Você está na página 1de 14

INTERNATIONAL JOURNAL OF ENERGY RESEARCH

Int. J. Energy Res., 22, 657 670 (1998)

A MODEL OF THE COMBUSTION OF A SINGLE SMALL COAL


PARTICLE USING KINETIC PARAMETERS BASED ON
THERMOGRAVIMETRIC ANALYSIS

D. VAMVUKA1* AND E. T. WOODBURN2


1 Department of Mineral Resources Engineering, Technical University of Crete, 731 00 Chania, Greece
2 Department of Chemical Engineering, P.O. Box 88, Manchester M60 1QD, U.K.

SUMMARY
A mathematical model for the combustion in air of a single entrained spherical coal particle, 30 lm in diameter, has been
developed incorporating thermogravimetric analysis data of Whitwick coal. The model is based on a set of ordinary
differential equations, describing the reaction rates and the mass and heat transport processes. The system of equations
was solved numerically. The combustion mechanism of the particle was described by locating the reaction zone at the
solid surface, where gas-phase combustion of volatiles and heterogeneous reaction between gaseous oxygen and the
carbon and hydrogen in the solid occurred in parallel. The combustion process was chemical-reaction-rate-controlled,
with the oxygen partial pressure at the surface almost that of the surrounding bulk gas. The simulation results using this
model, with the kinetic parameters for devolatilization and combustion derived from the experimental thermogravimet-
ric data, are consistent with previously reported combustion lifetimes of approximately 1 s, for particles of this size and
rank. They are also consistent with the anticipation that higher ambient gas temperatures should result in shorter
burn-out times. The use of thermogravimetric data in the modelling of the combustion of small particles of these
low-rank coals is a potentially valuable method for characterization of feedstocks for pulverized coal-fired boilers.
( 1998 John Wiley & Sons, Ltd.
KEY WORDS combustion; coal particle; thermogravimetric

1. INTRODUCTION
The combustion of isolated coal particles has been the subject of many theoretical and experimental studies,
because of its practical importance for the design and operation of pulverized-fuel furnaces. Most of these
studies (Howard and Essenhigh, 1967; Timothy et al., 1982; Jost et al., 1984; Gururajan et al., 1988; Beck and
Hayhurst, 1990; Lufei et al., 1993; Baum and Street, 1971) adopted the oil drop model for the combustion of
coal volatiles, according to which the volatiles are assumed to be burned in a spherical film which is stabilized
some distance around the particle, while the residual char is assumed to burn only after volatiles evolution
has ceased. However, this model has proved to be unsatisfactory by several investigators (Howard and
Essenhigh, 1967; Shadman and Cavendish, 1980; Saito et al., 1991; Saastamoinen et al., 1993) who showed for
high temperatures, high oxygen content of the gas or slow evolution of pyrolysis products, that ignition starts
adjacent to the particle surface, so that gas-phase and heterogeneous combustion take place in parallel.
Howard and Essenhigh (1967) were the first to recognize the possible overlap of volatiles and char
combustion and predicted that the rate of oxygen diffusion, for particles below 65 lm, was sufficient to
suppress the formation of a volatile flame. Such an occurrence for small particles was later on confirmed by

* Correspondence to: D. Vamvuka, Department of Mineral Resources Engineering, Technical University of Crete, 731 00 Chania,
Greece.

CCC 0363907X/98/070657 14$17.50 Received 29 July 1997


( 1998 John Wiley & Sons, Ltd. Accepted 7 November 1997
658 D. VAMVUKA AND E. T. WOODBURN

other researchers (Saastamoinen et al., 1993; Annamalai and Durbetaki, 1991; Kramlich et al., 1988; Du and
Annamalai 1994), including one of the authors Vamvuka, 1988.
Among the various models that have been developed to allow parallel combustion of volatiles and char,
some coupled combustion of volatiles in the gas-phase and/or gas-phase transport of reaction products with
reaction at the external particle surface (Beck and Hayhurst, 1990; Saastamoinen et al., 1993; Field et al.,
1967; Hamor et al., 1973; Caram and Amundson, 1977; Mon and Amundson, 1978). Others extended the
analysis of the heterogeneous combustion to include intraparticle reaction, diffusion within the particle and
pore growth (Srinivas and Amundson, 1980; Sotirchos and Amundson, 1984; Loewenberg et al., 1987; Hsuen
and Sotirchos, 1989, Chen et al., 1994). The difficulty of defining internal surfaces due to inadequate
information on the pore structure of a burning coal particle still restricts modelling at this degree of
sophistication.
All previous models have been based on data derived from flow experiments, under rapid heating
conditions. It was the purpose of this study to investigate whether a mathematical model for the combustion
of a single pulverized coal particle, incorporating parameters obtained from thermogravimetric analysis
(TGA), could predict to a satisfactory degree the various characteristics of the process, in particular the
burning time of the particle. As the TGA measurements are relatively easy to obtain they offer potentially
a convenient means of allowing for variations in particularly low-rank coal quality in the operation of
pulverized-fuel-fired boilers.
The model described in the paper postulates parallel gas-phase heterogeneous combustion and, at the
particle surface. The rate of combustion allows for the diffusion of oxygen from the bulk gas to the particles
surface. It is shown that this is relatively rapid and the partial pressure of oxygen at the surface is
approximately that of the bulk gas. The combustion is consequently limited by the release of volatiles and the
surface reaction with oxygen. The devolatilization is considered to be a pseudo-first-order reaction based on
the residual volatile content of the particle, while the surface reaction is based on the area of the remaining
char and the oxygen partial pressure at the surface. For the TGA tests the oxygen partial pressure at the
surface was taken to be that of the bulk air.
The validity of the postulate, that rate constants derived from thermogravimetric data could be used to
assess the combustion behaviour of fine coal particles, was tested by formulating a model incorporating these
parameters and then comparing its predictions with previously published data.

2. GENERAL DESCRIPTION OF MODEL POSTULATES


A spherical coal particle (30 lm in diameter) is considered to be immersed in an oxygen-containing gaseous
medium flowing past it at a uniform relative velocity. Previous work by Vamvuka (1988) provided the
basis for the assumption of parallel devolatilization and surface reaction. When the devolatilization of coal
had been presumed to occur first, followed sequentially by the heterogeneous combustion of the produced
char, the simulation results did not correspond with the experimental data. This implied that the rate of
volatiles evolution was slow in comparison to the inward rate of oxygen diffusion. Better agreement was
achieved by assuming that the radius of the shell of the volatiles combustion was the same as that of the
particle.
The surrounding boundary-layer was therefore considered to be stagnant gas, of thickness one particle
radius (Caram and Amundson, 1977; Mon and Amundson, 1978; Sotirchos and Amundson, 1984), as shown
schematically in Figure 1. Through this stationary film inward diffusion of oxygen, outward diffusion of
reaction products and heat took place.
The composition of the gas at the particle surface is defined by surface reactions, the transport processes
through the boundary layer and the oxygen partial pressure of the main stream. In the TGA experimental
tests however, the oxygen partial pressure at the surface was taken to be that of the main stream which was
constant during the combustion.

Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.
A MODEL OF THE COMBUSTION OF A SINGLE COAL PARTICLE 659

Figure 1. Schematic diagram of coal particle combustion

With excess oxygen the volatile material was assumed to burn completely as soon as it was evolved, while
the solid material reacted irreversibly with the oxygen at its surface, according to a second-order process. The
composition of the residual solid changed with the extent of the devolatilization. Chemical reactions within
the gaseous boundary layer were taken to be negligible.
The heat generated at the solid surface was the sum of the heats evolved by the gas phase and
heterogeneous reactions. Heat was lost by radiation and by conduction through the fluid boundary layer to
the main stream of gas. The surface temperature at any given time was calculated from a heat balance by

( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 657670 (1998)
660 D. VAMVUKA AND E. T. WOODBURN

assuming that the temperature of the particle was uniform throughout its volume (as shown by Vamvuka,
1988).
Other simplifying assumptions used in the model development are summarized below:
(a) Heat conduction through the boundary-layer flow was a quasisteady-state, one-dimensional laminar
flow in the r-direction.
(b) All body forces, gravity, viscous and thermal diffusivity effects, pressure gradient diffusivity, particle
interactions, convective losses, kinetic energy and Stefan flow were neglected.
(c) All gases obeyed the ideal gas law and binary diffusion coefficient pairs were considered equal.
(d) The pressure, temperature and oxygen concentration of the free stream were constant.
(e) Due to the lack of appropriate kinetic data and information on the change of internal surface area as
reaction proceeds, internal mass transport effects were ignored. (This is reasonable for low-rank coals, as
almost all their internal area is contained within pores less than 100 As , particularly when high combus-
tion temperatures are considered).
(f ) The particle radius was allowed to shrink, but the residual ash was not considered to remain on the
reacting coal surface, as required by the shrinking core model.
(g) The evolved volatiles consisted of methane, carbon monoxide and water vapour, in proportions 2:1:1.

3. EXPERIMENTAL TGA DATA


Devolatilization and differential combustion tests were done using a Linseis Thermo-Gravimetric Analyser
on three low-rank coals supplied by the National Coal Board (type NCB 902) and on a Lakhra lignite from
Pakistan. Proximate analyses are given in Table 1.
For the devolatilization tests, 190 mg of screen fraction samples with a mean size of approximately 30 lm
were dried at 110C under nitrogen before processing. After constant mass was achieved the furnace
temperature was ramped at a rate of 20C min~1 under nitrogen to a maximum of 850C. The instantaneous
devolatilization rates were obtained from the cumulative experimental weight loss curves by graphical
differentiation. The rate constant, as a function of temperature, could be inferred from the slopes and the
corresponding residual mass of volatiles:
!dM ()/dt
k ()" 7 (1)
7 [M ()!M ]
7 &
The devolatilization Arrhenius constants for the three low-rank coals and the lignite are reported in
Table 2(a).
The combustion data were obtained by alternating nitrogen and air flow at a constant furnace temper-
ature. The furnace was initially held at about 300C. Nitrogen was passed over the sample until constant
mass was recorded, at which point it was assumed that the sample was free of moisture and volatiles at the
furnace temperature. Once constant mass was achieved the nitrogen stream was replaced by an air stream
with the furnace temperature maintained at its previous value. The mass loss caused by combustion and the
time were measured and the maximum temperature reached noted. Once the mass had stabilized the air was
replaced by nitrogen and the next cycle started. In this the furnace temperature was set slightly greater
(2040C) than the previous value. These cycles were repeated until no further mass loss was noted in the
combustion phase. The final temperature reached with these coals was about 550C.
The rate of combustion at the furnace temperature was approximated from the time required to complete
the mass loss in the interval. A pseudo-first-order rate constant was calculated from the point rate divided by
the mass of unburnt devolatilized coal remaining:
!dM ()/dt
k ()" # (2)
# [M ()!M ]
# &$
Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.
A MODEL OF THE COMBUSTION OF A SINGLE COAL PARTICLE 661

Table 1. Proximate analysis of experimental coal samples

Moisture content Total volatiles Ash content Fixed carbon


Coal sample (wt%) (wt%) (wt%) (wt%)

S. Leicestershire 4)82 30)30 5)18 59)71


NCB 902
Whitwick 5)05 29)88 10)65 54)41
NCB 902
Willesley 5)21 31)50 13)13 50)17
NCB 902
Lakhra 7)37 36)47 12)30 43)86
Lignite

Table 2(a). Devolatilization rate constants for experimental coals

Pre-exponential factor Activation energy]10~3


Coal (s~1) (kJ kmol~1)

S. Leicestershire NCB 902 0)03 26)09


Whitwick NCB 902 0)04 28)29
Willesley NCB 902 0)02 25)19
Lakhra Lignite 0)02 25)45

Table 2(b). Combustion rate constants for experimental coals

Pre-exponential Pre-exponential
factor factor]104 Activation energy]10~3
Coal (s~1) (kg s~1 m~2 Pa~1) (kJ kmol~1)

S. Leicestershire NCB 902 302)3 1)41 85)29


Whitwick NCB 902 490)4 1)93 87)41
Willesley NCB 902 0)90 0)004 47)74
Lakhra Lignite 9)86 0)046 60)35

Table 2(c). A comparison of the initial rate constants for the various coals with the correspond-
ing value based on the Arrhenius regression over the entire temperature range

Furnace Combustion Experimental rate Regressed


temperature temperature constant rate constant
Coal (C) (C) (min~1) (min~1)

S. Leicestershire NCB 902 330 431)4 0)0357 0)0087


Whitwick NCB 902 330 336)7 0)0011 0)0010
Willesley NCB 902 300 337)6 0)0085 0)0045
Lakhra Lignite 290 412)6 0)0461 0)0150

The lignite and S. Leicestershire burned rapidly at the initial furnace temperature generating temperatures
significantly in excess of the furnace setting. These initial values were outliers on a otherwise satisfactory
linear regression of rate constants vs. 1/(K). The Whitwick coal had the slowest initial rate and burned
consistently with the Arrhenius regression over the whole temperature range. After the initial combustion

( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 657670 (1998)
662 D. VAMVUKA AND E. T. WOODBURN

above 400C$ all the chars rate constants were well represented by the Arrhenius equation:
*E 1
ln k ()"ln k ! (3)
# #0 R
The Arrhenius constants for the pseudo-first-order combustion, measured in this way, are shown in
Table 2(b). Table 2(c) shows the initial rate constants for the coals and the corresponding deviation from the
regression prediction.
From the pseudo-first-order approximation for the spherical coal particle of radius r , a relationship
#
appropriate for the combustion of a shrinking particle can be derived in terms of the surface reaction:
dM ()
! # "k ()p 4nr2 (4)
dt 4 O4 #

4. MATHEMATICAL FORMULATION
Based upon the above experimental data and the stated assumptions, a series of model equations governing
the combustion process of the coal particle was formulated.

4.1. Mass balances


Although it was assumed in the TGA tests that the oxygen partial pressure at the coal surface was that of
the ambient gas, in the model, the rate of diffusion of oxygen to the solid surface was related to the chemical
and devolatilization rates at the surface:

A B
dM dM
! f 7#f # "N 4nr2 . (5)
7 dt # dt O4 #
By assuming equimolar counterdiffusion, through a film thickness equal to the radius of the particle, the
flux of oxygen is
DM P
N " (y !y ) (6)
O4 RTM r O= O4
#
The rate of volatile combustion was taken to follow the experimental devolatilization rates with pseudo-
first-order kinetics and the rate of char combustion was given by the surface reaction with a shrinking
particle radius and at an oxygen partial pressure given by equation (4).
The ratio of solid material consumed to oxygen transported to the surface was dependent on the product
which was transported away, according to the reaction (Baum and Street, 1971; Field et al., 1967)

A B A B
1 1 2
C# O N2 1! CO# !1 CO (7)
u 2 u u 2
where the stoichiometric factor u for particle sizes less than 50 km was given by (Field et al., 1967)
2z#2
u" (8)
z#2
where
z"2500e~6249@T (9)

4.2. Heat balance


The thermal energy balance over the reaction zone was formulated by equating the rate of heat generated
at the particle surface, due to simultaneous homogeneous and heterogeneous combustion, to the rate of heat

Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.
A MODEL OF THE COMBUSTION OF A SINGLE COAL PARTICLE 663

transferred to the surrounding atmosphere by conduction (assuming Nu"2) and radiation:

A B C D
dM dM j
! 7 *H # # *H " ' ( ! )#e p(4!4 ) 4nr2 (10)
dt 7 dt # r !r # = # # = #
= #
For n volatiles species, the heat liberated from the combustion of volatiles was
n
*H " + y *H . (11)
7 i 7i
i/1
while the heat liberated from the combustion of solid carbon was (Baum and Street, 1971; Field et al., 1967)

A B A B
2 2
*H "*H !1 #*H 2! (12)
# CO2 u CO u
The dependence of the heat of reaction on temperature was described by Kirchhoff s equation.

5. METHOD OF SOLUTION
The initial boundary conditions for each of the first-order ordinary differential equations were expressed by
at t"0, "
# #0
r "r
# #0
M "M ( , t"0) (13)
70 7r #0
dM
#"0
dt
The radius of coal particle was divided into 10 equal increments defining 10 spherical shells. The changes in
particle mass and size, combustion rate and interface composition of reactant and product gases were
computed as a function of time. Briefly, the solution procedure was as follows.
From the postulated initial conditions (a cold spherical particle of 30 lm diameter is dropped in hot air at
1500C), the combustion rate, which was based solely on volatiles evolution, was calculated by equation (1).
A small increment of time was chosen. The heat balance equation (10) had to be solved iteratively for the
surface temperature by using a modified Powell (1970) hybrid method. The combustion rates of volatiles
#
was calculated from equation (1) and the combustion rate of solid carbon by equations (4)(6). This required
an iterative estimate of the oxygen partial pressure at the surface.
The above procedure was repeated, holding the particle radius constant at its equivalent shell value, over
subsequent time steps, until all spherical particle shells were consumed in sequence.
The logic flow diagram of the computer program is given in Figure 2. Input data to the computer program,
are summarized in Table 3.

6. MODEL PREDICTIONS

6.1. Prediction of burning time and its relation to combustion characteristics


Model predictions, including temperature of the coal particles, the changes in particle size, mass and
combustion rate, as well as the composition of gas and the heat of combustion, at each time step, for the
conditions indicated in Table 4, will be presented here, in order to illustrate its consistency with previous
data. If this consistency is acceptable, then the use of TGA data in a computational model of this type would
be a valid method of predicting the performance of various coals.

( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 657670 (1998)
664 D. VAMVUKA AND E. T. WOODBURN

Figure 2. Computer flow diagram for the model

Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.
A MODEL OF THE COMBUSTION OF A SINGLE COAL PARTICLE 665

Table 3. Solid-phase property and parameter values used in the simulation

Property/parameter Value/equation Reference

Density o (kg m~3) 1430 UMIST lab


#
Specific heat c (kcal kg~1 K~1) 0)222#2)18]10~4!9741)666/2 Chem. Eng. Handbook
1#
Thermal conductivity j (kcal s~1 m~1 K~1) 0)6]10~4 M. A. Field et al.,
#
Emissivity of coal surface e 0)9 M. Baum, P. J. Street
#
Initial particle diam d (lm) 30
#
Initial moisture (wt%) 5)05 D. Vamvuka
Initial volatiles (wt%, dry basis) 29)88 D. Vamvuka
Ash, content (wt%, dry basis) 10)65 D. Vamvuka
Fixed carbon (wt%) 54)41 D. Vamvuka
Rate constant of devolatilization, k (s~1) 0)04e~28>29@RT D. Vamvuka
7
Rate constant of combustion 1)926e~87>41@RT D. Vamvuka
k ]104 (kg s~1 m~2 Pa~1)
4

Table 4. Simulation results of coal particle combustion. Total pressure P, 105 Pa; ambient temperature , 1773 K;
=
boundary-layer thickness d, 15 lm

Surface Heat generated Oxygen mole Integer


Combustion Particle particle at particle fraction at Particle Combustion no. of
time radius temperature surface]109 particle mass]1011 rate]1011 Fractional shells
(s) (lm) (K) (kcal s~1) surface (kg) (kg s~1) conversion left

0 15 525)3 0 0)21 2)024 0 0 10


0)005 14)956 1746)7 465)856 0)2017 2)006 8)367 0)008 9
0)010 14)830 1843)5 584)890 0)1997 1)956 10)645 0)033 9
0)025 14)438 1843)5 513)769 0)2006 1)805 9)556 0)108 9
0)050 13)831 1843)5 415)072 0)2019 1)587 8)027 0)215 9
0)103 12)681 1843)5 265)560 0)2040 1)223 5)650 0)395 8
0)200 11)025 1843)5 126)851 0)2063 0)803 3)296 0)603 7
0)304 9)617 1834)5 63)501 0)2075 0)533 2)040 0)736 6
0)401 8)499 1827)7 37)767 0)2081 0)368 1)419 0)818 5
0)505 7)401 1825)5 24)059 0)2085 0)243 1)011 0)879 4
0)602 6)422 1823)2 16)459 0)2088 0)158 0)736 0)921 4
0)706 5)400 1823)2 11)043 0)2090 0)094 0)512 0)953 3
0)803 4)457 1823)2 7)328 0)2092 0)053 0)346 0)973 2
0)900 3)522 1820)9 4)486 0)2093 0)026 0)214 0)987 2
1)004 2)525 1820)9 2)292 0)2095 0)009 0)110 0)995 1
1)101 1)596 1820)9 0)916 0)2097 0)002 0)043 0)999 1
1)200 0)643 1820)9 0)153 0)2098 0)0001 0)007 0)999 0
1)234 0)320 1807)6 0)040 0)2099 0)00001 0)001 0)999 0

As it can be observed in Table 4, the particle temperature was higher than that of the ambient gas (1773 K)
after 0)01 s, which is in agreement with previously published data (Howard and Essenhigh, 1967; Baum and
Street, 1971; Saastamoinen et al., 1993; Srinivas and Amundson, 1977; Field, 1969). The temperature showed
an early peak when 3)3% of conversion was achieved, at which point a small diffusional resistance was noted,
as the mole fraction of oxygen at the surface dropped to 0)199.
In Table 4, the decrease in particle mass with time is reasonable. The combustion rate showed a peak at
a conversion of approximately 3%, due to the temperature rise of particle, after which it decreased with
particle size.

( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 657670 (1998)
666 D. VAMVUKA AND E. T. WOODBURN

Figure 3. Variations of volatile matter content, fixed carbon content and surface temperature of the coal particle with burning time

The surface mole fraction of oxygen was found to be only slightly below the ambient mole fraction. As it
was expected, for the small coal particle of this study boundary-layer diffusion was so fast that the overall rate
of combustion was completely controlled by the surface reaction. This finding is consistent with the reports of
other investigators (Howard and Essenhigh, 1967; Saastamoinen et al., 1993; Field et al., 1967; Mon and
Amundson, 1978).
The last column in Table 4 represents the degree of burning in terms of the disappearance of the spherical
shells, which modelled the particle shrinkage. On this basis, the lifetime of the particle was predicted to be
approximately 1)23 s, which is in reasonable agreement with the results obtained by previous workers.
Howard and Essenhigh (1967), who studied the combustion of pulverized coal particles in a flame at ambient
gas temperatures of 1400 K, found that a solid conversion of 70% was achieved in 0)8 s. Baum and Street
(1971), who modelled the combustion of a 53 lm coal particle in a tube furnace at 1300 K, adopting the oil
drop theory, reported burning time greater than 0)7 s. Also, Shadman and Cavendish (1980), who used
a shrinking density model to describe the combustion of a 26 lm char particle at 1600 K, predicted a burning
time of 0)8 s.
Variations in volatile matter and fixed carbon contents with burning time are represented in Figure 3,
together with the surface temperature history of the particle. As can be observed, after 0)4 s virtually all
volatiles had been burned, but there was still 27% of the original fixed carbon remaining. Clearly, the rather
long burning time of the volatiles reflects the slow liberation rate of these, as determined by the TGA
experimental tests. Smoot et al. (1977) and Solomon (1986) determined pyrolysis times of 0)05 s and 0)064 s,
respectively, at heating rates of the order of 104 K s~1, at ambient temperatures of 1000 K and for particle
sizes of 30 and 5374 lm, respectively, at which time this model predicts about 45% removal of volatiles.
The interface concentration profiles of the product gases during burn-off are shown in Figure 4. The mole
fractions reported are very small reaching a maximum at about 0)01 s. The largest peak was exhibited by the
curve corresponding to steam, as the latter, besides of being a volatile product, was also a product of methane
combustion, which constituted 50% of the volatiles evolved. The amount of carbon monoxide was higher
than that of carbon dioxide throughout the combustion process, agreeing with earlier conclusions (Field
et al., 1967; Mulcahy and Smith, 1969) that carbon monoxide is the dominant product at high temperatures.
After approximately 0)1 s, carbon monoxide became the main product. Finally, the concentration of oxygen
was only slightly below the ambient value, during the whole combustion period, and the concentration of
inert nitrogen was uniform at its ambient value.

Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.
A MODEL OF THE COMBUSTION OF A SINGLE COAL PARTICLE 667

Figure 4. Interface concentration profiles of product gases

Table 5. Effect of increased ambient temperature on combustion characteristics (cf. Table 4). Total pressure P, 105 Pa;
ambient temperature , 2000 K; boundary-layer thickness d, 15 lm
=
Surface Heat generated Oxygen mole Integer
Combustion Particle particle at particle fraction at Particle Combustion no. of
time radius temperature surface]109 particle mass]1011 rate]1011 Fractional shells
(s) (lm) (K) (kcal s~1) surface (kg) (kg s~1) conversion left

0 15 541)0 0)0 0)21 2)022 0 0 10


0)005 14)901 2046)5 953)143 0)1946 1)982 17)470 0)019 9
0)010 14)652 2135)9 1058)650 0)1928 1)884 19)741 0)068 9
0)025 13)923 2135)9 830)767 0)1954 1)617 16)212 0)200 9
0)050 12)846 2135)9 560)768 0)1988 1)270 11)911 0)372 8
0)103 11)014 2097)3 241)866 0)2036 0)800 6)275 0)604 7
0)200 8)677 2066)8 79)503 0)2066 0)391 2)847 0)806 5
0)304 6)640 2056)6 33)728 0)2078 0)175 1)476 0)913 4
0)401 4)873 2054)0 16)187 0)2085 0)0069 0)766 0)965 3
0)505 3)027 2051)4 5)972 0)2091 0)0017 0)291 0)992 2
0)602 1)363 2048)6 1)217 0)2095 0)0002 0)059 0)999 0
0)651 0)442 2048)6 0)150 0)2098 0 0)006 0)999 0

6.2. Effect of ambient temperature and combustion characteristics


As Table 5 shows, when the ambient temperature was increased from 1773 to 2000 K, all other conditions
being kept constant, the conclusion of surface reaction control still applied. Furthermore, as the surface
temperature of the particle increased with increasing ambient temperature, the burning time, required to
achieve complete conversion, decreased from 1)234 to 0)651 s. These findings agree well with analogous
studies made earlier on by other researchers (Field et al., 1967; Sotirchos and Amundson, 1984; Hsuen and
Sotirchos, 1989; Field, 1969).
The combustion rateparticle size and the lifetimeparticle size curves are compared in Figure 5.
The change in gas temperature from 1773 to 2000 K also affected substantially the heat generation curves,
as illustrated in Figure 6. As the surface temperature was raised, there was a gradual increase in the heat
generated, owing to the strong dependence of the surface reaction rate on temperature. Both curves passed

( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 657670 (1998)
668 D. VAMVUKA AND E. T. WOODBURN

Figure 5. Effect of ambient temperature on the burning rate and combustion time of the coal particle surface

Figure 6. Effect of ambient temperature on heat generated at the coal particle surface

through a maximum, but the broader peak was exhibited by the one corresponding to the higher ambient
temperature.
However, thermal effects are sensitive to boundary-layer thickness. The above findings were critically
dependent on the initial assumption of excess air in the bulk gas and might be different if the bulk gas
composition was changing during the course of combustion.

7. CONCLUSIONS
(1) The coal particle burned under a reaction rate control.
(2) While the combustion of volatile products was almost complete in 0)4 s, the lifetime of the coal particle
itself, with an ambient air temperature 1773 K, as estimated by the model, was approximately 1)23 s.

Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.
A MODEL OF THE COMBUSTION OF A SINGLE COAL PARTICLE 669

(3) The rate-controlling resistance was insensitive to variations in the ambient gas temperature.
(4) The surface temperature of the particle, and consequently the combustion rate and the heat generated by
the reaction, increased with increasing ambient gas temperature, resulting in shorter burn-out time of
0)65 s, for an ambient air temperature of 2000 K.
(5) The model predictions are reasonable, which suggests that thermogravimetric data are acceptable in
characterizing coals as feedstocks for pulverized coal-fired boilers.

NOMENCLATURE
c specific heat (kcal kg~1 K~1)
p
D1 diffusivity of oxygen in gas mixture, at the average coal particle boundary-layer temperature
(m2 s~1)
f ,f moles of oxygen required for complete combustion of a unit mass of volatiles and fixed carbon,
7 #
respectively (kmol kg~1)
k apparent rate constant for coal devolatilization (s~1)
7
k apparent rate constant for combustion of devolatilized coal (s~1)
#
k pre-exponential Arrhenius term for combustion (s~1)
#0
k rate constant for fixed carbon combustion (kg s~1 m~2 Pa O~1)
4 2
M() mass of coal at furnace temperature (kg s)
N diffusional flux of oxygen to the particle surface (kmol s~1 m~2)
O4
P total pressure (Pa)
p partial pressure (Pa)
r radial distance from centre of fuel particle at arbitrary time (m)
R universal gas constant (kJ kmol~1 K~1)
t time (s)
furnace temperature (K)
TM average temperature in boundary layer (K)
y mole fraction of gaseous component

Greek symbols
d Boundary-layer thickness (lm)
*E activation energy (kJ kmol~1)
*H heat of reaction (kcal kg~1)
e emissivity
j thermal conductivity (kcal s~1 m~1 K~1)
o density (kg m~3)
p StefanBoltzmann constant 13)55]10~12 (kcal s~1 m~2 K~4)

Subscripts
c coal, fixed carbon
f final
fd final devolatilized
g gas mixture
O oxygen
0 initial condition
s condition at the coal particle surface
v volatile products
vr residual volatiles
R freestream of gas

( 1998 John Wiley & Sons, Ltd. Int. J. Energy Res., 22, 657670 (1998)
670 D. VAMVUKA AND E. T. WOODBURN

REFERENCES
Howard, J. B. and Essenhigh, R. H. (1967). Proc. 11th Symp. (Int.) on Combustion, The Combustion Institute, Pittsburgh, p. 399.
Timothy, L. D., Sarofim, A. F. and Beer, J. M. (1982). Proc. 19th Symp. (Int.) on Combustion, The Combustion Institute, Pittsburgh, PA,
U.S.A., p. 1123.
Jost, M., Leslie, I. and Kruger, C. (1984). Proc. 20th Symp. (Int.) on Combustion, The Combustion Institute, Pittsburgh, PA, U.S.A.,
p. 1531.
Gururajan, V. S., Wall, T. F. and Truelove, J. S. (1988). Combust. Flame, 72, 1.
Beck, N. C. and Hayhurst, A. N. (1990). Combust. Flame, 79, 47.
Lufei, J., Becker, H. A. and Code, R. K. (1993). he Canadian J. Chem. Engng. 71, 10.
Baum, M. M. and Street, P. J. (1971). Combust. Sci. echnol., 3, 231.
Shadman, F. and Cavendish, J. C. (1980). Canadian J. Chem. Engng., 58, 470.
Saito, M. et al. (1991). Combust. Flame, 87, 1.
Saastamoinen, J. J., Aho, M. J. and Linna, V. L. (1993). Fuel, 72(5), 599.
Annamalai, K. and Durbetaki, P. (1991). Combust. Flame, 87, 1.
Kramlich, J. C., Seeker, W. R. and Samuelson, G. S. (1988). Fuel, 67, 1182.
Du, X. and Annamalai, K. (1994). Combust. Flame, 97(34), 330.
Vamvuka, D. (1988). Ph. D. Thesis, UMIST, Manchester, U.K.
Field, M. A., Gill, D. W., Morgan, B. B. and Hawksley, P. G. W. (1967). Combustion of Pulverised Coal, BCURA, Lleatherhead, England.
Hamor, R. J., Smith, I. W. and Tyler, R. J. (1973). Combust. Flame, 21, 153.
Caram, H. S. and Amundson, N. R. (1977). Ind. Engng. Chem. Fundam., 16(2), 171.
Mon, E. and Amundson, N. R. (1978). Ind. Engng. Chem. Fundam., 17(4), 313.
Srinivas, B. and Amundson, N. R. (1980). Canadian J. Chem. Engng., 58, 476.
Sotirchos, S. V. and Amundson, N. R. (1984). A.I.Ch.E. J., 30(4), 537.
Loewenberg, M., Bellan, J. and Gavalas, G. R. (1987). Chem. Engng. Commun., 58, 89.
Hsuen, H. K. D. and Sotirchos, S. V. (1989). Chem. Engng. Sci. 44(11), 2639.
Chen, H., Sun, X., Han, C. and Cui, H. (1994). J. Chem. Ind. Engng., 45(3), 327.
Powell, M. J. D. (1970). In Numerical Methods for Nonlinear Algebraic Equations, P. Rabinowitz (Ed.), Gordon and Breach, London,
England.
Perry, R. H. and Don Green, W. (1984). Perrys Chemical Engineers Handbook, 6th edn, McGraw-Hill, New York.
Field, M. A. (1969). Combust. Flame, 13, 237.
Smoot, L. D., Horton, M. D. and Williams, G. A. (1977). Proc. 26th Symp. (Int.) on Combustion, The Combustion Institute, Pittsburgh,
PA, U.S.A., p. 375.
Solomon, P. R., Serio M. A., Carangelo, R. M. and Markham, J. R. (1986). Fuel, 65, 182.
Mulcahy, M. F. R. and Smith, I. W. (1969). Rev. Pure Appl. Chem., 19, 81.

Int. J. Energy Res., 22, 657670 (1998) ( 1998 John Wiley & Sons, Ltd.

Você também pode gostar