You are on page 1of 22

Modelling the Complex Interactions Between

Reformer and Reduction Furnace in a


Midrex-Based Iron Plant
Khalid Alhumaizi, AbdelHamid Ajbar* and Mustafa Soliman
Department of Chemical Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
This article studies the complex mass and energy interactions between the reformer and the reduction furnace in an iron plant based on Midrex
technology. The methodology consists in the development of rigorous rst principle models for the reformer and the reduction furnace, in addition
to models for auxiliary units such as heat recuperator, scrubber and compressor. In this regard, a one-dimensional heterogeneous model for the
catalyst tubes which takes into account the intraparticle mass transfer resistance was developed for the reformer unit, while the furnace was
modelled with bottom-ring conguration. As for the reduction furnace, the mathematical model was based on the concept of shrinking core
model. The furnace was modelled as a moving bed reactor taking into consideration the effects of water gas shift reaction, steam reforming of
methane and carburisation reactions. The model was rst validated using data from a local iron/steel plant and was then simulated to determine
key output variables such as bustle gas temperature, degree of metalisation, carbon content, ratio of hydrogen to carbon monoxide, reductants
to oxidants ratio and required compression energy. The effects of key input parameters on the performance of the plant were studied. These
parameters included recycle ratio, scrubber exit temperature, injected oxygen ow rate, ow rate of natural gas after reformer, to transition
zone, to reformer and to cooling zone. Useful proles were compiled to illustrate the results of the sensitivity analysis. These results may serve as
guidelines for a further optimisation of the plant.
Keywords: direct reduced iron, Midrex plant, reformer, shrinking core model, modelling, simulations
INTRODUCTION
T
he direct reduction process has gained growing impor-
tance in the last decades as a source of metallic units
for electric arc furnaces used for the production of steel.
The Midrex technology is the most important gas-based direct
reduced iron (DRI) process and consists of around 58% of the
world DRI production (Midrex, 2011). Moreover, in today environ-
mentally focused industry, the gas-based direct reduction process
offers a clear advantage over the conventional BF(blast fur-
nace)/BOF(basis oxygen furnace) technology that has been the
backbone of iron/steel production. The Midrex process produces
only one-third of the carbon dioxide generated by a traditional BF
process (Midrex, 2011). The issue of greenhouse gas emissions is
of great importance given that the iron/steel industry is known
to be a large polluter that emits around 15% of carbon diox-
ide emissions within the industrial sector (Fujita et al., 2010).
The study on ways to make the iron/steel industry more prof-
itable by optimising the energy use is a subject that has received
increasing attention in recent years, since the iron/steel indus-
try is known to be one of the most energy-intensive industries
(Larsson and Dahl, 2003). The analysis of energy use in this
industry is, however, a complex task because there are a num-
ber of material and energy ows that interact in sometimes
unpredictable ways. Accordingly, different approaches have been
followed in the literature. Such methodologies consist of thermo-
dynamic analysis, economic models, methods based on process
integration (PI) and model-based optimisation. Thermodynamic-
based concept of exergy has been used to analyse the energy
efciency in many industrial sectors and in particular in the iron
and steel industry (Fraser et al., 2006). However, exergy account-
ing may not provide an answer to the analysis of all aspects of
material and energy ows since, being a tool for analysis, it can
not provide a framework for design or optimisation.

Author to whom correspondence may be addressed.


E-mail address: aajbar@ksu.edu.sa
Can. J. Chem. Eng. 9999:122, 2011
Copyright

2011 Canadian Society for Chemical Engineering


DOI 10.1002/cjce.20596
Published online in Wiley Online Library
(wileyonlinelibrary.com).
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 1 |
PI tools, on the other hand, have been used successfully in the
literature for the analysis of interactions between mass and energy
aspects. PI refers to the application of systematic methodologies
that facilitate the selection and/or modication of processing steps
and the analysis of interconnections within the process, with the
goal of minimising the use of resources. Schultmann et al. (2004),
for instance, presented a PI method for a basic oxygen furnace.
Mathematical programming methods are also tools that belong
to PI, and are increasingly being investigated in the iron/steel
industry (Larsson and Dahl, 2003; Larsson et al., 2006). Larsson
et al. (2006), for instance, presented a PI-based optimisation
approach to study material and energy systems in an iron/steel
plant. Due to the high temperatures and the different characteris-
tics of the different ows in the iron/steel industry, the application
of different PI methods is still an active area of research.
Model-based approach, followed in this article, is also a useful
tool for analysis. In process modelling, mathematical equations
are used to describe the mechanism of processing operations.
These equations are usually material, heat balances and design
equations. The task of developing a rigorous model for the plant
is, however, quite difcult. The plant consists essentially of the
reformer and the reduction unit. These units are quite complex
within themselves and are also connected to each other by recycle
loops making the modelling and the simulation challenging tasks,
which are exactly the objectives of the article, and represent the
main novelty of this work.
Before the models for the individual reactor units are described,
we present in the following a short overview of the modelling
approaches followed in the literature for the steam reforming
and reduction reactors. We also outline the main assumptions
made in our modelling work. The mathematical modelling of con-
ventional steam reformers used in chemical and petrochemical
industries (e.g. ammonia) has been well studied in the litera-
ture (Soliman et al., 1988; Elnashaie et al., 1992; Pedernera et
al., 2003). Various models with different degrees of complexity
were used for the design and simulation of these units. The mod-
elling of reformers used in DRI plants did not receive similar
attention in the literature. Murty and Murthy (1988), for instance,
used the ux method for radiative transfer to develop models for
reformer furnaces. Of notable contribution is the work of Farhadi
et al. (2003) who carried out a detailed one-dimensional heteroge-
neous model for the Midrex reformer. Pedernera et al. (2003), on
the other hand, studied the steady-state operation of a large-scale
primary reformer by means of a heterogeneous two-dimensional
model, which accounts for the diffusion limitations in the cat-
alyst particle. The authors uncovered strong radial temperature
gradients in the reformer tube, particularly close to the reactor
entrance. In a recent study, Shayegan et al. (2008) developed a
two-dimensional heterogeneous model for a Midrex reformer. The
authors presented arguments for the need to account for temper-
ature gradients in radial dimension for their studied unit. These
included low Reynolds number and low tube length to diameter
ratio.
In this work, a one-dimensional heterogeneous model for the
catalyst tubes which takes into account the intraparticle mass
transfer resistance is developed for the reformer. The furnace is
modelled with bottom-ring conguration. Besides its relative
simplicity, the choice of a one-dimensional model is based on
the recent work of Shayegan et al. (2008) who concluded that
the quality of predictions of the two-dimensional model depended
strongly on tted correlations for wall to uid heat transfer. The
authors also reported that one-dimensional models had better pre-
dictions for ue gas temperatures. Moreover, the lack of any radial
measurements for the studied plant limits the usefulness of any
two-dimensional model. Another worthy difference in this work
is the choice of kinetics. Previous studies (Farhadi et al., 2003;
Shayegan et al., 2008) used the simple rst order kinetic model of
Akers and Camp (1955) for which an analytical solution is avail-
able for the effectiveness factor, whereas this work uses a more
rigorous LangmuirHinshwold type of kinetics (Xu and Froment,
1989) for which a numerical solution for the effectiveness factors
is needed.
As for the reduction unit, a variety of studies on moving
gassolid bed reactors have been available in the literature since
the early 1950s. These studies used a number of approaches,
including simple linear models (Amundson, 1956), the shrink-
ing core model for solid particles with different temperatures in
solid and gas phases (Amundson and Arri, 1978), one and three-
interface pellet models (Negri et al., 1987; Rao and Pichestapong,
1994; Valipour et al., 2006) as well as the grain model with prod-
uct layer resistance (Nouri et al., 2011).
In this article, the concept of the shrinking core model (Rao and
Pichestapong, 1994; Valipour et al., 2006) is adopted, and a one-
dimensional non-isothermal moving bed model is derived for the
furnace. Unlike previous works (Parisi and Laborde, 2004; Nouri
et al., 2011), both the kinetics of steam reforming of methane
and carburisation are included in the model. Moreover, all the
reduction reactions are assumed to be reversible, which is more
reminiscent of the actual behaviour in the reactor.
Plant Characteristics
The major components of a Midrex plant are the reduction fur-
nace and the natural gas reformer with associated heat exchange
equipment (Figure 1). The furnace is fed at the top with a mixture
of lump ore and ore pellets. Reduction takes place as the charge
descends in counter-current contact with hot reducing gas. This
gas is generated by catalytic reforming of natural gas together with
a portion of the reducer offgas. The reformed gas temperature can
be raised to about 980

C by injecting oxygen in its stream. The hot


DRI is cooled to near ambient temperatures in the lower section
of the furnace by means of inert cooling gas which circulates in a
closed loop. Around two-thirds of this gas are mixed with natural
gas as feed to the reformer, and the remainder is burned as a fuel
to produce heat in the reformer.
The iron/steel plant, under study, is located in Saudi Arabia.
Unlike most plants in the world where the scarp constitutes the
major part of the feed, the plant under study uses a feed which
consists of large proportion of DRI, that is in the order of 80%,
with the rest being scarp. Slag, oxygen and carbon injection prac-
tices are different when melting DRI as compared to melting 100%
scarp. DRI and scrap have different metallic content, with scrap
inherently composed of iron. Intrinsic pieces of steel scrap usu-
ally have about 9899% metallic iron content. DRI, on the other
hand, has a 7989% metallic iron content. The rest is composed
of FeO, carbon, phosphorous, sulfur and gangue. Tables 13 sum-
marise the characteristics of the plant. These include the design
parameters of the reformer (Table 1), those of the reduction fur-
nace (Table 2), as well as the nominal operating parameters of the
plant (Table 3).
Models Development
In the following section, we present the derivation of mathemat-
ical models for the reformer unit, the reduction unit and the
various auxiliary units (heat recuperator, scrubber and compres-
sor). For the sake of clarity, only model assumptions and kinetics
| 2 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
Figure 1. Schematic simplied diagram for the Midrex process.
Table 1. Design parameters of reformer
Parameter Value
Number of tubes 521
Inside tube diameter (m) 0.2
Tube thickness (m) 0.024
Tube length (m) 7.9
Catalyst particles characteristic length (m) 0.0038
Catalyst pellets bulk density (kg/m
3
) 250
Catalyst pore radius (A) 80
Tortosity 2.74
Table 2. Design parameters of reduction furnace
Parameter Value
Diameter (m) 5.5
Height (m) 9.1
Iron ore pellets radius (m) 0.012
Porosity 0.2
are shown in the text, while the detailed model equations for the
different units are presented in the appendix.
Model for the Reformer
As was mentioned in the Introduction section, a one-dimensional
heterogeneous model for the catalyst tubes is derived for the
reformer while the furnace is modelled with bottom-ring con-
guration.
Table 3. Nominal values of operating parameters for combined
reformer-reduction furnace plant
Parameter Value
Flowrate of hematite (Fe
2
O
3
) (ton/h) 130.0
Flow rate of gangue (SiO
2
) (ton/h) 8.0
Natural gas ow rate to reformer (10
3
Nm
3
/h) 22.4
Natural gas after the reformer ow rate (10
3
Nm
3
/h) 2.9
Oxygen after the reformer (10
3
Nm
3
/h) 1.1
Natural gas ow rate to transition zone (10
3
Nm
3
/h) 5.0
Natural gas ow rate to cooling zone (10
3
Nm
3
/h) 3.0
Fraction of scrubber exit not sent to the burner 0.68
Scrubber exit temperature (

C) 53
% Methane in natural gas 0.95
Reaction Kinetics
The kinetic rate expressions considered in this work are
those developed by Xu and Froment (1989), based on
LangmuirHinshelwood (HougenWatson) approach. The fol-
lowing reactions are considered:
CH
4
+ H
2
O CO + 3H
2
(1)
CO + H
2
O CO
2
+ H
2
(2)
CH
4
+ 2H
2
O CO
2
+ 4H
2
(3)
The kinetics of the dry reforming reaction:
CH
4
+ CO
2
2CO + 2H
2
(4)
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 3 |
Table 4. Kinetic rates for reforming (Xu and Froment, 1989).
Parameter Expression
k
1
[kmol/bar/kgcat h] 9.4910
15
exp(240.1/RT)
k
2
[kmol/kgcat hbar] 4.3910
6
exp(67.13/RT)
k
3
[kmol/kgcat hbar] 2.2910
15
exp(243.9/RT)
k
CO
[bar
1
] 8.2310
5
exp(70.65/RT)
k
CH
4
[bar
1
] 6.6510
4
exp(38.28/RT)
k
H
2
O
[] 1.7710
5
exp(88.68/RT)
k
H
2
[bar
1
] 6.1210
9
exp(82.9/RT)
k
1
[bar
2
] exp(26830.0/T +30.114)
k
2
[bar
2
exp(4400/T 4.036)
are negligible with respect to the steam reforming reactions (reac-
tions 1 and 3). CO
2
mainly hinders the forward reactions (2) and
(3). The authors (Xu and Froment, 1989; Froment and Bischoff,
2010) used an adsorptiondesorption mechanism model consist-
ing of 13 steps and reached the following rates equations:
r
1
= k
1
_
P
CH
4
P
H
2
O
P
2.5
H
2

P
0.5
H
2
P
CO
K
1
_
,DEN
2
(5)
r
2
= k
2
_
P
CO
P
H
2
O
P
H
2

P
CO
K
2
_
,DEN
2
(6)
r
3
= k
3
_
P
CH
4
P
H
2
O
P
3.5
H
2

P
0.5
H
2
P
CO
2
K
1
K
2
_
,DEN
2
(7)
With
DEN = 1 + k
CO
P
CO
+ k
H
2
P
H
2
+ k
CH
4
P
CH
4
+
k
H
2
O
P
H
2
O
P
H
2
(8)
where P
i
is the partial pressure of component i. The expressions
for reaction rate constants k
1
, k
2
and k
3
, absorption constants k
CO
,
k
H
2
O
, k
CH
4
and k
H
2
and the equilibrium constants K
1
and K
2
are
summarised in Table 4.
Reformer Model Assumptions
The following are the main assumptions used for the development
of the reformer model:
The reactor is at steady state.
Radial distribution of the temperature and the concentrations
of the different components inside the reactor are uniform (i.e.
one-dimensional model).
Heat and mass diffusions in the longitudinal direction are negli-
gible considering the very high gas velocity at which the reactor
is operated (i.e. axial dispersion is negligible).
Ideal behaviour is assumed for the gases.
Mass and heat transfer resistances between the uid and the
particle surface are negligible.
The detailed model equations for the reformer unit are given in
the appendix.
Model for Reduction Furnace
In this section, we present the derivation of a model that describes
the direct reduction of iron oxides in the furnace. As was men-
tioned in the introduction, the concept of the shrinking core model
is adopted. A mathematical formulation is obtained that denes
the mass ux of the reactant gas in terms of bulk gas phase
and pellet properties. A complete pellet model is developed for
the reduction furnace using three consecutive reaction schemes.
Afterwards, a one-dimensional non-isothermal moving bed model
is derived for the furnace.
Iron Oxides Reduction
Three consecutive reactions are considered for the reduction of
iron oxide ore (hematite). The reduction reactions by hydrogen
are represented by:
Hematite reduction to magnetite:
3Fe
2
O
3
+ H
2
2Fe
3
O
4
+ H
2
O (9)
Magnetite reduction to wustite:
Fe
3
O
4
+
4x3
x
H
2

3
x
Fe
x
O +
4x3
x
H
2
O (10)
Wustite reduction to iron:
Fe
x
O + H
2
xFe + H
2
O (11)
In the case of reduction by CO, the three reactions are:
Hematite reduction to magnetite:
3Fe
2
O
3
+ CO 2Fe
3
O
4
+ CO
2
(12)
Magnetite reduction to wustite:
Fe
3
O
4
+
4x3
x
CO
3
x
Fe
x
O +
4x3
x
CO
2
(13)
Wustite reduction to iron:
Fe
x
O + CO xFe + CO
2
(14)
Unlike previous works (Parisi and Laborde, 2004; Nouri et al.,
2011), all the reactions are assumed to be reversible. This is specif-
ically the case for the most difcult reduction step from wustite
to iron which is highly reversible (Negri et al., 1991).
Pellet Model
The three consecutive reactions are used to develop a pellet model
based on the shrinking core scheme. The following assumptions
are made:
Quasi steady-state assumption.
The reduction reactions are assumed to take place on three
sharp interfaces; one between iron and wustite, the second
between wustite and magnetite and the third between mag-
netite and hematite.
The gas diffusivities between the different layers are the same
for the reducing gases and are calculated based on bulk gas
conditions. (A more rigorous model that assumes that gas dif-
fusivities change with composition from one layer to another
would complicate unnecessarily the model.)
The water gas shift reaction takes place only close to the top
gas exit, and is modelled as suggested by Negri et al. (1991).
| 4 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
Figure 2. Schematic representation of the three-interface pellet model.
There is no volume change during the reduction. The reduction
process occurs at three advancing interfaces inside the hematite
particle (Figure 2).
The rate of moving for each interface is determined in terms of
rate constants for surface reactions and bulk gas conditions. The
key assumption in the model is the quasi steady-state approxi-
mation, that is the mass ux of the reactant diffusing towards
an interface is equal to the amount of the gas reacted plus the
remaining amount which diffuses to the following interface. The
diffusion of the gaseous product is related to the reactant by
the stoichiometric ratio which is 1/1 for both hydrogen and car-
bon monoxide. In our derivation, the subscripts 15 are used to
indicate the hematite/magnetite, magnetite/wustite, wustite/Fe,
Fe/boundary layer, and boundary layer/bulk interfaces, as shown
in Figure 2. The reaction rate O
j
of the reducing gas at the pellet
interface j is expressed in the following form:
O
j
= k
j
_
4r
2
j
_
_
c
j

c
j.p
K
j
_
(15)
where k
j
is the rate constant at interface j, r
j
the radius, c
j
the
concentration of the reducing gas, c
j.p
the concentration of the
product gas and K
j
is the equilibrium constant of the reduction
reaction. For the two reducing gases (H
2
and CO) we have a total
of six rate equations.
Let x, y and z be the reaction rates of one reducing gas
(e.g. hydrogen) with hematite (layer 1), magnetite (layer 2) and
wustite (layer 3) respectively. The rates are dened by:
x =
c
1

_
c
1.p
_
K
1
_
R
1
. y =
c
2

_
c
2.p
_
K
2
_
R
2
. z =
c
3

_
c
3.p
_
K
3
_
R
3
(16)
where R
j
is the resistance of the interfacial reaction and is equal
to 1/(k
j
r
2
j
). The rates x, y and z are related to O
j
simply by:
O
1
= 4x. O
2
= 4y. O
3
= 4z (17)
The reduction rate in the rst layer (magnetite/hematite) can
also be expressed in terms of the resistances of the intraparticle dif-
fusion of the reactant (R
m
) and the product (R
m
p
) in the magnetite
region:
x =
c
2
c
1
R
m
=
c
1

_
c
1.p
_
K
1
_
R
1
=
c
1.p
c
2.p
R
m
p
(18)
where
R
m
=
r
2
r
1
D
m
e
r
1
r
2
. R
m
p
=
r
2
r
1
D
m
ep
r
1
r
2
(19)
D
m
e
and D
m
ep
are the effective diffusion coefcients of the reduc-
ing gas and the product gas respectively. They are computed using
the turtosity and porosity of the layer. Similar equations can be
written for the other layers.
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 5 |
Reduction Furnace Model
A moving bed model is developed for the DRI furnace. The fol-
lowing assumptions are used:
Steady-state operation.
Both gas and solid streams move in plug ow with no radial
temperature or concentration gradients, that is one-dimensional
model.
Ideal gas behaviour is assumed for the reducing gases.
Spherical pellets are assumed with equal size and with constant
voidage of packing.
Heat transfer from the reactor wall to the surrounding is
assumed to occur through an effective convective heat transfer
coefcient.
The detailed model equations for the reduction furnace are
given in the appendix. In addition to individual models for
reformer and reduction units, simple models were also developed
for the recuperator, scrubber and compressors, and are given in
the appendix.
Assumptions for Transition and Cooling Zones
The main role of the transition zone in the reduction furnace is to
carburise the product iron which saves energy in the subsequent
rening in the electrical arc furnace. As to cooling zone, its main
role is to cool the product iron. The gas temperature proles in
the transition and cooling zones are very steep because the ow
rate of gas is small while that of the solid is large, resulting in
large changes in temperature. The solution of models for both
transition and cooling zones lead to difculties in convergence
because of the severe steepness of the proles. For this purpose,
we made a simple analysis to approximate the events taking place
in the transition zone. In the present study, we assume that 95%
of the methane in the transition zone decomposes into carbon
and hydrogen, and that 5% of the hydrogen is used in the fur-
ther reduction of iron oxides. As to cooling zone, the methane
decomposition is very minor and could be neglected.
Solution Strategy
The developed computer code starts by reading the design and
operating data for reduction furnace unit, reformer, oxygen ow
rate, enrichment natural gas ow rate and composition. The pro-
gram then carries out calculations for reduction furnace material
and heat balances equations, calculate process gas composition
after scrubber, compressor energy consumption, and carries out
material balances on the scrubber, compressor and recycle loop.
The scrubbed gas after compression is divided into two streams:
one goes to the reformer furnace to be burned while the other goes
to the reformer where it is mixed with fresh natural gas. For the
reformer calculations, we have two sides: the furnace side calcu-
lations, where the adiabatic ame temperature and radiation heat
transfer equations are calculated. These calculations are based on
guessed values for the reformer tubes outside temperature. On the
reformer tube side, differential material and heat balances require
the calculations of the catalyst effectiveness factors. The program
thus calculates the effectiveness factors for CO
2
and CH
4
from the
catalyst particle module at each differential step length, and then
calculates bulk gas concentrations, temperature and pressure at
each step. It then calculates new tube outer wall temperature so
that the transferred heat fromthe furnace is equal to the heat trans-
ferred from the tube to the reacting mixture. Iterations are carried
out until the outer wall temperature becomes constant within a
predetermined tolerance. The program then carries out material
and heat balance calculations for oxygen injection, adds enrich-
ment natural gas and adds natural gas from cooling and transition
zones to calculate bustle gas mix ow rate and composition.
The method of spline orthogonal collocation (Soliman, 1992)
is used to solve the resulting boundary value problems. The main
advantage of this technique is its adequate representation of the
model with minimum number of variables and its ability to solve
moderately steep proles.
In order to validate the model against the industrial data, a num-
ber of parameters were tted. These include the wall to uid heat
transfer coefcient in the reformer unit. The tting is justied by
the recent work of Shayegan et al. (2008) who showed that the
quality of predictions of their two-dimensional model depended
strongly on the tted correlations for heat transfer coefcient. The
other parameters are the effectiveness factors for methane and
carbon dioxide in the reformer. As for the reduction reactor, the
tting parameters are kinetic parameters for the side reactions.
These include methane decomposition, carbon monoxide dispro-
portionation and carbon monoxide reduction by hydrogenation,
as explained in the appendix. The tting is justied by the fact
Table 5. Validation of the model
Process Reformed Bustle Overall
Parameter gas gas gas plant
T (

C)
Plant NA 948.0 980.0
Model 53 947.0 976.0
CO (dry) %
Plant 24.5 36.5 36.0
Model 26.4 34.1 34.0
CO
2
(dry) %
Plant 22.5 3.5 3.6
Model 14.8 2.6 2.5
H
2
(dry) %
Plant 48.0 57.5 56.0
Model 56.5 60.8 58.7
CH
4
(dry) %
Plant 3.0 1.2 2.8
Model 0.9 1.4 3.6
N
2
(dry) %
Plant 2.4 1.7 1.7
Model 1.4 1.1 1.2
Metalisation %
Plant 94.0
Model 95.1
CO (wet) %
Plant 31.6
CO
2
(wet) %
Plant 2.3
H
2
(wet) %
Plant 54.5
CH
4
(wet) %
Plant 3.4
H
2
O (wet) %
Plant 7.1
N
2
(wet) %
Plant 1.1
Carbon content %
Plant 2.5
Model 2.4
Compression energy (kwh)
Model 9068.0
| 6 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
that the experiments used to get these kinetics are at conditions
different from dry reforming conditions.
RESULTS AND DISCUSSION
The simulation results for the plant at the base case conditions are
shown in Table 5. Also shown are the plant data which consist
of dry gas composition, metalisation degree and the carbon con-
tent. The results compare well except for the process gas where
the simulated results showlowmole fractions of methane and car-
bon dioxide as well as a large mole fraction of hydrogen. It should
be noted that since the model predicts that too much methane is
being steam reformed then it is natural that the hydrogen mole
fraction is large. Another reason for these discrepancies could
be due to the fact that the chosen kinetics predict the reduction
to occur more by carbon monoxide than by hydrogen. Measure-
ments errors in the plant data could also explain these model-plant
mismatches.
Key performance indicators are the degree of metalisation
which is 94.0% from the plant data and 95.1% from simulations.
The discrepancy is most likely due to the assumptions made in
the transition zone, as explained in the previous section. The ratio
H
2
/CO (on a dry basis) predicted by the model is 1.72 and is
close to the value of 1.56 obtained from the plant. The Midrex
Figure 3. Proles of the effectiveness factors along the steam reformer.
Figure 4. Prole of the gas temperature in the reduction reactor.
Figure 5. Prole of the solid temperature in the reduction reactor.
Figure 6. Proles of mole fractions in the reduction reactor.
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 7 |
Table 6. Qualitative effect of the different operating parameters on the performance of the plant: (+) increases, () decreases, (Max) exhibits a
maximum, (Min) exhibits a minimum
An increase in Bustle Reformer inlet Metallisation Carbon (H
2
+CO)/ Compression
the variable temp. temp. degree content H
2
/CO (H
2
O+CO
2
) power
NG to reformer Max Max Max Max + Max +
NG after reformer Max Max Max Max + Max +
Oxygen after reformer + + +
Recycle ratio Max Min +
Scrubber exit temp. + +
NG ow rate to cooling zone + + + + + + +
NG ow rate to transition zone + + + + + + +
Ratio of all ow rates Max + +
Figure 7. Effects of recycle ratio on the performance of the plant.
| 8 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
technology recommends that the ratio (H
2
/CO) should be about
1.5 and not less than 1.4 to avoid carbon deposition in the
reformer. A larger value would mean lower reduction furnace
temperature but this can be dealt by, for instance, increas-
ing oxygen injection. The reductants to oxidants ratio, that is
(H
2
+CO)/(H
2
O+CO
2
) obtained from simulations is 9.16 while
these data are not available from the plant since the sampling is
done on a dry basis. Too small values of this ratio would reduce
metalisation while larger values would lead to carbon deposi-
tion on the reformer catalyst. The percentage of carbon is around
2.4% from simulations and 2.5% from the plant data. This value,
together with the mole fraction of CO
2
of 3.6% suggests that the
plant is operating at reasonable conditions. Very low values of
CO
2
can lead to carbon deposition in the reformer.
Before the results of the parametric study of the overall plant
are shown, it may be useful to discuss some results pertinent
to the individual performance of steam reforming and reduction
reactors. Figure 3a,b shows the proles of the effectiveness fac-
tors for methane and carbon dioxide along the steam reformer
length. The values of the effectiveness factors are very small,
indicating that the reactions are very fast and take place in
a very small layer at the surface of the catalyst. This also
indicates diffusional limitations which is consistent with the con-
clusions reached in previous studies (Elnashaie and Abashar,
Figure 8. Effects of scrubber exit temperature on the performance of the plant.
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 9 |
1993). Intraparticle convection could also be a factor in the
efciency of the process, especially given the small size of the
pores. Some investigators (Quinta Ferreira et al., 1992; Oliveira et
al., 2010) showed that intraparticle convection can be promoted
by using large-pore structured catalysts that reduce intraparticle
gradients.
As for the reduction reactor, the proles of gas and solid tem-
perature are shown in Figures 4 and 5. It can be seen that since
the gas is fed counter-currently to the solid, the gas temperature
decreases starting from the dimensionless distance of 1 (bustle
gas) to the dimensionless distance of 0 (top gas). This decrease is
due to the endothermic nature of the reaction and to the fact that
the solid is fed at colder temperature. Figure 5 shows the prole
of solid temperature. The prole is quite close to that of the gas
temperature.
Figure 6 shows the proles for the mole fractions (wet gas basis)
of hydrogen and carbon monoxide in the reduction reactor. For
hydrogen, starting frombustle gas inlet conditions (dimensionless
distance of 1), both steam reforming and methane decomposi-
tion produce hydrogen which leads to the increases in its mole
fraction. Subsequently, the reduction of the oxides dominate,
causing a decrease in the hydrogen mole fraction. It should be
noted that at top gas conditions (dimensionless distance of 0)
there is the constraint that water gas shift reaction is at equi-
librium which leads to an increase in the hydrogen mole
fraction.
Figure 9. Effects of injected oxygen ow rate on the performance of the plant.
| 10 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
As for carbon monoxide, the reforming leads to the increase in
its mole fraction at the bustle gas inlet conditions (dimensionless
distance of 1) followed by a decrease in the mole fraction as result
of its use in the reduction of the iron oxide.
Parametric Study
In the following, we present simulations showing the effects of a
number of operating parameters on the performance of the plant.
The effects of the following parameters are studied:
Recycle ratio.
Scrubber exit temperature.
Injected oxygen ow rate.
Flow rate of natural gas to reformer.
Flow rate of natural gas after reformer.
Flow rate of natural gas to transition zone.
Flow rate of natural gas to cooling zone.
Besides these input parameters, we will be interested in know-
ing the effect of increasing all the ow rates including the ore
ow rate on the performance of the plant. Thus, we introduce
as an other input variable an articial parameter (r
f
) dened as
the ratio of actual ow rates to the ow rates in the nominal
case.
Figure 10. Effects of natural gas ow rate to reformer on the performance of the plant.
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 11 |
As for plant performance, the following key output parameters
are selected:
Bustle gas temperature.
Reformer inlet temperature.
Degree of metalisation.
Carbon content.
Ratio H
2
/CO.
Reductants to oxidants ratio (H
2
+CO)/(H
2
O+CO
2
).
Required compression energy.
The variations of each input parameter around the nominal
value were made quite wide but keeping in perspective that some
bounds are to be placed on some output parameters, for example
the metalisation degree should not go below 90%, the temper-
ature of the reduction furnace should not exceed 1250 K while
the ratios H
2
/CO and (H
2
+CO)/(H
2
O+CO
2
) should remain in
the range mentioned in the previous section. Besides the plots,
a summary of the sensitivity analysis is provided in Table 6
where the qualitative effects of the different input parameters are
illustrated.
Effect of Recycle Ratio
The effects of recycle ratio on the plant performance are shown
in Figure 7. The recycle ratio was varied from 0.65 to 0.75, the
nominal value being 0.68. An increase in the recycle ratio leads
to a lower amount of fuel to the reformer furnace, which means
Figure 11. Effects of natural gas ow rate after reformer on the performance of the plant.
| 12 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
less reforming, less reductants to oxidants ratio, lower bustle
gas temperature, lower inlet temperature to the reformer, less
metalisation, less carbon content and more compressor energy.
However, we will have more dry reforming than steam reforming
and thus lower H
2
/CO, which yields higher temperature inside
the reduction furnace, since more reduction is taking place by the
exothermic CO reduction. Lower temperatures favour high H
2
/CO
ratio. This is why we see a minimum at the prole of H
2
/CO.
Higher reduction temperatures in the furnace, on the other hand,
cause higher metalisation which is in conict with lower met-
alisation due to lower bustle gas temperatures. This is why the
metalisation prole exhibits a maximum with the increase in the
recycle ratio.
Effect of Scrubber Exit Temperature
The effect of scrubber exit temperature is shown in Figure 8.
The values of this temperatures were varied from 49 to 56

C
around the nominal value of 53

C. An increase in the scrubber


exit temperature means more water vapour in the process gas,
more steam reforming, more ow rates, less bustle gas tempera-
ture and less reduction of iron ore. This leads to less metalisation,
Figure 12. Effects of natural gas ow rate to transition zone on the performance of the plant.
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 13 |
less carbon content, lower reductants to oxidants ratio and higher
compression energy. As for the H
2
/CO ratio, it increases due to
more steam reforming of the excess water. Although a low scrub-
ber temperature is desired, it should not decrease below certain
limits, since a minimum amount of water is needed to prevent
carbon formation in the reformer tubes.
Effect of Injected Oxygen Flow Rate
The effect of injecting oxygen is shown in Figure 9. The ow
rate was varied in a wide range from 300 to 1700 Nm
3
/h around
the nominal value of 1100 Nm
3
/h. As the ow rate of oxygen
increases, some reformed hydrogen will burn. This leads to an
increase in the bustle gas temperature leading to more reduction
and carburisation which leads to an increase in metalisation and
carbon content. This also leads to a decrease in the ratio of reduc-
tants to oxidants. Since we have less reductants in the process gas,
there will be less H
2
and CO to burn in the reformer furnace, lead-
ing to lower ue gas exit temperature and lower inlet temperature
of the reformer.
Increasing the injected oxygen ow rate could also lead (as in
this case) to less H
2
/CO because of the combustion of hydrogen
by oxygen. Lower reformer inlet temperatures mean less reform-
ing, less overall ow rates and hence less compression energy.
Although increasing the injected oxygen ow rate is desirable,
there is a maximum bustle gas temperature for the safe operation
of the reduction furnace.
Figure 13. Effects of natural gas ow rate to cooling zone on the performance of the plant.
| 14 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
Effect of Natural Gas Flow Rate to Reformer and After
Reformer
As it can be seen in the summary Table 6, both these ow rates
(NG to reformer and NG after reformer) have similar effects on
the plant performance. The natural gas ow rate to reformer was
varied from 21,000 to 26,500 Nm
3
/h around the nominal value
of 22,400 Nm
3
/h. Figure 10 shows the results for the changes in
natural gas ow rate to reformer. The ow rate after reformer
was, on other hand, varied from 1000 to 7100 Nm
3
/h around the
nominal value of 2900 Nm
3
/h. The results of sensitivity analysis
are shown in Figure 11.
An increase in natural gas ow rate will lead initially to more
reforming, more reduction and more fuel to the reformer furnace.
This means higher bustle gas temperature, higher reformer inlet
temperature, higher H
2
/CO ratio, higher (H
2
+CO)/(H
2
O+CO
2
)
ratio, higher metalisation and higher carbon content. Compres-
sion energy increases because of the increase in natural gas ow
rate and process gas ow rate. However, as the ow rate of natural
gas increases, conditions causing less reforming in the reduc-
tion furnace occur due to the limited capacity of the reformer.
This leads to lower reductants to oxidants ratio. Also, because of
the limited capacity of the recuperator, reformer inlet tempera-
ture starts to drop at certain ow rate. Consequently, bustle gas
Figure 14. Effects of ratio of all ow rates on the performance of the plant. The enlargement in the upper left gure shows the maximum exhibited by
the prole of reformer inlet temperature.
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 15 |
temperature starts to drop. Combination of lower reducants to oxi-
dants ratio and lower bustle gas temperature leads to a decrease in
metalisation and carburisation. This explains the maxima in the
proles of metalisation, carburisation and reductants to oxidants
ratio.
Effect of Natural Gas Flow Rate to Transition Zone
and Cooling Zone
As it can be seen in the summary Table 6, these two streams (NG
to transition zone and NG to cooling zone) have the same effect
on the process. The ow rate to transition zone was varied from
3000 to 6700 Nm
3
/h around the nominal value of 5000 Nm
3
/h.
The results of the sensitivity analysis are shown in Figure 12. The
ow rate to cooling zone was, on the other hand, varied from
2000 to 4750 Nm
3
/h around the nominal value of 3000 Nm
3
/h.
The sets of Figure 13 showthe results of the sensitivity analysis for
this stream. Both ow rates to transition or cooling zone provide
hydrogen for the reduction zone due to methane decomposition,
and thus increase the reductants to oxidants ratio. An increase
in natural gas ow rate will lead to more reforming, more reduc-
tion and more fuel to the reformer furnace. This means higher
bustle gas temperature, higher reformer inlet temperature, higher
H
2
/CO ratio, higher (H
2
+CO)/(H
2
O+CO
2
) ratio, higher metali-
sation and higher carbon content. Compression energy increases
because of the increase in natural gas ow rate and process gas
ow rate. Therefore, although it is desirable to increase ow rates
of natural gas to the cooling and transition zones, increasing these
ow rates above certain limits could cause carbon formation in
the reformer tubes.
Effect of Increase in all Inlet Flow Rates Including Ore
In this section, we present the simulation results for the increase
in all inlet ows including the ore ow rate. The parameter (r
f
)
is introduced to represent a ratio multiplying all the inlet ows
including ore ow rate. The value of (r
f
) was varied from 0.925 to
1.072, the unity being the nominal value. The results of changes in
r
f
are shown in Figure 14. For a xed design, increasing the value
of r
f
means less metalisation, less reforming, less reductants to
oxidants ratio and higher compressor energy. The bustle gas tem-
perature decreases but the inlet temperature to the reformer shows
a maximum (shown in the enlargement of the rst gure). This
maximumis due to the initial decrease in the amount of heat going
to the reformer which later increases. The ratio H
2
/CO increases
due to low exit reformer temperature (lower bustle gas temper-
ature), since at low temperatures the water gas shift reaction is
favoured, with CO going to CO
2
and H
2
O going to H
2
.
CONCLUSIONS
This article presented the modelling and simulations of a DRI
plant based on Midrex technology. The approach followed in this
work was the development of rigorous models for the individual
units of the plant followed by validation and simulations stud-
ies. Mathematical modelling, while sometimes difcult, is always
useful when the model incorporates the essential elements of the
process and can be validated against plant data. The level of rigor
in the development of these models was constrained by the com-
plexity of the numerical solution as well as the availability of plant
data to validate the resulting model. Given the high temperature
nature of the process involved, the only data available from the
plant are some temperature measurements, dry gas composition
and metalisation degree.
In this regard, a detailed one-dimensional heterogeneous model
was developed for the reformer. The need for a two-dimensional
model was deemed not useful given that practically no data are
available for radial measurements.
Adetailed model based on the well-known shrinking core pellet
was developed for the reduction furnace. The models for the tran-
sition and cooling zones were approximated by simple algebraic
equations since a detailed model in these zones would result in too
steep proles (and stiff problems) while small extent of methane
decomposition takes place in the transition zone and even less in
the cooling zone. When these individual models were integrated
together, the effect of a number of key operating parameters on
the performance of the plant was clearly identied in appropriate
plots.
The sensitivity study could provide some practical guidelines
for the optimisation of the plant, subject to the constraints on
the prevention of carbon deposition in the reformer tubes. In this
regard, the parametric study has revealed that the scrubber exit
temperature could be lowered to the extent of providing enough
water vapour to prevent carbon formation. Recycle ratio should
be optimised subject to the same constraint. There is an optimum
natural gas/iron ore ratio. If we would like to increase natural
gas ow rate, it is preferable that this addition to be done in
the transition zone where methane decomposes to carbon and
hydrogen, and thus increases carburisation and metalisation. It is
also preferable that this natural gas is at elevated temperatures.
Injected oxygen ow rate should be high enough to increase the
bustle gas temperature to the level where no iron ore coagulation
is possible.
ACKNOWLEDGEMENTS
This work was made possible by a generous grant from the
National Plan for Science and Technology (Project # 08-ENE337-
2).
APPENDIX: MODEL EQUATIONS
Model Equations for Reformer
A one-dimensional heterogeneous model is used to describe the
catalyst tube, which takes into account the intraparticle mass
transfer resistance. One reformer tube performance is assumed to
be representative of any other tube in the furnace, by assuming the
feed gas to be distributed equally among the reformer tubes, and
that heat ux passing through each tube wall is the same regard-
less of the location of the tube inside the furnace with respect to
the location of the burners.
Reactor Equations
The rate of disappearance and formation of CH
4
and CO
2
(taken
as key components) are given by:
R
CH
4
= r
1
+ r
3
. R
CO
2
= r
2
+ r
3
(A.1)
where r
1
and r
2
are reaction rates for reactions (1) and (2) already
presented in the text. Let the conversions of methane and carbon
dioxide be dened by:
x
CH
4
=
_
n
CH
4
.f
n
CH
4
_
n
CH
4
.f
(A.2)
| 16 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
x
CO
2
=
_
n
CO
2
n
CO
2
.f
_
n
CH
4
.f
(A.3)
The molar ow rate n
i
of the component i in the reacting mix-
ture can be expressed, using stoichiometric relations, in terms of
the molar feed rate n
i.j
and the conversions x
CH
4
and x
CO
2
. Stoi-
chiometric relations in terms of x
CH
4
and x
CO
2
can be written as
follows:
n
CH
4
= n
CH
4
.f
x
CH
4
n
CH
4
.f
(A.4)
n
H
2
O
= n
H
2
O.f

_
x
CH
4
+ x
CO
2
_
n
CH
4
.f
(A.5)
n
CO
= n
CO.f
+
_
x
CH
4
x
CO
2
_
n
CH
4
.f
(A.6)
n
CO
2
= n
CO
2
.f
+ x
CO
2
n
CH
4
.f
(A.7)
n
H
2
= n
H
2
.f
+
_
3x
CH
4
+ x
CO
2
_
n
CH
4
.f
(A.8)
The total molar ow rate n
T
is given by:
n
T
= n
T.f
+ 2x
CH
4
n
CH
4
.f
(A.9)
The material, energy and momentum balance equations can be
written as follows:
dx
CH
4
dl
= A,

j
CH
4
R
CH
4
n
CH
4
.f
(A.10)
dx
CO
2
dl
= A,

j
CO
2
R
CO
2
n
CH
4
.f
(A.11)
dT
dl
=
1
C
p
U
s
_
,

LH
1
j
CH
4
(r
1
+ r
2
) + ,

LH
2
j
CO
2
(r
2
+ r
3
) + 4
U
dt
i
(T
S
T)
_
(A.12)
dP
T
dl
= j
_
G
2
,
g
d
P
_
(A.13)
T is the temperature of the reacting mixture, T
s
the temperature
of inner tube skin, P
T
the total pressure along the reformer, A
the reformer tube cross sectional area, ,

the bulk density of the


catalyst, j
CH
4
and j
CO
2
are the effectiveness factors associated with
the reactions of CH
4
and CO
2
, respectively, LH
i
(i =1,2) the heats
of reactions, u
s
the supercial gas velocity, Gthe gas mass velocity,
d
p
the catalyst particle diameter and d
ti
the internal diameter of
the reformer tubes.
The heat transfer coefcient U needed in the energy balance
equation (Eq. A.12) is calculated using the correlation developed
by De Deken et al. (1982); and Froment and Bischoff (2010).
Ergun correlation is used to calculate the friction coefcient
(f) in the momentum balance Equation (A.13) for large Reynolds
number:
j = 1.75
1

3
(A.14)
where is the bed void fraction.
Catalyst Particle Equations
The catalyst pellet is assumed to be a slab with a characteristic
length l
c
. This is a simplication since the actual shape of the
catalyst is a cylinder with holes. However, the adopted assump-
tion will simplify the modelling. Besides, the effectiveness factors
computed based on the selected shape will be further tuned to
validate the model.
The material balance equations for the catalyst pellet take the
following form:
d
2
P
CH
4.p
d
2
=
R
CH
4
RTl
2
c
D
CH
4.e
(A.15)
d
2
P
CO
2.p
d
2
=
R
CO
2
RTl
2
c
D
CO
2.e
(A.16)
with the boundary conditions,
P
CH
4.P
= P
CH
4.s
and P
CO
2.P
= P
CO
2.s
at = 1.0 (A.17)
dP
CH
4.P
d
=
dP
CO
2.P
d
= 0 at = 0 (A.18)
P
i.p
and P
i.s
designate, respectively, the partial pressure of com-
ponent i in catalyst particle and on catalyst particle surface, the
dimensionless coordinate of the catalyst pellet and D
i.e
is the effec-
tive molecular diffusitivity of component i. The slab is assumed
to be isothermal while external mass and heat transfer resistances
are assumed negligible.
The physico-chemical parameters such as viscosity, thermal
conductivity, molecular diffusitivities are obtained using standard
correlations (Reid and Sherwood, 1958; Froment and Bischoff,
2010).
It should be noted that in the model development, we assumed
that the ux of each species depends only on its concentration gra-
dient, which is a simplication. Amore rigorous model such as the
dusty gas model would account for the contribution of the uxes
of other components. However, this will complicate the modelling.
Moreover, some investigators (e.g. Elnashaie and Abashar, 1993)
have compared the dusty gas model against simple Fickian type
models for the steam reforming under industrial conditions, and
concluded that in the case of low steam to methane ratios the
predictions of the different models are in good agreement.
Model of the Furnace
One type of ring will be considered which is the bottom-red
furnace. Roesler (1967) introduced the two-ux model for one-
dimensional furnaces. The model is capable of taking into account
the radiative heat along and normal to the gas ow. We will adopt
in this work the Roesler model that was modied by Filla (1984)
to take the following form:
d
2
E
1
dl
2
=
1
_

_
E
1
o[T
4
g
_
+
t
o
t
_
E
1
o[T
4
t.0
_
+
r
o
r
((1[)E
1
[E
2
)
_
(A.19)
d
2
E
2
dl
2
=
2
_

t
o
t
_
E
2
o(1[)T
4
t.0
_
+
r
o
r
([E
2
(1[)E
1
)
_
(A.20)
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 17 |
with the following boundary conditions:
At l =0,
1

1
dE
1
dl
=
1

2
dE
2
dl
=

r
2
r
{(1[)E
1
[E
2
} (A.21)
At l =L,

1
dE
1
dl
=
1

2
dE
2
dl
=

r
2
r
{(1[)E
1
[E
2
} (A.22)
E
1
and E
2
are, respectively, the grey and clear gas component heat
gas ux, T
g
the temperature of furnace gas, T
t.0
the outer tube
surface temperature, o the Stefan-Boltzman constant, [ the band
radiation fraction, the absorption coefcient and
t
, are, respec-
tively, the emissivities of refractory slab and tube wall. In this
model, it is assumed that the gas emissivity can be approximated
by:

g
= [
_
1exp
_
BL

2
__
(A.23)
where L

is the beam length and

1
= + o
t
+ o
r
.
2
= o
t
+ o
r
(A.24)
where a
r
and a
t
are, respectively, the refractory furnace area and
the tube side furnace area to furnace free volume half ratio. The
refractory temperature T
r
is given by:
T
r
=
_
E
1
+ E
2
o
_
1,4
(A.25)
The differential heat balance on the ue gas stream takes the
following form:
G
g
C
g
d
_
T
g
F
_
T

T
g.0
_
dl
= 2
_
E
1
[E
g

(A.26)
where G
g
is the process ue gas mass velocity, T* the adiabatic
ame temperature, F the fraction of the fuel burned along the
reformer and E
g
is the gas emissive powers in side red furnaces.
The ue gas inlet temperature T
g.0
is given by:
T
g.0
=
_
E
1.0
o[
_
1,4
(A.27)
where E
1.0
is the value of E
1
at the gas inlet.
The heat transfer Q
r
by radiation to the reformer tubes is given
by:
Q
r
= 2
t
o
t
(E
r
E
t
)V (A.28)
where E
t
is the tube emissive powers in side red furnaces, E
r
=
oT
4
r
and V is the free volume of the reformer.
Model Equations for Reduction Furnace
Mass balance-solid phase
In the rst step, mass balance equations are written for iron oxides
in the interface positions as particles move down through the reac-
tor. The changes in iron oxides along the bed are represented in
terms of the variation of the radii of the shrinking interfaces for the
three layers. For the magnetite/hematite interface, the change in
the hematite concentration is dened as the oxygen fraction con-
sumed during the reduction at this interface. Consider a volume
of element (,J)J
2
t
Z, the change in the hematite concentration in
the pellets of this element, in terms of the interface radius, can be
expressed as follows:
LC
Fe
2
O
3
=
solid volume inFe
2
O
3
layer
..
(1
h
)L
_
4,3r
3
1
_
,
h
(1
b
),4J
2
t
LZ
. .
volume of reactor solid phase
(A.29)
where Z is the axial coordinate, r
1
the effective radius of interface
between hematite and magnetite in the pellet, d
t
the diameter of
the furnace and
h
and ,
h
are, respectively, the porosity and molar
density of hematite.
The number of moles of hematite consumed in this element is
related to the reaction rates by the following balance equation:
_

4
J
t
2
_
(1
b
)u
s
LC
Fe
2
O
3
=
4
_
x
H
2
+ x
CO
_

hm
(A.30)
In this equation, we use the stoichiometric ratio: one atom of
oxygen is removed from the pellet for each mole of hydrogen
or carbon monoxide reacted. The parameter
hm
is simply the
relation between the consumed oxygen atoms and the reduced
hematite molecules. x
H
2
and x
CO
are reaction rates with respect
to H
2
and CO, respectively, u
s
the solid velocity and
b
is the bed
bulk porosity.
Introducing the dimensionless reactor height:
=
Z
L
(A.31)
the balance equation for the hematite concentration in terms of
the interface radius can be rewritten as a differential equation:

hm
u
s
(1
h
)(1
b
),
h
(1
b
)L
d
_
(4,3)r
3
1
_
d
= 4
_
x
H
2
+ x
CO
_
(A.32)
The molar ux rate for the solids is given by:
G
s
= u
s
,
h
(1
b
)(1
h
) (A.33)
where ,
h
and
h
are the density and porosity of the hematite.
Thus, Equation (A.32) can be rewritten for the radius r
1
of the
magnetite/hematite interface as follows:
dr
1
d
=
L(1
b
)
_
x
H
2
+ x
CO
_
r
2
1
G
s

hm
(A.34)
Similar equations can be written for the wustite/magnetite, and
Fe/wustite layers:
dr
2
d
=
L(1
b
)
_
y
H
2
+ y
CO
_
r
2
2
G
s

mw
(A.35)
dr
3
d
=
L(1
b
)
_
z
H
2
+ z
CO
_
r
2
3
G
s

wFe
(A.36)
where r
2
and r
3
are, respectively, the effective radius of interface
between magnetite and wustite and between wustite and iron.
| 18 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|

mw
and
wFe
denote, on the other hand, oxygen atoms react-
ing/mole of magnetite converted to wustite, and oxygen atoms
reacting/mole of wustite converted to iron.
For these equations, the initial conditions are given by:
= 0. r
1
=r
2
=r
3
=r
4
(r
4
is the particle diameter) (A.37).
Mass balance-gas phase
Two mass balance equations are written for the reducing gases,
one for hydrogen and the other for carbon monoxide. The hydro-
gen molar balance in the element (,4)J
2
t
Z is given by:
gas volumetric owrate
..

4
J
t
2
u
g

b
LC
H
2
=
molar ux of H
2
..
j
H
2
o
p

4
J
2
t
(1
b
)LZ (A.38)
The mass ux of hydrogen j
H
2
is related to the reduction rates
at all interfaces by the following relation:
J
H
2
o
p
=
moles of H
2
consumed inall interfaces per unit time
..
4
_
x
H
2
+ y
H
2
+ z
H
2
_
(4,3)r
3
4
. .
Pellet volume
(A.39)
The mass balance equation for hydrogen can be rewritten as
follows:

b
u
g
,
g
_
4
/
3
_
r
3
4
(1
b
)L,
g
dC
H
2
d
= 4
_
x
H
2
+ y
H
2
+ z
H
2
_
(A.40)
Simply, this equation relates the change in hydrogen concen-
tration along the bed with the amount of hydrogen used to reduce
the three types of iron oxides. The gas mass velocity is dened as
follows:
G =
b
u
G
,
G
(A.41)
Thus, Equation (A.40) can be rewritten as follows:
dC
H
2
d
=
3(1
b
)L,
g
r
3
4
G
_
x
H
2
+ y
H
2
+ z
H
2
_
(A.42)
Similar equation can be derived for CO:
dC
CO
d
=
3(1
b
)L,
G
r
3
4
G
(x
CO
+ y
CO
+ z
CO
) (A.43)
with the following boundary conditions:
= 1. C
H
2
= C
H
2
. inlet
. C
CO
= C
CO. inlert
(A.44)
Energy balance-solid phase
The differential heat balance for solid relates the change in solid
temperature along the bed with the heat transferred from the gas
phase into the solid phase, and the heat generated or consumed
by the reduction reactions.
(,4)d
2
t
molar fux rate of solid
..
u
s
(1
b
)(1
h
),
h
C
ps
LT
s
=
number of pellets inunit volume
..
(1
b
)(,4)J
2
t
LZ
(4,3)r
2
4
4r
2
4
h
_
T
g
T
s
_
. .
heat transfer per pellet
(A.45)
T
g
and T
s
are, respectively, gas and solid temperature along the
reactor, h the heat transfer coefcient between the gas and solid
and C
ps
the solid heat capacity.
The solid energy balance equation can be rewritten as:
dT
s
d
=
3hL(1
b
)
G
s
C
ps
r
4
_
T
g
T
s
_
+
3hL(1
b
)
G
s
C
ps
r
3
4
H
R
(A.46)
with the boundary conditions:
= 0. T
s
= T
s. inlet
(A.47)
H
R
denes the heat generated or consumed by the reduction
reactions, and it is given by:
H
R
= (LH
1
)x
H
2
+ (LH
2
)y
H
2
+ (LH
3
)z
H
2
+(LH
4
)x
CO
+ (LH
5
)y
CO
+ (LH
6
)z
CO
(A.48)
with LH
i
being the heat of reaction for reaction R
i
, (i =16).
Energy balance-gas phase
The heat balance for the gas phase is written as:
gas molar owrate
..
G(,4)J
2
t
C
pg
LT
g
=
_
4r
2
4
_
total area ofpellets
..
(,4)J
2
t
(1
b
)LZ
(4,3)r
3
4
h
_
T
g
T
s
_
. .
heat ux of solid
(A.49)
Model Modication to Account for Steam Reforming
and Carburisation
In this part, we include the effect of other side reactions that could
occur in the reduction furnace. All these reactions are assumed to
take place on the iron layer. These reactions are:
Steam reforming reaction (Takahashi et al., 1986):
CH
4
+ H
2
O CO + 3H
2
(A.50)
The rate equation for this reaction is given by:
r
mf
= A
1
e
(
6.7710
3
)
,RT
_
P
CH
4
P
H
2
O

P
CO
P
3
H
2
K
M
_
V
Fe
(A.51)
and in the reverse reaction:
r
mb
= A
2
e
(4.1810
3
),RT
_
P
CO
P
3
H
2
P
CH
4
P
H
2
O
K
M
_
V
Fe
(A.52)
Methane decomposition reaction (Sawai et al., 1998):
CH
4
C + 2H
2
(A.53)
with the rate equation:
r
d
=
A
3
e
(5510
3
),RT
P
O.5
H
2
_
P
CH
4

P
2
H
2
o
c
K
d
_
V
Fe
(A.54)
where a
c
is the carbon activity.
Carbon monoxide disproportionation (Grabke, 1965):
2CO C + CO
2
(A.55)
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 19 |
The reaction rate equation is:
r
c
=
_
A
4
e
(27.210
3
),RT
P
0.5
H
2
+ A
5
e
(8.810
3
),RT
_
_
P
2
CO

P
CO
2
o
c
K
c
_
(A.56)
Carbon monoxide reduction by hydrogen (Grabke, 1965):
CO + H
2
C + H
2
O (A.57)
r
r
= A
6
e
(16.4810
3
),RT
_
P
CO
P
H
2

P
H
2
O
o
c
K
r
_
(A.58)
The subscripts m, d, c, r denote steam reforming, methane
decomposition, carbon monoxide disproportionation, and carbon
monoxide reduction by hydrogenation respectively. As was men-
tioned in the text, the parameters A
i
(i =1,6) appearing in the
reaction rates will be used as tuning parameters to t industrial
data. Reactions involving carbon and iron to form carbides will
not be included in the model due to the scarcity of information.
The reactor model will include two additional mass differ-
ential equations balances in the gas phase: one for methane
consumption:
dC
CH
4
d
=
r
m
+ r
d
G(4,3)r
3
4
(1
b
)L,
g
(A.59)
and the second one for carbon formation:
dC
C
d
=
r
d
+ r
c
+ r
r
2G
s
(4,3)r
3
4
(1
b
)L,
G
(A.60)
The gas balance equations (Eqs. A.42 and A.43) for hydrogen
and carbon monoxide are modied to the following forms:
dC
H
2
d
=
3(1
b
)L,
g
r
3
4
G
_
x
H
2
+ y
H
2
+ z
H
2
+
r
r
2r
J
3r
m
4
_
(A.61)
dC
CO
d
=
3(1
b
)L,
G
r
3
4
G
_
x
CO
+ y
CO
+ z
CO
+
(r
r
+ 2r
c
r
m
)
4
_
(A.62)
The carbon activity a
c
is based on the formula given by Chip-
man (1972):
logo
C
= 2300,T0.92 + (3860,T)C + log
C
1-C
(A.63)
where C is atom C/atom Fe.
The solid differential heat balance equation (A.46) is modied
to include the heat of all side reactions:
dT
s
d
=
2hL(1
b
)
G
s
C
ps
r
r
_
T
g
T
s
_
+
3hL(1
b
)
G
s
C
ps
r
3
4
H
R
+
L(1
b
)
_
r
3
4
r
3
3
_
G
s
C
ps
r
3
4
((LH
1
)r
m
+ (LH
d
)r
d
+ (LH
c
)r
c
+ (LH
r
)r
r
)
(A.64)
Kinetics
The kinetics used in this work are those developed by Tsay et al.
(1976a, 1976b)
For hydrogen production, the kinetic rates constants are:
k
1
= 160 exp
_
22. 000
RT
_
(A.65)
k
2
= 23 exp
_
17. 000
RT
_
(A.66)
k
3
= 20 exp
_
15. 200
RT
_
(A.67)
For the reduction of carbon monoxide, the kinetic rate constants
(m/s) are:
k
1
= 2700 exp
_
27. 200
RT
_
(A.68)
k
2
= 25 exp
_
17. 600
RT
_
(A.69)
k
3
= 17 exp
_
16. 600
RT
_
(A.70)
Model for the Recuperator
The recuperator is a heat exchanger where a cold side of heat
capacity C
c
is heated from t
c
to T
c
with heat source of capacity C
h
fromtemperature t
h
to T
h
. For counter-current ow, and assuming
that C
h
-C
c
, the following relation holds:
T
h
t
c
t
h
T
C
= e
N(1C
h
,Cc)
= (A.71)
where
N =
UA
C
h
(A.72)
Therefore,
T
h
t
C
= (t
h
T
C
) (A.73)
T
C
= t
h

(T
h
t
C
) (A.74)
The heat balance, on the other hand, yields:
C
C
(T
C
t
C
) = C
h
(t
h
T
h
) (A.75)
Eliminating T
c
between these two equations yields:
T
h
= t
h
+ (1)t
C
(A.76)
with
=
C
C
C
h
(C
C
,)C
h
(A.77)
Equation for the Compressor
The compressor power is given by:
k
j(k1)
P
1
V
_
(C
R
)
k1
k
1
_
(A.78)
| 20 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|
where C
R
is the compressor ratio (P
2
/P
1
), k =1.35 the polytropic
coefcient and j =0.7 the efciency.
Equation for the Scrubber
The equation for the scrubber is provided by the vapor pressure
of water (P
V
) given by:
log
10
(760 P
v
) = 8.07131
1730.63
233.426 + t(

C)
(A.79)
where P
V
is in atm.
REFERENCES
Akers, W. W. and D. P. Camp, Kinetics of the Methane-Steam
Reaction, AIChE J. 1, 471475 (1955).
Amundson, N. R., Solid-Fluid Interactions in Fixed and Moving
Beds With Small Particles, Ind. Eng. Chem. 48, 2635
(1956).
Amundson, N. R. and L. E. Arri, Char Gasication in a
Countercurrent Reactor, AIChE J. 24, 87101 (1978).
Chipman, J., Thermodynamics and Phase Diagram of the Fe-C
System, Metall. Mater. Trans. B 3, 5564 (1972).
De Deken, J. C., E. F. Devos and G. F. Froment, Steam
Reforming of Natural Gas: Intrinsic Kinetics, Diffusional
Inuences, and Reactor Design, Chem. React. Eng., ACS
Symp. Ser 196, 181197, Boston (1982).
Elnashaie, S. S. E. H. and M. E. E. Abashar, Steam Reforming
and Methanation Effectiveness Factors Using the Dusty Gas
Model Under Industrial Conditions, Chem. Eng. Sci. 32,
177189 (1993).
Elnashaie, S. S. E. H., A. M. Adris, M. A. Soliman and A. S.
Al-Ubaid, Digital Simulation of Industrial Steam Reformers
for Natural Gas Using Heterogeneous Models, Can. J. Chem.
Eng. 70(4), 786793 (1992).
Farhadi, F., M. Y. M. Hashemi and M. B. Babaheidari, Modelling
and Simulation of Syngas Unit in Large Scale Direct
Reduction Plant, Ironmak. Steelmak 30, 1824 (2003).
Filla, M., An Improved Roesler Type Flux Method for Radiative
Heat Transfer in One-dimensional Furnaces, Chem. Eng. Sci.
39, 159161 (1984).
Fraser, S. D., M. Monsberger and V. Hacker, A Thermodynamic
Analysis of the Reformer Sponge Iron Cycle, J. Power Sour.
161, 420431 (2006).
Froment, G. F., L. B. Bischoff and J. De Wilde, Chemical
Reactor: Analysis and Design, 3rd ed., Wiley, New York
(2010).
Fujita, K., T. Harada, H. Michishita and H. Tanaka, CO
2
Emission Comparison between Coal-based Direct Reduction
Process and Conventional Blast Furnace Process,
International Symposium on Ironmaking for Sustainable
Development, Osaka, Japan (2010).
Grabke, H.-J., Die Kinetik der Entkohlung und Aufkohlung von
,-Eisen in Methan-Wasserstoff-Gemischen, Berichte der
Bunsengesellschaft 69, 409414 (1965).
Larsson, M. and J. Dahl, Reduction of the Specic Energy Use
in an Integrated Steel PlantThe effect of an Optimisation
Model, ISIJ Int. 43, 16641673 (2003).
Larsson, M., C. Wang and J. Dahl, Development of a Method
for Analysing Energy, Environmental and Economic
Efciency for an Integrated Steel Plant, Appl. Therm. Eng.
26, 13531361 (2006).
Midrex, Inc. :www.midrex.com. (2011).
Murty, C. V. S. and M. V. K. Murthy, Modelling and Simulation
of a Top-Fired Reformer, Ind. Eng. Chem. Res. 27,
18321840 (1988).
Negri, E. D., O. M. Alfano and M. G. Chiovetta, Direct
Reduction of Hematite in a Moving Bed. Comparison Between
One- and Three-Interface Pellet Models, Chem. Eng. Sci. 42,
24722475 (1987).
Negri, E. D., O. M. Alfano and M. G. Chiovetta, Direct
Reduction of Hematite in a Moving-Bed Reactor. Analysis of
the Water Gas Shift Reaction Effects on the Reactor
Behaviour, Ind. Eng. Chem. Res. 30, 474482 (1991).
Nouri, S. M. M., H. Ale-Ibrahim and E. Jamshidi, Simulation of
Direct Reduction Reactor by the Grain Model, Chem. Eng. J.
166, 704709 (2011).
Oliveira, E. L. G., C. A. Grande and A. E. Rodrigues, Methane
Steam Reforming in Large Pore Catalyst, Chem. Eng. Sci. 65,
15391550 (2010).
Parisi, D. R. and M. A. Laborde, Modelling of Counter Current
Moving Bed Gas-Solid Reactor Used in Direct Reduction of
Iron Ore, Chem. Eng. J. 104, 3543 (2004).
Pedernera, M. N., J. Pina, D. O. Borio and V. Bucala, Use of a
Heterogeneous Two-Dimensional Model to Improve the
Primary Steam Reformer Performance, Chem. Eng. J. 94,
2940 (2003).
Quinta Ferreira, R. M., M. M. Marques, M. F. Babo and A. E.
Rodrigues, Modelling of the Methane Steam Reforming
Reactor With Large Pore Catalysts, Chem. Eng. Sci. 47,
29092914 (1992).
Rao, Y. K. and P. Pichestapong, Modelling of the Midrex
Direct-Reduction Iron Making Process: Mass Transfer and
Virtual Equilibrium at Steady State, In: XVth CMMI
Congress, Johannesburg, SAIMM (1994), pp. 8
Reid, R. and T. Sherwood, The Properties of Gases and
Liquids, McGraw Hill, New York (1958).
Roesler, F. C., Theory of Radiative Heat Transfer in Co-Current
Tube Furnaces, Chem. Eng. Sci. 22, 13251336 (1967).
Sawai, S., Y. Iguchi and S. Hayashi, Iron Carbide Formation
With CO Gas and CO-H
2
Gas Mixture Under Low Sulfur
Potential Using the Two-Step Process of Metallisation and
Carbidisation, J. Iron Steel Inst. Jpn. 84, 844849
(1998).
Schultmann, F., B. Engels and O. Rentz, Flowsheeting-Based
Simulation of Recycling Concepts in the Metal Industry, J.
Clean. Prod. 12, 737751 (2004).
Shayegan, J., M. M. Y. Hashemi and K. Vakhshouri, Operation
of an Industrial Steam Reformer Under Severe Condition: A
Simulation Study, Can. J. Chem. Eng. 86(4), 747755
(2008).
Soliman, M. A., A Spline Collocation Method for the Solution of
DiffusionConvection Problems with Chemical Reactions,
Chem. Eng. Sci. 47, 42094213 (1992).
Soliman, M. A., S. S. E. H. Elnashaie, A. S. Al-Ubaid and A.
Adris, Simulation of Steam Reformers for Methane, Chem.
Eng. Sci. 43, 18011806 (1988).
Takahashi, R., Y. Takahashi, J. Yagi and S. Omori, Operation
and Simulation of Pressurised Shaft Furnace for Direct
Reduction, Trans. ISIJ 26, 765774 (1986).
Tsay, Q. T., W. H. Ray and J. Szekely, The Modelling of
Hematite Reduction with Hydrogen Plus Carbon Monoxide
Mixtures: Part I. The Behaviour of Single Pellets, AIChE J.
22, 10641072 (1976a).
Tsay, Q. T., W. H. Ray and J. Szekely, The Modelling of
Hematite Reduction with Hydrogen Plus Carbon Monoxide
|
VOLUME 9999, 2011
| |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| 21 |
Mixtures: Part II. The Direct Reduction Process in a Shaft
Furnace Arrangement, AIChE J. 22, 10721079 (1976b).
Valipour, M. S., M. Y. M. Hashemi and Y. Saboohi,
Mathematical Modelling of the Reaction in an Iron Ore Pellet
Using a Mixture of Hydrogen, Water Vapor, Carbon monoxide
and Carbon dioxide: An Isothermal Study, Adv. Powder
Technol. 17, 277296 (2006).
Xu, J. and G. F. Froment, Methane Steam Reforming,
Methanation and Water-Gas Shift: I. Intrinsic Kinetics,
AIChE J. 35, 8896 (1989).
Manuscript received March 11, 2011; revised manuscript
received April 12, 2011; accepted for publication April 19, 2011
| 22 |
THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
| |
VOLUME 9999, 2011
|

You might also like