Producing Aromatic-Enriched Oil from Mixed Plastics Using Activated

Feb 20, 2018 - ABSTRACT: Producing aromatic-enriched oil from mixed plastics through catalytic pyrolysis has been experimentally studied. The effect o...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Article

Producing aromatic enriched oil from mixed plastics using activated biochar as catalyst Kai Sun, Qunxing Huang, Mujahid Ali, Yong Chi, and Jianhua Yan Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b03710 • Publication Date (Web): 20 Feb 2018 Downloaded from http://pubs.acs.org on February 24, 2018

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 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 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 38 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

2

Producing aromatic enriched oil from mixed plastics using activated biochar as catalyst

3

Kai Sun, Qunxing Huang*, Mujahid Ali, Yong Chi, Jianhua Yan

4

State Key Laboratory of Clean Energy Utilization, Zhejiang University, Hangzhou, 310027,

5

People's Republic of China

6

KEYWORDS: Plastics, Oil, Aromatics, Activated biochar, Catalyst

7

ABSTRACT: Producing aromatic enriched oil from mixed plastics through catalytic pyrolysis

8

has been experimentally studied. The effect of biochar catalysts has been investigated and the

9

possible dominating catalytic mechanisms of biochars activated with different chemical agents

1

10

have been discussed. Results indicated that when waste plastics were pyrolyzed with raw

11

biochar, the alkene fraction in the oil product increased to 54.9%. When biochar was activated by

12

ZnCl2, KOH and H3PO4, the oil product showed high selectively towards aromatics, and the

13

proportions of aromatics were up to 47.6%, 44.7%, and 66.0%, respectively. Benzene, 1,1'-

14

(1,3-propanediyl) bis- was the main composition in aromatics, the proportion of which could be

15

up to 25.1% when KOH activated biochar was used. The enrichment part of aromatics was

16

mainly bicyclic aromatics and C15-C16 compositions, the maximum proportions of which could

17

reach 92.5% and 28.1% by KOH activated biochar. High surface functional group (e.g. C=O) 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 38

18

content and low metal content of KOH activated biochar promoted hydrogen transfer reaction of

19

alkenes to alkanes and aromatics. While aromatization process promoted by Lewis acid sites and

20

Brønsted acid sites on ZnCl2 and H3PO4 activated biochar respectively significantly increased

21

aromatic yield.

22

1. Introduction

23

As the production and consumption of plastic products grow rapidly, the waste plastics are

24

leading to increasingly serious pollution of environment and waste of energy due to their low

25

recovery rate and non-biodegradability1-4. Compared with the complicated and costly mechanical

26

recycling method 1-3 or polluted and energy-wasting incineration method 5, converting waste

27

plastics into valuable chemical products 6, 7 through catalytic pyrolysis is a more promising

28

technology suitable for the volume reduction and resource recovery of dirty and mixed waste

29

plastics 8, 9.

30

Commonly used catalysts for plastics pyrolysis process include conventional zeolites (e.g.

31

HZSM-5, HY), meso-structure catalysts (e.g. MCM-41), FCC catalysts, noble metal catalysts

32

and some other catalysts. Conventional zeolites usually show high selectivity to aromatic

33

compounds and have a low rate of deactivation. C Santellan et al. 6 found that adding HUSY and

34

HZSM-5 during pyrolysis of real waste plastics could increase the yields of styrene (17.5%) and

35

ethylbenzene (15%). However, conventional zeolites have single pore size distribution, resulting 2 ACS Paragon Plus Environment

Page 3 of 38 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

36

in poor adaptability to feedstocks. Besides, zeolites like HZSM-5 would decrease the oil yield

37

significantly 10. Spend FCC catalysts are low-cost and could increase oil yield. GDL Puente et al.

38

11

39

of waste polyethylene catalyzed by commercial FCC catalysts. But FCC catalysts are less active

40

than zeolites 12 and show low selectivity to aromatics 13. Noble metal catalysts such as Pt and Ga

41

are usually loaded on chars or zeolites, which show high selectivity to aromatics. Y Uemichi et

42

al. 14-16 obtained a maximum yield of aromatics up to 46.2 wt.% and 30 wt.% in pyrolysis oil of

43

PE and PP by using 3 wt.% Pt-loaded activated carbon catalyst. However, high cost, easy-coking

44

and deactivation make it difficult to scale up the process.

45

found that oil with gasoline component content as high as 80% could be produced by pyrolysis

In recent years, carbonaceous materials like activated biochar are widely used in catalytic

46

fields, such as hydrogen production 17, 18, methane reforming 19, 20, and tar reforming 21, etc., due

47

to their low production cost, well developed and controllable pore structures, and convenient

48

surface modification 22, etc. Some researches23, 24 have studied the catalytic effect of activated

49

biochar on the pyrolysis of plastics in the comparative study with molecular sieves and silica gel,

50

and found that activated biochar could decrease sulfur content and increase aromatic yield in oil.

51

However, to the best of our knowledge, few researchers have studied the differences among

52

different carbonaceous materials, or analyzed the possible catalytic mechanisms.

53

In this study, raw biochar and different chemical activated biochars were used as catalysts for

54

the pyrolysis of mixed waste plastics to produce aromatic enriched oil. Chemical activation could 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

55

improve the specific surface area of catalysts and support chemical agent on the surface

56

simultaneously. Thus, a large number of active sites can be introduced on char surface in single

57

modification process, which make chemical activated biochar a promising catalyst. The effects

58

of these catalysts on oil yield and aromatic compositions were comparatively investigated and

59

the possible dominating catalytic mechanisms of different activation were also discussed.

60

2. Materials and experimental methods

61

2.1 Raw Materials

62

Page 4 of 38

In this work, commonly used plastics, such as polyethylene (PE), polypropylene (PP), and

63

polystyrene (PS) were employed. Samples were purchased from Guanbu Electromechanical

64

Technology Co., China with particle size less than 0.18µm. Before use, plastics were dried at 105

65

°C for 24 h and mixed with 59 wt.% of PE, 22 wt.% of PP, and 19 wt.% of PS according to their

66

typical relative fraction in municipal solid waste 25.

67

2.2 Experimental preparation of the catalysts

68

In this paper, both inactivated raw biochar (BC) and activated biochar (ACs) were used as

69

catalysts for plastics pyrolysis. Raw biochar was prepared by pyrolyzing wood chips (Lishui,

70

Zhejiang, China) from room temperature to 500 °C at a heating rate of 10 °C/min and kept at 500

4 ACS Paragon Plus Environment

Page 5 of 38 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

71

°C for 40 min. Activated biochars were prepared respectively through KOH, ZnCl2 and H3PO4

72

activation methods (abbreviated as AC-KOH, AC-ZnCl2, and AC-H3PO4, respectively).

73

To obtain AC-ZnCl2 catalyst, 20 g of wood chips were impregnated with 300 ml aqueous

74

solution of ZnCl2 at an impregnation ratio (Defined as the weight ratio of activating agent to

75

feedstock) of 1:1 for 24 h. Then the mixtures were dried, calcinated under N2 atmosphere at 600

76

°C for 90min, washed with 0.1 M HCl solution and then distilled water until the pH reached

77

neutral. For KOH activation, 20 g of raw biochar were impregnated with 300 ml aqueous

78

solution of KOH at an impregnation ratio of 4:1 for 24 h. Then the mixtures were dried,

79

calcinated under N2 atmosphere at 800 °C for 90 min, washed with 0.1 M HCl solution and then

80

distilled water until the pH reached neutral. While for H3PO4 activation, 20 g of wood chips were

81

impregnated with 300 ml aqueous solution of H3PO4 an impregnation ratio of 2:1 for 24 h. Then

82

the mixtures were dried, calcinated under N2 atmosphere at 500 °C for 90 min, washed with 0.1

83

M NaOH solution and then distilled water until the pH reached neutral. Detailed preparation

84

parameters of ACs are summarized in Table 2. After preparation, all the catalysts were milled

85

and sieved to a particle size of less than 0.4 mm.

86

ACs were washed with chemical agents to remove active matters. 5 g AC-KOH and AC-ZnCl2

87

were respectively added into 40 ml 5 M HCl solution. 5g AC-H3PO4 was added into 40 ml 5 M

88

NaOH solution. Then the mixtures were stirred for 24 h, washed with distilled water until the pH

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

89

reached neutral, and dried at 105 °C. Washed samples were abbreviated as WAC-KOH,

90

WAC-ZnCl2, and WAC-H3PO4, respectively.

91

2.3 Chars characterization

Page 6 of 38

92

Ultimate analysis of raw and activated biochar was carried out by a vario MAX cube

93

elemental analyzer (Elementar, German). Proximate analysis was analyzed by a 5E-IRSII

94

industrial analyzer (Kaiyuan Instrument Co., Ltd, Changsha, China). Surface morphology and

95

element distribution were characterized by a SIRON field emission scanning electron

96

microscope equipped with EDAX-EDS (FEI, Holland). Specific surface area was measured by

97

an AUTOSORB-IQ2-MP automatic specific surface and micropore size analyzer

98

(Quantachrome, USA) using multipoint Brunauer-Emmett-Teller (BET) standard method.

99

Microporous and mesoporous parameters were calculated by HK and BJH methods, respectively.

100

Surface functional groups were recorded by a Nicolet 5700 Fourier transform infrared

101

spectroscopy (Thermo Fisher scientific, USA), and a VG ESCALAB MKII X-ray photoelectron

102

spectrophotometer (VG, UK). High-resolution spectra were curve-fitted by multiple Gaussian

103

function.

104

2.4 Experimental setup

6 ACS Paragon Plus Environment

Page 7 of 38 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

105

Energy & Fuels

2 g catalyst and 8 g mixed plastic samples were mixed and heated from room temperature to

106

500 °C at a heating rate of 20 °C/min and remained for 1.0 h using 0.3 L/min nitrogen flow as

107

carrier gas. Pyrolysis oils were considered water-free and mainly collected by condensing device

108

and two gas-washing bottles containing dichloromethane. The whole reaction pipe except the

109

quartz boat was weighed before and after the test, then it was washed three times with 400 ml

110

dichloromethane. All solutions were dried at 40 °C for 24 h and weighted respectively. Oil yield

111

was calculated by adding up the mass of the oil both in pipe and collectors. Weight change of the

112

char was calculated from the weight increment of the quartz boat after the reaction.

113

Condensable oil product was recovered in dichloromethane and then analyzed by Agilent

114

6890-5973 gas chromatography-mass spectrometry (GC-MS). For each experimental run, the

115

oven stayed at 80 °C for 2 min, then was heated up to 250 °C at 20°C/min and kept for 10min.

116

All tests were repeated and the results presented in tables were the average values.

117

3. Results and discussion

118

3.1 Characterization of catalysts

119

Results of ultimate analysis and proximate analysis are shown in Table 2. Raw BC had a

120

relatively high carbon and hydrogen content. The increase ratio of C/H in ACs indicated a high

121

carbonization degree. High oxygen content, mostly caused by chemical activation process, may

122

provide a large number of binding sites on the char surface. 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 38

123

SEM and EDS analysis results are presented in Fig. 1. EDS spectra shows the elemental

124

distribution on char surface. Minerals in BC came from the feedstock, while minerals in ACs

125

both came from the feedstocks and the activating agents. Activating agents not only reshaped the

126

physical morphology, but were also loaded on the char surface in specific form, endowing the

127

ACs with catalytic capabilities. After activated by AC-ZnCl2 and AC-H3PO4, cylindrical

128

channels along the direction of the wood fiber were formed. While after activated by AC-KOH,

129

more intensive pore distribution and thinner pore walls can be observed, indicating that KOH

130

could promote the formation of pores more efficiently. Acid or alkali washing didn’t change the

131

surface morphology of ACs apparently. However, the content of active matters Zn, K, and P was

132

respectively reduced by 53.6%, 92.1%, and 66.5% after washing process.

133

Surface area and pore properties of the catalysts are shown in Table 3. Among all the ACs,

134

AC-KOH had the highest specific surface area and minimum average pore size. The micropore

135

ratio was as high as 82.6%, indicating that micropores played a dominant role in the pore

136

structure of AC-KOH. AC-ZnCl2 was another typical microporous AC which had similar

137

micropore ratio but much lower specific surface area compared with AC-KOH. AC-H3PO4 was a

138

kind of mesoporous AC with a high mesopore ratio of 50.8%.

139

High resolution scans of C1s and O1s XPS spectra are presented in Fig. 2, respectively.

140

Compared with BC, ZnCl2 activation led to an increase in carbon content and a decrease in

141

oxygen content on the char surface, while KOH and H3PO4 treatments were opposite. For surface 8 ACS Paragon Plus Environment

Page 9 of 38 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

142

carbon atoms, activation treatment decreased the content of graphitized carbon, and increased the

143

content of carbon atoms bonded to oxygen. C=O was the most abundant oxygen-containing

144

functional group in all catalysts. Meanwhile, –OH and –COOH were also abundant in AC-KOH

145

and AC-H3PO4.

146

3.2 Product yields and composition distribution of oils

147

Product yields of the gas, oil and weight change of catalysts after pyrolysis are shown in Fig.

148

3. Significant differences can be observed in oil yields as well as the weight changes of different

149

catalysts. Compared with the direct thermal pyrolysis, the catalysts have increased gas yields and

150

reduced oil yields in various degree. High specific surface area could increase heat transmission

151

area between the catalyst and the volatiles during thermal pyrolysis 24, meanwhile provide more

152

spaces for active sites like AAEM species which could promote the secondary pyrolysis of the

153

pyrolysis oil 26, 27, resulting in the production of smaller molecules.

154

Alkanes, alkenes and aromatics were major components in pyrolysis oils. Addition of different

155

catalysts changed the reaction mechanisms, thus changed the distribution of oil products to

156

varying degrees. As showed in Fig. 4, BC showed a pronounced catalytic effect in promoting

157

alkene yield (54.9%) and decreasing aromatic yield (16.9%). While chemically activated

158

biochars showed obvious catalytic effects in increasing aromatic yield. Catalytic pyrolysis oil of

159

AC-ZnCl2 had a high yield of aromatics (47.6%) at the expense of alkenes (24.1%). Catalytic 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 38

160

effect of AC-KOH led to a significant decrease in alkene production, while the yield of alkanes

161

(43.1%) and aromatics (44.7%) both increased remarkably. Catalytic effect of AC-H3PO4 was

162

similar to AC-ZnCl2, which presented the highest selectivity of aromatics (66.0%) and the lowest

163

selectivity of alkenes (9.8%). Among all catalysts, AC-H3PO4 possessed the largest potential in

164

the enrichment of aromatics.

165

The final residues after catalytic pyrolysis was higher than the initial mass of the catalysts,

166

except when BC was used. Weight change of BC was mainly weight loss caused by further

167

interaction with plastics, which will be discussed detailedly in section 3.4. Weight increases of

168

ACs were more obvious compared with BC. The enrichment effects of ACs towards aromatics

169

were AC-H3PO4 > AC-ZnCl2 > AC-KOH. Aromatics in oils provided abundant feedstocks for

170

dehydrogenation and condensation process, which led to the formation of coke.

171

Carbon atoms number against the concentration of the oil compositions is shown in Fig. 5 (a).

172

The main compositions were hydrocarbons with carbon atoms between 6 and 26. For most

173

catalytic reactions, C15-C16 hydrocarbons were the main compositions in pyrolysis oils.

174

Compared with direct thermal pyrolysis oil, catalysis of BC resulted in a more dispersed carbon

175

atoms number distribution. This could be due to the poor selectivity caused by the wide variety

176

of minerals and surface functional groups in BC. However, catalysis by ACs improved the

177

concentration of C15-C16 compositions (mainly branched bicyclic aromatics) and decreased the

178

content of C10 compositions (mainly branched monocyclic aromatics), resulting in a narrower 10 ACS Paragon Plus Environment

Page 11 of 38 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

179

production distribution. According to Fig. 5 (b), the selectivities towards C15-C16 aromatics were

180

increased 20.8%, 20.0%, and 29.0% by AC-ZnCl2, AC-KOH, and AC-H3PO4, respectively.

181

These increasing proportions of aromatics may be both from the decomposition of aromatic

182

polymers and condensation of monocyclic aromatics, or from the aromatization of alkanes and

183

alkenes, which will be discussed in the following section.

184

3.3 Aromatics in oils

185

The compositions of aromatics classified by the number of aromatic rings are shown in Fig. 6.

186

The aromatics in direct thermal pyrolysis oil were mainly bicyclic aromatics (63.3%) and

187

monocyclic aromatics (36.7%). The increasing proportion of monocyclic aromatics (44.0%) and

188

the appearance of a relatively high proportion of tricyclic aromatics (12.6%) in catalytic

189

pyrolysis oil proved that the presence of BC affected the distribution of aromatics, though the

190

catalytic selectivity was not apparent. ACs all showed obvious selectivities towards bicyclic

191

aromatics, especially AC-KOH (92.5%). The occurrence of tricyclic aromatics and even

192

tetracyclic aromatics indicated the occurrence of condensation reaction during the catalytic

193

processes, which was different from direct thermal pyrolysis process.

194

In the direct thermal pyrolysis oil, monocyclic aromatics were probably both derived from the

195

secondary reaction (e.g. alkylation reaction, transalkylation reaction, and dehydrogenation of the

196

side chains) of the PS pyrolysis products (such as A1 and A4), and the dehydrocyclization and 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 38

197

aromatization of long-chain alkanes and alkenes (such as A2 and A3). Pyrolysis of PE and PP

198

produced a large number of n-alkenes, providing sufficient reactants for both alkylation reaction

199

of aromatics and dehydrocyclization process. For BC, AC-ZnCl2, and AC-KOH catalytic

200

pyrolysis oils, monocyclic aromatics were composed of alkyl and alkenyl benzenes, which were

201

similar to thermal pyrolysis oils. Main monocyclic aromatics in AC-KOH group were alkyl

202

benzenes which had a higher saturation degree than the other groups. While high content of

203

cyclized productions in AC-H3PO4 group showed that the dehydrocyclization reaction was

204

obviously enhanced.

205

Bicyclic aromatics were the main compositions of the aromatic components. Most of the

206

bicyclic aromatics were multi-phenyl alicyclic hydrocarbons like B11, and a few were fused-ring

207

aromatics or biphenyl type aromatics such as B1 and B4. B11 (Benzene,

208

1,1'-(1,3-propanediyl)bis-, 64.9%) was the main bicyclic aromatics in direct thermal pyrolysis

209

oil, followed by B15 (16.3%), B10 (10.3%), and B5 (8.5%). They were mostly derived from the

210

secondary reactions (e.g. alkylation, transalkylation, dehydrogenation, and condensation) of

211

monocyclic aromatics. Catalysis of ACs enriched the species of the bicyclic aromatics, especially

212

for AC-H3PO4. A high proportion of B8 and the emergence of B1, B12, and B14, etc. in

213

AC-H3PO4 group clearly showed the enhancement of dehydrogenation, cyclization, and

214

aromatization during secondary reactions.

12 ACS Paragon Plus Environment

Page 13 of 38 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

215

In tricyclic aromatics, C1 and C2 were fused-ring aromatics with higher condensation degree

216

compared with C3 and C4. Thus, higher proportions of C1 and C2 in AC-H3PO4 group suggested

217

AC-H3PO4 had more pronounced effects on condensation reaction. Presence of tetracyclic

218

aromatics (3.5%) in oil compositions of AC-H3PO4 group also supported this view.

219

In conclusion, catalysis of all ACs enriched the bicyclic aromatics. AC-H3PO4 significantly

220

increased the proportions of aromatics (especially bicyclic ones), and its catalytic effect on

221

dehydrogenation and condensation was the most obvious. As showed in Fig. 7, AC-KOH,

222

AC-ZnCl2, and AC-H3PO4 all showed remarkable selectivities to Benzene,

223

1,1'-(1,3-propanediyl)bis- (B11), the proportion of which in could be up to 25.1%, 20.3%, and

224

16.2%, respectively. This was beneficial to the production and extraction of aromatics from

225

plastic pyrolysis oil.

226

3.4 Discussion of reaction mechanism

227

It is believed that direct thermal pyrolysis of polyolefins followed the free radical

228

mechanism 28, 29, while the dominating mechanism of catalytic pyrolysis depends on the catalyst

229

types, which are summarized in Fig. 8.

230

BC was the only non-activated carbonaceous material, the preparation temperature of which

231

was the same as the pyrolysis temperature of plastics. Since the mass loss of BC was around 3%

232

after catalytic reaction (Fig. 3), it can be inferred that the presence of plastics promoted the 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 38

233

further pyrolysis of BC. Hydrogen transfer reaction from hydrogen-rich materials (polyolefin

234

chains), to hydrogen-deficient materials (BC) occurred during their interaction 30. The

235

dehydrogenation process was enhanced by the abstraction of hydrogen from hydrogen enriched

236

hydrocarbons like alkanes onto the carbon surface 31, thus the production of alkenes increased

237

substantially. The occurrence of unsaturated compositions like cycloalkanes, dienes and even

238

cycloalkene, as showed in Fig. 9, had obviously proved the existence of hydrogen transfer and

239

cycloaddition. What’s more, the high content of Ca species in inherent minerals may contribute

240

to the reduction of aromatic yield 32.

241

When WAC-ZnCl2 and WAC-H3PO4 were used as catalysts, the content of aromatics

242

decreased remarkably and alkanes became the main components in the oil product. The reason is

243

that the concentration of Zn2+ and H3PO4 decreased after washing, and the catalytic effect of acid

244

sites on Diels-Alder reaction, hydrogen transfer reaction and dehydrocyclization reaction to

245

produce aromatics were reduced. It indicated that the effect of active matters (Zn2+ and H3PO4)

246

was dominating the catalytic pyrolysis reaction when AC-ZnCl2 and AC-H3PO4 were used as

247

catalysts. On the contrary, when WAC-KOH was used as catalyst, the change of production

248

composition was relatively small, which suggested that the catalytic effect of AC-KOH was less

249

sensitive to the concentration of the introduced active sites, and biochar properties may play a

250

more important role in catalytic reaction.

14 ACS Paragon Plus Environment

Page 15 of 38 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

251

Zn species on AC-ZnCl2 could form Lewis acid sites 33, 34, as showed in Fig. 8. The

252

presence of Zn species increased the consumption of alkenes, and led to the formation of

253

aromatics through Diels-Alder reaction 34, hydrogen transfer reaction 33 and direct

254

dehydrocyclization 35. Meanwhile, Lewis acid sites could effectively catalyze the alkylation of

255

aromatics, leading to the further consumption of alkenes and formation of a large number of

256

aromatics with side chains.

257

AC-KOH was a relatively pure AC. Mineral content on carbon surface was very low due to

258

alkali treatment and the loss of potassium in the process of pre-washing. It was reported that pure

259

carbon support was also effective in promoting hydrogen transfer reaction 31, 36 and cyclization of

260

straight-chain intermediate 37. The dehydrogenation step in hydrogen transfer was enhanced by

261

the abundant surface functional groups on carbon surface such as carbonyl groups 38, 39, while the

262

step of hydrogen release was relatively limited due to the lack of metal sites 36, leaving more

263

hydrogen atoms on carbon surface available for the hydrogenation of hydrogen acceptors such as

264

alkenes. Therefore, hydrogen transfer reaction of alkenes was enhanced greatly, and the

265

proportions of alkanes and aromatics increased significantly. This may also be the main catalytic

266

mechanisms of WACs. The slight decrease in aromatic content was probably caused by the

267

restrained hydrogen transfer reaction due to the lack of metal site.

268 269

H3PO4 treatment also introduced large number of acid sites and increased the acidity of catalyst surface 40, 41. AC-H3PO4 was abundant in phosphorus and oxygen-containing functional 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

270

groups, especially acid ones like -P=OOH 42, -COOH and -OH. These functional groups could

271

act as Brønsted acid sites, which had obvious catalytic effect on hydrogen transfer reaction,

272

Diels-Alder reaction, and Friedel-Crafts alkylation 43. Thus, AC-H3PO4 significantly enhanced

273

the consumption of alkenes and alkanes and improved the yield of branched aromatics.

274

Intermediate products of apparent dehydrocyclization and dehydro-condensation in aromatic

275

compositions (mentioned in section 3.3) were a strong proof of dehydrogenation reaction.

276

Page 16 of 38

Production distribution will be affected by not only the pore structure but also catalytic active

277

sites. In this paper, AC-KOH had the highest BET surface area and micropore ratio. As KOH

278

activation didn’t introduce much external active sites on AC, the product distribution was mainly

279

affected by the physical properties like pore structure. While for AC-ZnCl2 and AC-H3PO4, the

280

acid sites introduced played more important roles in changing production distribution. Strong

281

acid sites on the surface of AC-ZnCl2 and AC-H3PO4 promoted the end-chain scission of

282

polymers and favored the yield of gaseous products 44. High density of strong acid sites not only

283

led to coke formation, but also increased gas yield. So, the production distributions of AC-ZnCl2

284

and AC-H3PO4 catalyzed reactions were affected more by the introduced active sites.

285

4. Conclusions

16 ACS Paragon Plus Environment

Page 17 of 38 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

286

Carbonaceous material, including biochar and chemically activated biochars, had obvious

287

catalytic effects on pyrolysis oils of mixed plastics. From the analysis and comparison among

288

pyrolysis oils and different catalytic pyrolysis oils, the main conclusions were as follows:

289

(1) The catalysis of raw and activated biochars resulted in an increase in the proportion of

290

pyrolysis gas and a slight decrease in the proportion of pyrolysis oil.

291

(2) The catalytic effect of BC greatly increased the proportion of alkenes in the oil (up to

292

54.9%). The selectivity of aromatics was greatly improved by chemically activated biochars at

293

the cost of alkenes. AC-H3PO4 showed the strongest enrichment effect towards aromatics,

294

which could reach 66.0%.

295

(3) C15-C16 compositions (bicyclic aromatics) contributed most to the increase of aromatic

296

fraction. The enrichment effect of AC-H3PO4 was the most obvious, which increased C15-C16

297

compositions by 29%. Benzene, 1,1'- (1,3-propanediyl) bis- was the main composition in oils,

298

the proportion of which in AC-KOH, AC-ZnCl2, and AC-H3PO4 groups could be up to 25.1%,

299

20.3%, and 16.2%.

300

(4) The presence of plastics promoted the further pyrolysis of BC. AC-KOH’s catalytic effect

301

was mainly characterized by promoting hydrogen transfer process, which increased the yield

302

of alkanes and aromatics at the cost of alkenes. ZnCl2 and H3PO4 treatment could form

303

Lewis/Brønsted acid sites on the char surface, which favored dehydrogenation process, 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 38

304

hydrogen transfer reaction, Diels-Alder reaction, etc., and promoted the conversion of alkenes

305

to aromatics.

18 ACS Paragon Plus Environment

Page 19 of 38 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

306

Associated content

307

Supporting Information

308

Main compositions in direct thermal and catalytic pyrolysis oils from GC-MS results (based on

309

area %). (Table S1)

310

Author information

311

Corresponding Author

312

*Telephone: +86-571-87952834. Fax: +86-571-87952438. E-mail: [email protected]

313

Notes

314

The authors declare no competing financial interest.

315

Acknowledgments

316

This work was supported by National Natural Science Foundation of China (Grant No.

317

51621005), the National Key Research and Development Program of China (2016YFE0202000),

318

the Environmental Protection Special Funds for Public Welfare (201509013) and the

319

Fundamental Research Funds for the Central Universities.

320 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 38

321

References

322

(1) Kunwar, B.; Cheng, H. N.; Chandrashekaran, S. R.; Sharma, B. K., Plastics to fuel: a review.

323

Renewable and Sustainable Energy Reviews 2016, 54, 421-428.

324

(2) Wong, S. L.; Ngadi, N.; Abdullah, T. A. T.; Inuwa, I. M., Current state and future prospects

325

of plastic waste as source of fuel: A review. Renewable and Sustainable Energy Reviews 2015,

326

50, 1167-1180.

327

(3) Anuar Sharuddin, S. D.; Abnisa, F.; Wan Daud, W. M. A.; Aroua, M. K., A review on

328

pyrolysis of plastic wastes. Energy Conversion and Management 2016, 115, 308-326.

329

(4) Panda, A. K.; Singh, R. K.; Mishra, D. K., Thermolysis of waste plastics to liquid fuelA

330

suitable method for plastic waste management and manufacture of value added products—A

331

world prospective. Renewable and Sustainable Energy Reviews 2010, 14, (1), 233-248.

332

(5) Fu, Z.; Zhang, S.; Li, X.; Shao, J.; Wang, K.; Chen, H., MSW oxy-enriched incineration

333

technology applied in China: combustion temperature, flue gas loss and economic

334

considerations. Waste management 2015, 38, 149-56.

335

(6) Santella, C.; Cafiero, L.; De Angelis, D.; La Marca, F.; Tuffi, R.; Vecchio Ciprioti, S.,

336

Thermal and catalytic pyrolysis of a mixture of plastics from small waste electrical and

337

electronic equipment (WEEE). Waste management 2016, 54, 143-52. 20 ACS Paragon Plus Environment

Page 21 of 38 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

338

(7) Nishino, J.; Itoh, M.; Fujiyoshi, H.; Uemichi, Y., Catalytic degradation of plastic waste into

339

petrochemicals using Ga-ZSM-5. Fuel 2008, 87, (17-18), 3681-3686.

340

(8) Kumagai, S.; Yoshioka, T., Feedstock Recycling via Waste Plastic Pyrolysis.

341

Journal of the Japan Petroleum Institute 2016, 59, (6), 243-253.

342

(9) Miandad, R.; Barakat, M. A.; Aburiazaiza, A. S.; Rehan, M.; Nizami, A. S., Catalytic

343

pyrolysis of plastic waste: A review. Process Safety and Environmental Protection 2016, 102,

344

822-838.

345

(10) Seo, Y. H.; Lee, K. H.; Shin, D. H., Investigation of catalytic degradation of high-density

346

polyethylene by hydrocarbon group type analysis. Journal of Analytical & Applied Pyrolysis

347

2003, 70, (2), 383-398.

348

(11) Puente, G. D. L.; Klocker, C.; Sedran, U., Conversion of waste plastics into fuels: Recycling

349

polyethylene in FCC. Applied Catalysis B Environmental 2002, 36, (4), 279-285.

350

(12) Aguado, J.; Serrano, D. P.; Escola, J. M., Fuels from Waste Plastics by Thermal and

351

Catalytic Processes: A Review. Industrial & Engineering Chemistry Research 2008, 47, (21),

352

7982-7992.

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 38

353

(13) Lin, Y. H.; Yang, M. H., Catalytic conversion of commingled polymer waste into chemicals

354

and fuels over spent FCC commercial catalyst in a fluidised-bed reactor. Applied Catalysis B

355

Environmental 2007, 69, (3), 145-153.

356

(14) Uemichi, Y.; Ayame, A.; Kanazuka, T.; Kanoh, H., Degradation of polyethylene over

357

activated carbon catalyst. Nippon Kagaku Zassi 1985, (7), 1429-1435.

358

(15) Uemichi, Y.; Makino, Y.; Kanazuka, T., Degradation of polyethylene to aromatic

359

hydrocarbons over metal-supported activated carbon catalysts. Journal of Analytical & Applied

360

Pyrolysis 1989, 14, (4), 331-344.

361

(16) Uemichi, Y.; Makino, Y.; Kanazuka, T., Degradation of polypropylene to aromatic

362

hydrocarbons over Pt- and Fe-containing activated carbon catalysts. Journal of Analytical &

363

Applied Pyrolysis 1989, 16, (3), 229-238.

364

(17) Lee, K. K.; Han, G. Y.; Yoon, K. J.; Lee, B. K., Thermocatalytic hydrogen production from

365

the methane in a fluidized bed with activated carbon catalyst. Catalysis Today 2004, 93-95,

366

81-86.

367

(18) Yao, D.; Hu, Q.; Wang, D.; Yang, H.; Wu, C.; Wang, X.; Chen, H., Hydrogen production

368

from biomass gasification using biochar as a catalyst/support. Bioresource technology 2016, 216,

369

159-64.

22 ACS Paragon Plus Environment

Page 23 of 38 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

370

(19) Song, Q.; Xiao, R.; Li, Y.; Shen, L., Catalytic Carbon Dioxide Reforming of Methane to

371

Synthesis Gas over Activated Carbon Catalyst. Industrial & Engineering Chemistry Research

372

2008, 47, (13), 4349-4357.

373

(20) Abbas, H. F.; Daud, W. M. A. W., Deactivation of palm shell-based activated carbon

374

catalyst used for hydrogen production by thermocatalytic decomposition of methane.

375

International Journal of Hydrogen Energy 2009, 34, (15), 6231-6241.

376

(21) Striūgas, N.; Zakarauskas, K.; Stravinskas, G.; Grigaitienė, V., Comparison of steam

377

reforming and partial oxidation of biomass pyrolysis tars over activated carbon derived from

378

waste tire. Catalysis Today 2012, 196, (1), 67-74.

379

(22) Chen, Y.; Zhu, Y.; Wang, Z.; Li, Y.; Wang, L.; Ding, L.; Gao, X.; Ma, Y.; Guo, Y.,

380

Application studies of activated carbon derived from rice husks produced by chemical-thermal

381

process--a review. Advances in colloid and interface science 2011, 163, (1), 39-52.

382

(23) Gonzalez, Y. S.; Costa, C.; Marquez, M. C.; Ramos, P., Thermal and catalytic degradation

383

of polyethylene wastes in the presence of silica gel, 5A molecular sieve and activated carbon.

384

Journal of hazardous materials 2011, 187, (1-3), 101-12.

385

(24) Namieśnik, J.; Miskolczi, N.; Wu, C.; Williams, P. T.; Kan, C. W.; Yang, D., Fuels by

386

Waste Plastics Using Activated Carbon, MCM-41, HZSM-5 and Their Mixture. MATEC Web of

387

Conferences 2016, 49, 05001. 23 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 24 of 38

388

(25) Ehrig, R. J., Plastics recycling : products and processes. Hanser Publishers , Distributed in

389

the U.S.A. and Canada by Oxford University Press: 1992.

390

(26) Nowakowski, D. J.; Jones, J. M.; Brydson, R.; Ross, A. B., Potassium catalysis in the

391

pyrolysis behaviour of short rotation willow coppice. Fuel 2007, 86, (15), 2389-2402.

392

(27) Wang, Z.; Wang, F.; Cao, J.; Wang, J., Pyrolysis of pine wood in a slowly heating fixed-bed

393

reactor: Potassium carbonate versus calcium hydroxide as a catalyst. Fuel Processing

394

Technology 2010, 91, (8), 942-950.

395

(28) Miskolczi, N., Co-pyrolysis of petroleum based waste HDPE, poly-lactic-acid biopolymer

396

and organic waste. Journal of Industrial & Engineering Chemistry 2013, 19, (5), 1549-1559.

397

(29) Sekine, Y.; Fujimoto, K., Catalytic degradation of PP with an Fe/activated carbon catalyst.

398

Journal of Material Cycles & Waste Management 2003, 5, (2), 107-112.

399

(30) Önal, E.; Uzun, B. B.; Pütün, A. E., Bio-oil production via co-pyrolysis of almond shell as

400

biomass and high density polyethylene. Energy Conversion & Management 2014, 78, (1),

401

704-710.

402

(31) And, Z. G. Z.; Yoshida, T., Behavior of Hydrogen Transfer in the Hydrogenation of

403

Anthracene over Activated Carbon. Energy & Fuels 2001, 15, (3), 708-713.

24 ACS Paragon Plus Environment

Page 25 of 38 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

404

(32) Song, H.; Long, J.; Yi, W.; Sheng, S.; Sun, L. S.; Xu, B. Y.; He, L. M.; Xiang, J., Effects of

405

inherent alkali and alkaline earth metallic species on biomass pyrolysis at different temperatures.

406

Bioresource technology 2015, 192, 23.

407

(33) Fanchiang, W. L.; Lin, Y. C., Catalytic fast pyrolysis of furfural over H-ZSM-5 and

408

Zn/H-ZSM-5 catalysts. Applied Catalysis A General 2012, 419-420, (12), 102-110.

409

(34) Escande, V.; Olszewski, T. K.; Grison, C., Preparation of ecological catalysts derived from

410

Zn hyperaccumulating plants and their catalytic activity in Diels–Alder reaction. Comptes rendus

411

- Chimie 2014, 17, (7-8), 731-737.

412

(35) Biscardi, J. A.; Meitzner, G. D.; Iglesia, E., Structure and Density of Active Zn Species in

413

Zn/H-ZSM5 Propane Aromatization Catalysts ☆. Journal of Catalysis 1998, 179, (1), 192-202.

414

(36) Sun, L. B.; Wei, X. Y.; Liu, X. Q.; Zong, Z. M.; Li, W.; Kou, J. H., Selective Hydrogen

415

Transfer to Anthracene and Its Derivatives over an Activated Carbon. Energy & Fuels 2009, 23,

416

(10), 4877-4882.

417

(37) Uemichi, Y.; Kashiwaya, Y.; Ayame, A.; Kanoh, H., Formation of aromatic hydrocarbons in

418

degradation of polyethylene over activated carbon catalyst. Chemistry Letters 1984, (1), 41-44.

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 38

419

(38) Pereira, M. F. R.; Órfão, J. J. M.; Figueiredo, J. L., Oxidative dehydrogenation of

420

ethylbenzene on activated carbon catalysts. I. Influence of surface chemical groups. Applied

421

Catalysis A General 1999, 184, (1), 153-160.

422

(39) Sui, Z. J.; Zhou, J. H.; Dai, Y. C.; Yuan, W. K., Oxidative dehydrogenation of propane over

423

catalysts based on carbon nanofibers. Catalysis Today 2005, 106, (1), 90-94.

424

(40) Kumar, S., Preparation and Characterization of Acids and Alkali Treated Kaolin Clay.

425

Journal of Veterinary Medical Science 2013, 67, (5), 503-508.

426

(41) Soh, J. C.; Chong, S. L.; Hossain, S. S.; Cheng, C. K., Catalytic ethylene production from

427

ethanol dehydration over non-modified and phosphoric acid modified Zeolite H-Y (80) catalysts.

428

Fuel Processing Technology 2017, 158, 85-95.

429

(42) Peng, H.; Peng, G.; Gang, C.; Bo, P.; Peng, J.; Xing, B., Enhanced adsorption of Cu(II) and

430

Cd(II) by phosphoric acid-modified biochars ☆. Environmental Pollution 2017.

431

(43) S. Bhadury, P.; Pang, J., Chiral BrOnsted Acid-Catalyzed Friedel-Crafts Reaction of

432

Indoles. Current Organic Chemistry 2014, 18, (16), 2108-2124(17).

433

(44) †, D. P. S., ,; José Aguado, A.; Escola, J. M., Catalytic Cracking of a Polyolefin Mixture

434

over Different Acid Solid Catalysts. Industrial & Engineering Chemistry Research 2000, 39,

435

(39), 1177-1184.

26 ACS Paragon Plus Environment

Page 27 of 38 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

436

Tables

437

Table 1. Preparation parameters of ACs.

Activation Activating

Impregnation Impregnation

method

ratio a

ACs agent

one-step

ZnCl2

activation

solution

two-step

KOH

activation

solution

one-step

H3PO4

activation

solution

AC-ZnCl2

AC-KOH

AC-H3PO4

438

a

Activation

Activation

temperature

time

(°C)

(min)

time (h)

1:1

24

600°C

90

4:1

24

800°C

90

2:1

24

500°C

90

Defined as the weight ratio of activating agent to feedstock.

27 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

439

Page 28 of 38

Table 2. Physico-chemical properties of catalysts. Catalysts Properties BC

AC-ZnCl2

AC-KOH

AC-H3PO4

Cd

84.0

77.6

78.1

64.5

Hd

3.4

1.1

0.0

2.4

Nd

2.9

1.5

1.2

1.3

Std

0.1

0.2

0.1

0.04

Od (diff.)

4.8

8.7

9.1

20.5

Ultimate analysis (wt.%)

Proximate analysis (wt.%) Md

4.8

11.0

11.5

11.2

Ad

7.5

11.8

8.4

21.9

Vd

14.7

8.8

7.0

23.5

FCd

73.0

68.4

73.1

43.4

Yield b (wt. %)

28.5

45.8

14.4

59.3

440

28 ACS Paragon Plus Environment

Page 29 of 38 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

441

Energy & Fuels

Table 3. Surface areas and pore properties of different catalysts a. BET

Total

Average

surface

pore

pore

Micropore Sample

volume area 2

442 443 444

volume 3

diameter

Mesopore

Mesopore

volume

ratio c

(cm3/g)

(%)

Micropore (cm3/g)

b

ratio (%)

(m /g)

(cm /g)

(nm)

BC

79.2

0.0587

2.965

0.0323

55.1

0.0157

26.7

AC-ZnCl2

1082.2

0.5859

1.983

0.4834

82.5

0.0662

11.3

AC-KOH

2110.7

1.1290

1.945

0.9327

82.6

0.1086

9.6

AC-H3PO4

600.1

0.5193

3.249

0.2400

46.2

0.2642

50.8

a

All data in Table 3 were calculated with adsorption isotherm. Defined as the ratio of micropore volume to total pore volume. c Defined as the ratio of mesopore volume to total pore volume. b

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 38

445

Figures

446

Figure 1. SEM and EDS results of (a) BC, (b) AC-ZnCl2, (c) AC-KOH, (d) AC-H3PO4, (e)

447

WAC-ZnCl2, (f) WAC-KOH, (g) WAC-H3PO4

448 449 30 ACS Paragon Plus Environment

Page 31 of 38 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

450

Energy & Fuels

Figure 2. High resolution C1s and O1s XPS spectra of different catalysts.

451 452

31 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

453

454 455

Page 32 of 38

Figure 3. Product yields of mixed plastics by direct thermal and catalytic pyrolysis.

(a)

456

32 ACS Paragon Plus Environment

Page 33 of 38 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

457

Figure 4. Content of alkanes, alkenes, aromatics, and other compounds in direct thermal

458

and catalytic pyrolysis oils.

459 460

33 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 34 of 38

461

Figure 5. Concentration of the compositions in (a) pyrolysis oils, (b) aromatic compounds

462

as a function of carbon atoms number.

463 464 465

(a)

(b)

34 ACS Paragon Plus Environment

Page 35 of 38 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

466

Figure 6. Compositions of aromatics in pyrolysis oils classified by number of aromatic

467

rings.

468 469

35 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 36 of 38

470

Figure 7. Hydrocarbons distribution in the gas chromatogram of the direct thermal and

471

catalytic pyrolysis oils.

472 473

36 ACS Paragon Plus Environment

Page 37 of 38 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

474

Energy & Fuels

Figure 8. Dominating reaction mechanisms of different catalytic processes.

475 37 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

476

Page 38 of 38

Figure 9. Cycloalkanes, dienes, and cycloalkene in BC catalytic pyrolysis oil.

477

38 ACS Paragon Plus Environment