You are on page 1of 18

Accepted Manuscript

Title: Effects of Temperature on Pyrolysis Characteristics of


Indonesia Oil Sands

Authors: JIA Chuxia, XIAO Yunpeng, YU Hao, GONG


Shishang, WANG Qing

PII: S0165-2370(18)30595-3
DOI: https://doi.org/10.1016/j.jaap.2018.11.021
Reference: JAAP 4486

To appear in: J. Anal. Appl. Pyrolysis

Received date: 4 July 2018


Revised date: 26 October 2018
Accepted date: 16 November 2018

Please cite this article as: Chuxia J, Yunpeng X, Hao Y, Shishang G, Qing W, Effects of
Temperature on Pyrolysis Characteristics of Indonesia Oil Sands, Journal of Analytical
and Applied Pyrolysis (2018), https://doi.org/10.1016/j.jaap.2018.11.021

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Effects of Temperature on Pyrolysis Characteristics of Indonesia Oil
Sands

JIA Chuxia1, XIAO Yunpeng1, YU Hao1, GONG Shishang2, WANG Qing1

(1Engineering Research Centre of Oil Shale Comprehensive Utilization Ministry of Education , Northeast Electric Power
University , Jilin 132012, Jilin, China; 2Energy China Northwest Electric Power Test Research Institute Co LTD , Xi’an
710032, Shaanxi, China)

T
————————————————

IP
First author: Jia Chunxia, Email:jiachunxia_215@126.com.
Contact author: Wang Qing, Email:rlx888@126.com.

R
SC
Highlights
 The pyrolysis products of oil sand samples were studied.
 The total organic matter yield first increased, and then, decreased with the increase in pyrolysis final


temperature.
U
The nitrogen adsorption capacity, BET specific surface area, and pore volume of the two oil sand
N
samples reached the maximum peak at 450 °C.
 The change in fractal dimension D was slightly lower than the nitrogen adsorption fractal value.
A
 SEM images provided visual pore structure information of the semi-focal surface of oil sands.
M
ED

Abstract: In this work, pyrolysis experiments were conducted on Indonesia oil sands in a Cracking
device at different pyrolysis temperatures. The effects of pyrolysis temperature on total hydrocarbon
yield, liquid hydrocarbon production rate and gas production rate were investigated. Below 300 °C,
PT

the pyrolysis products were mainly gaseous. However, the proportions of liquid and gas products
were identical above the temperature of 300 °C. The relationship between the volume fraction of
gaseous products and temperature was analyzed using gas chromatography. The presence of
E

functional groups of methyl, methylene, aldehyde, ketone, acid, fat, alkene and aromatic
hydrocarbons in liquid pyrolysis products was investigated at different pyrolysis temperatures. The
CC

pore structure, electron microscope and fractal dimension of the chars were also analyzed. The result
showed that the amount of adsorbed nitrogen reached the maximum value at 450 °C. The adsorption
loop of char was close to the B-type loop, and the pore was mainly the slit hole. In addition, the pore
A

structure was very well developed. The BET specific surface area and pore volume presented their
peak values at 450 °C and the fractal dimension was obtained using FHH equation, which peaked at
the temperature of 450 °C. The pore structure was loose and there was no obvious connection
between the particles. Finally, the overall fractal value was slightly lower than the nitrogen adsorption
fractal value.
Key words: oil sand; hydrocarbon generation; oil sand char; retorting temperature; adsorbent; SEM;
fractal dimension
1 Introduction
Due to the shortage of energy resources in the world, countries with higher energy consumptions in the world
are constantly seeking new energy resources to replace the conventional ones. Among unconventional energy
resources, oil sands have attracted worldwide attention. Oil sands are composed of a series of quartz sand and clay
composition Bonanza. The sand grains are filled with different amounts of heavy, high-viscosity bitumen. It is a
kind of sand or sandstone containing tar and bitumen. It is also called tar sand and belongs to unconventional
petroleum resources. Countries with high oil sand resources include Canada, Russia, Venezuela, the United States,
and China [1-3]. The oil sand resources of the world are abundant, accounting for nearly one-third of the world's oil
reserves. As an important alternative resource for future energy supply, oil sand pyrolysis is one of the best ways

T
to observe and study the characteristics of oil and gas production [4-5]. In the pyrolysis of oil sands, the temperature

IP
has a decisive influence on the total organic production, the yield of liquid organic matter, the yield of gaseous
organic matter, and the mechanism of generation of char during the pyrolysis of oil sands. Therefore, it is

R
necessary to explore the influence of temperature on the generation characteristics of organic matter during the
pyrolysis process. The char produced during the pyrolysis has low calorific value, low volatility, high ash content,

SC
and causes serious environmental pollution. Research on the strategic thinking of oil refining and building
materials is an effective way to realize the comprehensive utilization of oil sand resources. The analysis of pore
structure of char products during the pyrolysis of oil sands is an important part of the comprehensive utilization of

U
oil sand resources. The parameters, such as heating rate and final pyrolysis temperature, have a great influence on
the properties of char pore structure in the pyrolysis process [6-8].
N
In this study, pyrolysis experiments of oil sands were conducted in a small fixed-bed at the final temperatures
of 250 °C, 300 °C, 350 °C, 400 °C, 450 °C, 500 °C, and 550 °C. The experiments examine the effect of final
A
pyrolysis temperature on the total organic production, liquid organic matter yield, and yield of gaseous organic
M

matter during the pyrolysis process. Additionally, char produced at different final temperatures was collected. A
fast specific surface area measuring instrument was used to study the surface characteristics of char. The
characteristics of the char structure was described by the fractal theory. Because the application of fractal theory of
ED

char can be thought as an important method for describing the characteristics of the structure, and reflects the
degree of pore roughness on the surface of the char surface and the irregularity of the sample surface. Based upon
the experimental results, the influence of various factors on the char surface structure parameters was analyzed.
PT

Furthermore, the mechanism of formation of organic matter was explored during pyrolysis. The study provides
theoretical basis for comprehensive utilization of oil sands, such as fixed bed pyrolysis and fluidized combustion
of char products.
E

2 Experimental
CC

2.1 Samples
Two kinds of oil sands, supplied by Burton Island, Indonesia, were used and labeled as YN 1 and YN 2. The
results for the calorific value, proximate analysis and ultimate analysis of the oil sands are presented in Table 1.
A

Table 1 Proximate and ultimate analyses of YN 1 , YN 2 and the chars of YN 1 , YN 2

Calorific
Ultimate analysis / % / ad Proximate analysis /%/ad
Simple Value/kJ·kg-1

C H N S H/C Qnet,ar M V A FC CO2

YN 1 19.71 2.42 0.23 0.92 1.47 8814.22 1.41 20.63 49.46 0.71 27.79
YN 2 19.34 2.43 0.19 1.02 1.51 6714.71 1.76 19.39 48.15 0.85 29.85
Char of 7.03 0.37 0.17 0.06 0.63 2064.00 0.37 3.64 60.92 0.88 34.19
YN 1
Char of
6.20 0.36 0.15 0.06 0.70 1479.25 0.36 0 61.55 1.09 37.00
YN 2
The CO2 content in Table 1 is the decomposition of CaCO3 at high temperatures. The carbon content in the elemental analysis does
not include this portion of carbon. Oil sand char is the solid product of the analysis of oil content (Aluminum crucible experiment).
M: moisture, A: ash, V: volatile matter, Fc: fixed carbon.
Table 2 presents the results for oil content and asphaltene yield for the two kinds of oil sands samples.
Table 2 Results of Fischer assay and Dean Stark Soxhlet extraction

Simple Oil content/﹪ Water content/﹪ Char content /﹪ Gas +loss/﹪ Bitumen content/﹪

T
YN 1 16.80 1.80 78.38 3.02 23.11

IP
YN 2 13.28 2.28 81.30 3.14 18.26

R
The results of the Fischer assay and Soxhlet extraction experiments for the two samples are also provided in

SC
Table 2. The results show that, the oil yield of YN 1 was 8.40 g, which was higher than YN 2. The yield of asphalt
from YN 1 was 14.64%, which was around 1.5 times that of YN 2. Furthermore, YN 1 has a lower char yield,
water yield, and air loss than YN 2. In summary, the quality of YN 1 was better than that of YN 2.
2.2 Cracking device
U
Oil sand pyrolysis was conducted in a small pyrolysis furnace, as shown in Figure 1. The main part was the
N
heating furnace, where the test sample reacts. N2 was used as the heat carrier gas. In order to exclude the air from
the test section and remove the external water of the sample, at the beginning of the experiment, the nitrogen for
A
purging air was heated to 50 °C in the preheating furnace and then passed to the test section. The heat required for
pyrolysis of the sample was provided by the test section. The preheating furnace and the reaction furnace were all
M

made of 1Cr18Ni9Ti stainless steel heat-resistant tube, which was thermally resistant to temperature up to 700 °C.
The tube’s specifications were: φ53*3mm; length 100 mm. A temperature controller was used to control the
ED

heating of external heating electric wire. The reactor was sealed using two nuts. The samples were ground to a
particle size of less than 3 mm. The sample was dried in a dry box to a constant temperature and N2 was used as
the carrier gas. The samples were placed in the reactor, as shown in Figure 1. A temperature controller was used to
set the heating rate to values of 20 °C/min, while the final temperatures were varied through values of 250 °C,
PT

300 °C, 350 °C, 400 °C, 450 °C, 500 °C, and 550 °C. After reaching the final temperature of pyrolysis, the
temperature was kept for 20 minutes. After cooling to room temperature, chars were collected. A natural cooling
E

process at room temperature was applied, and the nitrogen environment was maintained throughout the cooling
process.
CC

2 3 4

6
5
A

1 7

11

12

10 9

1-N2 Bottle 2-Mass Flow Meter 3-Pressure Gauge 4-Preheater 5-Temperature Controller 6-Reaction Tube 7-Reactor 8-Heater Thermostat 9-Erlenmeyer
10-Cold Trap 11-pressure gauge 12-wet gas flow meter

Figure 1. Diagram of oil sand pyrolysis device


2.3 Gas Chromatography and Infrared Spectroscopic Analysis
Shimadzu's gas chromatograph (GC-2014) was used for composition analysis. The liquid product was
analyzed using US Spectrum-2 infrared absorption spectrometer. The experimental instrument of Infrared
spectrum analyzer is a Perkin-Elmer Fourier transform infrared spectrometer with attenuated total reflection
accessory(ATR). The spectral range is 4000cm-1~400cm-1. The resolution was set to be 4 cm-1, while the
wavelength accuracy was 0.01 cm-1. Reference gas is air. The cumulative number of scans is 32. A little oil sand
bitumen is taken out and completely covered on the entire surface of the ATR crystal. Pressurization allows the
bitumen sample to be tightly bonded to the ATR crystal. Measuring the infrared spectrum of a sample. The same

T
sample was repeatedly measured 3 times, and its average spectrum was taken as the infrared spectrum of the

IP
sample. The sample on the ATR crystal is cleaned with a cottonseed solution of the appropriate amount of xylene
before each loading.

R
2.4 Determination of low temperature nitrogen adsorption isotherm curves
The pore structure was measured using low-temperature nitrogen adsorption method. The experiments were

SC
performed on TriStat II 3020 high-performance multi-channel automatic surface area and porosity analyzer from
McMurray Tek Instruments Pty. Ltd., Canada. Nitrogen was at liquid nitrogen temperature (-196 °C). The relative
pressure p/p0 was within the range of 0.01 - 0.995 (p and p0 are the equilibrium and saturation pressures for

U
nitrogen cryogenic adsorption, respectively). Physical adsorption was performed under the above-mentioned
conditions to obtain nitrogen adsorption and desorption isotherms and other parameters of the char samples.
N
2.5 SEM experiments
The scanning electron microscope (SEM), used in these experiments, was the hot-field emission scanning
A
electron microscope of Japan’s JEOL JSM-7610. The instrument was used to directly observe the surface
M

microscopic morphology and structural characteristics of the sample. First, the sample was placed on the sample
holder and sprayed to make it electrically conductive. The resolution is secondary electron image: 1.0 nm (15kv).
The image was magnified 2,000 times to observe the changes in particle and pore morphology of the sample. The
ED

image was a grayscale image, while the storage format was BMP. The pixel matrix size was 1024×1280.

3 Results and discussion


PT

3.1 Analysis of oil sand organic matter formation characteristics


Organic matters produced from the pyrolysis of oil sands were divided into liquid and gaseous products.
Liquid organic matter consisted of the liquid produced by the pyrolysis gas generated during the pyrolysis process,
E

which was condensed by the cooling system. Gaseous organic matter was the organic matter, which remained in
the gaseous form (mostly methane). Figure 2 shows the variation in the yield of gaseous and liquid organic matters
CC

with the pyrolysis final temperature.


A
12 200 12
120
Gaseous 180 Gaseous
Liquid Liquid
10 10
Gaseous organic production( mg/g)

Gaseous organic production( mg/g)

Liquid organic production( mg/g)


Liquid organic production( mg/g)
160
100
140
8 8
80
120
6 100 6
60
80
4 4
40
60

2 40 2 20
20
0 0 0
0

T
250 300 350 400 450 500 550 250 300 350 400 450 500 550

Temperarure( ℃) Temperarure( ℃)

IP
(a)YN 1 (b)YN 2
Fig. 2. Production of gaseous and liquid organic matters at different final pyrolysis temperatures

R
As can be seen from Fig. 2, the yield of gaseous organic matters gradually increased with the increase in

SC
pyrolysis final temperature. At different final pyrolysis temperatures, the yield of gaseous organic matter in YN 1
oil sands was similar with that of YN 2 oil sands. When the final temperature of pyrolysis was 250 °C, the yield of
liquid organic matter was low, and the yield of gaseous organic matter was similar with that of the liquid organic
matter, indicating that the gaseous organic matter will begin to be produced at lower temperatures. When the
temperature was higher than 300 °C, the production of liquid organic matter sharply increased. When the final U
N
pyrolysis temperature reached 400 °C, the increment in liquid organic matter decreased, indicating that the oil
sand organic matter has basically been completely pyrolyzed. The production of liquid organic matter reached its
A
maximum value at the final temperature of 450 °C. When the pyrolysis final temperature was higher than 450 °C,
M

the production of liquid organic matter began to decline, which may be due to the higher pyrolysis temperature
and secondary cracking. From Figure 2, it can be concluded that the optimal pyrolysis final temperature was
450 °C, at which point, the liquid organic matter yield was maximum. Furthermore, a comparison with the data
ED

presented in Table 2 shows that YN 1 oil sands exhibited high oil content and strong organic matter production
capacity.
Figure 3 shows the relationship between the total organic matter yield and the final pyrolysis temperature. As
PT

can be seen from Fig. 3, the total organic matter yield of YN 1 oil sands gradually increased with the increase in
final temperature of pyrolysis. Within the temperature range of 300 - 450 °C, the content of total organic matter
increased by 172.94 mg/g. This phase is the main phase for the production of oil sand organic matter. At 450 °C,
E

the total output of YN 1 oil sand organic matter was 183.27 mg/g. When the temperature was higher than 450 °C,
the increments in total organic matter decreased. However, there was still a large amount of liquid organic matter,
CC

which belonged to the main pyrolysis stage. This was because the YN 1 oil sands had a high potential for
generating liquid organic matter below 450 °C. The formation of organic matter in the pyrolysis of YN 2 oil sands
was similar to that of YN 1 oil sands, however the yield of organic matter was lower than that of YN 1 oil sands.
A

With the increase in final pyrolysis temperature, the total organic yield of YN 2 oil sands first increased, and then
stabilized [9]. The total organic matter yield was 125.58 mg/g at 450 °C. The final pyrolysis temperatures for the
total organic matter yields of 106.45 and 99.08 mg/g were 500 °C and 550 °C, and showed a smooth trend with
time.
200
Total organic matter yield(mg/g)

150
YN 1
YN 2
100

50

T
250 300 350 400 450 500 550
Temperarure(℃)

IP
Fig. 3. Relationship between the total yield of organic matter and the final pyrolysis temperature

3.2 Oil sand pyrolysis gas product emission characteristics

R
In the pyrolysis experiments, gas generated from the pyrolysis of oil sands at different final temperatures was
collected using an aluminum foil gas collecting bag. A gas chromatographic analyzer was used to analyze the

SC
gaseous products. Due to the limitations of gas chromatography, only about 14 hydrocarbon gases were detected.
These gases were methane, ethane, isobutane, n-butane, isopentane, n-pentane, ethylene, propylene, butylene,
n-butene, cis butene, 1,3-butadiene, acetylene, and propyne. Two non-hydrocarbon gases detected were carbon
monoxide and carbon dioxide [10]. Figures 4 and 5 show the volume fractions of each of these gases.

20 Methane 20 MethaneU
N
Ethane Ethane
Ethylene Ethylene
15 15
A
Content( %)
Content( %)

10 10
M

5 5
ED

0 0

250 300 350 400 450 500 550 250 300 350 400 450 500 550
Temperature(℃) Temperature(℃)
PT

(a)YN 1 (b)YN 2
Fig. 4. Content of methane, ethane and ethylene at different final pyrolysis temperatures
E

The main components of gaseous hydrocarbons were methane, ethane and ethylene. Their molar ratio was
CC

around 20 mol/mol at 550 °C and were the main gaseous organic products. From Figure 5, it can be seen that the
YN 1 and YN 2 oil sands presented similar variation trends with temperature. The content of methane rapidly
increased at 300 °C. This is due to the initial decomposition of oil sands’ organic matter. Within the temperature
range of 350 - 400 °C, the molar ratio of methane remained unchanged, and the amount of methane gas produced
A

was less. Therefore, the temperature range of 400 - 550 °C was chosen as the main temperature range for pyrolysis.
At this stage, a large number of oil sand organics began to decompose, and the content of gaseous organic matter
increased. With the temperature range of 500 - 550 °C, the YN 2 oil sands tended to decrease. It is likely that the
methane gas participated in the secondary reaction within this temperature range, due to which, its molar ratio
decreased.
1.5 1.0
Isobutane
Isobutane
Butane
Butane 2-Methylbutane
0.8
2-Methylbutane n-Pentane
1.0 n-Pentane
0.6
Content( %)

Content( %)
0.4
0.5

0.2

0.0 0.0

T
250 300 350 400 450 500 550 250 300 350 400 450 500 550
Temperature(℃) Temperature(℃)

IP
YN 1 YN 2

R
(a)Alkanes
1.6

SC
1.4
1.4 Propylene Propylene
Trans-butene 1.2 Trans-butene
1.2 n-Butene n-Butene
Isobutene 1.0 Isobutene
1.0 Cis butene

U
Cis butene
Content( %)

1,3-Butadiene 1,3-Butadiene
Content( %)

0.8 0.8
N
0.6 0.6
0.4
0.4
A
0.2
0.2
0.0
M

0.0
-0.2
250 300 350 400 450 500 550
250 300 350 400 450 500 550
Temperature(℃) Temperature(℃)
ED

YN 1 YN 2

(b)Olefins

0.4 2.0
PT

Acetylene
Acetylene
Propyne 1.6 Propyne
0.3
E

1.2
Content( %)
Content( %)

0.2
CC

0.8

0.1 0.4
A

0.0 0.0

250 300 350 400 450 500 550 250 300 350 400 450 500 550
Temperature(℃) Temperature( ℃)

YN 1 YN 2
(c)Alkynes
Fig. 5. Content of gaseous organic matters at different pyrolysis temperatures
In addition to higher concentrations of methane, ethane and ethylene, very few other gaseous alkanes, olefins
and alkynes were detected. As shown in Fig. 5, the maximum molar ratio of the small amounts of alkanes and
olefins detected at 550°C was within the range of 1.2 - 1.5 mol /mol. Comparing Figures 5(a) and 5(b), it can be
seen that more paraffin and olefins were produced during the pyrolysis of YN 1 oil sands. Figure 5(a) shows that
the molar ratio of n-pentane and iso-butane were relatively high. The isobutane of YN 1 oil sand dramatically
increased at 400 °C, after which, it continued to increase and reached the maximum value. However, the molar
ratio of YN 2 oil sand increased to 0.8 mol/mol at 300 °C, and no significant change occurred with the increase in
temperature. The outputs of iso-pentane and n-pentane were very low. Fig. 5(b) shows that, with the increase in
pyrolysis temperature, the molar ratio of trans-butene increased, whereas its relative content was also higher. The
concentration of n-butene obtained from the pyrolysis of YN 1 oil sand was greater than that of YN 2 oil sand. The

T
other olefin concentration was almost zero. Figure 5(c) shows that the pyrolysis of YN 2 oil sands began at 400 °C

IP
with a sharp increase in propyne molar ratio, and at 450 °C, the rate of increase slowed down. However, its molar
ratio was higher than that of pyrolysis YN 1 oil sands [11-13]. The molar ratio of acetylene produced at each final

R
temperature during the pyrolysis of oil sands in the two regions was extremely low.

SC
3.3 Liquid organic matter emission characteristics
The change in the yield of liquid organic matter with the pyrolysis final temperature is shown in Figure 2. It
can be seen that, as the pyrolysis temperature gradually increased, the yield of liquid organic matter first increased,

U
and then, decreased. The yield of liquid organic matter in YN 1 oil sand was higher than that in YN 2 oil sand,
indicating that the YN 1 oil sand had a higher capacity for generating organic matter.
N
Figure 6 shows the infrared spectrum of liquid organic matter at different pyrolysis temperatures for the two
oil sands.
A
100 100

80
80
M

60 250℃ 250℃
60
4000 3500 3000 2500 2000 1500 1000 500 4000 3500 3000 2500 2000 1500 1000 500
100 100

80
80
ED
透射率/%

300℃ 300℃
透射率/%

60
60
4000 3500 3000 2500 2000 1500 1000 500 4000 3500 3000 2500 2000 1500 1000 500
100
100

80
80
PT

350℃ 350℃
60
60
4000 3500 3000 2500 2000 1500 1000 500
4000 3500 3000 2500 2000 1500 1000 500 100
100

80
80
E

400℃ 400℃
60 60
4000 3500 3000 2500 2000 1500 1000 500 4000 3500 3000 2500 2000 1500 1000 500
CC

100 100

80 80

450℃ 450℃
60
60 4000 3500 3000 2500 2000 1500 1000 500
A

4000 3500 3000 2500 2000 1500 1000 500 100


透射率/%

100
透射率/%

80
80
500℃
500℃ 60
60 4000 3500 3000 2500 2000 1500 1000 500
4000 3500 3000 2500 2000 1500 1000 500 100
100

80
80

550℃ 550℃
60 60
4000 3500 3000 2500 2000 1500 1000 500 4000 3500 3000 2500 2000 1500 1000 500

波长/cm-1 波长/cm-1
(a) YN 1 (b) YN 2
Fig 6. IR spectra of liquid products obtained from different final pyrolysis temperatures

As can be seen from Figure 6, there is almost no liquid precipitate within the temperature range of 250 -
300 °C for YN 1 during the pyrolysis process. It was mainly the stretching vibration of free hydroxyl O-H within
the wavelengths of 3700 - 3600 cm-1 and the stretching vibration of the intermolecular hydrogen bonds within the
wavelength range of 3600 - 3000 cm-1, showing a broad absorption peak. In general, with the increase in final
pyrolysis temperature, the peak value and the area of absorption peak showed an increasing trend, indicating that
the type of pyrolysis liquid organic matter was similar, though its content increased [14-15]. The liquid organic
matter was dominated by aliphatic hydrocarbons. The main performance was the stretching vibration of methyl
(-CH3) and methylene (-CH2-) within the wavelength range of 3100 - 2770 cm-1. Followed by -CH3 within the

T
wavelength ranges of 1385 - 1372 cm-1 and 1472 - 1431 cm-1, -CH2 within 1480 - 1440 cm-1, and -CH at 1340
cm-1 were also observed. These were caused by the bending vibration in the saturated C-H plane. When the

IP
temperature was 350 °C, there was no absorption peak within the wavelength range of 1788 - 1550 cm-1 for the
two types of oil sands. However, when the final pyrolysis temperature increased, the intensity and area of the

R
absorption peak within the range of 1788 - 1550 cm-1 increased significantly. This was caused by the carbonyl

SC
(C=O) stretching vibrations in aldehydes, ketones, greases, acids, and the stretching vibrations of C=C double
bonds in olefins and aromatic hydrocarbons [16-17]. Furthermore, higher the final pyrolysis temperature, more were
the yields of aldehydes, ketones, acids, lipids, olefins, and aromatic hydrocarbons in the liquid organic matter. The
absorption peak intensity and the peak area of aromatic hydrocarbons in the infrared spectrum of liquid organics
increased with the increase in final pyrolysis temperature. This was mainly caused by the out-of-plane bending U
N
vibration of =CH within the wavelength ranges of 995 - 985 cm-1, 915 - 905 cm-1, and 900 - 640 cm-1, and the
out-of-plane oscillation of ≡CH within the wavelength range of 665 - 625 cm-1 [18-19]. In addition, there was a
A
weaker absorption peak of fatty ether within the wavelength range of 1108 - 980 cm-1.
The intensity and the area of the total absorption peak of liquid organic matter in the infrared spectrum of YN
M

1 oil sand were higher than those of the YN 1 oil sand, which indicates that the content of liquid organic matter of
YN 1 oil sand pyrolysis was higher than that of YN 2 oil sand. This result was consistent with those shown in Fig.
ED

5. However, the intensity and the area of aromatic peaks of YN 2 oil sands were lower than those of YN 1 oil sand,
indicating that the liquid organic matter obtained from the pyrolysis of YN 1 oil sand contained more aromatic
compounds.
3.4 Adsorption-desorption isotherm analysis
PT

Adsorption and desorption isotherms of the solid products, namely the char, showed rich pore structure
information. Figure 7 shows the adsorption and desorption curves of YN 1 and YN 2 samples and their respective
E

chars.
12 12
CC

Original sample Original sample


10 250℃ 10 250℃
300℃ 300℃
8 350℃ 8 350℃
Adsorption cm3•g-1

Adsorption cm3•g-1

400℃
A

400℃
6 450℃ 6 450℃
500℃ 500℃
4 550℃ 4 550℃

2 2

0 0

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure(P/P0) Relative pressure(P/P0)

YN 1 YN 2
Fig. 7. Adsorption-desorption curves at different pyrolysis temperatures for oil sands and char
From the physisorption point of view, the sorption isotherms of YN 1 and YN 2 in Fig. 7 are slightly different
in morphology. However, both of them showed a reversed-s-shape, which was classified as the Type II adsorption
isotherm curve. As shown in the Figure 7, the first half of the curve, that is, the part where the relative pressure
p/p0 is less than 0.2, shows an upward convex shape. As the amount of adsorption increased, the isothermal
adsorption curve showed a slowly rising trend. When the relative pressure p/p0 was within the range of 0.2 - 0.6,
the isothermal adsorption line maintained a slow and steady upward trend. At this stage, the relative pressure was
not high, and micropores was mainly filled with N2. The micropore volume was small, while the gas was adsorbed
on the micropore wall. Additionally, the monolayer adsorption saturation occurred with the increase in adsorption
amount. During this process, the intermolecular van der Waals forces played a decisive role [20-22]. After

T
condensation, nitrogen can wet the pore walls, due to which, the relative pressure of the gas and the thickness of

IP
the adsorption layer of the gas in each pore wall exhibited a positive proportional relationship. Both the gas and
the thickness of the adsorption layer of the gas increased. When p/p0 reaches a critical relative pressure

R
corresponding to a certain pore size, the condensed liquid will first fill the pores with smaller radius due to the
capillary condensation phenomenon. Furthermore, smaller the pore radius, easier it was to be filled with the

SC
coagulation liquid. In the middle stage of the curve (relative pressure p/p0 lied within the range of 0.6 - 0.75), the
adsorption curve showed a fast upward trend, mainly because the relative pressure increased, and the condensed
liquid began to fill the hole with a larger radius. The pore walls of these large and medium pores continuously

U
adsorbed and became thicker. When the relative pressure p/p0 lied within the range of 0.75 - 1.0 in the latter half of
the curve, the curve showed a large gradient. This was due to the capillary condensation in the relatively large
N
pores, which led to a sharp increase in the amount of adsorption. This also proves that during the pyrolysis of YN
1 and YN 2 oil sands, the pore structure becomes very developed. It can be shown from Fig. 7 that the char
A
contains not only the micropores, but also the mesopores and macropores.
M

Figure 7 shows that, the adsorption amounts of YN 1 and YN 2 oil sands are very low and they are almost
close to zero. This is because the pores of oil sands are filled with organic matter in oil sands before pyrolysis.
Furthermore, it is difficult to adsorb adsorbents. When the final pyrolysis temperature was varied to values of 250,
ED

300, and 350 °C, the char absorption slightly changed, although it was still close to the original adsorption
amount. The reason for this is that the lower pyrolysis final temperature was not sufficient to release the volatile
matter present in the oil sands. However, the lower pyrolysis final temperature changed the oil sand organic
PT

matter, thus resulting in a change in the structure of tiny pores. Therefore, the isothermal adsorption/desorption
curves slightly changed. The final pyrolysis temperature continued to increase, and the adsorption of char of the
two types of oil sands changed significantly. When the final temperature of pyrolysis was 450 °C, the maximum
E

adsorption capacity of the two oil sands in each relative pressure section occurred. This showed that, at the final
pyrolysis temperature, the oil sand volatiles released, forming a rich pore structure. Furthermore, factors, such as
CC

violent collisions, oscillations, and friction were not involved in the process, leading to the destruction of the pore
structure and increasing the total pore volume. When the final temperature of pyrolysis continued to increase, the
relative amount of adsorption in the relative pressure sections decreased. The reason for this phenomenon was
A

that the relatively high pyrolysis final temperature enhanced the intermolecular thermal motion, whereas the char
was affected by gravity and surface tension. In addition, the volatilization analysis caused the upper char to drop,
causing melting and collapse. Due to the interconnect effect between the hole and the aperture, the hole’s
structure changed. The diameter of the hole with different radius reduced or even closed, resulting in the
phenomenon of obturator. At a higher final temperature, although a small amount of volatilization was observed
to increase the pore volume, the rate of increase of pore volume was much smaller than the rate of decrease of
pore volume. As a result, the char adsorption slightly decreased at higher final temperatures [23-24].
As we can see from the Fig. 7, in the adsorption and desorption curves of boost and depressurization, they did
not coincide at higher values of relative pressure and formed adsorption loops. This was caused by the fact that the
relative pressure in the adsorption and desorption process may be different in the same hole. According to the
classification criteria of De Boer's adsorption loop, oil sands and char adsorption loops were close to the B-type
loops, indicating that the pores are mainly slit-shaped pores, and the pore structure is very well developed.
3.5 Specific surface area and pore specific surface area
Due to the diversity of pore structures, it is difficult to classify the types of pores to understand the changes in
morphology, such as the pore distribution and size. However, a clear pattern of change can be seen in Figure 8
(change in the total BET specific surface area and pore volume against the final pyrolysis temperature).
6 7
Specific Surface Area 0.016 0.016
Specific Surface Area

T
5 Pore Volume 6 Pore Volume

Specific Surface Area(m2•g-1)


0.014
Specific Surface Area (m2•g-1)

0.014

Pore Volume(cm3•g-1)
0.012

IP
0.012

Pore Volume (cm3•g-1)


4
4 0.010
0.010

3 0.008
0.008 3

R
0.006
0.006
2 2
0.004

SC
0.004
1 0.002
1
0.002
0 0.000
0 0.000
250 300 350 400 450 500 550 250 300 350 400 450 500 550

YN 1
Temperature(℃)

YN 2 U
Temperature(℃)
N
Fig. 8. BET specific surface area and pore volume of chars at different final temperature
Fig. 8 shows that the total BET specific surface area and the pore volume of the two oil sands are in good
A
agreement. The total BET specific surface area and the pore volume of YN 2 oil sand showed a decreasing trend
M

within the range of 250 - 350 °C. This was due to the low final temperature of pyrolysis, which was insufficient to
analyze the volatilization in oil sands. However, within this temperature range, the organic matter in the oil sand
became fluidized due to heat, resulting in obstruction of original pores due to the flow of oil sand’s organic matter.
ED

This phenomenon did not appear in the other (YN 1 oil sand) sample. Within the temperature range of 350 -
450 °C, the total BET specific surface area and the pore volume of YN 2 oil sand dramatically increased, and
reached the maximum value at 450 °C. It was due to the increase in final temperature of pyrolysis and the
PT

devolatilization of the oil sands, which left the original pores filled with oil sand organic matter. The total BET
specific surface area and the pore volume increased from 300 °C for YN 1 oil sands, after which, the rate of
increase dropped. It reached the maximum value at 350 °C with a sharp increase till 450 °C, and then within the
E

range of 450 - 550 °C, the total BET specific surface area and the pore volume of YN 2 oil sands decreased. It was
due to the increase in final temperature of pyrolysis, the collapse of pores, and the emission of organic matter,
CC

which blocked a part of the pores. The YN 1 oil sand first decreased, and then, increased [25-26].
3.6 Fractal dimension
The application of fractal theory of char can be thought as an important method for describing the
A

characteristics of the structure, and reflects the degree of pore roughness on the surface of the char surface and the
irregularity of the sample surface. The characterization parameter is called the fractal dimension ‘D’. The effective
range of fractal dimension D is 2 < D < 3, where 2 represents absolutely smooth pores. Furthermore, larger the
period dimension, higher are the degree of surface roughness and the irregularity [27-28]. In addition, the value of D
was calculated using the FHH model to understand the surface roughness of oil sand char by detecting the
adsorption of nitrogen molecules (see Eq. (1)).
 D  3
  p 
a  k ln  0   (1)
  p 
After taking the natural logarithm of the two sides of Eq. (1), Eq. (2) is obtained.

  p 
ln a   D  3 ln ln  0    C (2)
  p 

where a is the amount of nitrogen adsorption (cm3/g), p0/p is the reciprocal of the relative pressure, D is the
fractal dimension, and C is a constant term. The relative pressure and the amount of nitrogen adsorption were
measured using a specific surface area analyzer. The discrete points were obtained and the origin was used for
linear fitting. The FHH equation’s fitting diagram of the char sample at each pyrolysis final temperature was
obtained, and is shown in Fig. 9. Using ln(a) to plot against ln[ln(p0/p)], the resulting slope can be used to
determine the fractal dimension D. The fitting correlation coefficient is found to be greater than 0.9.

T
2.5 2.5

IP
250℃
250℃ 300℃
2.0 300℃ 2.0 350℃
350℃ 400℃
400℃
1.5 450℃ 1.5 450℃
500℃
500℃

R
550℃
550℃
1.0
1.0
0.5

SC
0.5
lna
lna

0.0
0.0
-0.5
-0.5
-1.0
-1.0

-1.5
-1.5
U
N
-2.0
-5 -4 -3 -2 -1 0 -5 -4 -3 -2 -1 0
lnln(p0/p) lnln(p0/p)
A
YN 1 YN 2
Fig.9. FHH equation’s fitting diagram of char at different final pyrolysis temperatures
M

Figure 10 shows the variation in gas adsorption fractal dimension with the final temperature of pyrolysis. The
effective range of fractal dimension D was 2 < D < 3, where 2 represented the absolutely smooth pores.
ED

Furthermore, larger the period dimension, higher were the degree of surface roughness and the irregularity. The
study found that oil sand pyrolysis was divided into three stages. The first stage was below 250 °C, and belonged
to the dehydration degassing stage, during which, the sample was out of thermal stability. The third stage was the
PT

basic emission of volatiles within the temperature range of 550 - 900 °C, during which, the inorganics decomposed
at high temperature. The second stage was within the temperature range of 250 - 550 °C, which was the key
research stage and the final temperature lied within the range of 250 - 300 °C. The oil sand removed a small
E

amount of organic matter, whereas the residual thermoplastic components increased, resulting in the blockade of
orifice. The particles stuck together into a highly viscous plastic state, reducing the fractal dimension and
CC

smoothing the surface. As the final temperature of pyrolysis increased, the amount of gas released increased and
the sample began to undergo irreversible plastic deformation. It shows that the emission of oil sand volatiles
caused the change in pore structure on the surface of char, and the degree of complication was strengthened, which
A

increased it to the peak of 350 °C. At around 420 °C, the volatilization analysis basically ended. The gas and oil
production rates reached the minimum values. The fractal dimension showed a minimum at the temperature of
400 °C [29]. The increase in fractal dimension within the temperature range of 400 - 450 °C may be attributed to the
removal of thermoplastics and further decomposition (secondary decomposition) of asphaltenes to produce a large
number of micropores [26]. The third stage of the reaction lied above the temperature of 450 °C and belonged to the
main temperature range of coking reaction. The thermal decomposition of inorganic salts led to a continued
increase in gas production, while the production of oil products was basically over. Fig. 10 shows the portion of
fractal dimension after 450 °C, which proves the occurrence of above-mentioned process. On the other hand,
above the temperature of 450 °C, the pore structure of the particles began to reform. Due to the effect of thermal
stress and surface tension on the hole’s surface plastic deformation, char surface became smooth, and the uneven
distribution of pore size may also have caused the reduction of fractal dimension [30].

2.80 2.80

2.75 2.75

2.70
2.70
lna

lna
2.65
2.65

T
2.60
2.60

IP
2.55
2.55
250 300 350 400 450 500 550 250 300 350 400 450 500 550
lnln(p0/p) lnln(p0/p)

R
YN 1 YN 2

SC
Fig. 10. Variation in gas adsorption fractal dimension for chars at various final temperatures of pyrolysis
3.7 SEM fractal analysis
Scanning electron microscopy (SEM) is the observation of oil sands and char surface structures in the form of

U
pictures. It is widely used to study the pore distribution. The use of scanning electron microscopy to analyze the
char of oil sands helps in understanding the structural changes in char and the evolution of volatiles. Figure 11
N
shows the scanning electron microscopy analysis of YN 2 oil sands and oil sands’ char. All the SEM tests in this
article are performed under the same conditions, and therefore, the test errors can be ignored. The scanning
A
electron micrographs of YN 1 and YN 2 showed similar characteristics, and the SEM images of YN 2 and YN 2
M

char at different pyrolysis final temperatures are shown in Figure 11.


ED
PT

(a) Original sample (b) 250℃ (c) 300℃


E
CC
A

(d) 350℃ (e) 400℃ (f) 450℃


(g) 500℃ (h) 550℃

Fig. 11. SEM images of char at different final pyrolysis temperatures

T
It can be seen from Figure 11 that the surface structure of oil sand char is irregular, the pore structure is loose,

IP
and there is no obvious connection between the particles. When the final temperature of pyrolysis was 250 °C, the
organic matter begins to show a molten state, while the surface was broken by the gas, forming an outward

R
expansion hole. The pore structure was gradually formed, and the surface structure was stepped. With the gradual
increase in the final temperature, the granularity of surface structure became more and more obvious, whereas

SC
larger, deeper and wider grooves appeared. When the final pyrolysis temperature reached 450 °C, the char surfaces
were all fine-grained and accompanied by relatively large non-precipitated oil sand organic particles. However,
when the final pyrolysis temperature reached 550 °C, large particles of the organic matter significantly reduced on
the surface of the oil sands, indicating that with the increase in pyrolysis temperature, volatile matter was
continuously released. In addition, from the surface structure of the oil sand’s char, obtained using scanning U
N
electron microscopy, the volatiles have basically released at 450 °C. The char exhibited a rough surface and the
remaining pyrolysis solid product after volatilization release should be mineral.
A
The black edge below the SEM image was removed to reduce the effect on the grayscale segmentation value.
MATLAB® was used to process the SEM gray image of oil sands’ char with the YN 2 oil sands shown in Fig. 11.
M

The final pyrolysis temperature was maintained at 450 °C. Sobel filter was used for YN 2 oil sands as well as the
edge extraction of char skeleton. The results are shown in Figures 12 and 13. Fig. 12(a) shows that there is a
ED

certain change after using the sobel filter. The structure of the pores was clear, while the distinction was obvious.
The boundary contour lines were clear, and the pore space structure had a good three-dimensional structure. In
order to accurately measure the change in pore structure, MATLAB® was used to obtain the first derivative of
PT

cumulative number of image pixels (Figure 12(b)) under different gray levels. The gray value segmentation value
corresponding to the maximum value of the first derivative was taken as the binarization threshold. The
transformation of Figure 12(a) into Figure 12(c) was then achieved [31]. Similar process was repeated to transform
E

Fig. 13(a) to Fig. 13(c).


CC

1400000

1200000
The number of pixel unit

1000000
A

800000

600000

400000

200000

0 50 100 150 200 250


Gray segmentation value

(a)After sobel processing (b)Image pixel accumulation under (c)Binary picture

different grayscale segmentation values

Fig. 12. Marginal extraction and binarization of YN 2


1400000

1200000

The number of pixel units


1000000

800000

600000

400000

200000

0
0 50 100 150 200 250
Gray segmentation value

(a)After sobel processing (b)Image pixel accumulation under (c)Binary picture

different grayscale segmentation values

T
Fig.13. Extraction and binarization of YN 2 at final temperature of 400 °C

IP
The box dimension method was used to calculate the fractal dimension of SEM scanning binarization. The
fractals are usually of an irregular or inaccurate geometry. It was assumed that the fractal object was divided into N

R
similar smaller figures, whereas each smaller figure was reduced by r times the original figure, and then the fractal
dimension D was defined using Equation (3) [32].

SC
log  Nr 
D= (3)
log 1 r 
.
According to Eq. (3), the SEM scanning electron micrographs of YN 1 and YN 2 at different pyrolysis
temperatures were used to calculate the fractal dimension D. U
N
2.46 2.52
A
2.44
2.48
M
Fractal dimension D

Fractal dimension D

2.42
2.44
2.40
ED

2.38 2.40

2.36
2.36
PT

250 300 350 400 450 500 550 250 300 350 400 450 500 550
Temperature(℃) Temperature(℃)

YN 1 YN 2
E

Fig 14. Variation in fractal dimension with final pyrolysis temperature


The comparison of results shown in Fig. 14 and Fig. 10 shows that the variation in fractal dimension of
CC

samples is basically similar. The fractal dimension of nitrogen adsorption is generally higher than the fractal
dimension of SEM image, because the nitrogen adsorption can measure the internal structural information of the
pore, while SEM can only detect the change in pore structure on the surface of char, and cannot reflect the internal
A

structure of the pore [33].

4 Conclusion
The pyrolysis products of YN 1 and YN 2 oil sand samples included oil products, gas products, and char
products. The formation of YN 1 oil sands’ organic matter was basically similar to that of YN 2 oil sands. The
yield of YN 1 oil sands’ organics was higher than that of YN 2 oil sands, and the total organic matter yield
increased with the increase in final pyrolysis temperature. For the final pyrolysis temperature of above 300 °C, the
liquid was dominant in the products. The temperature range of 300 - 500 °C was the main production stage of oil
sand organic matter. The total organic matter yield first increased, and then, decreased with the increase in
pyrolysis final temperature. The nitrogen adsorption capacity, BET specific surface area, and pore volume of the
two oil sand samples reached the maximum peak at 450 °C. FHH equation was used to fit and the results showed
that the fractal dimension also reached its peak at 450 °C. MATLAB® was used to measure the fractal dimension
of oil sand char. The change in fractal dimension D was slightly lower than the nitrogen adsorption fractal value.
However, the pattern of change was basically the same. The SEM images provided visual pore structure
information of the char surface of oil sands.

Acknowledgment

T
This research was supported by the Natural Science Foundation of China (No. 51676032); The science and technology

IP
innovation and development projects of Jilin city (20166006).

R
Acknowledgements: The authors are grateful to the financial support from the China Scholarship Council (CSC) and the Natural
Science Foundation of China (No. 51676032).

SC
References
U
[1] Haiying Li, Guijie Zhang, GAO Xiang, Study on Pyrolysis Characteristics of Non -conventional Oil Resource[J]. Chemical Engineering of Oil & Gas,
N
2010, 39(03):189-192+177.

[2] Li Li. Oil sands - a new alternative energy source[J]. Pteroleum & Petrochemical Today, 2005,13(12):28-30
A
[3] Qiang Xiu, Hongyan Wang, Dewen Zheng, et al. Research Progress in Application of the Oil Sands[J]. Liaoning Chemical Industry,2008. 37(4):268-271
M

[4] Kaishuai Zhang. Study on the pyrolysis characteristics of oil sands in Indonesia [D]. Dalian University of Technology, 2016.
[5] Yanli Shang, Zhengxin Peng, Opportunities and Risks of the Canadian Oil Sand Resource Development[J]. CNPC Research Institute of Economics and
Technology,2006,3:32-36
ED

[6] Baizhong Sun, Qing Wang, Ping Tan, et al. Thermal Fragmentation Characteristic of Oil Shale and Char in Fluidized Bed Combustion[J]. Proceedings of
the CSEE, 2010, 30(23): 62-66.

[7] Xiangxin Han, Xiumin Jiang, Zhigang Cui, et al. Study Of Combustion On Performance Of Oil Shale Char[J]. Proceedings of the CSEE, 2005, 25(15):
PT

106-110.

[8] Chunxia Jia. Research on the pyrolysis characteristics and products formation mechanism of oil sand[D]. NCEPU North China Electric Power
University, 2014.
E

[9] Granada E, Eguía P, Vilan J A, et al. FTIR quantitative analysis technique for gases. Application in a biomass thermochemical process[J]. Renew Energy,
2012,41:416–421
CC

[10] Peng Fu, Weiming Yi, Xueyaun Bai, et al. Effect of temperature on gas composition and char structural features of pyrolyzed agricultural residues[J].
:8211–8219
Bioresoure Technology,2011,102(17)

[11] Moschopedis S. E., Parkash S., Speight J. G.. Thermal decomposition of asphaltenes[J]. Fuel,1978,57(7):431-434
A

[12] Ritchie R. G. S., Roche R. S., Steedman W.. Pyrolysis of Athabasca tar sands: analysis of the condensible products from asphaltene[J]. Fuel,1979,58
(7):523-530

[13] Ritchie R. G. S., Roche R. S., Steedman W.. Non-isothermal programmed pyrolysis studies of oil sand bitumens and bitumen fractions: 1. Athabasca
asphaltene[J]. Fuel,1985,64(3):391-399

[14] Qing Wang, Yuhe Yan, Chunxia Jia, et al. FTIR analysis and pyrolysis characteristics of oil shale from Gansu province. Chemical Industry and
:1730-1735
Engineering Progress,2014,33(7)

[15] Ningbo Gao, Aimin Li, Quan Cui, et al. TG–FTIR and Py-GC/MS analysis on pyrolysis and combustion of pine sawdust[J]. Journal of Analytical and
Applied Pyrolysis 2013,100:26–32
[16] Junwei Yan, Xiuming Jiang, Xiangxin Han, et al. A TG-FTIR investigation to the catalytic effect of mineral matrix in oil shale on the pyrolysis and
combustion of kerogen[J]. Fuel,2013,104:307-317

[17] Qing Wang, Chunxia Jia, Qianqian Jiang,et al. Pyrolysis model of oil sand using thermogravimetric analysis[J]. Journal of Thermal Analysis and
Calorimetry,2014,116(7):499-509

[18] Meng Meng, Haoquan Hu, Qiumin Zhang, et al. Pyrolysis Behaviors of Tumuji Oil Sand by Thermogravimetry (TG) and in a Fixed Bed Reactor.
:2245-2249
Energy& Fuels,2007,21(4)

[19] Qing Wang, Jianxin Ge, Chunxia Jia, et al. Influence of retorting end temperature on chemical structure of oil-sand oil[J]. CIESC Journal,2013,64(11):
4216-4222

[20] Aik Chong Lua,Ting Yang. Characteristics of activated carbon prepared from pistachio-nut shell by zinc chloride activation under nitrogen and vacuum
conditions[J]. Journal of Colloid And Interface Science, 2005, 290(2).

T
[21] Jimin Yan, Qiyuan Zhang. Adsorption and Agglomeration: The Surfaces and Pores of Solid[M]. Science Press, 1979.

IP
[22] Qing Wang,Guojun Jiao, Hongpeng Liu, Jingru Bai, Shaohua Li. VARIATION OF THE PORE STRUCTURE DURING MICROWAVE PYROLYSIS OF
OIL SHALE[J]. Oil Shale, 2010, 27(2).

R
[23] Jingru Bai, Qing Wang, Guojun Jiao. Study on the Pore Structure of Oil Shale During Low-Temperature Pyrolysis[J]. Energy Procedia, 2012, 17.
[24] Han X, Jiang X, Yu L, et al. Change of pore structure of oil shale particles during combustion. Part 1. Evolution mechanism[J]. Energy & fuels, 2006,

SC
20(6): 2408-2412.

[25] Cetin E, Gupta R, Moghtaderi B. Effect of pyrolysis pressure and heating rate on radiata pine char structure and apparent gasification reactivity[J]. Fuel,
2005, 84(10): 1328-1334.

U
[26] Schrodt J T, Ocampo A. Variations in the pore structure of oil shales during retorting and combustion[J]. Fuel, 1984, 63(11): 1523-1527.
[27] Ahmad A L, Mustafa N N N. Pore surface fractal analysis of palladium-alumina ceramic membrane using Frenkel–Halsey–Hill (FHH) model[J]. Journal
N
of colloid and interface science, 2006, 301(2): 575-584.

[28] Ismail I M K, Pfeifer P. Fractal analysis and surface roughness of nonporous carbon fibers and carbon blacks[J]. Langmuir, 1994, 10(5): 1532-1538.
A
[29] Ritchie R G S, Roche R S, Steedman W. Non-isothermal programmed pyrolysis studies of oil sand bitumens and bitumen fractions: 1. Athabasca
M

asphaltene[J]. Fuel, 1985, 64(3): 391-399.

[30] Peng Fu, Song Hu, Jun Xiang, et al. Evolution of pore structure of biomass particles during pyrolysis[J]. CIESC Journal, 2009, 60(07): 1793-1799.
[31] Filgueira R R, Fournier L L, Cerisola C I, et al. Particle-size distribution in soils: A critical study of the fractal model validation[J]. Geoderma, 2006,
ED

134(3): 327-334.

[32] Bisoi A K, Mishra J. On calculation of fractal dimension of images[J]. Pattern Recognition Letters, 2001, 22(6):631-637.
[33] Qing Wang, Jianxin Ge, Chunxia Jia, et al. Influence of final retorting temperature on the surface property of oil sand char[J]. Chemical Industry and
PT

Engineering Progress, 2014, 33(04): 891-895.


E
CC
A

You might also like