Вы находитесь на странице: 1из 9

09575820/06/$30.00+0.

00
# 2006 Institution of Chemical Engineers
Trans IChemE, Part B, November 2006
Process Safety and Environmental Protection, 84(B6): 420 428

www.icheme.org/psep
doi: 10.1205/psep06031

THE REACTIVITY AND KINETICS OF YANZHOU COAL


CHARS FROM ELEVATED PYROLYSIS TEMPERATURES
DURING GASIFICATION IN STEAM AT 900 120088C
S. WU, J. GU, L. LI, Y. WU and J. GAO
Department of Chemical Engineering for Energy Resources, East China University of Science and Technology, Shanghai, China

his paper studied on the steam gasification reactivity and kinetics for Yanzhou coal
chars prepared at elevated pyrolysis temperatures, using a fixed-bed reactor at the
atmospheric pressure and gasification temperature ranging from 9008C to 12008C.
The following conclusions were reached mainly. The effect of gasification temperatures on
the gasification reactivity was stronger than that of pyrolysis temperatures under the condition
of elevated pyrolysis temperatures. Changes of reaction rate and surface area had a similar
trend, but the relationship between the change of surface area and that of reaction rate was
not proportional during the reaction process. While simulating the steam gasification reaction
of Yanzhou coal chars at the temperatures of 900 12008C, the shrinking un-reacted core
model (SUCM) was superior to the volume reaction model (VRM) and the reaction diffusion
model based on the shrinking un-reaction core model was more favourable, which suggested
that the diffusion had an effect on the reaction process, but its effect was not rather strong.
Therefore, it may be considered that the reaction process was mainly controlled by the chemical reaction. Apparent activation energies obtained for the steam gasification of Yanzhou coal
chars ranged from 127.17 kJ mol21 to 196.55 kJ mol21. Moreover, it was also found that kinetic parameters were in accord with the isokinetic relationship, according to which the calculated isokinetic temperature for all the reactions was 11978C.
Keywords: coal char; elevated temperature pyrolysis; steam gasification; kinetics model.

INTRODUCTION

for coal chars produced at elevated pyrolysis temperatures


is highly significant and acquiring the kinetic parameters
of the gasification reaction is essential for designing an efficient gasifier.
So far, in spite of extensive studies (Jan et al., 1998;
Kajitani and Matsuda, 1998; Liu et al., 2000a, b, 2004;
Cetin et al., 2004; Kajitani et al., 2002a, b; Luo et al.,
2001; Devi and Kannan, 2000; Bayarsaikhan et al.,
2005a, b; Zhou et al., 2002; Sekine et al., 2006; Bhat
et al., 2001), the majorities are those on the CO2 gasification reaction for coal chars, therefore there is still a
strong need for more reactivity-related data on steam gasification for coal chars, particlarly for those prepared at
elevated pyrolysis temperatures. As a fundamental research
on coal gasification, the main objects of this work are to
investigate the gasification reactivity for chars prepared at
elevated pyrolysis temperatures and to obtain valid kinetics
data for a typical coal from China in steam.

Coal gasification is an efficient technology for coal utilization and becomes increasingly important, due to its high
carbon conversion and reduction of air pollutant emision.
The operation temperature in some gasfiers is as high as
14008C and even higher. Thus, coal particles in the gasifier
undergo elevated pyrolysis temperatures. It is well-known
that pyrolysis conditions strongly affects the reactivity of
resulting chars and hence the overall performance of a
coal gasifier. Comparing with low pyrolysis temperatures,
elevated pyrolysis temperatures as one of pyrolysis conditions, may result in the deformation or fusion of ash,
the change in intra-particle pore structure, a decrease of
active surface area of resulting chars and a better crystallinity of the carbonaceous part (Skodras and Sakellaropoulos,
2002; Leboda et al., 1998; Sekine et al., 2005; Jan et al.,
1998; Kajitani and Matssuda, 1998; Liu et al., 2004; Cetin
et al., 2004), and consequently a different behaviour or
mechanism of gasification reaction for resulting chars.
Accordingly, the clarification of gasification characteristics

EXPERIMENTAL
Preparation of Coal Chars

Correspondence to: Professor J. Gao, Department of Chemical Engineering for Energy Resources, East China University of Science and Technology, 200237 Shanghai, China.
E-mail: gjs@ecust.edu.cn

Yanzhou coal, one of the typical Chinese coals, was used


in this study. Four char samples were prepared under elevated pyrolysis temperatures ranging from 9508C to 14008C
420

THE REACTIVITY AND KINETICS OF YANZHOU COAL CHARS

421

by devolatilization of parent coal in the muffle. Coal


samples were submitted to a heating rate of 68C min21
up to a desired temperature at the nitrogen atmosphere
and held at this temperature for 20 min. Char samples
were ground, screened to 3 6 mm and stored for the
measurement of the reactivity. Char samples, whose pyrolysis temperatures were 9508C, 12008C, 13008C and
14008C, were respectively marked as SP950, SP1200,
SP1300 and SP1400 in the study. The proximate and ultimate analysis data of the coal sample and resulting char
samples are shown in Table 1.
Measurements of the Steam-Char Gasification
Reactivity
The experiment for coal char gasification was conducted
using a fixed-bed reactor, as shown in Figure 1. The reactor
was a corundum tube with the inner diameter of 20 mm.
Coal char samples with the height of 70 mm located
between the two layer fillings, which were ceramic rings
with the outer diameter of 4 mm, the inner diameter of
2 mm and the length of 4 mm. A thermocouple was
inserted into coal char samples to monitor its temperature.
The nitrogen gas and the water entered the reactor from the
top of the reactor. The water flow rate for steam generation
was controlled accurately by a dosing pump (WLB-78-A).
When the water was fed into the reactor through a stainless
steel tube, the steam was generated inside the reactor and
immediately mixed with a nitrogen flow. The steam concentration was held constant at 46% v/v.
In each run, 7 g of coal char was set in the reactor. The
reactor was heated at a rate of 208C min21 up to a prescribed holding temperature under downward a nitrogen
gas flow of 250 ml min21 as a sweeping gas. Once steam
generated, gasification products were collected by a
gasbag. At this time, the reading of the wet flowmeter
and the time respectively were marked as V0 and t0. In
turn, gaseous products were collected continuously by several gasbags and the reading of the wet flowmeter was
marked as Vn at tn. After a certain reaction time, the experiment was stopped and carbon oxides (CO and CO2) were
analysed by chemical absorption for each gasbag. The
basic carbon conversion was calculated by the following
equation:
P
Mc f ni0 (Vi1  Vi )(yi1,1 yi1,2 )
X(tn )
Vmol FCad W
 (n 0, 1, 2, . . . )
(1)

Table 1. Proximate and ultimate analysis of original coal and resulting


chars.
Raw coal

SP950

Proximate analysis (wt%, dry basis)


Ash
12.90
17.01
Volatile matter
38.82
0.92
Fixed carbon
48.28
82.07
Ultimate analysis (wt%, dry basis)
C
66.63
81.45
H
4.70
0.50
2.49
2.39
St
N
0.94
0.42

SP1200

SP1300

SP1400

17.22
0.76
82.02

17.28
0.62
82.10

17.31
0.60
82.09

81.7
0.21
2.35
0.11

81.75
0.18
2.21
0.12

81.78
0.17
2.00
0.09

Figure 1. The schematic diagram of the experimental set-up for steam


char gasification. 1: temperature controller; 2: flume; 3: dosing pump;
4: rubber plug; 5: stainless steel pipe; 6: furnace; 7: filling; 8: corundum
reactor; 9: absorption bottle; 10: wet flowmeter; 11: Gasbag.

Where tn is the reaction time, X(tn) is the carbon conversion at tn, MC is the atomic weight of carbon, f is the correctional factor of the wet flowmeter, Vi is the reading of
the wet flowmeter at ti, Vmol is the mole volume of gas at
the environmental temperature and the atmospheric
pressure, W represents the initial weight of coal char and
FCad is the content of fixed carbon in char. yi1,1 and
yi1,2 are respectively the volumetric contents of CO
and CO2 in the No. (i 1) gasbag.
Measurement of BET Surface Area
The analysis of BET surface area was conducted by N2
adsorption technique using the pore structure analyser
(ASAP 2400) produced by Micrometrics Instrument Co.
RESULTS AND DISCUSSION
In this study, experimental runs of the steam gasification
reaction for four chars, which were SP950, SP1200,
SP1300 and SP1400, were conducted at temperatures ranging from 9008C to 12008C, as shown in Figure 2. The
figure shows that the basic carbon conversion is very sensitive to gasification temperature variations, which suggests
that the chemical reaction is probably the rate controlling
step.
Effects of Gasification Temperatures and Pyrolysis
Temperatures on the Reactivity
The average reaction rate constant (KS), which was calculated through simulating the gasification reaction with
the SUCM (seen in the third part of Results and Discussion), was used as a reactivity index in this study to characterize the gasification reactivity of different char samples
for quantitative comparison. Figures 3 and 4 show respectively the effects of gasification temperatures and pyrolysis
temperatures on the average rate constant. From Figure 3, it
can be seen that with respect to any of resulting chars, its

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

422

WU et al.

Figure 2. Effect of gasification temperatures on basic carbon conversion of various chars (gasification temperature: B 9008C; 10008C; O 11008C; P
12008C). A: SP950; B: SP1200; C: SP1300; D: SP1400.

average rate constant increases in a large extent with the


increase of gasification temperatures and that the steam
gasification reactivity of all chars prepared at different
pyrolysis temperatures, increases with the increase of gasification temperatures ranging from 9008C to 12008C. It is
well-known that at higher gasification temperatures, the
reaction can acquire more energy which accelerates
the reaction in chemical controlling. Figure 4 shows that
the average rate constant of steam gasification reaction
for pyrolysis chars decreases with the increase of pyrolysis
temperatures and the decreasing extent is smaller and
smaller as gasification temperature increases. The results
suggest that the steam-char gasification reactivity decreases
with the increase of pyrolysis temperatures and the influencing extent of pyrolysis temperatures on the gasification

reactivity depends on gasification temperatures. The


reason that the reactivity is lower for the char prepared at
higher pyrolysis temperatures, may be mainly attributed
to some factors similar to those of the CO2 char gasification: a more ordered structure and better crystallinity of
the carbonaceous part (Kajitani and Matsuda, 1998; Liu
et al., 2004; Cetin et al., 2004; Devi et al., 2000), reducing
active sites concentration which is associated with specific
surface area values or micro-porous volume and gas
reactant accessing uneasily to active sites due to the shrinkage of macro-pores which is characterized by the porosity
value and pore size distribution of chars (Liu et al.,
2000b; Sekine et al., 2006).
Furthermore, while comparing the data in Figure 3 with the
data in Figure 4, it can be found that for Yanzhou coal, the

Figure 3. Effect of gasification temperatures on average rate constants B


SP950; SP1200; OSP1300; PSP1400.

Figure 4. Effect of pyrolysis temperatures on average rate constants


B 9008C; 10008C; O 11008C; P 12008C.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

THE REACTIVITY AND KINETICS OF YANZHOU COAL CHARS


Table 2. BET surface area of SP1400 at different carbon conversions.
Carbon conversion (%, daf)
BET surface area (m2 g21)

0a
0.89

36.28
30.81

63.90
40.09

73.34
49.08

91.60
43.12

Original char sample.

effect of gasification temperatures on resulting char gasification reactivity is stronger than that of pyrolysis temperatures
under the condition of elevated pyrolysis temperatures.

Changes of BET Surface Area and Reaction Rate


During the Reaction Process
Generally, it is thought that BET surface area of the char
changes as the reaction advances. However, there was also
such finding (Asanez and DeDiego, 1993) that was no any
variation in surface area. Therefore, the SP1400 was
selected as a representative sample in this work and the
BET surface areas at different carbon conversions during
steam gasification at 11008C, were measured to obtain
useful information about changes of BET surface area of
chars during the reaction process. The result is that the
BET surface area increases gradually and then declines
with carbon conversions, as shown in Table 2. The reasons
for the result are mainly related to changes of the porosity
of chars during the gasification process. As the reaction
processed, the porosity of chars increased gradually at
low carbon conversions due to the closed pores being
opened and the opening pores being enlarged simultaneously, while it decreased at high carbon conversions
due to the neighbouring pores amalgamating and the

423

pores disappearing, which were caused from the carbonaceous part of chars being consumed.
It is generally thought that the reaction rate of the char
changes with conversion. Here, it can also be seen that
the reaction rate increases initially with the increase of
carbon conversion, reaches a maximum at a certain conversion, and then declines after that (see Figure 5). Normally,
the change of the reaction rate would be related to changes
of surface area during reaction and the increasing surface
area can result in the increasing reaction rate, but it is not
the case. For example, making a comparison of the data
presented in Figure 5(d) and Table 2, those could be
found that at the temperature of 11008C, the reaction rate
of the SP1400 presents a maximum at the conversion of
about 30%, while after the point the surface area still
increases and the reaction rate decreases due to the
decrease of the active sites. The results indicate that
changes of reaction rate and surface area have the similar
trend, but the relationship between the change of surface
area and that of reaction rate is not proportional during
the gasification reaction process. The relationship between
surface area and reaction rate has been widely investigated,
but there is no general agreement. Adshiri et al. (1986) considered that the gasification rate is proportional to the surface area during gasification. Adanez and DeDiego
(1993) found that the reaction rate changed as the reaction
advances, while there was not any variation of surface
areas. As a matter of fact, most of studies (Yang and Watkinson, 1994; Alvarez et al., 1995; Kasaoka et al., 1987;
Hashimoto et al., 1986) found that surface area and reaction
rate were not proportional and proportionality was rather
found between reaction rate and active surface area related
to the number of active sites.

Figure 5. Changes of the reaction rate as the carbon conversion for four chars at different temperatures (gasification temperature: B 9008C; 10008C;
O11008C; P12008C). A: SP950; B: SP1200; C: SP1300; D: SP1400.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

424

WU et al.

Figure 6. 2 ln(1 2 X) versus t for various coal chars at different temperatures (gasification temperature: B 9008C; 10008C; O11008C; P12008C).
A: SP950; B: SP1200; C: SP1300; D: SP1400.

Gasification Kinetics Models for Coal Chars from


Elevated Pyrolysis Temperatures
Gasification of coal char particles is usually classified as
an irreversible gas solid reaction, which is commonly
described using the VRM and the SUCM. The two
models have been applied to simulating the gasification
process in the study. The basic conceptions of the two
models are respectively introduced as follows.
The VRM proposed by Wen (1968) assumed that the gas
reactant was reacting with char in all possible places, both
outside and inside the particle surface. That was to say, all

of the subparticles in chars reacted uniformly. It was


represented by the expression:
 ln (1  X) Kv t

(2)

The SUCM also proposed by Wen (1968) thought that


reacting char particles were spherical grains. Initially, the
reaction occurred at the external surface of char particle,
and the front gradually moved inside leaving an ash layer
behind. At the intermediate conversions of the solid, there
was essentially a shrinking core of unreacted solid, whose

Figure 7. 1 2 (1 2 X)1/3 versus t for various coal chars at different temperatures (gasification temperature: B 9008C; 10008C; O11008C; P12008C).
A: SP950; B: SP1200; C:SP1300; D: SP1400.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

THE REACTIVITY AND KINETICS OF YANZHOU COAL CHARS

425

Table 3. Correlative indexes obtained using the SUCM (RSUCM) and the VRM (RVRM), respectively.
SP950

SP1200

SP1300

SP1400

Gasification temperature

RSUCM

RVRM

RSUCM

RVRM

RSUCM

RVRM

RSUCM

RVRM

9008C
10008C
11008C
12008C

0.9878
0.9969
0.9889
0.9915

0.9864
0.9921
0.9758
0.9654

0.9922
0.9964
0.9932
0.9767

0.9918
0.9923
0.9834
0.9393

0.9885
0.9962
0.9940
0.9766

0.9881
0.9941
0.9881
0.9361

0.9940
0.9918
0.9961
0.9894

0.9938
0.9892
0.9893
0.9678

radius diminished as the reaction advanced. Only the


chemical reaction at the surface was the controlling step,
carbon conversion with t was described as:
1  (1  X)1=3

Ks
t
3

(3)

In equations (2) and (3), KV, KS, X and t are respectively the volume reaction rate constant, the surface reaction rate constant, the basic carbon conversion and the
reaction time.
Figures 6 and 7 respectively, show that the experimental
data of the gasification for four chars with steam fitted to
the VRM and the SUCM, and correlative indexes (R) calculated are shown in Table 3. The table represents that the fitting extent of the SUCM is better than that of the VRM for
any one of Yanzhou coal chars at any gasification temperature, which indicates that the SUCM is superior to the
VRM for simulating the steam gasification reaction for
Yanzhou coal chars at gasification temperatures ranging
from 9008C to12008C.
Moreover, from Figure 7 it can be found that the plots
show worse and worse linearity as the reaction temperature
increases, meaning that the gas diffusion affects the gasification reaction of coal chars with steam.
Szekely et al. (1979, 1980) and Park and Levenspiel
(1975) considered that a porous spherical pellet of solid
was made up of a large number of spherical grains with uniform initial radius, and proposed that while considering the
effect of the diffusion on reaction process, the relationship
of the conversion and the reaction time is approximately
expressed as the equation (4) based on the SUCM.
t A{1  (1  X))1=3  s2 1  3(1  X)2=3
2(1  X)}



1=2
3 r m Vg
Vp 3(1  1)k Ag
1
A
; s
1
2De
KE
kCA0 Ag
Ap
Fg Vg
(4)
where Ag and Ap are respectively surface area of grain and
external surface area of pellet, Vg and Vp are respectively
volume of grain and pellet, rm is molar density of chars,
1 is porosity of chars, CA0 is concentration of gaseous reactant, k is reaction rate constant, De represents effective diffusivity of gases in porous chars and KE is equilibrium
constant. s2 reflects the influencing extent of the diffusion
on the reaction. The larger value of s 2 indicates that the
gas diffusion has a stronger effect on the reaction. A and
s 2 are calculated by the least-square.

Consequently, considering the gas diffusion during the


steam gasification, the gasification process for Yanzhou
coal chars at temperatures of 900 12008C was simulated
using the reaction diffusion model proposed above. The
results of simulation are shown in Figure 8. Comparing
Figures 7 and 8, it can be seen that the simulation of
the reaction process using the reaction diffusion model
is more accurate than that only using the reaction
model. At the same time, from Table 4 it is shown that
the value of s2 increases with the increasing of gasification temperatures for the same coal char. This suggests
that the influencing extent of the gas diffusion on the reaction is larger and larger as the gasification temperature
increases. However, all values of s2 in Table 4 are
quite small, which indicates that the effect of the gas diffusion on the gasification reaction is limited. In a word,
there is a little effect of the diffusion on the steam gasification reaction of Yanzhou coal chars, especially at the
higher gasification temperature, but it may be thought
that speaking as a whole, the reaction process is mainly
controlled by the chemical reaction.
Activation Energies for Steam Gasification for Elevated
Temperature Pyrolysis Chars
Through experienced Arrhenius plots as shown in
Figure 9, pre-exponential factors (K0) and apparent activation energies were obtained for various chars using the
SUCM. The results are shown in Table 5. The table
shows that the value of apparent activation energy increases
with the increase of pyrolysis temperature, which also tells
us that the increase of pyrolysis temperatures diminishes
the steam gasification reactivity of Yanzhou coal chars.
This also approves the result as shown before.
Here, apparent activation energies obtained for the steam
gasification of Yanzhou coal chars are from
127.17 kJ mol21to 196.55 kJ mol21. For the gasification of
coal or coal chars in steam, wide variations in the values
of apparent activation energy were reported by many
researchers, as shown in Table 6. Thus, the values of activation energy obtained in this work for four coal chars are
at the higher end of all those reported in literatures.
Compensation Effects in Steam Char Gasification
It has been found in a number of cases, especially for
coal char gasification reaction, that the apparent activation
energy (Ea) and the pre-exponential factor (K0) change in
the same direction, which is referred to as compensation
effect. Essenhigh and Misra (1990) illustrated the correlations of the activation energy and the pre-exponential
factor with their data and the data reported in literatures

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

426

WU et al.

Figure 8. Carbon conversion versus time of reaction for various chars using the reactiondiffusion model (gasification temperature: B 9008C; 10008C;
O 11008C; P 12008C). A: SP950; B: SP1200; C: SP1300; D: SP1400.

Table 4. Values of s2 obtained using the reactiondiffusion model.

Table 5. Apparent activation energy (Ea) and pre-exponential factor (K0)


estimated from the SUCM.

Char samples
SP950

Gasification
temperature

SP950

SP1200

SP1300

SP1400

9008C
10008C
11008C
12008C

0.0133
0.0356
0.0390
0.0898

0.0028
0.0279
0.0646
0.1376

0.0044
0.0227
0.0560
0.1252

0.0024
0.0200
0.0286
0.0788

for char gas reactions, and reached a conclusion that for a


wide range of coal char gasification reaction, kinetic parameters were autocorrelated, which could be regarded as
compensation effect. Xie (2002) investigated the catalytic
and uncatalytic gasification kinetics of various coals with
CO2 and concluded that compensation effect of gasification
was a true one. The present results also show that a marked
compensation effect is exhibited between activation

21

SP1200
2

9.27  10
K0, min
Ea, kJ mol21 124.17

SP1300
4

5.19  10
172.72

SP1400
5

2.18  10
190.71

3.39  105
196.55

energies and pre-exponential factors, as shown in


Figure 10. Before, it was considered that resulting in
compensation effect was due to the pure mathematic calculation method in statistics, and presently the majority of
researchers considered that in effect, it indicated the
similarity of the reaction mechanism in the same reaction
and it was related to the property of the active complex
(carbon oxygen complex) formed during the reaction
process.
The relation of compensation effect can be achieved by
linear-regression as the following equation:
ln (K0 ) aEa b

(5)

where a and b are referred to as compensation parameters.


According to the Arrhenius equation, the following
relation is yielded:
ln (K0 )

Figure 9. Arrhenius plots for various coal chars (A SP950; W SP1200;


4 SP1300; 5 SP1400).

Ea
ln (Ki )
RTi

(6)

where Ti is the isokinetic temperature for the reaction process and Ki is the reaction rate constant at Ti. Equation (6)
means that at the isokinetic temperature, all the reactions
with a given gasification agent proceed at the same rate.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

THE REACTIVITY AND KINETICS OF YANZHOU COAL CHARS

427

Table 6. Apparent activation energies reported by some literatures for the steam gasification reaction of coals and coal chars.
Samples

Reaction temperatures

Apparent activation energy

coal residue
In-situ and ex-situ coal chars

1040 14308C
1000 14008C

Lignite char

750 8908C
890 9308C
800 10008C
,10008C
1000 14008C
1000 14008C
750 9008C

82.84 kJ mol21
56 85 kJ mol21 (in-situ chars)
63 99 kJ mol21 (ex-situ chars)
176 kJ mol21
75 kJ mol21
165.27 kJ mol21
159193 kJ mol21
91 107 kJ mol21
121125 kJ mol21
200 kJ mol21

High ash coal


Four coals
Bituminous and lignite chars
Rice husk char

Figure 10. Compensation effect of steamchar gasification.

Based on equations (5) and (6), the compensation effect


parameters are expressed as
a

1
;
RTi

b ln (Ki )

(7)

Thus, the isokinetic temperature for steam gasification


reaction in the study is found to be 11978C, which is
close to the top of the experimental temperatures. The
result is different with Essenhigh and Misras (1990) consideration that the isokinetic temperature was close to the
midpoint value of the experimental temperature range.
Likewise, Xie et al.s (2002) result is that the isokinetic
temperature is close to the initial reaction temperature.
These indicate that there is no last conclusion for the problem that which point in the experimental temperature
range the isokinetic temperature is close to. Hence, more
investigations should be conducted to obtain more data
for this problem in the future.
CONCLUSIONS
In this work, the following conclusions were obtained
through investigating the steam gasification reactivity and
kinetics of Yanzhou coal chars from elevated pyrolysis
temperature at temperatures ranging from 9008C to
12008C.

(1) The reactivity of resulting chars increases with the


increase of gasification temperatures in a large extent,
while that decreases with the increase of pyrolysis
temperature in a relatively small extent, which depends

Literatures
Jensen (1975)
Peng et al. (1995)
Linares-Solano et al. (1979)
Schmal et al. (1983)
Kasaoka et al. (1985)
Tamhankar et al. (1982)
Bhat et al. (2001)

on gasification temperature. The influencing of gasification temperatures on resulting char gasification reactivity is stronger than that of pyrolysis temperatures at
elevated pyrolysis temperatures.
(2) BET surface area of resulting chars increases gradually
and then decreases as the carbon conversion, as does
the reaction rate of resulting chars. Although changes
of reaction rate and surface area have a similar trend,
the relationship between surface area and reaction
rate is not proportional during the gasification reaction
process.
(3) For simulating the steam gasification reaction process
of Yanzhou coal chars, the SUCM is superior to the
VRM, and under considering the effect of the diffusion
on the reaction, the reaction diffusion model based on
the SUCM is more favourable. However, the effect of
the diffusion on the reaction process at experimental
temperature range is quite limited and only appears a
little at the higher temperature. Thereby, it can be considered that the reaction process is mainly controlled by
the chemical reaction at the temperature ranging from
9008C to 12008C.
(4) Apparent activation energies obtained for resulting
char gasification reaction range from 127.17 kJ mol21
to 196.55 kJ mol21, which are at the higher end of all
those reported. It is also found that the activation
energy and pre-exponential satisfy the isokinetic
relationship and the isokinetic temperature calculated
is 11978C for all the reactions. The isokinetic temperature is close to the top of the experimental temperature
range and this finding is different from those reported
in literatures.

REFERENCES
Alvarez, T., Fuertes, A.B., Pis, J.J. and Ehrburger, P., 1995, Influence of
coal oxidation upon char gasification reactivity, Fuel, 74(5): 729 735.
Adanez, J. and DeDiego, R.F., 1993, Int Chem Eng, 33: 656.
Adshiri, T., Shiraha, T., Kojima, T. and Furuzawa, P., 1986, Prediction of
CO2 gasification rate of char in fluidized bed gasifier, Fuel, 65(2):
16881693.
Bayarsaikhan, B., Hayashi, J., Shimada, T., Sathe, C., Li, C., Tsutsumi, A.
and Chiba, T., 2005a, Kinetics of steam gasification of nascent char
from rapid pyrolysis of a Victorian brown coal, Fuel, 84(1213):
16121621.
Bayarsaikhan, B., Hayashi, J., Shimada, T., Sathe, C., Li, C.Z., Tsutsumi, A.
and Chiba, T., 2005b, Kinetics of steam gasification of nascent char from
rapid pyrolysis of a Victorian brown coal, Fuel, 84(1213): 16121621.
Bhat, A., Ram Bheemarasetti, J.V. and Rajeswara Rao, T., 2001, Kinetics
of rice husk char gasification, Energy Conversion and Management,
42(18): 2061 2069.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

428

WU et al.

Cetin, E., Moghtaderi, B., Gupta, R. and Wall, T.F., 2004, Influence of
pyrolysis conditions on the structure and gasification reactivity of biomass chars, Fuel, 83(16): 2139 2150.
Devi, T.G. and Kannan, M.P., 2000, Gasification of biomass chars in
air-effect of heat treatment temperature, Energy & Fuel, 14(1):
127 130.
Essenhigh, R.H. and Misra, M.K., 1990, Autocorrelations of kinetic parameters in coal and char reactions, Energy & Fuels, 4(2):171 177.
Hashimoto, K., Miura, K. and Ueda, T., 1986, Correlation of gasification
rates of various coals measured by a rapid heating method in a steam
atmosphere at relatively low temperatures, Fuel, 65(11): 15161523.
Jan, W.N., John, P.H. and Steven, A.B., 1998, The role of physical factors
in mass transport during sintering of coal ashes and deposit deformation
near the temperature of glass transformation, Fuel Processing Technology, 56(12): 89 101.
Jensen, G.A., 1975, The kinetics of gasification of carbon contained in coal
minerals at atmospheric pressure, Ind Eng Chem Proc Des Dev, 14:
308 314.
Kajitani, S. and Matsuda, H., 1998, Influence of pyrolysis conditions on
coal gasification reactivity. Report No. W97020 (Yokosuka Research
Laboratory, Central Research Institute of Electric Power Industry,
Japan).
Kajitani, S., Hara, S. and Matsuda, H., 2002, Gasification rate analysis of
coal char with a pressurized drop tube furnace, Fuel, 81(5): 539 546.
Kasaoka, S., Sakata, Y. and Tong, C., 1985, Kinetic evaluation of the reactivity of various coal chars for gasification with carbon dioxide in comparison with steam, Int Chem Eng, 25(1): 160 175.
Kasaoka, S., Sakata, Y. and Shimada, M., 1987, Effects of coal carbon
conditions on rate of steam gasification of char, Fuel, 66(5): 697 701.
Leboda, R., Zieba, J.S. and Grzegorczyk, W., 1998, Effect of calcium catalyst loading procedure on the porous structure of active carbon from
plum stones modified in the steam gasification process, Carbon,
36(4): 417 425.
Linares-Solano, A., Mahajan, O.P., Walker, P.L. Jr, 1979, Reactivity of
heat-treated coals in steam, Fuel, 58: 327 332.
Liu, G., Tata, A.G., Bryant, G.W. and Wall, T.F., 2000a, Mathematical
modeling of coal char reactivity with CO2 at high pressures and temperatures, Fuel, 79(10): 11451154.
Liu, G.S., Benyon, P., Benfell, K.E., Bryant, G.W., Tate, A.G., Boyd,
R.K., Harris, D.J. and Wall, T.F., 2000b, The porous structure of bituminous coal chars and its influence on combustion and gasification
under chemically controlled conditions. Fuel, 79(6): 617 626.
Liu, H., Kaneko, M., Luo, C., Kato, S. and Kojima, T., 2004, Effect of
pyrolysis time on the gasification reactivity of char with CO2 at elevated
temperatures, Fuel, 83(78): 10551061.
Luo, C., Watanabe, T., Nakamura, M., Uemiya, S. and Kojima, T., 2001a,
Gasification kinetics of coal chars carbonized under rapid and slow

heating conditions at elevated temperatures, Energy Resources Technology, 123(1): 2126.


Luo, C., Watanabe, T., Nakamura, M., Uemiya, S. and Kojima, T., 2001b,
Development of FBR measurement of char reactivity to carbon dioxide
at elevated temperatures, Fuel, 80(2): 233 243.
Park, J.Y. and Levenspiel, O., 1975, The crackling core model for the reaction of solid particles, Chem Eng Sci, 30(10): 12071214.
Peng, F.F., Lee, I.C. and Yang, R.Y.K., 1995, Reactivities of in situ and
ex situ coal chars during gasification in steam at 100014008C. Fuel
Processing Technology, 41(3): 233251.
Schmal, M., Monteiro, J.L.F. and Toscani, H., 1983, Gasification of high
ash content coals with steam in a semibatch fluidized bed reactor, Ind
Eng Chem Proc Des Dev, 22(4): 563570.
Sekine, Y., Ishikawa, K., Kikuchi, E. and Matsukata, M., 2005, New
evaluation of method of carbonaceous structure on coal steam gasification, Energy & Fuels, 19(1): 326 327.
Sekine, Y., Ishikawa, K., Kikuchi, E., Matsukata, M. and Akimoto, A.
Reactivity and structural change of coal char during steam gasification,
Fuel, 85(2): 122 126.
Skodras, G. and Sakellaropoulos, G.P., 2002, Mineral matter effects in lignite gasification, Fuel Processing Technology, 7778(6): 151 158.
Szekely, J., Lin, C.I. and Sohn, H.Y., 1979, A changing grain size model
for gas-solid reactions, Chem Eng Sci, 34(1): 10721075.
Szekely, J., Lin, C.I. and Sohn, H.Y., 1980, A changing grain size model
for gas-solid reactions, Chem Eng Sci, 35(7): 12611262.
Tamhankar, S.S., Sears, J.T. and Wen, C.Y., 1982, Rates of coal gasification at high temperatures in CO2 and steam, Processings Joint AIChECIESC Meeting, Vol. II, 596 606.
Wen, C.Y., 1968, Noncatalytic heterogeneous solid-fluid reaction models,
Ind. Eng. Chem., 60(9): 3454.
Xie, K.C., 2002, Coal structure and Its Reactivity (Chinese), 343358
(Science Publishing House, Beijing, China).
Yang, Y. and Watkinson, A.P., 1994, Gasification reactivity of some Western Canadian coals, Fuel, 73(11): 17861791.
Zhou, J., Zhou, Z.J., Gong, X. and Yu, Z.H., 2002, Study of char-CO2
gasification (I) by isothermal thermogravimetry, Coal Conversion
(Chinese), 25(4): 6669.

ACKNOWLEDGEMENTS
The authors thank National Basic Research Program of China for
financial support for this work under Project No.2004CB217704.
The manuscript was received 5 June 2006 and accepted for publication
after revision 30 August 2006.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2006, 84(B6): 420 428

Вам также может понравиться