Fuel 79 (1145 1154)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Fuel 79 (2000) 1145–1154

www.elsevier.com/locate/fuel

Mathematical modeling of coal char reactivity with CO2 at high pressures


and temperatures
Gui-su Liu*, A.G. Tate, G.W. Bryant, T.F. Wall
Cooperative Research Centre for Black Coal Utilization, Department of Chemical Engineering, The University of Newcastle, Callaghan, NSW 2308, Australia
Received 26 August 1999; received in revised form 22 November 1999; accepted 23 November 1999

Abstract
Literature data on the gasification of coal chars with CO2 at moderate temperatures and high pressures has been reviewed, the focus being
the factors affecting the reactivity. A model has been developed to extrapolate this reactivity data to high-temperature conditions. The
intrinsic reactivity for various chars were predicted from the moderate temperature apparent reactivity data, assuming that reaction is in
regime I, and the effects of pressure, char type and temperature on intrinsic reactivity were obtained. The apparent char reactivity was
predicted at high temperature by incorporating the intrinsic reactivity with an effectiveness factor. It is shown that both apparent and intrinsic
reaction rates at 1123 K increase continually with CO2 partial pressure. The char type has a more significant effect on the intrinsic reaction
rate than the pressure. The char surface area is an important factor in obtaining char apparent reaction rate. Gasification temperature has the
greatest influence on reactivity. The large variation in predicted intrinsic reaction rate for various chars at high temperatures is observed due
to different activation energies. The activation energy for char–CO2 gasification generally decreases as coal char rank decreases. For
char–CO2 gasification of 100 mm particles in an entrained flow coal gasification system where the gasification temperature exceeds
1700 K, the char apparent gasification rate is limited by mass transfer in porous structure of char particles. The predicted apparent reactivity
showed a reasonable agreement with experimental measurements that were obtained under high-temperature low-pressure gasification
conditions. The extrapolated high-temperature gasification kinetics can be used in modeling the performance of an entrained flow gasifier.
q 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Gasification; Coal char; Intrinsic reaction rate

1. Introduction The effect of pressure on char–CO2 gasification has been


extensively studied [2–8]. Most of these experiments were
Gasification processes have been widely applied for coal performed at low temperatures, which is the case for fixed
utilization [1]. Both fixed bed and fluidized bed gasifiers bed or fluidized bed gasification. However, there is an
have been commercialized. Recently, gasification under absence of experimental data for high-pressure and high-
elevated pressures and temperatures, for example in an temperature gasification in the literature.
Integrated Gasification Combined Cycle (IGCC) entrained In this work, Langmuir–Hinshelwood expressions, which
flow gasifier, has received greater interest due to the high were derived on the basis of adsorption and desorption
conversion efficiency and low pollutant concentrations mechanisms, have been used to model the char reactivity
obtained from this technology [1]. at pressure [2,3]. A number of Langmuir–Hinshelwood
Understanding the kinetics of the gasification reaction is expressions have been proposed based on CO2 gasification
essential for designing an efficient gasifier. A number of experiments performed under different conditions. These
experiments on the char reactivity at high pressure have expressions can predict the apparent reactivity at low
been carried out for a variety of coal char types using temperatures, typically 900–1300 K. The kinetics of high-
pressurized thermogravimetry, pressurized drop-tube temperature and high-pressure gasification can be modeled
furnaces or pressurized fluidized bed reactors [2–8]. by extrapolating the high-pressure low-temperature data and
considering the effects of temperature. These effects
include: (1) the effect of temperature on intrinsic reactivity;
* Corresponding author. Tel.: 161-2-4921-7442; fax: 161-2-4921-8692. (2) incomplete gas penetration into the char particle; (3)
E-mail address: cggsl@cc.newcastle.edu.au (G.-s. Liu). surface area reduction due to graphitization [9]; and (4)
0016-2361/00/$ - see front matter q 2000 Elsevier Science Ltd. All rights reserved.
PII: S0016-236 1(99)00274-4
1146 G.-s. Liu et al. / Fuel 79 (2000) 1145–1154

Nomenclature
A; B; C; D intermediate parameters used in Eqs. (A2) and (A3)
Ai pre-exponential constant for ki
E activation energy (kJ mol 21)
Ei activation energy for ki (kJ mol 21)
Da macro-pore diffusivity (cm 2 s 21)
DAB binary diffusivity (cm 2 s 21)
De meso-pore diffusivity (cm 2 s 21)
Deff effective diffusivity (cm 2 s 21)
Deff;CO CO effective diffusivity coefficient (cm 2 s 21)
Deff;CO2 CO2 effective diffusivity coefficient (cm 2 s 21)
Di micro-pore diffusivity (cm 2 s 21)
Dk;a ; Dk;e ; Dk;i Knudsen diffusivity of macro-, meso- and micro-pores (cm 2 s 21)
k reaction rate constant in the nth order equation (g cm 22 s 21 atm 2n)
ki (i ˆ 1; 2, 3 and 4) reaction rate constant (i ˆ 1, g cm 22 atm; i ˆ 2; 3, atm 21; i ˆ 4; g cm 22 atm 22)
Ki (i ˆ 1; 2 and 4) transformed reaction rate constant (i ˆ 1; g cm 22 atm; i ˆ 2; atm 21; i ˆ 4; g cm 22 atm 22)
MA ; MB ; MC molecule weight of species A, B and carbon (g mol 21)
n pressure order of nth order equation
PCO CO partial pressure (atm)
PCO;0 CO partial pressure at center (atm)
PCO;S CO partial pressure at surface (atm)
PCO2 CO2 partial pressure (atm)
PCO2 ;0 CO2 partial pressure at center (atm)
PCO2 ;S CO2 partial pressure at surface (atm)
PT total pressure (atm)
r radial distance from particle center (cm)
R gas constant ( ˆ 8.314 × 10 23 kJ mol 21 K 21 ˆ 82.057 atm cm 3 mol 21 K 21)
R0 particle radius (cm)
Ra ; Re ; Ri pore radii of macro-, meso- and micro-pore, respectively (cm)
Rapp apparent reactivity (s 21)
Rin intrinsic reactivity (g cm 22 s 21)
S surface area (m 2 g 21)
S0 initial surface area (m 2 g 21)
T particle temperature (K)
V particle volume (cm 3 g 21)
Greeks
g reaction stoichiometry
1a macro-porosity
1e meso-porosity
1i micro-porosity
1T Total porosity
h effectiveness factor
sAB force constant (Å)
fM modified Thiele modulus
4 intermediate parameter
V AB collision integral

deactivation due to thermal annealing [10]. The last two effects extrapolate the high-pressure low-temperature gasification
contribute to a reduction in reactivity for the burnt char during reactivity data to high-pressure high-temperature con-
the latter stages of char burnoff [10]. These are not discussed as ditions. This study will provide kinetic data for estimating
they are beyond the scope of the current contribution. entrained flow gasifier performance and for use in modeling
The purpose of this paper was to develop a model to of coal gasification.
G.-s. Liu et al. / Fuel 79 (2000) 1145–1154 1147

2. Extrapolation of reactivity data The nth order intrinsic rate equation is expressed as:
Rin ˆ kPnCO2 …9†
The extrapolation of reactivity data from low temperatures
to high temperatures was conducted using the following steps: When compared to the nth order equation, the Langmuir–
1. derivation of the intrinsic reaction rate for various chars Hinshelwood expression has three important features: (a)
from published experiments; the intrinsic reactivity is a nonlinear function of CO2 partial
2. extrapolation of the intrinsic reactivity from low to high pressure, and does not use an uncertain pressure order n; (b)
temperature; it has a mechanistic basis due to the consideration of an
3. calculation of the effective diffusivity and effectiveness adsorption–desorption two-step reaction; and (c) it includes
factor at high temperature; the inhibiting effect of the product gas (CO), which may
4. generation of apparent reactivity at high temperature. become significant at a high CO partial pressure.

Some of the steps were formulated as follows. 2.2. Incomplete gas penetration at high temperatures

2.1. Intrinsic reactivity As gasification temperature exceeds 1700 K, which is the


case for entrained flow gasification, the char–CO2 gasifi-
Most char–CO2 high-pressure gasification experiments cation reaction is limited by gas diffusion to the porous
have been performed for char particles less than 1 mm matrix [14]. Hence, an effectiveness factor h , which is the
over the temperature range of 900–1300 K. Under these ratio of the actual rate per internal surface area to the rate
conditions gasification reactions are controlled chemical attainable if no pore diffusion resistance existed, is often
reactions [3,4]. The intrinsic reaction rate Rin (g cm 22 s 21) used. The apparent reaction rate Rapp is then calculated by:
is therefore defined as the ratio of the apparent reaction rate
Rapp ˆ hRin S …10†
Rapp (s 21) to the internal surface area S (m 2 g 21) of the
particle, as given below: The effectiveness factor h , which primarily depends on
Rin ˆ Rapp =S …1† particle temperature and size, can be calculated by the
well-known Thiele modulus approach [15]. In this approach
Several carbon–CO2 reaction mechanisms have been an nth order equation is used. In this study, however, the
proposed in the literature [5,8,11]. In recent studies, more Langmuir–Hinshelwood rate expression, rather than the nth
modern theories were proposed under atmospheric pressure order one, makes it difficult to calculate the effectiveness
conditions [12,13], and no rate equations were provided. In factor due to its nonlinear differential nature. A numerical
the current study, a high-pressure reaction mechanism was calculation, as given in Appendix A, is therefore presented.
used [8], for which the reaction is composed of the follow- The pore diffusion becomes dominant as the temperature
ing elemental reactions: increases. An effective diffusivity coefficient Deff at high
temperatures is significant for the calculation of the effec-
Cf 1 CO2 $ C…O† 1 CO …2†
tiveness factor, and is strongly dependent on the pore size
within the particle. Conventional methods for calculation of
C…O† ! CO …3† Deff use a uniform pore size that can be determined by
porosity and surface area measurements [15]. However, the
CO 1 Cf $ C…CO† …4† assumption of a single pore size would result in an erroneous
result for the case of char particles with a wide range of pore
CO2 1 C…CO† ! 2CO 1 C…O† …5† sizes, and hence inaccurate reactivity predictions.
Poly-pore size models, such as the bimodal model [16]
CO 1 C…CO† ! CO2 1 2Cf …6† and the poly-modal model [17], have been developed. The
effective diffusivity Deff ; based on the poly-modal model,
where C…O† and C…CO† denote the oxygen complexes on the has been predicted for high-temperature char oxidation
carbon surface. Furthermore, the intrinsic reaction rate process [18]. In the poly-modal model [17] the total porosity
equation was derived in the form of a Langmuir–Hinshel- of the char particle is expressed as a sum of the macro-,
wood expression as shown in Eq. (7) [5,8]: meso- and micro-porosity, which is written as:
k1 PCO2 1 k4 P2CO2 1T ˆ 1a 1 1e 1 1i …11†
Rin ˆ …7†
1 1 k2 PCO2 1 k3 PCO where 1a ; 1e and 1i are denoted as the macro-, meso- and
where k1 ; k2 ; k3 and k4 are the temperature-dependent constants micro-pore porosities, respectively. The effective diffusion
generated from the elemental reactions (2)–(7), which can be rate in the particle is therefore a sum of the contributions
represented in an Arrhenius-type equation, as given below: from the micro-, meso- and macro-pores. For a given initial
micro-, meso- and macro-pore size and porosity, the effec-
ki ˆ Ai e2Ei =RT …8† tive diffusivity can be calculated using following equation,
Table 1 1148
Published experimental investigation on char–CO2 gasification at high pressure

Investigation Char samples Temperature (K) Pressure (MPa) Particle diameter (mm) Surface area (m 2 g 21) Experimental method
a
Nozaki et al. [2] Yallourn sub-bituminous char 1123 0.02–2.5 500–590 282 PDTF b
Nozaki et al. [2] Baiduri sub-bituminous char 1123 0.02–2.5 500–590 278 a PDTF b
Nozaki et al. [2] Taiheiyo sub-bituminous char 1123 0.02–2.5 500–590 358 a PDTF b
Nozaki et al. [2] Hongei anthracite char 1123 0.02–2.5 500–590 115 a PDTF b
Sha et al. [3] Xiago Long Tan Lignite char 1173 0.12–3.1 420–840 682 a PFBR c
Li et al. [4] Lignite Char 1073–1223 1.96 250–420 200 g PPBR d
Blackwood et al. [5] Jarrah purified carbon 1063–1143 4.0 1200–2400 46 RR e
Boyd et al. [6] Drayton bituminous char 1138–1240 0.05–0.75 – 12 PDTF b
Adánez et al. [7] Spanish lignite char 1073–1273 2.5 100–200 211 a PFBR f
Mühlen et al. [8] German bituminuous Char 1173 , 6.0 – 100 g PTGA h
a
Values of initial surface area are calculated from the correlation S0 ˆ 218:4 × …VM=FC† 1 98:4; which is extracted from the data published in the literature.
b
Pressurized drop tube furnace.
c
Pressurized fixed bed reactor.
d
Pressurized packed bed reactor.
e
Pressurized bed.
f
Pressurized fluidized bed reactor.
g
Values of initial surface area are taken from those for the same char type given in the literature.
h
Pressurized thermogravimetrical analysis.

(16).
Intrinsic reactivity, Rin (g.cm-2.s -1) Apparent reactivity, Rapp (s -1)

1.E-11
1.E-10
1.E-09
1.E-08
1.E-05
1.E-04
1.E-03
1.E-02
121a 1e

0
0
G.-s. Liu et al. / Fuel 79 (2000) 1145–1154

1
1
2
1 21e …1 2 1a 2 1e †
1 21a …1 2 1a 2 1e †

2
…1=Da † 1 …1=De †

3
(a)

(b)
2
2
as presented in Appendix B [17]:

4
4
…1=De † 1 …1=Di †
…1=Da † 1 …1=Di †
Deff ˆ 12a Da 1 12e De 1 …1 2 1a 2 1e †2 Di

5
macro-, meso- and micro-pores, respectively.

Total pressure, Ptotal (MPa)

Total pressure, Ptotal (MPa)


6

6
7

7
…12†

[7] lignite char; —, prediction by this work using Eq. (7) and Eqs.(13)–
intrinsic reaction rate (g cm 22 s 21) of CO2 gasification. Particle tempera-

et al. [4] lignite char; O, Boyd et al. [6] bituminous char; × , Adánez et al.
trations were 100%. S, Nozaki et al. [2] sub-bituminous char; A, Nozaki et

wood et al. [5] purified carbon; 1, Mühlen et al. [8] bituminous char; –, Li
Nozaki et al. [2] anthracite char; p , Sha et al. [3] lignite char; W, Black-
tures were adjusted to 1123 K using original rate expressions; CO2 concen-
where Da ; De and Di denote the effective diffusivity for

al. [2] sub-bituminous char; K, Nozaki et al. [2] sub-bituminous char; B,


Fig. 1. Effect of total pressure on: (a) apparent reaction rate (s 21); and (b)
G.-s. Liu et al. / Fuel 79 (2000) 1145–1154 1149

3. Results and discussion Eq. (1) has been used to calculate the intrinsic reaction
rate of chars from coals of various ranks. Fig. 1(b) shows the
3.1. Apparent and intrinsic reaction rate intrinsic reaction rate varying with total pressure at a
corrected gasification temperature of 1123 K. The effect of
Table 1 summarizes the published experimental data for CO2 pressure on intrinsic reaction rate is almost similar to
high-pressure char–CO2 gasification for a variety of coal that on the apparent reaction rate. The effect of coal char
types under various conditions. Fig. 1(a) shows the apparent type, however, is reduced to about one order of magnitude,
reaction rate against total pressure at a CO2 concentration of indicating that the effect of char surface area on the apparent
100%. Particle temperatures were adjusted to 1123 K using reaction rate is significant. Smith [23] calculated the intrin-
the original rate expressions from the literature. It can be sic reactivity for char oxidation for various chars. Large
seen that for most of coal char types the apparent reaction differences were observed in the reaction rates obtained
rate increases as the CO2 partial pressure increases. For for the chars investigated in his study due to the variation
purified carbon and anthracite the reaction rate increases in atomic structure for carbon, and the different levels of
with CO2 pressure up to 2.5–4 MPa. It is suggested that catalytic and inhibiting impurities present in the particle.
the pressure effect on the reaction rate varies with coal The catalytic effect of CaO in low rank coal causes an
rank, and is significant for high rank coal. This conclusion increase in the number of active sites and hence the reac-
is inconsistent with the study by Cope et al. [19] who found tivity [23]. As these chars were obtained from different
that increased pressure produced greater increases in carbon studies, their preparation methods (pyrolysis conditions
conversion as the coal rank decreased. The apparent reac- and apparatus) may also have effects on their intrinsic reac-
tion rate also varies greatly with the coal char type. It gener- tivity [24]. The study of Lizzio et al. [25] has mentioned that
ally increases by two orders of magnitude when the char the CO2 surface area obtained by the physical adsorption
rank decreases from purified carbon to lignite coal char. technique does not adequately normalize the apparent reac-
To calculate the intrinsic reaction rate of char–CO2 tivity, which may also be the reason for the discrepancy in
gasification, particle surface area is necessary, which is intrinsic reactivity obtained in this study. However this is
absent in some original investigations. The initial surface beyond the scope of this study.
area of the char particle is determined by the properties of
the parent coal, and strongly depends on the devolatilization 3.2. Extrapolation of intrinsic reaction rate to high
process [20]. In order to approximate the initial surface areas a temperatures
correlation of the initial surface area with the ratio of volatile
matter to fixed carbon (VM/FC) based on experimental data Fig. 3 shows that the intrinsic reaction rate of char–CO2
[21,22] is presented in Fig. 2. The initial surface area for gasification depends on the temperature. The calculated
different chars was then calculated as indicated in Table 1. intrinsic reactivity from the experimental data is shown as
The chars derived from low-rank coals with high volatile points in the figure. The extrapolated intrinsic reactivity,
matter contents have the highest surface areas. over the temperature range of 1300–2100 K, is expressed

1600

1400
Initial surface area, S 0(m2g-1)

1200

1000

800 S 0 = 218 . 4 × ( FC / V M ) −1 + 98 . 4

600

400

200

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
VM/FC

Fig. 2. Correlation of initial BET surface area of char particles measured by CO2 isotherm with volatile matter (VM) and fixed carbon (FC) contents of coal. K,
Hassimoto et al. [21]; A, Adschiri et al. [22]; and —, correlation.
1150 G.-s. Liu et al. / Fuel 79 (2000) 1145–1154

1.E+00

IGCC
1.E-02 Gasifier

E1=221.2~317.7kJ/mol
1.E-04
E1=140~153.1kJ/mol
Rin (g.cm-2..s -1) 1.E-06

Fixed bed
1.E-08 or fluidized
bed

1.E-10

1.E-12

1.E-14
0.0004 0.0005 0.0006 0.0007 0.0008 0.0009 0.001
-1
1/T (k )
22 21
Fig. 3. Intrinsic reaction rate (g cm s ) of char–CO2 gasification as a function of gasification temperature. Total pressure, 2 MPa; CO2 concentration, 100%.
Points, experimental data; lines, extrapolated trends. K, – - - –, Blackwood et al. [5] purified carbon; A, – – – –, Mühlen et al. [8] bituminous char; S, - - - - - -,
Li et al. [4] lignite char; W, – - – - –, Boyd et al. [6] bituminous char; 1, – – (fine), Adánez et al. [7] lignite char. — (coarse), prediction by this work using Eq.
(7) and Eqs.(13)–(16).

as lines. A typical entrained flow gasification lies within the over the whole temperature range and are provided in Table
two dashed vertical lines where the temperature is in the 2. The activation energies for constants k1 ; k2 ; k3 and k4 in
range of 1700–2000 K. It is shown that the temperature Eq. (7) were provided by the literature (see Table 2). It was
has a pronounced effect on the intrinsic reactivity for all found that the true activation energy E for each char is only
chars. For low rank lignite char, for example, as the slightly greater than E1 ; which demonstrates that the extra-
temperature increases from 1100 up to 2000 K, the intrinsic polated intrinsic reactivity is dominated by the activation
reaction rate increases by approximately four orders of energy E1 : The activation energy E1 is also a function of
magnitude. Also observed in Fig. 3 is the effect of char coal rank, as indicated in Table 2. Among those chars
types. The differences in the intrinsic reaction rate between studied, purified carbon has higher activation energy E1
chars increase from an order of magnitude at a temperature (317.7 kJ mol 21) than lignite char (140.0 kJ mol 21). The
of 1100 K to four orders of magnitude at temperature of large variation for the intrinsic reaction rate observed
2000 K. At the higher temperature, the intrinsic reaction between chars at high temperature is attributed to the
rate is approximately 1:0 × 1022 g cm22 s21 for anthracite various activation energies E of the chars. The reported
char, but is only 1:0 × 1026 g cm22 s21 for lignite char, activation energy E ranges from 170 to 250 kJ mol 21 for
indicating a low reactivity for low rank char. most char–CO2 gasification at low pressures [26–28].
The intrinsic reactivity expressed by the Langmuir– These values are close to those for bituminous and lignite
Hinshelwood type rate equation does not vary linearly chars presented in this study.
with temperature. For quantitative analysis, the trend lines From whole source data, the averaged intrinsic reactivity
are approximately regarded as Arrhenius plots. The true and a set of parameters k1 ; k2 ; k3 and k4 can be obtained,
activation energies for these char reactions were calculated with an activation energy E1 of 212 kJ mol 21. The solid

Table 2
True activation energy E and activation energies Ei of k1 ; k2 ; k3 and k4 (kJ mol 21) shown in Eq. (7) (N/A ˆ not available)

Investigation Char Sample E E1 E2 E3 E4

Blackwood et al. [5] Purified carbon 357 317.7 0 0 359.5


Boyd et al. [6] Bituminous 306.5 221.2 296.5 2227 N/A
Mühlen et al. [8] Bituminous 171 153.1 223 248.1 68.5
Adánez et al. [7] Lignite 169.5 140.0 242 260 N/A
Li et al. [4] Lignite 149.1 N/A N/A N/A N/A
G.-s. Liu et al. / Fuel 79 (2000) 1145–1154 1151

1.E-01

1.E-02
1

Def f (cm2.s -1)


1.E-03

1.E-04

1.E-05

1.E-06
0 1 2 3 4
1000/T(k -1)

Fig. 4. Comparison of the effective diffusivity between the extrapolations from experimental results and the model predictions. K, experimental measurements
[29]; – – –-, extrapolations from experiments [29]; solid line 1, predictions using poly-modal model [17]; and solid line 2, predictions by Valix et al. [29] using
the average pore size model [15].

lines in Figs. 1(b) and 3 indicate the dependence of the at higher temperatures of 1400–2000 K, the Deff predicted
averaged intrinsic reactivity on pressure and temperature, using the poly-modal model shows good agreement with the
respectively. The Arrhenius expressions for k1 ; k2 ; k3 and extrapolated Deff : This implies that at high temperatures, the
k4 in Eq. (7) are: contribution of the meso-pores as well as the macro-pores in
the particles to the effective diffusivity becomes significant.
k1 ˆ 0:45e2212=RT …g cm22 atm21 s21 † …13† The poly-modal pore size structure of particles allows a
better understanding of the gas diffusion in the pore matrix
k2 ˆ 0:021e23=RT …atm21 † …14† at high temperatures.

k3 ˆ 0:038e48:1=RT …atm21 † …15†


3.4. Prediction of apparent reaction rate at high
210 268:5=RT 22 22 21 temperatures
k4 ˆ 2:03 × 10 e …g cm atm s † …16†
Prior to the calculation of the apparent reaction rate (s 21),
3.3. Effective diffusivity the effectiveness factors h at various temperatures were
obtained. The averaged intrinsic parameters discussed
The overall reactivity of char particles at high tempera- above k1 ; k2 ; k3 and k4 were used here. In the calculation
ture can be obtained with a known effective diffusivity. of the effectiveness factor, a 100 mm particle was employed
There has been little information reported on the measured and the CO2 partial pressure was taken as 2 MPa. As a
effective diffusivity for bituminous chars. The experimental result, the calculated effectiveness factor decreases from
effective diffusivity Deff for brown coal char was provided unity to 0.13 as the temperature increases from 1400 to
by Valix et al. [29], as shown by the points in Fig. 4; the 2000 K, indicating that an increase in temperature shifts
measured activation energy of Deff was also given as the reaction from a chemical reaction control to a pore
6.2 kJ mol 21. The extrapolated effective diffusivity was diffusion control. This result is in close agreement with
obtained over a wide range of temperatures using this the investigation by Weeda [14], who reported an effective-
activation energy, as expressed by the broken lines in Fig. ness factor of 0.2 at a temperature of 2000 K with the same
4. The predicted effective diffusivity Deff using the single particle size. In the entrained flow coal gasification system
pore model [15] and the poly-modal model [17] were also where gasification temperature exceeds 1700 K, the char
presented and are shown using the solid lines in Fig. 4. A gasification rate is likely to be limited by mass transfer in
large variation is observed between measured and predicted the porous structure of char particles.
Deff at low temperatures (,700 K) which may be attributed To date there has been limited experimental data for
to pore constrictions [29]. A difference of half an order of char–CO2 gasification at both high temperature and high
magnitude is also shown between the two models through- pressure. To validate the model presented two experimental
out the whole temperature range. At low temperatures of measurements of CO2 gasification at high and moderate
around 800 K, the Deff predicted using the single pore model temperatures but at low pressures were employed. The
[15] is close to the experimentally extrapolated value, while feasibility of these demonstrations is based on two reasons:
1152 G.-s. Liu et al. / Fuel 79 (2000) 1145–1154

Fig. 5. Comparison of the apparent reaction rate of char–CO2 gasification between experimental measurements and model predictions. CO2 partial pressure,
0.1 MPa. K, experimental by Lin et al. [9]; A, experimental by Dutta et al. [30]; —, predictions for Lin et al. [9]; and – – – –, predictions for Dutta et al. [30].

(1) the model presented is suitable over a wide range of reactivity. A large variation in predicted intrinsic reaction
pressures; and (2) temperature and pressure have an rate for various chars at high temperature is observed due
independent influence on the reactivity as indicated by the to different activation energies. The activation energy for
Langmuir–Hinshelwood expression. char–CO2 gasification generally decreases as the coal
Fig. 5 shows that the predicted apparent reactivity is in char rank decreases.
reasonable agreement with experimental measurements for 4. For char–CO2 gasification with 100 mm particles, the
Newlands char [9] and for Synthane char No.122 [30]. The calculated effectiveness factor decreases from unity to
apparent reaction rate of Synthane char is about half an order 0.13 as the gasification temperature is raised from 1400
of magnitude higher than that of Newlands char owing to a to 2000 K, indicating that the reaction transfers from
larger surface area. A reduction in the apparent reaction rate chemical reaction limited to pore diffusion limited. In
occurs at high temperature due to pore diffusion, which is an entrained flow coal gasification system where the
observed for both cases. It is also suggested that the averaged gasification temperature exceeds 1700 K, the char
activation energy E1 taken as 212 kJ mol 21 is appropriate. gasification rate is limited by mass transfer in porous
Further measurements of CO2 gasification reactivity structure of char particles.
under both high-temperature and high-pressure conditions, 5. An averaged activation energy E1 taken as 212 kJ mol 21
and a comparison between model predictions and measure- was found to be appropriate for CO2 gasification of
ments will be presented in subsequent studies. bituminous coal chars. The predicted apparent reactivity
showed a reasonable agreement with the experimental
measurements that were obtained under high temperature
4. Conclusions but low-pressure gasification conditions. This suggests
that the extrapolated high-temperature gasification
The following conclusions can be drawn from this study. kinetics can be used for the modeling of an entrained
flow gasifier.
1. For low rank coal char the apparent reaction rate for
char–CO2 at 1123 K increases as the CO2 partial pressure
increases.
2. The effect of char type is more significant compared to
pressure. The apparent reaction rate at 1123 K increases Acknowledgements
by two orders of magnitude as the coal rank decreases;
while the intrinsic reaction rate varies with coal char rank The authors wish to acknowledge the financial support
within one order of magnitude, suggesting that char provided by the Cooperative Research Centre for Black
surface area is important in obtaining char apparent Coal Utilization, which is funded in part by the Cooperative
reaction rate. Research Centres Program of the Commonwealth Govern-
3. Gasification temperature has the greatest influence on ment of Australia. The first comparison of effective
G.-s. Liu et al. / Fuel 79 (2000) 1145–1154 1153

diffusivity performed by Ms Valix [29] is also Appendix B. Effective diffusivity


acknowledged.
The diffusivity Da ; De and Di for macro-, meso- and
micro-pores are given by the following equations [17,18]:
Appendix A. Effectiveness factor 1 1
Da ˆ ; De ˆ and Di
1=DAB 1 1=Dk;a 1=DAB 1 1=Dk;e
For CO2 gasification reaction C 1 CO2 ! 2CO, the reac-
tion rate is expressed in a Langmuir–Hinshelwood type 12i 1
equation [5,8]: ˆ …B1†
…1 2 1a 2 1e † AB 1 1=Dk;i
2 1=D

k1 PCO2 1 k4 P2CO2 where DAB is the molecular diffusion calculated by


p
Rin ˆ …A1† DAB ˆ …0:000000266T 3=2 =…PT s2AB V AB 1=MA 1 1=MB ††, and
1 1 k2 PCO2 1 k3 PCO
the Knudsen diffusion Dk;a ; Dk;e and Dk;i are strongly depen-
dant on the macro-pore, meso-pore and micro-pore radii,
Similar to the derivation of the effectiveness factor respectively, as indicated in the following forms:
presented by Roberts et al. [31], the internal CO2 partial s s
pressure gradients along the particle radial distance is 2 8RT 2 8RT
obtained as follows: Dk;a ˆ Ra ; Dk;e ˆ Re and
3 pMA 3 pMA
s …B2†
( "
p f 1
dPCO2 M 2
Dk;i ˆ Ri
8RT
ˆ 2 2 B…P2CO2 2 P2CO2 ;0 † 1 C…PCO2
dr R0 A 3 pMA
#)1=2
1 1 AP
CO2
2 PCO2 ;0 † 1 D ln (A2)
1 1 APCO2 ;0 References

[1] Harris DJ, Patterson JH. Aust Inst Energy J 1995;13:22.


The effectiveness factor is then derived as: [2] Nozaki T, Adschiri T, Fujimoto K. Fuel 1992;71:213.
[3] Sha XZ, Chen YG, Cao J, Yang YM, Ren DQ. Fuel 1990;69:293.
p !( "
3 2 1 1 APCO2 ;S
[4] Li S, Sun R. Fuel 1994;73:413.
1
hˆ B…P2CO2 ;S 2 P2CO2 ;0 † [5] Blackwood JD, Ingeme AJ. Aust J Chem 1960;13:194.
fM PCO2 ;S 1 2BP2CO2 ;S A [6] Boyd RK, Benyon P, Nguyen Q, Tran H, Lowe A. Final report,
Australian Coal Association Research Program (ACARP), project
#)1=2
1 1 AP C3095. Australia: Pacific Power, 1998.
CO2 ;S
1 C…PCO2 ;S 2 PCO2 ;0 † 1 D ln [7] Adanez J, Miranda JL, Gavilan JM. Fuel 1985:801.
1 1 APCO2 ;0 [8] Mühlen HJ, van Heek KH, Jüntgen H. Fuel 1985;64:591.
[9] Lin SY, Hirato M, Horio M. Energy Fuels 1994;8:598.
…A3† [10] Hurt R, Sun J-K, Lunden M. Combust Flame 1998;113:181.
[11] Gadsby J, Long FJ, Sleightholm P, Sykes KW. Proc Royal Soc
where: fM represents the modified Thiele modulus and can 1948;193:357.
[12] Kapteijn F, Meijer R, Moulijn JA. Energy Fuels 1992;6:494.
be expressed as [31]
[13] Chen SG, Yang RT, Kapteijn FK, Moulijn JA. Ind Engng Res
1993;32:2835.
!1=2 [14] Weeda M, University of Amsterdam, Amsterdam, 1995.
K1 SRT
fM ˆ R0 …A4† [15] Wheeler A. Advances in catalysis, VIII. New York: Academic Press,
VDeff;CO2 MC 1951. p. 1797.
[16] Smith IW, Tyler RJ. Fuel 1972;51:793.
[17] Wakao N, Smith JM. Ind Engng Chem Fundam 1964;3:133.
and PCO2 ;0 is a partial pressure of CO2 at the particle center, [18] Reade WC, Morris K, Hecker WC. Coal Sci Technol 1995;24:222.
which can be obtained by solving Eq. (A2) using the [19] Cope RF, Smoot LD, hedman PO. Fuel 1989;68:807.
Runga–Kutta numerical method [32]. A, B, C and D in [20] Lee CW, Jenkins RG, Schobert HH. Energy Fuels 1992;6:40.
Eqs. (A2) and (A3) are the constants expressed as A ˆ K2 ; [21] Hashimoto K, Miura K, Ueda T. Fuel 1986;65:1516.
[22] Adschiri T, Shiraha T, Kojima T, Furusawa T. Fuel 1986;65:1688.
B ˆ K4 =2K1 ; C ˆ 1 2 …K4 =K1 K1 † and D ˆ 2……1=K2 † 2 [23] Smith IW. Fuel 1978;57:409.
…K4 =K1 K22 ††; respectively. K1 ; K2 and K4 are the alternated [24] Megaritis A, Messenbock RC, Collot A-G, Zhuo Y, Dugwell DR,
reaction constants, and are defined as K1 ˆ k1 =v; K2 ˆ Kandiyoti R. Fuel 1998;77:1411.
…k2 2 k3 gDeff;CO2 =Deff;CO †=v and K4 ˆ k4 =v; and intermedi- [25] Lizzio AA, Jiang H, Radovic LR. Carbon 1990;28:7.
ate constant 4 is expressed as v ˆ 1 1 k3 …PCO;S 1 [26] Osafune K, Marsh H. Fuel 1988;67:369.
[27] Knight AT, Sergeant GD. Fuel 1982;71:879.
PCO2 ;S gDeff;CO2 =Deff;CO †; where Deff;CO2 and Deff;CO are the [28] Hampartsoumain E, Murdoch PL, Pourkashian M, Trangmar DT,
effective diffusivities of CO2 and CO in the particle, Williams A. Combust Sci Technol 1993;92:105.
respectively. [29] Valix MG, Harris DJ, Smith IW, Trimm DL. Proceedings of the 24th
1154 G.-s. Liu et al. / Fuel 79 (2000) 1145–1154

Symposium (International) on Combustion. The Combustion [31] Roberts GW, Satterfield CN. Ind Engng Chem Fundam 1965;
Institute, 1992 p. 1461. 4:123.
[30] Dutta S, Wen CY, Belt RJ. Ind Engng Chem Process Des Dev [32] Liu G-S, Tate AG, Rezaei HR, Beath AC, Wall TF. Dev Chem Engng
1977;16:20. Miner Process 1999;7:525.

You might also like