ArticlePDF Available

Mechanisms of the central mode particle formation during pulverized coal combustion

Authors:

Abstract and Figures

Ash particles produced from pulverized coal combustion are considered to be tri-modally distributed. These include the well-known ultrafine and coarse modes, and a central mode that is less reported but attracts increasing attention. This work presents a preliminary study on the formation mechanisms of the central mode particles during pulverized coal combustion. Experiments of four sized and density-separated coal samples were carried out in a laboratory drop-tube furnace under various controlled conditions. Experimental data show that the ash particle size distributions have an evident central mode at ∼4μm for all coal samples. Increasing combustion temperature leads to an increase in the central mode particle formation, which is thought to be due to enhanced char fragmentation. The small-size coal sample produces a larger amount of the central mode particles, reasonably due to abundant fine particles in the parent coal sample. Under similar combustion conditions, both the Heavy (>2.0g/cm3) and Light (
Content may be subject to copyright.
This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
Mechanisms of the central mode particle
formation during pulverized coal combustion
Dunxi Yu, Minghou Xu
*
, Hong Yao, Xiaowei Liu,
Ke Zhou, Lin Li, Chang Wen
State Key Laboratory of Coal Combustion, Huazhong University of Science and Technology,
1037 Luoyu Road, Wuhan, Hubei 430074, China
Abstract
Ash particles produced from pulverized coal combustion are considered to be tri-modally distributed.
These include the well-known ultrafine and coarse modes, and a central mode that is less reported but
attracts increasing attention. This work presents a preliminary study on the formation mechanisms of
the central mode particles during pulverized coal combustion. Experiments of four sized and density-sep-
arated coal samples were carried out in a laboratory drop-tube furnace under various controlled condi-
tions. Experimental data show that the ash particle size distributions have an evident central mode at
4lm for all coal samples. Increasing combustion temperature leads to an increase in the central mode
particle formation, which is thought to be due to enhanced char fragmentation. The small-size coal sample
produces a larger amount of the central mode particles, reasonably due to abundant fine particles in the
parent coal sample. Under similar combustion conditions, both the Heavy (>2.0 g/cm
3
) and Light
(<1.4 g/cm
3
) coal fractions produce a central mode, indicating that not only the included minerals but also
the excluded minerals contribute to the formation of the central mode particles.
Ó2009 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
Keywords: Coal combustion; Formation mechanism; Central particle mode; Fragmentation
1. Introduction
Particulate matter (PM) from pulverized coal
combustion (PCC) is conventionally considered
to be bimodally distributed, including an ultrafine
mode (0.1 lm) formed primarily by the vapori-
zation–condensation mechanism [1–4] and a
coarse mode (1–20 lm) produced mainly by the
coalescence of molten ash droplets and char frag-
mentation [5–7]. However, some experimental and
modeling studies also reported an additional
central particle mode (0.3–5 lm) between the
commonly observed two modes [5,8–17]. The cen-
tral mode contributes a large portion to PM
2.5
(PM with an aerodynamic diameter of less than
2.5 lm) and has some practical implications. For
example, the central mode particles may have sub-
stantial influence on gas-particle mass transfer
and be responsible for a significant fraction of het-
erogeneous Hg
0
transformation within the electro-
static precipitator (ESP) [18]. Furthermore, these
particles seem to have higher surface-to-volume
ratios than typical smooth spheres [11,17] and
contain a larger mass fraction of some transition
metals [9] and trace elements [11,19,20] than
coarse mode particles. Once suspended in the
1540-7489/$ - see front matter Ó2009 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.proci.2008.07.037
*
Corresponding author. Fax: +86 27 87545526.
E-mail address: mhxu@mail.hust.edu.cn (M. Xu).
Available online at www.sciencedirect.com
Proceedings of the Combustion Institute 32 (2009) 2075–2082
www.elsevier.com/locate/proci
Proceedings
of the
Combustion
Institute
Author's personal copy
atmosphere and inhaled by humans, the particles
in the central mode size range can be retained in
the lung [21] and are expected to cause adverse
health effects.
So far, the origins of the central particle mode
have not been made as clear as those of the ultra-
fine and coarse modes. Therefore, sufficient
knowledge of the formation of these particles is
essential to the understanding of the effects of fly
ash on toxicant partitioning and the development
of advanced technologies for PM
2.5
emission con-
trol. The central mode particles have been implied
to be formed through char fragmentation
[5,13,15,22,23], ash bursting [24] and/or shedding
[8,11]. Fine particles [25,26] and excluded minerals
[27] in the raw coal are also suggested to be possi-
ble sources. Collision-induced fragmentation is
suspected to contribute as well. However, none
of them has been demonstrated by purposely
designed experiments.
This work is intended to explore several possi-
ble mechanisms contributing to the formation of
the central particle mode by burning pretreated
coal samples from a bituminous coal in a well-
designed drop tube furnace (DTF). Due to diffi-
culties in measurements, ash bursting and/or
shedding are not covered presently. In addition,
the role of collision-induced fragmentation is not
evaluated since there are insufficient particles in
the DTF for this purpose. Three possible origins,
including char fragmentation, transformation of
fine particles and excluded minerals, are investi-
gated under well-controlled conditions. Combus-
tion temperatures have significant effects on char
structures and hence particle fragmentation [5].
To verify the contribution of char fragmentation
to the central particle mode, one size-classified
coal sample was burnt at different temperatures.
Fine particles in the raw coal are also suggested
to be another source of the central mode [25].
To validate this, two size-classified coal samples,
i.e. the fine and coarse size fractions, were com-
busted under the same conditions. In order to
evaluate the role of excluded minerals, the raw
coal was density separated by using heavy liquids.
The light and heavy density fractions were burnt
under the same conditions.
2. Experimental
2.1. Coal sample preparation and their character-
istics
A Chinese bituminous coal (Pingdingshan) was
crushed, ground and separated into two batches
for subsequent sample preparations. One batch
was sieved into two size fractions, i.e., <63 and
100–200 lm (hereafter named as Fine and Coarse,
respectively). The other batch was further sepa-
rated into two density fractions using heavy
liquids, i.e., <1.4 and >2.0 g/cm
3
(hereafter named
as Light and Heavy, respectively). This method is
conventionally used to isolate excluded minerals
from the raw coal.
The proximate and ultimate analyses of the
coal samples are listed in Table 1. The data show
that particle sizing seems to have little effect on
coal properties, but density separation has a sig-
nificant influence. The Light coal fraction has a
very low ash content of 4.4% and is expected to
contain only fine included minerals. In contrast,
the Heavy coal fraction has a much higher ash
content of 79.5% and should mainly contain
excluded minerals. For further illustration, a small
amount of the density fractions was mounted in
an epoxy resin, polished, carbon coated, and then
examined by a Sirion 200 field emission scanning
electron microscope (SEM). SEM micrographs
in Fig. 1 do show that mineral particles in the
Light coal fraction (Fig. 1a) are fine and domi-
nantly included in nature, while those in the
Heavy coal fraction (Fig. 1b) are coarse and
mainly excluded in nature.
The particle size distributions (PSDs) of the
four coal samples, the low temperature ash
(LTA) of the Heavy coal fraction, as well as char
samples prepared (see next section) were analyzed
using a Malvern particle-size analyzer (Model:
MAM 5004).
2.2. Char and PM sample preparation and char-
acterization
A high temperature drop-tube furnace (DTF)
was used to prepare PM samples from the com-
bustion of the four coal samples. To help explain
the role of char properties in the central mode for-
mation, pyrolysis experiments were also carried
out in the N
2
atmosphere. The obtained char sam-
ple is assumed to be comparable to that formed at
the devolatilization stage of coal combustion.
Table 1
Characteristics of coal fractions
Size (lm) Density (g/cm
3
)
Fine Coarse Light Heavy
<63 100–200 <1.4 >2.0
Proximate analysis
a
% (wt, dry)
VM 41.8 36.4 27.3 16.5
FC 39.6 44.7 68.3 4.0
A 18.6 18.9 4.4 79.5
Ultimate analysis % (wt, dry)
C 69.5 67.8 81.3 6.7
H 5.2 5.4 5.4 0.5
N 1.1 1.1 1.4 0.2
S 3.9 3.9 3.3 7.2
O
b
1.7 2.9 4.2 5.9
a
VM, volatile matter; FC, fixed carbon; A, ash.
b
By difference.
2076 D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082
Author's personal copy
Details of the DTF system have been pro-
vided elsewhere [16,28]. Briefly, coal particles
were fed via a Model MFEV-10 Micro Feeder
(Sankyo Piotech Co., Ltd.) and introduced into
a laminar flow tube reactor through a water-
cooled injection probe. The tube reactor has a
diameter of 56 mm and a length of 2000 mm
and is heated electrically. PM or char samples
were extracted isokinetically via a water-cooled
sampling probe and quenched immediately by
a stream of nitrogen at the inlet of the probe
to prevent further reactions. For PM collection,
a Dekati cyclone (Model SAC-65) was first used
to remove particles larger than 10 lm. The
entrained particles less than 10 lm were then
size-segregated by a Dekati low pressure impac-
tor (DLPI) into 13 fractions. More information
can be found elsewhere [16,28]. For char collec-
tion, glass fiber filters were used and the proce-
dure has been detailed previously [29].
A coal feeding rate of about 0.2 g/min was
used in all experiments. The particle residence
time in the heated zone of the reactor was esti-
mated to be 1–2 s. Pyrolysis and combustion
experiments of the Fine coal fraction were con-
ducted at 1373 and 1673 K, while those of the
other coal samples were carried out at 1673 K.
Pyrolysis experiments were completed in the N
2
atmosphere. Combustion experiments were car-
ried out in a 50% O
2
in N
2
mixture to ensure com-
plete combustion. Loss-on-ignition (LOI)
measurements indicated that negligible organic
matter was detected and PM samples consisted
almost entirely of inorganic matter. The porosity
and pore size distributions of all char samples
were analyzed using a Micromeritics ASAP 2000
Analyser. Some char samples were prepared and
observed in the above-mentioned SEM. PM sam-
ple collected on each DLPI stage was weighed
using a Sartorius M2P Microbalance (0.001 mg).
3. Results and discussion
3.1. The modality of particle mass fraction size
distributions
Differential PM mass fraction size distributions
for the sized and density-separated coal fractions
are plotted in Figs. 2 and 3, respectively. Three
distinct modes are clearly observed for all coal
fractions. An ultrafine mode occurs at 0.1 lm
and a coarse mode beyond 6 lm. Besides the
two commonly reported particle modes, a central
mode is also observed at 4lm for all samples.
This further demonstrates the tri-modal nature
of coal fly ash particle size distributions, as sug-
gested previously [5,8–13,15,17]. Since these size
distributions are based on mass fraction rather
than mass concentration, they should not be used
to indicate relative changes in particle modes for
different cases. The formation of the ultrafine
and the coarse modes are believed to be due to
Fig. 1. SEM micrographs of density fractions: (a) Light, <1.4 g/cm
3
; (b) Heavy, >2.0 g/cm
3
.
Fig. 2. PM mass fraction size distributions for sized coal
fractions.
D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082 2077
Author's personal copy
vaporization–condensation [3,6,7] and char frag-
mentation [22,23], respectively. However, the for-
mation mechanisms of the central mode particles
are still not clear. For the convenience of the fol-
lowing discussion, the three observed modes are
separated by the transition point from one mode
to another [12], as shown by the vertical lines in
Figs. 2 and 3. The central mode particles are
defined as particles in 0.5–6 lm and their mass is
estimated as the sum of particle mass collected
on impactor stages 7–11. The central mode con-
centration (conc.) used in related figures is repre-
sented as the mass of the central mode particles
produced per gram of coal (mg/g_coal).
3.2. Formation of the central particle mode
3.2.1. Char fragmentation
Char fragmentation has been shown to control
the formation of residual ash particles [22,23,30–
32]. However, there are limited reports emphasizing
its role in the formation of the central mode. Several
studies [5,13,15,22,23] have suggested that char
fragmentation most likely contributes to the central
mode, which has not yet been illustrated in detail by
experiments. To justify the role of char fragmenta-
tion, the Fine coal fraction (<63 lm) was com-
busted in the DTF at 1373 and 1673 K,
respectively. The concentrations of the generated
central mode are compared in Fig. 4. It is clear that
the central mode concentration measured at
1673 K is much higher, approximately 4 times that
at 1373 K. This shows that the higher temperature
leads to an increase in the central mode concentra-
tion. Such phenomenon is most possibly due to dif-
ferent fragmentation patterns during char
combustion.
It has been shown that char particle size [23]
and the porosity of macropores (>50 nm) [13,22]
are important variables in controlling the extent
of fragmentation. Figure 5 compares the average
particle size of all coal fractions and their resul-
tant char samples. The two char samples prepared
from the Fine coal fraction have a much larger
average particle size than their parent coal sample,
suggesting significant swelling during coal pyroly-
sis. They generally contain a large number of par-
ticles with structures as shown in Fig. 6. These
char particles are primarily spherical in shape
and have a large central void surrounded by a
very thin shell with a non-uniform distribution
Fig. 3. PM mass fraction size distributions for density-
separated coal fractions. Fig. 4. Central mode concentrations for the Fine coal
fraction at 1373 and 1673 K.
Fig. 5. Average particle size of all coal and char
samples.
Fig. 6. SEM micrographs of char cenospheres: (a)
surface; (b) cross-section.
2078 D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082
Author's personal copy
of macropores, often known as cenospheres. This
type of char particles is believed to fragment
extensively during combustion [5,8,11,13,22,23,
32] and possibly contributes to the formation of
the central mode.
As shown in Fig. 5, the 1673 K char prepared
from the Fine coal fraction has a larger average
particle size than the 1373 K char, indicating more
significant swelling at 1673 K. Though quantifica-
tion is not available, the 1673 K char is observed
to qualitatively contain more cenospheres under
SEM. As a consequence, the 1673 K char pre-
pared from the Fine coal fraction has a higher
porosity of macropores (>50 nm) as shown in
Fig. 7, nearly twice as much as the 1373 K char.
Baxter [23] found that the extent of char fragmen-
tation was strongly dependent on particle size and
coal rank. Larger bituminous char particles frag-
mented more extensively and generated much
more fine ash particles. Both experimental and
modeling results [5,22] showed that when the
macroporosity was increased, char fragmentation
also increased and more fine particles were pro-
duced. Therefore, the higher concentration of
the central mode at 1673 K (Fig. 4) is believed
to be primarily due to more extensive fragmenta-
tion of char particles caused by their larger size
and higher macroporosity. These results show
that char fragmentation is an important contribu-
tor to the central particle mode.
3.2.2. Contribution of original fine particles
Relative to char fragmentation, the role of fine
particles in the raw coal in the central mode forma-
tion has received much less concern. But the utility-
grind coal generally has a wide size distribution
and often contains a large number of fine particles,
which are also expected to contribute to the central
mode after combustion. The lack of such informa-
tion is probably due to the unavailability of a quan-
titative measurement technique [25] or narrowly
size-classified samples used [22].
To examine the contribution of the original
fine particles, the Coarse coal fraction (100–
200 lm) assumed to have eliminated the effects
of fine particles was also combusted in the DTF
at 1673 K. The obtained central mode concentra-
tion for the Coarse coal fraction is compared in
Fig. 8 with that for the Fine coal fraction com-
busted at 1673 K. It is evident that the central
mode for the Fine coal fraction has a much higher
concentration, more than 3 times that for the
Coarse coal fraction. This demonstrates that the
Fine coal fraction enriched in fine particles con-
tributes more to the central mode than the Coarse
coal fraction depleted of fine particles. Kramlich
and Newton [14] separated both the raw and
cleaned coals into different size cuts and measured
their resultant ash size distributions. A central
mode at approximately 2 lm was observed for
all cases. The more important result was that the
mode height for the smaller size fraction was
higher than that for larger size fractions, indicat-
ing a higher concentration of the central mode
for the smaller size fraction. It is qualitatively con-
sistent with the present work.
Char fragmentation has been shown previously
to be an important source of the central mode
particles. It is strongly dependent on particle size
[23].Figure 5 indicates that the 1673 K char from
the Coarse coal fraction has a larger average par-
ticle size than that from the Fine coal fraction. Its
lower macroporosity (Fig. 7) may be partially due
to more violent fragmentation of larger coal par-
ticles during pyrolysis [29]. It has been pointed
out that larger char particles fragment more
extensively and a larger number of fine fly ash
would be produced [23]. If assuming that char
fragmentation is the only mechanism for the cen-
tral mode formation, the 1673 K char from the
Coarse coal fraction should fragment more vio-
lently and generate more central mode particles
due to its larger average particle size (Fig. 5).
But it is not the case. As indicated in Fig. 8, the
Fig. 7. Macroporosity of all char samples.
Fig. 8. Central mode concentrations for sized coal
fractions at 1673 K.
D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082 2079
Author's personal copy
Fine coal fraction produced more central mode
particles at the same combustion condition. This
obvious inconsistency is thought to be due to no
consideration of the contribution of the original
fine particles.
There are at least two studies that have
addressed the contribution of original fine parti-
cles to fine fly ash. For example, Sadakata et al.
[26] studied the formation of the submicron fly
ash (SFA) between 0.1 and 1 lm during PCC,
and concluded that approximately half of the
SFA was simply carried over from the submicron
coal fragments, while the remaining SFA was
newly formed through the breakup of larger coal
particles in the devolatilization region. Measure-
ments by Holve [25] showed that the convention-
ally ground power plant coal had a high number
density of very fine particles, whose size distribu-
tion and composition were similar to that of the
fly ash in 0.1–2 lm range. It implied that the
resulting fine fly ash size distribution was derived
directly from the input size distribution. This
again suggested that the original fine particles
were reasonably another important source of fine
fly ash. In the present study, the Fine coal fraction
is abundant in fine particles while the Coarse frac-
tion depleted of fine particles. Therefore, in addi-
tion to char fragmentation, the much larger
number of fine particles in the Fine coal fraction
may also contribute to the central mode. This
can explain why the Fine coal fraction produced
more the central mode particles (Fig. 8).
It is of great practical significance for quantita-
tively determining the contributions of fragmenta-
tion and original fine particles to the central mode
formation. If char fragmentation controls the cen-
tral mode formation, it suggests that major efforts
should be concentrated on combustion processes
and PM control equipments. In contrast, if the
original fine particles are the dominant source as
shown by Holve [25], then pre-combustion
removal of the fine materials might be an efficient
PM emission control choice. However, for the
present work, the degree to which the central
mode formation is induced by fragmentation
and the combustion of the original fine particles
is hard to quantify. This remains a challenge for
future research.
3.2.3. Contribution of excluded minerals
Previous studies of ash formation often focus
on the behavior of included minerals that are
embedded in the carbon matrix. Excluded miner-
als that are outside the carbon matrix behave
quite differently during combustion. They are con-
sidered to contribute to ash particles larger than
1lm either through evolving independently or
fragmentation [33–36]. However, very limited
data are available regarding the role of the
excluded minerals in the central mode formation.
In one study, Kramlich and Newton [14] con-
cluded that the observed central mode around
2lm was not due to excluded minerals, but due
to included minerals that escaped agglomeration.
In contrast, our data in Fig. 3 also show an evi-
dent central mode for the Heavy coal fraction
(>2.0 g/cm
3
) that is comprised primarily of
excluded minerals, similar to that for the Light
coal fraction (<1.4 g/cm
3
) consisting mainly of
included minerals. This indicates that excluded
minerals also contribute to the central mode. In
a later report [27], the data for various cleaned
fractions from an Illinois bituminous coal showed
an appreciable central mode for all coal fractions,
including the heaviest fraction (ash content, 61%)
enriched in excluded minerals. It implied that the
central mode could also be produced from
excluded minerals, consistent with the results
obtained in this study.
The central mode concentrations for the Light
and Heavy coal fractions are compared in Fig. 9.
It can be seen that the central mode concentration
for the Heavy coal fraction is comparable to that
for the Light coal fraction. It should be noted that
this is in terms of the amount of the central mode
particles per gram of coal. But the Heavy fraction
has a much higher ash content of 79.5% (Table 1),
nearly 18 times that of the Light fraction (4.4%,
Table 1). Simple calculation indicates that the cen-
tral mode particles produced from per gram of
excluded minerals is only one twentieth of that
from per gram of included minerals. In this point,
the excluded minerals contribute less to the central
mode than the included minerals. But they may
have a noticeable contribution for utility coals
with a high content of the excluded minerals.
The central mode formation for the Light coal
fraction is reasonably due to particle fragmenta-
tion and combustion of original fine coal particles
as proposed previously. As shown in Figs. 5 and
7, the char prepared from the Light coal fraction
at 1673 K has a larger average particle size than
Fig. 9. Central mode concentrations for density-sepa-
rated coal fractions at 1673 K.
2080 D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082
Author's personal copy
its parent coal sample and a macroporosity of
0.011 cm
3
/g, indicating particle swelling during
pyrolysis. The produced macroporous or even
cenospherical char particles are believed to frag-
ment extensively during combustion and form
the central mode particles. Since the Light coal
fraction was prepared without particle sizing, it
possibly contains a large number of fine coal par-
ticles that also contribute directly to the central
mode. The Heavy coal fraction has been shown
to consist primarily of excluded minerals, and
therefore its average particle size remains almost
unchanged after pyrolysis at 1673 K (Fig. 5).
The very low macroporosity of the pyrolyzed sam-
ple (Fig. 7) is possibly due to fractures and defects
in large particles. The central mode for the Heavy
coal fraction is likely formed through simple car-
ryover of fine minerals and fragmentation of large
particles. The particle size measurement of the
LTA prepared from the Heavy coal fraction
shows that approximately 5% on a volume basis
of the total minerals is less than 6 lm. These par-
ticles may be transformed directly into the central
mode [25,37]. On the other hand, some kinds of
minerals (e.g., pyrite and calcite) can fragment
during combustion due to thermal shock or gas
evolution [33,35,38–40] and likely generate the
central mode particles.
The obtained results also have important prac-
tical implications. If excluded minerals contribute
significantly to the central mode, then pre-com-
bustion separation by coal cleaning techniques
would reduce fine PM emissions greatly, especially
for coals with a significant fraction of excluded
minerals.
4. Conclusions
This work conducted a preliminary study on
the formation mechanisms of the central particle
mode during pulverized coal combustion. Com-
bustion experiments were purposely designed
using sized and density-separated coal fractions
and well-defined conditions. The results showed
that the fly ash particle size distributions had a
general central particle mode at 4lm for all coal
samples. A larger amount of the central mode par-
ticles was produced at the higher combustion tem-
perature due to enhanced char fragmentation. The
higher concentration of the central mode particles
for the small-size coal sample suggested that fine
particles present in the parent coal also contrib-
uted to the formation of the central mode. Exper-
imental results for two density-separated coal
samples indicated that the central mode particles
could also be produced from the excluded miner-
als. But their contribution was less than the
included minerals if in terms of per gram of ash
contained in the coal samples. The quantification
of the contributions of individual formation
mechanisms to the central mode is essential to
the prediction of fine PM emissions and the selec-
tion of appropriate control strategies, which is
deserving of future efforts.
Acknowledgments
This work was supported by the National Key
Basic Research and Development Program of
China (Grant No. 2002CB211602) and the Na-
tional Natural Science Foundation of China
(Grant Nos. 50706013 and 50720145604). The
authors thank Dr. Hongwei Wu at Curtin Univer-
sity, Australia for beneficial discussions and his
help in polishing this manuscript.
References
[1] R.C. Flagan, Proc. Combust. Inst. 17 (1) (1979) 97–
104.
[2] R.E. Desrosiers, J.W. Riehl, G.D. Ulrich, A.S.
Chiu, Proc. Combust. Inst. 17 (1) (1979) 1395–1403.
[3] R.C. Flagan, D.D. Taylor, Proc. Combust. Inst. 18
(1) (1981) 1227–1237.
[4] M. Neville, A.F. Sarofim, Proc. Combust. Inst. 19
(1) (1982) 1441–1449.
[5] S.G. Kang, A.F. Sarofim, J.M. Beer, Proc. Com-
bust. Inst. 24 (1) (1992) 1153–1159.
[6] R.C. Flagan, S.K. Friedlander, in: D.T. Shaw
(Ed.), Recent Developments in Aerosol Science,
Wiley-Interscience, New York, 1978, p. 25.
[7] A.F. Sarofim, J.B. Howard, A.S. Padia, Combust.
Sci. Technol. 16 (3–6) (1977) 187–204.
[8] W.P. Linak, C.A. Miller, W.S. Seames, et al., Proc.
Combust. Inst. 29 (1) (2002) 441–447.
[9] W.P. Linak, C.A. Miller, J.O.L. Wendt, J. Air
Waste Manag. 50 (8) (2000) 1532–1544.
[10] W.S. Seames, A. Fernandez, J.O.L. Wendt, Environ.
Sci. Technol. 36 (12) (2002) 2772–2776.
[11] W.S. Seames, Fuel Process. Technol. 81 (2) (2003)
109–125.
[12] C. Wang, W.S. Seames, M. Gadgil, J. Hrdlicka,
G. Fix, Aerosol Sci. Tech. 41 (12) (2007) 1062–
1075.
[13] S.G. Kang, J.J. Helble, A.F. Sarofim, J.M. Beer,
Proc. Combust. Inst. 22 (1) (1989) 231–238.
[14] J.C. Kramlich, G.H. Newton, Fuel Process. Tech-
nol. 37 (2) (1994) 143–161.
[15] S.G. Kang, A.R. Kerstein, J.J. Helble, A.F. Sar-
ofim, Aerosol Sci. Tech. 13 (4) (1990) 401–412.
[16] D. Yu, M. Xu, H. Yao, et al., Proc. Combust. Inst.
31 (2) (2007) 1921–1928.
[17] D. Yu, M. Xu, H. Yao, X. Liu, K. Zhou, Powder
Technol. 183 (1) (2008) 105–114.
[18] H.L. Clack, Fuel Process. Technol. 87 (11) (2006)
987–996.
[19] W.S. Seames, J.O.L. Wendt, Fuel Process. Technol.
63 (2–3) (2000) 179–196.
[20] W.S. Seames, J.O.L. Wendt, Proc. Combust. Inst.
31 (2) (2007) 2839–2846.
[21] International Commission on Radiological Protec-
tion (ICRP), Human Respiratory Tract Model for
D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082 2081
Author's personal copy
Radiological Protection. ICRP Publication 66,
Annals of the ICRP, 24 (1–3), Pergamon Press,
Oxford, 1994, pp. 1–300.
[22] J.J. Helble, A.F. Sarofim, Combust. Flame 76 (2)
(1989) 183–196.
[23] L.L. Baxter, Combust. Flame 90 (2) (1992) 174–184.
[24] R.D. Smith, J.A. Campbell, K.K. Nielson, Atmos.
Environ. 13 (5) (1979) 607–617.
[25] D.J. Holve, Combust. Sci. Technol. 44 (5–6) (1986)
269–288.
[26] M. Sadakata, M. Mochizuki, T. Sakai, K. Okazaki,
M. Ono, Combust. Flame 74 (1) (1988) 71–80.
[27] J.C. Kramlich, B.C. Chenevert, J. Park, D.A.
Hoffman, E.K. Butcher, Suppression of fine ash
formation in pulverized coal flames, Report No.
DE-FG22-92PC92546, Department of Mechani-
cal Engineering, University of Washington,
1996.
[28] Y. Yu, M. Xu, H. Yao, et al., Proc. Combust. Inst.
31 (2) (2007) 1947–1954.
[29] D. Yu, M. Xu, Y. Yu, X. Liu, Energ. Fuel. 19 (6)
(2005) 2488–2494.
[30] L.J. Wibberley, T.F. Wall, Combust. Sci. Technol.
48 (3–4) (1986) 177–190.
[31] H. Wu, T. Wall, G. Liu, G. Bryant, Energ. Fuel. 13
(6) (1999) 1197–1202.
[32] G. Liu, H. Wu, R.P. Gupta, J.A. Lucas, A.G. Tate,
T.F. Wall, Fuel 79 (6) (2000) 627–633.
[33] L. Yan, R.P. Gupta, T.F. Wall, Fuel 81 (3) (2002)
337–344.
[34] L. Yan, R.P. Gupta, T.F. Wall, Fuel 80 (9) (2001)
1333–1340.
[35] L. Yan, R. Gupta, T. Wall, Energ. Fuel. 15 (2)
(2001) 389–394.
[36] B.J.P. Buhre, J.T. Hinkley, R.P. Gupta, P.F.
Nelson, T.F. Wall, Fuel 85 (2) (2006) 185–193.
[37] S.A. Benson, E.A. Sondreal, J.P. Hurley, Fuel
Process. Technol. 44 (1–3) (1995) 1–12.
[38] S.F. Miller, H.H. Schobert, Energ. Fuel. 7 (4) (1993)
532–541.
[39] S. Srinivasachar, J.J. Helble, A.A. Boni, Prog.
Energ. Combust. 16 (4) (1990) 281–292.
[40] H.M. ten Brink, S. Eenkhoorn, M. Weeda, Fuel
Process. Technol. 47 (3) (1996) 233–243.
2082 D. Yu et al. / Proceedings of the Combustion Institute 32 (2009) 2075–2082
... In recent years, extensive research has focused on determining the effect of different coal mineral forms, their occurrence and thermochemical behavior on particle size distribution (PSD), and composition and morphology of the generated PM. The physicochemical properties of PM 2.5 and PM 10 have been reported, and theories regarding the formation mechanisms have been discussed [4,[8][9][10][11][12][13][14][15][16][17]. It is noted that PM is transformed from minerals that mainly contain Si, Al, Na, K, Ca, Fe, and Mg [4,16]. ...
... Various types of mineral matter present in coals and modes of their occurrence are quite complex, differing in coal ranks and local areas. There is extensive information in the available literature on the transformation of various forms of inorganic species during the combustion of high-grade coal (e.g., sub-bituminous and bituminous coal) and its contribution to PM 10 emissions [12,[14][15][16][23][24][25][26][27][28][29], and there are only few studies on the formation of PM from mineral matter of low-rank coal (i.e., lignite) [25,[29][30][31]. It should be noted that most elements identified in low-rank coal are characterized by varying degrees of organic association, and some major elements (e.g., Ca, Mg, Fe, Al, and Ti) largely occur in nonmineral forms [18]. ...
Article
Full-text available
This article presents the results of a systematic study on the composition and origin of PM1-2 microspheres in high-calcium fly ash. The composition of individual microspheres was studied by scanning electron microscopy and energy-dispersive X-ray spectroscopy. It is shown that the compositions of the analyzed microspheres satisfy the general dependency with a high correlation coefficient: [SiO2 + Al2O3] = 88.80 − 1.02 [CaO + Fe2O3 + MgO], r = −0.97. The formation pathway is parallel to the general trend: anorthite, gehlenite, esseneite, tricalcium aluminate, ferrigehlenite, and brownmillerite. The microspheres were classified into four groups depending on the content of major components: Group 1 (CaO > 40, SiO2 + Al2O3 ≤ 35, Fe2O3 < 23, MgO < 16 wt %); Group 2 (30 < CaO < 40, SiO2 + Al2O3 ≤ 40, Fe2O3 < 27, MgO < 21 wt %); Group 3 (CaO ≤ 30, 40 ≤ SiO2 + Al2O3 ≤ 75, Fe2O3 < 10, MgO < 10 wt %); and Group 4 (14 < CaO < 40, SiO2 + Al2O3 < 14, Fe2O3 > 30, MgO ≤ 14 wt %). A comparative analysis of the relationship between major component concentrations suggests the routes of PM1-2 formation from feldspars and Ca–, Mg–, and Fe–humate complexes during lignite combustion.
... It is worth noting that the identification of the PM modal based on both mass-based and composition-based PSD curves suggested a consistent ultrafine PM (i.e., PM0.1) and further revealed that the PM2.5 was made up of both ultrafine PM and central PM. As pointed out by Yu et al. [39][40][41], the central modal PM was formed via the heterogeneous condensation of the mineral vapor on the fine mineral fragments, which was well consistent with the transition of the content of the volatile mineral matter such as Na and S in the central PM. Compared with the PM2.5 derived from coal SX, there was more Ca in the PM2.5 from coal WCW, which agreed well with the high Ca content in coal WCW (see Table 2) and contributed to the higher PM2.5 yield. ...
Article
Full-text available
Clay minerals composed of Si and Al could help reduce ultrafine particulate matter (PM) formation as an additive during coal combustion while currently unacceptable high adding dosages (normally 3–5 wt.%) are required due to their inadequate capture efficiency. To find additives that could effectively reduce the formation of ultrafine PM, coal combustion with a novel nano SiO2 additive (<100 nm) was performed to evaluate its effects on reducing ultrafine PM. The generated PM10 was sampled to characterize their particle size distribution, mass yield, size-resolved composition and micromorphology. The results showed that adding a small dosage (0.6%) of nano SiO2 reduced the mass yield of ultrafine PM by 30.70%, showing a much higher ultrafine PM capture efficiency than an existing micron-sized natural clay mineral. However, its performance on different coals varied due to disparities in ash content and composition in coal. A composition analysis revealed that the Na content in the ultrafine PM was decreased after adding nano SiO2, indicating that nano SiO2 inhibited the migration of volatile alkali metals such as Na into ultrafine PM because the Na-containing mineral vapor reacted with the nano SiO2 additive particles with a large specific surface area at a high temperature and inhibited their transformation into ultrafine PM via homogenous nucleation. Changes in the element size distributions and micromorphology showed that the majority of the added nano SiO2 particles reacted or coalesced with each other and/or the minerals embedded in coal, finally growing into a larger PM.
... μm in size 12−14 are formed via the vaporization−condensation mechanism, 10,21 while coarse-mode particles with the size >5.0 μm 12,13 result from char fragmentation and mineral coalescence. 7,20 The formation of central-mode particles (0.3− 5.0 μm) 14,18 depends on a combination of different mechanisms, including the char and mineral fragmentation, 18,19 direct transformation, and coalescence of fine included minerals. 19,20 The mechanisms of PM formation significantly depend on the structure of char particles, which are formed as a result of the coal type-dependent conversions during devolatilization. ...
Article
Scanning electron microscopy and energy-dispersive X-ray spectroscopy were used to analyze individual microspheres, 1–2 μm in size, located in coal char particles of Inertoid and Fusinoid/Solid morphological types. It was shown that PM1–2 (where PM = inorganic particulate matter) is formed in the porous structure of the carbon matrix, which controls the microsphere size, from authigenic minerals that determine their composition. Depending on the contents of SiO2, Al2O3, and FeO, the studied microspheres fall into various groups differing in mineral precursors. The precursor of the Group 1 microspheres with the specific composition of SiO2 + Al2O3 > 95 wt % and FeO ≤ 1.5 wt % is NH4–illite. Microspheres containing SiO2 + Al2O3 < 95 wt % and FeO in increasing content amounts up to 4, 6, and 10 wt %, included in Group 2, Group 3, and Group 4, respectively, are formed from mixed-layer K–illite–montmorillonite minerals subjected to cationic substitution with iron, followed by the entry of Fe3+ in interlayer sites. Calcite, dolomite, gypsum, magnesite, rutile, and siderite are involved in the formation of Group 5 microspheres with a high content of Ca, Mg, Ti, or Fe. The significant part of PM1–2 is represented by microspheres of Groups 2, 3, and 4 regardless of the type of coal char particles (62% for Inertoid ones and 75% for Fusinoid/Solid ones). About one-third of the microspheres for both char morphotypes refer to Group 5. Microspheres of Group 1 (8%) are located only in the Inertoid char particles, which results from the characteristic effect of the maceral–mineral composition of original coal. It has been suggested that Inertoid and Fusinoid/Solid char particles are formed from various macerals, semifusinite and fusinite, respectively. Due to the closed-cell structure, semifusinite contains noncation-exchanged NH4–illite, the mineral precursor of microspheres with low contents of Fe, K, Na, and Mg. The fusinite structure allows cationic substitution in NH4–illite with the formation of mixed-layer K–illite and montmorillonite, the mineral precursors of a significant part of PM1–2.
Article
At present, the output of municipal sludge in China is increasing year by year, and co-combustion of sludge and pulverized coal for power generation is a means to treat sludge in large quantities. However, the co-combustion of sludge and pulverized coal will generate a lot of particulate matter (PM). In this paper, the formation characteristics of particulate matter (PM10) produced by the co-combustion of sludge and pulverized coal were studied. The effect of the furnace temperature and the sludge mixing ratio on the mass-based particle size distribution and elemental composition of PM10 in sludge, coal combustion and co-combustion was discussed. The experiment results show that the mass-based particle size distribution of PM10 produced by pulverized coal and sludge combustion are both bimodal, while the elemental composition and particle size distribution of PM10 are different. Whether burning alone or co-firing, increasing the furnace temperature will increase the mass yield of PM10. Increasing the temperature promotes particle fragmentation more than the melting aggregation of particulate minerals, increasing PM0.4–10 yield. The mass yield of PM10 produced by coal combustion is higher than that of sludge. In the sludge addition ratio range ≤ 30%, increasing the sludge addition ratio will reduce the mass yield of PM10. The synergistic effect exists between sludge and pulverized coal co-combustion. Compared to the weighted calculation results, the content of main elements Si, Ca, Fe and S in PM0.4 decreases, while in PM0.4–10 the content of Si decreases and the content of Ca, Fe, Al, and P increases. At a low sludge addition ratio of 10%, the elemental composition of PM10 with the particle size distribution trend is consistent with that of pulverized coal, and sludge co-firing takes negligible effect on PM10 formation.
Article
With the increasing production of municipal sewage sludge (MSS) in China every year, the co-firing of MSS and pulverized coal is getting more and more widely applied in large coal-fired power plants. The co-firing of MSS and pulverized coal will produce a large amount of particulate matter (PM) emissions, especially submicron particles. In this paper, the formation characteristics of submicron particles in the co-firing process of coal and MSS were studied in a drop tube furnace. The influence of the furnace temperature and the addition ratio of sludge on the particle size distribution and element composition of submicron particles in MSS, pulverized coal combustion and co-firing was mainly studied. The experimental results show that the furnace temperature has an influence on the formation of PM0.4. For sludge combustion, increasing the furnace temperature will promote the formation of PM0.4. The main reason is that increasing the furnace temperature promotes the gasification of Si, S, Fe, and P to form the precursor of PM0.4 or PM0.4. At same furnace temperature, the volume concentration and mass concentration of PM0.4 produced from pulverized coal combustion are less than that of sludge. Different from sludge combustion, co-firing of pulverized coal and sludge has a synergistic effect on eliminating PM0.4 formation. Increasing the addition ratio of sludge can decrease the volume concentration and mass concentration of PM0.4. This is because that aluminosilicates formed during co-firing promotes the scavenge Si, Ca, Fe, thereby reducing the precursors of PM0.4 and the mass yield of PM0.4. Increasing the furnace temperature in co-firing can inhibit the formation of PM0.4. When the furnace temperature is between 1100 °C and 1300 °C, increasing the furnace temperature will reduce the Fe content and increase the content of Si, Ca, Na, K, and P in PM0.4. However, the reduction of Fe and the increase of Si, Ca, Na, K, and P in PM0.4 offset each other, resulting in an insensitive relationship between the mass yield of PM0.4 and the furnace temperature.
Article
Current additives generally have lower efficiency on reducing PM2.5 than the ultrafine PM as they could not effectively reduce the central PM. To find additives that could effectively reduce the formation of central PM, two kinds of nano particles (namely Al2O3 and TiO2) were screened out, and coal combustion with the nano additives were performed. The generated PM10 were sampled to characterize their particle size distribution, mass yield, size-resolved composition and micro morphology with element map. The results showed that adding a small dose (0.6%) of nano Al2O3 and TiO2 reduced the mass yields of PM0.1-2.5 (i.e., central PM) and PM2.5 by 24.38%-26.06% and 12.85%-19.59% respectively, indicating quite different PM reduction characteristics than the conventional clay-mineral-derived additives. The nano additives on the fuel surface promoted the melting and coalescence of small ash particles, promoting their migration into coarse PM. Moreover, the added nano additives covering the coal particle scavenged the mineral vapors releasing from the coal particle during its pyrolysis/combustion, and thereby reduced their migration into central PM via the following heterogenous condensation. Finally, the added nano additives mainly migrated into the large ash particles and enhanced heterogeneous condensation/reaction of the released mineral vapor via providing larger specific surface area.
Article
Biomass can be torrefied to improve the fuel quality prior to its utilization. In addition, a large number of soluble alkali metals can be removed by water washing, to solve slagging and other ash-related problems. In this study, computer-controlled scanning electron microscopy (CCSEM) was used to analyze the mineral composition and elemental distribution of straw, torrefied straw char, and torrefied-water washed straw char. The mineral transformation behavior and the distribution of main elements, when these biomass species were individually co-combusted with coal, were also studied in detail. The comparison of the mineral composition of bulk ash obtained from torrefied straw combustion and straw combustion revealed that the quartz phase decreased and the aluminosilicates of K-, Na-, and Ca–Fe–K increased in the former compared to that in the latter. After the torrefied straw was washed with water, the quartz and Si-Rich phases increased. K aluminosilicate decreased and the quartz phase increased when the torrefied sample was co-combusted with coal at both mixing ratios of 1:1 and 1:4 compared to straw/coal co-combustion. When the torrefied-water washing sample and coal were co-combusted at a 1:1 ratio, the quartz phase decreased, and the Si-Rich and mullite phases increased. When the straw/torrefied char/torrefied-washed char was co-fired with coal at the 1:4 ratio, the amount of K aluminosilicate was less compared to that observed at the 1:1 ratio. It can be concluded that both torrefaction and torrefaction-washing can reduce the content of low-melting-point aluminosilicates in the mineral composition of the studied biomass after co-combustion, which can be beneficial to alleviate the ash slagging problem.
Article
An important environmental issue is the emission of semi-volatile toxic metals, such as arsenic, selenium, and cadmium, from the combustion of coal. These materials may vaporize in the hot portions of the combustor then return to the solid phase in cooler zones of the process downstream. Understanding the mechanisms by which toxic metals partition between the vapor and solid phases is an important step for predicting and mitigating the effect of these metals upon the environment. Particulate ash samples were withdrawn from a 17-kW pilot-scale downflow combustor in which pulverized coal was burned under self-sustaining conditions. The samples were size segregated in a Berner low pressure impactor, and then analyzed using neutron activation. This research approach has suggested mechanisms, which govern the partitioning of arsenic, selenium, and cadmium in practical pulverized coal combustion processes. The results suggest that volatilization and subsequent transfer of selenium to submicron particle surfaces appears to be an important post-combustion phase mechanism for Illinois #6 coal but not for Pittsburgh seam coal. Most of the selenium in the Pittsburgh submicron fly ash, cadmium in Illinois #6 submicron fly ash, and arsenic in both Pittsburgh and Illinois #6 submicron fly ash enters the post-combustion zone in the solid phase. The dominant heterogeneous partitioning mechanism for transformation to large, supermicron particles is the reaction of metal vapor on the surface or within the pores of an ash particle. The results also suggest that the rate of transformation is dominated either by an exterior surface reaction-controlled regime or a pore diffusion-controlled regime. A relationship between the concentration of solid phase arsenic, selenium, and cadmium to calcium in supermicron particles was also observed, suggesting the formation of As–Ca, Se–Ca, and Cd–Ca reaction products.
Article
Submicron particles resulting from pulverized coal combustion in coal-fired steam plants have been chemically characterized and the results interpreted in terms of various mechanisms for their formation. Elemental analyses using XRF, AA, and INAA techniques are reported for approximately 60 elements for flyash from two large coal-fired plants. The concentrations of elements volatilized during combustion are independent of particle size in the submicron size range, in contrast to the larger particle sizes where an inverse relationship with particle size applies. Scanning electron micrographs show the submicron particles are also much more homogeneous than larger flyash particles, while still showing evidence of some surface enrichment in ESCA studies. Various mechanisms for formation of submicron particles have been considered. The present results are most consistent with formation of submicron particles by the bursting of larger particles due to gas release during rapid heating, followed by coagulation and condensation of volatilized elements to form particles in the 0.1–1.0 μm size range. The importance of this mechanism in other coal-fired plants is probably dependent on both temperature and ash composition.
Article
The submicron fraction of coal fly-ash, though comprising a minor fraction of the total mass, has become increasingly recognized as a health hazard and a possible link in irreversible boiler fouling. An outgrowth of continuing research on submicron particle behavior, this document describes efforts to predict the composition and morphology of condensation ash. A model based on a process of metal oxide vaporization from a burning char particle, diffusion and recondensation has been constructed and employed. A submicron particle concentration equaling five percent by weight of the fly-ash from a cyclone-fired boiler is predicted. Because of lower combustion temperatures in pulverized-fuel-fired boilers, less than one tenth of one percent of the total fly-ash is attributed to vapor transport. Enrichment of silicon and iron is predicted in the submicron fraction. This is verified experimentally for silicon but not for iron. Experimental examination reveals aggregates containing primary particles, of about 60 nm in diameter. Particles of this size are evidence that the submicron ash precipitates in the combustion zone of the boiler to grow by coagulation for several seconds before quenching.
Article
This article presents the differential mass size distributions of coal combustion particulate matter (PM) determined with the Berner low-pressure impactor (BLPI, Hauke Model 25-4/0.015) and a newer generation of low pressure impactor, the Dekati low-pressure impactor (DLPI, Dekati Ltd Model 6281). The collection characteristics of the BLPI and DLPI are compared and cutoff diameters are calculated. Samples were collected in the post-combustion zone of a 19 kW vertical downflow combustor from two coal types. Both BLPI and DLPI represent a tri-modal distribution and give statistically similar characterizations of the coal ash particle size distribution. Distributions generated from DLPI data have higher fractions of submicron particles compared to those generated from BLPI data. The DLPI's two additional stages may provide greater resolution in the submicron region than the BLPI.
Article
Ash formation during pulverized coal combustion has been simulated using a Monte Carlo model to follow the ash coalescence and char fragmentation processes that govern the final particle size distribution. As a consequence of the competition between the particle breakup by fragmentation and the ash coalescence enhanced by the shrinking peripheral area with the progress of char reaction, a bimodal size distribution is obtained. A simplified two-column model is also presented in order to explain the influence of the ash redistribution scheme used in the model on the evolution of the bimodality.
Article
An Australian bituminous coal was burnt in a laboratory drop-tube combustor at 1400, 1500 and 1600°C to determine the effect of p.f. properties and combustion temperature on the character and particle size of the flyash. The experiments showed that the mass mean particle size of the flyash was approximately proportional to that of the p.f., and almost independent of combustion temperature. In contrast, the proportion of fine ash (< 10 μ) was independent of p.f. size but increased markedly with increased combustion temperature. The upper size of the ash was determined by unburnt char for the 175 μm and 240 μm p.f. and by the mineral matter for the 29 μm p.f. Cenosphere formation increased with combustion temperature and dominated the ash formed at 1600°C. The work emphasises the need for more detailed p.f. analysis during laboratory and pilot-scale combustion studies into flyash formation and related phenomena such a's fouling, filtration and precipitability. Additional research is required to quantify cenosphere formation and its influence on the formation of fine flyash.
Article
Optical methods for particle size distribution measurements in practical high temperature environments have achieved feasibility and offer significant advantages over conventional sampling methods. The present paper describes a mobile electro-optical system which has been designed for general use in a wide range of research and industrial environments. The instrument has demonstrated capability for measuring individual particles in the size range 0.25-100 microns at number densities up to 10/m.- The system incorporates an in situ alignment procedure that is easy to set up and use in a wide variety of research and industrial environments. In addition to demonstration of the system's wide dynamic range, we show the utility of the in situ alignment method in hot (1100 K) turbulent flows where beam steering can be a problem. A major focus of the work described here is the application of the particle sizing technique to measurements of flyash formation in laboratory reactors. Number and mass frequency distribution measurements of flyash and pulverized coal obtained in a turbulent and a laminar flow reactor show that the raw pulverized coal contains large numbers of submicron particles similar to the flyash formed after combustion. These results suggest another source for much of the submicron flyash formed in pulverized combustion, with significant implications for control of fine particle emissions.
Article
A suite of six coals, of widely differing As, Se, Ca, Fe, and sulfur contents, was burned under self-sustaining conditions in a 17kW downflow laboratory combustor. Size segregated ash-laden aerosol samples were isokinetically withdrawn and collected on a Berner low pressure impactor. Correlations between trace element concentration (As or Se) and that of major elements (as functions of particle size) were then used to infer chemical associations between trace metals and Ca and/or Fe, and how these depended on sulfur. These baseline data led to formation of the following hypotheses, namely:(1)dominant As and Se partitioning mechanisms depend on the availability of Ca and/or Fe active sites for surface reaction;(2)increasing combustion temperature increases the availability of active cation sites, and increases partitioning of As and Se to fly ash by surface reaction;(3)sulfur competes with these surface reactions, decreasing As and Se partitioning to fly ash surfaces.These hypotheses were tested by manipulating the As, Se, Ca, Fe, and S contents for various coals by doping. Temperature was adjusted in order to achieve comparisons of different coals and different coal constituents at similar thermal conditions, through O2 and CO2 addition, as required. These results confirmed the hypotheses above, and allowed an association regime map to be constructed. This map shows that both As and Se associate with Fe and Ca, provided active sites are available. Se reacts preferentially with Fe over Ca when both are available while As reactions with both Fe and Ca are comparable. Sulfur can prevent association of both As and Se, by preferentially reacting with active sites, especially those on Fe. When sufficient sites are not available, the release of vapor-phase As and Se species is promoted.
Article
The transformations and deposition characteristics of pyrite (FeS2) under pulverized coal combustion conditions were examined in an entrained flow reactor. Mössbauer spectroscopy, scanning electron microscopy, and energy dispersive X-ray analysis, were used to monitor the evolution of particle composition and morphology. Pyrite initially decomposed to form pyrrhotite (Fe0.877S). Fissures resulting within the particle during this step led to limited bulk fragmentation. Subsequently, the pyrrhotite particles melted to form an iron oxysulfide droplet. Magnetite (Fe3O4) crystallized out of the melt once the melt oxide content exceeded 85%. Hematite (Fe2O3) was formed at longer residence times. A comparison of kinetic model calculations with experimental data revealed the controlling resistances for each of the stages during the transformation process. Deposition experiments established the iron oxysulfide phase to be responsible for particle adhesion, and the duration of this melt phase was determined to be a function of pyrite particle size, gas temperature, oxygen concentration, and the extent of fragmentation. Combustion experiments were also conducted with two coals, one with predominantly included pyrite grains and the other with a significant fraction of excluded pyrite, to discern the limiting behavior of pyrite in forming the final ash.