Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Energy

Energy Procedia 4 (2011) 892–899 Procedia


www.elsevier.com/locate/procedia
Energy Procedia 00 (2010) 000–000
www.elsevier.com/locate/XXX

GHGT-10

Lean Flammability Limit for Oxy-Fuel Fired Pulverized Coal


Combustion Systems.
Masayuki Taniguchia*, Tsuyoshi Shibataa, Kenji Yamamotoa, Christian Kuhrb,
Osamu Itoa 1*
a Energy and Environmental Systems Laboratory, Power Systems Company, Hitachi, Ltd.,
7-2-1 Omika-cho, Hitach, Ibaraki, 319-1292, Japan
b Hitachi Power Europe GmbH, Schifferstraβe 80, 47059 Duisburg, Germany
Elsevier use only: Received date here; revised date here; accepted date here

Abstract

We developed a model based on fundamental experimental data to predict lean flammability limit, L, and flame propagation
velocity, Sb, for oxy-fuel combustion conditions. The basic model system consisted of two particles. One side of the two
particles burns first, then, the other particle is ignited by the heat of combustion of the one burning particle. This phenomenon
was defined as flame propagation. We analyzed at what distance the first burning particle could ignite the next particle (flame
propagation distance d, and related to L), and how fast the first burning particle could ignite the next particle (flame propagation
time s, and related to Sb) under various conditions. The proposed model was verified with data of both fundamental and pilot-
and actual-scale experiments. We also applied the model to develop burner systems for lignite-fired oxy- fuel combustion. Local
Sb and L near the ignition points of the burner systems could be analyzed from the concentration and temperature profiles of the
general CFD results (k-ε method). Flame stability was judged by the calculated Sb and L profiles, and past results of blow-off
limits obtained with actual- and pilot-scale experiments. We call this proposed technique flammability analysis. By using
combination with the technique and Large Eddy Simulation, we could quickly clarify points for improvement of the burner
systems. The calculated results were applied to a DS ®T-burner designed by Hitachi Power Europe, installed at Schwarze Pumpe
pilot plant.
©
⃝c 2010 Elsevier Ltd.
2011 Published All rights
by Elsevier reserved
Ltd.

Keywords; pulverized coal; oxy -fuel combustion ; large eddy simulation; flame propagation velocity ; lean flammability limit

1. Main text

1. Introduction
Recently, development of oxy-fuel combustion technology has been particularly active using pilot-scale plants
[1,2]. Fuel ignition properties are fundamental combustion performance parameters for engineering design of
combustion systems. Improving ignition performance leads to improved combustion performances including such
items as, expanding the turn-down ratio, expanding the application fu el properties, maintaining safety of the fu el

* Corresponding author. Tel.: 81-29-276-5889; fax: 81-29-276-5622.


E-mail address: masayuki.taniguchi.xc@hitachi.com.

doi:10.1016/j.egypro.2011.01.134
M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899 893
2 Author name / Energy Procedia 00 (2010) 000–000

supplying system, improving the combustion efficiency, and reducing environmental pollutants such as NOx.
Usually, for oxy-fu el combustion systems, a mixture of exhaust flue gas and oxygen is used as the combustion
supporting gas [1]. Oxygen concentrations are usually variable for oxy-fuel combustion systems; therefore, ignition
performances can be varied significantly.
Flame propagation velocity is one of the most important ignition performance parameters. Suda et al. [3] have
studied flame propagation velocities under air and oxy-fu el combustion conditions. Lean flammability limit is also
a very important ignition performance. However, the lean fl ammability limits of pulverized coals for oxy -fu el
combustion have not been reported.
We developed a laser ignition experiment technique [4] to study both fl ame propagation velocity and lean
fl ammability limit in N2 /O2 atmospheres [4, 5]. In the present investigation, we extended the model to oxy-fu el
combustion. We analyzed the experimental results of Suda et al. [3] and our own results, and developed a model for
estimating lean flammability limit and flame propagation velocity for oxy -fuel combustion. We also applied the
model to develop an engineering design for actual and/or pilot-scale burner systems for oxy-fuel combustion.

2. Experimental equipment and the basic model


2.1 Ignition and flame propagation phenomena for pulverized coal combustion
Fig.1 shows a model for flame stabilization of pulverized coal burners [6]. Generally, pulverized coal and carrier
gas (primary air) are supplied from the center of the burner. The secondary air is supplied around the primary air. A
recirculation region of the high temperature burning gas is fo rmed between the flows of the primary air and
secondary air by the effect of the fl ame stabilizer. The burning gas of the recirculation region heats the pulverized
coal particles in the primary air and these particles ignite and burn. When the coal concentration is high, the burning
coal particles heat and ignite other nearby coal particles. The flame moves from the coal particles directly heated by
the gas of the recirculation region to the surrounding coal particles. This phenomenon is considered flame
propagation. The coal/oxidizer (such as air) mixture should be under flammable condition to stabilize the flame.

Furnace wall
Fundamental experimental data
Heat loss Heat loss
Ignition Flame propagation
Secondary air H igh tem perature tim e
recirculation region
Flam e
stabilizer 2
ignition Flam e
Prim ary air
+ Coal Burning particle
Unignited particle

Fig.1 Ignition and flame propagation phenomena of pulverized coal combustion.

2.2 Laser ignition experiments


Fig.2(a) shows a schematic of the experimental equipment [4, 5]. Uniformly sized pulverized coal particles were
suspended in a laminar upward flow and rapidly heated by a single-pulsed YAG laser. Velocity of the upward flow
was controlled according to the particle diameter. The heated pulverized coal particles were burned in the quartz
test section (50mm square cross section). The particle concentration was measured from the intensity of particle
scattering by a He-Ne sheet laser.
Another continuous laser was used to determine the effect of radiant heat loss on the ignition characteristics. Its
beam diameter was usually 15mm around the ignition point.
Photos of the burning particles are shown in Fig.2(b). Two kinds of phenomena were observed. The mixture of
Fig.2(b-1) was under the non-flammable condition. Immediately after the pulse laser shot, light emission was
observed from a small number of particles. These were burning particles that were directly ignited by the laser shot.
The mixture of Fig.2(b-2) was under the fl ammable condition. A small number of particles ignited at fi rst in the
same way as in Fig.2(b-1). However, the number of burning particles then increased, as revealed by the laser area of
light emission. That is, the flame moved from the particles directly heated by the laser to the surrounding particles.
This phenomenon was considered fl ame propagation. Flame propagation velocity was defined as the growing rate
of the flame radius [4].
894 M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899
Author name / Energy Procedia 00 (2010) 000–000 3

(a) Laser ignition equipm ent (b) Flam e pictures (c) B asic Phenom ena
Pulsed particle feeder (b-1) non-flam m a b le condition
3m s after
ignition
Suspended Tim e = 0
laser shot 5ms
Coal P articles 11ms
Single pulsed
Y A G laser
(for ignition) V o latile flam e
L a m inar Flow ignition
d
Tim e = s
H e-Ne sheet laser (b-2 ) flam m a b le condition Flame
(coal concentration 3 m s propagation
measurem ent ) 10m s
30ms
Flow conditioning Distance betw een particles; d
Continuous
section Tim e of flam e propagation; s
laser
(adjustm ent of Combustion Model; relation betw een d and s
the heat loss) supporting gas
(oxidizer) 10mm Obtained from experim ents

Fig.2 Laser ignition experiments.

Fig.2(c) summarizes the basic phenomena of flame propagation. One side of two particles burns first, then, the
other particle is ignited by the heat of combustion of the one burning particle. When the first particle ignites, volatile
matter is pyrolized. A volatile matter flame is formed around the first particle. The flame grows due to
volatilization, and the flame heats the next particle which has not ignited yet. Flame propagation is observed if the
fi rst burning particle can transfer the fl ame to the next particle before the volatile matter combustion of the first
particle has finished. We defined the distance between particles as d and the time of flame propagation as s. Flame
propagation velocity, Sb, was defined as the value of d divided by s. In this study, we analyzed experimental data to
obtain the relationships between d and s under various experimental conditions.

3. Results and Discussion


3.1 Lean flammability limit and flame propagation velocity
Coal;bitum inous, anthracite, petroleum coke
=(
d/s) 3
Diam eter: 22μm ,O xygen: 21-100% (N 2 /O 2 )
Heat flux from w all: room tem p. - 3x104 W / m 2
(c) hv bitum inous
propagation flam e propagation

(O 2 ;1 00%)
probability (-) velocity: Sb (-)

1/Lean flam m ability lim it:1/L

(a)
1.0
Norm alized

hv bitum inous
2 (O2; 61%, N2; 39%)
0.5 hv bitum inous
(arb. unit)

(O 2 ; 21%, N2; 79%)


lv bitum inous (O2; 100%,),
0 flam e w as heated by the
100 1 continuous laser
(b)
Flam e

lv bitum inous (O2; 100%)


Lean flam m ability lim it
50 anthracite (O2; 100%)
∝ (
1/d)3 lv bitum inous (O2; 100%,), large particle
0
0
0 2 4 0 1 2 3
Norm alized particle concentration (-) Maxim um flam e propagation velocity: Sb-m a x (arb. unit)

Fig.3 Correlation between lean flammability limit and flame propagation velocity
and experimental analyses based on that.

Examples o f relationships between coal concentration and fl ame propagation velocity and between coal
concentration and fl ame propagation probability are shown in Fig.3. Coal concentrations and fl ame propagation
velocities are shown as normalized value. The coal concentration is in inverse proportion to 3 power of the distance:
M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899 895
4 Author name / Energy Procedia 00 (2010) 000–000

d. Fig.3(a) shows normalized flame propagation velocity. When coal concentration increased, flame propagation
velocity increased. But there was an upper limit value to the fl ame propagation velocity. The flame propagation
probability is shown in Fig.3(b). The flame propagation probability was calculated as the ratio of the number of
experiments in which flame propagation was observed and the number of experiments in which ignition was
observed. [4]. We defined the lean flammability limit as the coal concentration when the flame propagation
probability was zero.
The results of Figs.3(a) and (b) meant that lean fl ammability limit correlated with flame propagation velocity.
Flame propagation velocities were almost zero at the lean flammability limit concentration. Fig.3(c) shows analyses
of experimental results obtained by the relationships shown in Figs.3(a) and (b). Lean fl ammability limit L was
inversely proportion to the maximum flame propagation velocity, Sb-max, obtained by varying coal concentration.

3.2 Effects of experimental conditions on flame propagation performances


Fig.4(a) shows the difference of the flame propagation phenomena with coal with different volatile content. For
high volatile content coal, the growth rate of the volatile fl ame was large, because pyrolysis rate was large. The
fl ame could be transmitted in a short time from one burning particles to another one, even though the distance
between the particles was large. Flame propagation velocities became large fo r high volatile content coals. The
time of flame propagation s became long. Fig.4(b) shows the relationship between pyrolysis rates o f coals and
fl ame propagation velocities. The pyrolysis rates were calculated results for the same heating rate condition as that
o f actual boilers (20000K/s). Each pyrolysis rate constant was obtained by laborato ry-scale experiments [5]. Flame
propagation velocities strongly depended on pyrolysis rates. Fig.4(c) relates coal properties and pyrolysis rates.
The influence of the coal properties on flame propagation velocity and lean flammability limit could be analyzed by
getting the database which expressed relationships between coal properties, pyrolysis rates and fl ame propagation
velocities for various kinds of coal and other solid fuels.
3 3
(a) M odel (b) (c)
P yro ysis ra te (arb. unit)

high volatile coal Peat


S b-m ax (arb. unit)

Volatile fla m e
2 2 B itum inous
W ood
L ignite
1 1
ignition
low volatile coal Anthracite
0 0
0 1 2 0 50 100
Pyrolysis rate (arb. un it) Volatile m atter (wt%, dry ash-free)
ignition
Fig.4 Effects of coal properties.

The flame propagation velocities in the N2 /O2 and CO2 /O2 atmospheres were experimentally obtained by Suda et
al [3]. Their results are shown in Fig.5(a). The flame propagation velocities strongly depended on composition of
combustion supporting gas (oxidizer). The flame propagation velocity increased significantly with O2 concentration.
The flame propagation velocities in the CO2 /O2 atmosphere were smaller than those in the N2 /O2 atmosphere when
oxygen concentrations were the same. Fig.5(b) shows temperatures of flames form ed around the particles; T vf
shows the flame temperature, and T vf* is the standard value. The experimental results [4] were measured with two
color pyrometers. The calculated fl ame temperature was defined as adiabatic flame temperature when volatile
matter was burnt under the stoichiometric condition. The flame temperature also strongly depended on composition
of combustion supporting gas. The difference of this fl ame temperature produced the difference of the flame
propagation velocity. Fig.5(c) shows the relationship between the fl ame temperatures and the fl ame propagation
velocities. Both sets of results in the N2 /O2 and CO2 /O2 atmospheres are shown. The flame propagation velocity
was strongly influenced by the fl ame temperature. When the flame temperature was fixed, the flame propagation
velocity had hardly any influence from the gas composition.
We developed a model to predict both fl ame propagation velocity and lean fl ammability limit, based on these
results. The model can analyze effects of coal properties, coal particle diameter, coal concentration, oxygen
896 M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899
Author name / Energy Procedia 00 (2010) 000–000 5

concentration, radiant heat loss from flame to surroundings, and composition o f combustion supporting gas.
Relationships between coal concentrations and flame propagation velocities were cal culated under various
compositions of combustion supporting gas. The results are compared in Figs.5(d) and (e) with the experiment
results [3,4]. Fig.5(d) shows the effects of oxygen concentration in the N2 /O2 atmosphere . Fig.5(e) compares N2 /O2
and CO2 /O2 combustion when the oxygen concentration was 40%. The calculated results agreed with the
experimental ones.
Coal: hv bitum inous, Diam eter: 58μm for (a), (c) and (e), 2 2μm for (b) and (d)
Symbols; exp. Lines; calc. Experim e n tal data of (a),(c) and (e): Suda et al [3], (b) and (d): Taniguchi et al [4]
1.2 1.2 (d)

Sb (arb. Unit)
1.2
(a) (c) O 2 100%
S b (m /s)

N 2/O 2 0.8
0.8 1
Oxy-fuel 0.4
0.4 0.8 N 2/O 2
N 2 /O 2 : O 2 61%
Sb (m /s)

0 0
0 0.5 1
0.6 Volatile flam e, Coal concentration (arb. unit)
Tvf - T vf* ( (K )

1000 Temp: Tvf 1.5

Sb (arb. U nit)
(b) N /O (e) O 2 40l%
2 2 0.4 N 2 /O 2
500 1
0.2
0 0.5
Oxy-fuel Oxy-fuel Oxy-fuel
0 0
-500 -500 0 500 1000 0 0.5 1.0
0 50 100
O (vol%) Tvf - Tvf* ( (K ) Coal concentration (arb. unit)
2

Fig.5 Effects of composition of combustion supporting gas.


(a) Laser ignition (b) P ilot scale (c) Pilot → A ctual scale
Experim ents: K iga et al [8]
coal feed line W ater w all Caster w all Experim ents : K iyam a et al [7]
Pilot scale A ctual
(w all tem p . (w all tem p . (w all tem p . Pilot scale
600-900K) 1000-1300K) Caster (coal feed rate ≒ 3t/h) boiler
300-400K) W ater w all
Operating W ater
3 Oxy
Oxy -fuel Caster burner wall
Lean flam m ability lim it (arb. unit)

fuel (O 2 21%) burner


hv
hv -bitum iinous
nous 3
B low -off lim it (arb. unit)

60°
a m mabl
Flam m able
Oxy-
Oxy -fuel
fuel Fine coal
B low -off Lim it (arb. unit)

(O 2 20%)
(O 5
2 SStabl
tablee
Lignite 2 4
A ir, hv
ir, hv - Stable
coarse 3
bitum iinous
nous
1 2
1
Pure oxygen, 1 Unstable
lv-
v -bitum
bitum inous
inous x3 Unstable
Interm ediate
0
non-
non -Fla m mmaabl
b le 0
0 0 10 20 30 40 0 0.2 0.4 0.6 0.8 1
10 2 10 4 10 6 Norm alized heat loss rate
Volatile m atter content
2
Heat flux from w all (W /m ) (w t%, as fired) - Standard value (-)

Fig.6 Effects of heat loss rate from the flame to the furnace wall. Symbols, experimental; lines, calculated.

For engineering design of actual boilers, it is important to consider the effect of heat loss from the flame to the
furnace wall. Lean fl ammability limit and flame propagation velocity were varied with the difference of the heat
loss rate by the furnace conditions, such as whether the furnace was covered with a water wall or caster. The basic
expression of the radiant heat loss was obtained from the laser ignition experiments [6]:

1/L =a Qwall (1/L0 ) (1)


M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899 897
6 Author name / Energy Procedia 00 (2010) 000–000

where, L is lean flammability limit, a is a constant, Qwall is radiant heat flux from a wall or from the continuous laser,
and L 0 is standard lean fl ammability limit for standard heat flux condition (usually, wall temperature = room
temperature). Example effects of heat flux from a wall on lean flammability limit are shown in Fig.6(a).
Experimental data of Fig.6(a) were obtained by the laser ignition experiment. When the wall temperature was low,
the lean fl ammability limit concentration was high. As wall temperature became higher, lean fl ammability limit
decreased, and ignition became easy. The difference between the air and oxy -fuel combustion is also shown in the
fi gure. Lean flammability limit concentration became higher fo r oxy-fuel combustion, when oxygen concentration
was the same.
The model was verified by experiments in pilot- and actual-scale furnaces (Figs.6(b) and (c)). The effects of coal
properties on lean blow-off limit concentration are shown in Fig.6(b). The lean blow-off limits were obtained
experimentally fo r six kinds of coals by using the same burner and furnace conditions [7]. The effects of heat loss
are shown in Fig.6.(c) The lean blow-off limit was obtained experimentally by varying the area ratio of the caster
and water wall that covered the fu rnace wall [8]. We assumed that the flame became unstable when Sb was lower
than the constant value that was decided for each burner structure. We defined the constant value as minimum-Sb to
form a stable fl ame by the burners. The minimum-Sb value was around 0.05m/s for the present experimental
conditions. The lines in Figs.6(b) and (c) mean the coal concentrations when the calculated fl ame propagation
velocities were equal to the minimum-Sb. Calculated results of Figs.6(b) and (c) agreed with the experimental ones.
A more detailed explanation is offered in reference [6].

3.3. Application of the model: proposal of flammability analysis


We applied the model to develop the engineering design of actual and/or pilot -scale burner systems. We call our
proposed technique fl ammability analysis. This technique is a post-processing analysis of CFD calculations. A
general CFD method, such as the k-ε method can be used for this analysis. Calculated results of coal concentration,
gas composition and temperature profiles were read at first. Next, experimental conditions, such as coal properties,
particle diameter distribution and fu rnace wall temperature, were read. Flame propagation velocity profiles were
calculated by using the information. Flame stability was judged by the calculated flame propagation velocities and
past results of actual- and pilot-scale experimental results. If the flame propagation velocity Sb was larger than the
minimum-S b, the condition of the mixture of coal and oxidizer was judged as fl ammable. Flame propagation
velocity near the recirculation region was very important for considering the flame stability.
Flam e picture Detailed C FD com bustion GeneralC FD com bustion Flam m ability analysis
calculation (LE S) calculation (k-ε)
CASE I few w eeks few days few hours
A ir com bustion
Flam mability; Good
Flam e propagation velocity

C A S E II Difficult
O x y-fuel
com bustion
(initial)

C A S E III
Oxy-fuel com bustion Good
(revised)
R Tem perature
300 1700K
position
C A S E III;a DSR T-burner designed by Hitachi Pow er E urope,
installed at Schwarze Pumpe pilot plant.

Fig.7 Typical results obtained by using the flammability analysis.


898 M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899
Author name / Energy Procedia 00 (2010) 000–000 7

Typical application examples are shown in Fig.7. Fig.7 shows the following, from the left:
(1) a flame picture which was provided by the experiment;
(2) calculated temperature profiles obtained by using the LES (Large Eddy Simulation) method;
(3) calculated temperature profiles obtained by using the k-ε method; and
(4) results of flammability analysis
Calculations by k-ε and LES methods were described in the literature [9, 10].
CASE I is the results of air combustion. The flame picture showed that flame stability was good. The results by
the LES method also showed that a stable flame could be obtained. LES was a good analytical technique to evaluate
the stability of pulverized coal fl ames [9]; however, the calculation load was large. CASE II was an example of
oxy-fuel combustion. According to the analysis results by the LES method, stability of the fl ame was inferior to that
o f CASE I. The difference between CASES I and II was clear. Revision of burner structure would be needed for
CASE II. Calculated results of the k-ε method were also compared. The difference between CASES I and II was
not clear. The calculation load of k-ε was small; however, it was hard to evaluate the flame stability. The results of
fl ammability analysis are also shown in the figure. A clear difference was obtained fo r the calculated flame
propagation velocities of CASES I and II. Flame propagation velocity of CASE II was inferior to that of CASE I.
This conclusion accorded with that of the LES method. The flammability analysis also used the k-ε method, but
evaluation of the flame stability was enabled.
CASE III is another example of oxy-fuel combustion. The burner structure was revised based on the results of
the flammability analysis. Clear difference was observed in flame propagation velocity of CASES II and III. Flame
stability of CASE III was improved. CASE III was a DS ®T-burner designed by Hitachi Power Europe, installed at
Schwarze Pumpe pilot plant.

3.4 More case studies


(a) Burner structure (b) Initial design (c) A fter revision
enlarged
Recirculation region

=0 Recirculation 1
Coal:
Norm alized Coal

region
Oxygen
Sb (m /s) Norm alized C oal (vol%)

lignite r 20 FSR
concentration

Gas FSR Coal flow


Coal Flow Coal Flow Flow 0.5 moved
(revision) 0
(initial) outside
concentration

1
Gas
Flow 0
Flam e
by controlling
stabilizer 0 0.4 local oxygen
Sb (m /s)

(FSR) distribution
0.4
0.2 m inim um -Sb
0.2 m inimum -Sb
Ignition
point 0 0
0 1 2 0 1 2
Exhaust
gas + O 2 Norm alized r (-) Normalized r (-)

Exhaust gas Exhaust gas + O 2 Exhaust gas Exhaust gas + O 2


Exhaust gas + coal + O2 Exhaust gas + coal + O2
Exhaust gas + coal

Fig.8 Examples of CASE studies.

Figs.8 show examples of CASE studies that applied this technique to a burner design for lignite-fi red oxy-fuel
combustion. Fig.8(a) shows examples of the burner design. Pulverized coal was supplied by combustion flue gas,
and injected from a burner exit. Usually, oxygen concentration of the flue gas was less than 10 vol%. Oxygen for
combustion was mixed with the flue gas, and supplied by roughly two systems. Some gas was supplied from the
central part of the burner. The remaining gas (secondary gas) was supplied from the circumference side of the coal
fl ow. In many cases, the secondary gas was injected as swirl flow. A flame stabilizer was arranged between the
coal fl ow and the secondary gas flow. Recirculation regions were formed downstream from the fl ame stabilizer.
Coal concentration, gas composition and temperature profiles were obtained by the CFD calculation (k-ε method).
M. Taniguchi et al. / Energy Procedia 4 (2011) 892–899 899
8 Author name / Energy Procedia 00 (2010) 000–000

Flame propagation velocity profile was obtained from this information in the burner neighborhood shown as the
dotted line. The position and size of the recirculation region were judged from calculated flow velocity profiles.
Flame stability became good when flame propagation velocity near the recirculation region increased.
The flame propagation velocity profile obtained with the initial design is shown in Fig.8(b). Flame propagation
velocity was lower than the minimum-Sb value in most of the area. It was hard to obtain a stable flame. Coal and
oxygen concentration profiles are also shown in Fig.8(b). Oxygen concentration was low where coal concentration
was high and vice versa. In this system, coal was supplied by combustion flue gas. It was easy to lower the
oxygen concentration where many coal particles were fl owing. Coal concentration should be increased where
oxygen concentration was high.
Fig.8(c) shows flame propagation velocity and coal concentration profiles obtained after revision (detailed results
o f CASE III in Fig.7). A part of the coal particle flow was separated from the flue gas flow (lower oxygen
concentration). These coal particles could flow to the recirculation region (higher oxygen concentration) by
modification of burner structure. The recirculation region was enlarged by modification of secondary flow. Flame
propagation velocity became considerably larger than the minimum-Sb value because the coal concentration
increased where the oxygen concentration was large. By using the flammability analysis, we could quickly clarify
improvement points of the burner.

4. Conclusion
We developed a model to predict lean fl ammability limit L and fl ame propagation velocity Sb fo r oxy-fuel
combustion conditions based on fundamental experimental data. The proposed model was verified with data of both
fundamental and pilot- and actual-scale experiments. The model could predict both fl ame propagation velocities
and lean flammability limits for N2 /O2 and CO2 /O2 combustion systems.
We also applied the model to develop burner systems for lignite-fi red oxy-fuel combustion. Local Sb and L near
the ignition points of the burner systems could be obtained from concentration and temperature profiles of the
general CFD results (k-ε method). Flame stability was judged by the calculated Sb and L profiles, and past results of
blow-off limits obtained from actual - and pilot-scale experiments. We proposed this technique as flammability
analysis. By using the technique, we could quickly clarify improvement points of the burner systems. The
calculated results were applied to a DS ®T -burner designed by Hitachi Power Europe, installed at Schwarze Pumpe
pilot plant.

References
1) Tigges KD, Klauke F, Bergins, Busekrus CK, Niesbach, Ehmann JM, et al. Conversion of existing coal-fired
power plants to oxyfuel combustion: case study with experimental results and CFD-simulations. Energy Procedia;
2009; 1: 549-56.
2) Strömberg L, Lindgren G, Jacoby J, Giering R, Anheden M, Burchhardt U, et al. Update on Vattenfall’s 30MWth
oxyfuel pilot plant in Schwarze Pumpe. Energy Procedia; 2009; 1: 581-9.
3) Suda T, Masuko K, Sato J, Yamamoto A, Okazaki K. Effect of Carbon dioxide on fl ame propagation of
pulverized coal clouds in CO2 /O2 combustion. Fuel; 2007; 86: 2008-15.
4) Taniguchi M, Kobayashi H, Azuhata S. Laser ignition and flame propagation of pulverized coal dust clouds. In:
26th Symp. (Int.) on Combustion. Pittsburgh: The Combustion Institute; 1996, pp. 3189-95.
5) Taniguchi M, Kobayashi H, Kiyama K, Shimogori Y. Comparison of flame propagation properties of petroleum
coke and coals of different rank. Fuel; 2009; 88: 1478 -84.
6) Taniguchi M, Yamamoto K. Fundamental research on oxy-fuel combustion: The NOx and coal ignition reactions.
In: Grace CT, editor. Coal Combustion Research, New York: Nova Science Publishers Inc; 2010, in press.
7) Kiyama K, Tsumura T , Sei N, Taniguchi M, Nomura S. Combustion technology fo r high fu el ratio coal.
Proceedings of International Conference on Power Engineering -97 2, Tokyo: 1997, pp. 477-81.
8) Kiga T, Makino K, Okushima K, Abe T, Hirano T, Takagi T. Development of IHI wide range pulverized coal
burner. Ishikawajima Harima Giho; 1987; 27: 333-8.
9) Yamamoto K, Taniguchi M, Kobayashi H, Sakata T, Kudo K. Validation of coal combustion model by using
experimental data of utility boilers. JSME International Journal Series B; 2005; 48: 571-8.
10) Yamamoto K, Murota T, Okazaki T, Taniguchi M. Large eddy simulation of a pulverized coal jet flame ignited
by a preheated gas flow. In: 33rd Symp. (Int.) on Combustion. Pittsburgh: The Combustion Institute; 2010, in press.

You might also like