Você está na página 1de 12

Minerals Engineering 63 (2014) 4556

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Numerical study of hot charge operation in ironmaking blast furnace


S.B. Kuang a,, Z.Y. Li a, D.L. Yan b, Y.H. Qi b, A.B. Yu a
a
b

Laboratory for Simulation and Modelling of Particulate Systems, School of Materials Science and Engineering, The University of New South Wales, Sydney, NSW 2052, Australia
State Key Laboratory for Advanced Iron and Steel Processes and Products, Central Iron and Steel Research Institute, Beijing 100081, China

a r t i c l e

i n f o

Article history:
Available online 28 November 2013
Keywords:
Extractive metallurgy
Mathematical modelling
Blast furnace
Multiphase ow

a b s t r a c t
Charge of hot coke and iron-bearing materials into an ironmaking blast furnace (BF) may bring signicant
energy and environmental benets to the BF process. However, there is little information about the quantitative effects of hot charge operation on BF ow and performance. This paper presents a numerical
study of multiphase ow, heat and mass transfer in a BF by a process model. The applicability of the
model in predicting BF performance is rst conrmed by different applications. It is then used to study
the effects of hot charge operation at different temperatures. The results are analyzed in detail with
respect to BF ow and performance. It is shown that compared to the conventional operation, hot charge
operation can lead to an increased productivity, decreased coke rate and CO2 emission, and at the same
time, increased gas pressure and top gas temperature. These effects vary with hot charge temperature.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Blast furnace (BF) ironmaking is the most important technology
by which iron is rapidly and efciently reduced from iron-bearing
materials (Biswas, 1981). Its primary energy source and reducing
agent are mainly coal in form of coke and pulverized coal, which
is nally released as CO2 to the environment. Also, BF ironmaking
system consumes 70% of the energy input in an integrated steelmaking works. Therefore, BF, as the core of the system, is usually
featured with extensive energy consumption and massive greenhouse gas emission. Furthermore, such a reactor demands a significant amount of coke to maintain adequate furnace permeability
and provide thermal and chemical energy sources. The consumed
coke, as a kind of noble material, shares a large portion (25%) of
the production cost of hot metal (HM). As such, coke rate (coke
consumption per tonne of hot metal, also referred to as CR for convenience) is critical to BF performance with regard to energy efciency, CO2 emission and production cost.
In recent years, various technologies have been developed to
improve BF performance. These include, for example, top gas recycling (Austin et al., 1998; Nogami et al., 2006; Chu and Yagi, 2010;
Helle et al., 2010), injection of pulverized coal, hydrogen bearing
materials, natural gas, and coke oven gas (Slaby et al., 2006; Li
et al., 2007; Shen et al., 2009), charge of novel burden materials
such as scrap and carbon composite agglomerate (Nogami et al.,
2006; Kawanari et al., 2011), and hot charge (Biswas, 1981).
Although being examined at various levels, many of these technol Corresponding author. Tel.: +61 2 9385443; fax: +61 2 93856565.
E-mail address: s.kuang@unsw.edu.au (S.B. Kuang).
0892-6875/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.mineng.2013.11.002

ogies are still on trial, with the long-term practical feasibility largely remaining unknown. This is especially true for hot charge
operation, where coke and iron-bearing materials, usually referred
to as burden, are alternatively charged into a BF at a higher temperature than the ambient temperature as used in a conventional
operation. This high temperature may be achieved through the following two ways. One directly charges the hot stock materials
from the upstream of the BF, which avoids the massive energy loss
related to cooling process. Another makes use of the unutilized
sensible heat and chemical energy of materials within the integrated steelmaking works to preheat the burden materials to a certain temperature before charging. With the help of the extra heat
input from the furnace top, hot charge operation may have great
potential in improving BF performance. However, to date, our
knowledge about the effect of such a technology on BF ow and
performance is little, especially at a quantitative level. This problem is further complicated by the fact that conveying and charging
systems required by hot charge operation to withstand the high
temperature environment at the furnace top have not been fully
established yet.
On the other hand, in order to secure a successful running of
new operations, it is a necessary prerequisite to predict and understand BF ow and performance over a wide range of conditions.
This is difcult to achieve experimentally, because ironmaking BF
is a very complex multiphase reactor accompanying with high
temperature and hazardous conditions. In principle, this problem
can be overcome by numerical simulation. In this direction, various
mathematical models have been developed in the past decades to
describe localized or global particulate and multiphase ow behaviors in BFs (see, for example, the recent review by Dong et al.

46

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

Nomenclature
aFeo
Ac
cp
d
Deg;CO
Ef
fo
F
g
hij
H
DH
k
kf
ki
K1
Mi
Msm
p
pct
pr
R
S
Shr


Shr

T
u
Vb
Vg
Volcell

activity of molten wustite


effective surface area of coke for reaction, m2
specic heat, J kg1 K1
diameter of solid particle, m
effective diffusivity of carbon monoxide, m2 s1
effectiveness factors of solution loss reaction
fraction conversion of iron ore
interaction force per unit volume, kg m2 s2
gravitational acceleration, m s2
heat transfer coefcient between i and j phases,
W m2 K1
enthalpy, J kg1
reaction heat, J kg1
thermal conductivity, W m1 K1
gas-lm mass transfer coefcient, m s1
reaction constant of ith chemical reaction, m s1
equilibrium constant of indirect reduction of iron ore by
CO
molar mass of ith species in gas phase, kg mol1
molar mass of FeO or ux in solid phase, kg mol1
pressure, Pa
percentage
prandtl number, cp lK1
reaction rate, mol m3 s1
source term
shrinkage ratio dened as the ratio of the decreased volume, caused by softening and melting, to the original
volume occupied by iron-bearing material

normalized shrinkage ratio, Shr Shr =Shr;max ,
Shr,max = 0.7
temperature, K
interstitial velocity, m s1
bed volume, m3
gas volume, m3
volume of control volume, m3

(2007)). Generally speaking, the existing approaches can be classied into two categories: continuum approach at a macroscopic level and discrete approach at a microscopic level. The former is
suitable for process modelling and applied research because of
its computational convenience and efciency. Indeed, most of the
BF modelling is based on continuum approach (Dong et al.,
2007). However, to date, comprehensive numerical studies of hot
charge operation in BF have not been found in the literature.
This paper presents a numerical study of BF ow and performance at different hot charge temperatures by a continuum-based
process model. It is organized as follows. First, the numerical model is introduced. The applicability of the model is examined by different applications. On this base, the effects of hot charge on
process performance are quantied, followed by a detailed study
of the ow and heat and mass transfer for better understanding.
The ndings from this study should be useful not only for establishing a full picture about the hot charge operation but also for
developing some guides for possible implementation of this technology in practice.

2. Model description
The current BF process model is a two-dimensional (2-D) mathematical model which considers mass, momentum and enthalpy
conservations for gas and solid phases at steady state. It is in principle the same as that recently reported by Dong et al. (2010). For

yi
yi

mole fraction of ith species in gas phase


mole fraction of ith species in equilibrium state

Greek letters
C
diffusion coefcient
I
identity tensor
/
general variable
U
shape factor
a
specic surface area, m2 m3
b
mass increase coefcient of uid phase associated with
reactions, kg mol1
d
distribution coefcient
e
volume fraction
g
fractional acquisition of reaction heat
l
viscosity, kg m1 s1
q
density, kg m3
s
stress tensor, Pa
x
mass fraction
nore, ncoke local ore, coke volume fraction
Subscripts
e
effective
g
gas
i
identier (g or s)
i, m
mth species in i phase
j
identier (g or s)
k
kth reaction
s
solid
sm
FeO or ux in solid phase
Superscripts
e
effective
g
gas
s
solid

brevity, we only describe the key features of this model below,


with the new developments emphasized.
2.1. Governing equations
Table 1 summarizes the governing equations for uid ow as
well as heat and mass transfer considered in this study. Gas is described by the well-established volume-averaged, multiphase, NavierStokes equations (Dong et al., 2007). Solids are assumed to be
a continuous phase that can be modelled based on the typical viscous model used in multiphase ow modelling (Austin et al., 1997),
coupled with the method proposed by Zhang et al. (1998) for
determination of the deadman boundary. General convectiondiffusion equations are applied to describe heat and mass transfer
among the phases.
2.2. Momentum, heat and mass transfer between phases
The gassolid interaction as gas ows through a packed bed is
described by the Erguns expression (1953):

F sg F gs af qg jusg j bf usg :

The interphase mass transfer, which occurs due to reactions and


phase changes, is evaluated from simple mass balances. Accordingly, three crucial chemical reactions and two important phase

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556


Table 1
Governing equations.
Equations

Description

Mass conservation

r  ei qi ui Si ; where Si  k bi;k Rk


r  eg qg ug ug r  sg  eg rp qg eg g F sg

Momentum
conservation

sg eg lg rug rug T   23 eg lg r  ug I

Gas

r  (esqsusus) = r  ss - esrps + qsesg


ss es ls rus rus T   23 es ls r  us I
Solid

r  ei qi ui ui;m  r  ei Ci rui;m Sui;m


If /i,m is Hi,m, Ci ckp;ii
Heat and species
conservation

Sui;m di hij aT i  T j gi


k Rk DHk

If /i,m is xi,m, Ci = qiDi, Sui;m

Phase volume fraction


State equation


k i;m;k Rk

where
u xg;CO ; xg;CO2 ; xs;Fe2 O3 ; xs;Fe3 O4 ; xs;FeO ; xs;flux
Pi;m
iei = 1
P
p=
i(yiMi)RTg/Vg

Table 2
Chemical reactions.
Reaction formula

Reaction rate

Fe2O3(s) + CO(g) = Fe(s) + CO2(g)

R1

FeO(l) + C(s) = Fe(l) + CO(g)

R2
Ac
Vb

C(s) + CO2(g) = 2CO(g)


FeO(s) ? FeO(l)
Flux(s) ? Slag(l)

Refs.

12nore eore PyCO yCO =8:314T s

1
1
3 1dore fk1 11=K 1 g

dore =Deg;CO 1f0

k2 VAcb

aFeO ,

0:468eucoke dcoke 

R3

6ncoke ecoke pyco =8:314T s


2
dcoke =kf 6=qcoke Ef k3
T T

1
i
min;sm
R4 hT max;sm
T min;sm i0

xsm ui qi ei dA
M sm Volcell

Muchi
(1967)
Omori
(1987)
Omori
(1987)
Austin
et al.
(1997)

changes are considered, including indirect reduction of iron, direct


reduction of wustite, solution loss reaction, and melting of Fe and
FeO, as listed in Table 2. Because the hydrogen proportion in the
current simulation system is small, hydrogen reduction and gas
water reactions are not considered.
The gassolid convective heat transfer coefcient hgs is calculated using the RanzMarshall equation:

hgs c

kg
0:333
2:0 0:6Re0:5
:
g Pr
ds

As the permeability of the softening phase is lost within the


cohesive zone, heat transfer between gas and particles here
changes to between gas and slab. Thus, the corresponding heat
transfer coefcient is expressed as follows (Maldonado, 2003):

hgslab

kg
0:4
0:203Re0:33
Pr0:33 0:22Re0:8
g
g Pr :
ds

Heat loss through BF wall is characterised by Newtons law of


cooling, in which the temperature gradient normal to furnace wall
is employed to quantify the heat transfer amount. Considering the
current BF refractory materials and wall thickness, the heat conductivity coefcient of BF wall is set to 5 W m1 K1.
2.3. Modelling of cohesive zone (CZ)
The softening and melting zone within a BF, i.e., the so-called
CZ, contributes signicantly to the process complexity. This zone
is of critical importance for efcient operation of the BF because
its shape and position determine the permeability, uid ow, gas
usage, thermal and chemical efciency, and hot-metal quality in
the furnace. Within this region, iron-bearing materials are
gradually transformed from the lumpy, to the softening, to the

47

half-molten states before nally melting down. Instead of passing


through the low permeability portion of this region, the reducing
gas can ow radially through adjacent coke layers, which form a
low resistance path between dripping and lumpy zones. This nding implies a possible retardation of heating and reduction rates for
ore in the CZ, with direct gas contact only occurring at the ore-coke
interface. Therefore, careful consideration must be given to the
internal structure of the CZ.
Different numerical CZ treatments have been proposed, such as
isotropic and anisotropic nonlayered treatments and layered treatment. In this study, the layered treatment recently proposed by
Dong et al. (2010) is adopted. In such a model, the CZ is treated
as layered structure and subdivided into different states according
to the BF dissection studies reported in the literature. To achieve
this, the stratied structure of coke and ore layers is rst calculated
based on the solid ow eld. Then, the CZ is dened to start and
nish within the temperature range of 14731673 K. Finally, the
CZ region is divided into alternate coke and softening and melting
ore layers.
For simplicity, the following three states are specied for the

ore layer: (1) state I, 0.7 < Shr < 1.0 corresponds to the portion with
molten state and liquid source in which the ore layer voidage is

occupied fully by the liquid phase; (2) state II, 0.5 < Shr < 0.7 corresponds to the combined portion with softening and melting of ores
in which the pressure drop may have increased signicantly; and

(3) state III, 0 < Shr < 0.5 corresponds to the softening stage in
which macropores of the bed remain open so the variation of the
pressure drop is limited. In the CZ, the solid conduction and gas
solid heat transfer coefcient are specied according to the different heat and mass transfer mechanisms in each of the states, as
discussed by Dong et al. (2010). Note that in this study, a similar
treatment is extended to the lumpy zone (the region above the
CZ) consisting of ore and coke layers.
2.4. Modelling of variation of stockline
The coke charged into a BF is consumed mainly through the following ways: chemical reactions such as direct reduction of wustite and solution loss reaction, combustion in the raceway in
front of tuyeres, and carburization, as illustrated in Fig. 1a. The
coke consumption due to the combustion is constant because of
the constant oxygen in the blown-in blast in a normal operation.
The coke consumption due to the carburization is small and can
be largely regarded as constant. But the coke consumption by
chemical reactions varies with in-furnace state. To adapt to this
variation, the solid charge rate needs to be changed by varying
the stockline in BF practice to ensure that the furnace is not overloaded. To date, modelling of varying stockline has not been discussed in the literature and is not clear. In this study, this
problem is overcome by introducing a simple numerical treatment
according to the nature of BF operation, as discussed in the
following.
Generally, in a BF process model, it is difcult to directly model
the raceway. Alternatively, the condition of reducing gas generated
in the raceway is approximated using the heat and mass balance
and treated as the input of gas phase, with the outlet for solids
(coke) introduced. The outlet boundary essentially represents the
simplied boundary of the raceway. Thus, the solid owrate at
the outlet is the rest of the charged coke after chemical reactions
and carburization, and should correspond to the tuyere coke
amount (the amount of coke combusted in the raceway per tonne
of hot metal). If not, the coke balance is not satised. In this case,
we can adjust the productivity and thus the solid charge rate at
the furnace top to control the coke input. This is to some degree
similar to the variation of stockline in practice. Here, the adjusted
amount of productivity is determined by denition according to

48

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

(a)

(b)

Fig. 1. Schematic illustrations of: (a) coke balance in a BF, and (b) solution procedure of the present BF process model.

the difference between the tuyere coke amounts obtained respectively according to the blast conditions and solid owrate. This ensures that the proposed method can quickly obtain the expected
productivity to satisfy the coke balance. In this study, the number
of adjustments is less than four for all the cases simulated.
3. Numerical solution
The well-established sequential solution procedure is employed
to calculate the coupled uid ow, heat and mass transfer through
the following steps as shown in Fig. 1b:
(1) The input conditions in simulation are determined according
to the global heat and mass balance under given material,
operational and geometrical conditions.
(2) The initial gas and solid ow, temperature, as well as concentration elds are determined.
(3) The layered structure is determined based on the burden
distribution at the furnace top (e.g., the ore and coke batch
weights as well as the radial distribution of orecoke ratio)
and the timelines of solid ow.
(4) The ow, temperature, and concentration elds are calculated without considering the chemical reactions until an
approximate convergence (i.e., high tolerance) in terms of
ow, mass and energy residuals is obtained.
(5) Thermochemical behaviors are taken into account, which
leads to the CZ determination, and with this information,
uid ow, heat and mass transfer, as well as chemical reactions are recalculated until the CZ position is converged or
does not change much.

(6) The coke balance is examined to adjust the productivity. If


the difference between the tuyere coke amounts calculated
respectively according to the blast conditions and solid owrate is less than 0.1 kg/tHM, the simulation stops. If not, go
to Step 1 and repeat the whole procedure.
In the above proposed procedure, because the uid ow and
heat and mass transfer are solved in sequence, their numerical
solutions are in principle similar to those used for a single-phase
ow, whose convergence conditions have been well discussed
and documented in the literature (Ferziger and Peric, 2002). However, a general convergence criterion as used in CFD cannot ensure
that the present model generates physically meaningful results. In
fact, our tests indicated that the convergence condition of the model is mainly determined by the position of CZ where both ow and
thermochemical behaviors are signicantly different from those in
other regions of BF. Therefore, in this study, the criterion for the
convergence is set to 10 pct of the relative CZ positions in two consecutive iterations. The error is reasonable considering that the
problem is complicated and the variation of the process performance predictions within this error is small.
4. Simulation conditions
Fig. 2 shows the computational domain and an enlarged area
presenting the representative computational cells. A BF geometry
with the hearth diameter of 10.6 m and inner volume of 2327 m3
is used in the simulation. Assuming the symmetrical distribution
of process variables, only half the BF is considered in the simulation and is treated as a two-dimensional slot model for simplicity.

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

The whole computational domain is divided into 519  118 nonuniform control volumes in the Cartesian coordinates. Each computational cell is around 5 cm  6 cm, which should ensure that
the mesh is small enough to properly capture the complicated multiphase ow in different layers in the lumpy and CZ regions. Note
that the thickness of ore or coke layer at the furnace top is around
50 cm. Our trial numerical results indicate that this mesh size gives
mesh-independent numerical solutions.
The hot blast (air) is blown into the BF at the owrate of 4917
Nm3/min and temperature of 1050 C. Corresponding to the blast
conditions, the components, temperature and owrate of the
reducing gas generated in the raceway are determined using the
mass and heat balance, and used as the inlet conditions of gas
phase, as listed Table 3. Note that there are no injectants (e.g. pulverised coal) through the tuyeres, and the water content in the
blast is negligible under the condition considered. Thus, the reducing gas mainly consists of CO and N2.
Solids (ore, coke and ux) are charged from the furnace top with
the ore batch weight of 30 tonnes, coke rate of 500 kg/tHM. Their
components listed in Table 3 are needed to determine solid heat
capacity and thermal conductivity, as suggested by Austin et al.
(1997). The typical radial distributions of orecoke ratio and sizes
of ore and coke as used by Dong et al. (2010) are adopted in this
study. This information, together with the productivity, determines
the solid inlet velocity that is assumed constant along the horizontal direction. Note that the productivity is a simulation output in
the present model, which to some degree varies around a target value depending on the in-furnace state under a given condition, as
discussed in Section 2.4. The solid temperature at the furnace top
varies from 300 to 1300 K to cover a wide range of hot charge temperature (HCT). Note that HCT is the only independent variable
considered in the present study.
With the input information of solids, the physical properties of
the mixture of ore and coke, such as particle diameter and porosity
can be obtained in the computational domain for modelling the
ow and heat and mass transfer of solids. In the present study,
the ore-to-coke ratio is low at the center, and the coke particle size

(a)

49

is small near the furnace wall. This burden distribution provides a


high permeability at the furnace center. Initially, the particle properties imposed at the furnace top are extended to the lower part so
that particle size and porosity distributions in the entire furnace
could be obtained. During the calculation, with the identication
of the lumpy zone, CZ, dripping zone, and deadman, the properties
in these regions are re-estimated to replace the initial values based
on the following rules: (1) in the lumpy zone, porosity in the ore or
coke layer is a function of particle size (see Table 3); (2) in the CZ,
porosity and particle size of iron-bearing materials are functions of
normalized shrinkage ratio as used by Dong et al. (2010); (3) in the
CZ and dripping regions, porosity in the areas occupied by coke is
calculated in the same way as used for coke layers in the lumpy
zone; and (4) in the deadman, coke size and porosity are assumed
as ds = 0.02 m and e = 0.65 according to a normal BF practice.
5. Results and discussion
5.1. Model applicability
It is necessary to examine the applicability of the proposed
mathematical model before its application for numerical experiments. This is here done using three cases, focused mainly on
process performance because it is impossible to measure the
in-furnace ow and heat and mass transfer at this stage of development. Three major parameters used to describe process performance are considered: utilization efciency and temperature of
top gas as well as productivity. The top gas utilization efciency
is dened as the volume percentage of CO2 in the mixture of CO
and CO2 in the gas owing out of the furnace top. It reects the
utilization efciency of reducing gas CO inside a BF. The top gas
temperature is usually used to assess the thermal energy utilization efciency of the furnace. The productivity is the amount of
hot metal produced per furnace volume per day.
In the rst case, the applicability of the proposed model in predicting process performance at different coke rates is examined.
Note that adjustment of coke rate is a basic BF operation frequently

(b)

Fig. 2. Computational domain (a), and its representative grid arrangement (b).

50

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

Table 3
Inlet conditions in the base case simulated.
Variables

Values

Gas
Inlet velocity, m/s
Inlet gas component, mole percentage
Inlet gas temperature, K
Top pressure, atm

177
34.959 pct CO; 0.0 pct CO2; 0.813 pct H2; 0.0Pct H2O; 64.228 pct N2
2313.6
2.0

Solid
Ore, t/tHM
Ore components, mass fraction
Average ore particle size, m
Coke, kg/tHM
Coke components, mass fraction
Average coke particle size, m
Flux, t/tHM
Flux components, mass fraction
Ore voidage
Coke voidage
Average ore/(ore + coke) volume ratio
Burden temperature, K

1.64
Fe2O3 0.6566; FeO 0.1576; CaO 0.0652; MgO 0.0243; SiO2 0.06; Al2O3 0.0295; MnO 0.0061; P2O5 0.008
0.03
500
C 0.857; Ash 0.128; S 0.005; H 0.005; N 0.005
0.045
0.0264
CaO 0.438; MgO 0.079; SiO2 0.024; Al2O3 0.033; CO2 in CaO 0.344; CO2 in MgO 0.082
0.403(100dore)0.14
0.153logdcoke + 0.742
0.5923
300

(a)

(b)
700

40

Top gas temperature (K)

Top gas utilization efficiency (%)

45

35
30
25
20
15

600
500
400
300
200

10
100

5
0

0
420

500

420

580

CR (kg/tHM)

500

580

CR (kg/tHM)

(c)

(d)

Unit: K

Productivity (tHM/day/m3)

2.5

1.5

0.5

0
420

500

CR (kg/tHM)

580
CR=420 kg/tHM

500 kg/tHM

580 kg/tHM

Fig. 3. Effect of coke rate on BF performance: (a) top gas utilization efciency, (b) top gas temperature, (c) productivity, and (d) solid temperature.

51

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

(a)

(b)
520

40

Top gas temperature (K)

Top gas utilization efficiency (%)

45

35
30
25
20
15

500

480

460

440

10
420
5
0

400
3663

4360

3663

4884

Blast rate (m3/min)

4360

4884

Blast rate (m3/min)

(c)

(d)

Unit: K

Productivity (tHM/m3/day)

2.5

1.5

0.5

0
3663

4360

4884

Blast rate (m3/min)

Blast rate=3663 m3/min

4360 m3/min

4884 m3/min

Fig. 4. Effect of blast rate on BF performance: (a) top gas utilization efciency, (b) top gas temperature, (c) productivity, and (d) solid temperature.

practiced in ironmaking plants, and the effect of coke rate on process performance is well recognized. Fig. 3 shows the calculated
performance parameters at different coke rates. Here, the simulation conditions are based on the base case with consideration of
coke rate variation. As seen from Fig. 3(a)(c), when coke rate is increased, both productivity and top gas utilization efciency decrease, however, the top gas temperature increases. This is
because a larger coke rate leads to an increased amount of coke
combusted in the raceway, generating more heat and reducing
gas CO. Fig. 3 also includes the solid temperature distribution at
different coke rates to show two un-expected in-furnace states
widely encountered in practice. When coke rate is too high (e.g.
CR = 580 kg/tHM in Fig. 3d), the BF is over heated and the CZ position is very high. Thus, the BF becomes too hot. Conversely, when
coke rate is too low (e.g. CR = 420 kg/tHM in Fig. 3d), the thermal
energy is not enough to maintain a normal operation and the CZ
position is very low and the BF may cool down. Overall, the predicted results shown in Fig. 3 are consistent with the practical
application of BF (Biswas, 1981).
The second case considers process performance at different
blast rates. It is known that given a certain coke rate, the

productivity of a BF is determined by the amount of oxygen blown


through the tuyeres. The more oxygen is blown into the furnace,
the more coke is consumed and forms CO, and the higher productivity becomes (Geerdes et al., 2009). This relationship is used to
examine the validity of the proposed model. For such a purpose,
process performance is predicted at different blast rates based on
the condition in the base case, and the results are given in Fig. 4.
As seen from this gure, the productivity increases with the blast
rate (Fig. 4c). This is because a larger blast rate leads to an increased amount of oxygen blown into the furnace. Furthermore,
it is observed that other performance parameters such as temperature and utilization efciency of top gas (Fig. 4a and b) and the
thermal state inside the BF as reected by solid temperature distribution (Fig. 4d) are affected slightly by blast rate. This is in line
with the basic requirement in an intensied smelting operation
for high productivity. Again, the predicted results agree with the
practical application of BF (Biswas, 1981; Geerdes et al., 2009).
In the above two cases, only qualitative comparison is made because it is difcult to nd some published experimental or plant
data which include the details adequate enough for model validation. Recently, the present model has been used to directly

52

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

Fig. 5. Calculated distributions of porosity for the BF operated at: (a) HCT = 300 K, CR = 500 kg/tHM; (b) HCT = 1300 K, CR = 500 kg/tHM; and (c) HCT = 1300 K, CR = 460 kg/
tHM.

simulate a few BFs in operation. The measured and calculated results in terms of productivity, top gas temperature, and top gas utilization efciency are compared quantitatively. The prediction
errors are less than 10%, as reected elsewhere (Kuang et al.,
2012). Nonetheless, the above results shown in Figs. 3 and 4 suggest the proposed model can be used to describe the key behaviors
of BF, at least qualitatively. In the following, the effects of hot
charge operation are investigated based on the numerical results
generated by the model.
5.2. Effect of hot charge operation on BF performance
Fig. 5 shows the representative porosity distributions within
the BF charged with ambient-temperature (Fig. 5a) and hot
(Fig. 5b) burdens, respectively, corresponding to conventional
and hot charge operations. According to such distributions, four regions can be identied including lumpy zone, CZ, dripping zone
and stagnant zone (or deadman). The results reveal that favourable
inverse-V shaped CZs are expectedly obtained in both cases; however, the CZ position rises signicantly due to the hot charge operation. This suggests that excessive energy is available in the hot
charge operation. Note that given the same CZ, a BF operated under
different conditions could lead to similar ow and thermal conditions at the lower part of the furnace, producing hot metal with
similar qualities. Based on such an understanding, some amount
of coke is reduced to achieve the same CZ prole and position as
the conventional operation (or original operation without hot
charge), and the result is given in Fig. 5c. Here, the conventional
operation representing a normal operation in BF practice is treated
as the base case. As seen from this Fig. 5c, a nearly same CZ prole
and position can indeed be obtained for the operations with and
without hot charge. Notably, to achieve this, the coke rate for the
hot charge operation needs to be reduced by 40 kg/tHM. This represents a signicant amount of coke reduction, leading to much
less CO2 emission and production costs. Evidently, this large
amount of coke reduction is due to the high hot charge temperature (=1300 K) used in the considered case.

Fig. 6 shows the inuence of hot charge operation on process


performance in terms of coke rate, productivity, top gas utilization
efciency, and top gas temperature at different burden temperatures. Here, hot charge operations are considered, respectively, at
the same coke rate (without coke reduction) and the same CZ (with
coke reduction) as the conventional operation. It can be seen from
Fig. 6 that when hot charge temperature is increased, the coke rate
decreases and slows down at the same CZ (Fig. 6a). Conversely, the
top gas temperatures under two different conditions of hot charge
operation both initially increase and then slow down (Fig. 6d). The
results suggest that a higher hot charge temperature brings more
benets in view of coke reduction but more challenges in conveying/charging gas and burden materials of high temperature. Moreover, the gradual decrease of coke rate indicates that the benets
are less signicant when charge temperature is too high. In fact,
a too high hot charge temperature is also unfavourable considering
that it may lead to decreased permeability and deteriorate BF performance as a result of powder generation due to degradation of
raw materials and/or softening of ore and ux. Similar problems
may occur to other BF operations. Modelling of such an adverse effect in a BF process model should be useful but challenging, and deserves comprehensive studies in the future. At present, it is
neglected in this study. On the other hand, it is of interest to note
that the productivity (Fig. 6b) and top gas utilization efciency
(Fig. 6c) show different trends under two different conditions. Given the same coke rate, with increasing hot charge temperature,
the productivity initially remains the same and then increases
slightly, and the top gas utilization efciency decreases, particularly at relatively high hot charge temperatures. Conversely, given
the same CZ, both productivity and top gas utilization efciency
rapidly increase and then slow down.
5.3. Effect of hot charge operation on in-furnace state
Process performance is essentially governed by the in-furnace
state such as ow and heat and mass transfer. Analysis of the infurnace state can generate some insights into BF ow and thermal

53

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

500

At the same CZ
At the same coke rate

44

(c)

At the same CZ
At the same coke rate

490

Coke rate (kg/tHM)

45

(a)
Top gas utilization efficiency (%)

510

480
470
460
450
440

43
42
41
40
39
38
37
36
35

430
300 400 500 600 700 800 900 1000 1100 1150 1200 1250 1300

300 400 500 600 700 800 900 1000 1100 1150 1200 1250 1300

HCT (K)

HCT (K)

2.6

1200

(b)

At the same CZ

(d)

At the same CZ
At the same coke rate

At the same coke rate

1000

Top gas temperature (K)

Productivity (tHM/m3/day)

2.5

2.4

2.3

2.2

2.1

800

600

400

200

2
300 400 500 600 700 800 900 1000 1100 1150 1200 1250 1300

HCT (K)

0
300 400 500 600 700 800 900 1000 1100 1150 1200 1250 1300

HCT (K)

Fig. 6. Process performance as a function of hot charge temperature: (a) coke rate, (b) productivity, (c) top gas utilization efciency, and (d) top gas temperature.

behaviors, hence providing a better understanding of process performance as well as some information guiding the design and operation of BF. Such analysis is difcult to achieve experimentally but
can be readily performed based on the numerical simulations. For
clarity, the analysis here mainly focuses on the conventional operation and the hot charge operations at HCT = 1300 K with and
without coke reduction, similar to Fig. 5.
Fig. 7 shows the gas and solid ow elds for the three cases considered. As seen from the gure, the difference among different
cases for gas ow is mainly observed in the CZ region, where the
gas ows mainly through the coke window (Fig. 7a). Compared
to the conventional operation, both hot charge operations regardless of coke reduction generally lead to larger gas velocities, which
is not desirable for preventing erosion of wall refractory materials.
On the other hand, all the solid ow elds show similar distributions. The solid velocities are observed to considerably decrease
in the dripping zone as only coke is left there. In the lumpy zone,
the solid velocities in both hot charge operations are generally larger than the conventional operation, corresponding to the productivity shown in Fig. 6b. As such, the slip velocities between gas and
solid phases are increased by a hot charge operation, particularly
when coke rate is high.

Fig. 8 shows the solid and gas temperature elds. Compared to


the conventional operation, both hot charge operations have higher solid temperatures but mainly in the lumpy zone, which also
show more uniform distributions in this region (Fig. 8a). Such high
solid temperatures in the lumpy zone should promote solution loss
reaction which is an intensively endothermic reaction, leading to
more coke consumption. In addition, it is noted that the solid temperature near the wall at the upper part of the furnace in the hot
charge operation with coke reduction is lower than the hot charge
operation without coke reduction. This may be due to that the heat
transfer from the hot burdens to the gas is intensied by the coke
reduction, leading to less energy generated inside the furnace. Corresponding to the solid temperature distributions, the gas temperatures at the upper part of the furnace in both hot charge
operations are higher than the conventional operation (Fig. 8b),
achieved by the extra heat brought into the furnace by the hot burdens. This leads to gas expansion and larger gas velocities in the
two cases as observed in Fig. 7a. Note that the gas or solid temperatures in Fig. 8 (also Figs. 3d and 4d) are largely uniform in the
deadman or stagnant zone because here the gas ow through this
region is small and the heat exchange between gas and solid
phases is also small.

54

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

(a)

HCT=300K
CR=500 kg/tHM

1300 K
500 kg/tHM

(a)

1300 K
460 kg/tHM

(b)
Fig. 7. Gas (a) and solid (b) ow elds, corresponding to Fig. 5.

Fig. 9 shows the gas pressure distributions and reveals that hot
charge operations regardless of coke reduction generally result in
larger gas pressure in the entire furnace compared to the conventional operation. This should be attributed to the increased slip
velocity between gas and solid phases as a result of gas expansion
and increased productivity. Note that a large pressure drop may
cause unfavourable phenomena for smooth running of the BF
including uidization, hanging and channelling. Interestingly, it is
observed that coke reduction, coupled with hot charge operation,
can to some degree decrease gas pressure, although coke adjustment itself may increase gas pressure signicantly, if not controlled reasonably. It should be pointed out that the ndings

HCT=300 K
CR=500 kg/tHM

1300 K
500 kg/tHM

1300 K
460 kg/tHM

(b)
Fig. 8. Distributions of: (a) solid temperature, and (b) gas temperature, corresponding to Fig 5.

from the analysis of in-furnace state can also apply to other hot
charge temperatures, whose results are not included in this paper
for brevity.
In order to better link the in-furnace state to the BF performance, analysis of coke consumption in the furnace is performed
and the results are given in Fig. 10. As seen from this gure, the
coke rate is nearly equal to the sum of the coke consumptions by
the combustion in the raceway (=the amount of tuyere coke) and
reactions. Overall, the tuyere coke is the dominating contributor
to the coke rate. With increasing hot charge temperature, the

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

HCT=300 K
CR=500 kg/tHM

1300 K
460 kg/tHM

1300 K
500 kg/tHM

Fig. 9. Distributions of gas pressure, corresponding to Fig. 5.

560

Filled: at the same CZ


Hollow: at the same coke rate

Amount of coke (kg/tHM)

520

440
400
360
320
140
120
400

600

800

1000

Fig. 10 can also be used to explain the top gas utilization efciency shown in Fig. 6c. At the same coke rate, the variation of
tuyere coke amount is small, the decreased top gas utilization efciency at a higher hot charge temperature is mainly attributed to
the increased coke consumption by reactions. However, with the
same CZ, the tuyere coke amount and coke consumption by reactions show opposite trends at different hot charge temperatures.
Their effects on top gas utilization efciency are cancelled out by
each other. This is particularly evident at a high temperature
where the decrease of tuyere coke slows down, whereas the increase of coke consumption by reactions largely remains linear.
Note that the tuyere coke amount contributes to a larger portion
of the coke rate and plays a dominating role. Consequently, the decreased tuyere coke amount at a higher hot charge temperature
essentially accounts for the decrease of coke rate and thus the increase of top gas utilization at the same CZ. However, it is the
opposite effect of the two coke consumptions that accounts for
the gradual variation of top gas utilization efciency at a high
hot charge temperature. In addition, it should be pointed out that
a decreased tuyere coke amount leads to less heat brought from
the lower part to the upper part of the furnace. This explains
why given the same burden temperature, the solid temperatures
and CZ position generally decrease in the hot charge operation at
a smaller coke rate, as shown in Figs. 5 and 8.

6. Conclusions
The multiphase ow, heat and mass transfer in a BF have been
studied by a BF process model, with special reference to hot charge
operation. The model can satisfactorily predict BF performance
such as productivity, top gas utilization efciency and top gas temperature under different conditions, although further developments may be needed, particularly for more complicated BF
operations, e.g. combustion of injected pulverized coal in the raceway. The ndings from this study can be summarized as follows:

480

100
200

55

1200

1400

HCT (K)
Fig. 10. Different coke consumptions as a function of hot charge temperature (A:
Coke charged, R: Coke consumed by reactions, and D: Coke combusted in the
raceway).

tuyere coke amount decreases at the same CZ, however, it decreases slightly at the same coke rate. Meanwhile, the coke consumptions by reactions under two different conditions of hot
charge operation both increase. It is known that the productivity
of a BF increases with the decrease of tuyere coke amount in a stable situation (Geerdes et al., 2009). This is because every charge of
coke at the top of a furnace brings with it an amount of iron-bearing materials, and the hot metal is produced as soon as the coke is
consumed under the stable condition. On the other hand, the top
gas utilization efciency may decrease with the increase of tuyere
coke amount or coke consumption by reactions because more
reducing gas CO is generated from coke. Based on this understanding, the calculated ow and thermal behaviors as well as process
performance at different hot charge temperatures can be further
explained using the results shown in Fig. 10. At the same CZ or
coke rate, the increased productivity at a higher hot charge temperature (Fig. 6b) is essentially attributed to the decreased tuyere
coke amount. The variation of tuyere coke amount at the same
CZ is more considerable than that at the same coke rate, leading
to more signicant variation of the productivity.

(1) Compared to the conventional operation, hot charge operation can lead to an increased productivity, and decreased
coke rate and CO2 emission, and at the same time, increased
gas pressure and top gas temperature. In general, with
increasing hot charge temperature, these effects are more
evident. However, the gains do not vary much at high hot
charge temperatures.
(2) Hot charge operation may lead to a higher solid temperature
and a larger gas pressure than the conventional operation,
particularly at the upper part of the furnace. The effects
are unfavorable in view of coke consumption and smooth
operation. This problem can be overcome by decreasing coke
rate to produce a cohesive zone of the same position. This
technique can also be used to quantify the coke reductions
with different hot charge temperatures.
(3) The coke reduction as a result of hot charge operation is
mainly due to the reduced amount of coke combusting in
the raceway, which also accounts for the increases of gas utilization efciency and productivity.
Finally, it should be pointed out that to date, hot charge operation has not been practiced due to its high requirements in equipment as a result of high temperature environment at the furnace
top. If this problem can be overcome, and the properties of burden
materials do not change much, hot charge operation should be
benecial to BF performance, as demonstrated in this study. Therefore, related aspects need to be investigated carefully. At this stage,
caution should be taken in developing and implementing this technology in practice.

56

S.B. Kuang et al. / Minerals Engineering 63 (2014) 4556

Acknowledgements
The authors are grateful to the Australia Research Council (ARC)
and Central Iron & Steel Research Institute (CISRI) for the nancial
support of this work, and the National Computational Infrastructure (NCI) for the use of its high performance computational
facilities.
References
Austin, P.R., Nogami, H., Yagi, J.I., 1997. A mathematical model for blast furnace
reaction analysis based on the four uid model. ISIJ International 37 (8), 748
755.
Austin, P.R., Nogami, H., Yagi, J.I., 1998. Prediction of blast furnace performance with
top gas recycling. ISIJ International 38 (3), 239245.
Biswas, A.K., 1981. Principles of Blast Furnace Ironmaking: Theory and Practice.
Cootha Publishing House.
Chu, M.S., Yagi, J.I., 2010. Numerical evaluation of blast furnace performance under
top gas recycling and lower temperature operation. Steel Research International
81 (12), 10431050.
Dong, X.F., Yu, A.B., Yagi, J.I., Zulli, P., 2007. Modelling of multiphase ow in a blast
furnace: recent developments and future work. ISIJ International 47 (11), 1553
1570.
Dong, X.F., Yu, A.B., Chew, S.J., Zulli, P., 2010. Modeling of blast furnace with layered
cohesive zone. Metallurgical and Materials Transactions B Process Metallurgy
and Materials Processing Science 41 (2), 330349.
Ergun, S., 1953. Pressure drop in blast furnace and in cupola. Industrial and
Engineering Chemistry 45 (2), 477485.
Ferziger, J., Peric, M., 2002. Computational Methods for Fluid Dynamics. Springer,
New York.

Geerdes, M., Toxopeus, H., Vliet, V.D., 2009. Modern Blast Furnace Ironmaking: An
Introduction. IOS Press BV, Amsterdam, the Netherlands.
Helle, H., Helle, M., Pettersson, F., Saxen, H., 2010. Multi-objective optimization of
ironmaking in the blast furnace with top gas recycling. ISIJ International 50 (10),
13801387.
Kawanari, M., Matsumoto, A., Ashida, R., Miura, K., 2011. Enhancement of reduction
rate of iron ore by utilizing iron ore/carbon composite consisting of ne iron ore
particles and highly thermoplastic carbon material. ISIJ International 51 (8),
12271233.
Kuang, S.B., Li, Z.Y., Yu, A.B., 2012. Process modelling and analysis of ironmaking
processes for improving energy efciency in Baosteel. Internal Report to
Baosteel.
Li, J.X., Wang, P., Zhou, L.Y., Cheng, M., 2007. The reduction of wustite with high
oxygen enrichment and high injection of hydrogenous fuel. ISIJ International 47
(8), 10971101.
Maldonado, D., 2003. Heat and mass transfer in blast furnace: low temperature
cohesive layer experiments and mathematical modelling at low and high
temperates. PhD thesis, University of New South Wales, Sydney.
Muchi, I., 1967. Mathematical model of blast furnace. Transactions ISIJ 7 (1), 223
237.
Nogami, H., Yagi, J.I., Kitamura, S., Austin, P.R., 2006. Analysis on material and
energy balances of ironmaking systems on blast furnace operations with
metallic charging, top gas recycling and natural gas injection. ISIJ International
46 (12), 17591766.
Omori, Y., 1987. Blast furnace phenomena and modelling. Elsevier Applied Science,
London, UK.
Shen, Y.S., Guo, B.Y., Yu, A.B., Zulli, P., 2009. A three-dimensional numerical study of
the combustion of coal blends in blast furnace. Fuel 88 (2), 255263.
Slaby, S., Andahazy, D., Winter, F., Feilmayr, C., Burgler, T., 2006. Reducing ability of
CO and H2 of gases formed in the lower part of the blast furnace by gas and oil
injection. ISIJ International 46 (7), 10061013.
Zhang, S.J., Yu, A.B., Zulli, P., Wright, B., Tuzun, U., 1998. Modelling of the solids ow
in a blast furnace. ISIJ International 38 (12), 13111319.

Você também pode gostar