Вы находитесь на странице: 1из 10

Article

Cite This: Energy Fuels XXXX, XXX, XXX−XXX pubs.acs.org/EF

Computational Fluid Dynamics Study of Biomass Combustion in a


Simulated Ironmaking Blast Furnace: Effect of the Particle Shape
Yiran Liu and Yansong Shen*
School of Chemical Engineering, University of New South Wales, Sydney, New South Wales 2052, Australia

ABSTRACT: Biomass is a carbon-neutral fuel and has potential to be used in pulverized coal injection (PCI) technology in
ironmaking blast furnaces (BFs). In comparison to pulverized coal particles, biomass particles vary considerably in particle shape,
and thus, the change of the aspect ratio of biomass particles may affect the motion and conversion of biomass particles. In this
study, a computational fluid dynamics (CFD) model is developed to simulate the flow and thermochemical behaviors related to
biomass injection into BFs. The model features non-spherical particle shapes and an improved devolatilization model. The model
is then applied to a pilot-scale test of charcoal injection using a pilot-scale PCI test rig under simulated BF conditions for model
validation. The burnout comparisons between simulation and measurement indicate that, in comparison to spherical particles,
the non-spherical particles show smaller burnouts as a result of the shorter traveling time in the chamber; moreover, the burnouts
predicted by the model considering cylindrical particles and improved devolatilization model are more comparable to the
measurements. This confirms the model validity and also concludes that it is necessary to include the effect of the particle shape
in the modeling of biomass injection in BFs. Then, to give a full picture of charcoal injection, other typical phenomena
of flow and combustion behaviors are analyzed in detail for charcoals in cylindrical particle shape, aspects of gas−charcoal flow,
their temperature distribution, and gas composition distribution. This model provides an effective way for understanding and
optimizing biomass injection in BF practice.

1. INTRODUCTION exceeded coal at the same volatile matter (VM) content,13


As an environmentally friendly and carbon-balance renewable and the solution loss reaction for biomass went faster than
energy, biomass comes from a variety of sources in the world.1,2 pulverized coal,14 making full substitution possible.15 BHP
It has been widely used in coal-fired boilers for power developed a pilot-scale PCI test rig, tested the charcoal
generation.3 On the other hand, blast furnaces (BFs) are the injection (Figure 1a), and found that charcoal injection into
predominant route of ironmaking, representing the main BF lowered the pressure drop and its fluctuation amplitude.16,17
energy consumer, nearly 70% of the whole plant as a result Moreover, charcoal injection was tested under actual produc-
of coke combustion.4 By means of pulverized coal injection tion conditions using mini-BFs in Brazil,18 where the charcoal
(PCI), coal is injected into the raceways to partially replace injection rate of 100−190 kg/tonne of heavy metal (HM) was
expensive metallurgical coke.5−7 As a result, a large amount achieved19 and 200−225 kg/tonne of HM should be feasible
of CO2 is emitted from the combustion of coal and coke. in large BFs.20 However, all of these experiments conducted
Therefore, biomass, as a carbon-balance source, has potential to at either lab or plant scale failed to provide the detailed
be used in ironmaking BFs. That is, substituting traditional information on multiphase flow and thermochemical processes
pulverized coal with biomass is significant for not only meeting related to charcoal injection, such as flow distribution, turbulent
the energy needs of ironmaking BFs but also the mitigation of mixing, reaction zones, and species distribution.21
greenhouse gases.8,9 Practically, biomass injection technology At present, mathematical modeling, especially computational
was reported to decrease CO2 emission by 21−28%10 up to fluid dynamics (CFD), was used to overcome these problems.
37−45%11 compared to PCI technology. However, biomass Wijayanta et al.22,23 reported a two-dimensional (2D) CFD
and coal differ significantly from physicochemical properties, investigation of pulverized biochar injection into a BF and
such as the size, density, and especially non-spherical shape, found that a longer raceway was required to obtain the same
where the latter could greatly affect the motion and conversion combustibility of biochar as coal. However, this 2D model did
of fuel particles in combustors.12 On the other hand, the flow not consider the three-dimensional (3D) features and can
and thermochemical behavior of PCI operation in BF differs only generate useful results for practice. Wu et al.24 developed a
significantly from coal combustion in boilers, such as a higher 3D CFD model to evaluate the flow characteristics of charcoal
temperature, more intense heating rate and mixing, larger utilization in a PCI blowpipe and tuyere and optimize the posi-
velocity, and more complex gas concentration. Therefore, it is tion of the blowpipe. However, the raceway was not considered.
of importance to understand the thermochemical behavior
of biomass injection in BFs and, in particular, consider the Special Issue: 6th Sino-Australian Symposium on Advanced Coal and
irregular shape in studying the biomass injection in BFs. Biomass Utilisation Technologies
Biomass combustion has been studied under BF conditions
using various methods. Lab- or pilot-scale experimental studies Received: October 15, 2017
of biomass combustion in BF indicated that the combustion Revised: December 7, 2017
performance of the charcoal, a biomass-derived product, Published: December 12, 2017

© XXXX American Chemical Society A DOI: 10.1021/acs.energyfuels.7b03150


Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

Figure 1. Schematic of the PCI test rig: (a) schematic of the PCI test rig and (b) geometry details of the computational domain (mm).30

Table 1. Governing Equations for the Gas and Particle Phases in the Model25
for the Gas Phase

mass
∇(ρU) = ∑ ṁ
np

⎛ 2 ⎞
∇(ρUU) − ∇((μ + μt )(∇U + (∇U)T )) = −∇⎜p + ρk ⎟ + ∑ fD
momentum ⎝ 3 ⎠ np

⎛ ⎛ λ μ ⎞ ⎞
energy ∇⎜⎜ρUH − ⎜ + t ⎟∇H ⎟⎟ = ∑q
⎝ ⎝ CP σH ⎠ ⎠ np

⎛ ⎛ μ ⎞ ⎞
gas species i ∇⎜⎜ρUYi − ⎜⎜Γi + t ⎟⎟∇Yi ⎟⎟ = Wi
⎝ ⎝ σYi ⎠ ⎠

⎛ ⎛ μ⎞ ⎞
turbulent kinetic energy ∇⎜⎜ρUk − ⎜μ + t ⎟∇k ⎟⎟ = (Pk − ρε)
⎝ ⎝ σ k⎠ ⎠
⎛ ⎛ μ⎞ ⎞ ε
turbulent dissipation rate ∇⎜⎜ρUε − ⎜μ + t ⎟∇ε⎟⎟ = (C1Pk − C2ρε)
⎝ ⎝ σε ⎠ ⎠ k
for a Particle in the Particle Phase
dm p
mass = − ṁ
dt
dUp
mp = −fD
dt
momentum
1
−fD = πd p2ρC D|U − Up|(U − Up)
8
dTp
mpCp = −q
dt
energy
dm p
− q = πd pλNu(Tg − Tp) + ∑ Hreac + A pεp(πI − σBTp 4)
dt

Shen et al.25 developed a 3D CFD model to simulate the the charcoal type, such as the VM content. Castro et al.26,27
injection of different charcoals in BFs to evaluate the effect of reported CFD studies of simultaneous injection of pulverized
B DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Table 2. Reactions Considered in This Charcoal Injection Model with Respective Reaction Kineticsa
reaction reaction rate expression reaction kinetics38

k
⎛ E ⎞
moisture(l) → H 2O(g) k = A exp⎜− ⎟ A = 5.13 × 108 s−1; E = 88 kJ mol−1
⎝ RTP ⎠
Energy & Fuels

k v1
⎛ E ⎞
charcoal ⎯→
⎯ α1fuel gas + (1 − α1)char k = A exp⎜− ⎟ A1 = 4.3 × 107 s−1; E1 = 136 kJ mol−1; α1 = VM (daf); CS = 0
⎝ RTP ⎠

k v2 d ṁ
charcoal ⎯→
⎯ α2 fuel gas + (1 − α2)char d = CSd0 ref A2 = 1.46 × 1013 s−1; E2 = 251 kJ mol−1; α2 = Q × VM (daf) + 0.14; CS = 0
dt mref,0
⎛ E ⎞
CH4 + 0.5O2 → CO + 2H2 k = [i]A exp⎜− ⎟ A = 2.3 × 107 s−1; E = 125.5 kJ mol−1; [i] = [CH4]−0.3[O2]1.3
⎝ RTP ⎠

ε ⎛ [i] ⎞
CO + 0.5O2 → CO2 ri = CA min⎜ ⎟ A = 1014.6 s−1 mol−1 cm2.25; E = 167.3 kJ mol−1; [i] = [CO][O2]0.5
κ ⎝ vt′ ⎠
CO2 → CO + 0.5O2 A = 5 × 108 s−1; E = 167.3 kJ mol−1; [i] = [CO2]1
H2 + 0.5O2 → H2O A = 540 m3 kg−1 s−1; E = 15.1 kJ mol−1; [i] = [H2][O2]0.5
ε ⎛ [i] ⎞
H2O → H2 + 0.5O2 ri = CA min⎜ ⎟ CA = 4.0; [i] = [H2O]
κ ⎝ vt′ ⎠

2(ϕ − 1) ⎛ T⎞
ϕchar + O2 → 2(ϕ − 1)CO + (2 − ϕ)CO2 = AS exp⎜⎜− s ⎟⎟ Ac = 140 m s−1 K−1; Tc = 21580 K; As = 2500; Ts = 6240 K

C
2−ϕ ⎝ Tp ⎠
dmC 3ϕ MC ρ∞ −1
char + CO2 → 2CO =− (k1 + (k 2 + k 3)−1mC) Ac = 202300 m s−1 K−1; Tc = 39743 K; As = 0.0004; Ts = 6240 K
dt 1− e MO2 ρC
char + H2O → CO + H2 Ac = 6069 m s−1 K−1; Tc = 32406 K; As = 0.0004; T s = 6240 K
a 2 2 0.5
k1 = D/RP ; k2 = (1 − e)kc/RP; k3 = kcTp(β coth β − 1)/β α; kc = AcTp exp(−Tc/Tp); β = R(kc/Dpea) ; Dp = effic × D; D = Dref/ρfluid((Tp + Tg)/2Tref)α.
Article

DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

coal and charcoal into a BF. However, in these model studies, (ii) Devolatilization is modeled using two-competing reaction
the irregular shape of biomass particles that may greatly affect models30,35 rather than a one-step global model used in the past.25
the motion and conversion of fuel particles in combustors28,29 In the experimental study,10 it was found that the charcoal particles
was not considered. To date, the studies considering irregular used in the test do not swell much during the devolatilization. Thus, in
this numerical study based on the conditions of ref 10, the swelling
particle shape have not been conducted for biomass injection coefficient of charcoal particles is set to zero (Cs = 0). (iii) Gaseous
into BFs. combustion reactions of fuel gas are modeled by the three-fuel gas
In this study, a 3D CFD model is improved to describe the model and controlled by the combined eddy dissipation model and
flow and thermochemical behaviors related to biomass injec- finite rate chemistry. (iv) Char oxidation and gasification are modeled
tion in BFs and study the effects of non-spherical particle by Gibb’s model. The models are summarized in Table 2. The models
shapes. The model is applied to charcoal in a pilot-scale PCI are developed on the basis of the framework of software package
test rig. The comparisons against measurements show that the ANSYS-CFX, version 17.2.36
simulations using the cylindrical particle shape give the most
comparable burnouts to the measurements. Then, other key phe- 3. SIMULATION CONDITIONS
nomena of injecting charcoal in cylinder shape are described, 3.1. Properties of Charcoal. In this study, the model of
including gas−particle flow, temperature, and concentrations. biomass injection is applied to charcoal in the base cases for
model validation. Charcoal, derived by carbonizing biomass, is
2. MODEL DEVELOPMENT typically low in oxygen, VM, and ash contents compared to
2.1. Model Outline. The mathematical model of the biomass biomass/cellulosic materials, while they all tend to remain the
injection is based on our previous PCI models.30−32 The governing original shape. The proximate (ad) and ultimate (daf) analyses
equations for the gas−solid flow and associated heat transfer and the and mean particle size (de) of the pulverized charcoal are listed
reactions of charcoal are outlined below for completion. in Table 3. For the injectant, 50 particle size groups ranging
The gas phase is modeled using a 3D steady-state Reynolds-
averaged Navier−Strokes equation closed by a standard k−ε
Table 3. Proximate (ad) and Ultimate (daf) Analyses of the
turbulence model. The particle phase is modeled using the Lagrangian
method, where the trajectories of the discrete particles are determined Pulverized Charcoal Used in This Study25
by integrating Newton’s second law of motion, where drag force moisture ash VM fixed carbon calorific value density
and turbulence dispersion are included. The heat transfer of the (%) (%) (%) (%) (MJ/kg) (kg/m3)
gas−particle phase is modeled by considering three physical processes: 9.3 1.8 4.3 84.6 32.48 1150
convective heat transfer, latent heat transfer associated with mass
C (%) H (%) O (%) N (%) S (%)
transfer, and radiative heat transfer. Full coupling of continuity,
momentum, and energy are applied to the particle−gas phase inter- 95.0 0.99 3.7 0.33 0.03
action. The models are summarized in Table 1.
2.2. Model Improvements. Some model improvements specific
to biomass combustion are also included in this part, including the from 1 to 250 μm are sampled and 500 representative particles
model of the non-spherical particle shape and the improved devolatil- are tracked in the simulations. In this model, de is the volume-
ization model. equivalent sphere diameter for different particle shapes,
2.2.1. Non-spherical Particle Shape. Unlike coal that will soften 27 μm in this study. In this study, charcoal is in the shape
and form a spherical particle shape during devolatilization, biomass of a cylinder, as shown in the scanning electron microscopy
particles will largely retain their initial irregularly shaped form.33 (SEM) characterization.10
This large aspect ratio may affect gas−particle drag force and heat and 3.2. Geometry and Boundary Conditions. The model is
mass transfer correlations. This necessitates the introduction of a non- applied to a pilot-scale PCI test rig (Figure 1a).17 The test
spherical shape factor. In this model, two shape factors, namely, cross-
rig was designed to simulate the flow and combustion in the
sectional area factor and surface area factor, are used to describe the
influence of the particle shape on the drag force and mass and heat region of lance−tuyere−raceway (along the coal plume)
transfer correlations, respectively. The former factor is multiplied by related to PCI operation under simulated BF conditions.
the calculated cross-sectional area when assuming spherical particles. The computational domain (Figure 1b) includes lance,
The latter factor is defined as the ratio of the surface area of the non- tuyere, and raceway centerline. The main duct is used for
spherical particle to the surface area of the spherical particle with the injecting the hot blast (i.e., oxygen-enriched air). The coaxial
equivalent diameter. With the cylindrical spherical particle taken for lance is installed at an inclination angle of 6° with respect
example, the two shape factors are defined as eqs 1 and 2. The particle to the centerline of the duct. The inner tube is used for injec-
shape factors are considered in the governing equations in the form of ting pulverized charcoal and the conveying gas (i.e., 100%
the volume-equivalent sphere diameter of particle, de. For consistency
nitrogen), and the outer tube is used for injecting cooling gas
and tidiness, their definitions (i.e., eqs 1 and 2) are not specified in the
governing equations in Table 1. (i.e., air). The geometry is plane-symmetric. The exit at the
end of the chamber is set as an outlet, and the wall of the
scross‐cylinder πr 2 chamber is assumed non-slippery and adiabatic. A boundary-
fcross‐sectional = = fitted, multi-block structured finite volume mesh is used, with
Scross‐sphere π(1/2de)2 (1)
a highly fine mesh around the lance and tuyere and along the
centerline of the chamber. Table 4 shows main parameters of
ssurface cylinder 2πr 2 + 2πrd operating conditions.
fsurface = =
Ssurface sphere 4π(1/2de)2 (2) 3.3. Model Validation. The charcoal injection model is
validated against the measured burnouts obtained from the
2.2.2. Improved Devolatilization Model. Similar to pulverized
coal combustion, biomass combustion is considered to be a four-stage pilot-scale test rig using two charcoal materials, as shown in
chemical reaction; however, the parameters of reactivity and inner Table 5.10,25 The burnout is calculated according to ash
particle structure are quite different, which gives highly different balance, as below. The ash data were collected at three positions,
reaction rates.34 They are modeled by means of the following: (i) The +50 mm, centerline, and −50 mm, at 925 mm downstream from
moisture evaporation process is controlled by finite rate chemistry. the injection point.
D DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

Table 4. Operating Conditions in the PCI Test Rig Used in (case 2), and flat chip (case 3), where the cylindrical shape is
This Study26 common for charcoal particles and the flat chip is common for
wood chip particles: case 1, spherical charcoal injection; case 2,
temperature
flow rate (Nm3/h) (K) O2 (%) cylindrical charcoal injection; and case 3, flat-chip-shaped
charcoal injection.
blast gas 300 1473 20.9
cooling gas 3.2 600 20.9 eat‐upCO = mfCO /(mfCO + mfCO ) (4)
2 2
conveying gas 2 323 0 (i.e., 100% N2)
coal/charcoal 38 kg/h 320 Figure 2 compares the burnout evolutions along the centerline
of the raceway between the three cases. They are discussed in
Table 5. Experimental Conditions for Model Validation
blast
charcoal blast rate temperature blast O2 charcoal rate
case materiala (Nm3/h) (K) (%) (kg/h)
1 charcoal 1 301 1474 22.4 33.5
2 charcoal 1 304 1475 22.5 25.2
3 charcoal 2 302 1478 22.4 29.4
4 charcoal 2 301 1477 22.3 48.4
a
The injectants: charcoal 1, referenced in Table 3; charcoal 2, 9.5%
moisture, 2.9% ash, 10.2% VM, 77.4% fixed carbon, and de of 25 μm.

⎛ ma,0 ⎞
burnout = ⎜1 − ⎟ /(1 − ma,0)
⎝ ma ⎠ (3)
Table 6 compares the measured burnouts with the predicted
burnouts of particles in different shapes, i.e., sphere and cylinder.
Figure 2. Comparison of burnout evolutions along the centerline
Table 6. Model Validation in Terms of Burnout for Different among spherical (case 1), cylindrical (case 2), and flat-chip-shaped
Particle Shapes (case 3) charcoals.

predicted burnout (%)


two aspects: overall trend and difference between them. First, it
measured burnout sphere-shaped cylinder-shaped is indicated that the overall trends are similar for all three curves
number (%) charcoal charcoal
of the three cases. That is, there are two fast increases along the
1 70.1 72.12 70.79 burnout curves around 0.1−0.3 and 0.7−1.0 m. They resulting
2 80.8 81.44 79.82 from the release and combustion of VM at a lower temperature
3 70.4 75.10 73.75 upstream and higher temperature downstream, along the two
4 72.1 80.40 71.61 competing reaction paths in the two-competing reaction model
of devolatilization (Table 2), respectively. Second, the burnout
It is indicated that the improved model can reasonably predict evolutions along the centerline and burnout at 1.1 m, traveling
burnouts for both spherical and cylindrical particles. Moreover, time, and maximum temperature are then compared among
it is also indicated that, in comparison to spherical particles, spherical (case 1), cylindrical (case 2), and flat-chip-shaped
the burnout predicted by cylindrical particle shapes is more (case 3) charcoals. It is indicated that the burnout evolution of
comparable to the measurements; that is, it is necessary to con- spherical charcoal particles is always larger than those of the
sider the particle shape when injecting non-spherical materials two irregularly shaped charcoal particles along the centerline.
into the BF. The same trend can be found in burnout at the end of the
raceway, i.e., 1.1 m, and maximum temperature. The differ-
4. RESULTS AND DISCUSSION ence may result from the longer traveling time of the sphe-
After the model is validated against the measurements from rical particle of 0.404 s than those of the corresponding
the pilot-scale test rig, the model is then used to investigate the irregularly shaped cases, 0.328 and 0.241 s, as shown in Table 7.
effects of the particle shape on the flow and thermochemical
behaviors of biomass injection in BFs and the significance of Table 7. Comparisons of Burnout at 1.1 m, Traveling Time,
improving the devolatilization model from a one-step global and Maximum Temperature among Spherical (Case 1),
reaction model to a two-completing reaction model in the Cylindrical (Case 2), and Flat-Chip-Shaped (Case 3)
modeling of biomass injection in BFs. Finally, to give a full Charcoals
picture of charcoal injection into BFs specifically, in additional
predicted burnout at traveling time maximum
burnout and eat-up, other key flow and thermochemical behaviors number 1.1 m (%) (s) temperature (K)
are overviewed for cylindrical charcoals under simulated BF
1 65.50 0.404 1900
conditions. 2 64.74 0.328 1882
4.1. Effects of the Particle Shape. To clarify the effects of 3 64.39 0.241 1872
particle shapes in the modeling of biomass injection in BFs,
burnout and the so-called “eat-up” (defined as the rate of gas
conversion, shown in eq 4) along the centerline are compared This is because the deviation of a particle from the spherical
for charcoal particles in the shapes of a sphere (case 1), cylinder shape causes a decrease of its terminal velocity in the fluid.
E DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

This implies a higher drag force of the non-spherical particles


compared to the spherical particles.37 It should be noted that
the small difference in traveling time as a result of the PCI
process is a highly intense process, where the mixing rate is very
high and the traveling time is very short.
Figures 3 and 4 compare the effect of particle shapes on gas
concentrations among the three cases, along the symmetry

Figure 5. Comparison of “eat-up” among sphere (case 1), cylinder


(case 2), and flat chip (case 3).

Figure 6. Comparison of burnout evolution of cylindrical charcoals


using two different devolatilization models.

Figure 3. Comparison of the CO2 concentration among sphere (case 1),


cylinder (case 2), and flat chip (case 3).

Figure 4. Comparison of gas composition along the centerline among Figure 7. (a) Gas velocity vector and (b) charcoal particle trajectories
sphere (case 1), cylinder (case 2), and flat chip (case 3). colored by the particle size in the case of cylindrical charcoal injection.

F DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

concentrations than those of the spherical particle cases.


The non-spherical charcoal particle cases with a relatively
smaller burnout, especially the cylindrical case, show a slower
conversion from O2 to CO2 and CO, as shown in Figure 4.
However, in Figure 5, the comparison between “eat-up” curves
(rate of gas conversion) along the longitudinal direction of
different shaped particles shows that, when the particle shape is
taken into consideration, the rate of gas consumption of the
cylindrical particle is higher than those of other two cases in
the upstream (in the range of 0−0.4 m). In the downstream,
although the “eat-up” of non-spherical particles has a larger
value than that of the spherical counterpart, the final value
of eat-up is similar. In view of “eat-up” of one particle shape, a
rapid gas conversion upstream and slow convergence down-
stream are found for charcoal, indicating that homogeneous
reactions are the dominant reactions in determining the burnout
of charcoal.
4.2. Effect of the Improved Devolatilization Model.
The one-step global model is a relatively simple model
widely used in the modeling of coal combustion. This may
lead to insufficient accuracy in the prediction of the product
yield distributions with temperature and heating conditions.
The two-competing reaction model includes a primary path and
a secondary path, during which biomass materials undergo
Figure 8. (a) Gas-phase temperature contours and (b) radiation inten-
additional cracking to produce gases, and, thus, has potential to
sity in the case of cylindrical charcoal injection. overcome the inaccuracy in the one-step global model.
Figure 6 compares the burnout evolutions predicted by
the two models, i.e., one-step global devolatilization and
two-competing reaction devolatilization, and the average
burnout measurement at 0.925 m in the validation case 1
(Table 6), i.e., 70.1%. It can be seen that the predicted burnout
using the two-competing reaction model is comparable to the
measurement, while the model using the one-step global model
gives a much lower burnout evolution. That is, the present
model using the two-competing reaction model with a suitable
Q factor is adequate to illustrate the impacts of shape on
charcoal combustion. Thus, the two-competing reaction model
Figure 9. Gas-phase temperature contours in the case of pulverized rather than the one-step global model is suggested for the
coal injection.30 modeling of biomass combustion in the future.
4.3. Typical Results of Other Flow and Thermochem-
plane and along the centerline, respectively. It is indicated that ical Behaviors of Charcoal Particles in a Cylindrical
the non-spherical charcoal injection cases obtain smaller CO2 Particle Shape. In this study, the charcoal particles are largely

Figure 10. Contours of gas concentrations of cylindrical charcoal injection, in terms of (a) H2O, (b) H2, (c) CO, and (d) O2.

G DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

4.3.1. Flow Patterns. Figure 7a shows the velocity vector of


the gas phase along the symmetry plane for cylindrical charcoal
injection. The results show the velocity of the conveying gas at
the lance tip to be ∼9 m/s and the gas velocity at the tuyere to
be ∼125 m/s. The gas stream forms a high-speed inclined jet
after exiting the tuyere, and then the jet starts to expand in the
radial direction and finally dispersing near the exit. A low-speed
recirculation occurs near the wall. Similar phenomena are also
observed in the particle flow (Figure 7b), where the charcoal
plume is concentrated until exiting the jet and then starts
to rapidly disperse, mainly fine particles toward the exit of
the chamber. Some fine particles recirculate above, behind, and
below the jet.
4.3.2. Temperature Field. Figure 8a shows the gas-phase
temperature along the symmetry plane for cylindrical charcoal
injection. The asymmetric temperature distribution shows
that the temperature in the lower part is higher toward the
exit, which is due to the inclined lance arrangement. The high-
temperature part surrounding the lance tip corresponds to the
combustion of volatile, which starts to release after a preheating
stage of a high heating rate in front of the tuyere. Different from
PCI simulations, the charcoal injection does not show a flame-
front high temperature over the coal plume surface in front of
tuyere, as observed in Figure 9.30 In the chamber, the high
temperature resulted from the combustion of the recirculated
small charcoal particles. In the downstream, the highest temper-
ature was achieved because of the combustion of char released
from the charcoal particles. From Figure 8b, similar to the
temperature field, the radiation intensity contour shows that
there are different dominant reactions along the distance from
the lance tip.
4.3.3. Gas Concentrations. Figure 10 shows the gas species
distributions (H2O, H2, CO, and O2) along the symmetry plane
in charcoal injection. As the combustion proceeds, the moisture
evaporation process for charcoal combustion resulted in releas-
ing more H2O at the upstream, which is observed at the upstream
for the charcoal injection (Figure 10a). The significant moisture
content makes the evaporation process obvious and may delay
the ignition process. As one of the products of devolatilization
of charcoal, H2 has two relatively high-concentration parts
along the symmetry plane (Figure 10b), which may result from
the two-competing reaction model of devolatilization. Partial
O2, CO, and CO2 gasification conditions produced similar con-
centration profiles (panels c and d of Figure 10 and Figure 3b),
and the following changes in the concentration were observed:
rapid consumption of the O2 concentration and a rapid increase
of CO and CO2 at the downstream. This indicates a fast devola-
tilization reaction at a low temperature and a fast char com-
Figure 11. Combustion characteristics along the centerline for bustion at a higher temperature.
different size groups for cylindrical charcoal injection: (a) burnout,
(b) VM content, and (c) particle temperature. 4.3.4. Combustion Characteristics Analyzed by Particle
Size Groups. Combustion characteristics related to charcoal
in a cylindrical shape, as observed in ref 10, as confirmed in the injection are further analyzed by particle size groups. Figure 11
model validation that the predicted burnout using a cylindrical shows the combustion characteristics in terms of burnout, VM
particle shape is more comparable to the measurements. Some content, and particle temperature in six particle size groups.
typical results of a cylindrical particle shape have been discussed Although similar to the overall performance of burnout, the
above, including burnout evolution and gas conversation. burnout evolution curve for each size group (Figure 11a)
In this section, other typical results of flow and thermochemical varies for different size groups. The particle of a smaller size
phenomena related to charcoal injection using a cylindrical has a larger specific surface area and faster devolatilization
shape are studied under simulated BF conditions, including (Figure 11b) and, thus, achieves a higher particle temperature
spatial distribution of flow and temperature distributions in the (Figure 11c). The smaller particles are going through a higher
entire chamber and, more quantitatively, evolutions of burnout, heating rate (Figure 11c) and start the devolatilization process
VM content, and particle temperature along the centerline. earlier than larger particles (Figure 11b); thus, the burnout of
H DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

smaller particles starts earlier and achieves a higher value than d = particle diameter (μm)
that of larger particles. de = volume-equivalent sphere diameter (μm)
e = void fraction of char particles
5. CONCLUSION E1 and E2 = activation energy of devolatilization reactions
A 3D CFD model is developed for describing the injection of (kJ/mol)
irregularly shaped charcoal particles and applied to a pilot-scale fD = drag force from a particle (N)
PCI test rig to simulate the charcoal injection under simulated fcross‑sectional = cross-sectional area factor
BF conditions. The model is validated against the measure- fsurface = surface area factor
ments in terms of burnout. The model is then used to H = enthalpy (J kg−1)
investigate the effects of particle shapes on irregularly shaped Hreac = reaction heat (J kg−1)
charcoal particles and the significance of the improved I = radiation intensity on the particle surface (W m−2)
devolatilization model and illustrate other flow and thermo- [i] = molar concentration of component i
chemical behaviors of charcoal particles in a cylindrical par- K = turbulent kinetic energy (m2 s−2)
ticle shape. The following is found: (1) This model with the kv1 and kv2 = devolatilization rate constants (s−1)
introduced shape factors can better describe the evolutions k1 = rate of external diffusion in the Gibb model (s−1)
of irregularly shaped charcoal particles in terms of burnout. k2 = rate of surface reaction in the Gibb model (s−1)
For the charcoal injection, the burnouts using a cylindrical k3 = rate of internal diffusion and surface reaction in the
shape are more comparable to the measurements, consistent Gibb model (s−1)
with the practice. (2) The present model using the two-competing kc = carbon oxidation rate in the Gibb model (m s−1)
reaction model with a suitable Q factor is more suitable to ṁ = mass transfer rate from a particle (kg s−1)
illustrate the impacts of particle shape on charcoal combustion ma = ash mass fraction
than that using the one-step global model. (3) In comparison to ma,0 = original ash mass fraction
spherical charcoal cases, the two cases of irregularly shaped mc = mass of char (kg)
charcoal particles have a slower combustion process and lower mfCO = mass fraction of carbon monoxide
burnouts. (4) The significant moisture content of charcoal mfCO2 = mass fraction of carbon dioxide
makes the evaporation process obvious at the upstream and Mc = molecular weight of carbon
may delay the ignition process for the charcoal injection. MO2 = molecular weight of the oxygen molecule
(5) Different from PCI simulations, the charcoal injection does np = particle number per unit volume (m−3)
not show a flame-front high temperature over the coal plume Nu = Nusselt number
surface in front of tuyere. Pk = turbulence production by viscous force


p = pressure (Pa)
AUTHOR INFORMATION q = heat transfer from a particle (W)
Corresponding Author Q = Q factor for the devolatilization model
*E-mail: ys.shen@unsw.edu.au. rp = particle radius (m)
ri = reaction rate of gas species i (mol m−3 s−1)
ORCID
Re = Reynolds number
Yansong Shen: 0000-0001-8472-8805 T = temperature (K)
Notes Tblast = blast temperature (K)
The authors declare no competing financial interest. Tc = activation temperature (E/R) in the Gibb model (K)

■ ACKNOWLEDGMENTS
The authors acknowledge the financial support from the
Tg = gas temperature (K)
Tp = particle temperature (K)
Tref = reference temperature in the Gibb model (293 K)
Australian Research Council and Baosteel (LP150100112) and Ts = constant in the Gibb model (6240 K)
the first author wishes to acknowledge the financial support U = mean velocity of gas (m s−1)
from China Scholarship Council. Up = mean velocity of the particle (m s−1)

■ NOMENCLATURE
A1 and A2 = pre-exponential factors of devolatilization
u, v, and w = gas velocity components (m s−1)
VM = volatile matter
vi = stoichiometric coefficient of species i.
reactions (s−1) Wi = reaction rate of species i (per unit volume) (kg m−3 s−1)
Ac = pre-exponential factors in the Gibb model (m s−1 K−1) Yi = mass fraction of species i
Ap = particle area (m2) Greek Letters
As = constant in the Gibb model (0.0004) α = volume/internal surface area ratio in the Gibb model
a = exponent in the Gibb model (0.75) α1 and α2 = volatile yields
C0 = mass of raw charcoal (kg) ε = turbulent dissipation rate (m2 s−3)
C1 and C2 = turbulent model constants εp = particle emissivity
CD = drag coefficient λ = thermal conductivity (W m−1 K−1)
Cp = particle heat capacity (J kg−1 K−1) σB = Stefan−Boltzmann constant (5.67 × 10−8 W m−2 K−4)
D = external diffusion coefficient of oxygen in the Gibb σk and σε = turbulence model constants
model (m2 s−1) σH = turbulent Ptandtl number for enthalpy
Dref = reference dynamic diffusivity in the Gibb model (1.8 × ϕ = mechanism factor in the Gibb model
10−5 kg m−1 s−1) ρ = density (kg m−3)
Dp = pore diffusivity μ = dynamic viscosity (Pa s)
daf = dry and ash free μt = turbulent viscosity (Pa s)
I DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX
Energy & Fuels Article

Γi = molecular diffusivity of species i (kg m−1 s−1) (27) de Castro, J. A.; Araújo, G. M.; da Mota, I. O.; Sasaki, Y.; Yagi, J.
J. Mater. Res. Technol. 2013, 2 (4), 308−314.
Subscripts (28) Zhang, X.; Chen, Q.; Bradford, R.; Sharifi, V.; Swithenbank, J.
C = char Fuel Process. Technol. 2010, 91 (11), 1491−1499.
G = gas (29) Momeni, M.; Yin, C.; Kaer, S. K.; Hansen, T. B.; Jensen, P. A.;
P = particle Glarborg, P. Energy Fuels 2013, 27, 507−514.


(30) Shen, Y.; Guo, B.; Yu, A.; Maldonado, D.; Austin, P.; Zulli, P.
ISIJ Int. 2008, 48 (6), 777−786.
REFERENCES (31) Shen, Y.; Maldonado, D.; Guo, B.; Yu, A.; Austin, P.; Zulli, P.
(1) Jin, X.; Ye, J.; Deng, L.; Che, D. Energy Fuels 2017, 31, 2951− Ind. Eng. Chem. Res. 2009, 48, 10314−10323.
2958. (32) Shen, Y.; Guo, B.; Yu, A.; Zulli, P. ISIJ Int. 2009, 49 (6), 819−
(2) Thanapal, S. S.; Chen, W.; Annamalai, K.; Carlin, N.; Ansley, R. 826.
J.; Ranjan, D. Energy Fuels 2014, 28, 1147−1157. (33) Backreedy, R. I.; Fletcher, L. M.; Jones, J. M.; Ma, L.;
(3) Li, P.; Wang, F.; Tu, Y.; Mei, Z.; Zhang, J.; Zheng, Y.; Liu, H.; Pourkashanian, M.; Williams, A. Proc. Combust. Inst. 2005, 30, 2955−
Liu, Z.; Mi, J.; Zheng, C. Energy Fuels 2014, 28, 1524−1535. 2964.
(4) Kuang, S.; Li, Z.; Yu, A. Steel Res. Int. 2017, 87 (9999), 1−25. (34) Shen, F.; Ding, Z.; Yang, W.; Jiang, X.; Mu, L.; Shen, Y. J. Iron
(5) Kim, J.-H.; Kim, R.-G.; Kim, G.-B.; Jeon, C.-H. Exp. Therm. Fluid Steel Res. Int. 2009, 16, 753−757.
Sci. 2016, 79, 266−274. (35) Guo, B.; Zulli, P.; Rogers, H.; Mathieson, J. G.; Yu, A. ISIJ Int.
(6) De Girolamo, A.; Grufas, A.; Lyamin, I.; Nishio, I.; Ninomiya, Y.; 2005, 45 (9), 1272−1281.
Zhang, L. Energy Fuels 2016, 30, 1858−1868. (36) ANSYS, Inc. ANSYS-CFX Documentation; ANSYS, Inc.:
(7) Gao, Q.; Li, S.; Yuan, Y.; Zhao, Y.; Yao, Q. Energy Fuels 2016, 30, Canonsburg, PA, 2017.
1815−1821. (37) Tran-Cong, S.; Gay, M.; Michaelides, E. E. Powder Technol.
(8) Fitzpatrick, E. M.; Bartle, K. D.; Kubacki, M. L.; Jones, J. M.; 2004, 139, 21−32.
Pourkashanian, M.; Ross, A. B.; Williams, A.; Kubica, K. Fuel 2009, 88 (38) Shiozawa, T. Numerical Modelling of Multiphase Flow in
(12), 2409−2417. Raceway of Ironmaking Blast Furnace. Ph.D. Thesis, Materials Science
(9) Söderman, J.; Saxén, H.; Pettersson, F. Comput.-Aided Chem. Eng. & Engineering, Faculty of Science, The University of New South
2009, 26, 567−571. Wales, Sydney, New South Wales, Australia, 2013.
(10) Rogers, H.; Mathieson, J. G. BlueScope Internal Report;
BlueScope Steel Research: Port Kembla, New South Wales, Australia,
2010.
(11) Babich, A.; Senk, D.; Fernandez, M. ISIJ Int. 2010, 50 (1), 81−
88.
(12) Tabet, F.; Gökalp, I. Renewable Sustainable Energy Rev. 2015, 51,
1101−1114.
(13) Norgate, T.; Haque, N.; Somerville, M.; Jahanshahi, S. ISIJ Int.
2012, 52 (8), 1472−1481.
(14) Feliciano-Bruzual, C.; Mathews, J. A. Rev. Metal. 2013, 49 (6),
458−468.
(15) Jahanshahi, S.; Mathieson, J. G.; Somerville, M. A.; Haque, N.;
Norgate, T. E.; Deev, A.; Pan, Y.; Xie, D.; Ridgeway, P.; Zulli, P. J.
Sustain. Met. 2015, 1, 94−114.
(16) Mathieson, J.; Rogers, H.; Somerville, M.; Ridgeway, P.;
Jahanshahi, S. Proceedings of the 1st International Conference on Energy
Efficiency and CO2 Reduction in the Steel Industry (EECR Steel 2011);
Dusseldorf, Germany, June 27−July 1, 2011; pp 1−10.
(17) Mathieson, J. G.; Rogers, H.; Somerville, M. A.; Jahanshahi, S.
ISIJ Int. 2012, 52 (8), 1489−1496.
(18) Braga, R. N. B.; Goncalves, H. T.; Santiago, R.; Neto, J. Met.
ABM 1986, 42 (343), 389−394.
(19) Machado, J. G. M. S.; Osório, E.; Vilela, A. C. F.; Babich, A.;
Senk, D.; Gudenau, H. W. Steel Res. Int. 2010, 81 (1), 9−16.
(20) Nogami, H.; Yagi, J.; Sampaio, R. S. ISIJ Int. 2004, 44 (10),
1646−1652.
(21) Ahmed, P.; Habib, M. A.; Ben-Mansour, R.; Ghoniem, A. F.
Energy Fuels 2016, 30 (3), 2458−2473.
(22) Wijayanta, A. T.; Alam, M. S.; Nakaso, K.; Fukai, J.; Kunitomo,
K.; Shimizu, M. ISIJ Int. 2014, 54 (7), 1521−1529.
(23) Wijayanta, A. T.; Alam, M. S.; Nakaso, K.; Fukai, J.; Kunitomo,
K.; Shimizu, M. Fuel Process. Technol. 2014, 117, 53−59.
(24) Wu, B.; Roesel, T.; Zhou, C. Q. Proceedings of the ASME 2009
International Mechanical Engineering Congress and Exposition; Lake
Buena Vista, FL, Nov 13−19, 2009; pp 445−451, DOI: 10.1115/
IMECE2009-13107.
(25) Shen, Y.; Shiozawa, T.; Yu, A.; Austin, P. Proceedings of the 7th
International Symposium on Multiphase Flow, Heat Mass Transfer and
Energy Conversion; Xi’an, Shaanxi, China, Oct 26−30, 2012; pp 78−87,
DOI: 10.1063/1.4816855.
(26) de Castro, J. A.; da Silva, A. J.; Sasaki, Y.; Yagi, J. ISIJ Int. 2011,
51 (5), 748−758.

J DOI: 10.1021/acs.energyfuels.7b03150
Energy Fuels XXXX, XXX, XXX−XXX

Вам также может понравиться