in n-

146

heptane. This solvent has an appropriate boiling point that ensures complete vaporization at reaction

147

conditions and almost quantitative condensation in the sampling cylinders. The blank experiments, in

148

which n-heptane with no acetophenone was fed to the reactor, showed negligible n-heptane

149

conversion at reaction conditions, which is very important to study the HDO of acetophenone.

150

In the kinetic analysis, the catalytic bed has been modelled as an ideal plug flow reactor, as shown in

151

eq. (1).  =   /

(1)

152

Where  is the molar concentration of compound (acetophenone or reaction products),  is the

153

reaction rate per catalyst weight, / is the space time,  is the weight of catalyst in the reactor, 

154

is the feed total molar flow rate, and  is the feed total molar density (calculated at the reaction

155

temperature and pressure, using the ideal gas law).

156

Kinetic parameters have been calculated by fitting eq. (1) to the experimental results by the least-

157

square method. The concentrations of the different compounds have been appropriately scaled in

158

the objective function. Calculations have been carried out with the help of MATLAB using lsqnonlin

159

(trust-region-reflective algorithm), ode15 sand lnparci functions, respectively, to solve the least-

160

square problem, the set of differential equations and determine the confidence intervals.

161

162

3. Results and Discussion

163

3.1. Selection of precious metal catalyst

164

The performance (activity and stability) of different precious-metal alumina-supported catalysts for

165

the gas phase hydrodeoxygenation of acetophenone has been studied in a fixed-bed reactor. As

7 ACS Paragon Plus Environment

Energy & Fuels

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 30

166

described in the experimental section, experiments consisted of reactions carried out for 60 h at

167

constant operating conditions: temperature 325ºC, pressure 0.5 MPa, space time W/F = 1.00 kgcat

168

s/moltot and feed hydrogen to oxygen in acetophenone molar ratio, H2/O = 23. Results are depicted in

169

Figure 2 in terms of acetophenone conversion, defined as X = 1-CA/CA0 (where CA is concentration of

170

acetophenone and sub-index 0 indicates at reactor inlet). As observed, all the catalysts deactivate

171

markedly during the first 15 h of reaction. Conversion for Pd, Ru and Rh decreases from 25-50%

172

initially to 1-5% by the end of the experiment, while for Pt, conversion decreases from the initial 97%

173

to around 30% at 40 h on stream, remaining then nearly constant till the end of the experiment.

174

Analysis of the reactor effluent revealed that the main reaction products were ethylbenzene and

175

styrene. The presence of these compounds is in agreement with the reaction scheme proposed for

176

this reaction in the literature, Figure 3.31, 39, 40 Precious metals catalyze the hydrogenation of

177

acetophenone to 1-phenyl ethanol, then alcohol dehydration takes place in the presence of an acid

178

catalyst (the Al2O3 support), and finally styrene is hydrogenated to ethylbenzene. 1-Phenyl ethanol

179

was not identified among the reaction products, because the alcohol dehydration is very fast in

180

comparison with the other reactions.32 Hydrogenation of the aromatic ring is also possible according

181

to the reaction scheme, but the corresponding reactions products were not identified at our

182

experimental conditions.

183

Catalysts tested can be classified in two groups, according to the product selectivity of the aged

184

catalysts (after 55 h on stream). Selectivity to product “i” is defined as Si = Ci/(CA0-CA), where Ci is

185

concentration of product “i”, CA is concentration of acetophenone, and sub-index 0 indicates at

186

reactor inlet. On one hand, Pt and Pd exhibit 80-85 % selectivity to ethylbenzene, 12-14% selectivity

187

to styrene, and very small selectivities towards benzene and toluene. On the other hand, Ru and Rh

188

show lower ethylbenzene selectivity (41-58%), slightly higher styrene selectivity (19-20%), and an

189

important selectivity towards benzene (10-23%), and toluene (10-15%). Benzene and toluene are

190

associated to hydrogenolysis of C-C bonds in styrene and ethylbenzene products. These cracking

8 ACS Paragon Plus Environment

Page 9 of 30

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

Energy & Fuels

191

products can be formed through a bifunctional mechanism which requires both acid and metallic

192

centers on the catalyst surface, as evidenced in studies regarding ethylbenzene hydrogenation and

193

dehydrogenation.45, 46 It has been checked that most reaction products were identified and

194

quantified by means of the total mass balance, with a maximum error of 6% for Pd. Concerning to

195

the support, it must be noted that alumina has a low-medium strength acidity, which present a good

196

balance between the acid sites needed for the main reactions and the low concentration of the

197

strong acids leading to by-products by C-C bond cleavage.47,48

198

The aged catalysts have been characterized by nitrogen adsorption and temperature programmed

199

oxidation (TPO). The BET surface area of the aged catalysts is lower than the corresponding surface

200

area of the fresh catalysts. The surface area decreased from 82 to 72 m2/g for Pd/Al2O3, from 106 to

201

104 m2/g for Ru/Al2O3, from 107 to 98 m2/g for Rh/Al2O3 and from 87 to 75 m2/g for Pt/Al2O3. In

202

general terms, a decrease in the surface area is observed, although it is not conclusive. This fact is

203

expected for mesoporous catalysts where porous structure is unlikely to be blocked by the

204

carbonaceous deposits.

205

The CO2 release profiles obtained for the TPO of the spent catalysts, shown in Figure 4, provide more

206

evidences on coke formation. On increasing temperature, carbon dioxide evolves from the catalysts

207

in the form of two marked peaks. The high temperature peak, generated in all the catalysts at 520-

208

540ºC, suggests the presence of carbon deposits on the catalyst surface. The low temperature peak

209

appears at different temperatures depending on the catalyst: for Pt and Pd at around 240ºC, while

210

for Ru and Rh at 350ºC. The low temperature peaks can be associated to the presence of adsorbed

211

heavier products, which may differ on nature for the different catalysts, and hence the difference

212

oxidation temperature observed. It should be noted that there is a correlation between the low

213

temperature peak and the product distribution, as discussed previously for the different catalysts.

214

Thus, Ru and Rh catalysts presented important selectivity towards benzene and toluene, which was

215

negligible for Pd and Pt catalysts. Also the selectivity towards styrene was higher. In fact, styrene is

9 ACS Paragon Plus Environment

Energy & Fuels

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 30

216

known to generate coke precursors (in hydrogen media) that cause catalyst deactivation and would

217

explain the observed behavior.45, 49 For example, in the commercial process of styrene production

218

from dehydrogenation of ethylbenzene, the catalyst deactivation is controlled using an excess of

219

steam that reacts with the coke precursors. In the present case, acetophenone hydrodeoxygenation,

220

steam is generated as the product of deoxygenation, but the concentration is low.

221

The adsorbed products explain the deactivation of the catalysts observed in the stability test. The

222

size of the peaks is higher for Ru and Rh, which are the catalysts with faster deactivation. On the

223

contrary, Pt, which was found to be stable after 40 h of stabilization, presents the smallest peaks.

224

225

3.2. Optimization of operating conditions for Pt/Al2O3

226

In the previous section, it was determined that Pt/Al2O3 is the most stable and active catalyst for

227

acetophenone HDO in gas phase. Now, the influence of operating conditions (pressure and

228

temperature) in the process with this catalyst is studied. For this purpose, stability experiments with

229

60 h duration, and using fresh catalyst in each one, have been carried out.

230

The influence of pressure in the range 0.5-1.5 MPa on the catalyst stability has been determined at

231

325ºC, H2/O molar ratio 23 and a space time 1.00 kgcat s/moltot. Results are summarized in Figure 5a

232

and Table 1. As observed, initial acetophenone conversion was high (95-97%) in all experiments.

233

Acetophenone conversion decreases upon time, very markedly in the experiment at 0.5 MPa, and

234

more gradually at 1.0 and 1.5 MPa. For 1.0 and 1.5 MPa, the behavior is very similar, with conversion

235

at the end of experiment (t > 45 h) around 64%.

236

Regarding product distribution, the main product was ethylbenzene, with selectivity 85-92% for all

237

pressures. Styrene was only detected at 0.5 MPa, with 12% selectivity at 55 h. Styrene is an

238

intermediate in the hydrodeoxygenation of acetophenone to ethylbenzene, and hence it is only

239

detected at low conversion. Benzene and toluene, generated by hydrogenolysis of C-C bonds as

10 ACS Paragon Plus Environment

Page 11 of 30

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

Energy & Fuels

240

explained in section 3.1, were detected at all pressures, with similar low selectivities (1-3%).

241

Ethylcyclohexane was detected at the higher pressures, with selectivity increasing as pressure

242

increases (1% selectivity at 1.0 MPa and 4% at 1.5 MPa). Ethylcyclohexane is produced by

243

hydrogenation of the aromatic ring, favored at high pressure. Hydrogenation of the aromatic ring is

244

usually undesired in bio-oil de-oxygenation, as it consumes valuable hydrogen. For this reason, and

245

given the similar acetophenone conversions obtained at 1.0 and 1.5 MPa, the recommended

246

operating pressure in the range studied is 1.0 MPa.

247

The influence of temperature on the hydrodeoxygenation of acetophenone was studied at 275 and

248

375ºC for pressure 1.0 MPa (found optimal, as explained previously), H2/O molar ratio 23 and space

249

time 0.75 kgcat s/moltot. It should be noted that in this set of experiments, space time is 25% lower

250

than in the pressure dependence experiments. This value was selected in order to avoid too high

251

conversions, which could mask catalyst deactivation. Results in Figure 5b indicate that acetophenone

252

conversion decreases upon time during the first 20 h, and then remains approximately constant for

253

both temperatures. At 375ºC, conversion decreases faster than at 275ºC. Conversion achieved after

254

55 h is slightly higher at 375ºC: 27% at 275ºC and 36% at 375ºC.

255

Results of products distribution are summarized in Table 1. At 275ºC, the main product is

256

ethylbenzene with 83% selectivity, followed by cyclohexyl methyl ketone with 7% selectivity and

257

ethylcyclohexane with 1% selectivity. The presence of cyclohexyl methyl ketone evidences that at

258

275ºC the hydrogenation the aromatic ring of acetophenone is hydrogenated to some extent.33 This

259

compound was not detected at 375ºC, neither at 325ºC (in the previous set of experiments).

260

Formation of cyclohexyl methyl ketone is undesired from the point of view of bio-oil upgrading,

261

because the oxygen content of the molecule does not decrease, and its formation is hydrogen

262

consuming. Ethyl cyclohexane found among the reaction products can be produced by

263

hydrogenation of ethylbenzene or hydrogenation followed by dehydration-hydrogenation of the

264

resulting cyclohexyl methyl ketone, as shown in the reaction scheme of Figure 3. At 375ºC, ethyl

11 ACS Paragon Plus Environment

Energy & Fuels

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 30

265

benzene is also the main product, with 80% selectivity, but styrene is formed with 11% selectivity. As

266

explained before, styrene is a reaction intermediate in the formation of ethylbenzene.

267

As shown in Table 1, the presence of an important selectivity towards styrene (experiment at 0.5

268

MPa and 325ºC, and experiment 1.0 MPa and 375ºC) is correlated to a stronger loss of activity during

269

the catalyst stabilization experiment. As explained in section 3.1, the characterization of the aged

270

catalysts revealed that the presence of high selectivity towards styrene is correlated to a higher

271

carbon dioxide emission during the temperature programmed oxidation tests. Thus, styrene is known

272

to be responsible of the formation of coke precursors at hydrogen media. The adsorption of these

273

higher molecular weight products is the most likely responsible of observed catalyst deactivation.

274

275

3.3. Reaction kinetics

276

The kinetics of the hydrodeoxygenation of acetophenone on the Pt/Al2O3 catalyst has been studied

277

at 1.0 MPa, H2/O molar ratio 23, and 275 and 375ºC. Hydrogen/oxygen ratio is high enough as to

278

assume constant H2 concentration inside the reactor. Experiments have been done using 0.1875·10-3

279

kg of catalyst, previously aged for 40 h at the corresponding temperature. This way, the catalyst

280

activity is constant during the experiments, as previously found in the stability tests. Space velocities

281

ranged from 0 to 1.5 kgcat s/moltot, obtained by varying the gas and liquid flow rates while

282

maintaining constant the H2/O ratio. Experiments have been done in absence of diffusional

283

limitations, as indicated below.

284

Experimental results are depicted as symbols in Figure 6 and 7. Product distributions obtained agree

285

with the reaction scheme in Figure 3. Cyclohexyl methyl ketone is a primary product from

286

acetophenone, because important amounts are formed at low space times, when acetophenone

287

conversion is low. On the contrary, the concentration of ethyl cyclohexane is very low at low space

288

times, and increases only when acetophenone conversion increases at higher space times, which

12 ACS Paragon Plus Environment

Page 13 of 30

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

Energy & Fuels

289

suggests that it is a secondary product. Styrene was only detected at 375ºC, with very low

290

concentration, that first increases and then decreases as space time increases, behavior expected for

291

an intermediate product. Finally, the main product, ethylbenzene, behaves as an end-product, with

292

concentration increasing with space time. Benzene and toluene have been excluded from this

293

analysis, because their selectivity is always very low (< 1%). According to these results, at 275ºC the

294

kinetically relevant steps are reactions 1, 4 and 5 (reactions 2, 3 and 6 are very fast, so the products

295

are not detected, and reactions 3 and 8 very slow). At 375ºC, the kinetically relevant steps are

296

reactions 1 and 3, since hydrogenation of the aromatic ring does not take place.

297

Kinetic models, based on the reaction scheme in Figure 3, have been fitted to the experimental

298

results. The kinetic models consider only the reactions kinetically relevant at the corresponding

299

temperature, as discussed before. The reactions are supposed pseudo-first order with respect to the

300

corresponding organic reactant, as proposed in the literature.43 Hydrogen concentration is the same

301

in the different kinetic tests, and in great excess. Hence, it is found constant along the reactor length,

302

and its influence on the reaction kinetics can be included in the apparent rate constants. As

303

mentioned in section 2.3, the catalytic bed has been modelled as a plug flow reactor. Results of the

304

least-square fitting of the models to the experimental data are summarized in Table 2. The regression

305

coefficients obtained for the fittings are R2 = 0.98 and 0.97, respectively, for 275ºC and 375ºC.

306

Model predictions for 275ºC are depicted as lines in Figure 6, showing that the model fits fairly well

307

the experimental data. The reaction step with the higher rate constant (1.7·10-3 m3/kgcat s) is reaction

308

1, hydrodeoxygenation of the acetophenone carbonyl group.

309

At 375ºC, the model also fits the experiment data well, see Figure 7. At this temperature, reaction 1 is

310

faster than at 275ºC (kinetic constant 3.7·10-3 m3/kgcat s), and for this reason reaction 3 is kinetically

311

relevant and styrene was detected in the reaction products. Hydrogenation of styrene (rate constant

312

130·10-3 m3/kgcat s) is much faster than hydrodeoxygenation of acetophenone, hence the

313

concentration of styrene is maintained very low.

13 ACS Paragon Plus Environment

Energy & Fuels

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 30

314

The experiments were performed in absence of diffusional limitations. This has been confirmed by

315

calculating the intraparticle effectiveness factor and Carberry number at the worst conditions, i.e.

316

highest reaction rate (highest temperature, feed concentration and lowest flow rate). Physical

317

properties were evaluated at these conditions using correlations from the literature.50 The effective

318

diffusion coefficient in the porous catalyst particles was estimated as a contribution of molecular and

319

Knudsen diffusion. The values obtained are 0.91 for the intraparticle effectiveness factor and 0.0022

320

< 0.05 for the Carberry number. The isothermicity of the system at the catalyst particle level was also

321

evaluated. The maximum temperature difference inside the catalyst particle is estimated as 0.28ºC

322

(considering the worst case scenario, corresponding to complete conversion), and the maximum

323

temperature difference in the gas film outside the catalyst particle is 0.04ºC.

324

325

4. Conclusions

326

Hydrodeoxygenation of oxygenated aromatics present in biomass-derived bio-oils, to produce de-

327

oxygenated compounds is of interest for the upgrading of biofuels. Acetophenone is an interesting

328

model compound for the study of catalysts activity and optimization of operating conditions, because

329

of the carbonyl group and aromatic ring are chemical structures representative of most bio-oils.

330

Thus, the reactivity of the carbonyl group and aromatic ring in acetophenone leads to a rather

331

complex reaction scheme with competitive parallel and series steps. Experiments of acetophenone

332

hydrodeoxigenation, carried out in gas phase in a fixed-bed reactor with different precious metal

333

catalysts and operating conditions lead to the following conclusions:

334

-

Pt/Al2O3 was found the most active and stable catalyst at 0.5 MPa and 325ºC, the main

335

reaction products being ethylbenzene and styrene. Pd, Ru and Rh on Al2O3 catalysts

336

deactivates drastically in the first 15 h.

14 ACS Paragon Plus Environment

Page 15 of 30

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

Energy & Fuels

337

-

In the range 0.5-1.5 MPa and at 325ºC, the optimum pressure for hydrodoexygenation of

338

acetophenone on Pt/Al2O3 was 1.0 MPa. At 0.5 MPa, the activity and stability of the catalyst

339

is lower, while at 1.5 MPa the selectivity towards ethylcyclohexane increases.

340

-

341

Operation at 375ºC (Pt/Al2O3 catalyst, 1.0 MPa) is preferred in order to avoid hydrogenation of the aromatic ring, observed at 275ºC. In addition, at 375ºC the reaction is faster.

342

-

The reaction kinetics for the Pt/Al2O3 catalyst at 1.0 MPa and temperature 275ºC and 375ºC

343

fits well to a pseudo-first order kinetic model based on the kinetically relevant steps of the

344

reaction scheme.

345

346

Acknowledgements

347

This work was financed by the Asturian Local Government (Spain, project reference GRUPIN14-078).

348

C. González also thanks the Asturian Local Government for a Ph.D Grant (Severo Ochoa Program).

349

350

List of abbreviations

351

B

352

CHMK Cyclohexyl methyl ketone

353

EB

Ethylbenzene

354

ECH

Ethylcyclohexane

355

S

Styrene

356

T

Toluene

Benzene

357

15 ACS Paragon Plus Environment

Energy & Fuels

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 30

358

References

359 360 361

1. Arun, N.; Sharma, R. V.; Dalai, A. K., Green diesel synthesis by hydrodeoxygenation of bio-based feedstocks: Strategies for catalyst design and development. Renewable and Sustainable Energy Reviews 2015, 48, (0), 240-255.

362 363

2. Guo, Y. W., Y. Wei, F., Research progress in biomass flash pyrolysis technology for liquids production. Chemical industry engineering progress 2001, 8, 13-17.

364 365

3. French, R.; Czernik, S., Catalytic pyrolysis of biomass for biofuels production. Fuel Processing Technology 2010, 91, (1), 25-32.

366 367 368

4. Moraes, M. S. A.; Georges, F.; Almeida, S. R.; Damasceno, F. C.; Maciel, G. P. d. S.; Zini, C. A.; Jacques, R. A.; Caramão, E. B., Analysis of products from pyrolysis of Brazilian sugar cane straw. Fuel Processing Technology 2012, 101, (0), 35-43.

369 370

5. Zhang, Q.; Chang, J.; Wang, T.; Xu, Y., Review of biomass pyrolysis oil properties and upgrading research. Energy Conversion and Management 2007, 48, (1), 87-92.

371 372

6. Zhang, L.; Liu, R.; Yin, R.; Mei, Y., Upgrading of bio-oil from biomass fast pyrolysis in China: A review. Renewable and Sustainable Energy Reviews 2013, 24, (0), 66-72.

373 374 375

7. Imran, A.; Bramer, E. A.; Seshan, K.; Brem, G., High quality bio-oil from catalytic flash pyrolysis of lignocellulosic biomass over alumina-supported sodium carbonate. Fuel Processing Technology 2014, 127, (0), 72-79.

376 377

8. Choudhary, T. V.; Phillips, C. B., Renewable fuels via catalytic hydrodeoxygenation. Applied Catalysis A: General 2011, 397, (1–2), 1-12.

378 379

9. Adjaye, J. D.; Bakhshi, N. N., Production of hydrocarbons by catalytic upgrading of a fast pyrolysis bio-oil. Part I: Conversion over various catalysts. Fuel Processing Technology 1995, 45, (3), 161-183.

380

10. Furimsky, E., Catalytic hydrodeoxygenation. Applied Catalysis A: General 2000, 199, (2), 147-190.

381 382

11. Elliott, D. C., Historical Developments in Hydroprocessing Bio-oils. Energy & Fuels 2007, 21, (3), 1792-1815.

383 384

12. French, R. J.; Hrdlicka, J.; Baldwin, R., Mild hydrotreating of biomass pyrolysis oils to produce a suitable refinery feedstock. Environmental Progress and Sustainable Energy 2010, 29, (2), 142-150.

385 386 387

13. Grilc, M.; Likozar, B.; Levec, J., Hydrotreatment of solvolytically liquefied lignocellulosic biomass over NiMo/Al2O3 catalyst: Reaction mechanism, hydrodeoxygenation kinetics and mass transfer model based on FTIR. Biomass and Bioenergy 2014, 63, 300-312.

388 389 390

14. Grilc, M.; Likozar, B.; Levec, J., Hydrodeoxygenation and hydrocracking of solvolysed lignocellulosic biomass by oxide, reduced and sulphide form of NiMo, Ni, Mo and Pd catalysts. Applied Catalysis B: Environmental 2014, 150–151, 275-287.

391 392 393

15. Grilc, M.; Veryasov, G.; Likozar, B.; Jesih, A.; Levec, J., Hydrodeoxygenation of solvolysed lignocellulosic biomass by unsupported MoS2, MoO2, Mo2C and WS2 catalysts. Applied Catalysis B: Environmental 2015, 163, 467-477.

16 ACS Paragon Plus Environment

Page 17 of 30

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

Energy & Fuels

394 395 396

16. Brillouet, S.; Baltag, E.; Brunet, S.; Richard, F., Deoxygenation of decanoic acid and its main intermediates over unpromoted and promoted sulfided catalysts. Applied Catalysis B: Environmental 2014, 148–149, 201-211.

397 398 399

17. Bykova, M. V.; Ermakov, D. Y.; Kaichev, V. V.; Bulavchenko, O. A.; Saraev, A. A.; Lebedev, M. Y.; Yakovlev, V. А., Ni-based sol–gel catalysts as promising systems for crude bio-oil upgrading: Guaiacol hydrodeoxygenation study. Applied Catalysis B: Environmental 2012, 113–114, 296-307.

400 401 402

18. Fisk, C. A.; Morgan, T.; Ji, Y.; Crocker, M.; Crofcheck, C.; Lewis, S. A., Bio-oil upgrading over platinum catalysts using in situ generated hydrogen. Applied Catalysis A: General 2009, 358, (2), 150156.

403 404 405

19. de Wild, P.; Van der Laan, R.; Kloekhorst, A.; Heeres, E., Lignin valorisation for chemicals and (transportation) fuels via (catalytic) pyrolysis and hydrodeoxygenation. Environmental Progress & Sustainable Energy 2009, 28, (3), 461-469.

406 407

20. Wildschut, J.; Melián-Cabrera, I.; Heeres, H. J., Catalyst studies on the hydrotreatment of fast pyrolysis oil. Applied Catalysis B: Environmental 2010, 99, (1–2), 298-306.

408 409 410

21. Wildschut, J.; Iqbal, M.; Mahfud, F. H.; Cabrera, I. M.; Venderbosch, R. H.; Heeres, H. J., Insights in the hydrotreatment of fast pyrolysis oil using a ruthenium on carbon catalyst. Energy & Environmental Science 2010, 3, (7), 962-970.

411 412 413

22. Chen, C.; Chen, G.; Yang, F.; Wang, H.; Han, J.; Ge, Q.; Zhu, X., Vapor phase hydrodeoxygenation and hydrogenation of m-cresol on silica supported Ni, Pd and Pt catalysts. Chemical Engineering Science 2015, 135, 145-154.

414 415

23. Yao, G.; Wu, G.; Dai, W.; Guan, N.; Li, L., Hydrodeoxygenation of lignin-derived phenolic compounds over bi-functional Ru/H-Beta under mild conditions. Fuel 2015, 150, 175-183.

416 417 418

24. Valle, B.; Gayubo, A. G.; Aguayo, A. T.; Olazar, M.; Bilbao, J., Selective Production of Aromatics by Crude Bio-oil Valorization with a Nickel-Modified HZSM-5 Zeolite Catalyst. Energy & Fuels 2010, 24, (3), 2060-2070.

419 420 421

25. Yang, H.-M.; Zhao, W.; Norinaga, K.; Fang, J.-J.; Wang, Y.-G.; Zong, Z.-M.; Wei, X.-Y., Separation of phenols and ketones from bio-oil produced from ethanolysis of wheat stalk. Separation and Purification Technology 2015, 152, 238-245.

422 423 424

26. Undri, A.; Abou-Zaid, M.; Briens, C.; Berruti, F.; Rosi, L.; Bartoli, M.; Frediani, M.; Frediani, P., A simple procedure for chromatographic analysis of bio-oils from pyrolysis. Journal of Analytical and Applied Pyrolysis 2015, 114, 208-221.

425 426

27. Demirbas, A., The influence of temperature on the yields of compounds existing in bio-oils obtained from biomass samples via pyrolysis. Fuel Processing Technology 2007, 88, (6), 591-597.

427 428 429

28. Huang, J.; Jiang, Y.; Van Vegten, N.; Hunger, M.; Baiker, A., Tuning the support acidity of flamemade Pd/SiO2-Al 2O3 catalysts for chemoselective hydrogenation. Journal of Catalysis 2011, 281, (2), 352-360.

430 431

29. Lin, S. D.; Sanders, D. K.; Albert Vannice, M., Influence of metal-support effects on acetophenone hydrogenation over platinum. Applied Catalysis A: General 1994, 113, (1), 59-73.

17 ACS Paragon Plus Environment

Energy & Fuels

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

Page 18 of 30

432 433 434

30. Schimmoeller, B.; Hoxha, F.; Mallat, T.; Krumeich, F.; Pratsinis, S. E.; Baiker, A., Fine tuning the surface acid/base properties of single step flame-made Pt/alumina. Applied Catalysis A: General 2010, 374, (1–2), 48-57.

435 436

31. Chen, C.-S.; Chen, H.-W.; Cheng, W.-H., Study of selective hydrogenation of acetophenone on Pt/SiO2. Applied Catalysis A: General 2003, 248, (1–2), 117-128.

437 438

32. Alharbi, K.; Kozhevnikova, E. F.; Kozhevnikov, I. V., Hydrogenation of ketones over bifunctional Ptheteropoly acid catalyst in the gas phase. Applied Catalysis A: General, (0).

439 440 441

33. Chen, M.; Maeda, N.; Baiker, A.; Huang, J., Molecular insight into Pt-catalyzed chemoselective hydrogenation of an aromatic ketone by in situ modulation-excitation IR spectroscopy. ACS Catalysis 2012, 2, (9), 2007-2013.

442 443 444

34. Jiang, Y.; Büchel, R.; Huang, J.; Krumeich, F.; Pratsinis, S. E.; Baiker, A., Solvent-free Hydrodeoxygenation of Bio-oil Model Compounds Cyclopentanone and Acetophenone over Flamemade Bimetallic Pt-Pd/ZrO(2) Catalysts. ChemSusChem 2012, 5, (7), 1190-1194.

445 446

35. Chen, C.-S.; Chen, H.-W., Enhanced selectivity and formation of ethylbenzene for acetophenone hydrogenation by adsorbed oxygen on Pd/SiO2. Applied Catalysis A: General 2004, 260, (2), 207-213.

447 448 449 450

36. Wang, Y.; Su, N.; Ye, L.; Ren, Y.; Chen, X.; Du, Y.; Li, Z.; Yue, B.; Tsang, S. C. E.; Chen, Q.; He, H., Tuning enantioselectivity in asymmetric hydrogenation of acetophenone and its derivatives via confinement effect over free-standing mesoporous palladium network catalysts. Journal of Catalysis 2014, 313, (0), 113-126.

451 452 453

37. Xiang, Y.-z.; Lv, Y.-a.; Xu, T.-y.; Li, X.-n.; Wang, J.-g., Selectivity difference between hydrogenation of acetophenone over CNTs and ACs supported Pd catalysts. Journal of Molecular Catalysis A: Chemical 2011, 351, (0), 70-75.

454 455

38. Sumner, C., Jr.; Burchett, W., Developments in the Pd Catalyzed Hydrogenation of Oxygenated Organic Compounds. Topics in Catalysis 2012, 55, (7-10), 480-485.

456 457 458

39. Bergault, I.; Fouilloux, P.; Joly-Vuillemin, C.; Delmas, H., Kinetics and Intraparticle Diffusion Modelling of a Complex Multistep Reaction: Hydrogenation of Acetophenone over a Rhodium Catalyst. Journal of Catalysis 1998, 175, (2), 328-337.

459 460 461

40. Malyala, R. V.; Rode, C. V.; Arai, M.; Hegde, S. G.; Chaudhari, R. V., Activity, selectivity and stability of Ni and bimetallic Ni–Pt supported on zeolite Y catalysts for hydrogenation of acetophenone and its substituted derivatives. Applied Catalysis A: General 2000, 193, (1–2), 71-86.

462 463 464

41. Rajashekharam, M. V.; Bergault, I.; Fouilloux, P.; Schweich, D.; Delmas, H.; Chaudhari, R. V., Hydrogenation of acetophenone using a 10% Ni supported on zeolite Y catalyst: kinetics and reaction mechanism. Catalysis Today 1999, 48, (1–4), 83-92.

465 466

42. Wang, W.; Yang, Y.; Luo, H.; Hu, T.; Liu, W., Amorphous Co–Mo–B catalyst with high activity for the hydrodeoxygenation of bio-oil. Catalysis Communications 2011, 12, (6), 436-440.

467 468 469

43. Trasarti, A. F.; Bertero, N. M.; Apesteguía, C. R.; Marchi, A. J., Liquid-phase hydrogenation of acetophenone over silica-supported Ni, Co and Cu catalysts: Influence of metal and solvent. Applied Catalysis A: General 2014, 475, (0), 282-291.

18 ACS Paragon Plus Environment

Page 19 of 30

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

Energy & Fuels

470 471 472

44. Ordóñez, S.; Sastre, H.; Díez; F.V.; Thermogravimetric determination of coke deposits on alumina-supported noble metal catalysts used as hydrodechlorination catalysts. Thermochimica Acta 2001, 379, 25-34.

473 474

45. Cavani, F.; Trifirò, F., Alternative processes for the production of styrene. Applied Catalysis A: General 1995, 133, (2), 219-239.

475 476 477

46. Seoane, X. L.; Arcoya, A.; Gonzalez, J. A.; Travieso, N., Hydrogenation of ethylbenzene over a nickel/mordenite catalyst. Catalytic decay by thiophene poisoning. Industrial & Engineering Chemistry Research 1989, 28, (3), 260-264.

478 479 480

47. Faba, L.; Díaz E.; Ordóñez, S.; Role of the support on the performance and stability of Pt-based catalysts for furfural-acetone adduct hydrodeoxygenation. Catalysis Science and Technology 2015, 5, 1473-1484.

481 482 483

48. Faba, L.; Díaz E.; Ordóñez, S.; Hydrodeoxygenation of acetone-furfural condensation adducts over alumina-supported noble metal catalysts. Applied Catalysis B: Environmental. 2014, 160-161, 436444.

484 485 486

49.Irún, O.; Sadosche, S. A.; Lasobras, J.; Soler, J.; Francés, E.; Herguido, J.; Menéndez, M., Catalysts for the production of styrene from ethylbenzene: Redox and deactivation study. Catalysis Today 2013, 203, (0), 53-59.

487 488

50. Polling, B. E.; Prausnitz, J. M.; O'Connell, J. P., The properties of gases and liquids. 5th ed.; McGraw-Hill: Boston MA, 2001.

489 490

19 ACS Paragon Plus Environment

Energy & Fuels

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

Page 20 of 30

491 492

Caption to figures

493 494

Figure 1

Diagram of the experimental device.

495

Figure 2

Stability of alumina supported precious metal catalysts for the hydrodeoxygenation

496

of acetophenone: Pd (), Ru (), Rh () and Pt (). Conditions: 325ºC, 0.5 MPa,

497

H2/O = 23.2 molar, W/F =1.00 kgcat s/moltot.

498

Figure 3

General scheme of the hydrodeoxygenation of acetophenone.

499

Figure 4

Temperature programmed oxidation (TPO) of aged catalysts (temperature ramp

500 501

5ºC/min): Pd (▬), Ru (▬), Rh (▬) and Pt (▬). Figure 5

Stability of Pt/Al2O3 catalyst for the hydrodeoxygenation of acetophenone.

502

a) Influence of pressure: 0.5 MPa (), 1.0 MPa () and 1.5 MPa (). Conditions:

503

325ºC, H2/O = 23.2 molar, W/F = 1.00 kgcat s/moltot.

504

b) Influence of temperature: 275ºC () and 375ºC (). Conditions: 1.0 MPa, H2/O =

505

23.2 molar, W/F = 0.75 kgcat s/moltot.

506

Figure 6

507 508 509

Results of the reaction kinetic experiments and fitting of the kinetic model for the Pt/Al2O3 catalyst at 275ºC, 1.0 MPa and H2/O = 23 molar.

Figure 7

Results of the reaction kinetic experiments and fitting of kinetic model for the Pt/Al2O3 catalyst at 375ºC, 1.0 MPa and H2/O = 23 molar.

510

20 ACS Paragon Plus Environment

Page 21 of 30

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

Energy & Fuels

511 512

List of tables

513 514

Table 1

Summary of acetophenone conversion and product selectivities for the Pt/Al2O3

515

catalyst at different operating conditions (X = conversion. S = selectivity at 55 h. Sub-

516

index indicates compounds: S = styrene, EB = ethylbenzene, ECH = ethylcyclohexane,

517

CHMK = cyclohexyl methyl ketone, B = benzene, T = toluene)

518 519

Table 2

Results of the fitting of the kinetic models for the Pt/Al2O3 catalyst at 1.0 MPa and

520

temperature 275ºC and 375ºC (Sub-indexes of the rate constants refers to the steps

521

of the reaction scheme in Figure 3. Sub-indexes of compounds: S = styrene, EB =

522

ethylbenzene, ECH = ethylcyclohexane, CHMK = cyclohexyl methyl ketone)

523 524

21 ACS Paragon Plus Environment

Energy & Fuels

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

Page 22 of 30

525 526

Table 1

527 p T (ºC) (MPa)

W/F (kg s/mol)

X1h

X55h

SS

SEB

SECH

SCHMK

SB

ST

Total

0.5

325

1.00

97%

30%

12%

85%

-

-

3%

-

100%

1.0

325

1.00

97%

64%

-

85%

1%

-

1.2%

0.3%

88%

1.5

325

1.00

95%

64%

-

92%

4%

-

1.5%

0.4%

98%

1.0

275

0.75

97%

27%

-

83%

1%

7%

-

-

91%

1.0

375

0.75

79%

36%

11%

80%

-

-

1%

1%

93%

528 529 530

22 ACS Paragon Plus Environment

Page 23 of 30

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

Energy & Fuels

531 532

Table 2

533 275ºC

375ºC

 = −  −  

 = − 

 =   −  

 =   −  

 =  

 =  

 =   Kinetic constants (m3/kgcat s) · 103 k1 = 1.7 ± 0.2

k1 = 3.7 ± 0.3

k4 = 0.22 ± 0.02

k3 = 130 ± 10

k5 = 1.4 ± 0.3 R2 = 0.98

R2 = 0.97

534 535

23 ACS Paragon Plus Environment

Energy & Fuels

536 537

Figure 1

Cylinder 2

538

Cylinder 1

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

Page 24 of 30

539 540 541 542

24 ACS Paragon Plus Environment

Page 25 of 30

543 544

Figure 2

545 100

Pd Ru Rh Pt

80

Conversion (%)

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

Energy & Fuels

60 40 20 0 0

546

10

20

30

40

50

60

Time-on-steam (h)

547 548

25 ACS Paragon Plus Environment

Energy & Fuels

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

Page 26 of 30

549 550

Figure 3

551 552 553 554

26 ACS Paragon Plus Environment

Page 27 of 30

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

Energy & Fuels

555 556

Figure 4

557

558 559 560 561 562

27 ACS Paragon Plus Environment

Energy & Fuels

563 564

Figure 5

565

100

Conversion (%)

80 60 40 0.5 MPa 1.0 MPa 1.5 MPa

20 0 0 566

10

20

30

40

50

60

Time-on-stream (h)

a) 100

275ºC

Conversion (%)

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

Page 28 of 30

80

375ºC

60 40 20 0 0

567

b)

10

20

30

40

50

60

Time-on-stream (h)

568 569 570 571

28 ACS Paragon Plus Environment

Page 29 of 30

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

Energy & Fuels

572 573

Figure 6

574

a)

575

b)

576

29 ACS Paragon Plus Environment

Energy & Fuels

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

Page 30 of 30

577 578

Figure 7

579

580 581 582 583 584

30 ACS Paragon Plus Environment