Energy Comparative Study 1-S2.0-S0360544223010071-Main

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Energy 276 (2023) 127613

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Comparative study on the chemical structure characteristics of lump coal


during superheated water vapor pyrolysis and conventional pyrolysis
Chao Zhang a, b, Yangsheng Zhao a, b, Zijun Feng a, b, *, Lei Wang b, Qiaorong Meng a, b, Yang Lu b,
Qiang Gao b
a
Department of Mining Engineering, Taiyuan University of Technology, Taiyuan, 030024, China
b
Key Laboratory of In-situ Property Improving Mining of Ministry of Education, Taiyuan University of Technology, Taiyuan, 030024, China

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

Handling Editor: Wojciech Stanek Underground coal thermal treatment is an in-situ pyrolysis process that converts underground coal into synthetic
liquid and gaseous fuels using either nitrogen or superheated water vapor as the heat-and-mass-transfer medium.
Keywords: Therefore, comparing the effects of both pyrolysis methods on the coal pyrolysis characteristics is essential. The
Lump coal present study examines proprietary superheated water vapor pyrolysis and conventional pyrolysis of Tashan
Pyrolysis
lump coals. The functional groups and carbon microcrystal structures of the solid residue were detected using
Functional group
Fourier transform infrared, Raman, and X-ray diffraction spectroscopies. Volatile components were quickly
Carbon microcrystalline structures
Carbon stacking structure removed from coal in the 350◦ C-500 ◦ C temperature range and were more effectively removed by superheated
water vapor than by conventional pyrolysis. Increasing the temperature decreased the percentages of aliphatic
–O, C–O, and aromatic C–
structures, C– –C groups, increased the percentage of –OH groups, and increased and
then decreased the percentage of aromatic C–H groups. The superheated water vapor facilitated the removal of
aliphatic structures and increased the content of aromatic C–H groups in the solid residue, but did not signifi­
cantly change the evolution of oxygen-containing functional groups. Increasing the temperature of the super­
heated water vapor increased the degree of ordering of the carbon microcrystalline structure in the coal while
gradually decreasing the aromatic stacking height. At temperatures between 300 ◦ C and 500 ◦ C, the superheated
water vapor more effectively increased the aromaticity and rank of the coal than conventional pyrolysis.

1. Introduction in-situ pyrolysis of coal using superheated water vapor (which has a
higher specific heat than gaseous media) as the heat-and-mass-transfer
Since the beginning of the 21st century, social and economic medium. They injected superheated water vapor through a well into
development has increased the global energy demand by 55.04% [1]. In the coal seam. The technical scheme of this method is depicted in Fig. S1
China, the proportion of renewable energy is annually increasing but in the supporting information.
fossil fuels are expected as the main energy source over the long-term [2, Coal pyrolysis is a critical component of the UCTT process. As
3]. As oil and gas resources are scarce, coal accounts for over 65% of pointed out by many researchers [13,17–20], the coal pyrolysis reaction
China’s energy structure [4]. Although coal resources in China are is very complex and involves chemical reactions such as bond breaking,
abundant, approximately 48% of the total coal reserves are low-rank gasification, condensation, and crosslinking. Coal pyrolysis is affected
coal reserves [5], defined as coals with high ash and volatile matter by the heating rate, pressure, particle size, and reaction atmosphere. Yu
contents and low calorific value [6]. Underground coal thermal treat­ et al. [21] demonstrated that high heating rates can increase the tem­
ment (UCTT) is an in-situ pyrolysis process that converts underground perature range of the pyrolysis reaction and the rate of volatile
coals into synthetic liquid and gaseous fuels with higher H/C ratios than component removal from coal. However, a low heating rate facilitates
coals [7–9]. At present, coal is mainly heat-treated by internal com­ secondary reactions. Mathews et al. [22] investigated the pyrolysis
bustion, wall conduction, or externally generated hot gas [10,11]. characteristics of bituminous coals in the 20–250 μm range of particle
Externally generated hot-gas methods mainly employ CO2, N2, and sizes. They reported that particle size did not significantly affect the
combustion exhaust gas [8]. Alternatively, Zhao et al. [12–16] proposed removal of volatile components from coal. In contrast, Anthony et al.

* Corresponding author. Department of Mining Engineering, Taiyuan University of Technology, Taiyuan, 030024, China.
E-mail address: [email protected] (Z. Feng).

https://doi.org/10.1016/j.energy.2023.127613
Received 2 November 2022; Received in revised form 19 April 2023; Accepted 20 April 2023
Available online 21 April 2023
0360-5442/© 2023 Elsevier Ltd. All rights reserved.
C. Zhang et al. Energy 276 (2023) 127613

[23] revealed that the size of bituminous coal particles in the Fourier transform infrared (FTIR), Raman, and X-ray diffraction (XRD)
53–1000-μm range significantly influences the pyrolysis of coals. In fact, spectroscopies.
increasing the size of millimeter-scale particles increases the
heat-transfer distance to the particle center. Moreover, increasing the 2. Material and methods
particle size increases the transfer path of volatiles from the particles
and promotes secondary reactions, thereby increasing the gas yield and 2.1. Coal samples
reducing the tar yield [8]. Wall et al. [24] revealed that increasing the
pressure decreases the total yield of volatiles and tar while increasing Coal was obtained from the Tashan coal mine in the Datong coal field
the total gas yield. Jun et al. [25,26] numerically investigated the pri­ located in the northern region of Shanxi Province, China. Table 1 pre­
mary reactions of coal volatiles for soot formation mechanism under sents the proximate and ultimate analyses of raw coal. Coal was crushed
pyrolysis, O2 gasification, and O2/CO2 gasification conditions. They into lumps of approximate diameter 20–30 mm for the experimental and
found that the product concentration more significantly depends on analytical tests. The raw and pyrolyzed coal samples were measured
temperature than on pressure. Jiang et al. [27] and Wu et al. [27,28] according to the ISO 333, ISO 334, ISO 562, ISO 1171, and ISO 1172
compared the pyrolysis characteristics of low-rank coals under CH4 and standards [45] and the results are shown in Table 1.
N2 atmospheres. They reported that a CH4 atmosphere increases the tar
yield and percentage of light oil in tar. Meanwhile, Ma et al. [29] found 2.2. Pyrolysis procedure
that a CO atmosphere inhibits volatile-product removal during the rapid
pyrolysis stage of coal. Yu et al. [21,30] showed that a steam atmosphere 2.2.1. Superheated water vapor pyrolysis
reduces the molecular weight of the gas components and increases the The superheated water vapor pyrolysis experiments on coal were
H/C ratio during coal pyrolysis. performed in a self-designed superheated water vapor pyrolysis system
As residual solids are essential components in the coal pyrolysis (Fig. 1(a)). The system included a water-vapor generator, a superheater,
process, they must be studied in detail to comprehend the coal pyrolysis a reactor, a cooling system, and a temperature recorder. The superheater
mechanism. The functional-group structure and carbon-microcrystalline employed three independent heating modules for graded water vapor
structure are important aspects of the chemical structure of residual heating at 5 ◦ C/min. The reactor was 200 mm long and its internal and
solids. Researchers have observed that increasing the pyrolysis tem­ external diameters were 80 and 100 mm, respectively. Prior to the
perature gradually removes the oxygen-containing functional groups, experiment, approximately 1000 g of the crushed coal sample was
aliphatic hydrocarbon chains, and side chains, which are converted to loaded into the reactor. Approximately 5.0 L of pure water was then
oil and gas products [31–33]; specifically, the degree of condensation of added to the tank on the right side of the water vapor generator,
aromatic hydrocarbons increases and small amounts of hydrogen are launching both the water vapor generator and superheater. When the
released [34,35]. Chang et al. [36] and Valdés et al. [36,37] found that a superheater reached a furnace temperature of 150 ◦ C, the valve between
CO2 atmosphere promotes the decomposition of aliphatic groups but less the water vapor generator and superheater was opened and the super­
effectively enhances the coal aromaticity than an N2 atmosphere. Dong heater was heated stepwise to convert the saturated water vapor into
et al. [38] showed that coal pyrolysis in a CH4 atmosphere facilitates the superheated water vapor. The superheated water vapor was then
release of aliphatic groups. Yu et al. [39,40] and Li et al. [39–41] injected into the reactor for preheating. Once the preheating was com­
observed that increasing the temperature significantly decreases the plete, the superheater continued heating until the reactor temperature
total band area of the Raman spectrum, gradually decreases the reached the target temperature (200◦ C-600 ◦ C). It was then held at the
peak-area intensity ratio of the D and G bands (the ID1/IG ratio), and target temperature for approximately 120 min. The superheated water
gradually reduces the number of extensive defects and amorphous vapor was passed through the reactor and thereafter carried the oil and
structures in coal while increasing the degree of ordering of the carbon gas mixture from the pyrolysis reaction into the cooling system (shell
microcrystalline structure. Wu et al. [42] revealed that the lateral size and tube cooler). After cooling the mixtures, the cooling system
(La) of the carbon microcrystalline structure gradually increases with exhausted the gaseous products from its upper port and the mixture
pyrolysis temperature. Meanwhile, Liang et al. [43] found that the consisting of pyrolysis-derived oil and water from the cooled super­
carbon stacking height (Lc) decreases and increases with temperature in heated water vapor from its lower port. At the end of the reaction, the
the lower and higher ranges of pyrolysis temperatures, respectively. Sun water vapor generator and superheater were switched off. The coal
et al. [44] investigated the temperature-induced variations in carbon samples were removed from the reactor and weighed when the reactor
microcrystalline structure during coal pyrolysis in supercritical water had cooled to room temperature in air. Table 2 lists the quantities of
from 350 ◦ C to 750 ◦ C. They concluded that the number of defects and pure water required in the pyrolysis experiments with superheated
amorphous structures in coal first increases and then decreases with water vapor at different temperatures. The experiment was repeated
increasing temperature. The carbon stacking height gradually decreases twice under each experimental condition.
while the lateral size gradually increases.
The evolution of the functional group structure and carbon micro­ 2.2.2. Conventional pyrolysis
crystalline structure in solid coal residues after pyrolysis have been Fig. 1(b) presents the experimental system for the conventional py­
comprehensively studied under different conditions. However, previous rolysis of coal. The system comprises an electric heater, a reactor, a
studies were limited to coal powders with particle sizes below 1.0 mm. cooling system, and a temperature controller. The reactor was the same
The coal in the UCTT process usually consists of large-particle lumps size as that used for the superheated water vapor pyrolysis. Prior to the
whose size cannot be controlled. Few studies have explored the evolu­ experiment, approximately 1000 g of the crushed coal sample was
tion of functional group structures and carbon microcrystalline struc­ loaded into the reactor. After turning on the nitrogen-cylinder valve,
tures in residual solids after lump-coal pyrolysis. Moreover, studies on nitrogen was continuously introduced to the reactor at 0.6 L/min to
lump-coal pyrolysis have been performed under N2, hot air, and CO2 exhaust the air from the reactor. After 5 min of continuous nitrogen
atmospheres, but never under a superheated water vapor atmosphere. introduction, the electric heater was switched on and the temperature
The adjustable parameters in the UCTT process are the temperature and was increased at 5 ◦ C/min until the reactor reached the target temper­
atmosphere of pyrolysis. Hence, this paper investigates the effects of ature (200◦ C-600 ◦ C). The target temperature was maintained for
pyrolysis temperature (200◦ C-600 ◦ C) and pyrolysis atmosphere (N2 and approximately 120 min and the gas and liquid phase products were
superheated water vapor) on the functional-group and carbon- subsequently collected. At the end of the experiment, the electric heater
microcrystalline structures in the residual solids after highly volatile was switched off and the reactor was cooled to room temperature in air.
bituminous lump-coal pyrolysis. The analysis was performed using The nitrogen cylinder valve was closed and the coal samples were

2
C. Zhang et al. Energy 276 (2023) 127613

Table 1
Results of proximate and ultimate analyses of raw and pyrolyzed coals.
Samples Mad/% Ad/% Vd/% FCd/% Sd/% Cd/% Hd/% Nd/% Oda/%

raw coal 1.34 29.92 24.99 45.09 0.34 57.17 3.63 0.95 7.99
SWV200b 0.56 29.94 24.76 45.30 0.35 57.24 3.62 0.98 7.87
SWV250 0.42 30.18 24.45 45.37 0.30 57.40 3.60 0.96 7.56
SWV300 0.35 30.55 24.09 45.36 0.28 57.58 3.57 0.98 7.04
SWV350 0.27 30.97 23.34 45.69 0.26 57.82 3.51 0.95 6.49
SWV400 0.24 31.69 21.03 47.28 0.22 58.21 3.36 0.98 5.54
SWV450 0.26 32.48 17.45 50.07 0.20 58.81 2.97 0.99 4.55
SWV500 0.23 33.09 13.98 52.93 0.20 59.39 2.41 0.99 3.92
SWV550 0.20 32.86 12.29 54.85 0.18 60.25 2.09 0.98 3.64
SWV600 0.21 32.71 11.25 56.04 0.19 61.01 1.92 0.99 3.18
C200c 0.45 29.91 24.81 45.28 0.37 57.20 3.64 0.97 7.91
C250 0.39 30.09 24.44 45.47 0.32 57.31 3.62 0.95 7.71
C300 0.34 30.52 24.15 45.33 0.30 57.53 3.61 0.98 7.06
C350 0.28 30.99 23.76 45.25 0.25 57.78 3.54 0.93 6.51
C400 0.23 31.63 21.65 46.72 0.23 58.17 3.45 0.99 5.53
C450 0.24 32.44 18.07 49.49 0.20 58.76 3.12 0.97 4.51
C500 0.20 33.05 14.53 52.42 0.22 59.34 2.55 0.98 3.86
C550 0.22 32.93 12.60 54.47 0.20 60.17 2.21 0.99 3.50
C600 0.20 32.82 11.36 55.82 0.18 60.89 2.01 0.99 3.11

Ad: air-dried basis, d: dry basis, M: moisture, A: ash, V: volatile matter, FC: fixed carbon.
a
By difference.
b
SWV200 indicates that the sample was subjected to superheated water vapor pyrolysis at 200 ◦ C.
c
C200 indicates that the sample was subjected to conventional pyrolysis at 200 ◦ C.

Fig. 1. Schematic of the test equipment: (a) Superheated water vapor pyrolysis system; (b) conventional pyrolysis experimental system.

Table 2
Quantities of pure water required in the pyrolysis experiments with superheated water vapor at different temperatures.
Temperature (◦ C) 200 250 300 350 400 450 500 550 600

Pure water quantities (L) 20 21 22 23 25 27 29 32 36

removed from the reactor and weighed. The experiment was repeated lapped and placed in a mold. The mold was positioned and pushed onto
twice under each experimental condition. a tablet pressing machine. Next, a clear slice between 0.1- and 1.0-mm
In addition, the exergy consumption in both pyrolysis processes has thick was created from the samples. The thin sheet was placed inside
been calculated in detail, as shown in section 2 of the supporting the FITR for analysis. The scan region ranged from 4000 to 400 cm− 1.
documentation. The results show that the exergy consumption of coal The FTIR spectra were recorded by coadding 32 scans at 4.0-cm− 1
pyrolysis was approximately equal in both pyrolysis processes. resolution.

2.3. Analysis methods 2.3.2. Raman measurements


The Raman spectra of the coal samples were recorded using a Lab­
2.3.1. FTIR measurements RAM HR Evolution Raman spectrometer (HORIBA Scientific, France).
The infrared spectra of the coal samples were measured using a Each coal sample was pulverized to a particle size of no more than 75
Nicolet iS5 infrared spectrometer (Thermo Scientific, USA). Each coal μm. The excitation wavelength was 532 nm and the source power
sample was pulverized to a particle size of no more than 75 μm. A few (provided by a semiconductor) was 1.7 mW. The experiment was con­
coal-particle samples and KBr (mass ratio of coal sample: KBr = 1:100) ducted at room temperature within the frequency band 800–2000 cm− 1.
were added to a grinding bowl. The mixed sample was then evenly The data acquisition time was 35 s. Every measurement was recorded at

3
C. Zhang et al. Energy 276 (2023) 127613

three different positions to account for the heterogeneous nature of coal. 3.2. Evolution of functional groups

2.3.3. XRD measurements Coal has a complex macromolecular structure. The functional-group
The XRD profiles of the coal samples were determined by a Pan­ evolution during coal pyrolysis was effectively revealed at the molecular
alytical Empyrean X-ray diffractometer (Panalytical, The Netherlands) level by FTIR spectroscopy. Fig. 3 shows the FTIR spectra of the coal
with Cu Kα radiation (40 kV, 40 mA). Each coal sample was ground into samples obtained after pyrolysis by the two methods at various tem­
particles measuring approximately 45 μm. Subsequently, 0.1 g of the peratures. After pyrolysis at different temperatures, the coal samples
coal sample was taken and placed evenly on the slide. The slide was then exhibited similar absorption bands and characteristic absorption peaks
inserted into a loading table with a scanning speed of 5◦ /min over a but with different intensities. These results indicate that after pyrolysis,
5◦ –90◦ (2θ) interval. the types of functional groups in the coal samples were unchanged but
the contents of each functional group decreased [47]. Next, the quantity
3. Results and discussion of each functional group in the coal was measured using the zonal
approach of Ibarra et al. [19,48–51]. Following this approach, the FTIR
3.1. Solid yield evolution spectrum was split into four absorbance bands: I: 700–900 cm− 1 (aro­
matic C–H out-of-plane deformation); II: 1000–1800 cm− 1 (C–O
Panels (a) and (b) of Fig. 2 depict the changes in solid yields and stretching, aromatic C– – C ring stretching, and C– – O stretching); III:
weight-loss rates of the coal samples, respectively, after superheated 2800–3000 cm (aliphatic C–H stretching); and IV: 3000–3750 cm− 1
− 1

water vapor pyrolysis and conventional pyrolysis at different tempera­ (aromatic C–H and –OH stretching). The experimentally obtained FTIR
tures. At 250 ◦ C, the solid yield of the coal samples in both pyrolysis spectra were curve-fitted to the four regions using PeakFitV4.12 soft­
methods was 98.1% and the weight loss rate was 0.01%/◦ C. In ware. Fig. 4 shows the fitted curves of the four regions in the FTIR
conjunction with the data in Tables 1 and it can be observed that the spectra of raw coal. The fitted curves in all regions after both pyrolysis
volatile contents of the coal were not significantly decreased but the methods were similar to those in the FTIR spectrum of raw coal. The
moisture was almost completely removed. Hence, at this stage, the solid- correlation coefficients R2 for all samples exceeded 0.99 and all residuals
yield reduction was mainly attributable to the desorption of the adsor­ were within ±5%. Tables S7–S10 in the Supporting Information present
bed water and gases from coal [46]. At temperatures above 300 ◦ C, the the assignments and integral areas of the functional groups. However,
solid yield of the coal sample clearly decreased but the weight-loss rate the absorption intensity of each peak in the FTIR spectrum was sensitive
first increased and then decreased. The weight-loss rate increased in the to the quantity of samples used in the pelleting process [52]. To more
350◦ C-500 ◦ C range, indicating that the coal pyrolysis reaction was more accurately and quantitatively compare the evolutions of the functional
dramatic at these temperatures and that volatile components were groups during the two pyrolysis methods, the integrated area of each
rapidly removed. This finding is generally consistent with the results in functional group was calculated as a percentage using Eq. (1):
Table 1. During the rapid pyrolysis stages of the coal (350◦ C-500 ◦ C), the
Am, T
superheated water vapor pyrolysis gave lower solid-product yields than AP, m, T = × 100%. (1)
Atotal, T
conventional pyrolysis. The temperature at which the weight-loss rate
began increasing was also significantly lower in the former than in the Here, AP, m, T and Am, T are the percentage and integrated areas,
latter. As evidenced in Table 1, the volatile content of coal from super­ respectively, of the absorption peak of functional group m at tempera­
heated water vapor pyrolysis was significantly lower than that from ture T, and Atotal, T is the integrated area of the absorption peaks of all
conventional pyrolysis at this stage, indicating that superheated water functional groups in the four absorption bands at temperature T.
vapor pyrolysis more effectively removes volatile components than
conventional pyrolysis. This result is probably related to two phenom­ 3.2.1. Evolution of aliphatic structures
ena. First, whereas conventional pyrolysis is a heat-conductive process, The coal macromolecule includes three aliphatic structures: methyl,
superheated water vapor pyrolysis is a convective heat-transport process methylene, and methylidyne. Fig. 4(c) shows the fitted curves in the
that more effectively transfers heat to the coal lumps coals, enabling absorption band of aliphatic stretching vibrations in pyrolyzed coal. Six
faster reaching of the pyrolysis temperature inside the coal lumps. curves are plotted with peaks at 2830, 2854, 2875, 2903, 2925, and
Second, the superheated water vapor atmosphere may facilitate the 2960 cm− 1. Fig. 5 depicts the temperature variations of the percentage
conversion of part of the functional group structure in the coal to oil and contents of aliphatic structures in the coal after both pyrolysis methods.
gas products (see subsection 3.2 for details of the reaction mechanism).

Fig. 2. Solid yields (a) and weight-loss rates (b) of the coal samples after superheated water vapor pyrolysis and conventional pyrolysis at different pyrolysis
temperatures.

4
C. Zhang et al. Energy 276 (2023) 127613

Fig. 3. Fourier transform infrared (FTIR) spectra of coals after superheated water vapor pyrolysis (a) and conventional pyrolysis (b).

Fig. 4. Fitted curves to the FTIR spectra of raw coal. (a) Aromatic C–H out-of-plane deformation vibrations. (b) Oxygen-containing functional groups and aromatic
C–
–C. (c) Aliphatic structures stretching vibrations. (d) Hydroxy and aromatic C–H stretching vibrations.

The aliphatic-structure contents in both pyrolysis approaches increased free radicals in coal at temperatures as low as 200 ◦ C. Yu et al. [21]
at 200 ◦ C. At this temperature, small aliphatic molecules and moistures showed that both water vapor and H2 can participate in various
adsorbed in the macropores and micropores of the coal were volatilized coal-pyrolysis reactions by providing H⋅ free radicals. As the H⋅ free
[53] but the removal of aliphatic small molecules was surpassed by the radical content is higher under a superheated water-vapor atmosphere
removal of hydroxyl groups in the moisture evaporated from coal. When than under an N2 atmosphere, the superheated water-vapor pyrolysis
the temperature was further increased to 250 ◦ C, the increase in facilitates the coupling reaction between CH2⋅ and H⋅ and increases the
aliphatic structure content of superheated water vapor pyrolysis was number of aliphatic functional groups. When the temperature reached
almost twice that of conventional pyrolysis. Seehra et al. [54,55] 350 ◦ C, the aliphatic-structure contents were slightly reduced in both
revealed that the breakage of weak bonds in coal forms stable free pyrolysis methods, but the effect was less severe and began at a higher
radicals between 200 ◦ C and 300 ◦ C. Furthermore, Niu et al. [19] found temperature in superheated water vapor pyrolysis than in conventional
that during the initial stages of coal pyrolysis, CH2⋅ in coal couples with pyrolysis. Therefore, the coal maintained its methylene-to-methyl con­
H⋅ to form CH3⋅ free radicals. Hence, the aliphatic-structure contents at version reactions after superheated water vapor pyrolysis at this tem­
this stage are probably increased via conversion of methylene in coal to perature. At this stage, the aliphatic-structure decline was mainly caused
methyl by binding to H⋅. Niu et al. [56] demonstrated that during by breakage of oxygen-containing bridge bonds connecting the struc­
non-catalyzed coal pyrolysis, gaseous hydrogen can react with stable tural units of coal with the surrounding alkyl side chains, which are

5
C. Zhang et al. Energy 276 (2023) 127613

band of C– – O was fitted to three peaks at 1655, 1699, and 1725 cm− 1.
Fig. 6(a) depicts the temperature-dependent changes in the percentage
–OH contents of coal after different pyrolysis processes. The hydroxyl-
group content slightly decreased after pyrolysis at 200 ◦ C, mainly
because the moisture and small molecules containing hydroxyl groups
are removed from coal at this temperature [19,58]. At 350 ◦ C, the hy­
droxyl contents after superheated water vapor pyrolysis and conven­
tional pyrolysis increased to 35.1% and 32.3%, respectively. This
increase was mainly attributed to thermal decomposition of the C–O and
C–– O groups with low thermal stability in coal [59]. When the tem­
perature was further increased to 450 ◦ C, the hydroxyl group contents of
both pyrolysis processes slightly decreased. According to Zhou et al.
[60], 450 ◦ C is the phenolic release temperature, indicating that the
hydroxyl-group content was decreased by removal of phenolic groups
from the coal. At the temperature increased from 450 ◦ C to the final
temperature of 600 ◦ C, the hydroxyl contents of both pyrolysis methods
increased by approximately 13%. During this stage, the
hydroxyl-content increase was mainly caused by pyrolytic removal of
the C–O and aliphatic groups from coal. In conclusion, hydroxyl groups
are pyrolyzed almost identically by the superheated water vapor and
conventional pyrolysis methods.
Fig. 5. Temperature-dependent changes in the percentage contents of aliphatic
structures in coal after superheated water vapor and conventional pyroly­ Fig. 6(b) plots the temperature-dependent changes in the percentage
sis processes. contents of C–O groups in coal after both pyrolysis methods. The content
of C–O groups slightly increased when the pyrolysis temperature
reached 200 ◦ C, mainly because moisture was evaporated from coal.
thermally unstable [50].
When the temperature increased from 200 ◦ C to 350 ◦ C, the content of
Although the aliphatic structure contents from room temperature
C–O groups decreased from 19.3% to 16.2% in superheated water vapor
(RT) to 350 ◦ C in the 5.3%–6.4% range differed between the pyrolysis
pyrolysis and from 19.6% to 18.0% in conventional pyrolysis. Shi et al.
methods, the changes were within 25% those of the overall pyrolysis
[60] revealed that the C–O bond energy in coal ranges from 220 to 330
processes. In other words, the aliphatic structure was relatively stable at
kJ/mol and that C–O bonds should not break at temperatures below
low temperatures but its content rapidly decreased from approximately
350 ◦ C. Therefore, the reduced C–O group contents at this temperature
5.5% to approximately 1.5% as the temperature increased from 350 ◦ C
cannot be attributed to thermal self-decomposition. The main
to 500 ◦ C. This rapid removal of aliphatic content is mainly attributable
hydrogen-bond interactions in coal are π–π interactions between the
to breaking of the aliphatic bridge bonds and C–C bonds in the long-
aromatic rings in the coal structure [61]. Because highly volatile bitu­
chain aliphatic hydrocarbons [21,57]. However, in superheated water
minous coals have abundant oxygen-containing groups and low
vapor pyrolysis, the initial and final temperatures of the rapid drop in
aromatization, their hydrogen bonds are much stronger than π–π in­
aliphatic-structure content were lower than in conventional pyrolysis
teractions. Weak interactions such as hydrogen bonding have low
and the decrease was slightly higher in superheated water vapor py­
thermal stability and are easily disrupted at low temperatures. Niu et al.
rolysis. Moreover, superheated water vapor pyrolysis obtained a higher
[19,62] found that the C–O groups in coal are partially linked to the
hydrocarbon proportion of the gas components than conventional py­
macromolecular network of coal via hydrogen bonding. Therefore, the
rolysis in the 400◦ C-500 ◦ C temperature range (see Table S6 in the
C–O groups at this temperature are mainly removed via hydrogen-bond
Supporting Information). Previous studies on the hydrogenation pyrol­
breakage. The content of C–O groups in coal significantly increased as
ysis of coal [56] have shown that H2 can be added to the products by
the temperature rose to 500 ◦ C. At this temperature, the C–O-group
hydrogen abstraction reactions; furthermore, smaller radicals have
content was negatively correlated with the aliphatic-group content in
stronger hydrogen abstraction capacity than larger radicals. As
both pyrolysis approaches, although the C–O and aliphatic group con­
mentioned above, water vapor can also provide H⋅, which explains the
tents were more variable in superheated water vapor pyrolysis than in
higher H⋅ content in the reaction atmosphere of superheated water
conventional pyrolysis, indicating that the C–O-group content increased
vapor pyrolysis than of conventional pyrolysis. Aliphatic radicals with
via rapid removal of aliphatic groups. At the final pyrolysis temperature
low molecular weight, such as methyl and methylene, also facilitate
(600 ◦ C), the contents of the C–O groups decreased to 15.5% and 16.9%
hydrogen acquisition for conversion to various hydrocarbon gases.
in superheated water vapor pyrolysis and conventional pyrolysis,
Combining these results, it is concluded that superheated water vapor
respectively, but the same proportional decrease (3.2%) was observed in
pyrolysis at 400◦ C-500 ◦ C facilitates the removal of aliphatic compounds
both pyrolysis approaches. At this stage, the C–O groups were rapidly
from coal. This conclusion is generally consistent with the critical tem­
decomposed to CO2, CO, and H2O under thermal action [63]. Both
perature range in which the UCTT process obtains light aliphatic gases.
methods exhibited the same trends and temperature turning points
The aliphatic-structure content remained almost stable in the
during the whole pyrolysis process, although superheated water vapor
500◦ C-600 ◦ C range, suggesting that the removable aliphatic structures
pyrolysis obtained slightly lower content of C–O groups than conven­
in the coal were almost completely pyrolyzed at these temperatures.
tional pyrolysis.
Fig. 6(c) presents the temperature-dependent changes in the per­
3.2.2. Evolution of the oxygen-containing functional groups
centage contents of C– – O groups in the coal after both pyrolysis pro­
Coal includes three oxygen-containing functional groups: OH, C–O,
cesses. At 200 ◦ C, the C– – O-group contents slightly increased because
and C– – O. The –OH, C–O, and C– – O stretching vibrations manifested as
moisture was removed from coal. From 200 to 350 ◦ C, the content of
bands in the 3100–3750, 1100–1300, and 1650–1750 cm− 1 regions of
C–– O groups rapidly decreased from 2.6% to 1.6% in superheated water
the FTIR spectra of coal, respectively. The absorption band of –OH was
vapor pyrolysis and from 2.7% to 1.1% in conventional pyrolysis. The
fitted to seven peaks positioned at 3186, 3291, 3388, 3459, 3531, 3607,
C–– O groups were mainly eliminated through pyrolytic removal of
and 3686 cm− 1 (Fig. 4(d)), the absorption band of C–O was fitted to five
compounds containing C– – O groups. Feng et al. [64], who studied the
peaks at 1113, 1157, 1191, 1230, and 1271 cm− 1 and the absorption
conversion of organic oxygen in coal, revealed that the thermal stability

6
C. Zhang et al. Energy 276 (2023) 127613

Fig. 6. Temperature-dependent changes in the percentage contents of oxygen-containing functional groups in coal after the superheated water vapor and con­
ventional pyrolysis processes.

of the C–– O-containing compounds decreases in the following order: – O contents [57]. The decline rate of the C–
for the decline in C– –O
quinones > carbonyl groups > carboxyl groups. Moreover, the intra­ groups slowed after rapid removal of the aliphatic structures. After
molecular carboxylic anhydride in coal undergoes the decarboxylation dropping to a certain level, the content of the C–– O groups in coal sta­
reaction above 250 ◦ C. Hence, the decreased content of C–– O groups at bilized in both pyrolysis approaches. Although the temperature and
this stage was mainly attributable to removal of the carboxylic acids content at which the C–
– O groups stabilized were higher in superheated
from coal in the form of small-molecule gases (e.g., CO2) under thermal water vapor pyrolysis than in conventional pyrolysis, the percentage
action [63,65]. As the temperature continued to increase, the contents of the C–– O groups after stabilization differed by less than
C–– O-group content maintained a decreasing trend but at a significantly 0.35% in both pyrolysis processes. That is, the two pyrolysis approaches
slower rate. During this stage, quinone removal was mainly responsible showed no significant differences in their removal of C–– O groups from

Fig. 7. Temperature-dependent changes in the percentages of aromatic structures in coal after the superheated water vapor and conventional pyrolysis processes. (a)
Aromatic C––C. (b) Aromatic C–H out-of-plane deformations.

7
C. Zhang et al. Energy 276 (2023) 127613

coal. water vapor pyrolysis than in conventional pyrolysis. Previously [54], it


was demonstrated that low levels of stable free radicals form via
3.2.3. Evolution of the aromatic groups weak-bond breakage at temperatures below 300 ◦ C. Consequently, the
The main aromatic structures in coal are aromatic C– – C and C–H removal of C–H-containing coal molecules and fragments can explain
groups. In the FTIR spectrum, the absorption bands of aromatic C– –C the decrease in C–H-group contents at temperatures below 300 ◦ C.
stretching vibrations, aromatic C–H out-of-plane deformation vibra­ Above 300 ◦ C, many free radicals are formed via breakage of the C–C or
tions, and aromatic C–H stretching vibrations appeared at 1490–1650, C–O bonds linking aromatic and aliphatic branched chains or
700–900, and 3000–3100 cm− 1, respectively. The absorption band of oxygen-containing functional groups in coal [19]. Therefore, the con­
aromatic C– – C was fitted to three peaks at 1494, 1563, and 1606 cm− 1 tents of aromatic C–H-group contents increase between 300 ◦ C and
(Fig. 4(b)). The absorption band of aromatic C–H out-of-plane defor­ 500 ◦ C. This trend is mainly influenced by the aromatic substitution
mation vibrations was fitted to eight peaks at 729, 746, 762, 790, 812, reaction of stable free radicals with H⋅ free radicals. At the final tem­
836, 860, and 876 cm− 1 (Fig. 4(a)). Finally, the absorption band of ar­ perature of 600 ◦ C, the contents of the aromatic C–H groups decreased in
omatic C–H stretching vibrations was fitted to a single peak at 3050 both pyrolysis approaches via condensation reactions of the aromatic
cm− 1 (Fig. 4(d)). Fig. 7(a) illustrates the temperature-dependent rings in coal.
changes in the percentage contents of aromatic C– – C groups in coal
after both pyrolysis approaches. The content of the aromatic C– –C 3.3. Evolution of the carbon microcrystalline structures
groups slightly increased when the pyrolysis temperature reached
200 ◦ C. This increase was related to desorption of the adsorbed moisture Raman spectroscopy is widely used to characterize the crystalline
from coal. When the temperature rose to 300 ◦ C, the aromatic C– –C structures of various carbonaceous materials, including coal [70–72].
content showed slightly fluctuating trends in both pyrolysis approaches, Panels (a) and (b) of Fig. 8 display the baseline-corrected Raman spectra
indicating that the aromatic C– – C groups in coal were essentially stable. (800–2000 cm− 1) of the coals after superheated water vapor pyrolysis
Between 300 C and 500 C, the aromatic C–
◦ ◦ – C group content almost and conventional pyrolysis, respectively. All spectral curves present two
linearly decreased in conventional pyrolysis but remained stable and widely and strongly overlapping peaks: one positioned at 1350 cm− 1 (D
then rapidly decreased in superheated water vapor pyrolysis. During this band), the other located at 1595 cm− 1 (G band). As the D and G bands
stage, the aromatic C– – C groups were removed via breakage of the overlap, analysis of either band alone would provide incomplete infor­
oxygen-containing and aliphatic bridging bonds between the coal mol­ mation on the characteristics of the obtained carbonaceous material,
ecules and fragments containing those fragments [19,53]. Why the ar­ which is highly disordered [44,73,74]. Hence, to understand the hidden
omatic C– – C-group contents trend differently in the two pyrolysis carbon-microcrystalline structure in the overlapping regions, the
approaches must be elucidated in further research. As the temperature experimentally obtained Raman spectra were decomposed into five
further increased from 500 ◦ C to 600 ◦ C, the content of the aromatic Raman bands (i.e., D4, D1, D3, G, and D2) using the Gaussian/Lorentz
C–– C groups decreased from approximately 10.9% to approximately function in PeakFitV4.12 software (the method is described in Liu et al.
8.4% in both pyrolysis approaches, likely due to aromatic ring [39,44,60,73,75]). Many parameters related to the carbon microcrys­
condensation in coal. talline structure (e.g., peak position, peak area, and full width at
Fig. 7(b) displays the temperature variations in the percentage con­ half-maximum (FWHM)) were obtained by curve fitting. Table 3 pre­
tents of aromatic C–H out-of-plane deformation vibrations in coal after sents the peak positions and assignments of the different Raman bands
the two pyrolysis approaches. The absorption band of aromatic C–H and Fig. 9 shows the fitted curves of the Raman spectrum of raw coal.
stretching vibrations was close to that of hydroxyl stretching vibrations The fitted curves of the Raman spectra of coal after both pyrolysis pro­
but showed a narrower peak width. Therefore, the absorption peak of cesses were similar to those of raw coal. The correlation coefficients R2
aromatic C–H stretching was largely affected by hydroxyl stretching for all samples exceeded 0.99 and all residuals were within ±5%.
vibrations. As previously reported [66,67], the large hydrogen loss from
coal during pyrolysis more severely limits the absorption band intensity 3.3.1. Evolution of the area ratio of the major bands
of the aromatic C–H stretching vibrations than of the aromatic C–H The degree of order or disorder of a carbon microcrystal structure
out-of-plane deformation vibrations. Hence, to more accurately analyze can be derived from the area ratios of the D1 band to the G band (ID1/IG)
the evolution of the aromatic C–H groups, the following discussion fo­ and of the G band to all bands (IG/IAll) [39,74,81]. The ID1/IG ratio is
cuses on the variation in the aromatic C–H out-of-plane deformation inversely proportional to the lateral size of the crystalline structure;
vibrations. When the temperature increased from 200 ◦ C to 300 ◦ C, the indicating that the ID1/IG ratio increases with decreased ordering of the
aromatic C–H-group contents tended to decrease in both pyrolysis pro­ carbon microcrystalline structure [75,79]. Conversely, an increase in
cesses; however, the decrease was lower in superheated water vapor IG/IAll reflects an increase in the ordering degree of the carbon micro­
pyrolysis than in conventional pyrolysis. As the temperature increased crystalline structure because the G band is the vibrational band of the
from 350 ◦ C to 500 ◦ C, the aromatic C–H group contents increased from aromatic ring. Panels (a) and (b) of Fig. 10 present the ID1/IG and IG/IAll
2.7% to 4.1% in superheated water vapor pyrolysis and from 2.8% to evolutions, respectively, during both pyrolysis processes. As the tem­
3.9% in conventional pyrolysis. Mathews et al. [68] demonstrated that perature increased to 300 ◦ C, the ID1/IG in both pyrolysis approaches
the hydrogen and ether bonds inter-linking the coal molecules or frag­ displayed a slightly decreasing trend while the IG/IAll displayed a gently
ments are cleaved at low temperatures, with subsequent removal of the increasing trend. The temperature variations in the ID1/IG and IG/IAll
coal molecules and fragments containing the aromatic C–H groups. ratios were similar in the two pyrolysis approaches. According to these
Furthermore, as mentioned in subsection 3.2.1, H2 reacts with the stable results, the degree of ordering of the carbon microcrystalline structure
free radicals in coal at temperatures above 200 ◦ C. Water vapor and H2 slightly increases in this temperature range and both pyrolysis ap­
undergo similar reactions in pyrolysis. Wilson et al. [69] found that the proaches elicit the same effects. As the macromolecular structure of the
non-volatile components of coal react with H2 primarily through coal remained stable at pyrolysis temperatures below 300 ◦ C, the
hydrogen–hydrogen exchange reactions and aromatic substitution re­ increased ordering degree of the carbon microcrystalline structure was
actions. The contents of aromatic C–H groups in coal are not changed by likely associated with the desorption of small molecular compounds
the hydrogen–hydrogen exchange reactions but are increased by aro­ from the coal. When the temperature increased from 300 to 350 ◦ C, the
matic substitution reaction of the stable free radicals in coal with H⋅ free ID1/IG ratio increased from 1.58 to 1.75 in superheated water vapor
radicals. During superheated water vapor pyrolysis, aromatic substitu­ pyrolysis and from 1.59 to 1.82 in conventional pyrolysis. In addition,
tion reactions are facilitated by the high content of H⋅ free radicals, the IG/IAll ratio trended oppositely to ID1/IG. During this stage, the
which explains the higher aromatic C–H group content in superheated side-chain groups in the macromolecular structure began to be gradually

8
C. Zhang et al. Energy 276 (2023) 127613

Fig. 8. Raman spectra of coal after superheated water vapor pyrolysis (a) and conventional pyrolysis (b).

significant increases in IG/IAll, revealing a significant increase in the


Table 3
lateral size and hence the ordering of the crystalline structure. However,
Band assignments derived from the Raman spectra.
the reaction mechanism differed in different temperature stages. Be­
Band Position Assignments tween 350 and 500 ◦ C, the increased degree of ordering of the carbon
(cm− 1)
microcrystalline structure was mainly attributed to the decreased
D4 ~1250 Cross-linked structure between the aromatic rings such as quantity of substituted structures (e.g., aliphatic structures, C–H, C–O,
Caromatic–Calkyl, aryl–alkyl ethers, C–C, and C–H on the
and C– – O groups in the aromatic ring) [74]. The superheated water
hydroaromatic rings [44,76]
D1 ~1385 Edge or other defects (e.g., edge carbon atoms or
vapor atmosphere not only contained many H⋅radicals that facilitated
heteroatoms) caused by vibrations of the isolated sp2 hybrid the substitution reaction, but also reduced the deposition of small mo­
bonds; C–C between the aromatics (≥ six rings) [44,77,78] lecular compounds on the carbon surface. Consequently, the ID1/IG and
D3 ~1535 Sp2–sp3 hybrid amorphous carbon structure and aromatics IG/IAll ratios were lower and higher, respectively, after superheated
with three–five rings [73,77]
water vapor pyrolysis than after conventional pyrolysis. Between 500 ◦ C
G ~1595 E2g vibration of graphite and aromatic rings [79,80]
D2 ~1675 Lattice vibration analogous to the E2g symmetry of the G and 600 ◦ C, the increased degree of ordering of the carbon microcrys­
band, but involving graphene layers [72,76] talline structure was mainly attributable to condensation of the aromatic
rings and conversion of the amorphous carbon structure into a large
aromatic structure containing ordered crystalline sp2 carbon atoms [44,
77,84]. Although the degrees of ordering of the carbon microcrystal
structures were almost identical after both pyrolysis methods at 600 ◦ C,
superheated water vapor pyrolysis achieved a higher ordering degree
throughout the reaction process than conventional pyrolysis. Therefore,
superheated water vapor pyrolysis can more effectively improve the
degree of ordering of the carbon microcrystal structures than conven­
tional pyrolysis at temperatures below 600 ◦ C.

3.3.2. Evolution of the position and full-width-at-half-maximum of the


major bands
The degree of ordering of carbon microcrystal structures can be
derived not only from the area ratio of the D1 and G bands, but also from
the peak position and full-width-at-half maximum (FWHM) of the D1
and G bands [85,86]. Panels (a) and (b) of Fig. 11 plot the
temperature-dependent changes in the peak positions of the D1 and G
bands, respectively, during both pyrolysis methods. As the pyrolysis
temperature increased from RT and 600 ◦ C, the peak position of the D1
band gradually decreased to lower wavenumbers (shifting by approxi­
Fig. 9. Curves fitted to the Raman spectrum of raw coal. mately 25 cm− 1), whereas that of the G band fluctuated between 1594
and 1596 cm− 1. The leftward shift in peak position of the D1 band
released as gas. However, some small molecular compounds were during pyrolysis was caused by the reduced bandwidth of the amor­
deposited on the carbon surface, resulting in defects and amorphous phous structure [73]. Meanwhile, the peak positions of the D1 bands
structures that reduced the degree of ordering of the carbon micro­ were significantly lower after superheated water vapor pyrolysis than
crystalline structure [39,82,83]. The superheated water vapor facili­ after conventional pyrolysis in the 350◦ C-500 ◦ C temperature range.
tated mass transfer and hence the removal of small molecular Superheated water vapor pyrolysis improved the removal of volatile
compounds, reducing their deposition on the carbon surface. For these products and reduced their adsorption/deposition to/from the carbon
reasons, the carbon microcrystalline structure was slightly more ordered surface. Moreover, the peak positions of the D1 bands at the initial
after superheated water vapor pyrolysis than after conventional pyrol­ (RT–300 ◦ C) and final (600 ◦ C) temperatures did not significantly differ
ysis at 350 ◦ C. between the two pyrolysis approaches, consistent with the area ratios of
When the temperature increased from 350 ◦ C to 600 ◦ C, both py­ the bands.
rolysis approaches exhibited significant decreases in ID1/IG and Panels (a) and (b) of Fig. 12 depict the temperature-dependent

9
C. Zhang et al. Energy 276 (2023) 127613

Fig. 10. Temperature-dependent changes in ID1/IG (a) and IG/IAll (b) during superheated water vapor pyrolysis and conventional pyrolysis.

Fig. 11. Temperature-dependent changes in peak positions of the major bands during superheated water vapor pyrolysis and conventional pyrolysis.

Fig. 12. Temperature-dependent change in full-width-at-half-maximum of the major bands during superheated water vapor pyrolysis and conventional pyrolysis.

10
C. Zhang et al. Energy 276 (2023) 127613

FWHM changes of the D1 and G bands, respectively, during both py­


rolysis processes. In both pyrolysis processes, the FWHM of the D1 and G
bands decreased from 207 to 83 cm− 1, respectively, in raw coal to
approximately 167 and 79 cm− 1, respectively, in coal pyrolyzed at
600 ◦ C. The D1 and G band widths are functions of the lateral crystal size
and ordering degree of the carbon microcrystalline structure [87]. The
observed decreasing trend demonstrated that the carbon microcrystal­
line structure became increasingly ordered with increasing temperature.
The FWHM variation of the D1 band far exceeded that of the G band
during the whole pyrolysis process, probably because the D1 and G
bands exhibit different FWHM sensitivities to ordering degree of the
carbon microcrystal structure [74]. The FWHM values of the D1 bands
were consistently lower in superheated water vapor pyrolysis than in
conventional pyrolysis at temperatures between 450 ◦ C and 550 ◦ C. This
result shows that superheated water vapor pyrolysis enhanced the
ordering of the carbon microcrystal structures in coal.

3.4. Evolution of the carbon stacking structure

Coal is a complex mineral composed of various organic and inorganic Fig. 14. XRD spectra of raw coal fitted with two Gaussian peaks in the
materials. The changes in inorganic minerals and carbon microcrystal­ 16◦ –30◦ range.
line structures in coal are commonly observed using XRD [88–91].
Panels (a) and (b) of Fig. 13 display the XRD patterns of coal after su­ follows [44, 89, 91,97]:
perheated water vapor pyrolysis and conventional pyrolysis, respec­
λ
tively. Inorganic mineral components (mainly kaolinite and boehmite d002 = , (2)
with percentage contents of 30 and 70%, respectively) manifest as sharp 2 sin θ002
characteristic peaks in the XRD patterns [50,71]. The heights and 0.89λ
quantities of these peaks decreased with increasing temperature. The Lc = , (3)
β002 cos θ002
profiles of all samples showed high background intensities, indicating a
sizeable proportion of highly disordered substances in the coal samples A002
[90,92]. Coal also contains a number of aromatic microcrystalline fa = , (4)
A002 + Aγ
structures similar to that of graphite. The XRD spectra display a distinct
002-band positioned between 24◦ and 26◦ . A γ-band appears between I002
R= . (5)
20◦ and 22◦ at the left part of the 002-band due to band asymmetry. The Iγ
002-band is caused by the layer spacing of the aromatic ring micro­
crystalline structure [51,93,94] whereas the γ-band is attributable to a here, λ is the X-ray wavelength (λ = 0.15406 nm), and θ002 and β002 are
saturated structure (e.g., aliphatic side chains) connected to the the position and FWHM of the 002-band, respectively. Note that the
boundary of the carbon microcrystalline structure [44,71,95,96]. The aromaticity (fa) and coal rank (R) are functions of the areas (A002 and Aγ)
lack of a distinct 100-band at 42◦ –50◦ in Fig. 13 can probably be and peak intensities (I002 and Iγ), respectively, of the 002- and γ-bands.
explained by the high background level, suggesting a low growth level of Fig. 15 presents the temperature evolutions of the relevant XRD
the graphite structure on the basal surface in coal [90]. To quantitatively profile parameters during superheated water vapor pyrolysis and con­
analyze the evolution of the carbon stacking structures in pyrolyzed ventional pyrolysis. As the temperature increased, the d002 in both py­
coal, the XRD profiles in the 16◦ –30◦ range were curve-fitted with Origin rolysis approaches fluctuated between 0.352 and 0.355 nm (Fig. 15(a)),
software. The positions, FWHMs, intensities, and areas of the Gaussian indicating that the aromatic layer spacing did not significantly change
fitted peaks in the γ- and 002-bands were obtained from the fitted curves during the pyrolysis process. Moreover, the d002 was much higher in coal
of raw coal (Fig. 14). The stacking height (Lc), layer spacing (d002), than in pure graphite (0.337 nm), indicating low ordering of the carbon
aromaticity (fa), and coal rank (R) were respectively calculated as microcrystalline structure and no dense graphite layer in coal [51,96].

Fig. 13. XRD patterns of coal after superheated water vapor pyrolysis (a) and conventional pyrolysis (b) at different temperatures.

11
C. Zhang et al. Energy 276 (2023) 127613

Fig. 15. Temperature-dependent evolutions of the relevant XRD profile parameters during superheated water vapor pyrolysis and conventional pyrolysis: layer
spacing d002 (a), stacking height Lc (b), aromaticity fa (c), and coal rank R (d).

At low temperatures, the Lc values in the two pyrolysis approaches pyrolysis, indicating that superheated water vapor pyrolysis facilitates
fluctuated around 2.07 nm with no obvious changes. This behavior was the aliphatic structure removal from coals as previously concluded in
primarily associated with the stable macromolecular structure of coal at subsection 3.2.1. Between 500 ◦ C and 600 ◦ C, the aromaticity and coal
low temperatures [19]. At higher temperatures, the Lc values showed a rank in both pyrolysis approaches slightly decreased, possibly because
significantly decreasing trend reflecting the reduction in carbon stacking much of the aromatic hydrogen was lost during aromatic structure
height. Below 500 ◦ C, the decrease in Lc was mainly attributable to sharp condensation [67].
depolymerization of the macromolecular structure of coal and the
removal of saturated structures such as aliphatic side chains [44,96]. 4. Conclusions
Between 500 ◦ C and 600 ◦ C, the decrease in Lc was probably caused by
hydrogenated ring dehydrogenation, heterocycle breaking, formation of The pyrolysis behaviors of lump coals heated by superheated water
more aromatic structures, and condensation and extension of the aro­ vapor and electrical equipment were investigated in specialized super­
matic structures that can produce polycyclic aromatic hydrocarbons heated water vapor and conventional pyrolysis systems, respectively,
[44,97]. Superheated water vapor pyrolysis obtained lower Lc values which were designed for this purpose. The effects of pyrolysis temper­
than conventional pyrolysis in the temperature range 400◦ C-600 ◦ C; ature and atmosphere on the evolutions of the functional groups and
moreover, the temperature at which the Lc began decreasing was lower carbon microcrystalline structures in coal were investigated. The con­
in superheated water vapor pyrolysis. This suggests that superheated clusions are outlined below.
water vapor pyrolysis contributed to longitudinal thinning of the aro­
matic stacking structure. (1) During the coal pyrolysis process, the moisture was removed
Panels (c) and (d) of Fig. 15 present the temperature-dependent before 300 ◦ C and the volatile components were removed at high
changes in aromaticity and coal rank, respectively, during the two py­ rates in the 350◦ C-500 ◦ C range. Both pyrolysis processes exerted
rolysis methods. In both methods, the aromaticity and coal rank showed very similar effects on moisture removal, but superheated water
no significant changes at temperatures below 300 ◦ C, but significantly vapor pyrolysis proved more effective than conventional pyrol­
increased from approximately 0.48 and 1.25, respectively, at 300 ◦ C to ysis for volatile-component removal at 350◦ C-500 ◦ C.
0.83 and 4.00, respectively, at 500 ◦ C. The aromaticity increase can be (2) As the temperature increased, the percentages of the aliphatic
explained by depolymerization of the macromolecular structure of coal, structures C–– O, C–O, and aromatic C– – C groups in coal generally
which accompanies the pyrolysis and gasification of many aliphatic decreased while the –OH percentages generally increased and the
chains. Consequently, the number of aliphatic carbons decreases and the aromatic C–H-group percentages tended to decrease and then
aromaticity increases [44,90]. The aromaticity and coal rank were increase. Between 400 ◦ C and 500 ◦ C, superheated water vapor
higher after superheated water vapor pyrolysis than after conventional pyrolysis facilitated the removal of aliphatic structures from coal.

12
C. Zhang et al. Energy 276 (2023) 127613

By promoting aromatic substitution reactions during the whole [3] Yang Z, Khatri D, Verma P, Li T, Adeosun A, Kumfer BM, et al. Experimental study
and demonstration of pilot-scale, dry feed, oxy-coal combustion under pressure.
pyrolysis process, superheated water vapor pyrolysis yields a
Appl Energy 2021;285:116367. https://doi.org/10.1016/j.apenergy.2020.116367.
higher percentage content of aromatic C–H groups than con­ [4] Zhang D, Wang J, Lin Y, Si Y, Huang C, Yang J, et al. Present situation and future
ventional pyrolysis. However, the evolution patterns of the prospect of renewable energy in China. Renew Sustain Energy Rev 2017;76:
oxygen-containing functional groups are essentially the same 865–71. https://doi.org/10.1016/j.rser.2017.03.023.
[5] Xu R, Dai B, Wang W, Schenk J, Xue Z. Effect of iron ore type on the thermal
under both pyrolysis methods. behaviour and kinetics of coal-iron ore briquettes during coking. Fuel Process
(3) The degree of ordering of the carbon microcrystalline structure in Technol 2018;173:11–20. https://doi.org/10.1016/j.fuproc.2018.01.006.
coal gradually increased with temperature. This result was [6] Umar DF, Usui H, Daulay B. Change of combustion characteristics of Indonesian
low rank coal due to upgraded brown coal process. Fuel Process Technol 2006;87:
attributed to removal of the side chain groups (e.g., the highly 1007–11. https://doi.org/10.1016/j.fuproc.2006.07.010.
disordered aliphatic groups) and condensation of aromatic rings. [7] Kelly KE, Wang D, Hradisky M, Silcox GD, Smith PJ, Eddings EG, et al.
Superheated water vapor pyrolysis obtains a more ordered car­ Underground coal thermal treatment as a potential low-carbon energy source. Fuel
Process Technol 2016;144:8–19. https://doi.org/10.1016/j.fuproc.2015.12.006.
bon microcrystalline structure than conventional pyrolysis [8] Zhang HR, Li S, Kelly KE, Eddings EG. Underground in situ coal thermal treatment
because it removes the side-chain groups and reduces the depo­ for synthetic fuels production. Prog Energy Combust Sci 2017;62:1–32. https://doi.
sition of small molecular compounds on the carbon surface. org/10.1016/j.pecs.2017.05.003.
[9] Wang D, Fletcher TH, Mohanty S, Hu H, Eddings EG. Modified CPD model for coal
(4) Increasing the temperature gradually decreased the aromatic devolatilization at underground coal thermal treatment conditions. Energy Fuel
stacking height (Lc) but did not significantly change the layer 2019;33:2981–93. https://doi.org/10.1021/acs.energyfuels.8b04425.
spacing (d002) of the aromatic layers. Longitudinal thinning of the [10] Bolonkin A, Friedlander J, Neumann S. Innovative unconventional oil extraction
technologies. Fuel Process Technol 2014;124:228–42. https://doi.org/10.1016/j.
aromatic stacking structure was more promoted under super­
fuproc.2014.01.024.
heated water vapor pyrolysis than under conventional pyrolysis. [11] Abdul Waheed Bhutto, Aqeel Ahmed B, Gholamreza Z. Underground coal
Both the aromaticity and coal rank apparently increased by gasification: from fundamentals to applications. Prog Energy Combust Sci 2013;39:
different degrees as the temperature increased. Compared with 189–214. https://doi.org/10.1016/j.pecs.2012.09.004.
[12] Wang Y, Yangsheng Z, Feng Z. Study of evolution characteristics of pore structure
conventional pyrolysis, superheated water vapor pyrolysis during flame coal pyrolysis. Yanshilixue Yu Gongcheng Xuebao/Chinese J Rock
slightly improves the aromaticity and coal rank in the 300◦ C- Mech Eng 2010;29(9):1859–66.
500 ◦ C range. [13] Zhang C, Zhao Y, Feng Z, Zhao P, Wang X. Investigation on the effect of
superheated water vapor on gas production from pyrolysis of long-flame coal.
Chem Eng Commun 2021;209:1151–64. https://doi.org/10.1080/
Credit author statement 00986445.2021.1922894.
[14] Wang L, Yang D, Kang Z, Zhao J, Meng Q. Journal of Analytical and Applied
Pyrolysis Experimental study on the effects of steam temperature on the pore-
Chao Zhang: Conceptualization, Methodology, Writing - Review & fracture evolution of oil shale exposed to the convection heating. J Anal Appl
Editing. Zijun Feng: Visualization, Funding acquisition. Yangsheng Pyrolysis 2022;164:105533. https://doi.org/10.1016/j.jaap.2022.105533.
Zhao: Investigation, Supervision. Lei Wang: Data Curation. Yang Lu: [15] Kang Z, Zhao Y, Yang D, Tian L, Li X. A pilot investigation of pyrolysis from oil and
gas extraction from oil shale by in-situ superheated steam injection. J Pet Sci Eng
Formal analysis. Qiaorong Meng: Software. Qiang Gao: Validation. 2020;186:106785. https://doi.org/10.1016/j.petrol.2019.106785.
[16] Zhang C, Zhao Y, Feng Z, Meng Q, Wang L, Lu Y. Thermal maturity and chemical
structure evolution of lump long-flame coal during superheated water vapor –
Declaration of competing interest based in situ pyrolysis. Energy 2023;263:125863. https://doi.org/10.1016/j.
energy.2022.125863.
The authors declare that they have no known competing financial [17] Wu C, Zhuo Y, Xu X, Farajzadeh E, Dou J, Yu J. A combined experimental and
numerical study of coal briquettes pyrolysis using recycled gas in an industrial
interests or personal relationships that could have appeared to influence scale pyrolyser. Powder Technol 2022;404:117477. https://doi.org/10.1016/j.
the work reported in this paper. powtec.2022.117477.
[18] Xu K, Hu S, Zhang L, Li H, Chen Y, Xiong Z, et al. Effect of temperature on Shenfu
coal pyrolysis process related to its chemical structure transformation. Fuel Process
Data availability Technol 2021;213:106662. https://doi.org/10.1016/j.fuproc.2020.106662.
[19] Niu Z, Liu G, Yin H, Zhou C, Wu D, Tan F. A comparative study on thermal
Data will be made available on request. behavior of functional groups in coals with different ranks during low temperature
pyrolysis. J Anal Appl Pyrolysis 2021;158:105258. https://doi.org/10.1016/j.
jaap.2021.105258.
Acknowledgements [20] Krumm RL, Gneshin KW, Deo M. Adsorption characteristics of coals pyrolyzed at
slow heating rates. Energy Fuel 2017;31:1803–10. https://doi.org/10.1021/acs.
energyfuels.6b03116.
This work was supported by the National Natural Science Foundation [21] Yu J, Lucas JA, Wall TF. Formation of the structure of chars during devolatilization
of China (Grant Nos. 51974191, 52122405, and 52104144), China Na­ of pulverized coal and its thermoproperties: a review. Prog Energy Combust Sci
tional Postdoctoral Program for Innovative Talents (Grant No. 2007;33:135–70. https://doi.org/10.1016/j.pecs.2006.07.003.
[22] Mathews JP, Hatcher PG, Scaroni AW. Particle size dependence of coal volatile
BX2021209), and China Postdoctoral Science Foundation (Grant No. matter: is there a non-maceral-related effect? Fuel 1997;76:359–62. https://doi.
2022M712337). Moreover, we thank Shiyanjia Lab (www.shiyanjia. org/10.1016/S0016-2361(96)00220-7.
com) for its linguistic assistance during the preparation of this manu­ [23] Anthony DB, Howard JB, Hottel HC, Meissner HP. Rapid devolatilization and
hydrogasification of bituminous coal. Fuel 1976;55:121–8. https://doi.org/
script. The authors greatly appreciate the suggestions from the anony­
10.1016/0016-2361(76)90008-9.
mous reviewers for improving the paper’s quality. [24] Wall TF, Liu GS, Wu HW, Roberts DG, Benfell KE, Gupta S, et al. The effects of
pressure on coal reactions during pulverised coal combustion and gasification. Prog
Energy Combust Sci 2002;28:405–33. https://doi.org/10.1016/S0360-1285(02)
Appendix A. Supplementary data 00007-2.
[25] Alam SM, Wijayanta AT, Nakaso K, Fukai J, Norinaga K, Hayashi J ichiro.
Supplementary data to this article can be found online at https://doi. A reduced mechanism for primary reactions of coal volatiles in a plug flow reactor.
Combust Theor Model 2010;14:841–53. https://doi.org/10.1080/
org/10.1016/j.energy.2023.127613.
13647830.2010.517273.
[26] Wijayanta AT, Saiful Alam M, Nakaso K, Fukai J. Numerical investigation on
References combustion of coal volatiles under various O 2/CO 2 mixtures using a detailed
mechanism with soot formation. Fuel 2012;93:670–6. https://doi.org/10.1016/j.
fuel.2011.10.003.
[1] Zhong S. Structural decompositions of energy consumption between 1995 and
[27] Jiang H, Wang M, Li Y, Jin L, Hu H. Integrated coal pyrolysis with steam reforming
2009: evidence from WIOD. Energy Pol 2018;122:655–67. https://doi.org/
of propane to improve tar yield. J Anal Appl Pyrolysis 2020;147:104805. https://
10.1016/j.enpol.2018.08.017.
doi.org/10.1016/j.jaap.2020.104805.
[2] Wang M, Wan Y, Guo Q, Bai Y, Yu G, Liu Y, et al. Brief review on petroleum coke
[28] Wu L, Zhou J, Zhou J, Liang K, Song Y, Zhang Q, et al. Temperature-rising
and biomass/coal co-gasification: syngas production, reactivity characteristics, and
characteristics and product analysis of low-rank coal microwave pyrolysis under
synergy behavior. Fuel 2021;304:121517. https://doi.org/10.1016/j.
fuel.2021.121517.

13
C. Zhang et al. Energy 276 (2023) 127613

CH4 atmosphere. J Anal Appl Pyrolysis 2019;141:104632. https://doi.org/ [56] Niu B, Niu M, Zhang J, Liu R, Zhong H, Hu H. Novel insight into the mechanism of
10.1016/j.jaap.2019.104632. coal hydropyrolysis using deuterium tracer method. Fuel 2022;321:124109.
[29] Ma C, Zou C, Zhao J, Shi R, Li X, He J, et al. Pyrolysis characteristics of low-rank https://doi.org/10.1016/j.fuel.2022.124109.
coal under a CO-containing atmosphere and properties of the prepared coal chars. [57] Xiong G, Li Y, Jin L, Hu H. In situ FT-IR spectroscopic studies on thermal
Energy Fuel 2019;33:6098–112. https://doi.org/10.1021/acs. decomposition of the weak covalent bonds of brown coal. J Anal Appl Pyrolysis
energyfuels.9b00860. 2015;115:262–7. https://doi.org/10.1016/j.jaap.2015.08.002.
[30] Yardim MF, Ekinci E, Minkova V, Razvigorova M, Budinova T, Petrov N, et al. [58] Xu T, Srivatsa SC, Bhattacharya S. In-situ synchrotron IR study on surface
Formation of porous structure of semicokes from pyrolysis of Turkish coals in functional group evolution of Victorian and Thailand low-rank coals during
different atmospheres. Fuel 2003;82:459–63. https://doi.org/10.1016/S0016- pyrolysis. J Anal Appl Pyrolysis 2016;122:122–30. https://doi.org/10.1016/j.
2361(02)00295-8. jaap.2016.10.009.
[31] Xiong X, Miao Y, Lu X, Tan H, ur Rahman Z, Li P. C1~C2 hydrocarbons generation [59] Miura K, Mae K, Li W, Kusakawa T, Morozumi F, Kumano A. Estimation of
and mutual conversion behavior in coal pyrolysis process. Fuel 2022;308:121929. hydrogen bond distribution in coal through the analysis of OH stretching bands in
https://doi.org/10.1016/j.fuel.2021.121929. diffuse reflectance infrared spectrum measured by in-situ technique. Energy Fuel
[32] Zhong M, Zhang Z, Zhou Q, Yue J, Gao S, Xu G. Continuous high-temperature 2001;15:599–610. https://doi.org/10.1021/ef0001787.
fluidized bed pyrolysis of coal in complex atmospheres: product distribution and [60] Shi L, Liu Q, Guo X, Wu W, Liu Z. Pyrolysis behavior and bonding information of
pyrolysis gas. J Anal Appl Pyrolysis 2012;97:123–9. https://doi.org/10.1016/j. coal — a TGA study. Fuel Process Technol 2013;108:125–32. https://doi.org/
jaap.2012.04.009. 10.1016/j.fuproc.2012.06.023.
[33] Qiu Q, Pan D, Zeng F, Liu L, Qiu X. Catalytic effect of metal chlorides on coal [61] Iino M. Network structure of coals and association behavior of coal-derived
pyrolysis using TG & PY-gc/MS. Combust Sci Technol 2020:1. https://doi.org/ materials. Fuel Process Technol 2000;62:89–101. https://doi.org/10.1016/S0378-
10.1080/00102202.2020.1863954. –13. 3820(99)00120-4.
[34] Zhang K, Lu P, Guo X, Wang L, Lv H, Wang Z, et al. High-temperature pyrolysis [62] Niu Z, Liu G, Yin H, Zhou C, Wu D, Yousaf B, et al. In-situ FTIR study of reaction
behavior of two different rank coals in fixed-bed and drop tube furnace reactors. mechanism and chemical kinetics of a Xundian lignite during non-isothermal low
J Energy Inst 2020;93:2271–9. https://doi.org/10.1016/j.joei.2020.06.010. temperature pyrolysis. Energy Convers Manag 2016;124:180–8. https://doi.org/
[35] Yu G, Fan X, Liang P, Zhao GM, Hu X. Online characterization of pyrolysis products 10.1016/j.enconman.2016.07.019.
and kinetics study for the pyrolysis of a coal. J Anal Appl Pyrolysis 2021;160: [63] Li H, Shi S, Lin B, Lu J, Ye Q, Lu Y, et al. Effects of microwave-assisted pyrolysis on
105376. https://doi.org/10.1016/j.jaap.2021.105376. the microstructure of bituminous coals. Energy 2019;187:115986. https://doi.org/
[36] Chang Q, Gao R, Li H, Dai Z, Yu G, Liu X, et al. Effects of CO2 on coal rapid 10.1016/j.energy.2019.115986.
pyrolysis behavior and chemical structure evolution. J Anal Appl Pyrolysis 2017; [64] Feng XB, Cao JP, Zhao XY, Song C, Liu TL, Wang JX, et al. Organic oxygen
128:370–8. https://doi.org/10.1016/j.jaap.2017.09.012. transformation during pyrolysis of Baiyinhua lignite. J Anal Appl Pyrolysis 2016;
[37] Valdés CF, Chejne F. Effect of reaction atmosphere on the products of slow 117:106–15. https://doi.org/10.1016/j.jaap.2015.12.010.
pyrolysis of coals. J Anal Appl Pyrolysis 2017;126:105–17. https://doi.org/ [65] Cui X, Li X, Li Y, Li S. Evolution mechanism of oxygen functional groups during
10.1016/j.jaap.2017.06.019. pyrolysis of Datong coal. J Therm Anal Calorim 2017;129:1169–80. https://doi.
[38] Dong C, Jin L, Tao S, Li Y, Hu H. Xilinguole lignite pyrolysis under methane with or org/10.1007/s10973-017-6224-5.
without Ni/Al2O3 as catalyst. Fuel Process Technol 2015;136:112–7. https://doi. [66] Chen Y, Mastalerz M, Schimmelmann A. Characterization of chemical functional
org/10.1016/j.fuproc.2014.10.037. groups in macerals across different coal ranks via micro-FTIR spectroscopy. Int J
[39] Yu J, Guo Q, Ding L, Gong Y, Yu G. Study on the effect of inherent AAEM on char Coal Geol 2012;104:22–33. https://doi.org/10.1016/j.coal.2012.09.001.
structure evolution during coal pyrolysis by in-situ Raman and TG. Fuel 2021;292: [67] Li W, Zhu YM, Wang G, Jiang B. Characterization of coalification jumps during
120406. https://doi.org/10.1016/j.fuel.2021.120406. high rank coal chemical structure evolution. Fuel 2016;185:298–304. https://doi.
[40] Yu J, Guo Q, Ding L, Gong Y, Yu G. Studying effects of solid structure evolution on org/10.1016/j.fuel.2016.07.121.
gasification reactivity of coal chars by in-situ Raman spectroscopy. Fuel 2020;270: [68] Mathews JP, Chaffee AL. The molecular representations of coal - a review. Fuel
117603. https://doi.org/10.1016/j.fuel.2020.117603. 2012;96:1–14. https://doi.org/10.1016/j.fuel.2011.11.025.
[41] Li T, Zhang L, Dong L, Qiu P, Wang S, Jiang S, et al. Changes in char structure [69] Deuterium as a tracer in coal liquefaction part 1. The incorporation of deuterium
during the low-temperature pyrolysis in N2 and subsequent gasification in air of into liquid products. Fuel Process Technol 1982;5:281–98.
Loy Yang brown coal char. Fuel 2018;212:187–92. https://doi.org/10.1016/j. [70] Wang M, Roberts DG, Kochanek MA, Harris DJ, Chang L, Li CZ. Raman
fuel.2017.10.026. spectroscopic investigations into links between intrinsic reactivity and char
[42] Wu S, Gu J, Li L, Wu Y, Gao J. The reactivity and kinetics of yanzhou coal chars chemical structure. Energy Fuel 2014;28:285–90. https://doi.org/10.1021/
from elevated pyrolysis temperatures during gasification in steam at 900-1200◦ C. ef401281h.
Process Saf Environ Protect 2006;84:420–8. https://doi.org/10.1205/psep06031. [71] Ullah H, Chen B, Shahab A, Naseem F, Rashid A, Lun L, et al. Influence of
[43] Liang D, Xie Q, Wan C, Li G, Cao J. Evolution of structural and surface chemistry hydrothermal treatment on selenium emission-reduction and transformation from
during pyrolysis of Zhundong coal in an entrained-flow bed reactor. J Anal Appl low-ranked coal. J Clean Prod 2020;267:122070. https://doi.org/10.1016/j.
Pyrolysis 2019;140:331–8. https://doi.org/10.1016/j.jaap.2019.04.010. jclepro.2020.122070.
[44] Sun J, Feng H, Kou J, Jin H, Chen Y, Guo L. Experimental investigation on carbon [72] Liu H, Jiang B. Differentiated evolution of coal macromolecules in localized
microstructure for coal gasification in supercritical water. Fuel 2021;306:121675. igneous intrusion zone: a case study of Zhuxianzhuang colliery, Huaibei coalfield,
https://doi.org/10.1016/j.fuel.2021.121675. China. Fuel 2019;254. https://doi.org/10.1016/j.fuel.2019.115692.
[45] Zhang Y, Wu J, Wang Y, Miao Z, Si C, Shang X, et al. Effect of hydrothermal [73] Jiang J, Yang W, Cheng Y, Liu Z, Zhang Q, Zhao K. Molecular structure
dewatering on the physico-chemical structure and surface properties of Shengli characterization of middle-high rank coal via XRD, Raman and FTIR spectroscopy:
lignite. Fuel 2016;164:128–33. https://doi.org/10.1016/j.fuel.2015.09.055. implications for coalification. Fuel 2019;239:559–72. https://doi.org/10.1016/j.
[46] Wang Z, Shui H, Pei Z, Gao J. Study on the hydrothermal treatment of Shenhua fuel.2018.11.057.
coal. Fuel 2008;87:527–33. https://doi.org/10.1016/j.fuel.2007.03.017. [74] Xu J, Liu J, Zhang X, Ling P, Xu K, He L, et al. Chemical imaging of coal in micro-
[47] Liu X, He X. Effect of pore characteristics on coalbed methane adsorption in scale with Raman mapping technology. Fuel 2020;264:116826. https://doi.org/
middle-high rank coals. Adsorption 2017;23:3–12. https://doi.org/10.1007/ 10.1016/j.fuel.2019.116826.
s10450-016-9811-z. [75] Liu M, Bai J, Kong L, Bai Z, He C, Li W. The correlation between coal char structure
[48] Ibarra JV, Muñoz E, Moliner R. FTIR study of the evolution of coal structure during and reactivity at rapid heating condition in TGA and heating stage microscope.
the coalification process. Org Geochem 1996;24:725–35. https://doi.org/10.1016/ Fuel 2020;260:116318. https://doi.org/10.1016/j.fuel.2019.116318.
0146-6380(96)00063-0. [76] Wu D, Liu G, Sun R, Chen S. Influences of magmatic intrusion on the
[49] Guanhua N, Zhao L, Qian S, Shang L, Kai D. Effects of [Bmim][Cl] ionic liquid with macromolecular and pore structures of coal: evidences from Raman spectroscopy
different concentrations on the functional groups and wettability of coal. Adv and atomic force microscopy. Fuel 2014;119:191–201. https://doi.org/10.1016/j.
Powder Technol 2019;30:610–24. https://doi.org/10.1016/j.apt.2018.12.008. fuel.2013.11.012.
[50] Xin L, An M, Feng M, Li K, Cheng W, Liu W, et al. Study on pyrolysis characteristics [77] Zhang XP, Zhang C, Tan P, Li X, Fang QY, Chen G. Effects of hydrothermal
of lump coal in the context of underground coal gasification. Energy 2021;237: upgrading on the physicochemical structure and gasification characteristics of
121626. https://doi.org/10.1016/j.energy.2021.121626. Zhundong coal. Fuel Process Technol 2018;172:200–8. https://doi.org/10.1016/j.
[51] Xu M, Xin L, Liu W, Hu X, Cheng W, Li C, et al. Study on the physical properties of fuproc.2017.12.014.
coal pyrolysis in underground coal gasification channel. Powder Technol 2020; [78] Li X, Hayashi J ichiro, Li CZ. FT-Raman spectroscopic study of the evolution of char
376:573–92. https://doi.org/10.1016/j.powtec.2020.08.067. structure during the pyrolysis of a Victorian brown coal. Fuel 2006;85:1700–7.
[52] Liu S, Zhao H, Liu X, Li Y, Zhao G, Wang Y, et al. Effect of hydrothermal upgrading https://doi.org/10.1016/j.fuel.2006.03.008.
on the pyrolysis and gasification characteristics of baiyinhua lignite and a [79] Sadezky A, Muckenhuber H, Grothe H, Niessner R, Pöschl U. Raman
mechanistic analysis. Fuel 2020. https://doi.org/10.1016/j.fuel.2020.118081. microspectroscopy of soot and related carbonaceous materials: spectral analysis
[53] Niu Z, Liu G, Yin H, Zhou C. Devolatilization behaviour and pyrolysis kinetics of and structural information. Carbon N Y 2005;43:1731–42. https://doi.org/
coking coal based on the evolution of functional groups. J Anal Appl Pyrolysis 10.1016/j.carbon.2005.02.018.
2018;134:351–61. https://doi.org/10.1016/j.jaap.2018.06.025. [80] Xu Y, Chen X, Wang L, Bei K, Wang J, Chou IM, et al. Progress of Raman
[54] Seehra MS, Ghosh B. Free radicals, kinetics and phase changes in the pyrolysis of spectroscopic investigations on the structure and properties of coal. J Raman
eight american coals. J Anal Appl Pyrolysis 1988;13:209–20. https://doi.org/ Spectrosc 2020;51:1874–84. https://doi.org/10.1002/jrs.5826.
10.1016/0165-2370(88)80023-8. [81] Zhu H, Wang X, Wang F, Yu G. In situ study on K2CO3-catalyzed CO2 gasification
[55] Seehra MS, Ghosh B, Zondlo JW, Mintz EA. Relationship of coal extraction with of coal char: interactions and char structure evolution. Energy Fuel 2018;32:
free radicals and coal. Macerals 1988;18:279–86. 1320–7. https://doi.org/10.1021/acs.energyfuels.7b03255.

14
C. Zhang et al. Energy 276 (2023) 127613

[82] Song Q, Zhao H, Jia J, Yang L, Lv W, Gu Q, et al. Effects of demineralization on the [90] Shi Q, Qin B, Bi Q, Qu B. An experimental study on the effect of igneous intrusions
surface morphology, microcrystalline and thermal transformation characteristics of on chemical structure and combustion characteristics of coal in Daxing Mine,
coal. J Anal Appl Pyrolysis 2020;145:104716. https://doi.org/10.1016/j. China. Fuel 2018;226:307–15. https://doi.org/10.1016/j.fuel.2018.04.027.
jaap.2019.104716. [91] Kamble AD, Mendhe VA, Chavan PD, Saxena VK. Insights of mineral catalytic
[83] Zubkova V, Prezhdo V, Strojwas A. Comparative analysis of structural effects of high ash coal on carbon conversion in fluidized bed Co-gasification
transformations of two bituminous coals with different maximum fluidity during through FTIR, XRD, XRF and FE-SEM. Renew Energy 2022;183:729–51. https://
carbonization. Energy Fuel 2007;21:1655–62. https://doi.org/10.1021/ doi.org/10.1016/j.renene.2021.11.022.
ef060562p. [92] Zhao L, Guanhua N, Hui W, Qian S, Gang W, Bingyou J, et al. Molecular structure
[84] Jones SP, Fain CC, Edie DD. Structural development in mesophase pitch based characterization of lignite treated with ionic liquid via FTIR and XRD spectroscopy.
carbon fibers produced from naphthalene. Carbon N Y 1997;35:1533–43. https:// Fuel 2020;272:117705. https://doi.org/10.1016/j.fuel.2020.117705.
doi.org/10.1016/S0008-6223(97)00106-1. [93] Wu D, Zhang W. Evolution mechanism of macromolecular structure in coal during
[85] Wu D, Chen B, Sun R, Liu G. Thermal behavior and Raman spectral characteristics heat treatment: based on FTIR and XRD in situ analysis techniques. J Spectrosc
of step-heating perhydrous coal: implications for thermal maturity process. J Anal 2019;2019. https://doi.org/10.1155/2019/5037836.
Appl Pyrolysis 2017;128:143–55. https://doi.org/10.1016/j.jaap.2017.10.015. [94] Sonibare OO, Haeger T, Foley SF. Structural characterization of Nigerian coals by
[86] Marques M, Suárez-Ruiz I, Flores D, Guedes A, Rodrigues S. Correlation between X-ray diffraction, Raman and FTIR spectroscopy. Energy 2010;35:5347–53.
optical, chemical and micro-structural parameters of high-rank coals and graphite. https://doi.org/10.1016/j.energy.2010.07.025.
Int J Coal Geol 2009;77:377–82. https://doi.org/10.1016/j.coal.2008.06.002. [95] Li N, Te G, Liu Q, Ban Y, Wang Y, Zhang X, et al. Effect of metal ions on the steam
[87] Zhang Y, Li Z. Raman spectroscopic study of chemical structure and thermal gasification performance of demineralized Shengli lignite char. Int J Hydrogen
maturity of vitrinite from a suite of Australia coals. Fuel 2019;241:188–98. https:// Energy 2016;41:22837–45. https://doi.org/10.1016/j.ijhydene.2016.09.018.
doi.org/10.1016/j.fuel.2018.12.037. [96] Peng Z, Ning X, Wang G, Zhang J, Li Y, Huang C. Structural characteristics and
[88] Pan R kun, Li C, gao Yu M, Xiao Z jun, Fu D. Evolution patterns of coal micro- flammability of low-order coal pyrolysis semi-coke. J Energy Inst 2020;93:
structure in environments with different temperatures and oxygen conditions. Fuel 1341–53. https://doi.org/10.1016/j.joei.2019.12.004.
2020;261:116425. https://doi.org/10.1016/j.fuel.2019.116425. [97] Burket CL, Rajagopalan R, Foley HC. Overcoming the barrier to graphitization in a
[89] Xu Q, Liu R, Ramakrishna S. Comparative experimental study on the effects of polymer-derived nanoporous carbon. Carbon N Y 2008;46:501–10. https://doi.
organic and inorganic acids on coal dissolution. J Mol Liq 2021;339:116730. org/10.1016/j.carbon.2007.12.016.
https://doi.org/10.1016/j.molliq.2021.116730.

15

You might also like