Characteristics of Thermal Bitumen Structure as ... - ACS Publications

Aug 3, 2017 - Shandong Energy Longkou Mining Group Company, Ltd., Longkou 265700, ... of thermal bitumen as the pyrolysis intermediate of Longkou...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Florida | Smathers Libraries

Article

Characteristics of thermal bitumen structure as the pyrolysis intermediate of Longkou oil shale Jian Shi, Yue Ma, Shuyuan Li, and Lei Zhang Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b01542 • Publication Date (Web): 03 Aug 2017 Downloaded from http://pubs.acs.org on August 8, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 35

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

Energy & Fuels

1

Characteristics of thermal bitumen structure as the

2

pyrolysis intermediate of Longkou oil shale

3

Jian Shi†, Yue Ma†, Shuyuan Li*,† and Lei Zhang‡

4



5

Heavy Oil Processing , China, Beijing, 102249.

6



7

ABSTRACT: The effects of pyrolysis temperature on the structure of thermal

8

bitumen as the pyrolysis intermediate of Longkou oil shale are investigated through

9

electron paramagnetic resonance spectroscopy, Fourier transform infrared (FTIR)

College of Science, China University of Petroleum-Beijing, State Key Laboratory of

Shandong Energy Longkou Mining Group Co., Ltd., Longkou 265700, China

10

spectroscopy,

nuclear

magnetic

resonance

(NMR)

11

chromatography–mass spectrometry (GC–MS), and X-ray photoelectron spectroscopy

12

(XPS). Results indicate that the free radical concentration of thermal bitumen

13

increases at a maximum temperature of 410 °C and then decreases between 410 °C

14

and 450 °C. Moreover, the g factors of thermal bitumen are slightly higher than 2 and

15

decrease with increasing temperature because of the aromatization of saturates and

16

decarboxylation. FTIR analysis indicates that decarboxylation is completed before the

17

temperature reaches 370 °C. NMR analysis shows that aliphatic and aromatic

18

compounds comprise more than 80% and 15%–16% of thermal bitumen, respectively.

19

During pyrolysis, the fraction of aliphatic compounds, branched alkane, and the

20

average methylene chain length decrease because of secondary cracking, and the C–C

21

bond at the branch chainis easily cracked into gas and light oil.

22

ACS Paragon Plus Environment

spectroscopy,

gas

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 35

23

1. INTRODUCTION

24

Declining oil price restricts the development of the oil shale industry; nonetheless,

25

researchers still focus on the research and use of oil shale, which is an important

26

alternative energy resource, because of energy demand1,

27

macromolecular organic matter in oil shale is called kerogen, which can be pyrolyzed

28

to produce shale oil and gas through a series of complex chemical reactions. To

29

investigate the pyrolysis mechanism of oil shale kerogen, various studies analyzed the

30

formation and molecular structure3, 4, pyrolysis product yield5, 6, retorting technologies

31

2, 7, 8

32

compounds of oil shale12, 13. The pyrolysis of oil shale is affected by numerous factors,

33

such as its origin and different pyrolysis modes14. Nevertheless, the pyrolysis

34

processes can still be divided into two steps. The first step involves the

35

depolymerization of macromolecular organic matter in kerogen into soluble extracts,

36

including thermal bitumen and tiny gas products. In this step, macromolecules and

37

thermally unstable heterocyclic compounds in kerogen disintegrate into small

38

molecules15. The second step involves the further decomposition of thermal bitumen

39

into shale oil, gas, and semicoke as the temperature increases.

2

. The complex

, pyrolysis mechanism and kinetics model9-11, and heteroatom-containing

40

Previous studies characterized the structure and composition of kerogen through

41

thermogravimetry, Fourier transform infrared (FTIR) spectroscopy, nuclear magnetic

42

resonance (NMR) spectroscopy, pyrolysis gas chromatography–mass spectrometry

43

(Py-GC/MS), and X-ray diffraction16-18. However, the molecular structure of oil shale

44

kerogen remains to be completely described because of its complex composition. In

ACS Paragon Plus Environment

Page 3 of 35

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

45

addition, kerogen from different sources considerably varies in constitution and

46

construction, and these factors complicate the analysis of the pyrolysis process.

47

Recent studies have focused on the thermal decomposition intermediates (thermal

48

bitumen) of oil shale kerogen and explored preliminarily the characteristics of thermal

49

bitumen9. However, the influence of pyrolysis temperature on the structure of thermal

50

bitumen has yet to be clarified. The formation and decomposition of thermal bitumen

51

are continuous during pyrolysis; hence, exploring the composition of thermal bitumen

52

obtained at different temperatures can further explain the pyrolysis behavior of

53

kerogen.

54

The present study measured the yields of pyrolysis products at different

55

temperatures and explored the structural features of thermal bitumen at pyrolysis

56

temperatures of 330 °C, 370 °C, 410 °C, and 450 °C. Quantitative information about

57

the chemical structure and distributions of samples was obtained through electron

58

paramagnetic resonance (EPR) spectroscopy, FTIR spectroscopy, NMR spectroscopy,

59

GC–MS, and XPS. The information derived from the chromatograms provides new

60

insights into the pyrolysis process of oil shale.

61

2. EXPERIMENTAL SECTION

62

2.1. Materials. Oil shale samples were obtained from Longkou City, Shandong

63

Province, China. The oil content was 16.67%. Proximate and ultimate analyses of the

64

samples are illustrated in Table 1 and Table 2. The samples were crushed and screened

65

to 100 mesh size (particles less than 0.15 mm in size). Soluble bitumen in oil shale

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

66

was extracted for two days by a Soxhlet extractor with chloroform (CHCl3) as solvent

67

at 65 °C to 70 °C to ensure that the samples did not contain extracted bitumen. The oil

68

shale samples were dried at 50 °C to a constant weight and then stored for use in a

69

desiccator.

70

2.2. Thermal bitumen preparation and characterization. A retorting reactor

71

made of stainless steel was selected and modified based on the Fischer assay-type

72

retort. A thermocouple was used to measure the temperature of the sample center. For

73

each test, 50g of oil shale was loaded into the retorting reactor, which was then slowly

74

heated from room temperature to the pyrolysis temperature at a heating rate of

75

2 °C·min−1 to guarantee uniform temperature distribution in the sample19, 20. Upon

76

reaching the temperature, the reactor was quickly cooled down to room temperature,

77

and then the gas, shale oil, and semicoke were weighed. Thermal bitumen was

78

extracted from the semicoke by using the Soxhlet extractor with chloroform solvent

79

(CHCl3) at 65−70 °C for 48 h. The chloroform solution was finally evaporated to

80

obtain thermal bitumen and then maintained in a cold storage for further analysis.

81

EPR measurements were performed by using a Bruker A200 spectrometer with a 100

82

kHz field modulation. In this experiment, the scanned parameters were kept at room

83

temperature, and the operation conditions were used as follows: sweep time, 4min;

84

center field, 323.2 mT; frequency, 9.06 GHz; and microwave power, 0.5 mW. The

85

signals from Mn2+ dispersed in MnO2 were used to monitor EPR sensitivity. The g

86

values play an important role in reflecting the characteristics of the magnetic field

87

within the molecules. The change in g values mainly depends on the coupling effect

ACS Paragon Plus Environment

Page 4 of 35

Page 5 of 35

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

88

of the spin and orbital motions. The EPR signals were located in experiments by their

89

g value defined by

90

g=

91

where h is the Planck’s constant,  the resonance frequency,  the Bohr magneton

92

and H the magnetic field at which resonance occurs21. The free radical concentration

93

of thermal bitumen served as a quantified reference to the standard curve of

94

1,1-diphenyl-2-picrylhydrazyl (DPPH, g=2.0036)22, 23. For each test, 0.05 g of the

95

sample was diluted 40 times with methylbenzene, and this step was repeated three

96

times to ensure repeatability.

 

97

FTIR was used to analysis the factional group of the thermal bitumen. FTIR

98

spectra were obtained using a Bruker IFS 66 V/S FTIR spectrometer (Germany). The

99

samples were prepared by grinding 1 mg of thermal bitumen and homogenizing it

100

with 100 mg of ground KBr (dried under an infrared lamp). The spectra were

101

collected in the mid-IR region from 4000 cm−1 to 400 cm−1.

102

13

C and

1

H NMR curves were obtained at 100.62 MHz by a Bruker

103

AVANCEIII−400 MHz with a 5 mm probe insert and TMS as the internal standard.

104

To guarantee the total solution organic matters, CDCL3 was used as the solvent. The

105

acquisition times of

106

6.1.1 was used to analyze the NMR spectra and quantify the relative proportion of

107

different carbon and hydrogen types in the thermal bitumen samples.

13

C and 1H NMR were 3 and 0.2 s, respectively. MestReNova

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

108

GC–MS was performed with ACION TQ (Bruker) to determine the composition of

109

thermal bitumen. The type of chromatographic column was HP-5MS (30 m × 0.25

110

mm × 0.25 μm) fused silica capillary column. The oven temperature was maintained

111

at 20 °C for 5 min and then slowly increased to 280 °C at a rate of 10 °C min−1,

112

followed by holding the pyrolysis temperature for 20 min. The analysis was

113

performed by Qual Browser and Technology mass spectral library search24-26. To

114

qualitatively and quantitatively analyze the peaks in total ion chromatogram, the

115

curves was compared with the spectrum in NIST. The chromatographic column was

116

cleaned by CHCl3, there was about 10% remained in the column, it revealed that the

117

macromolecular structures in thermal bitumen were hardly detected by GC–MS.

118

Compared with the spectrum in NIST, only the compositions with a carbon number

119

that is smaller than C30 were analyzed in the experiments.

120

The existing form of heteroatoms in thermal bitumen was measured using XPS by a

121

Kratos Axis Ultra DLD spectrometer with Al Kα X-ray source (hm = 1486.6 eV) at

122

15 kV and 150 W. The parameters of different element peaks were fitted and

123

established with XPSPEAK41.

124

125

Table 1. Proximate and ultimate analyses of Longkou oil shale (wt.%).

126

127

Table 2. Ultimate Analyses of Longkou oil shale and thermal bitumen (wt.%).

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

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

128

129

Figure 1. Separation and structure analyses of thermal bitumen

130

3. RESULTS AND DISCUSSION

131

3.1. Effects of pyrolysis temperature on pyrolysate yields. The yields of

132

semicoke, thermal bitumen, shale oil, and gas as a function of temperature from

133

290 °C to 480 °C are shown in Fig.2. The yields of shale oil, water and gas increase,

134

whereas that of semicoke significantly decreases as the temperature is increased from

135

290 °C to 480 °C and reaches the peak at 480 °C. The yield of thermal bitumen

136

reaches the maximum value of 7.45% at 370 °C and then decreases when the

137

pyrolysis temperature is further increased. With increasing temperature, kerogen in

138

the oil shale breaks down into thermal bitumen, oil, and gas, while the thermal

139

bitumen decomposes into oil and retorting gas. The formation of thermal bitumen

140

starts at 290 °C, and the rate of thermal bitumen formation is faster than that of

141

decomposition at 370 °C. Moreover, the decomposition rate of thermal bitumen

142

increases at 370–450 °C and converted to the final pyrolysates after 480 ℃ ,

143

completely. This results is consistent with the previous studies27, 28.

144

145

Figure 2. Effect of pyrolysis temperature on pyrolysate yields.

146

147

3.2. EPR analysis. The typical EPR spectra of thermal bitumen are illustrated in

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 35

148

Fig. 3(a), and the free radical concentration (Ng) of DPPH solution in different

149

volumes can be calculated as follows:

150

N = (ω × m × 6.02 × 10 )⁄M

151 152

(1)

where m is the mass,ω is the percentage (wt. %), and M is the molecular mass of DPPH solution.

153

The relationship of Ng versus signal/marker (area of the DPPH sample absorption

154

curve vs. Mn2+ curve) and the standard curve of the radical concentrations based on

155

numerical analysis is given by the following:

156

Signal⁄marker = 8.887 × 10#$% N − 1.571

(2)

157

The Ng and g factors of thermal bitumen at different temperatures are shown in Fig.

158

3(b). Similar to the field trend of thermal bitumen, the Ng of thermal bitumen

159

increases with increasing temperature, reaches the peak at 410 °C, and then decreases

160

between 410 °C and 450 °C. The dynamic change in Ng indicates that the unpaired

161

electrons in thermal bitumen are constantly forming and disappearing during

162

pyrolysis.

163

164

Figure. 3. Effect of temperature on the radical concentrations and g factors of thermal

165

bitumen. (a) EPR spectra of thermal bitumen, (b) radical concentrations and g factors

166

of thermal bitumen.

167

ACS Paragon Plus Environment

Page 9 of 35

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

168

The original paramagnetic centers in coals mainly exist in aromatic units, and their

169

unpaired electrons may be found in C, N, or O atoms26. The gfactors of thermal

170

bitumen are slightly higher than 2 and decrease with increasing temperature. Petrakis

171

and Grandy29 found that the g values for aliphatic hydrocarbon π radicals are

172

2.0025–2.0026, 2.0025 for aromatic hydrocarbon π radicals, 2.0008−2.0014 for

173

σ-type oxygen-containing radicals, 2.0038−2.00469 for π-type oxygen-containing

174

radicals,

175

sulfur-containing radicals. The g factors shown in Fig.3(b) indicate that the unpaired

176

electrons in thermal bitumen are mainly localized in aliphatic hydrocarbons, aromatic

177

hydrocarbons, and O atoms, and that the reduction of g factors is caused by the

178

aromatization of saturates and decarboxylation.

2.0031

for

nitrogen-containing

radicals,

and

2.0080−2.0081

for

179

180

3.3. FTIR analysis. Fig.4 shows the FTIR spectra of thermal bitumen at different

181

temperatures. The strong peaks at 2829 and 2850cm−1 are attributed to the asymmetric

182

and symmetric stretching of the C−H bond of methylene groups, respectively

183

The absorption bands at 1463 and 1384 cm−1 correspond to the symmetric bending of

184

methyl groups. The peak at 740 cm−1 is attributed to the in-plane bending vibration of

185

-CH2-. As the temperature increases, the FTIR spectra exhibit minimal changes at the

186

peak of 1722 cm−1, which is related to the C=O stretching of carbonyl and/or carboxyl

187

groups (Fig. 4), and this peak weakens as the temperature increases. Moreover, the

188

peaks at 1463 cm−1 and 1384 cm−1 FTIR analysis indicates that thermal bitumen

189

mainly consists of aliphatic alkyl groups. The peaks at 1722 cm-1 and 1619 cm-1

ACS Paragon Plus Environment

14, 30

.

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

190

decrease from 410℃ to 450℃, which indicates that the end methyl start decompose

191

at this temperature. In addition, the carbonyl and/or carboxyl groups decrease as the

192

temperature increases because of decarboxylation. The strong peak at 3425 cm−1 is

193

likely contributed by adsorbed water.

194

195

Figure 4. FTIR spectrogram of thermal bitumen at different temperatures.

196

3.4. NMR analysis. The 13C NMR and 1H NMR curves of the different samples are

197

similar, indicating that the compounds of all samples demonstrate an overall similar

198

distribution. Thus, the 13C NMR and 1H NMR techniques are only employed for the

199

sample at 370 °C (Fig.5), and the distribution of carbon and hydrogen is illustrated in

200

Tables 3 and 4.

201

The

13

C NMR spectra exhibit signals at 15, 20, 21–25, 25–36, and 36–50 ppm,

202

which are assigned to the carbons of n-alkanes, where as the signals at 109–130,

203

130–137, 137–148, and 148–165 ppm correspond to aromatic carbons (Fig. 5(a), the

204

signal at 77 ppm is assigned to the solvent of the NMR measurement). The 1H NMR

205

spectra are dominated by the signals corresponding to aliphatic protons on saturated

206

carbons between 0.4 and 1.9 ppm, and the range of peak between 1.9 and 9.0 ppm is

207

caused by the protons in the aromatic carbons (Fig. 5(b), the signal at 7.25 ppm is

208

assigned to the solvent of the NMR measurement) 31-33.

209

Fig. 5(a) shows that the aliphatic compounds comprise over 80% of the integral

210

area in the spectrogram. This phenomenon indicates that the carbon skeletal structure

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

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

211

of thermal bitumen is mainly composed of aliphatic carbons and aromatic (110–165

212

ppm) and carboxyl carbons (188–220 ppm). The structure parameters of the carbon

213

skeletal structure can be calculated by using the following formulas:

214

. / 0 f)* = f)*+ + f)* + f)* + f)* + f)*

(3)

215

2 . 3 2 f)1 = f)1 + f)1 + f)1 + f)1

(4)

216

0 B5 = f)* /f)*

(5)

217

C8 = 9@:;=9:;
:?

218 219

(6)

:?

Where fal and far refer to the alkyl and aromatic carbons, respectively; BI is the degree of branching, and Cn stands for the average chain length.

220

The same result is also supported by the 1H NMR spectra shown in Fig. 5(b), and

221

the chemical shift ranges of 0.4–1.0 ppm and 1.0–1.9 ppm are assigned to the methyl

222

and methylene groups in the aliphatic structures, respectively, accounting for over 80%

223

of the total protons

224

formulas:

225

H)* = H)* + H)*

β

(7)

226

α ) H)1 = H)1 + H)1 + H)1

B

(8)

227

The calculation results are shown in Table 4.

228

Figure 5. 13C NMR and 1H NMR of the thermal bitumen at 370 °C

229

Table 3. Distribution of various carbons in thermal bitumen from 13C NMR.

γ

34-36

. The proton types can be calculated by using the following

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

230

Table 4. Distribution of various hydrogens in thermal bitumen from 1H NMR.

231

Table 5. Structural parameters of thermal bitumen.

232

The fraction of aliphatic structure decreases and the fraction of aromatic carbon

233

increases as the temperature increases (Table 5). Moreover, the pyrolysis processing

234

reduces the fraction of branched alkane and the average methylene chain length. In

235

consideration of prior research about the composition of oil shale kerogen, shale oil,

236

and semicoke, the branched alkane and the average methylene chain length of thermal

237

bitumen are less than those of kerogen and significantly greater than those of shale oil

238

and semicoke. This result proves that the bitumen is the intermediate of oil shale

239

pyrolysis. The formation and decomposition of thermal bitumen simultaneously occur.

240

Compared with other compounds, branched alkanes are fewer because the C–C bond

241

at the branch point is easier to crack, and the branch group disintegrates and tends to

242

form short chains.

243

The fraction of aromatic compounds in thermal bitumen is about 15%–16%, and

244

about 15–16 aromatic carbons per 100 carbons are calculated. This result indicates the

245

presence of 2.5 homologs of benzene or naphthalene and benzene ring compounds.

246

During pyrolysis, the fraction of aromatic compounds increases because of the

247

aromatization of aliphatic and aromatic compounds composited by the aromatization

248

of saturates that accumulate in thermal bitumen.

249

3.5. GC–MS analysis. GC–MS analysis indicates that thermal bitumen can be

250

classified into aliphatic hydrocarbons, aromatic compounds, and hetero atomic

ACS Paragon Plus Environment

Page 12 of 35

Page 13 of 35

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

37

251

compounds according to their structures

. As shown in Fig.6, over 70% of the

252

detectable substances in thermal bitumen are straight carbon chained aliphatic

253

hydrocarbons mainly ranging from C6–C30. The weak peaks at the front part of the

254

GC–MS spectra are possibly caused by phenols, ketones, alkyl benzenes, alkenes, and

255

alcohols. Polycyclic aromatic compounds and nitrogen and sulfur compounds appear

256

at the adjacent carbon number of aliphatic hydrocarbon chromatographic peaks. Other

257

compounds, including naphthalene, anthracene, phenanthrene, carboxylic, and lipid,

258

can also be detected by GC–MS.

259

260

Figure 6. GC–MS of thermal bitumen at different pyrolysis temperatures

261

Table 6. Relative contents of each compound of thermal bitumen.

262

263

Figure 7. Carbon number distribution of n-alkanes

264

265

The constitution of thermal bitumen and the intensive overlap in the gas

266

chromatogram are complex. Thus, the quantitative information of different

267

compounds from GC–MS may not be reliable. However, the variation tendency of

268

compounds with temperature changes can be analyzed by GC–MS. The relative

269

contents of n-alkanes, α-alkenes, monocyclic aromatics, polycyclic aromatics, and

270

oxygenated compounds are highly variable at 330 °C to 370 °C (Table 6). Compared

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

271

with the GC/MS curves, some low molecular monocyclic aromatics can be detected at

272

330 °C, such as phenolic homologues but disappear gradually with increasing

273

temperature. Moreover, the relative content of aromatic compounds increases between

274

370 °C and 450 °C. This phenomenon indicates that thermal bitumen is mainly

275

generated from 330 °C to 370 °C. The low molecular monocyclic aromatics are hardly

276

accumulated within thermal bitumen.

277

The carbon-number distribution of thermal bitumen is summarized in Fig. 7. The

278

relative content of low molecular n-alkanes increases with increasing temperature

279

from 330 °C to 370°C and then significantly decreases from 370 °C to 450 °C. The

280

amounts of ΣC>25 decrease obviously from 330 ℃ to 370℃ and the relative

281

content of low molecular hydrocarbons (