Sie sind auf Seite 1von 8

ISIJ International, Vol. 40 (2000), No. 10, pp.

935942

An Experimental Study on the Kinetics of Fluidized Bed Iron Ore


Reduction
Arno HABERMANN, Franz WINTER, Hermann HOFBAUER, Johann ZIRNGAST1) and
Johannes Leopold SCHENK1)
Christian Doppler Laboratory, Institute of Chemical Engineering, Fuel and Environmental Technology, Vienna University of
Technology, Getreidemarkt 9/159, A-1060 Vienna Austria. E-mail: ahaber@mail.zserv.tuwien.ac.at.
1) VOEST-ALPINE Industrieanlagenbau GmbH (VAI), Turmstrasse 44, P. O. Box 4, A-4031 Linz Austria.
(Received on November 22, 1999; accepted in final form on June 5, 2000 )

To optimize existing iron ore reduction processes or to develop new ones, it is necessary to know the reduction kinetics of the iron ore of interest under the relevant operating conditions. In this work the reduction
kinetics of hematite fine iron ore was studied for industrial-scale processes using the fluidized bed technology. Especially designed batch tests were performed in a laboratory-scale fluidized bed reactor fluidized with
H2, H2O, CO, CO2, N2 at atmospheric and elevated pressures to simulate the relevant process conditions. To
obtain the reduction rates and the degree of reduction, the concentrations of H2O, CO, and CO2 in the outlet gas were analyzed by FT-IR spectroscopy.
Preliminary reduction tests showed a strong effect of the sample weight on the reduction rates, especially in the early stages of reduction. The optimum sample weight was determined by partly replacing the
hematite with silica sand. Additionally, the silica sand provided a constant and stable flow pattern throughout the reduction tests. The effects of temperature, gas composition, particle size and pressure on the rates
of reduction were tested and discussed.
Rate analysis showed the existence of two phases with different rates during the reduction tests.
Additional investigations (microscope analysis, SEM) demonstrated that in the first phase the rates were
controlled by mass transport in the gas phase and in the second phase by the reduction process within the
small grains of the iron ore particles.
KEY WORDS: iron ore reduction; high temperature fluidized bed; reduction kinetics; elevated pressure;
H2CO gas mixture.

1.

In the 1960s intensive investigations on the fluidized bed


reduction of iron ore fines by pure hydrogen or hydrogen/
nitrogen-mixtures were started.513) The reduction processes
use a reducing gas produced by steam reforming of natural
gas. Therefor the reducing gas is a mixture of H2, CO, CO2,
CH4, N2 and H2O. Thus a strong interest in the kinetics of
reduction by a reducing gas containing CO and CH4 exists.
Some authors tested the kinetics of reduction by mixtures
of H2, CO, CO2 and H2O,1416) others the carburization of
iron ore fines by adding CH4 to the reducing gas.17,18)
Industrial scale fluidized bed reduction units operate at elevated pressure, but only a few kinetic studies for the reduction of iron ore in fluidized beds were done at elevated pressures.1922)
The scope of this work was to investigate the reduction
kinetics of iron ore fines under conditions similar to industrial scale units. That means that a laboratory fluidized bed
reactor for reduction tests had to be operating at elevated
pressure, in wide temperature range and with a reducing
gas containing all species of reformed natural gas.

Introduction

In the near future many so-called mini mills will replace


the conventional route of steelmaking in integrated mills.
Mini mills use scrap or scrap substitutes (e.g. direct reduced iron) as feed material in electric arc furnaces.1) To
satisfy the increasing demand on scrap substitutes new
processes for the direct reduction of iron ores have to be developed. Direct reduction processes operate in shaft furnaces, rotary kilns and fluidized beds. The reduction in fluidized bed reactors offers the advantages of no feed agglomeration, uniform temperature in the reactor and excellent heat and mass transport. Some of these fluidized bed
processes (e.g. FIOR process, FINMET process2) and
Circored process) produce hot briquetted iron, whereas others (e.g. Iron Carbide process) produce Fe3C. Both products
can be melted in electric arc furnaces. The second application of fluidized bed reactors in iron ore reduction is the
prereduction stage in a smelting reduction process (e.g.
FINEX process,3,4) DIOS process. For efficient use of raw
materials fundamental knowledge on reduction kinetics is
required.
935

2000 ISIJ

ISIJ International, Vol. 40 (2000), No. 10

Fig. 1. The experimental setup.

2.

differential pressure between the fluidized bed reactor and


the pressure vessel. At the exit valve the pressure was reduced to atmospheric pressure and the gas was analyzed,
burned and sent to the chimney.
To minimize iron ore losses during the reduction tests,
the upper part of the reactor was designed of a conical
shape to reduce the superficial velocity. The cyclone located within the reactor separated the iron ore particles from
the exhaust gas. One important aspect of the reactor design
was to guarantee the same operating conditions along the
height of the reactor. Therefore, the heating shells had to be
fixed up to the top flange of the reactor.
The iron ore samples were placed into the reactor and the
reactor purged with pure nitrogen as inert gas while heating
up to the desired temperature. As soon as the desired operating temperature and pressure had been reached the fluidizing gas was switched to the reducing gas composition.
While the reduction test gas analysis was obtained by using
FT-IR spectroscopy in combination with a long-path length,
low-volume heated gas cell. The following species were
measured: CO2, CO, CH4 and H2O. Since H2O was analyzed, the reactor exit tubes and the total sample line
(pump, filter, gas cell) had to be heated to prevent condensation. At the end of the reduction test the fluidizing gas
was switched back to pure nitrogen again and the reactor
was cooled down to ambient temperature. The reduced iron
ore samples were removed and analyzed by a titriometric
method (iron chloride method) for total Fe, Fe and Fe12.
To obtain knowledge about the structural changes and
spatial element (Fe, O) distribution within the iron ore particles during reduction, partly and totally reduced particles
were analyzed by optical and scanning electron microscopy

Experimental

2.1. Experimental Setup


The operating conditions of the newly developed fluidized bed reactor can be summerized as follows:
absolute pressure up to 10 bar
temperature range from 773 to 1 173 K
the reducing gas contains H2, CO, CO2, CH4, H2O and
N2
The concept of the experimental setup is shown in Fig. 1:
The fluidized bed reactor and the electrical heating shells
were set into a pressure vessel to avoid pressure diffference
at the hot walls of the reactor (up to 1 173 K). This concept
guarantees the fulfillment of requirements and safe operation under the conditions listed above. The welding examination and pressure test of the apparatus were performed by
the Technical Inspection Association (TUEV). The reducing gas was stored in gas bottles and the demanded mixture
was provided by mass flow controllers. If the reducing gas
contained H2O the gas was sent through a heated water bath
before entering the reactor, i.e. the reducing gas was enriched with H2O to the demanded content. To heat up the
reducing gas before entering the fluidized bed where the reduction of the iron ore takes place, the gas was preheated in
a fluidized bed filled with pure silica sand. Both fluidized
beds were heated by electrical heating shells. The temperatures were controlled by PID-controllers and the temperatures of the fluidized bed for reduction, the preheating bed,
the freeboard (cyclone height), the exhaust gas and inside
the pressure vessel were recorded. The pressure within the
fluidized bed reactor was controlled by the exit valve. The
pressure in the pressure vessel was regulated to minimize
2000 ISIJ

936

ISIJ International, Vol. 40 (2000), No. 10


Table 1.

Chemical analysis of the hematite ore in mass%.

Table 2.

Physical properties of the hematite ore.

Fig. 3. Concentration of exhaust gas measured by FT-IR spectroscopy for conditions stated in Fig. 2.

ORed [mol]
Oout [mol]
Oin [mol]

number of oxygen atoms reduced from the


iron ore
total number of oxygen atoms leaving the
reactor in the form of H2O, CO, and CO2
total number of oxygen atoms entering the
reactor

ORed in each time step i divided by the total number of oxygen atoms initially present in the iron ore and the time step
gives the actual rate of reduction:
dR ORed, i
...............................(2)
5
dt O0 t
O0 [mol]
Fig. 2. Typical FT-IR spectrum obtained during the first stage of
iron ore reduction (Temperature: 1 053 K, pressure: 1.2
bar, particle size: 0.1250.5 mm, inlet gas composition:
55 vol% H2, 9 vol% CO, 5 vol% CO2, 6 vol% CH4, 25
vol% N2)

dR
[1/min]
dt
D t [min]

FexOy1COFexO(y21)1CO2

Reaction 2

time step

R5
R [2]
D O [mol]

2.2. Determination of the Rate of Reduction


FT-IR spectra were recorded every 10 s (Figure 2 shows
a typical spectrum obtained during the first stage of reduction.). Using predefined calibration functions the progress
of the height or area of a selected peak led to the concentration plot shown in Fig. 3. All oxygen containing species
(H2O, CO, and CO2) were detected by FT-IR spectroscopy.
The number of oxygen atoms reduced from the iron ore by
Reactions 1 and 2 can be calculated by the following balance of the oxygen atoms (Eq. (1)):
Reaction 1

rate of reduction

The fractional reduction can be obtained by:

techniques.
The experiments were carried out with Mt. Newman
hematite ore from Western Australia. Chemical analysis
and physical properties of the ore are listed in Table 1 and
Table 2. The particle size was 0.1250.5 mm, unless stated
differently.

FexOy1H2FexO(y21)1H2O

total number of oxygen atoms initially


present in the iron ore

O
...................................(3)
O0

fractional reduction
loss of oxygen

D O is calculated by summarizing the number of oxygen


atoms reduced from the iron ore (ORed in [mol]) in each
time step i:
t

O5

Red, i

..............................(4)

i50

Using Eq. (4) the rate of reduction and the fractional reduction can be determined (Fig. 4). For a better comparison of
the reduction tests carried out at different conditions the
rate of reduction is plotted versus fractional reduction (R)
(Fig. 5).
The fractional reduction of the iron ore samples at the
end of a reduction test was calculated as shown above and

ORed5Oout2Oin ....................................................(1)
937

2000 ISIJ

ISIJ International, Vol. 40 (2000), No. 10

Fig. 4. Rate of reduction obtained from gas concentration measurements (Fig. 3) and fractional reduction plotted versus
time for conditions stated in Fig. 2.

Fig. 6. Mixing of iron ore particles (dark) in a fluidized bed of


silica sand observed in the cold model. The photo shows
the state of mixing after 60 min. of fluidization.

Fig. 5. Rate of reduction versus fractional reduction for conditions stated in Fig. 2.

Fig. 7. Effect of sample weight on reduction rate of hematite reduced at 1 bar and 1 053 K (reducing gas: 55 vol% H2, 9
vol% CO, 5 vol% CO2, 31 vol% N2).

the result was compared with the titriometric analysis performed with the sample. The difference between these two
different methods was always less than 5%.

ore so that the minimum fluidization velocity for both materials was equal and the iron ore particles mixed well with
the silica sand. This was proved by tests performed in a
cold model plexiglass unit (Fig. 6). To obtain hydrodynamic similarity between cold model tests and reduction tests
the operating conditions for the cold model tests were defined using scaling criteria.23) Optical observations showed
perfect mixing of the iron ore and the silica sand.
The mass of the iron ore sample was reduced stepwise
from 300 g to 7 g to find the optimum sample weight: As
shown above a high mass of iron ore led to gas supply controlled reactions whereas a small mass of iron ore produced
inaccurate results especially at the end of the reduction test
when the concentration changes in the exhaust gas were
very low. In Fig. 7 the effect of the iron ore sample weight
is shown: The rate of reduction increased strongly with decreasing sample weight. At a sample weight less than 30 g
inaccurate results (caused by very low changes in the exhaust gas) arose. The maximum H2O concentration in the
exhaust gas for a sample weight of 30 g was 8 vol%. This
means that the reducing potential of the exhaust gas was

2.3. Variation of Sample Weight


Preliminary reduction tests were performed with a sample weight of 300 g of iron ore. This sample weight (at a superficial velocity of 0.25 m/s) led to a well mixed bubbling
fluidized bed with a bed height of 70 mm. The quantitative
analysis of the exhaust gas showed that in the first stages of
reduction the H2O content exceeded 40 vol%, i.e. most of
the H2 at the inlet (55 vol%) was consumed by the reduction
of the iron ore and the rate of reduction was controlled by
the supply of H2 in the reducing gas. Under those conditions the determination of the reduction rate will give misleading results. To prevent this effect, the mass of iron ore
in the reactor had to be decreased. The iron ore was partly
replaced by the inert material silica sand to maintain constant hydrodynamic conditions in the fluidized bed along
the whole reduction process. The bed of silica sand provided a constant matrix for the iron ore particles in terms of
hydrodynamic properties. However, the size of silica sand
particles had to be adapted to the higher density of the iron
2000 ISIJ

938

ISIJ International, Vol. 40 (2000), No. 10

Fig. 8. Effect of temperature on reduction rate of hematite reduced by H2CO gas mixtures (55 vol% H2, 9 vol% CO, 5
vol% CO2, 31 vol% N2).

Fig. 9. Effect of absolute pressure on reduction rate of hematite


reduced at 1 053 K.
A: pH250.55 bar, pCO50.09 bar, pCO250.05 bar
B: pH250.55 bar, pCO50.09 bar, pCO250.05 bar
C: pH251.65 bar, pCO50.27 bar, pCO250.15 bar
Balance is N2.

still high and a differential operation of the reactor was possible, i.e. the conditions (e.g. gas composition, temperature)
in the reactor did not significantly change along the height
of the reactor. The first stage of the reduction was controlled by gas-phase mass transport (identified by a relatively constant rate of reduction, see also Sec. 3). For these reasons 30 g iron ore (combined with 270 g silica sand) was
considered to be the optimum sample weight. All further
experiments were carried out with this sample weight.
3.

stant H2 partial pressure led to no change in the rate of reduction during the early and medium stages of reduction,
whereas the increase of the H2 partial pressure led to significantly higher rates of reduction.
An increase of the rate of reduction with increasing pressure was also observed by Ahner et al.,19) Ashie et al.21) and
Sato et al.22) Most of these tests were carried out with an increased molar flow at increased pressure and gave the same
results as shown above. Reduction tests carried out at increased pressure and constant total molar flow (i.e. at constant molar flow H2) led to higher rates of reduction and
seem to be in disagreement to test B in Fig. 9. But these experiments at constant total molar flow led to a decrease in
the superficial velocity (i.e. increase in residence time of
the gas in the reactor) at increased pressure19) and can not
be compared with test B for that reason. In agreement to
the results presented in this paper Kawasaki et al.25) found
that the variation of absolute pressure had no effect on the
rate of reduction.

Results and Discussion

3.1. Variation of Temperature


In Fig. 8 the effect of temperature on the reduction rate is
presented: Increasing the reaction temperature led to higher
rates of reduction. The strongest effect was observed between 823 K and 973 K, while this effect was less strong at
higher temperatures. It was noticed that the rate of reduction dropped to almost zero for the test performed at 823 K
when the fractional reduction reached 0.5. Formation of
carbon was observed at the tests carried out at 823 K and
898 K. These results are in agreement with those of experiments performed by several researchers.5,911,22,24)
For all tests at temperatures higher than 898 K two phases with different rates of reduction can be seen (Fig. 8): The
first very fast peak at the beginning of the experiment is a
result of a start effect and the very fast reduction of
hematite to magnetite and is not considered as a special
phase of the reduction. The following first phase of reduction is characterized by constant rates of reduction and ends
with a strong decrease in the rate of reduction. During the
last phase the fractional reduction increases from 0.6 to
0.85 at very small rates of reduction.

3.3. Variation of Gas Composition


In Figs. 10 and 11 the influence of the gas composition
on the rate of reduction is given. Higher concentrations of
H2 in the reducing gas gave higher rates of reduction, addition of H2O to the reducing gas decreased the rates of reduction. If the reducing gas contained 55 vol% H2 the addition of 9 vol% CO and 5 vol% CO2 did not affect the rate of
reduction. The reason for this effect is the slow rate of reduction by only COCO2 containing mixtures (Fig. 11).
The addition of small amounts (e.g. 6 vol%) of CH4 did not
change the rate of reduction.
The same efffect of changing H2 and H2O concentrations
in the reducing gas was observed by Hutchings et al.15)
Reduction by COCO2 gas mixtures was also studied by
Hutchings et al., Jess et al.17) and Kawasaki et al.25) leading
to similar result as shown above.

3.2. Variation of Pressure


The effect of absolute pressure is shown in Fig. 9. The
tests were carried out with constant superficial gas velocity,
i.e. the overall molar flow for a test at 3 bar was three times
higher than for the tests at atmospheric pressure. For test B
the molar flow (or partial pressure) of H2 was kept constant
(at 0.55 bar), for test C the molar flow was increased to
keep the H2 concentration of the reducing gas constant (at
55 vol%). The increase of the absolute pressure with con-

3.4. Variation of Particle Size


Most industrial applications use iron ores with a wide
particle size distribution. Figure 12 shows the rate of reduction for particles larger than 0.5 mm compared with the
939

2000 ISIJ

ISIJ International, Vol. 40 (2000), No. 10

Fig. 10. Effect of reducing gas composition on the reduction rate


of hematite at 1 bar and 1 053 K.
A: 55% H2, 9 % CO, 5 % CO2
B: 75% H2
C: 55% H2
D: 55% H2, 9% CO, 5 % CO2, 6 % CH4
E: 55% H2, 9 % CO, 5 % CO2, 10 % H2O
Balance is N2.

Fig. 12. Effect of particle size on reduction rate of hematite reduced at 1 053 K, 1 bar and by H2CO gas mixtures (55
vol% H2, 9 vol% CO, 5 vol% CO2, 31 vol% N2).

Fig. 13. Calculated rate of reduction for different rate controlling


steps: A: Gas transport control; B: Interface reaction
control; C: Pore diffusion control.

caused by the transition from magnetite reduction to


wustite reduction. Samples taken during the second (slow)
phase of reduction (80 min of reduction by 55 vol% H2 at
1053 K) were analyzed and consisted of 91% total Fe,
75.2% Fe, 11.5% Fe12 and 4.3% Fe31. Fe31 was found in
samples taken during the slow phase of reduction of all
tests and indicates the existence of magnetite. Different kinetics of magnetite reduction and wustite reduction can
therefor not cause the strong decrease of the rate and a different mechanism has to be found. Hence the possible rate
controlling steps for iron ore reduction will be analyzed in
the next section.
The reduction of iron oxide particles in a fluidized bed
reactor must proceed through the following steps:
1. Transport of the gaseous reactants (H2 or CO) from
the bubble phase of the fluidized bed into the emulsion phase.
2. Transport of the gaseous reactants from the emulsion
phase to the external surface of the iron ore particle.
3. Diffusion of the gaseous reactants through the pores
of the particle to the internal surface.
4. Reduction of the iron ore
H21O22H2O22e2
Fe2112e2Fe
5. Diffusion of products (H2O, CO2) from the pores of

Fig. 11. Effect of reducing gas composition on the reduction rate


of hematite ore reduced at 1 bar and 1 053 K.
A: 55% H2, 9 % CO, 5 % CO2
B: 20% CO, 5 % CO2
C: 30% CO, 2.5 % CO2
Balance is N2.

usually used particle size of 0.1250.5 mm. It can be recognized that the larger particles reach the same fractional reduction at the end of the reduction, but the rates of reduction decreased for increasing particle size. This is in agreement with the results obtained by McKewan26) and differs
from the results obtained by Meissner et al.14) This apparent
disagreement may be caused by diffferences in structure
and porosity of the analyzed ores.
3.5. Rate Controlling Step
Most of the diagrams rate of reduction versus fractional
reduction showed a peak at the beginning of the experiment
(as a result of a start effect and the very fast reduction of
hematite to magnetite) followed by two phases with significant different rates of reduction. These two phases are separated by a strong decrease of the rate at a fractional reduction of 0.5. One might interpret that this decrease was
2000 ISIJ

940

ISIJ International, Vol. 40 (2000), No. 10

the particles to the external surface.


6. Transport of products from the external surface into
the emulsion phase
7. Transport of products from the emulsion phase into
the bubble phase of the fluidized bed.
Steps 1, 2, 6 and 7 can be identified as gas transport resistances, steps 3 and 5 are referred to as shell layer resistance, step 4 as interface resistance.
These steps offer resistance in series of the overall reduction reaction. If one of these steps is considerably slower
than the others, it may be called the rate controlling step.
The rate of reduction is given by dR/dt (12R)2/3 for interface reaction control, by dR/dt (12R)1/3/{12(12R)1/3}
for pore diffusion control and by dR/dt const. for gas
transport control.27) In Fig. 13 the calculated rate of reduction versus fractional reduction is shown for either gas
transport control, interface reaction control and pore diffusion control. If one compares the calculated rates of reduction (Fig. 13) with the ones obtained by experiment (Figs.
812) the following conclusions can be drawn: The first
phase of reduction is characterized by constant rates of reduction (after an initial peak from a fractional reduction of
0 to approx. 0.15), i.e. for most of the reduction tests, during the first phase, the rate of reduction is controlled by the
transport of the reduction gas from the bubble phase of the
fluidized bed to the iron ore particle and the transport of
product gas from the particle to the bubble phase. During
the second phase of reduction the rate of reduction seemed
to be controlled by interface reaction control or by pore diffusion, as the curves obtained by the reduction tests are
similar to curves B and C in Fig. 13.

3.6. Microscope Analysis


To get detailed knowledge about the structural changes
during reduction of iron ore particles and to verify the assumptions stated above scanning electron micrographs
(SEM) were made of cuts through partly reduced particles.
Cuts through an iron ore particle reduced by H2 for 10 min.
at 1 053 K showed that a high amount of pores is spread
through the whole particle (Fig. 14). The particle is subdivided into small grains (210 m m in diameter) surrounded
by the pores. All grains in the particles consist of a bright
shell on the outer surface (very likely metallic iron) and a
dark core (iron oxide). As this structure of the grains is the
same throughout the particle it is evident that the removal
of oxygen takes place throughout the particle and the reduction is not influenced by the pores.
A detailed view at the structure of the grains is shown in
Fig. 15: The thickness of the shell is 0.21.0 m m, that leads
to a total removal of oxygen from the smallest grains. To
verify the assumption that the shell consists of metallic iron
and the core of iron oxide the distribution of selected elements (Fe, O) was determined. It can be seen in Fig. 16 that
the concentration of iron is the same throughout the grains
and that the concentration of oxygen is considerable higher
in the core of the grains. This shows that the shell consists
of metallic iron while the iron oxide is present in the core.
Since no micro-pores through this shell can be detected, a
direct access of the reducing gas to the iron oxide via the
pores is not possible. Hence the removal of the oxygen
from the core of the grains is carried out by solid state diffusion and the rate of reduction in the second phase is controlled by this solid state diffusion.

Fig. 14. Scanning electron micrograph of a cut through a single


iron ore particle (0.1250.5 mm) after 10 min of reduction by 55 vol% H2 at 1 053 K.

Fig. 15. Scanning electron micrograph of a cut through an iron


ore particle partly reduced by H2 at 1 053 K. Bright
areas at the outer surface of the single grains indicate
metallic iron.

Fig. 16. Distribution of oxygen (middle) and iron (right) in single grains (Fig. 15) surrounded by black pores. Dark-grey
areas mark higher concentration of selected elements.

941

2000 ISIJ

ISIJ International, Vol. 40 (2000), No. 10

4.

REFERENCES

Conclusions
1)

Fine iron ore was reduced in a wide range of fluidized


bed conditions. The rate of reduction was measured by
means of FT-IR spectroscopy and the following conclusions
could be drawn:
(1) Exhaust gas analysis by FT-IR spectroscopy is a
very useful technique for rate measurements of fluidized
bed iron ore reduction because all important oxygen containing gaseous species (H2O, CO, CO2) can be analyzed.
Another advantage of FT-IR spectroscopy is the high time
resolution, that allows sampling every 8 seconds (important
at the first, very fast phase of reduction).
(2) It was shown by the analysis of the exhaust gas that
the dilution of the iron ore sample with the silica sand
(90% silica sand and 10% iron ore) in the reactor offers the
possibility to measure the rate of reduction not affected by
the supply of H2 in the reducing gas.
(3) The rate of reduction increased with higher temperatures and higher molar flow of reducing gases (H2 and
CO), but was not affected by the absolute pressure in the reactor and small quantities of CH4 in the reducing gas (6
vol%). Particles with a size of 0.54.0 mm showed significantly lower rates of reduction than particles with a size between 0.125 and 0.5 mm, but reached the same fractional
reduction at the end of the reduction.
(4) Highly reduced particles are subdivided into small
grains (210 m m in diameter) surrounded by pores. All
these grains consist of a shell of metallic iron on the outer
surface and a core of iron oxide.
(5) The reduction of iron ore particles proceeds
through two phases with highly different rates of reduction:
In the first, fast phase of reduction the rate is controlled by
the transport of the reducing gas from the bubble phase of
the fluidized bed to the iron ore particle and the transport of
product gas from the particle to the bubble phase. During
the second, slow phase the rate of reduction is controlled by
solid state diffusion in the small grains of the iron ore particles.

2)
3)

4)

5)
6)
7)
8)
9)
10)
11)
12)
13)
14)
15)
16)
17)
18)
19)
20)
21)
22)
23)
24)
25)
26)
27)

Acknowledgements

The authors wish to express their thanks to Mr. W.


Steinbach, Institute of Chemical Engineering, for experimental assistance and to the Christian Doppler Forschungsgesellschaft for the financial support.

2000 ISIJ

942

K.-H. Schubert, H. B. Lngen and R. Steffen: Stahl Eisen, 116


(1996), 71.
S. Zeller, G. Deimek, K. Milionis and L. Gould: La Revue de
Metallurgie-CIT, 95 (1998), 426.
J. L. Schenk, W. L. Kepplinger, F. Wallner, H. G. Kim, S. Joo and I.
O. Lee: Proc. of 2nd ICSTI / 57th Ironmaking Conf., ISS,
Warrendale, PA, (1998), 1556.
S. Joo, H. G. Kim, I. O. Lee, J. L. Schenk, U. R. Gennari and F.
Hauzenberger: Proc. of SCANMET I, 1st Int. Conf. on Process
Development in Iron and Steelmaking, MEFOS, Lulea, Sweden
(1999), 293.
S. Y. M. Ezz: Trans. Metall. Soc. AIME, 218 (1960), 709.
J. Feinman and T. D. Drexler: AIChE J, 7 (1961), 584.
W. Wenzel, F. R. Block and E. Wortberg: Arch. Eisenhttenwes., 43
(1972), 805.
J. S. Sheasby and J. F. Gransden: Can. Metall. Q., 13 (1974), 479.
L. Krol and W. Zymla: Arch. Eisenhttenwes., 49 (1978), 463.
C.-H. Koo and J. W. Evans: Trans. Iron Steel Inst. Jpn., 19 (1979),
95.
N. Inoue, Y. Nakano, M. Ishida and T. Shirai: Int. Chem. Eng, 19
(1979), 484.
K. Sakuraya, K. Kamiya and M. Tanaka: Trans. Iron Steel lnst. Jpn.,
20 (1980), 92.
M. V. Srinivasan and J. S. Sheasby: Metall. Trans. B, 12 (1981), 177.
H. P. Meissner and F. C. Schora: Trans. Metall. Soc. AIME, 221
(1961), 1221.
K. M. Hutchings, J. D. Smith, S. Yrk and R. J. Hawkins:
Ironmaking Steelmaking, 14 (1987), 103.
D. Neuschtz and T. Hoster: Steel Res., 60 (1989), 113.
A. Jess and H. Depner: Steel Res., 69 (1998), 77.
S. Hayashi and Y. Iguchi: ISIJ Int., 38 (1998), 1053.
W. D. Ahner and J. Feinman: AIChE J, 10 (1964), 653.
C. S. Kim: Diss. Abstr. Int. B, 31 (1970), 1248.
T. Ashie, Y. Shinohara, R. Watanabe, M. Onoda and K. Mori: Trans.
Iron Steel lnst. Jpn., 25 (1985), B169.
K. Sato, Y. Ueda, Y. Nishikawa and T. Goto: Trans. Iron Steel Inst.
Jpn., 26 (1986), 697.
L. R. Glicksman: Chem. Eng. Sci., 39 (1984), 1373.
J. S. Sheasby and J. F. Gransden: Can. Metall. Q., 13 (1974), 479.
E. Kawasaki, J. Sanscrainte and T. J. Walsh: AIChE J, 8 (1962), 48.
W. M. McKewan: Trans. Metall. Soc. AIME., 218 (1960), 2.
L. von Bogdandy and H.-J. Engell: Die Reduktion der Eisenerze,
Springer Verlag, Berlin/Verlag Stahleisen, Dsseldorf, (1967), 156.

Das könnte Ihnen auch gefallen