Sei sulla pagina 1di 10

G Model

JIEC 3577 1–10

Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

1 Catalytic conversion of glycerol to lactic acid over graphite-supported


2 nickel nanoparticles and reaction kinetics
3 Q1 Haixu Yin, Hengbo Yin* , Aili Wang* , Lingqin Shen
4 Faculty of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China

A R T I C L E I N F O A B S T R A C T

Article history:
Received 26 September 2016 Catalytic conversion of biomass glycerol to lactic acid is an alternative to the conventional fermentation
Received in revised form 8 August 2017 process starting from carbohydrate. Graphite-supported metallic Ni0 nanoparticles effectively catalyzed
Accepted 14 August 2017 the hydrothermal conversion of glycerol to lactic acid in NaOH aqueous solution. Ni0 nanoparticle and
Available online xxx NaOH synergistically catalyzed the glycerol conversion to lactic acid. When the reaction was carried out
with the initial glycerol and NaOH concentrations of 1.0 and 1.1 mol L 1 at 230  C for 3 h over Ni0.3/
Keywords: graphite catalyst, the selectivity of lactic acid was 92.2% at the glycerol conversion of 97.6%. The reaction
Glycerol activation energy, Ea, was 69.2 kJ mol L 1.
Lactic acid
© 2017 Published by Elsevier B.V. on behalf of The Korean Society of Industrial and Engineering
Graphite
Chemistry.
Nickel nanoparticles
Reaction kinetics

5 Introduction conventional petroleum-based poly-ethylene terephthalate plas- 29

tics due to its biocompatibility and biodegradability [14]. The 30


6 Biodiesel produced from biomass has been used as a sustain- production capacity of polylactic acid is estimated to be 830 kt in 31
7 able energy source to replace fossil fuel because the utility of 2020 [16]. 32
8 biodiesel can avoid natural resource depletion and solve green- Commercial lactic acid is mainly produced by the anaerobic 33
9 house gas problem [1–8]. Nowadays, to satisfy the increasing fermentation using starch-derived glucose as the raw material 34
10 energy demand, the biodiesel production is on a large scale [17]. Although the fermentation process gives a high lactic acid 35
11 worldwide. Meanwhile, glycerol as a by-product, is largely yield of 90%, the production cost is at relatively high level due to 36
12 produced in the biodiesel production via the transesterification the specific requirement of bacteria [18] and complex separation 37
13 of vegetable oil or animal fat. In the transesterification process, the steps in downstream [19]. As an alternative to the conventional 38
14 production capacity of glycerol is ca. 10–20% of raw material in fermentation process, catalytic conversion of glycerol to lactic acid 39
15 weight [9,10]. In 2016, according to the international market, the becomes a hot research topic because biomass glycerol is supplied 40
16 production capacity of biodiesel is estimated to be 37 billion liters, with a lower price of 0.04–0.11$/kg (crude glycerol) in the United 41
17 4 billion liters of crude glycerol will be produced [11,12]. However, States [1]. 42
18 glycerol is oversupplied in market and unavoidably affects Kishida et al. firstly reported that the lactic acid yield of 90% 43
19 biodiesel market [13]. Recently, catalytic conversion of biomass was obtained when the hydrothermal conversion of glycerol 44
20 glycerol into high valued chemicals has attracted researcher’s great (0.33 mol L 1) was carried out at 300  C using NaOH as the 45
21 attention [13,14]. homogeneous catalyst in an aqueous solution [20,21]. Ramírez- 46
22 The glycerol derivatives, such as lactic acid, 1,3-propanediol, López et al. reported that the lactic acid yield of 84.5% was obtained 47
23 1,2-propanediol, and succinic acid, are useful chemicals, which can when the hydrothermal conversion of glycerol (2.5 mol L 1) was 48
24 be synthesized by various catalytic processes [9]. Notably, lactic carried at 280  C for 2.5 h with a NaOH/glycerol mole ratio of 1.1:1 49
25 acid, as an important building block chemical in the chemical [22]. Hydrothermal conversion of glycerol using NaOH as the 50
26 platform, has been widely used in food, pharmaceutical, leather, homogeneous catalyst gave high lactic acid yield. However, 51
27 cosmetic, textile, and biodegradable polymer (polylactic acid) relatively high reaction temperature probably is the drawback 52
28 fields [9,15]. Moreover, polylactic acid has the potential to replace for the homogeneous catalysis method. 53

When Au-Pt/TiO2, Au/CeO2, and Pt-Au/CeO2 were used as the 54

heterogeneous catalysts to catalyze aerobic oxidation of glycerol 55

* Corresponding authors. Fax: +86 (0)511 88791800. (0.17–0.22 mol L 1) at 90–100  C in a NaOH aqueous solution, the 56

maximum yields of lactic acid were 26%, 81%, and 79%, respectively 57
E-mail addresses: yin@ujs.edu.cn (H. Yin), alwang@ujs.edu.cn (A. Wang).

http://dx.doi.org/10.1016/j.jiec.2017.08.028
1226-086X/© 2017 Published by Elsevier B.V. on behalf of The Korean Society of Industrial and Engineering Chemistry.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

2 H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx

58 [23–25]. When Pt/ZrO2 was used as the catalyst for the conversion photocatalytic degradation of trichloroethylene and Rhodamine B, 100
59 of glycerol (5%) to lactic acid at 180  C under anaerobic condition in respectively [44–46]. The addition of activated carbon in metallic 101
60 a NaOH aqueous solution, the highest yield of lactic acid was 84% aluminum significantly enhanced the methylation of methyltri- 102
61 [26]. The noble metals exhibited high catalytic activities for the chlorosilane [47]. The use of carbon materials as catalyst supports 103
62 conversion of glycerol to lactic acid at a lower reaction temperature or active components is worth of investigation. 104
63 under aerobic or anaerobic condition. However, the high cost of In our present work, catalytically hydrothermal conversion 105
64 noble metal catalyst and low glycerol concentration will cause a of glycerol to lactic acid over fine graphite-supported nickel 106
65 high production cost in industrial scale. nanoparticles was investigated in a batch reactor. The graphite- 107
66 Instead of noble metals, copper oxide and metallic copper have supported nickel nanoparticle catalysts were prepared by the wet 108
67 been investigated for the catalytically hydrothermal conversion of chemical reduction method in an ethanol solution. The effect of 109
68 glycerol to lactic acid. When Cu/SiO2, CuO/Al2O3, and Cu2O were reaction parameters, such as reaction time, reaction temperature, 110
69 used as the catalysts for the hydrothermal conversion of glycerol glycerol concentration, NaOH/glycerol mole ratio, and catalyst 111
70 (1 mol L 1) in a NaOH aqueous solution at 240  C for 6 h, the lactic loading, on the catalytic conversion of glycerol to lactic acid was 112
71 acid selectivities were 79.7%, 78.6%, and 78.1% at the glycerol investigated in detail. The roles of metallic Ni0 nanoparticles and 113
72 conversions of 75.2%, 97.8%, and 93.6%, respectively [27]. The lactic NaOH in the catalytic reaction were discussed. A power-function 114
73 acid selectivity of 90% was obtained at the glycerol conversion of type reaction kinetic model was used to evaluate the effect of 115
74 91% in the hydrothermal conversion of glycerol over Cu(16)/HAP reaction parameters, glycerol and NaOH concentrations and 116
75 catalyst at 230  C for 2 h with the initial glycerol and NaOH reaction temperature, over the supported nickel nanoparticle 117
76 concentrations of 1.0 and 1.1 mol L 1, respectively [13]. The catalysts. 118
77 heterogeneous transition metal catalysts exhibited good catalytic
78 activities for the catalytically hydrothermal conversion of glycerol Experimental 119
79 to lactic acid at a lower reaction temperature than the homoge-
80 neous alkaline catalysts. Materials 120
81 Metallic and oxide nanomaterials have attracted a great
82 attention of researchers because of their excellent performances Fine graphite slices with the sizes of 1–2.5 mm were supplied by 121
83 in catalysis and environmental protection fields [28–36]. Sup- Celtig Co. Ltd. USA. Glycerol, lactic acid, 1,2-propanediol, oxalic 122
84 ported metallic Ni0 and Ni0 nanoparticles exhibited high catalytic acid, formic acid, acetic acid, anhydrous ethanol, isopropyl alcohol, 123
85 activities for hydrogenation reactions [37–42]. They should also sodium hydroxide (NaOH), hydrazine hydrate (N2H4H2O, 85%), 124
86 have good catalytic activity for dehydrogenation reaction accord- and nickel acetate (C4H6O4Ni4H2O), were purchased from 125
87 ing to the catalysis principle. As a transition metal catalyst, metallic Sinopharm chemical reagent Co., Ltd. All the chemicals were of 126
88 nickel should have catalytic activity in the hydrothermal conver- reagent grade and were used as received without further 127
89 sion of glycerol to lactic acid, in which dehydrogenation reaction is purification. Deionized water was used throughout all the 128
90 included. To the best of our knowledge, metallic Ni0 catalyst has experiments. 129
91 not been used for the hydrothermal conversion of glycerol to lactic
92 acid. Preparation of catalysts 130
93 Carbon materials have been widely used in catalysis, electro-
94 chemistry, heavy metallic ion adsorption, and chemical purifica- The fine graphite-supported nickel nanoparticle catalysts were 131
95 tion fields. It was reported that activated carbon black-supported prepared by the wet chemical reduction method. The metallic Ni0 132
96 Ni/NiO exhibited high capacitance and energy density [43]. nanoparticles were prepared by reducing nickel acetate with 133
97 Polyacrylonitrile fiber-supported BiOClx/BiOBry/BiOIz composites, hydrazine hydrate in anhydrous ethanol in the presence of fine 134
98 graphene oxide/BiOCl/PAN nanofibers, and nitrogen-carbon dots/ graphite powder. Firstly, 1.0 g of fine graphite powder was 135
99 (3D)BiOBr composites exhibited high catalytic activities for the dispersed in 60 mL of anhydrous ethanol by ultrasonic treatment 136

Table 1
The properties and catalytic activities of Ni/graphite catalysts.a Q4

Catalysts Amount of Crystallite sizes Average particle sizes Specific Glycerol Selectivities Carbon TOFsf
Nib of Ni0 (111)c and size distributionsd surface areas conversions balancese (h 1)
(mmol g 1) (nm) (nm) 2
(m /g) (%) (%)
Lactic Oxalic Formic Acetic 1,2-
acid acid acid (%) acid propanediol
(%) (%) (%) (%)
Ni0.1/ 0.97 11.3 22.4, 9.3–45.7 7.7 62.2 83.2 0.3 0.5 0.6 3.1 87.7 59.6
graphite
Ni0.2/ 1.95 12.2 29.5, 12.9–44.6 11.7 77.4 88.9 0.4 1.0 0.9 2.3 93.5 39.2
graphite
Ni0.3/ 2.95 12.7 32.9, 21.3–67.4 13.0 95.1 92.1 0.7 2.0 1.3 1.0 97.1 33.8
graphite
(Spent) 2.70
Ni0.4/ 3.91 13.2 35.1, 18.9–61.5 13.5 99.3 76.8 1.2 2.8 2.0 0.4 83.2 26.5
graphite
graphite \ \ \ 5.9 trace 100 0 0 0 0 100 \

Reaction conditions: glycerol aqueous solution, 100 mL; glycerol concentration, 1.0 mol L ; NaOH/glycerol mole ratio,1.1:1; catalyst, 0.552 g; reaction temperature 230  C;
a 1

reaction time, 2.0 h.


b
The amount of Ni (mmol) was analyzed by atomic absorption spectrophotometer.
c
The crystallite sizes of Ni0 (111) were calculated by XRD.
d
The particle sizes of Ni0 nanoparticles were determined by TEM.
e
Carbon balances were calculated according to both detected products and reacted glycerol.
f
TOF (h 1) = Converted glycerol (mol) divided by the mole of Ni supported on graphite after reacting for 2.0 h.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx 3

137 for 30 min, and then given amount of nickel acetate was dissolved charged with 100 mL of aqueous solution of glycerol and NaOH and 199
138 in it. The mixture was preheated to 60  C in a thermostatic bath an appointed amount of Ni/graphite catalyst. The autoclave was 200
139 under stirring. A saturated NaOH anhydrous ethanol solution was flushed with N2 to replace air inside for 10 min. The stirring speed 201
140 added dropwise to adjust the pH value of the reaction solution to was set at 300 rpm. The reaction mixture was heated to the desired 202
141 12. Then, a hydrazine hydrate ethanol solution (10 mL in 150 mL temperature in ca. 0.5 h. The reaction time was counted from the 203
142 anhydrous ethanol) was added dropwise to the mixture and heated reaching of the setting temperature. After reacting for an 204
143 to 70  C for 4 h under mild stirring. After reaction, the reaction appointed time period, the reaction mixture was cooled down 205
144 mixture was cooled down to room temperature. The as-prepared to room temperature with cooling water through a coil in the 206
145 Ni/graphite catalysts were filtrated and washed with anhydrous reactor. 207
146 ethanol for three time and kept in an anhydrous ethanol solution Before analysis of the product, the reaction mixture was filtered 208
147 before they were characterized and used as the catalysts. The and the filtrate was acidified with hydrochloric acid (37%) to a pH 209
148 catalyst content in the ethanol suspension was analyzed by drying value of ca. 3 and diluted with deionized water for HPLC analysis 210
149 a given amount of suspension in a vacuum oven at 50  C for 12 h. [42]. The concentrations of the unreacted glycerol and produced 211
150 According to the weight percentage of catalyst in ethanol 1,2-propanediol were analyzed on a gas-phase chromatograph (SP- 212
151 suspension, given amount of ethanol suspension was taken out 6800A) equipped with a PEG–20 M packed capillary column 213
152 and filtrated. The wet cake of catalyst was directly used for the (0.25 mm  30 m) and a FID by using isopropyl alcohol as the 214
153 reaction to prevent the oxidation of metallic Ni0. According to the internal standard. The products, such as lactic acid, acetic acid, 215
154 amount of metallic Ni0, the as-prepared Ni/graphite catalysts (Nix/ oxalic acid, and formic acid, were analyzed on an Agilent HPLC 216
155 graphite, x, mole of Ni0 to 100 g of graphite) were denoted as Ni0.1/ system equipped with a tunable absorbance UV detector and a 217
156 graphite, Ni0.2/graphite, Ni0.3/graphite, and Ni0.4/graphite, respec- reverse-phase column (Innoval ASB C18, 5 mm, 100 Å, 4.6 mm 218
157 tively. The properties and compositions of the Ni/graphite catalysts  250 mm) at 30  C. The mobile phase was a methanol aqueous
158 are listed in Table 1. solution (10:90, V/V) with a flow rate of 0.8 mL min 1. The pH value 219

of the mobile phase was 2.3, which was adjusted with phosphate 220
159 Characterization of catalyst buffer. The detection wavelength was 210 nm. The product 221

selectivity was calculated by carbon balance. The catalytic test 222


160 The XRD patterns of the Ni/graphite catalysts and fine graphite for each catalyst was repeated for at least twice, and the relative 223
161 slices were recorded on a diffractometer (D8 super speed Bruker- deviations were in a range of 5%. 224
162 AEX Company) using Cu Ka radiation (l = 1.54056 Å) with Ni filter
163 and scanning (2u) from 10 to 85 . The crystallite sizes of Ni0 Results and discussion 225
164 nanoparticles (1 1 1) in the Ni/graphite catalysts were calculated
165 according to the Scherrer’s equation: D = Kl/(Bcosu). The value of K XRD analysis 226
166 was taken as 1.0 and B was the full width of the diffraction line at
167 half of the maximum of XRD peak of metallic Ni0 (1 1 1). The The XRD patterns of the Ni/graphite catalysts and fine graphite 227
168 crystallite sizes are listed in Table 1. slices are shown in Fig. 1. For the Ni/graphite catalysts, the XRD 228
169 The specific surface areas of the Ni/graphite catalysts and fine peaks appearing at 2u = 44.5, 51.8, and 76.4 were indexed as the 229
170 graphite slices were measured on a NOVA 2000e physical (1 1 1), (2 0 0), and (2 2 0) planes of the face centered-cubic (fcc) 230
171 adsorption apparatus by N2 adsorption/desorption technique. nickel (JCPDS 04-0850), respectively. No diffraction peaks of nickel 231
172 The samples were degassed at 150  C for 3 h under vacuum before oxides or nickel hydroxides were detected, indicating that phase- 232
173 the measurement. The N2 adsorption-desorption isotherms of the pure metallic Ni0 nanoparticles on graphite support were prepared 233
174 samples were measured at 196  C and the specific surface areas under the present experimental conditions. The peak intensities of 234
175 were calculated by the BET method. metallic Ni0 in Ni/graphite catalysts increased with the increase in 235
176 The microstructures and the average particle sizes of the the nickel loadings. The metallic Ni0 crystallite sizes (1 1 1) 236
177 metallic Ni0 nanoparticles on the graphite support were examined estimated by Scherrer’s equation ranged from 11.3 to 13.2 nm, 237
178 by transmission electron microscopy (TEM) on a microscope (JEM- revealing that the crystallite sizes of the metallic Ni0 in the 238
179 2100) operated at an acceleration voltage of 200 kV. The TEM catalysts were at a nanometer magnitude (Table 1). The Ni0 239
180 specimens were prepared by placing a drop of Ni/graphite
181 anhydrous ethanol suspension onto a copper grid coated with a
182 layer of amorphous carbon. For the preparation of the TEM
183 specimens, the Ni/graphite catalysts were dispersed in an
184 anhydrous ethanol via ultrasonic treatment for 30 min. The average
185 particle sizes of the metallic Ni0 nanoparticles on the graphite
186 support were measured from the TEM images by counting at least
187 150 individual particles. The average particle sizes of metallic Ni0
188 nanoparticles on the graphite support were calculated by a
189 weighted-average method according to the individual particle
190 sizes of all the counted particles.
191 The metallic Ni0 contents of the fresh Ni/graphite catalysts and
192 the spent Ni0.3/graphite catalyst were analyzed on an atomic
193 absorption spectrophotometer (TAS-986). The metallic Ni0
194 contents are listed in Table 1.

195 Catalytic test

196 The catalytic conversion of glycerol to lactic acid was carried out
197 in a 300 mL stainless steel autoclave with a mechanical stirrer and
198 the autoclave was heated in an electric furnace. The autoclave was Fig. 1. XRD patterns of fine graphite and Ni/graphite catalysts. *, Ni0; ^, graphite.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

4 H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx

240 crystallite sizes of the metallic Ni0 in the catalysts increased with that metallic Ni0 nanoparticles were well dispersed on the graphite 256
241 the increase in nickel loadings. surfaces, giving large specific surface areas. 257
242 The XRD pattern of the fine graphite slices showed a strong and
243 sharp diffraction peak at 26.6 , indexed to the reflection (0 0 3) of TEM and HRTEM analyses 258
244 graphite, and a weak diffraction peak appearing at 54.8 , ascribed
245 to the (0 0 6) plane of graphite (JCPDS 26-1079). The high intensity Fig. 2 shows the TEM and HRTEM images of the pure graphite 259
246 of XRD peak (0 0 3) indicated that the fine graphite sample had slices and Ni/graphite catalysts. It was observed that the graphite 260
247 high crystallinity. sample had thin lamellar structure. The TEM images show that the 261

metallic Ni0 nanoparticles with irregular shapes were formed and 262
248 N2 adsorption/desorption analysis well dispersed on the surface of graphite support. The average 263

particle sizes and particle size distributions of Ni0 nanoparticles in 264


249 The specific surface areas of the graphite slices and supported the Ni0.1/graphite, Ni0.2/graphite, Ni0.3/graphite, and Ni0.4/graphite 265
250 nickel catalysts are listed in Table 1. The specific surface areas of the catalysts were 22.4, 9.3–45.7; 29.5, 12.9–44.6; 32.9, 21.3–67.4 and 266
251 pure graphite slices and graphite-supported Ni0 catalysts were in 35.1, 18.9–61.5 nm, respectively. The average particle sizes of Ni0 267
252 the range of 5.9–13.5 m2 g 1. The specific surface areas of the nanoparticles increased with the increase in nickel loadings. 268
253 samples were in the order of Ni0.4/graphite > Ni0.3/graphite > Ni0.2/ The HRTEM images of the graphite-supported Ni nanoparticles 269
254 graphite > Ni0.1/graphite > graphite. The specific surface areas of show that the lattice fringes were around 0.102, 0.176, and 270
255 the samples increased with increasing the Ni0 loadings, indicating 0.203 nm, which were close to the (2 2 2), (2 0 0), and (1 1 1) lattice 271

Fig. 2. TEM and HRTEM images of pure graphite and Ni/graphite catalysts. (a,b) Ni0.1/graphite, (c,d) Ni0.2/graphite, (e,f) Ni0.3/graphite, (g,h) Ni0.4/graphite, and (i) graphite.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx 5

Fig. 2. (Continued)

272 spacings of fcc nickel, respectively. The resultant metallic Ni reacting at 230  C for 2 h. The Ni0.3/graphite catalyst gave the 305
273 nanoparticles had polycrystalline structure. highest lactic acid yield among the investigated catalysts. 306

Combining the lactic selectivity and glycerol conversion, it was 307


274 Catalytic conversion of glycerol to lactic acid found that both particle size and loading of Ni0 nanoparticles 308

affected the lactic acid yield. The Ni/graphite catalyst with the 309
275 Catalytic activities of supported nickel catalysts smallest Ni0 nanoparticle size or the highest Ni0 loading gave a low 310
276 To investigate the effect of metallic nickel loading and support lactic acid yield, probably due to the catalytic decomposition of the 311
277 on the catalytic conversion of glycerol to lactic acid, the catalytic resultant lactic acid to CO2. The Ni0.3/graphite catalyst with the Ni0 312
278 reaction over the Ni/graphite catalysts was carried out in 100 mL of nanoparticle sizes centered at ca. 33 nm favored the glycerol 313
279 glycerol (1 mol L 1) aqueous solution with the NaOH/glycerol mole conversion to lactic acid. 314
280 ratio of 1.1:1 and 0.552 g of catalyst at 230  C for 2 h. The Over the Ni/graphite catalysts, the carbon balances firstly 315
281 experimental results are listed in Table 1. increased and then decreased with the increase in the nickel 316
282 When NaOH or pure graphite was added in the reaction loadings. The Ni0.3/graphite catalyst gave the highest carbon 317
283 solution, only trace glycerol was converted and lactic acid was balance value. As explained above, small-sized Ni0 nanoparticles 318
284 solely formed. Pure graphite support or NaOH had a little catalytic formed at a low Ni0 loading or more metallic Ni0 active sites 319
285 activity for the conversion of glycerol to lactic acid at 230  C. When obtained at a high Ni0 loading probably resulted in the formation of 320
286 the Ni/graphite catalysts were used in the reaction, the glycerol CO2 through the decomposition of the resultant products. On the 321
287 conversions increased from 62.2% to 99.3% while the TOF values other hand, at a lower nickel loading, undetected by-products may 322
288 decreased from 59.6 to 26.5 h 1 with the increase in nickel be also formed, giving a lower carbon balance value. 323
289 loadings. The selectivities of lactic acid over the Ni0.1/graphite,
290 Ni0.2/graphite, Ni0.3/graphite, and Ni0.4/graphite catalysts were Effect of reaction time 324
291 83.2%, 88.9%, 92.1%, and 76.8%, respectively. The selectivities of Fig. 3 shows the conversions of glycerol and selectivities of 325
292 oxalic acid, formic acid, acetic acid, and 1,2-propanediol were less products over the Ni/graphite catalysts at 230  C for different 326
293 than 3.0%, respectively. reaction time periods. When the Ni0.1/graphite, Ni0.2/graphite, 327
294 The results revealed that the Ni0 nanoparticles supported on Ni0.3/graphite, and Ni0.4/graphite were used as the catalysts, with 328
295 graphite exhibited good catalytic activities for the catalytic prolonging the reaction time periods from 1 to 4 h, the conversions 329
296 conversion of glycerol to lactic acid. The TOF value of the Ni0.1/ of glycerol increased from 43.4% to 75.3%, 56.8% to 89.5%, 77.3% to 330
297 graphite catalyst with the smallest average particle size was 99.9%, and 88.6% to 100%, respectively. The maximum lactic acid 331
298 2.25 times that of the Ni0.4/graphite catalyst with the largest selectivities over the Ni/graphite catalysts were 86.8%, 91.3%, 332
299 average particle size, indicating that small-sized Ni0 nanoparticles 92.2%, and 89.2% after reacting for 4, 4, 3, and 1 h, respectively. The 333
300 exhibited high catalytic activity for the conversion of glycerol. selectivities of oxalic, formic, acetic acid, and 1,2-propanediol were 334
301 It was interesting to find that the selectivities of lactic acid less than 5%, respectively. 335
302 firstly increased and then decreased with the increase in nickel Over the Ni0.3/graphite catalyst, the lactic acid selectivities were 336
303 loadings. The Ni0.3/graphite catalyst gave the highest lactic acid ca. 92% at the glycerol conversions of 95.1%–97.6% after reacting for 337
304 selectivity of 92.1% at the glycerol conversion of 95.1% after 2–3 h. As compared with other Ni/graphite catalysts, the Ni0.3/ 338

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

6 H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx

Fig. 3. Catalytically hydrothermal conversions of glycerol over Ni/graphite catalysts. Reaction conditions: glycerol aqueous solution (100 mL) with the glycerol concentration
of 1.0 mol L 1, NaOH/glycerol mole ratio of 1.1:1, amount of catalyst of 0.552 g, and reaction temperature of 230  C.

339 graphite catalyst exhibited high catalytic activity in a wider range selectivities, and carbon balances decreased from 98.9% to 364
340 of reaction time for the hydrothermal conversion of glycerol to 76.5%, 93.1% to 76.6%, and 96.8 to 83.8%, respectively (Table 2). 365
341 lactic acid. The Ni0.3/graphite catalyst was selected as the model The selectivities of oxalic acid, formic acid, acetic acid, and 366
342 catalyst to investigate the effect of the other reaction parameters 1,2-propanediol were less than 3%, respectively. The results 367
343 on the catalytic hydrothermal reaction at a fixed reaction time revealed that at a low glycerol concentration, relative sufficient 368
344 period of 2 h.

345 Effect of reaction temperature


346 The glycerol conversions and product selectivities in the
347 hydrothermal conversion of glycerol catalyzed by the Ni0.3/
348 graphite at different reaction temperatures are shown in Fig. 4.
349 With increasing the reaction temperatures from 180 to 250  C, the
350 conversions of glycerol increased from 33.8% to 100% while the
351 selectivities of lactic acid decreased from 94.2% to 80.4%. The
352 selectivities of by-products, such as oxalic acid, formic acid, acetic
353 acid, and 1,2-propanediol, were less than 3%, respectively. The
354 results revealed that high reaction temperature favored the
355 catalytic conversion of glycerol to lactic acid. However, when
356 the reaction temperatures were increased from 230 to 250  C, the
357 selectivity of lactic acid rapidly decreased by 11.7%. The suitable
358 reaction temperature should be less than 250  C.

359 Effect of glycerol concentration and NaOH/glycerol mole ratio


360 When the catalytic conversion of glycerol was catalyzed by the
361 Ni0.3/graphite catalyst at 230  C for 2 h with the NaOH/glycerol Fig. 4. Effect of reaction temperature on the conversion of glycerol over Ni0.3/
362 mole ratio of 1.1:1, with increasing the glycerol concentrations graphite catalyst. Reaction conditions: glycerol aqueous solution (100 mL) with the
363 glycerol concentration of 1.0 mol L 1, NaOH/glycerol mole ratio of 1.1:1, amount of
from 0.5 to 3.0 mol L 1, the glycerol conversions, lactic acid
catalyst of 0.552 g, and reaction time of 2.0 h.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx 7

Table 2
Effect of glycerol and NaOH concentrations on hydrothermal conversion of glycerol.a

Glycerol NaOH Glycerol conversions Selectivities Carbon balancesb


(molL 1) (molL 1
) (%) (%)
Lactic acid Oxalic acid Formic acid Acetic acid 1,2-propanediol
(%) (%) (%) (%) (%)
0.5 0.55 98.9 93.1 0.6 0.8 1.0 1.3 96.8
1.0 1.1 95.1 92.1 0.7 2.0 1.3 1.0 97.1
2.0 2.2 88.3 83.6 1.0 2.4 1.5 0.7 89.2
3.0 3.3 76.5 76.6 1.2 2.9 2.4 0.7 83.8
1.0 1.0 94.1 91.4 0.6 1.8 1.3 1.1 96.2
1.0 1.3 97.0 90.0 0.9 2.0 1.4 0.9 95.2
1.0 1.5 99.0 89.4 1.0 2.3 1.5 0.8 95.0
a
Reaction conditions: glycerol aqueous solution, 100 mL; catalyst, 0.552 g; reaction temperature 230  C; reaction time, 2.0 h.
b
Carbon balances were calculated according to both detected products and reacted glycerol.

369 Ni0 active sites were available for the conversion of glycerol to before next recycling. As shown in Fig. 6, the glycerol conversion 395
370 lactic acid, giving a high lactic acid yield. and lactic acid selectivity over the fresh Ni0.3/graphite catalyst 396
371 When the mole ratios of NaOH to glycerol increased from were 95.1% and 92.1%, respectively. After recycling for 4 times, the 397
372 1.0:1.0 to 1.5:1.0, the glycerol conversions increased from 94.1% to glycerol conversion and lactic acid selectivity were 92.0% and 398
373 99.0%. The lactic acid selectivities and the carbon balances were 91.0%, respectively. According to the recycling experimental 399
374 around 90% and 96% respectively. The selectivities of all the results, it was found that the Ni0.3/graphite catalyst exhibited 400
375 measured by-products were less than 2.5%. The results revealed good recycling performance. 401
376 that high NaOH/glycerol ratio favored the glycerol conversion to After recycling for 4 times, the ratio of Ni0 mole number (mol) to 402
377 lactic acid. the weight of graphite (g) in the spent Ni0.3/graphite catalyst was 403

0.27:100, indicating only a little loss of Ni0 in the consecutive use of 404
378 Effect of catalyst loading Ni0.3/graphite catalyst. 405
379 Fig. 5 shows the conversions of glycerol and the selectivities of The catalytic performance of the Ni0.3/graphite was also 406
380 products over the Ni0.3/graphite catalyst at 230  C for 2 h with compared with those of the representative catalysts in published 407
381 different catalyst loadings. With increasing the catalyst loadings papers (Table 3). The results showed that the Ni0.3/graphite 408
382 from 0.092 to 0.644 g, the conversions of glycerol increased from catalyst gave a higher lactic acid yield than the noble metal and 409
383 60.2% to 99.3% while the selectivities of lactic acid decreased from copper-based catalysts. The reaction temperature was also 410
384 94.2% to 87.2%. The selectivities of oxalic acid, formic acid, acetic obviously lower than that in the hydrothermal reaction system 411
385 acid, and 1,2-propanediol were less than 2.5%, respectively. The without the use of heterogeneous catalyst. 412
386 results revealed that high catalyst loading favored the glycerol
387 conversion to lactic acid, but excessive catalyst loading resulted in Reaction kinetics 413
388 the decrease in lactic acid selectivity. The catalyst loading around
389 0.6 g favored the glycerol conversion to lactic acid. Preliminary consideration 414

Considering that the Ni0.3/graphite catalyst exhibited high 415


390 Recycling performance of Ni0.3/graphite catalyst catalytic activity for the hydrothermal conversion of glycerol to 416

lactic acid, it was selected as the model catalyst to investigate the 417
391 The recycling performance of the Ni0.3/graphite catalyst in the reaction kinetics. 418
392 catalytic conversion of glycerol to lactic acid was investigated In order to eliminate the effect of diffusion, the Ni0.3/graphite 419
393 (Fig. 6). After reacting at 230  C for 2 h, the used catalyst was catalyst with different loadings in the range of 0.276 g–0.552 g was 420
394 separated from reaction solution by centrifugation at 8000 rpm used for the catalytic conversion of glycerol at 170  C with the 421

Fig. 5. Effect of catalyst loading on the conversion of glycerol over Ni0.3/graphite Fig. 6. Recycling performance of Ni0.3/graphite catalyst. Reaction conditions for
catalyst. Reaction conditions: glycerol aqueous solution (100 mL) with the glycerol each run: glycerol aqueous solution (100 mL) with the glycerol concentration of
concentration of 1.0 mol L 1, NaOH/glycerol mole ratio of 1.1:1, reaction tempera- 1 mol L 1, NaOH/glycerol mole ratio of 1.1:1, amount of catalyst of 0.552 g; reaction
ture of 230  C, and reaction time of 2.0 h. temperature of 230  C; reaction time, 2 h.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

8 H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx

Table 3
Summary of published results for catalytic conversion of glycerol to lactic acid.a

Catalysts Reaction conditions Glycerol conversions (%) Lactic acid yields (%) Refs

Glycerol (molL 1
) NaOH (molL 1
) 
Temperature ( C) Time (h)
/ 0.33 1.25 300 1.5 / 90 [20]
AuPt/TiO2 0.22 0.88 90 / 30 26 [23]
AuPt/CeO2 0.17 0.68 100 0.5 99 79 [25]
Pt/ZrO2 0.54 0.98 180 24 94 84 [26]
Cu/SiO2 1.0 1.1 240 6 75.2 60 [27]
CuO/Al2O3 1.0 1.1 240 6 97.8 77 [27]
Cu2O 1.0 1.1 240 6 93.3 73 [27]
Cu/HAP 1.0 1.1 230 2 91 82 [13]
Ni0.3/graphite 1.0 1.1 230 2 95.1 88 This work
a
The yields of lactic acid were calculated according to the experimental data reported in literatures.

422 initial glycerol and NaOH concentrations of 1.0 and 1.1 mol L 1, ln( rA) = lnk + alnCA + blnCB (3) 458
423 respectively. A straight line was obtained by plotting the catalyst
424 loading versus the glycerol conversion (Fig. S1), indicating that the
425 reaction was controlled by chemical reaction rather than mass 459
lnk = lnA Ea/(RT) (4)
426 diffusion [42].
To calculate the reaction orders of a and b according to Eq. (3), 460
427 Before estimating the reaction kinetics, several experiments
the initial conversion rates of glycerol were calculated according to 461
428 were carried out to determine the reaction time and the effect of
the data shown in Figs. S2a and b, which show the glycerol 462
429 stirring speed. Raising the autoclave temperature to the prescribed
430 conversions with reaction time at 170  C under different initial 463
reaction temperature needed ca. 0.5 h without stirring. Without 464
431 concentrations of glycerol and NaOH. The initial conversion rates of
stirring, glycerol was almost not converted during the time period
432 glycerol at 170  C under different initial concentrations of glycerol 465
for raising reaction temperature. Furthermore, when the reaction 466
433 and NaOH were calculated at the first hour (Table 4). By using the
was carried out over the Ni0.3/graphite catalyst under stirring at 467
434 multiple linear regression method, the reaction orders and
500 rpm or 1000 rpm for 1 h, the glycerol conversions were close to 468
435 reaction rate constant were calculated with a good regression
each other, indicating that the diffusion effect was eliminated 469
436 coefficient of 0.9845.
under stirring at 500 rpm. 470
437 The initial conversion rates of glycerol at different reaction
A power-function type reaction kinetic equation (Eq. (1)) was 471
438 temperatures were calculated at the first hour according to the
used to investigate the effect of reaction parameters, such as 472
439 data shown in Fig. S2c. And then the values of the rate constant k at
reaction temperature, glycerol concentration, and NaOH concen- 473
440 different reaction temperatures were calculated according to
tration, on the glycerol consumption rate [13]. 474
Eq. (1). According to the linear Eq. (4), the pre-exponential
441 rA = dnA/(mcatdt) = kCA CBb a
(1) 475
(frequency) factor, A, and the reaction activation energy, Ea, were
calculated by the linear regression method with a good regression 476
443
442 where rA is the glycerol consumption rate, mol gcat 1 h 1; mcat is
coefficient of 0.9813 (Table 5). 477
444 the amount of catalyst, g; nA is the mole number of glycerol, mol; t
According to the regression coefficients, all the experiment data 478
445 is the reaction time, h; CA and CB are the glycerol and NaOH
were well simulated by using Eqs. (3) and (4), revealing that the 479
446 concentrations, mol L 1; a and b are the reaction orders for glycerol
power-function type kinetic model was appropriate for the 480
447 and NaOH, respectively.
evaluation of the effect of glycerol concentration, NaOH concen- 481
448 The rate constant k follows the Arrhenius equation.
tration, and reaction temperature on the hydrothermal conversion 482
449 k = Aexp ( Ea/(RT)) (2) of glycerol. The overall reaction kinetics was expressed as follows. 483

451
450 where k is the rate constant; A is the pre-exponential (frequency) rA = A exp( 69.2/RT) CA0.41CB0.91 [mol gcat 1
h 1
] (5) 484
452 factor; R is the ideal gas constant, 8.314  10 3 kJ mol 1 K 1; Ea is
453 the reaction activation energy, kJ mol 1; and T is the reaction
Reaction routes 485
454 temperature, K.
When sole Ni0.3/graphite catalyst or sole NaOH was present in 486
455 Reaction orders and reaction activation energy
456 the reaction system at 230  C for 4 h, the glycerol conversions were 487
By taking the natural logarithm of both sides of the Eqs. (1) and 488
457 less than 3% and lactic acid was formed as the main product
(2), linear Eqs. (3) and (4) are obtained as follows. 489
(Table S1). When the reaction was catalyzed by both graphite and

Table 4
Reaction rate constant and reaction orders over Ni0.3/graphite catalyst.a
1 1
Glycerol (mol L ) NaOH (mol L ) Glycerol conversions (%) Reaction rates kb a b R2
0.5 1.1 23.9 0.0216 0.0285 0.41  0.05 0.91  0.06 0.9845
1.0 1.1 17.1 0.0310
1.5 1.1 13.2 0.0358
2.0 1.1 10.5 0.0381
1.0 0.5 8.4 0.0153
1.0 1.0 15.5 0.0281
1.0 1.5 23.7 0.0429
a
The experimental conditions: glycerol aqueous solution, 100 mL; catalyst loading, 0.552 g; reaction time, 1 h; reaction temperature, 170  C.
b
Reaction rate constant k, mol1 (a+b)gcat(a+b) 1h 1.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx 9

Table 5
Rate constants, frequency factor, and activation energy over Ni0.3/graphite catalyst.

Temperatures ( C) Conversions (%) Reaction rates ki A Ea (kJmol 1


) R2
140a 4.6 0.0084 0.0077 4.5  106 69.2 0.9813
150a 7.3 0.0132 0.0121
160a 13.4 0.0242 0.0222
170b 0.0285
a 1 1
The experimental conditions: glycerol aqueous solution, 100 mL; glycerol concentration, 1 mol L ; NaOH concentration, 1.1 mol L ; catalyst loading, 0.552 g; reaction
time, 1 h.
b
The rate constant was taken from Table 4.

490 NaOH at 230  C for 2 h, only trace amount of glycerol was converted formate anions in an alkaline solution, accompanied with the 516
491 to lactic acid (Table 1). However, the co-presence of Ni/graphite formation of carbonate anions [22]. Oxalate anions were probably 517
492 catalyst and NaOH effectively catalyzed the glycerol conversion to formed by the cleavage of glyceraldehyde in an alkaline solution 518
493 lactic acid. Based on the above findings and the reaction pathways [50]. 519
494 for alkali-catalyzed hydrothermal conversion of glycerol in an Under the present experimental conditions, the intermediates, 520
495 aqueous solution [20,22,48], the reaction routes for the catalytic such as glyceraldehyde, 2-hydroxypropenal, and pyruvaldehyde, 521
496 conversion of glycerol to lactic acid over the supported nickel were not detected, indicating that these intermediates could be 522
497 nanoparticle catalysts in an alkaline solution are proposed as rapidly converted to subsequent chemicals and finally to lactate. 523
498 Scheme 1. Compared with the hydrothermal conversion of glycerol in alkaline 524
499 For the glycerol conversion to lactic acid, the catalytic solution without heterogeneous catalyst, the supported nickel 525
500 dehydrogenation of glycerol to glyceraldehyde was the first step catalyst could effectively catalyze the conversion of glycerol to 526
501 [26,49]. The supported nickel catalyst played an important role in lactic acid via the glycerol dehydrogenation step. Furthermore, the 527
502 the catalytic dehydrogenation of terminal hydroxyl group of empirical reaction kinetic equation showed that over the Ni/ 528
503 glycerol to glyceraldehyde. Subsequently, 2-hydroxypropenal was graphite catalyst, the reaction could be carried out with a moderate 529
504 formed via the intramolecular dehydration of glyceraldehyde in a activation energy. Meanwhile NaOH concentration showed a more 530
505 basic environment [13,48]. Pyruvaldehyde was readily formed important effect on the reaction rate than that of glycerol 531
506 from 2-hydroxypropenal via the keto-enol tautomerization [49]. concentration. 532
507 The resultant pyruvaldehyde was converted to lactate via the
508 internal Cannizaro reaction [13,22,24,25]. Conclusions 533
509 In our present experiments, the by-products detected were
510 probably formed via the routes 2–4 as shown in Scheme 1. The graphite-supported metallic Ni0 nanoparticle catalysts 534
511 1,2-Propanediol was produced by the catalytic hydrogenolysis of were prepared by the wet chemical reduction method. The Ni0 535
512 glycerol with resultant H2 over the supported nickel catalyst. nanoparticles with the average particle sizes of 22.4–35.1 nm were 536
513 Additionally, both 2-hydroxypropenal and pyruvaldehyde could supported on the surfaces of fine graphite slices. Ni0 nanoparticles 537
514 also be hydrogenated to 1,2-propanediol over the supported nickel and NaOH synergistically catalyzed the glycerol conversion to 538
515 catalyst. Lactate anions could be decomposed to form acetate and lactic acid. Both Ni0 nanoparticle size and Ni0 loading affected the 539

Scheme 1. Reaction routes for the catalytic conversion of glycerol catalyzed by supported Ni catalyst in a NaOH aqueous solution.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028
G Model
JIEC 3577 1–10

10 H. Yin et al. / Journal of Industrial and Engineering Chemistry xxx (2017) xxx–xxx

540 lactic acid yield. When the catalytic conversion of glycerol was [18] H. Oh, Y.J. Wee, J.S. Yun, H.W. Ryu, Appl. Biochem. Biotechnol. 105 (2003) 603.
541 carried out over Ni0.3/graphite catalyst at 230  C for 3 h, the [19] W.Z. He, G.M. Li, L.Z. Kong, H. Wang, J.W. Huang, J.C. Xu, Resour. Conserv. Recycl.
564
542 52 (2008) 691.
selectivity of lactic acid was 92.2% at the glycerol conversion of [20] H. Kishida, F. Jin, Z. Zhou, T. Moryia, H. Enomoto, Chem. Lett. 34 (2005) 1560.
543 97.6%. The Ni0.3/graphite catalyst exhibited good recycling perfor- [21] Z. Shen, F. Jin, Y. Zhang, B. Wu, A. Kishita, K. Tohji, H. Kishida, Ind. Eng. Chem.
544 565
mance. A power-function type reaction kinetic model well fitted Res. 48 (2009) 8920.
545 [22] C.A. Ramirez-Lopez, J.R. Ochoa-Gómez, M. Fernández-Santos, O. Goméz-
the experimental data, rA = A exp( 69.2/RT) Cglycerol0.41CNaOH0.91. Jiménez-Aberasturi, A. Alonso-Vicario, J. Torrecilla-Soria, Ind. Eng. Chem. Res. 566
546 The reaction activation energy was 69.2 kJ mol 1. 567
49 (2010) 6270.
[23] Y.H. Shen, S.H. Zhang, H.J. Li, Y. Ren, H.C. Liu, Chem. Eur. J. 16 (2010) 7368.
547 [24] P. Lakshmanan, P.P. Upare, N.T. Le, Y.K. Hwang, D.W. Hwang, U.H. Lee, H.R. Kim,
Acknowledgements J.S. Chang, Appl. Catal. A 468 (2013) 260. 568
[25] R.K.P. Purushothaman, J.V. Haveren, D.S.V. Es, I. Melián-Cabrera, J.D. Meeldijk,
548 Q2 569
The work was financially supported by the funds from National H.J. Heeres, Appl. Catal. B 147 (2014) 92.
549 [26] J. Ftouni, N. Villandier, F. Auneau, M. Besson, L. Djakovitch, C. Pinel, Catal. Today
Natural Science Foundation of China (21506078 and 21506082) 570
550 257 (2015) 267.
and Science and Technology Department of Jiangsu Province of [27] D. Roy, B. Subramaniam, R.V. Chaudhari, ACS Catal. 1 (2011) 548.
551 China (BY2014123-10). [28] F. Razi, S. Zinatloo-Ajabshir, M. Salavati-Niasari, Mater. Lett. 193 (2017) 9.
[29] M. Bazarganipour, M. Salavati-Niasari, Appl. Catal. A: Gen. 502 (2015) 57.
552 [30] S. Zinatloo-Ajabshir, M.S. Morassaei, M. Salavati-Niasari, J. Colloid Interface Sci.
Appendix A. Supplementary data 497 (2017) 298. 571
[31] F. Razi, S. Zinatloo-Ajabshir, M. Salavati-Niasari, J. Mol. Liq. 225 (2017) 645.
553 Supplementary data associated with this article can be found, in [32] S. Zinatloo-Ajabshir, S. Mortazavi-Derazkola, M. Salavati-Niasari, J. Mol. Liq.
234 (2017) 430. 572
554 the online version, at http://dx.doi.org/10.1016/j.jiec.2017.08.028. [33] S. Zinatloo-Ajabshir, S. Mortazavi-Derazkola, M. Salavati-Niasari, Ultrason.
Sonochem. 39 (2017) 452. 573
555 References [34] S. Mandizadeh, M. Salavati-Niasari, M. Sadri, Sep. Purif. Technol. 175 (2017)
399. 574
[35] S. Zinatloo-Ajabshir, M. Salavati-Niasari, Sep. Purif. Technol. 179 (2017) 77.
[1] A.B.F. Moreira, A.M. Bruno, M.M.V.M. Souza, R.L. Manfro, Fuel Process. Technol.
556 [36] S. Zinatloo-Ajabshir, Z. Zinatloo-Ajabshir, M. Salavati-Niasari, S. Bagheri, S.B.A.
144 (2016) 170. 575
Hamid, J. Energy Chem. 26 (2017) 315.
[2] A. Abuadala, I. Dincer, Int. J. Energy Res. 36 (2012) 415.
[37] Y.H. Feng, H.B. Yin, A.L. Wang, T. Xie, T.S. Jiang, Appl. Catal. A 425 (2012) 205.
[3] N.S. Bentsen, C. Felby, Biotechnol. Biofuels 5 (2012) 25.
[38] Y. Du, H.L. Chen, R.Z. Chen, N. Xu, Appl. Catal. A 277 (2004) 259.
[4] A.V. Bridgwater, Chem. Eng. J. 91 (2003) 87.
[39] A.L. Wang, H.B. Yin, M. Ren, H.H. Lu, J.J. Xue, T.S. Jiang, New J. Chem. 34 (2010)
[5] L. Faba, E. Díaz, S. Ordóñez, Renew. Sustain. Energy Rev. 51 (2015) 273. 576
708.
[6] S.E. Hosseini, M.A. Wahid, M.M. Jamil, A.A.M. Azli, M.F. Misbah, Int. J. Energy
557 [40] A.L. Wang, H.B. Yin, H.H. Lu, J.J. Xue, M. Ren, T.S. Jiang, Catal. Commun. 10
Res. 39 (2015) 1597. 577
(2009) 2060.
[7] G.W. Huber, S. Iborra, A. Corma, Chem. Rev. 106 (2006) 4044.
[41] A.L. Wang, H.B. Yin, H.H. Lu, J.J. Xue, M. Ren, T.S. Jiang, Langmuir 25 (2009)
[8] E.D. Revellame, W.E. Holmes, T.J. Benson, A.L. Forks, W.T. French, R. Hernandez, 578
558 12736.
Top. Catal. 55 (2012) 185.
[42] Y.H. Feng, H.B. Yin, D.Z. Gao, A.L. Wang, L.Q. Shen, M.J. Meng, J. Catal. 316 (2014)
[9] M.R.A. Arcanjo, I.J. Silva Jr, E. Rríguez-Castellón, A. Infantes-Molina, R.S. Vieira, 579
559 Q3 67.
Catal. Today (2016), doi:http://dx.doi.org/10.1016/j.cattod.2016.02.015 in
560 [43] Y. Zhang, M. Park, H.Y. Kim, S.-J. Park, J. Colloid Interface Sci. 500 (2017) 155.
press.
[44] Y. Zhang, M. Park, H.Y. Kim, B. Ding, S.-J. Park, Appl. Surf. Sci. 384 (2016) 192.
[10] S. Adhikari, S.D. Fernando, A. Haryanto, Renew. Energy 33 (2008) 1097.
[45] Y. Zhang, M. Park, H.-Y. Kim, S.-J. Park, J. Alloys Compd. 686 (2016) 106.
[11] R. Moita, A. Freches, P.C. Lemos, Water Res. 58 (2014) 9.
[46] Y. Zhang, M. Park, H.Y. Kim, B. Ding, S.-J. Park, Sci. Rep. 45086 (2017).
[12] R.K. Saxena, P. Anand, S. Saran, J. Isar, Biotechnol. Adv. 27 (2009) 895.
[47] Y. Liu, Y. Jiang, H. Yin, W. Chen, Y. Shen, Y. Feng, L. Shen, A. Wang, T. Jiang, Z. Wu,
[13] H.X. Yin, C.H. Zhang, H.B. Yin, D.Z. Gao, L.Q. Shen, A.L. Wang, Chem. Eng. J. 288 580
561 Korean J. Chem. Eng. 28 (2011) 2250.
(2016) 332.
[48] L. Chen, S.J. Ren, X.P. Ye, Fuel Process. Technol. 120 (2014) 40.
[14] C. Zhang, T. Wang, X. Liu, Y.J. Ding, Chin. J. Catal. 37 (2016) 502.
[49] F. Auneau, C. Michel, F. Delbecq, C. Pinel, P. Sautet, Chem. Eur. J. 17 (2011) 14288.
[15] R.E. Drumright, P.R. Gruber, D.E. Henton, Adv. Mater. 12 (2000) 1841.
[50] F. Auneau, L.S. Arani, M. Besson, L. Djakovitch, C. Michel, F. Delbecq, P. Sautet, C.
[16] G. Dreschke, M. Probst, A. Walter, T. Pümpel, J. Walde, H. Insam, Bioresour. 581
562 Pinel, Top. Catal. 55 (2012) 474.
Technol. 176 (2015) 47.
[17] R.P. John, K.M. Nampoothiri, A. Pandey, Appl. Microbiol. Biotechnol. 74 (2007)
563 524.

Please cite this article in press as: H. Yin, et al., Catalytic conversion of glycerol to lactic acid over graphite-supported nickel nanoparticles and
reaction kinetics, J. Ind. Eng. Chem. (2017), http://dx.doi.org/10.1016/j.jiec.2017.08.028

Potrebbero piacerti anche