Вы находитесь на странице: 1из 11

Low-temperature oxidation of single and blended coals

Y.S. Nugroho
1
, A.C. McIntosh
*
, B.M. Gibbs
Department of Fuel and Energy, University of Leeds, Leeds LS2 9JT, UK
Received 23 July 1999; received in revised form 3 March 2000; accepted 9 March 2000
Abstract
The oxidation of different types of coal under self-heating conditions is considered using the crossing point method performed in a cubical
wire-mesh basket. This study investigates the effect of particle size and the physical structure of coals (including pore size and surface area)
on the self-heating character of high and low rank Indonesian coals and their blends. The results conrm that both particle size and surface
area give considerably different effects on critical ambient temperature, the activation energy and the product of exothermicity and the pre-
exponential factor of low and high rank coals. It was found that the self-heating characteristics of high rank coals were strongly dependent
upon the particle size of the coals. For low rank coals, the dependence was weaker. A coal bed with mixed sizes is thus more vulnerable than
one with segregated sizes. This means that one must carefully consider the particle size distribution to judge the spontaneous ignition
behaviour of coal. It is suggested that the potential for spontaneous combustion of blended coal is greater following the increased percentage
of more reactive coal. 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Coal blend; Particle size; Surface area; Oxidation; Self-heating; Crossing point method
1. Introduction
It is widely accepted that low-temperature oxidation is
the primary source of heat release leading to the sponta-
neous ignition of stored coal. If the heat generated by oxida-
tion is not dissipated as fast as it is accumulated, then there
is an increase of temperature and a runaway ignition event
can ensue [1]. The ignition delay time can be large and can
easily run into days for coal stockpiles [2]. Once a re is
initiated in stored coal, the task of extinguishing the re can
be very difcult because of the amount of coal involved
(often hundred of tons).
The spontaneous ignition of coal poses serious problems
in its mining, transportation, storage and treatment [115].
Some major damage related to spontaneous ignition of coal
is summarised as follows: up to 10 res per year arise in the
German hard coal mines of the Ruhr area [3]; in China,
about 200 million tons per year is consumed by spontaneous
combustion of the coal when still in the ground [4]. Uncon-
trolled res remain a problem throughout the coalelds of
the USA [5]. Siberian deposits suffer from ignition in situ
despite the cold surface conditions and the recent forest res
in Indonesia have reportedly caused damage to surface peat
resources [6]. These events show the importance of studying
the self-heating behaviour of different coals, and in particu-
lar lower-rank coals, since they have a greater tendency to
spontaneous combustion.
For many reasons such as security of supply, economic
and environmental considerations, most users of coal
(including power generating companies, steel and cement
industries) prefer to have a number of suppliers. This has led
to a greater role of thermal coal as an internationally traded
fuel and it is common for power generating companies to
import coals from different world locations. But the safety
records of some of these coals is not good, and it is then
necessary to closely monitor the spontaneous combustion
behaviour of such coals in stockpiles and in coal milling
systems [7]. In the case of the emerging low sulphur
low rank coals, quality improvement by thermal drying
before shipment and utilisation is desirable. However,
drying can increase the natural spontaneous combustion
tendency of these coals [8], so the problem can be
exacerbated.
The reaction between coal and oxygen at low temperature
depends on many factors including temperature, particle
size, surface area, coal-pore structure, moisture content,
coal rank and the composition of ambient air. It is clear
that low temperature oxidation of coal is strongly dependent
on temperature, and that the rate of oxygen consumption
follows an Arrhenius-type law [9]. In fact, many workers
Fuel 79 (2000) 19511961
0016-2361/00/$ - see front matter 2000 Elsevier Science Ltd. All rights reserved.
PII: S0016-2361(00)00053-3
www.elsevier.com/locate/fuel
* Corresponding author. Tel.: 44-0113-233-2498; fax: 44-0113-
244-0572.
E-mail address: a.c.mcintosh@leeds.ac.uk (A.C. McIntosh).
1
Permanent address: Department of Mechanical Engineering, University
of Indonesia Kampus UI Depok, Jakarta 16424, Indonesia.
[6,1014] have applied the Arrhenius-type plot to determine
the values of the kinetic parameters for low temperature
oxidation of coal (i.e. the activation energy E
a
and the
pre-exponential factor A). Their ndings showed that for a
specic coal, the experimental values of E
a
and A are only
useful for the one coal, and different values of E
a
and A are
obtained for different coals.
The role of particle size, surface area and pore structure
on oxidation rate of coal has been indicated by several
authors [710,14,15]. The experimental ndings imply
that consumption of oxygen by coal at low temperatures is
governed by pore diffusion and chemical kinetics, depend-
ing on the intrinsic reactivity of the coal itself and the capa-
city of coal pores to transport oxygen [8,14,15]. In addition,
low internal surface area is considered as a major
contributor to the low reactivity of materials with high
carbon content [16].
Recently [6] the authors have tested the crossing point
method [6,13,17,18] for estimating the kinetic parameters of
oxidation of coal from the transient data of oven heating
experiments, where a coal sample was heated in a 50 mm
side length, cube basket at constant oven temperature. In
this paper, the crossing-point method is applied to investi-
gate the tendency of the lower and the higher rank coals and
blends of these same coals, to undergo self-heating as they
are heated in the oven heating experiments. The tendency to
self-heat was measured against factors such as critical ambi-
ent temperature (T
a crit
), the activation energy (E
a
) and the
product of the exothermicity and the pre-exponential factor
(QA) of the coals. The experiments were devised to
highlight the effect of particle size and the physical structure
of the coals including pore size and surface area.
2. Experimental
2.1. Coal selection and preparation
Four Indonesian coals ranging from lignite, sub-bitumi-
nous and bituminous coals were used in this work. They
were South Bangko, Tanjung Enim (South Sumatra),
Prima, and Pinang coals (East Kalimantan). The criteria
for the selection of coals was that they represented low-
sulphur coals (1%) and that the rank should span the
range from low to high rank coals with different self-heating
behaviours. All samples were air-dried for more than 24 h at
room temperature, having been crushed to a number of
separate size fractions. To examine the effect of drying,
some samples were de-moisturised in a nitrogen atmosphere
at 105C prior to the tests. The properties of the coals used
are given in Table 1. The values of the specic heat of coal
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1952
Nomenclature
A pre-exponential factor, s
1
C
ps
specic heat capacity of coal, J kg
1
K
1
d
p
average particle diameter, mm
E
a
activation energy, J mol
1
k thermal conductivity of coal, W m
1
K
1
l slab half-width, m
Q heat of oxidation, J kg
1
R universal gas constant 8.314 J mol
1
K
1
s surface area, m
2
g
1
t time, s
T temperature, K
T
a crit
critical ambient temperature, K
T
cp
crossing point temperature, K
T
i
local temperature of the sample, K
T
a
oven or ambient temperature, K
d
c
Frank-Kamenetskii parameter
r packing density of coal, kg m
3
Table 1
Properties of the coal samples
Properties Coal types (location)
S. Bangko coal (Sumatra) Tj.Enim coal (Sumatra) Prima coal (Kalimantan) Pinang coal (Kalimantan)
Proximate analysis (wt%)
Moisture 14.2 17.5 5.2 8.8
Volatile matter 43.7 39.7 46.8 39.2
Fixed carbon 37.5 39.1 49.1 47.5
Ash 4.6 3.7 2.9 4.5
Ultimate analysis (daf. wt%)
Nitrogen 1.25 1.05 1.76 1.55
Carbon 62.54 65.36 76.78 72.02
Hydrogen 4.20 4.51 5.50 4.99
Oxygen (by difference) 32.01 29.08 15.96 21.44
Caloric value (MJ kg
1
) 23.0 28.7 31.9 29.0
Thermal conductivity
(W m
1
K
1
)
0.13 0.14 0.11 0.11
Specic heat (J kg
1
K
1
) 1480 1440 1240 1310
Packing density (kg m
3
) 635 ^4 620 ^5 725 ^6 720 ^6
C
p
were estimated using a procedure developed by Melchior
and Luther [19] where a typical oven temperature of 130C
was chosen for calculation.
2.2. Pore size and surface area characterisation
For this low temperature oxidation work, surface area and
porosity measurements were carried out using a Quanta-
chrome Quantasorb QS-13 analyser using a N
2
/He mixture.
Nitrogen adsorption is conducted at 77 K, the normal boil-
ing point of N
2
. At this temperature, small micropores of
coal with radii less than 1 nm cannot be readily accessed.
The activation energy of diffusion in these micropores of
coal is too high for the adsorption to effectively reach equi-
librium, and there is also thermal shrinkage of the pores at
this temperature [20,21]. However, because the measure-
ment using the N
2
/He mixture relates to pores with radii
greater than 5 A

(excluding the area containing the micro-


pores), it gives a more representative surface area available
for reaction with molecular oxygen [22]. Smith et al. [21]
pointed out that although nitrogen does not adequately char-
acterise the total surface areas of coals and chars, which are
highly microporous, the method is valuable in measuring
the larger pores relevant to transport of oxygen during
combustion. In addition, Kaji and co-workers [14]
concluded that the small pores (100 A

) did not contribute


to the low-temperature oxidation reaction, indicating the
lack of micro-diffusion effects.
2.3. Activation energy and pre-exponential factor
estimations
The crossing-point method [6,13,17,18] was applied for
evaluating QA and E
a
from a plot of lndT=dt at T
cp
against
1/T
cp
, where T
cp
is the crossing point temperature when the
central temperature T
1
is the same as T
2
(an off-centre
temperature), and (dT/dt) is the corresponding rate of
change in temperature at this crossing point. The apparatus
required for the crossing-point method, consisted of a
temperature-controlled oven with a wire-mesh basket
made into a cube for holding the samples, with 50 mm
sides and an open top. The basket was suspended in the
re-circulating air oven which was capable of sustaining a
temperature up to 250C, whilst maintaining a stability of
^1C over the long period of the experiments. The stability
of the temperature and the airow during the experiments
was enhanced by putting the oven in a fume cupboard with a
constant airow. Five pre-calibrated thermocouples were
placed inside the 50 mm cubical basket to get the tempera-
ture proles of the sample. The rst point was the centre of
the basket, the second at 5 mm away from the centre, the
next 5 mm farther away, and then two farther ones 7.5 mm
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1953
Fig. 1. SEM micrographs of the coal samples: (A) Prima air-dried; (B) South Bangko air-dried; (C) Tanjung Enim coal; and (D) Tanjung Enim de-moisturised.
apart. All the thermocouples were aligned in a straight line.
The temperature prole obtained from the experiments was
used to dene an off-centre temperature where between this
and the central temperature, one could not get a peak occur-
ring. A sixth thermocouple was placed in the oven to
measure the ambient (oven) temperature (T
a
). Once the
off-centre temperature (T
2
) has been dened, i.e. 10 mm
away from the centre for a 50 mm side cube basket, only
two thermocouples were inserted into the basket for the next
experiments (T
1
and T
2
). Therefore, the effect of the inser-
tion of two thermocouples on the heat transfer situation
within the sample basket is insignicant. Having checked
the position of the thermocouples, the screws of the thermo-
couple carrier were fastened. In maintaining constant pack-
ing densities (Table 1) coal samples of known weight (for
example about 90 g of Pinang coal sample for a 125 cm
3
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1954
Fig. 2. (A) Cumulative pore volume for pore in the 1.530 nm ranges determined from N
2
adsorption isotherms for the South Bangko and the Prima coals. (B)
Surface area of coals at different particle size. The surface area of the coals with the particle diameters d
p
1:5 mm are as follows: the Prima coal s
0:36 m
2
g
1
; the Pinang coal s 4:71 m
2
g
1
; the South Bangko coal s 5:04 m
2
g
1
; and the Tanjung Enim coal s 6:29 m
2
g
1
:
cube basket) were poured carefully into the basket. At inter-
vals during this lling, the basket was gently tapped to
ensure that no air pockets developed around the tips of the
thermocouples. The thermocouples were connected to a
relay isolated multiplexer and amplier board, which was
then connected to an analogue to digital converter linked to
a computer.
In this paper, comparison is also made with the classical
Frank-Kamenetskii (F-K) approach [6,13,23,24] for evalu-
ating A and E
a
from a plot of lnd
c
T
2
a crit
=l
2
against 1/T
a crit,
where T
a crit
is the critical ambient temperature for ignition.
This classical approach required different basket sizes to be
used, e.g. 35, 40, 50, 60, 70 and 80 mm cube baskets.
3. Results and discussion
3.1. Physical structure of the coals
Micrographs for the coal samples obtained using the SEM
analyser are shown in Fig. 1AD. The surface of the hard
Prima coal particles were nearly at with only a few open
pores, while those of the softer South Bangko coal and
Tanjung Enim coals sample had more open pores. The
removal of water brings about the larger opening of pores
particularly for a high moisture coal such as Tanjung Enim
coal (Fig. 1D). This leads to a higher accessibility of oxygen
to reactive sites.
Fig. 2A shows the cumulative pore volume of the South
Bangko and the Prima coals from N
2
adsorption. It can be
seen that the South Bangko coal has a higher pore volume
than the Prima coal in the 1.530 nm ranges. The Prima coal
apparently has fewer meso-pores (radii between 1 and
20 nm) and macro-pores (radii larger than 20 nm) than the
South Bangko coal. This is in agreement with the general
trend of decreasing porosity with increasing rank.
The nitrogen BET surface area is used for characterising
the surface area (s) available for low temperature oxidation
reaction of coals. The surface areas for individual coals at
different particle sizes are given by a normalisation value,
which is dened as:
Norm s
s
i
s
1
100% 1
where s
1
is the surface area of the smallest particle size used.
The results are presented in Fig. 2B.
Fig. 2B clearly shows two distinct trends, for lower and
higher rank coals. For the lower rank coals (South Bangko/
Tanjung Enim), there is no change in surface area with
particle size. This is due to the porous structure of these
coals which mainly comprise of meso and macro-pores
(see Fig. 2A). There are, however, completely different
trends for the higher rank coals (Prima/Pinang). Since
these hard coals have only a few open pores, for a reduction
of particle sizes the surface area increases substantially
(almost ve fold for Prima coal, see Fig. 2B).
3.2. Self-heating phenomena
The measurement of local temperatures within the sample
is conducted by immersing thermocouples at ve points on
the centre line of the sample basket where coal particles are
packed to a certain density. The tip of the thermocouples is
considered to have excellent physical contact with the
sample particles. A closer investigation of the temperature
uctuation during these oven heating tests shows that the
oven temperature uctuates by about ^0.8 K, and that the
local temperatures (represented by T
1
and T
2
) seem to have
better stability of about ^0.4 K. This is because the local
temperatures are less affected by any instability in the oven
temperature due to fairly high thermal capacity of the coal.
This nding is important for the crossing-point method,
which relies on the stability of the local temperatures (T
1
and T
2
) for the determination of the crossing-point tempera-
ture T
cp
.
From the spatial temperature prole, it is clear that the
temperatures inside the basket rise following the increase of
surface temperature (close to oven temperature). After a
certain time, a at temperature prole through the centre
line of the basket develops, at most, a few degrees above
the oven temperature. Whether sub-critical (no ignition) or
supercritical (leading to ignition), it was found that the
central temperature continued to rise above the other local
temperatures in all the runs we made. This evidence
suggests that in self-heating conditions, the coal ignition
starts from the centre of the bed involving the low tempera-
ture slow oxidation reaction as the main reaction. However,
the situation is different under greatly supercritical condi-
tions, where the at temperature prole does not develop,
and the central temperature is continuously below the other
local temperatures. The critical ambient temperature T
a crit
for spontaneous ignition to occur is dened very sharply by
these marked changes in sub-critical and supercritical beha-
viour, obtained from the central temperaturetime observa-
tions [6]. The results for T
a crit
involving all coals are given
altogether in the following section, when discussing the
effect of particle size.
3.3. Kinetic oxidation parameterscomparison with F-K
The results for the Prima (air-dried) coal with an average
particle diameter d
p
0:18 mm and a density r
725 ^6 kg m
3
are shown in Fig. 3A. In Fig. 3A the
main line represents the results from the crossing-point
method, and the dotted line indicates the results from the
classical F-K approach. Horizontal and vertical error bars on
the main line are due to the uncertainty in the crossing-point
temperature T
cp
(^0.5 K), and in the corresponding rate of
change in temperature at this crossing point (dT/dt) at T
cp
(^0.04 K min
1
). A slight increase of the vertical error is
noticed, following a decrease of the crossing point tempera-
ture. However, the regression analysis agrees within the
error bars. For the F-K approach (the dotted line), the
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1955
horizontal and the vertical error bars represent the uncer-
tainty of measured oven temperature owing to the uncertain-
ties in the quantity lnd
c
T
2
a crit
=l
2
and the critical ambient
temperature T
a crit
(^2 K). The activation energy E
a
is
calculated to be 90 ^4 kJ mol
1
using regression analysis
and the product of the exothermicity and the pre-exponen-
tial factor, that is QA, is estimated as 4:48 10
12
J kg
1
s
1
:
The F-K approach gives an E
a
value of 90 ^5 kJ mol
1
and
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1956
Fig. 3. Plot of lndT=dtT
cp
against (1/T
cp
). (A) For Prima coal (air-dried) with an average particle diameter d
p
0:18 mm and a density r 725 ^6 kg m
3
:
The plots show the activation energies and the reactivity calculated from the crossing point method [6,13,17,18] and with the classical F-K approach
[6,13,23,24]. (B) For Pinang coal with an average particle diameter d
p
1:18 mm and a density r 720 ^6 kg m
3
:
QA 1:88 10
12
J kg
1
s
1
: The good agreement of the
two methods for estimating E
a
and QA conrms the ability
of the crossing point method to accurately measure the
kinetic parameters for low temperature oxidation. There is
another approach using the oven temperature (called the
heat release technique [11,12]), but we agree with Chen
[18] that the crossing point method gives better accuracy.
We now, consider the effect of basket size on the
measured values of the kinetic oxidation parameters. The
kinetic parameter for the oxidation of Pinang coal with d
p

1:18 mm and density r 720 ^6 kg m
3
are measured
using different basket sizes. The crossing-point method
allows the transient results from sub- and just super-critical
conditions to be used. Fig. 3B illustrates this, since the rst
two points on the graph are in fact for super-critical condi-
tions. It is clear that they are an extension of the line-t to
the sub-critical data. From the regression analysis, it is
evident that the value of E
a
is not affected by the size of
the basket used. However, as the smaller basket sizes shift
the critical ambient temperatures to higher levels, there is
also a slight reduction in QA during the early stage of self-
heating, before the central temperature crosses T
cp
(see the
two points for 40 mm basket on Fig. 3B). The crossing point
method gives an E
a
value of 57 ^4 kJ mol
1
and a QA
value of 2:25 10
8
J kg
1
s
1
for Pinang coal.
In general the crossing-point method is more convenient
to use than the F-Kapproach, and yet sufciently reliable for
the determination of the activation energy and the heat
release times reactivity (QA). The results for other particle
sizes and coal types were measured using this method.
3.4. Effect of particle size
Shown in Fig. 4A is the effect of particle size on the
critical ambient temperature T
a crit
obtained from the oven
heating tests using a 50 mm basket size. It can be seen that
T
a crit
for all coals increases with the particle size. The rate of
increase of T
a crit
, which is determined under a constant
basket size and uniform density for different particle size,
varies, depending on the particle size and physical structure
of the coals. Fig. 4A shows that the effect of the particle size
is very signicant on the Prima coal, moderate on Pinang
coal and almost negligible on Tanjung Enim coal and South
Bangko coal. This is due to the inuence of the surface area
of the coal on the balance between the rate of heat genera-
tion and the rate of heat loss to the surroundings. As shown
in Fig. 2B the nitrogen surface area of the higher rank coal
(Prima) has increased by almost vefold for average particle
diameters d
p
1:5 mm to d
p
0:06 mm: Whilst the
surface area of the lower rank coals is independent of the
particle size.
In a given experimental condition, an increase in the
critical ignition temperature with particle size could also
be attributed to the temperature gradients within the bed
and within the particle [25]. The spatial temperature prole
suggested that ignition occurs after the temperature within
the particle has equilibrated (i.e. a at temperature prole
has been achieved) and that larger particles have steeper
gradients and take a longer time for equilibration, resulting
in longer ignition delay and thus a higher ignition tempera-
ture.
The values of QA and E
a
for different particle sizes are
shown in Fig. 4B and C. It is clear that the value of QA and
E
a
of the coals are signicantly different, with generally a
higher E
a
corresponding to a higher QA (there is a possible
correlation between E
a
and QA which is explored elsewhere
[26,27]). The activation energy may be regarded as a
measure of the energy barrier to reaction [28]. From this
alone one would expect there to be a higher T
a crit
for these
smaller coal particles. However, it is the QA that is all-
important. For smaller particles, the effectiveness factor
for heat transfer (dened as the ratio of the actual heat
transfer rate to the maximum possible heat transfer rate) is
closer to one, leading to greater oxidation rates per unit
volume. For given airow and physical properties, the
mass transfer coefcient increases as particle size decreases
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1957
Fig. 4. Effect of particle size on critical ambient temperature (A), reactivity
(B) and activation energy (C) for Prima, Pinang, South Bangko and
Tanjung Enim coals.
and, in conjunction with higher specic surface area, leads
to greater rates of moisture depletion [29]. The higher effec-
tiveness factors and mass transfer coefcients for smaller
particles are reected by the higher QA as shown in Fig. 4B.
In this gure it is evident that for the mid-range size with a
particle diameter d
p
1:5 mm; the more reactive coal (e.g.
South Bangko), with a lower T
a crit
has a higher E
a
than those
of higher rank coals (e.g. Prima coal which has a higher
T
a crit
). In general, Fig. 4B and C show an overall trend of
lower critical ambient temperatures giving higher oxidation
reactivity and activation energy. However, the inuence of
particle size on pre-exponential factor is more prominent
than on the activation energy (compare Fig. 4Bgraph of
ln QA to Fig. 4Cgraph of E
a
). It is evident that the effect
of particle size and surface area of the coals on their self-
heating behaviour is very signicant for the higher rank
coals and almost negligible for the lower rank coals.
By considering the critical ambient temperature T
a crit
at
different particle sizes, it is observed that the oxygen content
of the coal can have a strong inuence on the self-heating
behaviour of these coals. It is seen from Fig. 5 that in
general T
a crit
decreases with the increase of oxygen percen-
tage in the coal. However, the effect of coal oxygen on T
a crit
becomes less signicant for ner particles. The results indi-
cate that the reason for this effect is probably due to the fact
that meso and macro-pores are responsible for the low-
temperature oxidation reactions, and that gaseous oxygen
is not able to penetrate the micro-pores.
3.5. Effect of coal blending
For all experiments under self-heating conditions, igni-
tion began in the centre of the bed involving the low
temperature oxidation reaction between the coal and oxygen
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1958
Fig. 5. Critical ambient temperature against oxygen in coal. The critical ambient temperatures of four different particle sizes were determined using a 50 mm
cubical basket.
Fig. 6. Effect of size fraction on the critical ambient temperatures of coals at different mixture ratio.
in the coal and from the air. A coal bed of coarse particles
may allow a higher airow rate into the centre. However,
the effective surface area encountered with coal of large size
fraction, (particularly the higher rank coal) is low. Oxygen
consumption with consequent heat build-up is therefore
poor and is likely to be dissipated away with the high rate
of air owoccurring within the inner coal layer. If, however,
the coarse coals are mixed with ne ones, the larger surface
area available for the presence of ne particles can promote
oxidation with the consequent higher heat generation rate at
the centre of the bed. This is shown in Fig. 6 where coarse
particles d
p
2:67 mm are mixed with ne particles d
p

0:06 mm: The gure shows that for higher rank coals with
few open pores in its structure (Fig. 1A) and with a rela-
tively low oxygen content in the coal (Fig. 5), the presence
of ne particles drastically decreases the critical ambient
temperature needed for ignition. Therefore, coal beds with
mixed up sizes of particles are more vulnerable than those
with segregated sizes. In practice, stockpiling of coals with
mixed particle sizes should be avoided.
The signicance of particle size effects on the sponta-
neous ignition of coals is not only apparent for the case of
higher rank coals, but also for lower rank coals as mentioned
earlier. Therefore, in attempts to investigate the effect of
coal blending, it is important to choose an appropriate parti-
cle size of the coals. In this study a mid range value with an
average particle diameter of 1.5 mm was used since at this
particle size, the activation energy E
a
, the product of the
exothermicity and the pre-exponential factor QA and the
critical ambient temperature (T
a crit
) of the lower rank
(South Bangko) coal are: E
a
89 ^3 kJ mol
1
; QA
2:76 10
12
J kg
1
s
1
; T
a crit
402 K; but the values for
higher rank (Prima) coal are: E
a
51 ^3 mol
1
; QA
1:12 10
7
J kg
1
s
1
; T
a crit
470 K and the values for
Pinang coal are: E
a
53 ^2 kJ mol
1
; QA 4:35
10
7
J kg
1
s
1
; T
a crit
425 K: Thus the values for lower
rank coal are very different to those of higher rank coals.
Fig. 7A shows the plots of temperature (measured at the
centre of sample basket) against time for the individual
coals and of their blends. This gure suggests that the
blend reects the properties of the individual coals. In addi-
tion, the curve representing the corresponding rate of
change in central temperature dT/dt of a 1:1 blend is located
in between those pertaining to the two individual compo-
nents of the blend. For this blend, the curve closely follows
the midpoints of the parent coals (Fig. 7B).
Fig. 8AC shows the plots of T
a crit
, QA, and E
a
,
respectively, obtained for blends of Prima and Pinang
coals with South Bangko coal. As expected, the
presence of South Bangko coal in the mixtures
decreases the critical ambient temperatures of the blends
(Fig. 8A). The higher QA value of the lower rank coal
leads to a sharper increase in particle temperature, and
this shortens the ignition delay time. Both QA, and E
a
increase with the percentage of the more reactive coal
(the South Bangko coal) in the blend. It is clear that
the increase of E
a
is not linear andFig. 8Bshows that the values
of QA increase almost exponentially (notice the log scale
on the vertical axis). Therefore, in terms of the heat release
rate, the inuence of the pre-exponential factor is more
prominent than the activation energy (compare Fig. 8B
graph of ln QA to Fig. 8Cgraph of E
a
). This reects the
inuence of the pore structure and surface area of the lower
rank coal on the effective rate of oxygen diffusion, the
oxidation of the coal particles and the reactivity of blended
coals. Fig. 2A and B conrm this, since the lower rank coal
has relatively a much higher pore volume and larger
surface area for the particle diameter concerned d
p

1:5 mm:
The presence of South Bangko coal also decreases the
critical ambient temperatures as shown in Fig. 8A.
Higher reactivity leads to a sharper increase in particle
temperature and this shortens the ignition delay time
[29]. These compounding factors suggest that the
presence of a more reactive coal enhances the potential
for the spontaneous combustion of blended coals by
increasing the rate of heat generation and lowering the
critical ambient temperature.
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1959
Fig. 7. Comparison of temperaturetime curve (A) and dT/dt curve (B)
obtained with South Bangko coal, Pinang coal and their blend at ambient
temperature, T
a
122C and d
p
1:5 mm:
4. Conclusions
Early work on the crossing-point method has been
extended to investigate the effect of particle size and the
physical structure of coal, including pore size and surface
area, on the self-heating character of single and blended
coals. The physical structure of coal determines the varia-
tion of surface area with particle size. For higher rank coals,
which have at surfaces and fewer meso and macro-pores,
the increase of surface area with particle size is very
obvious, but it is almost negligible for lower rank coals
(porous structure).
It is shown that particle size has considerable inuence
on the self-heating character of coal. On the one hand, a
smaller particle reduces the critical ambient temperature
for spontaneous ignition to occur. A typical critical ambi-
ent temperature is about 400 K for all coals at an average
particle diameter of 0.06 mm. On the other hand, the
product of the exothermicity and the pre-exponential
factor QA and the activation energy E
a
of the coals
increases with decreasing particle size.
The change of the critical ambient temperature with
particle size is almost negligible for porous coals
(South Bangko and the Tanjung Enim), but signicant
for hard coals (Pinang and Prima). It was observed that
nely crushed (pulverised) bituminous coals could exhi-
bit a strong tendency to spontaneous combustion similar
to those of low rank coals.
A coal bed with mixed up sizes is much more vulnerable
than one with segregated sizes. The effect is very non-
linear. That is one only needs a small amount of nely
crushed particle and/or strongly reactive coal to have a
large reduction on the critical ambient temperature. This
means that one must carefully consider the particle size
distribution to judge the spontaneous ignition behaviour
of coal. This has serious implications for case of coal
blending in industry.
The higher surface area found in the low rank coals is
thought to be one of the inuential factors for the increase
in QA and the decrease in T
a crit
with increasing low rank
coal content in the blend.
References
[1] Bowes PC. Self-heating: evaluating and controlling the hazards,
HMSO, 1984.
[2] Krishnaswamy S, Agarwal PK, Gunn RD. Low-temperature oxidation
of coal-modelling spontaneous combustion in coal stockpiles. Fuel
1996;75:35362.
[3] Pilarczyk E, Leonhardt P, Wanzl W. Characterisation of coals with
respect to their self-ignition tendency. Eighth International Confer-
ence on Coal Science, 1995. p. 497500.
[4] International Institute for Aerospace Survey and Earth Sciences (ITC)
homepage [http://www.itc.nl/~coalre/index.html#prob].
[5] Walker S. Uncontrolled res in coal and coal wastes, IEACR-London,
1999.
[6] Nugroho YS, McIntosh AC, Gibbs BM. Using the crossing-point
method to assess the self-heating behaviour of Indonesian coals.
Twenty-seventh Symposium (International) on Combustion, The
Combustion Institute, Pittsburgh, 1998. p. 29819.
[7] Ren TX, Edwards JS, Clarke D. Adiabatic oxidation study on the
propensity of pulverised coals to spontaneous combustion. Fuel
1999;78:161120.
[8] Carras JN, Young BC. Self-heating of coal and related materials:
model, application and test methods. Progress in Energy and Combus-
tion Science 1994;20:115.
[9] Krishnaswamy S, Bhat S, Gunn RD, Agarwal PK. Low-temperature
oxidation of coala single-particle reactiondiffusion model. Fuel
1996;78:33343.
[10] Petarca L, Tognotti L, Zanelli S, Bertozzi G. An experimental study
on oxidation of coal at low temperature. Twenty-rst Symposium
(International) on Combustion, The Combustion Institute, Pittsburgh,
1988. p. 193201.
[11] Jones JC, Chiz PS, Koh R, MatthewJ. Kinetic parameters of oxidation
of bituminous coals from heat-release rate measurements. Fuel
1996;75:17557.
[12] Jones JC, Henderson KP, Littlefair J, Rennie S. Kinetic parameters of
oxidation of coals by heat-release measurement and their relevance to
self-heating tests. Fuel 1998;77:1922.
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1960
Fig. 8. Critical ambient temperatures (A), reactivities (B) and activation
energies of the blended coals (C) at different blending ratio. Particle
diameter d
p
1:5 mm:
[13] Sujanti W, Zhang D-K, Chen XD. Low-temperature oxidation of coal
studied using wire-mesh reactors with both steady-state and transient
methods. Combustion and Flame 1999;117:64651.
[14] Kaji R, Hishinuma Y, Nakamura Y. Low temperature oxidation of
coals: effects of pore structure and coal composition. Fuel
1985;64:297302.
[15] Wang H, Dlugogorski BZ, Kennedy EM. Experimental study on low-
temperature oxidation of an Australian coal. Energy & Fuels
1999;13:11739.
[16] Hurt RH. Structure, properties, and reactivity of solid fuels. Twenty-
seventh Symposium (International) on Combustion, The Combustion
Institute, Pittsburgh, 1998. p. 2887904.
[17] Chen XD, Chong LV. Some characteristics of transient self-heating
inside an exothermically reactive porous solid slab. Transaction of
Institution of Chemical Engineers 1995;73 (Part B):1017.
[18] Chen XD. On basket heating methods for obtaining exothermic reac-
tivity of solid materials: the extent and impact of the departure of the
crossing point temperature from the oven temperature. Transaction of
Institution of Chemical Engineers 1999;77 (Part B):18792.
[19] Melchior E, Luther H. Measurement of true specic heats of bitumi-
nous coals of different rank, and of a high-temperature coke, in the
temperature range 30350C. Fuel 1982;61:10719.
[20] White WE, Bartholomew CH, Hecker WC, Smith DM. Changes in
surface area, pore structure and density during formation of high
temperature chars from representative US coals. Adsorption Science
and Technology 1991;7:180209.
[21] Smith KL, Smoot LD, Fletcher TH, Pugmire RJ. The structure and
reaction processes of coal. New York: Plenum Press, 1994.
[22] Hardgrave G, Pourkashanian M, Williams A. The combustion and
gasication of coke and coal chars. Twenty-rst Symposium (Inter-
national) on Combustion, The Combustion Institute, Pittsburgh, 1988.
p. 22130.
[23] Frank-Kamenetskii DA. Diffusion and heat exchange in chemical
kinetics. 2nd ed. New York: Plenum Press, 1969.
[24] Gray BF, Grifths JF, Hasko SM. Spontaneous ignition hazard
in stockpiles of cellulosic materials: criteria for safe storage.
Journal of Chemical Technology and Biotechnology 1984;
34A:45363.
[25] Katalambula H, Hayashi J, Chiba T. Dependence of single coal parti-
cle ignition mechanism on the surrounding volatile matter cloud.
Energy and Fuels 1997;11:10339.
[26] Nugroho YS. McIntosh AC, Gibbs BM. On the prediction of thermal
runaway of coal piles of differing dimension by using a correlation
between heat release and activation energy. Twenty-eighth Sympo-
sium (International) on Combustion, The Combustion Institute, 2000,
accepted for publication.
[27] Nugroho YS. Self ignition of Indonesian coal, PhD thesis, Leeds
University, 2000, submitted.
[28] Grifths JF, Barnard JA. Flame and combustion. 3rd ed. London:
Blackie Academic, 1995.
[29] Bhat S, Argawal PK. The effect of moisture condensation on the
spontaneous combustibility of coal. Fuel 1996;75:152332.
Y.S. Nugroho et al. / Fuel 79 (2000) 19511961 1961

Вам также может понравиться