You are on page 1of 17

Powder Technology 132 (2003) 167 183

www.elsevier.com/locate/powtec

Operating concept of circulating fluidized bed gasifier


from the kinetic point of view
Huachung Tsui a, Chung-Hsing Wu b,c,*
a

Research and Development Department, Cotech Engineering Corporation, Fl.1, No. 13, Wu Chuan 1st Road, Wuku Industrial Park,
Hsin Chung, Taipei Hsien, Taiwan, ROC
b
Department and Graduate Institute of Bio-Industrial Mechatronics Engineering, National Taiwan University, Taipei, Taiwan, ROC
c
College of Science and Engineering, National United University, Miaoli, Taiwan, ROC
Received in revised form 10 October 2001; accepted 19 February 2003

Abstract
Circulating fluidized bed gasifiers used in integrated gasification and combined cycle (IGCC) possess unique design and operational
features: high gas velocity, high solid loading, and a large amount of circulating chars. As a consequence, their performance is noticeably
different from the conventional fluidized bed gasifier operating at lower gas velocity. At the same time, limited information has been
published to show how to operate the circulating fluidized bed from the kinetic point of view.
In order to study these aspects, kinetic models are developed using kinetic data from previous literature and two concepts of operating the
circulating fluidized bed gasifier from the kinetic point of view are presented in this paper. The results of the kinetic analysis using Johnsons
kinetic data of a constant specific gasification rate of char (0.001 1/s) show that all species approach their equilibrium values after travelling a
distance of 2.5 m, which is set as the height of the gasifier. The resident time of the char in the bed is 0.86 s and the required circulation rate of
char and lime are 11.8 and 4.5 kg/s, respectively. In addition to Johnsons kinetic data, Wens kinetic model is also applied in the kinetic
analysis. It shows a way to operate the gasifier in the fast fluidization regime with a superficial velocity between 3.6 and 4.9 m/s and a solid
loading between 50 and 100 kg/m3, these operating conditions are recommended by industrial experience. It points out the necessity to
predict the performance of the gasifier from the kinetic point of view in order to design the optimal geometry and operating methods that will
allow the gasifier to play a successful role in the clean coal technology system such as IGCC, high-performance power system (HIPPS), etc.
D 2003 Elsevier Science B.V. All rights reserved.
Keywords: Circulating fluidized bed; Gasifier; Kinetic model; Solid loading; Circulating chars

1. Introduction
The world has large coal reserves; in the US, in particular, they constitute 90% of all known fossil energy resources in that country. Although natural gas-fired combined
cycles currently predominate the power generation market,
and interest is increasing in renewable energy sources, coal
will likely maintain its share of the power generation market
simply because it has the largest reserves and no alternative
is foreseen to replace it in the foreseeable future [1].
However, coal usage has been hampered by pollution
caused by its transport, storage, and combustion. Partly for
* Corresponding author. Department and Graduate Institute of BioIndustrial Mechatronics Engineering, National Taiwan University, Taipei,
Taiwan, ROC. Tel./fax: +886-2-23693159.
E-mail address: chwu@ccms.ntu.edu.tw (C.-H. Wu).

solving the transport and storage problem of coal and for


aiming at displacing fuel oil completely, heavily loaded
coal water slurries (f 70% coal by weight) treated with
additives (f 1% by weight) have been developed. Several
demonstration projects have already proven the feasibility of
such coal utilization schemes [2]. To cope with the pollution
problems caused by coal combustion, development programs for clean coal technologies have been adopted
worldwide [3]such as the integrated gasification and
combined cycle (IGCC), the pressurized fluidized bed combustor (PFBC) combined cycle, British Coal topping
cycle in the U.K. [4], the low emission boiler system
(LEBS) and the high-performance power system (HIPPS)
by the U.S. Department of Energy [5], etc. One of the most
important chemical processes in the clean coal technologies
is coal gasification converting solid coal into gaseous fuel
used to power turbine and subsequently to power generator.

0032-5910/03/$ - see front matter D 2003 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0032-5910(03)00060-3

168

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

An advanced gasifier is required in IGCC to accomplish such


work. Conventional gasifiers that operate in either a bubbling
or entrained mode are not suitable choices because they do
not provide efficient solid gas mixing. A novel gasifier
consisting of a circulating fluidized bed and a cyclone
combustor, shown schematically in Fig. 1, as developed by
Foster Wheeler [6], is one of the most promising choices
because it provides good contact efficiency between gas and
solids and reduces the size of gasifier.
The air enters the circulating bed at velocities high enough
for the fluidized bed to operate in the fast or circulating
regime. At these velocities, solid entrainment increases significantly, and circulation of solids becomes necessary to
preserve a constant solid inventory in the bed. Thus, a cyclone
is introduced to collect solids from the flue gas and recycle
them into the circulating fluidized bed. The cyclone is also
used as a combustor to supply energy for the endothermic
gasification reactions. Therefore, the cyclone is referred to as
a cyclone combustor. When the flue gas enters the cyclone
combustor tangentially, large solid particles in the flue gas are

forced to move towards the wall by centrifugal force and are


collected for recycling, whereas small solid particles are
carried over by the flue gas. Another advantage of tangential
injection is that the injected air stays close to the cylinder wall
and most of the oxygen in the air can be expected to react with
the char instead of the flue gas that is relative far from the
cylinder wall. The energy released in the cyclone combustor
will be carried by the circulating solids to the circulating
fluidized bed to provide the required heat for the endothermic
chemical reactions of coal gasification.
A three-level approach, which involves thermodynamic,
kinetic, and hydrodynamic models, was used to analyze the
operation of the gasifier shown in Fig. 1 [7]. This paper only
presents the kinetic analysis. The feed rates of materials, the
generation rate of product gas, and the circulation rate of char
used in the kinetic analysis come from the thermodynamic
analysis by balancing the mass and energy of the gasifier at
equilibrium. The kinetic analysis will predict how the
chemical reactions approach equilibrium in the gas and
provides an estimate for the required circulating rate of char.

Fig. 1. Sketch of a novel gasifier consisting of a circulating fluidized bed and a cyclone combustor.

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

Coal water slurry fuel is used to show its feasibility in this


coal utilization scheme. Gasification of coal water slurry fuel
consists of three steps: evaporation, devolatilization, and the
following char gasification. As the rate of evaporation and
devolatiliaztion are fast compared to gasification, the gasification rate of char is considered as the factor that dominates the gasification time of coal water slurry fuel in this
kinetic analysis [8].

2. Problem formulation

Table 1
Reaction mechanisms and rate expressions of char steam reaction (source:
Lowry [15])
Source

Reaction
mechanism

Walker
et al.

1
Cf H2 Og K!


Rate expression

K2

Rate

H2 g CO

K1 H2 O
1 K 1 H2 O K 2 H2
K3
K3

3
CO K!
COg

Ergun and
Menster

1
Cf H2 Og K!


K2

Rate

H2 g CO

K1 H2 O Kd CO2
K1
K2
K5
1 H2 O H2 CO
K3
K3
K4

K3

2.1. Kinetics of coal gasification


Numerous experiments have been performed to study the
kinetics of coal gasification [9 12]. Kinetic information has
been obtained from either the amount of coal converted or
weight of gaseous products formed during gasification.
Wide variations exist in the kinetic data reported in the coal
gasification literature. The variation appears to be caused by
differences in chemical and physical properties of the coal
and the experimental systems used. The chemical properties
of coal, such as carbon, hydrogen, and oxygen contents,
have significant effects on carbon conversion during gasification. As carbon content increases or as hydrogen and
oxygen present in coal decrease, the rank of the coal
increases. In general, the reactivity of coal decreases as
the carbon content increases or the oxygen and hydrogen
contents decrease. In other words, lower rank coals are more
reactive than higher rank coals [12].
The physical properties of coal, such as particle size and
porosity, have significant effects on the kinetics of coal
gasification. As the particle size becomes smaller the specific contact area between the coal and the reacting gases
increases, resulting in faster reaction. In general, coal gasification processes with entrained and fluidized bed reactors
use coal with an average size of 75 Am. Processes with fixed
bed reactors use larger size coal with diameter exceeding
300 Am. Porosity of coal is related to the pore volume and
surface area present in the coal. It has been found [12] that
reactivity in gasification increases with an increase in pore
volume and surface area of low and medium rank coals.
However, reactivity is not affected by changes in pore
volume and surface area in high rank coals with carbon
content larger than 85%. This is due to the smaller pores
present in higher rank coals. As pores become smaller, the
access of reacting gases to the internal surface of the coal is
reduced, i.e. the process is controlled by the rate of diffusion
into the coal.
In experimental systems, coal is pyrolyzed to become
char, which undergoes conversion to gas by reaction with
steam, carbon dioxide, and hydrogen. The kinetics and
mechanisms of reactions of coal char with steam, carbon
dioxide, and hydrogen were discussed in the previous literature [13,14]. In general, Langmuir-type adsorption equations have been used to represent gasification rates [9]. Table

169

CO ! COg
4
CO CO K!


K5

CO2 g Cf
Wen

C + H2O(g) !
CO + H2



CH2 CCORT
1  x where:
CH2 O 
K
Kv = volumetric rate constant;
K = equilibrium constant

dx
dt kv

1 contains some of the rate expressions and mechanisms


developed for the char steam reaction [15]. Carbon monoxide and hydrogen are the principal products of a char
steam reaction and their concentration may have a retarding
effect on the rate.
The rate of char steam reaction depends mainly on the
temperature and steam pressure. Fig. 2 shows the variation
of char gasification rates with temperature at 1 atm steam
pressure [10]. Although there are wide variations in gasification rates of different chars, they all increase exponentially with temperature. The variations in rates are due to
difference in chemical and physical properties of the chars.
Fig. 3 shows the variation of char gasification rate with
steam pressure at different temperatures [12]. The rates
increase with steam pressure but become independent of
pressure at pressures higher than about 10 atm.
2.2. Desulfurization during coal gasification
In developing coal gasification processes, the removal of
the sulfur compounds has been of great concern because of
their detrimental effects on air quality and their role in
corrosion of equipment. The principal sulfur compound that
evolves during gasification of coal is H2S with lesser
portions of COS, CS2. Hot gas desulfurization proceeds in
two steps when calcium-based sorbent is used. First, the
limestone (mostly CaCO3) is calcined to produce lime
(CaO) and CO2 as shown in Eq. (1). In the second step,
lime reacts with H2S to produce a stable sulfide CaS as
shown in Eq. (2).
CaCO3s ! CaOs CO2

CaOs H2 S ! CaSs H2 O:

170

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

between CaO and H2S depends on the concentration of H2S


(C) and the concentration of available CaO (A) as follows:


dN
K3 AC;
dt

where K3 is the rate constant of Eq. (2). The concentration of


available lime in the calcined limestone decreases as the
reaction of desulfurization proceeds in two ways: the available lime reacts with H2S (K3AC) or the surface of the
available lime is covered by CaS layer. The decrease rate of
available lime through the adsorption and desorption of CaS
layer equals the subtraction of the adsorption rate CaS layer
on the exterior of available lime (K1AC) to the desorption
rate of CaS layer on the exterior of inactivated lime (K2B).
Thus, the decrease rate of available lime (  dA/dt) could be
written as follows:


Fig. 2. Variation of char gasification rates with temperature (source:


Johnson [10]).

According to Abel et al. [16], the rate of Eq. (2) is fast, but
the rate of Eq. (1) is relatively slow. In addition, CaCO3 is
chemically inert to H2S. One potential reason for limestone
not being able to remove H2S effectively is the relatively slow
calcinations, Eq. (1), that occurs under gasification conditions. This problem can be solved in the proposed reactor
configuration. A large amount of CaO is recycled in the
gasifier, making the removal of H2S from the gas stream
relatively rapid. According to Schrodt and Hahnn [17], CaS
formed on the exterior of the sorbent prohibits further
reactions of the sorbent with H2S. Thus, CaS appears to form
a diffusion barrier for H2S penetration and this limits the
calcium utilization efficiency. This effect is taken into
account in modeling the process of desulfurization as follows.
A theoretical model for the reaction of CaO with H2S was
developed and experimentally verified by Pell et al. [18].
This model assumes that there is nothing adsorbed on the
surface of CaO at the beginning of reaction of desulfurization. As reaction (2) proceeds, the rate slows down, and this
suggests that CaS is adsorbed upon the surface thereby
inactivating the surface. The total amount of lime (CaO) in
the calcined limestone could be categorized into two groups,
according to Pell et al. [18]. One group is the available lime
that can react with H2S to form CaS through reaction (2). The
other group is inactivated lime that cannot react with H2S
due to the adsorption of CaS layer on the surface of lime. The
concentrations of the total amount of lime, available lime,
and inactivated lime in the calcined limestone are specified
as N, A, and B, respectively. The rate of reaction (dN/dt)

dA
K1 AC  K2 B K3 AC;
dt

where K1 is the rate constant of adsorption of CaS on the


surface of available lime and K2 is the rate constant of
desorption of CaS on the surface of inactivated lime. Substituting the expression A from Eq. (3) into Eq. (4) and
rearranging, Eq. (5) is obtained. From Eq. (5), the rate of
decrease of CaO(s) (dN/dt) is obtained as shown in Eq. (6),
where Ke = K1/K2. The decrease rate of CaO(s) (dN/dt) equals
the desulfurization rate because 1 mol of CaO(s) reacts with 1
mol of H2S, as indicated in Eq. (2). The kinetic constants K1,
K2, and K3 are determined from experimental results via
Arrhenius plots [19]. The pre-exponential constants and
activation energy for the Arrhenius law are described in
Table 2. The general Arrhenius expression is given in Eq.
(7). Form Table 2 and Eq. (7), the rate constants K1, K2, and
K3 can be determined. Substituting these rate constants into

Fig. 3. Variation of char gasification rates with pressure (source: Quader


[12]).

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183


Table 2
Coefficients of Arrhenius expressions for rate constants (source: Pell et al.
[18])
Constants
(Ki)

Preexponential
constants ( Pi)

Activation energy
(Ei, kcal/gmol K)

K1
K2
K3
Ke

8.3
9.5
13.1
 1.2

13.1
28.2
22.9
 15.1

Eq. (6), the desulfurization rate (dN/dt) can be obtained if the


concentration of H2S (C) and the amount of unconverted
CaO(s) (N) are known.
d2 N
dN
K2 K3 CN 0
K1 C K3 C K2
dt 2
dt
dN
K3 N C


dt
1 Ke C

Ei
Ki Pi exp
:
RT

The kinetic model of char gasification is derived from the


mechanism of the chemical reactions among the char, water
vapor, and CO2. According to Johnson [10] and Ergun [11],
the mechanism of the reaction between the char and water
vapor in the presence of CO2 can be written as Eqs. (8)
(10). They assume that there are many free sites Cf on the
char, which can adsorb oxygen from the dissociative chemisorption of CO2(g) and steam as shown in Eqs. (8) and (9),
respectively. As this happens, the free site is energized and
becomes an occupied site (C(o)), which in turn is gasified to
produce CO(g), as shown in Eq. (10). The energized site C(o)
is a chemical complex on the surface of the char which can
be transferred into reaction product CO(g) without receiving
additional energy. However, the process of desorption of the
chemical complex C(o) (reaction (10)) is slower than Eqs.
(8) and (9), therefore the rate of desorption controls the
reaction rate.
k1
CO2g Cf !
 COg Co
kV

k2

A negative sign is added to X in the third term of Eq. (11)


to represent the generation of energized sites C(o) because
the decomposition of CO2 (  X) will implicate the generation of C(o) through Eq. (8).
In this reaction system, the total number of oxygen atoms
is constant, being either present as energized sites C(o) or
contained in the gas species CO, CO2, and H2O. According
to Lindermanns [20] theory, the number of energized sites
C(o) reaches a stationary value rapidly, implying that the
oxygen concentration in C(o) complexes will reach a stationary number. Thus, the number of oxygen atoms contained in the gas species (H2O, CO2, CO) will also reach a
constant value because the total number of oxygen atoms in
the chemical system is a constant. The statement indicating
a constant number of oxygen atoms in the gas species is
represented by Eq. (12), where Y is the total number of
atoms of CO generated through either Eq. (8) or Eq. (10)
during the time interval Dt.
Z disappearance of oxygen through the generation of H2
2X appearance of oxygen through the generation of CO2
Y appearance of oxygen through the generation of CO:
12

k3

Z generation of Co through Eq: 9


X Generation of Co through Eq: 8 0: 11

2.3. Kinetic model of char gasification

Co ! COg :

(10) is slow compared with the generation of C(o) by


collisions via Eqs. (8) and (9). The kinetic analysis utilizes
Lindemanns theory in the way that the gasification process
proceeds step by step and the number of C(o) complexes
reaches a stationary value after each time step. To reach a
stationary value of C(o), the number of energized sites C(o)
desorbed through Eq. (10) should equal the number of the
energized sites C(o) generated through Eqs. (8) and (9) in a
small time interval Dt. Thus, the statement of Lindemenns
theory
can be written as Eq. (11), where negative values of
_
N c (atoms/s) represent the number of C(o) complexes
desorbed through Eq. (10), and positive values of X and
Z represent the number of atoms of CO2 and H2 generated
through Eqs. (8) and (9) in the time interval Dt, respectively.

Nc Dt decomposition of Co through Eq: 10

H2 Og Cf !
 H2g Co
kV

171

10

As Eq. (10) is the rate-controlling step, the gasification


process can be represented as a first-order reaction, i.e. the
reaction rate is proportional to the concentration of C(o).
According to Lindemanns theory [20], for a first-order
reaction, the number of energized molecules C(o) will reach
a stationary value if the conversion of C(o) into CO via Eq.

_
There are four unknowns, N c, X, Y, and Z, in_Eqs. (11)
and (12). Even if the rate of char gasification (N c) can be
obtained from experiments, there are still three unknowns
X, Y, and Zin Eqs. (11) and (12). In order to solve the
problem, another governing equation was applied as follows. The equilibrium constants, K1 and K2, of Eqs. (8) and
(9), respectively, can be written as Eqs. (13) and (14).
Dividing Eq. (14) by Eq. (13) and simplifying it will result
to Eq. (15). It is interesting to note that Eq. (15) represents
the equilibrium condition for the water gas shift Eq. (16).

172

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

In the next section, the concentration of H2O, H2, CO and


CO2 in Eq. (15) are related to X, Y, and Z in the smallcompartment scheme. Thus, (Eqs. (11), (12) and (15) are the
governing equations, which can be solved for X, Y, and Z
when Nc is known and a small-compartment scheme is used
in the next section.
CO Co
k1

K1
CO2 Cf
k1V


Wc Wc0 Nc Dt

17

0
Y
WCO WCO

18

WH2 WH0 2 Z

19

WH2 O WH0 2 O  Z

20

13

H2 Co
k2

k2V H2 O Cf

14

K2
CO2 H2

K1 CO H2 O

15

0
WCO2 WCO
X
2

21

16


0
WCaO WCaO
NH2 S Dt

22


WH2 S WH0 2 S NH2 S Dt:

23

K2

velocity are mainly due to the chemical reaction between


char and steam.

CO H2 O CO2 H2 :
2.4. One-dimensional control volume analysis

A one-dimensional control volume analysis was developed to determine the extent to which the chemical reactions
of char gasification approach equilibrium in the region from
X1 = 0 to X1 = 2.5 m (the gasifier height) as shown in Fig. 1.
The gasification of CWSF is assumed to occur in three
consecutive steps: evaporation, combustion, and slow reaction between char and steam. The processes of water
evaporation, coal pyrolysis, and combustion complete at
X1 = 0. This assumption may not fit reality because the
evolution rate of the volatiles, which depends on various
factors such as particle size, pressure, and heating rate, may
be too slow to assume that coal pyrolysis has been completed
at X1 = 0. A complete pyrolysis model would describe the
composition and physical state of the char particle and the
composition and evolution rate of the volatiles at all stages of
pyrolysis. However, the detailed mechanism of the evolution
and combustion of volatiles is not very well understood and
detailed experimental work is still required. Nevertheless,
this assumption is still a good approximation because in
general the reaction rates of water evaporation, coal pyrolysis, and combustion are much faster than that of the
gasification of the char. Based on this assumption, the initial
composition and superficial velocity of the gas mixture can
be obtained from the thermodynamic analysis.
The variation of the composition and the superficial gas
velocity as X1 increases can be determined using Eqs. (17)
(23). The variation of the number of mole of species (based
on 1 mol of carbon in feed CWSF) as X1 increases can be
calculated from these equations. This information is then
utilized to obtain the composition of the gas mixture (mole
fraction) and the superficial gas velocity. The variation in
the gas composition and the increase in the superficial gas

To conduct the one-dimensional control


volume analy_
sis, the five unknowns X, Y, Z, and N H2S appearing in the
above equations should be determined. The first three
unknownsX, Y, Zare obtained using the kinetic model
developed
_
_ in the previous section. The last two unknowns,
N c and N H2S, are determined by using experimental data in
the literature [8 10]. Eqs. (11) (13) derived in the previous section can be rearranged to Eqs. (14) and (15),
respectively. Eq. (24) is obtained by substituting Eqs.
(18) (21) into Eq. (13).
K

0
WCO
X WH0 2 Z
K2
2

0
K1 WCO
Y WH0 2 O Z

24


Z X  Nc Dt

25


Y X  Nc Dt:

26

Substituting Eqs. (25) and (26) into Eq. (24), results to Eq.
(27), which contains only X.
aX 2 bX c 0

27

where
a=1K
_
0
0
b = KW H
KW CO
+ W 0CO2_+ W0H2 N cDt _
_
_
2O
0
0
0
0
0
0
N cDt + K(N cDt)2+ W CO
W CO
N CDt.
c = KW H2OWH2O +KW H2O N cDt  KWCO
2
2

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

The solution of Eq. (27) is written as Eq. (28)


X

b

p
b2  4ac
:
2a

2.5. Circulation rate and shrinkage of chars


28

Substituting the value of X into Eqs. (15) and (16), the


values of Z and Y are obtained.
_
_
The rates of gasification N c and desulfurization N H2S are
determined from experiments conducted by Wen [9], Johnson [10], and Pell et al. [18]. Thus, two sets of kinetic data
were used to indicate the influence of the rate of gasification
on the performance of the circulating fluidized bed gasifier:
one set from Johnson [10] as described in Fig. 2 and the
other from Wen [9] as described in Eq. (29). Johnsons [10]
experiments indicate that the rate of the char steam reaction
depends mainly upon temperature and steam pressure for a
specific kind of char. Fig. 2 contains specific gasification
rates for different chars in steam at 1 atm. It shows that the
gasification rate increases as the temperature increases.
Wens [9] experiments indicate that the rate of char gasification also depends on the concentration of H2O, H2, and
CO. This is shown in his kinetic model, Eq. (29), which
includes the concentrations of these species. In the equation,
KC H2O is the equilibrium constant of char stream reaction
and Kv is the rate constant for the gasification of coal char in
the temperature range from 1273 to 1473 K.


CH2 CCO RT

Wc :
Nc Kv CH2 O 
KCH2 O

29

_
The rate of desulfurization N H2S is obtained from the
kinetic model of Pell et al. [18], as indicated by Eq. (30).
The rate constants K3 and equilibrium constant Ke are
obtained from experiments [19] of the reaction between
CaO and H2S from 700 to 1200 K
 K3 WCaO CH2 S :
N
H2 S
1 Ke CH2 S

173

30

The kinetic model of desulfurization described by Eq. (30)


does not consider the effect of the water concentration on
the rate of desulfurization. The effect of water concentration
on the rate of desulfurization is still not well understood.
However, water concentration should restrict the extent of
desulfurization from the thermochemical_point
_ of view.
Substituting the values of X, Y, Z, N c, N H2S into Eqs.
(17) (23), the composition of the product gas can be
determined after a time interval Dt or a distance DX1 = UgDt.
The final composition of the product gas at X1 = 2.5 m is
influenced_ by the circulation rate of the char because the
value of N C is related to the circulation rate of the char.
Using a computer simulation, the required circulation rate of
the char for the gasification reactions to reach equilibrium at
X1 = 2.5 m was determined.

The circulation rate of char is calculated in the following


two steps. First, the cut-off diameter of the cyclone combustor is calculated according to the cyclone geometry and
the Rankine vortex flow condition [21]. Second, the transient mass equations of char particles are developed. A trial
and error method is applied to solve these equations by
varying the initial diameter of char particle and the number
of circulation of char particle in the gasifier until the
diameter of char particle shrinks to the cut-off diameter
and the circulation rate of char is accumulated to the
required amount. In the analysis using Johnsons kinetic
data, the circulation rate of char reaches such amount that
the chemical reaction of char gasification reach equilibrium
at X1 = 2.5 m. In the analysis using Wens kinetic model, the
circulation rate of char reaches such amount that the fluidized bed has a superficial velocity between 3.6 and 4.9 m/s
(12 16 ft/s) and a solid loading density between 50 and 100
kg/m3, which are recommended for the bed to operate in the
circulating region by industrial experience [22]. The following assumptions were made in conducting the analysis.
1. The char is a spherical solid particle with a constant
density.
2. The shrinkage of the char is mainly due to the chemical
reactions between the char and the gas species (steam and
oxygen), i.e. the effects of attrition and agglomeration on
the size of the particles have been neglected.
3. A Rankine vortex [23] is formed in the cyclone
combustor, which is a combination of a central forced
vortex and a peripheral vortex.
4. The cut-off diameter of the cyclone combustor can be
determined by considering the motion of the particle on a
imaginary cylindrical surface as described by Ogawa [23]
and Barth and Trunz [24].
5. All particles larger than the cut-off diameter are separated
from the product gas and all particles smaller than the
cut-off diameter are carried out of the cyclone combustor
by the product gas.
2.5.1. Calculation of cut-off diameter of the cyclone
combustor
The separation of the particles and the gas in the cyclone
combustor is caused mainly by the centrifugal force
imposed on the solid particles. The equation of motion of
a solid particle in the cyclone combustor in the radial
direction can be written as Eq. (31)
Mp

dUr
1
Mp Uh2  FD :
r
dt

31

The term on the left-hand side is the inertial force. The


first term on the right-hand side is the centrifugal force,
and the second term on the right-hand side is the drag
force that is caused by the gas flow in the negative radial

174

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

direction. If the centrifugal force is larger then the drag


force, the solid particle will be accelerated in the radial
direction and be forced to move towards the wall of the
cyclone combustor. The drag force FD in Eq. (31) can be
written as Eq. (32).
FD Cd Ap

q
Ur  Vr 2 :
2

32

The expression for the drag coefficient Cd depends on the


type of flow. According to Ogawa [23], the drag coefficient Cd in Stokes flow can be written as Eq. (33). The
flow is considered as Stokes flow when the Reynolds
number based on the diameter of the particle Rep is less
than 4
Cd

24
:
Rep

33

Substituting Eqs. (32) and (33) into Eq. (31) results to Eq.
(34), which is the equation of motion in the radial
direction in Stokes flow.
Mp

dUr
1
Mp Uh2  3pldp Ur  Vr :
r
dt

34

The cut-off diameter of the cyclone combustor can be


determined by considering the motion of the particle on an
imaginary cylindrical surface as described by Ogawa [23]
and Barth and Trunz [24]. The imaginary cylindrical surface
has a diameter of De and a height of H, as shown in Fig. 4.
The cut-off diameter is the diameter of those particles that
can stay on the imaginary surface in a state of mechanical
balance, i.e. the radial velocity is zero (Ur = 0) and the
centrifugal force equals the drag force. When the particles
has diameters larger than the cut-off diameter, the centrifugal force is larger than the drag force because the centrifugal
force is proportional to the mass of the particle or dp3 but the
drag force is proportional to dp, as indicated by the last term
of Eq. (34). As a result, particles larger than the cut-off
diameter cannot pass through the imaginary cylindrical
surface. For the same reason, when the particles have
diameters smaller than the cut-off diameter, the centrifugal
force is smaller than the drag force. As a result, the particles
smaller than the cut-off diameter pass through the imaginary
cylindrical surface and are carried out of the cyclone
combustor by the flue gas.
The expression for the cut-off diameter is obtained by
considering the equation of motion of particles that stay on
the imaginary cylindrical surface and in a state of mechanical balance under the following conditions:
1. Ur2 = 0.
2. the centrifugal force equals the drag force.
3. the angular velocity of the particle equals that of the gas
(Uh = Vh).

Thus, Eq. (34) can be written as Eq. (35)



Vr2


qp
d2 V 2 :
9l De p h2

35

The velocity Vr2 in Eq. (35) can be obtained by assuming


that the inward radial velocity of gas is uniformly distributed
over the imaginary cylindrical surface. Thus,
_ Vr2 is obtained
by dividing the volume flow rate of gas Q by the area of the
imaginary cylinder surface
Vr2


Q
:
pDe H

36

Substituting Eq. (36) into Eq. (35) results to Eq. (37), which
is the expression for the cut-off diameter

dpc

1
Vh2


9Ql
pqp H

!1=2
:

37

The cut-off diameter dpc of the cyclone combustor with


the geometry indicated in Figs. 1 and 4 is calculated as
follows. The calculation is based on the following specified
conditions:
(1) The gasifier operates at 1150 K with a thermal capacity
of 586 KW and a CWSF containing 30% water by
weight.
(2) The gasifier is fed with rates of CWSF, air, and
limestone which were determined in the thermodynamic analysis.
(3) The product gas reaches chemical equilibrium on the
surface of the imaginary cylindrical surface.
(4) The feed air to the cyclone combustor is uniformly
distributed through 12 pipes of 2.54 cm (1 in.) diameter
at 298 K.
(5) The tangential velocity on the imaginary cylindrical
surface falls in the region of the free vortex of the
Rankine vortex.
(6) The mean density qp and the mean diameter of the char
are assumed to be 1200 kg/m3 and 50 Am, respectively.
Condition (5) has generally been found to be true from
experimental results [21]. The density and the diameter of
the char vary over a wide range depending on the physical
and chemical properties of parent coal and the coal pyrolysis
conditions. Thus, condition (6) is an approximation used to
facilitate the calculation.
Based on (1), (2) and (3), the mass flow rate and the
volume flow
rate of product gas through the imaginary
_
surface Q are calculated to be 0.103 kg/s and 0.43 m3/s,
respectively. The height of the imaginary cylinder H_ is 1.54
m, as indicated in Fig. 4. Substituting the values of Q , l, qp,

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

175

Fig. 4. The imaginary cylindrical surface in the cyclone combustor.

and H into Eq. (37), the relation between dpc and Vh2 is
written as Eq. (38).
 
1
4
dpc 1:816 10
:
38
Vh2

is calculated to be Vh = 6.13 m/s based on condition (4). Fig.


5 shows a sketch of Rankine vortex [25] of a combination of
a central force vortex and a peripheral free vortex. The
relation between the tangential velocity Vu and the radius r
is given in Eq. (39).

The value of Vh2 can be calculated under conditions (4) and


(5). From the thermodynamic analysis, 0.0435 kg/s of air
was required to be fed into the cyclone combustor. Thus,
tangential velocity of the air entering the cyclone combustor

Vh rh constant:

39

For the forced vortex, h equals  1 and for the free vortex h
is found between 0.45 and 0.8 [25]. Based on condition (5),

176

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

Fig. 6. Mole ratio of gas species as a function of axial distance along the
gasifier.
Fig. 5. Sketch of the Rankine vortex [25].

the tangential velocity Vu2 can be calculated from Eq. (39)


with h = 0.5 as follows:
Vh2


DWi

 0:5
 0:5
r
9
Vh
6:13
9:19 m=s:
r2
4

From Fig. 1, the values of r and r2 are obtained to be


0.2285 m (9 in.) and 0.1015 m (4 in.), respectively.
Finally, the cut-off diameter of the cyclone combustor is
calculated to be 19.7 Am by substituting Vh2 = 9.19 m/s
into Eq. (38).
2.5.2. Calculation of the char diameter dpn and circulation
rate of char Wn after n circulations
The shrinkage of a single char in each circulation is due
to both gasification and combustion. From previous studies
[26], it has been found that the rate of consumption of char
is the product of its surface area and the rate coefficient Ks
(kg/s m2). The rate coefficient Ks (kg/s m2) for gasification
is estimated from Eq. (40) as follows:
cMp Ks As

As a result, the total mass consumption of a single


particle char in the ith circulation DWi (kg/s) can be written
as Eq. (41).

0:331 0:1

1
DWgi 3:194 104 dp2i1 : 41
0:9
0:9

A shrinking particle model is used to estimate the


shrinkage of the char particle after n circulations. The char
is assumed to have a constant density and to shrink
uniformly when it reacts with either steam or oxygen. The
diameter of the char dpn after n circulations can be written as
Eq. (42).
6
dpn
pqp
6

pqp

W0 

n
X

! 13
DWi

i1
n
X
p
qp dp30  3:194 104
dp2i1
6
i1

! 13
42

Under the assumption that the char particle has a constant


density qp and a uniform diameter dp0 when it enters the

40

The specific rate of gasification c at 1150 k for a bituminous


char can be obtained as 0.001 (1/s) from Fig. 2 [10]. The
mean density qp and the mean diameter of char dp were
measured as 1200 kg/m3 and 50 Am, respectively. Substituting these values of c, dp, and qp into Eq. (40). Ks is
calculated to be 2.0 10 5 (kg/s m2). The required residence time of the char in the gasifier is found to be 0.86 s.
Thus, the mass consumption of a single char particle due to
the gasification in the ith circulation is DWgi = 0.00086Ks
(4pd 2pi  1). The energy balance of this system requires that it
has to burn 0.331 mol of char in the cyclone combustor and
0.1 mol of char in the circulating fluidized bed in order to
sustain 0.9 mol of endothermic steam coal gasification
process.

Fig. 7. Mole ratio of gas species as a function of axial distance along the
gasifier.

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

177

gasifier, the_ number of char particles fed into the gasifier per
unit time N is obtained by dividing the feeding rate of char
by the mass of a single char particle

Wf Fc

:
43
Np
dp30 qp
6
The circulation rate of char after n circulations can be
written as Eq. (44)


Wn N W0 W0  DW1 W0  DW1  DW2
W0  DW1  DW2  DW3 K W0  DW1
 DW2  DW3  K  DWn
!
n
X

N n 1W0 
n  i 1DWi

Fig. 9. Composition of product gas species as a function of axial distance


along the gasifier.

i1


wf Fc
p
n 1W0  3:194
dp30 qp
6
!
n
X
4
2
10
n  i 1dpi1 :

44

(44) by varying the initial diameter of char dp0 and the


number of circulation of char in the gasifier n until the
system reaches steady or both Eqs. (45) and (46) are
satisfied.

i1

Eqs. (42) and (44) are the transient mass equations of char
particles circulating in the fluidized bed and cyclone combustor. The system will reach steady state when the initial
diameter of char shrinks to become the cut-off diameter of
the cyclone combustor after n circulations

3. Results and discussion


3.1. Johnsons kinetic data

As the system reaches steady state the particle size distribution and the circulating rate of char will not vary with
time. A trial-and-error method is used to solve Eqs. (42) and

The results of the kinetic analysis using Johnsons


kinetic data of a constant specific gasification rate (0.01
1/s) are presented in Figs. 6 11. The constant specific
gasification rate of 0.001 1/s was obtained from Fig. 2 at
1150 K. In the kinetic analysis, the reactor temperature, the
thermal capacity of the gasifier, and the water concentration of the CWSF, were fixed at 1150 K, 586 KW (2 106
Btu/h) and 30%, respectively. The kinetic models derived
in previous section were used to predict the required
circulation rates of char and lime for the chemical reactions to reach equilibrium within the given height of the

Fig. 8. Mole ratio of gas species as a function of axial distance along the
gasifier.

Fig. 10. Gas velocity as a function of axial distance along the gasifier.

dpn dpc

45

and the circulating rate of chars becomes the amount


required of operating the gasifier simultaneously
W
n Wre :

46

178

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

Fig. 11. Solid loading density as a function of axial distance along the
gasifier.

gasifier. It was found that the required circulation rates of


carbon and lime were 11.8 and 4.5 kg/s, respectively, and
the required residence time of the char in the gasifier was

0.86 s. Figs. 6 8 show the approach of the chemical


reactions to equilibrium
from the curve of the mole ratio of
_ _
)
vs.
the
axial distance along the gasifier,
gas species
(X
/X
eq
_
where X eq is the number of moles of the species in the
gasifier as the char gasification reaches equilibrium. These
figures show
_ _ that all species approach their equilibrium
values (X /X eq = 1) after the gas has traveled a distance of
2.5 m, which was set as the height of the gasifier as
indicated in Fig. 1. Fig. 9 shows the mole fraction of the
species as a function of the axial distance. It shows the
mole fractions of H2 and CO increase and the mole
fractions of H2O and N2 decrease with distance. This
increase in the mole fractions of H2 and CO and decrease
of that of H2O are caused by the chemical reaction between
the steam and the char. The decrease in the mole fraction of
N2 is mainly due to the generation of gas species of CO and
H2 during char gasification. Fig. 9 also shows that the slope
or the mole fraction of CO2 increases slightly and then
decreases with distance beyond X1 = 0.55 m. This phenom-

Fig. 12. Operating concept of the gasifier with equilibrium feed rates at 1150 K using Johnsons [10] kinetic data.

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

enon is caused by the different shift direction of water gas


shift reaction
CO H2 OVCO2 H2 :

47

Initially, the chemical kinetic makes reaction (47) to shift


to the right, this will generate CO2 and make the slope of
CO2 curve positive. However, beyond X1 = 0.55 m the
chemical kinetic makes reaction (47) to shift to the left,
this will consume CO2 and make the slope of CO2 curve
negative. The comparison of the slope of CO curve with
that of H2 curve in Fig. 9 also shows the same shift
tendency. Initially, the slope of H2 curve is larger than that
of CO curve, but becomes smaller than that of CO curve
when X1 is beyond 0.55 m. The larger curve slope
represents the larger generation rate of gaseous species
and vice versa. Thus, the chemical kinetic initially makes
reaction (47) to shift to the right, which makes the
generation rate of H2 to be larger than that of CO, but
beyond X1 = 0.55 m the chemical kinetic makes reaction

179

(47) to shift to the left, which makes the generation rate of


H2 to be smaller than that of CO. Figs. 10 and 11 show the
gas velocity and the solid loading density as a function of
the axial distance, respectively. The gas velocity is
obtained by dividing the volume flow rate of gas by the
cross-sectional area of the gasifier and the solid loading
density is the ratio of the circulation rate of the solids to
the volume flow rate of gas. These figures show that the
gas velocity increases and the solid loading density
decreases as the axial distance increases. This phenomenon
is caused by the increase in the volume flow rate of gas as
the char gasification reaction
gasif ication

Cs H2 O 

! CO H2

48

proceeds along the axial distance. The char gasification


reaction consume 1 mol of H2O(g) but generates total two
mol of CO and H2.
All three operating parameters (reactor temperature,
thermal capacity, and water concentration in the CWSF)

Fig. 13. Operating concept of the gasifier with equilibrium feeding rates using Wens [9] kinetic model.

180

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

together with feed rates of materials are shown in Fig. 12.


The recycle ratio, the ratio of the circulation rate of lime to
the feed rates of CWSF, is calculated to be 203. When the
reactor temperature, the composition of the CWSF, and the
thermal capacity of the gasifier are specified, the maximum
volume flow rate in the gasifier is determined, i.e. the
maximum gas velocity in the gasifier is fixed. Industrial
experience indicates that a circulating fluidized bed should
operate at higher gas velocities than those indicated by Fig.
10 for improved performance [20]. In order to obtain higher
gas velocities, the unit should be operated at a higher
temperature and thermal capacity than those indicated in
Fig. 12. Thus, the following section demonstrates a way to
operate the facility at higher gas velocities by increasing the
reactor temperature and the gasifiers thermal capacity.
Also, Wens [9] kinetic model is used in order to show
the effect of a more sophisticated and realistic gasification
scheme.

Fig. 15. The composition of gas as a function of axial distance along the
gasifier using Wens [9] kinetic model.

This section demonstrates how to obtain a superficial


velocity between 3.6 and 4.9 m/s (12 16 ft/s) and a solid
loading density between 50 and 100 kg/m3, which are
recommended by industrial experience [20]. In this analysis,
Wens [9] kinetic model (Eq. (29)) was used to predict the
gasification rate of the char in the gasifier. Wens kinetic
model (Eq. (29)) describes the gasification rate of char as a
function of the reactor temperature and the concentration of
the species, steam, CO, and H2. A trial-and-error technique
was used to determine the operating parameters and the
circulation rates of char and lime so that the superficial gas
velocity and solid loading density fell within the specified
range. In this method, the water concentration of the CWSF
was fixed at 30%, but the other two operating parameters,
the reactor temperature and the gasifiers thermal capacity,
were varied. The results of this trial and error method show

that the gasifier is required to operate at 1450 K with a


thermal capacity of 879 KW (3 106 Btu/h) and circulation
rates of char and lime of 15 and 4.5 kg/s, respectively, in
order to reach the specified range of gas velocity and solid
loading density. The chemical reactions, however, do not
reach equilibrium at a height of X1 = 2.5 m and only 88% of
the water is consumed. The emissions of H2S (0.15 lb of
SO2 per million Btu) based on the total conversion of H2S to
SO2 after combustion were found still to be below the
allowed SO2 emission levels (1.2 lb per million Btu).
Fig. 13 shows the sketch of the gasifier with the feed
rates of materials and the operating conditions that are
needed to reach the expected ranges of gas velocity and
solid loading. The recycle ratio, the ratio of the circulation
rate of lime to the feed rate of CWSF, is calculated to be
163. The equilibrium feed rates were calculated at 1450 K
for the gasifiers thermal capacity of 879 KW (3 106 Btu/
h) and a CWSF containing 30% of water by weight. Fig. 14

Fig. 14. The consumption of steam as a function of axial distance along the
gasifier using Wens [9] kinetic model.

Fig. 16. Gas velocity as a function of axial distance along the gasifier using
Wens [9] kinetic model.

3.2. Wens kinetic model

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

181

Fig. 19. Circulating rate of char as a function of axial distance along the
gasifier.
Fig. 17. Solid loading density as a function of axial distance along the
gasifier using Wens [9] kinetic model.

shows the consumption of steam as a function of axial


distance. It shows that after the gas has traveled a distance of
2.5 m, 88% of the steam has been consumed in the gasifier.
From the thermodynamic analysis, it was found that the
consumption of steam should be larger than 99% as the
gasification reaches equilibrium. Thus, this figure indicates
that the gasification reactions have not reached equilibrium
by the time the product gas reaches the exit at X1 = 2.5 m.
The mole fractions of the gas species as a function of axial
distance are shown in Fig. 15. The superficial gas velocity
and the solid loading density are shown in Figs. 16 and 17,
respectively. They show that the superficial gas velocity and
solid loading density are within the specified ranges of 3.6
4.9 m/s and 100 50 kg/m3, respectively.
It is interesting to compare Fig. 16 with Fig. 10. Fig. 16,
which was obtained using Wens [9] kinetic model, shows
that the gradient of the superficial gas velocity in the flow
direction decreases as the axial distance increases. However,
Fig. 10, obtained from Johnsons [10] kinetic model, indicates that the superficial gas velocity increases linearly as
the axial distance increase. This difference is due to the

decrease in the gasification rate of the char as the concentrations of CO and H2 increase in Wens kinetic model.
3.3. Circulation rate and shrinkage of chars
The transient Eqs. 42 and 44 developed in Section 2.5
are solved by varying the initial diameter and the number
of circulation of the chars until the char particle shrinks to
be the cut-off diameter and the circulation rate of chars
equals the required circulation rate of chars. Figs. 18 and
19 show the results obtained by using Johnsons kinetic
data at 1150 K. The circulation rate of char is accumulated
to such amount that the chemical reaction of char gasification reach equilibrium at X1 = 2.5 m. The diameter of
the char is plotted as a function of the number of circulations in Fig. 18, and the circulation rate of the char is
plotted as a function of the number of circulations in Fig.
19. Figs. 18 and 19 show that the circulation rate of char
will reach a steady value of 11.8 kg/s after 2860 circulations from the startup if the initial diameter of the char is
80.5 Am. The char particle is expected to shrink from 80.5
Am to the cut-off diameter of the cyclone combustor, 19.7
Am, and be carried out of the cyclone combustor after 2860
circulations.

4. Conclusions and suggestions


The following conclusions have been made:

Fig. 18. Shrinkage of char as a function of axial distance along the gasifier.

(1) The char steam reaction in the gasifier does not reach
equilibrium at low char circulation of 11.8 kg/s and low
temperature of 1150 K. It needs to operate at higher char
circulation rate (15 kg/s) and higher temperature (1450
K) for the char steam reaction to reach equilibrium. It is
suggested to predict the performance of the gasifier from
the kinetic point of view in order to design the optimal
geometry and operating methods that will allow the
gasifier to play a successful role in the clean coal
technology system such as IGCC, HIPPS, etc.

182

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

(2) It shows a concept to operate the gasifier from transient


to steady condition after the diameter of char particle
shrinks to equal the cut-off diameter of the cyclone
combustor and the circulation rate of char equals the
required amount simultaneously. Whenever the system
reaches steady condition the particle size distribution
and the circulation rate of char will not change.
(3) The use of coal water slurry as fuel has restricting effects
on the performance of the gasifier to be used in IGCC
because in this way the water feeding rate becomes
proportional to the char feeding rate and cannot be varied
independently. Accordingly, if the gasifier is operated
under steady condition in IGCC its thermal capacity
cannot be modified by only varying the water feeding
rate without the change of operating temperature.
(4) A large amount of circulating lime CaO in the gasifier
makes the removal of H2S from the flue gas relatively
easy. The emissions of H2S (0.15 lb of SO2 per million
Btu) based on the total conversion of H2S to SO2 after
combustion were found still to be below the allowed
SO2 emission levels (1.2 lb per million Btu).
List of symbols
A
Concentration of available lime CaO in the
calcined limestone, any units
As
Surface area of single particle (m2)
Ap
Cross-section of char particle (m2)
B
Concentration of inactivated lime CaO covered by
CaS layers in the calcined limestone, any units
C
Concentration of H2S (gmol/E)
Cd
Drag coefficient
Cf
Free site on the surface of char available for
reaction
C(o)
Occupied site or activated complex on the surface
of the char
De
Diameter of the imaginary cylinder
dp
Diameter of the solid particle (m)
dpc
Cut-off diameter of the cyclone combustor (m)
dpi
Diameter of the solid particle after ith circulation
(m)
d pn
Diameter of char particle after n circulations (m)
dp0
The initial diameter of char (Am)
Ei
Activation energy (Kcal/gmol K)
FC
Mass fraction of carbon in the CWSF
FD
Drag force upon the solid particle in the radial
direction (N)
H
Height of the imaginary cylinder (m)
h
Exponential constants of Rankine vortex in Eq.
(39)
k1
Rate constant of Eq. (8)
k1V
Rate constant of Eq. (8)
k2
Rate constant of Eq. (9)
k2V
Rate constant of Eq. (9)
k3
Rate constant of Eq. (10)
K1
Rate constant of adsorption of CaS on the surface
of available lime (E/gmol s)

K2
K3
Ke
K1
K2
K
kC H2O
Ks
Kv
n
N_
N
_
N_c
N H 2S
Mp
P
_i
Q
r
R
Rep
T
t
Dt
Ur
Uh
Ug
Vr
Vr2
Vh
Vh2
W
_
W
_c
Wf
W0
W0
DWgi
DWi
X
X1
DX1
_
X_
Xeq
Y
Z
q

Rate constant of desorption of CaS on the surface


of inactivated lime (l/s)
Rate constant of Eq. 2 (E/gmol s)
Equilibrium constant [0.3012 105 exp(7603.2/
T)] (E/mol)
Equilibrium constants of Eq. (8)
Equilibrium constants of Eq. (9)
Equilibrium constants of Eq. (16)
Equilibrium constant [exp(17.64 16810/T)]
Surface rate coefficient (kg/s m2)
Rate constant of char gasification[exp(24.3
25,120/T)] (cm3/mol s)
Number of circulations of char
Total unconverted lime
Feed rate of the char (particle/s)
Rate of char gasification (atom/s)
Rate of the reduction of H2S (atom/s)
Mass of a single particle [6qpdp3/p] (kg)
Pre-exponential constant
Gas volume flow rate through the imaginary
surface (m3/s)
Coordinate in the radial direction
Universal gas constant (cal/mol K)
Reynolds number [(Ur  Vr)dp/m]
Temperature (K)
Time (s)
Time interval (s)
Solid particle velocity in the radial direction (m/s)
Solid particle velocity in the angular direction (m/s)
Superficial gas velocity (m/s)
Gas velocity in the radial direction (m/s)
Gas velocity on the imaginary surface (m/s)
Air tangential entering velocity (m/s)
Gas velocity in the angular direction on the
imaginary surface (m/s)
Number of moles of gas species, char, or lime
(mol) based on 1 mol of carbon in feed CWSF
(mol)
Circulation rate of char (kg/s)
Feeding rate of coal water slurry fuel (kg/s)
Initial mass of feeding char particle (kg)
Moles of gas species or lime at the beginning of
each time interval (mol)
Gasification of a single char particle in the ith
circulation (kg)
Total mass reducation of a single char particle in
the ith circulation (kg)
Converted atoms of in each time interval (mol)
Axial coordinate along the gasifier (m)
Traveling distance along the gasifier after a time
interval Dt
Mole of gas species (mol)
Mole of gas species at equilibrium (mole)
Converted atoms of CO in each time interval (mol)
Converted atoms of in each time interval (mol)
Density of flue gas (kg/m3)

H. Tsui, C.-H. Wu / Powder Technology 132 (2003) 167183

qp
m
l
c

Density of solid particle (kg/m3)


kinetic viscosity of gas
Viscosity of gas (N s/m2)
Specific rate of char gasification (l/s)

References
[1] L.A. Ruth, Advanced coal-fired power plants, ASME, J. Energy Resour. Technol. 123 (2001) 4 9.
[2] E.A.J. Gandolfi, G. Papachristodoulou, O. Trass, Preparation of coal
slurry fuels with the Szego mill, Powder Technol. 40 (1984) 269 282.
[3] S. De, P.K. Nag, Thermodynamic analysis of a partial gasification
pressured combustion and supercritical steam combined cycle, Proc.
Inst. Mech. Eng., A J. Power Energy 214 (2000) 565 574.
[4] J.M. Topper, P.J.I. Cross, S.H. Goldthorpe, Clean coal technology for
power and cogeneration, Fuel 73 (7) (1994) 1056 1063.
[5] L.A. Ruth, The US department of energys combustion 2000 Program: clean efficient electricity from coal, Energy Convers. Manag.
38 (10 13) (1997) 1249 1257.
[6] S.J. Goidich, M. Seshamani, The Foster Wheeler Approach to Circulating Fluidized Bed Technology, The Third International Power
Generation Industries Conference and Exhibition, Orlando, FL, USA,
1990.
[7] H.D. Tsui, Thermodynamic, kinetic, and hydrodynamic analyses of
the gasification of coal water slurry fuels in a circulating fluidized bed
gasifier, PhD Dissertation, Penn State University, 1990.
[8] C. Luo, T. Watanabe, M. Nakamura, S. Uemiya, T. Kojima, Gasification kinetics of coal chars carbonized under rapid and slow heating
conditions at elevated temperature, J. Energy Resour. Technol. 123
(2001) 21 26.
[9] C.Y. Wen, E.S. Lee, Coal Conversion Technology, Addison Wesley,
Massachusetts, USA, 1979, p. 586.
[10] J.L. Johnson, Kinetics of Coal Gasification, Wiley, NY, 1979.
[11] S. Ergun, Kinetics of the reactions of carbon dioxide and steam with
coke, Bur. Mines Bull. 598 (1961) (R622.05 v53-1 No. 598 605).
[12] S.A. Quader, Natural Gas Substitutes from Coal and Oil, Coal Science
and Technology, vol. 8, Elsevier, 1985.

183

[13] R.J. Dry, I.N. Christensen, C.C. White, Gas solids contact efficiency in a high-velocity fluidized bed, Powder Technol. 52
(1987) 243 250.
[14] J.R. Howard, Fluidized Beds Combustion and Applications, Science
Publish, 1983.
[15] H.H. Lowry, Chemistry of Coal-UtilizationSupplementary Volume,
Wiley, New York, 1963.
[16] W.T. Abel, F.W. Shultz, P.F. Langdon, Removal of Hydrogen Sulfide from Hot Producer Gas by Solid Absorbents, Bureau of Mines,
U.S. Department of the Interior, Morgantown, 1974, Publication RI7947, 1974.
[17] J.T. Schrodt, O.J. Hahnn, Hot Fuel Gas Desulfurization, Institute for
Mining and Minerals Research, University of Kentucky, Lexington,
KY, 1976, Publication IMMR/15-PD11-76.
[18] M. Pell, R.A. Graff, A.M. Squires, Desulfurization of fuel with calcined dolomite, Chemical Engineering Progress Technical Manual,
AIChE, New York, NY, 1971.
[19] B.J. Overmoe, S.L. Chen, L. Ho, W.R. Seeker, M.P. Heap, D.W.
Pershing, Boiler simulator studies on sorbent utilization for SO2 control, Proceedings: First Joint Symposium on Dry SO2 and Simultaneous SO2/NOx Control Technologies, vol. 1, Fundamental Research
and Process Development, July, 1985, pp. 15-1 15-7, EPA-600/9-85020a (NTIS No. PB85-232353).
[20] K.K. Kou, Principles of Combustion, Wiley-Interscience, New York,
1986.
[21] D.F. Ciliberti, B.W. Lancaster, Performance of rotary flow cyclones,
AIChE 22 (1976) 394 398.
[22] D. Kunii, O. Levenspiel, Fluidization Engineering, Robert E. Krieger
Publishing, New York, 1977.
[23] A. Ogawa, Separation of Particles from Air and Gases, vols. 1,2, CYC
Press, 1984.
[24] W. Barth, K. Trunz, Modellversuche mit wasserdurchstromten zyklonab-scheidern zur vorausbest timmung der abscheideleistung, Z. Angew. Math. Mech. 3 (819) (1950) 255.
[25] A.K. Cupta, D.G. Lilley, N. Swirl, Swirl Flows, Energy and Engineering Science Series, ABACUS Press, 1984.
[26] V.S. Semenov, N.A. Semenenko, Pattern of combustion and gasification of coal in a cyclone chamber, Teploenergetika 16 (12) (1969)
79 81.

You might also like