Sei sulla pagina 1di 11

A Kinetic Model of the PeirceSmith Converter: Part I.

Model Formulation and Validation


A.K. KYLLO and G.G. RICHARDS A kinetics-based mathematical model of the PeirceSmith converter has been developed. The model considers mass transfer, heat transfer, and reactions between each of the phases present in the converter. Model validation is carried out using industrial data obtained from both copper and nickel converters. The model is generally able to predict the temperature and compositional variations of the converters to within the errors of the industrial data. However, the interactions between the white metal and slag during the copper blow are not understood sufciently to model well.

I.

INTRODUCTION

AS part of an overall project aimed at improving the understanding of the PeirceSmith converter operation, a mathematical model has been developed. The model was designed to follow the converting process using a combination of kinetic and thermodynamic calculations. Previous work has shown that the oxygen efciency of the converter is signicantly less than 100 pct and varies throughout the operation,[1] so the inclusion of kinetics is required. The model consists of two distinct sections relating to the overall heat and mass balance calculations and the gas ow through the bath. This article considers the development of the heat and mass balance section of the model and the validation of the overall model. The gas ow model is dealt with elsewhere.[2]
II. PREVIOUS WORK

A. Converter Modeling Most models relating to the copper converter deal with the distribution of impurities between matte, metal, and slag. Three models attempt to reproduce the overall material balances and, of these, only one attempts to reproduce the heat balance as well. A recent model of the nickel converter reproduces both the material and heat balances. All of these rely on the assumption that the converter is in thermodynamic equilibrium. A model of the copper converter, developed by Goto, has been gradually improved over the last 15 years, most recently being applied to the copper ash smelter.[37] The base model assumed that the converter is in thermodynamic equilibrium. This allowed the bath composition to be calculated from a set of simultaneous equations based on the equilibrium equations and the mass balance.[3] A heat balance was added later to complete the representation of the entire converter.[4] It was claimed that the model was able

to predict temperature variations fairly well,[4] but no direct comparison with plant data was published, and no comparison of matte or slag composition was given. More recent developments of the model have included its extension by Shimpo et al. to cover the copper ash furnace[5] and the addition of a calculation of oxygen consumption using kinetic considerations.[6] A model of the heat and mass balances in the nickel converter has recently been produced by the authors.[8,9] It is based on the work of Goto and coworkers and has been found to be able to predict both the bath temperature and composition fairly accurately. A model of the Noranda continuous converting process has been developed by Nagamori and co-workers, which concentrates primarily on the minor element behavior.[1014] This model calculates the equilibrium phase compositions at a given temperature, partial pressure of sulfur dioxide, magnetite activity, and, in the case of matte making, matte grade. The most important ndings of these articles with respect to the present project relate to impurity removal, particularly removal to the gas phase. More recently this model has been extended to the ash furnace and converting with high oxygen enrichment and calcium ferrite slags by Sohn and co-workers.[1518] B. Process Kinetics 1. Copper converting kinetics The kinetics of copper converting have undergone little study. Calculations by Ashman et al.[19] have shown that the reactions in both the slag and copper blows are under gas-phase mass-transfer control. The high oxygen efciencies reported for the converter suggest, therefore, that the gas residence time in the bath is sufcient to allow the reactions to come close to or attain equilibrium. A recent study of the oxidation of molten iron sulde assumed that the rate was under mixed control through the gas and liquid boundary layers around a bubble.[20] A physical model of mass transfer in the copper converter has shown that kinetic calculations cannot be limited to the bubbles after detachment from the tuyere, as a signicant portion of the reaction occurs during bubble growth.[21] Unfortunately, this study neglected the effect of mass transfer at the bath surface, which can be considerable, especially at low tuyere submergences.[22] Other kinetic studies related to copper converting are of limited practical use. They either demonstrate that the process is controlled by gas phase mass transfer[23,24]

A.K. KYLLO, formerly Graduate Student, the University of British Columbia, is Research Fellow, G.K. Williams Cooperative Research Centre for Extractive Metallurgy, The University of Melbourne, Parkville, Victoria, Australia 3052. G.G. RICHARDS, formerly Associate Professor, University of British Columbia, Vancouver, BC, Canada, V6T 1Z4 is Senior Research Scientist, Cominco Metals, Trail, BC, Canada V1R 4L8. Manuscript submitted January 6, 1997.

METALLURGICAL AND MATERIALS TRANSACTIONS B

VOLUME 29B, FEBRUARY 1998239

mass-transfer coefcient with increasing gas ow, probably related to the increase in bath mixing velocity; (2) above a critical gas ow rate, there is a rapid increase in mass-transfer rates, due to the formation of droplets of slag in the metal; and (3) with further increases in gas ow, the increase in masstransfer rate is reduced, since the entire slag has been emulsied.[34,35,36] The values of the critical ow rates are dependent upon the slag depth, reaction dimensions, and tuyere position, as well as the physical properties of the phases involved.[36] As such, the values obtained in physical modeling studies are not transferable directly to the copper converter; however, the three mass-transfer regimes are likely to be found. Also, the observation that off-center tuyeres have a considerably greater critical velocity for emulsion formation than central tuyeres suggests that the conditions in a PeirceSmith converter are far from ideal for matte-slag mass transfer. This conclusion is supported by the observation of a large stagnant region of slag in a two-phase physical model with an off-center tuyere.[37] C. Process Thermodynamics 1. Matte thermodynamics Industrial mattes are usually a complex mixture of a number of different components. However, most of these are only present in small amounts, so mattes are generally considered as ternary Cu-Fe-S systems. The rst reported thermodynamic measurements on this system were carried out by Krivsky and Schuhmann in 1957.[38] They determined that the Cu2S-FeS psuedobinary deviated negatively from ideality. Later researchers determined that the Temkin[39] and Flood[40] models could be used to calculate activities within the psuedobinary system, but that the presence of oxygen invalidated their use.[40] Work carried out by Sinha and Nagamori determined the activity coefcients of cobalt, iron, and their suldes in copper-saturated mattes.[41] The authors also determined that the results of Krivsky and Schuhmann were a factor of 1.32 too high, due to uncertainties in the free energy data when the original experiments were made. In developing a mathematical model of the converter, Goto and co-workers[36] derived a set of equations for the activity coefcients of FeS, FeO, and Fe3O4 in the matte. The relationship obtained for the activity coefcient of FeS is plotted in Figure 1. The same gure shows the experimental points of Sinha and Nagamori[41] in copper-saturated matte. The agreement is very poor, demonstrating the extreme difculty that exists in characterizing the solution thermodynamics of mattes. 2. Slag thermodynamics There has been a considerable amount written regarding slag thermodynamics, and an extensive review was produced by Mackey in 1980.[42] More recent studies appear to have concentrated on ferrite slags, as are used in the Mitsubishi process.[4347] The activity coefcients for the major constituents in fayalite slags have also been calculated by Goto.[3] 3. Internal phase equilibrium The distribution of various elements between the matte and the slag at equilibrium is an important topic, since it
METALLURGICAL AND MATERIALS TRANSACTIONS B

Fig. 1Comparison of reported values of the FeS activity coefcient in copper mattes.[6,41]

or concern the controlling stage of the chemical reaction,[25,26] which is not relevant, since the reaction is considerably faster than mass transfer in the system. 2. Kinetics of gas/liquid reactions There have been some studies relating to the mass transfer around submerged gas jets.[21,22,2732] These were primarily low-temperature physical models, a majority dealing with the reaction of CO2 with an NaOH solution.[2830] Two studies have considered the kinetics of the deoxidation of copper by carbon monoxide.[31,32] The rst determined that the deoxidation rate was controlled by gas-phase mass transfer down to an oxygen concentration of 0.1 pct, below which it was under liquid-lm control.[31] The second study found that liquid-phase control remained in effect down to 0.005 pct oxygen, below which the reaction rate became signicant. A model of mass transfer between a horizontal jet and a liquid was used to analyze the experimental results from a physical model, in which a mixture of air and SO2 was injected into a bath containing H2O2.[27] The experimental work was designed to ensure that the process was limited by gas-phase mass transfer. Unfortunately, the lowest modied Froude number used in this study was 88, which is considerably higher than that found in the copper converter. Another study combining mathematical and physical modeling used the absorption of CO2 in water during vertical injection. In this case, the rate of absorption was controlled by liquid-phase mass transfer, which was represented in the model using Higbies penetration theory. One of the more interesting results of this study was the determination of the effect of surface reactions. In particular, it was found that at low ow rates and low tuyere submergences, surface reactions could account for well over 50 pct of the total CO2 absorbed.[22] 3. Kinetics of liquid/liquid reactions A number of studies have been carried out to determine the rates of mass transfer between two immiscible phases under gas-stirred conditions, and those prior to 1989 are reviewed by Mori.[33] All of these are studies related to bottom injection, particularly ladle rening, and have determined that the rate of mass transfer is dependent upon the gas ow rate. There are three mass-transfer regimes: (1) at low ow rates, there is a slow increase in the
240VOLUME 29B, FEBRUARY 1998

Fig. 2Schematic cross sections of a converter: (a) idle, (b) no gas ow through the slag, (c) gas ow through matte and slag, and (d ) gas ow through slag only.

relates both to the extent of valuable metal losses and of slagging of unwanted components of the matte. The importance of copper losses to the slag is evidenced by the relatively large body of literature written about the subject.[4855] The form of the copper in the slag appears to depend primarily upon the matte grade.[48] It would probably be more correct to consider the form of dissolution as a function of oxygen potential, with oxidic dissolution being favored by the high oxygen potentials that exist with high matte grades.[56] In fact, measurements made in ferrite slags show a large variation in S 0.5 with oxygen partial presCuO sure.[46] In order to deal with this, the Temkin or Herasymenko model has been applied to the slag.[55] The Herasymenko model has been shown to represent copper slags well, including the transition between suldic and oxidic dissolution of copper.

so it will be considered rst. The matte/slag interfacial area is given by


I A M/S 2 2 rc

(r

hM )

1/2

[1]

The slag/top gas interfacial area is given by a similar equation. Unfortunately, values for hM and hS are not directly calculable. Given the mass of matte, MM, and slag, MS, present in the converter, hM can be determined from the volume of matte present: MM
M

L rc

r2 c 2 hM 2

r 2 sin c

rc rc
2 hM)

hM [2]

(2 r

hM

1/2

III.

CONVERTER GEOMETRY

For mass- and heat-transfer calculations within the converter, interfacial areas between phases and solid/liquid contact areas are required. The horizontal cylinder form of the converter adds some complication to this. Figure 2 shows idealized schematic cross sections of a converter, both idle and blowing. The idle case (Figure 2 (a)) is the easiest to deal with,
METALLURGICAL AND MATERIALS TRANSACTIONS B

and the value of hS is calculated in a similar manner, using both matte and slag volumes. When air is being blown into the converter a spout is formed. The possible variations in spout form are shown schematically in Figures 2(b) through (d). Other motion of the interfaces is not shown in this gure. If the spout height is greater than the slag thickness (Figure 2(b)), only the slag thickness and matte/slag interfacial area are affected. The value of hS must be modied to account for the spout volume. If the modied slag thickness is greater than the spout height, then the conguration conforms to Figure 2(c). To calculate the slag/matte interfacial area, the extra area due to the spouts must be inVOLUME 29B, FEBRUARY 1998241

Table I. Reactions and Corresponding Free Energy[58] Required for the Model Equilibrium and Interphase MassTransfer Calculations Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 2Cu Ni 2Ni Fe Co Pb Zn 3FeO 1 /2 S2 Cu2S
1 1 1

IV.

MASS BALANCE

A. Equilibrium within Phases It is assumed for the calculations that each phase is in internal equilibrium. As such, each must be considered separately in the model. Table I lists the reactions considered in each phase with the standard free energy for the reaction. 1. Blister copper The copper phase is the easiest to deal with, since it can be assumed to only contain components in their elemental forms. Thus, there are no reactions and the composition will be governed only by mass-transfer considerations. 2. Matte The matte is considerably more complex, since it contains a mixture of suldes, oxides, and elements. The equilibrium composition within the matte will be governed by the sulfur potential. If it is assumed that all oxygen is present as magnetite, and all FeO produced by reaction is carried directly to the slag by the rising bubbles, then no reactions involve oxygen and an oxygen potential is not required. All components not included in Table I are considered nonreacting, so they are not subject to equilibrium calculations. The equilibrium calculations are combined with mass balances for each element to give the nal composition. 3. White metal The white metal is a special case of the matte phase, in which Cu2S is the only remaining sulde component. As such, its sulfur potential will be governed by Reaction 1 in Table I. 4. Slag While the slag is a complex mixture of oxides, it can be assumed that iron is the only component that has more than one oxide form present. As such, the oxygen potential of the slag will be controlled by Reaction 8 in Table I, and the composition is determined by the iron and oxygen mass balances. 5. Gas Like the slag, the equilibrium composition within the gas is dominated by a single reaction (9 in Table I). Combining this with sulfur and oxygen mass balances and setting the total pressure to atmospheric allows the composition to be calculated. B. Calculation Technique For each of the phases, except for the blister copper, there is a system of simultaneous equations that must be solved. In some cases, these will be solved directly. In the case of the matte, the system is more complex and requires the use of a nonlinear equation-solving routine. As in a previous work, a QuasiNewton technique is used.[59,60] C. Gas-Liquid Mass Transfer Over the majority of a cycle, the gas within the bath will only be in contact with the matte. However, reactions with the slag may also take place within the spout and on the surface of the bath. Within the matte, the rate of the gasmatte reaction is controlled by gas-phase mass transfer. This suggests that the reaction will be
METALLURGICAL AND MATERIALS TRANSACTIONS B

Reaction /2 S2 Cu2S /2 S2 NiS 1 /2 S2 Ni2S 1 /2 S2 FeS 1 /2 S2 CoS 1 /2 S2 PbS 1 /2 S2 ZnS 1 /2 O2 Fe3O4 O2 SO2 1 /2 O2 Cu2O /2 S2
1 1

G T (J mol 1) 147,100 129,300 152,500 115,300 125,000 138,400 264,800 402,000 363,000
1 1

42.05T 54.06T 46.06T 30.63T 48.12T 64.14T 98.20T 169.77T 72.42T 2.59T 41.67T 21.13T 22.61T 21.38T 3.35T

NiS FeS CoS PbS ZnS

/2 O2 /2 O2 1 /2 O2 1 /2 O2 1 /2 O2

NiO FeO CoO PbO ZnO

/2 S2 /2 S2 1 /2 S2 1 /2 S2 1 /2 S2

28,360 122,800 128,200 107,800 57,950 80,330

cluded. Unfortunately, the surface area of an ellipsoid is calculable only for a few special cases, so an approximation must be used. Therefore, it is assumed that the spout has a surface area equal to a regular ellipsoid, with minor axes equal to the average of the maximum bubble width and the spout height and eccentricity , so A
I M/S

2 (r nt

2 c

(rc
a2

hM) )

2 1/2

[3] sin
1

L where

(ab
b b

ab

)
[4]

hM sp 2

If the matte spout height is less than the slag thickness (Figure 2(c)), the slag will form part of the spout. The slag spout height can be calculated in a similar manner to the hM) as the matte spout height, using the value (hS hM sp submergence. A third possibility is shown in Figure 2(d). In this case, the gas is blowing directly into the slag, which may occur near the end of a blow when there is a relatively large slag volume present. Under these conditions, the matte/slag interfacial area will be given by Eq. [1] and the slag/gas interfacial area will be the same as the previous case. When blister copper is present, it is also possible that the gas will pass through it as well. In this case, the previous equations must be further modied. The values of interfacial areas calculated in this fashion will be an underestimate in most cases. Generally, there will be a certain amount of wave motion, and, in the extreme case, slopping, when the bath surface can be approximated by a standing wave. In this case, it is likely that the extent of interphase mixing will be increased considerably. To obtain a better estimate of the matte/slag interfacial area during blowing, a theoretical model of emulsion formation at the edge of the spout is used.[57] This model uses a force balance approach to determine the radius of the slag droplets formed, and an energy balance is carried out to determine the rate of droplet generation.
242VOLUME 29B, FEBRUARY 1998

FeS

3/2O2

FeO

SO2

[5]

because the partial pressure of oxygen at the bubble surface will be very low, favoring the least oxidized iron compound. The method for calculating the gas/liquid interfacial area and the gas-phase mass transfer is detailed elsewhere.[2] D. Mass Transfer between Liquid Phases Within any phase, I, the mass-transfer rate of component i to an interface is given by ni A k Ii
M I

passing between phases is calculated from the heat content of the material and its mass transfer rate. VI. DATA

(X

I i

X Ii * )

[6]

In a model of this sort, a wide range of data is required. Unfortunately, much of these data are unavailable in the literature or are only present for a narrow range of temperatures and/or compositions. Therefore, it is necessary to assume that whatever data are available are valid over the entire range of conditions present in the model and, if nothing is available, required values must be assumed or tted. A. Physical and Thermal Properties Data concerning the physical and thermal properties are required for the three condensed phases. Most of these properties vary considerably with composition and temperature, but it is unusual to nd a study that considers both together. The required properties and the equations used to calculate them are given in Table II. The equation for the slag viscosity is derived directly from measured viscosities,[64] while the equations for matte and blister copper viscosities are estimated from measured diffusivities. B. Thermodynamic Data 1. Free energy and enthalpy of reaction To calculate the equilibrium composition of the matte, slag, and gas phases, and the equilibrium constants used to determine the rates of interphase mass transfer, the free energies of the reactions given in Table I are used. Values required to calculate the heat produced by the reactions are given in Table III. 2. Activity coefcients The equilibrium calculations within each phase and the mass-transfer calculations between phases require values for activity coefcients. Equations for calculating these for most of the species present in copper and nickel converters can be found from the literature, and those for the major components are given in Table IV. However, as noted in Section II, the accuracy of some of these equations is limited, particularly by the conditions under which they were measured. While more extensive data are available, there are no simple correlations that can be readily applied to the present model. C. Kinetic Data 1. Diffusivities The diffusivity of each species in each phase is required for the mass-transfer calculations. Unfortunately, very little data are available for the system under consideration. Therefore, a theoretical approach must be used. Assuming spherical particles following Stokes law in a viscous liquid, the NernstEinstein relation for the diffusivity is[75] D kT 6 r [10]

It is important to note that the interfacial concentration of a component will be different on either side of the interface. At steady state, the rate of mass transfer to the interface will be equal to the rate of mass transfer away from the interface in the other phase. In general, at an interface between two liquid phases, I and J, there is a relationship (assuming instantaneous chemical kinetics) X I* i K (X J* ) i
n

[7]

where K is a constant determined by equilibrium considerations. In the second phase, J, the mass-transfer rate of i away from the interface is ni A kJ i
M J

(X

J* i

XJ) i

[8]

Combining Eqs. [9] through [11] and rearranging, assuming n 1, gives ni A k Ii 1 1


m I

1 k iJ
m J

(X

I i

KX J i

[9]

This is a general equation that can be used for mass transfer between any two liquid phases, including cases where there is a reaction. If n is not equal to one, the equation becomes more complex. However, this does not occur in the present case. V. HEAT BALANCE

Without the assumption of thermal equilibrium, each of the component phases within the converter may be at different temperature. While this adds to the complexity of the model, the temperature variation between phases may be signicant. Heat transfer from the bath to the gas is calculated as part of the overall gas ow model, so it does not need to be considered further. However, the cooling effect of the gas on the bath is important. The basic converter heat balance model, without interphase heat transfer, has been described elsewhere.[9] In the present model, interphase heat transfer is assumed to be due to convection and mass transfer. The convective heat transfer is calculated based on the temperature difference, interfacial area, and a heat-transfer coefcient calculated from the thermal conductivities of the phases. The heat transferred with materials
METALLURGICAL AND MATERIALS TRANSACTIONS B

where r is the atomic/molecular radius of the particle and k is the Boltzman constant. This equation requires atomic or molecular radii of the diffusing species, which poses another problem: while the
VOLUME 29B, FEBRUARY 1998243

Table II. Phase Slag Matte Blister copper

Physical and Thermal Properties of the Condensed Phases Property density (kg m 3) 0.115TS [SiO2]S 3297 128.5 [FeO]S 3 3880 404 [Cu]M 4590 [Cu] 2 3750 [Cu] M M 7800 viscosity (kg m 2.06 (10 3) exp 9380 TS
1

Reference 61 62 63

s 1) [Fe]S [SiO2]S

Slag Matte Blister copper

0.714 5000 TM 4866 TB

64 63 65

3.36 (10 4) exp 7.69 (10 5) exp

surface tension (N m 1) Slag XCu2S Matte Blister copper XFeS XCu2S 0.7148 2.271 5 3.17 (10 4) (TS 273) 1.188XNiS 2XNiS 61

1.673XFeS 3XFeS 1.136

XFeS

XFeS 1.149 XCu2S 3


5

66 67

1.6 (10 )TB


1

thermal conductivity (W m Slag Matte Blister copper 2.09 25.6 [Cu] 2 M 134

K 1) 42 68 68

13.4

17.9 [Cu]M

13.7 [Cu] 3 M

Table III. Phase Blister copper White metal Matte Slag

Enthalpies of Reaction Required for Energy Generation Calculation[69] Reaction 2Cu /2 O2 Cu2O Cu2S O2 2Cu SO2 FeS 1.5O2 FeO SO2 1 3FeO /2 O2 Fe3O4
1

HR (kJ mol 1) 166.7 190.4 463.4 317.6

s 1 is calculated for Fe2+ diffusion in a slag with an ironto-silica ratio of 0.25 at 1400 K, while a value of 1.13(10 9) m2 s 1 is obtained for Fe2+ diffusion in matte at the same temperature. These are lower than reported values, which range between 5(10 11) m2 s 1 and 5(10 8) m2 s 1 in slag[76 79] and between 2.9(10 8) m2 s 1 and 1.4(10 8) m2 s 1 in matte.[80]

minor elements and blister copper constituents are present in elemental form, the other species are present as compounds. The radius values used to calculate the diffusivity in the matte and slag will depend on the nature of the solution. If the solution is ionic, then the ionic radii of the diffusing elements must be used, whereas if the solution is covalent, then the molecular radius of the diffusing species should be used. To determine which case should be used, the approximate degree of ionic bonding can be found using the Pauling electronegativity scale. This suggests that the bonding in mattes is less than 12 pct ionic, while that in slags is closer to 50 pct.[69] It is important to note that these numbers are for solid compounds, and there may be some differences for liquids. The relatively low amount of ionic bonding in mattes suggests that the diffusing species will be covalent molecules, while in slags, diffusion of ions is likely to predominate. Values of diffusivity calculated using Eq. [10] for Fe2+ diffusion in slag and matte can be compared with measured values. A diffusivity of 1.4(10 11) m2
244VOLUME 29B, FEBRUARY 1998

VII.

MODEL VALIDATION

A. Copper Converter Validation The validity of the model can be tested by comparing predicted temperatures and compositions with those measured in the plant trials. The temperature variations for the three charges are given in Figure 3. Although there are only a small number of measured temperatures, it can be seen that the model predictions are reasonable in the slag blows. The matte and slag temperatures predicted in the copper blows are generally low, but the blister copper temperature ts the measured temperatures well. A comparison of the measured and predicted slag iron and silica contents is given in Figure 4. Considering the uncertainty in the weights of materials added, the iron contents of the slags are predicted well. However, the tendency to underpredict the silica content and overpredict the iron content suggests that more ux was added to the converter than was estimated during operation. Figure 5 shows a
METALLURGICAL AND MATERIALS TRANSACTIONS B

Table IV. Phase Slag i FeO Fe3O4 NiO Cu2O CoO Matte Fe Ni Cu Co FeS Ni2S

Activity Coefcients of the Major Constituents of the Matte and Slag


i

Reference 0.2) 6 3 70 71 72 49 59 3 49 0.52XFeS) 6

exp

1543 log (1.42 XFeO TS 56.8 XFe3O4 3980 TS 9 0.66 40 15 14 25

0.69 exp

5.45 XSiO2 1.62

exp

1458 ln (0.54 TM 10

1.4XFeS log XFeS


1840 0.6

TM

74 59 3 41
2

NiS Cu2S CoS FeO Fe3O4 exp exp

1 1 0.4

1573 TM

(5.1 (4.96

6.2 log XCu2S 9.9 log XCu2S

6.41 (log XCu2S)

2.8 (log XCu2S)

)
3

}
)

6 6

1573 TM

7.43 (log XCu2S)

2.55 (log XCu2S)

comparison of the measured and predicted iron and copper contents of all sampled mattes. The gure shows that the predicted iron and copper in the matte are close to the measured values, particularly at high matte grades. The one sample of metal taken during the copper blow of charge 586 has a much higher copper content and correspondingly lower sulfur content than is predicted by the model; this sample is not included in the gure. This is most likely explained by incomplete separation of blister copper from the matte. B. Nickel Converter Validation As with the copper converter, there are phase composition and temperature data available for the nickel converter, which may be compared to the model predictions. In this case, however, there are considerably more temperature data available, but the compositions are less accurate. A comparison of the measured and predicted bath temperatures is given in Figure 6. The measured temperatures were obtained using a two-color pyrometer mounted in the hood and aimed at the bath surface, so in most cases, this should give the slag temperature. However, the gures contain both the predicted matte and slag temperatures, because, at the beginning of a blow, there is usually only a thin layer of slag present; so, depending on where the pyrometer is aimed, the measured temperature may be matte rather than slag. It can be seen from the gure that the temperature t is
METALLURGICAL AND MATERIALS TRANSACTIONS B

quite good, especially considering the lack of accurate weights of the materials charged to and skimmed from the converter. In most cases, the largest disparity between the measured and predicted temperatures is at the beginning of the charge. This is most likely caused by the unknown quantity of mush remaining in the converter from the previous charge. Generally, the mush is a very high silica slag with a large amount of entrained Bessemer matte. It usually provides all the silica for the rst blow of the following charge, but its composition, weight, and temperature are all unknown. In a number of cases, the matte temperature is close at the beginning of the blow and the slag temperature is close at the end of the blow. This suggests that at some point during the blow, the slag thickness becomes sufcient to completely cover the bath surface, including the spout. There is often an abrupt change in the slope of the measured temperatures, and this coincides with the change from matte temperature to slag temperature. This is particularly evident at the beginning of charge 106 and in the second blow of charge 108. A comparison of measured and predicted iron and silica contents of the slags is given in Figure 7. The silica in the gure includes the other inert materials, such as alumina and lime. In all charges, the silica values predicted are close to those measured, as are the iron contents up until the miss blow, the penultimate blow, before which furnace matte is not added. After this point, the predicted iron content is considerably lower than the measured value and the nickel and cobalt contents are considerably higher. There
VOLUME 29B, FEBRUARY 1998245

(a)
Fig. 4Comparison of measured and predicted iron and silica contents in copper converter slags, charges 586, 588, and 595.

(b) (a)

(c)
Fig. 3Comparison of measured and predicted (lines) copper converter temperatures: (a) charge 586, (b) charge 588, and (c) charge 595.

(b)
Fig. 5Comparison of measured and predicted (a) iron and (b) copper contents in copper converter mattes, charges 586, 588, and 595, Feb. 1994.

are a number of possible reasons for this, including insufcient iron being added to the model, the model predicting more reaction than is actually occurring, and the value of liquid-phase diffusivity of iron being too low. The model generally underpredicts the amount of copper, nickel, and cobalt in the early slags. This discrepancy, for the most part, can be explained by the entrainment of matte in the
246VOLUME 29B, FEBRUARY 1998

slag, which is not accounted for in the model. There is a relatively large amount of sulfur in the early slags, which indicates a correspondingly large amount of entrainment. A comparison of measured and predicted iron, nickel, and copper contents of all matte samples is given in Figure 8.
METALLURGICAL AND MATERIALS TRANSACTIONS B

(a)

Fig. 7Comparison of measured and predicted iron and silica contents in nickel converter slags, #3 converter, charges 105, 106, 107, and 108, May 1988.

(b)
Fig. 6Comparison of model predicted matte and slag temperatures with nickel converter plant data: (a) charge 106 and (b) charge 108. Fig. 8Comparison of measured and predicted iron, nickel, and copper contents in nickel converter mattes, #3 converter, charges 105, 106, 107, and 108, May 1988.

The gure indicates that the model predictions are relatively close for these elements. C. Discussion 1. Phase compositions The overall compositions of the condensed phases are generally predicted well for both the copper and nickel converters, with some errors in the minor element compositions. There are a number of possible sources for these errors, including problems with the activity coefcients, diffusivities, interfacial areas, and assayed compositions. 2. Sulfur in slag Neglecting the presence of sulfur in the slag will have some effect on the predicted slag compositions. The sulfur content of the slags will be mainly due to matte entrainment, because the slags were skimmed while in contact with a high grade matte and so should have a dissolved sulfur content under 0.9 pct.[51] There are two main mechanisms by which matte entrainment may occur; emulsication of matte in slag and splashing from the spout. The former mechanism is not usually reported in physical modeling studies, but has been reported in one experimental system,[81] and is found in top blowing and combined blowing systems.[82,83] This suggests that the majority of enMETALLURGICAL AND MATERIALS TRANSACTIONS B

trained matte is either carried out of the spout by its own momentum or ejected from the spout by collapsing bubbles.[84] The extent of this will vary considerably with the injection conditions, and the residence time of the matte in the slag will also depend on the slag viscosity. Thus, a single value of a suspension index, as used by Nagamori and Mackey for the Noranda reactor,[10] will not be valid for this system. While it may be possible to produce a formula for suspension indexes to give a better t between the predicted and assayed slag compositions, it would not be particularly meaningful considering the other inaccuracies involved in the plant trials. 3. Oxygen in matte For the modeling, it was assumed that all of the oxygen in the matte was in the form of Fe3O4; however, it is probable that there is also some FeO present. The equilibrium mole fraction of FeO in matte may be obtained from Reaction 8 in Table I. The oxygen potential of the matte may be obtained from the magnetite formation reaction, and the activity coefcient of FeO in matte is given in Table IV. Using these equations, it can be calculated that while the mole fraction of FeO is as high as 50 pct of the magnetite mole fraction, the relative proportions of iron and oxygen present as FeO will be considerably less. The effect of inVOLUME 29B, FEBRUARY 1998247

cluding FeO will be to increase the iron content in the matte slightly. 4. Copper blow Overall, the model is able to predict the bath temperature reasonably well for both the copper and nickel converters, although the measured matte and slag temperatures during the copper blow are suspect. Toward the end of the copper blow, it is probable that the small amounts of matte and slag remaining are well mixed, increasing the extent of matte-slag reaction. Since the slag at this stage is highly oxidized, the increased reaction will reduce the sulfur content of the matte signicantly. A thermodynamic analysis indicates that Cu2O is not stable at converter operating temperatures, with the oxidized copper being in a copper ferrite form. In fact, the formation of the ferrite considerably reduces the range of stability of copper metal at lower temperatures. The composition of the copper slags indicates that their oxygen potential is higher than 0.05 atm, so the product of the matte-slag reaction is most likely to be Cu2O Fe2O3. This suggests that the main reaction between the matte and the slag at this stage is likely to be Cu2S 4Fe3O4 Cu2O Fe2O3 10FeO SO2 [11]

b D h K k K L M n n r T X Greek

This reaction will reduce both the oxygen and sulfur potentials of the mixture and will continue until one of these values drops below a level at which the copper ferrite is no longer stable. Whether any reaction occurs beyond this point will depend on the sulfur potential of the system. The presence of the ferromagnetic copper ferrite also explains the high values measured for magnetite content in the copper slags, since there is insufcient iron present to form the amount of magnetite measured. Its presence has also been found in accretions formed during the copper blow of a test converter.[85] The formation of a matte-slag emulsion will also have a stabilizing effect on the temperature of the two phases. During the majority of the copper blow, the predicted slag temperature is very low due to a combination of the radiation losses and absence of a reaction with the gas. In the actual converter, it is probable that the radiation losses would be from the emulsion rather than just the slag. Also, the presence of the slag within the matte would allow some gas-slag contact and, hence, reaction. The increased matte-slag reaction will also produce extra heat, which is not included in the model. VIII. SUMMARY

major axis of ellipse (m) diffusivity (m2 s 1) height (m) constant mass-transfer coefcient (m s 1) Boltzman constant converter length (m) mass (kg) number (moles) molar ux (mol s 1) radius (m) temperature (K) mole fraction Letters activity coefcient ellipse eccentricity dynamic viscosity (kg m 1 s 1) density (kg m 3) m molar density (mol m 3) Subscripts c converter I,J phase i species, initial, interfacial M matte phase R reaction, refractory S slag phase sp spout t tuyere Superscripts I* interfacial 1 I,J phase M matte REFERENCES

A mathematical model of the PeirceSmith converter has been developed, which considers both kinetics and thermodynamics. The model considers each phase separately and includes interphase heat and mass transfer between the phases. A validation of the model using data obtained in industrial trials[1] indicates that it is generally able to predict the bath composition and temperature within the accuracy of the plant data. In Part II of this article,[86] a sensitivity analysis of the model will be discussed, and the operation of the converter will be analyzed. NOMENCLATURE area (m2) minor axis of ellipse (m)

A a

1. A.K. Kyllo and G.G. Richards: The University of British Columbia, Vancouver, unpublished research, 1994. 2. A.K. Kyllo: Ph.D. Thesis, The University of British Columbia, Vancouver, 1994. 3. S. Goto: Copper-MetallurgyPractice and Theory, M.J. Jones, ed., IMM, London, 1975, pp. 23-34. 4. S. Goto: Copper and Nickel Converters, R.E. Johnson, ed., TMSAIME, Warrendale, PA, 1979, pp. 33-54. 5. R. Shimpo, S. Watanabe, S. Goto, and O. Ogawa: Advances in Sulphide Smelting, TMS-AIME, Warrendale, PA, 1983, vol. 1, pp. 295-316. 6. N. Kemori, T. Kimura, Y. Mori, and S. Goto: Pyrometallurgy 87, IMM, London, 1987, pp. 647-66. 7. M. Bustos and M. Sanchez: Pyrometallurgy of Copper, C. Diaz, C. Landolt, and A. Luraschi, eds., University of Chile, Santiago, 1987, pp. 473-87. 8. A. Kyllo, G.G. Richards, and S.W. Marcuson: Metall. Trans. B, 1992, vol. 23B, pp. 573-82. 9. A. Kyllo and G.G. Richards: Metall. Trans. B, 1991, vol. 22B, pp. 153-61. 10. M. Nagamori and P.J. Mackey: Metall. Trans. B, 1978, vol. 9B, pp. 255-65. 11. M. Nagamori and P.J. Mackey: Metall. Trans. B, 1978, vol. 9B, pp. 567-79. 12. M. Nagamori and P.C. Chaubal: Metall. Trans. B, 1982, vol. 13B, pp. 319-29. 13. M. Nagamori and P.C. Chaubal: Metall. Trans. B, 1982, vol. 13B, pp. 331-38. 14. P.C. Chaubal and M. Nagamori: Metall. Trans. B, 1982, vol. 13B, pp. 339-48. 15. H.G. Kim and H.Y. Sohn: in Pyrometallurgy of Copper, C. Diaz, C.

248VOLUME 29B, FEBRUARY 1998

METALLURGICAL AND MATERIALS TRANSACTIONS B

16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50.

Landolt, A. Luraschi, and C.J. Newman, eds., Pergamon Press, New York, NY, 1991, pp. 617-36. P.C. Chaubal, H.Y. Sohn, D.B. George, and L.K. Bailey: Metall. Trans. B, 1989, vol. 20B, pp. 39-51. K.W. Seo and H.Y. Sohn: Metall. Trans. B, 1991, vol. 22B, pp. 79199. H.G. Kim and H.Y. Sohn: EPD Congr. 91, D.R. Gaskell, ed., TMS, Warrendale, PA, 1991, pp. 437-67. D.W. Ashman, J.W. McKelliget, and J.K. Brimacombe: Can. Metall. Q., 1981, vol. 20, pp. 387-95. Y. Fukunaka, T. Nishihara, Z. Asaki, and Y. Kondo: Todays Technology for the Mining and Metallurgical Industries, MMIJ/IMM, London, 1989, pp. 579-88. E. Adjei: Masters Thesis, University of British Columbia, Vancouver, 1989. S. Taniguchi, A. Kikuchi, H. Matsuzaki, and N. Bessho: Trans. Iron Steel Inst. Jpn. 1988, vol. 28, pp. 262-70. F. Ajersch and J.M. Toguri: Metall. Trans., 1972, vol. 3, pp. 218793. Z. Asaki, S. Ando, and Y. Kondo: Metall. Trans. B, 1988, vol. 19B, pp. 47-52. J.J. Byerley, G.L. Rempel, and N. Takebe: Metall. Trans., 1974, vol. 5, pp. 2501-06. P.T. Morland, S.P. Matthew, and P.C. Hayes: Metall. Trans. B, 1991, vol. 22B, pp. 211-17. J.K. Brimacombe, E.S. Stratigakos, and P. Tarasoff: Metall. Trans., 1974, vol. 5, pp. 763-71. S. Inada and T. Watanabe: Trans. Iron Steel Inst. Jpn., 1977, vol. 17, pp. 21-27. R.J. Fruehan and L.J. Martonik: 3rd Int. Iron and Steel Congr., ASM, Metals Park, OH, 1978, pp. 229-38. T. Stapurewicz and N.J. Themelis: Can. Metall. Q., 1987, vol. 26, pp. 123-28. N.J. Themelis and P.R. Schmidt: Trans. AIME, 1967, vol. 239, pp. 1313-18. C.R. Nanda and G.H. Geiger: Metall. Trans., 1971, vol. 2, pp. 110106. K. Mori: Trans. Iron Steel Inst. Jpn., 1988, vol. 28, pp. 246-61. S.-H. Kim, R.J. Fruehan, and R.I.L. Guthrie: Scaninject V, MEFOS, Lulea, Sweden, 1989, pp. 385-418. M. Hirasawa, K. Mori, M. Sano, A. Hatanaka, Y. Shimatami, and Y. Okazaki: Trans. Iron Steel Inst. Jpn., 1987, vol. 27, pp. 277-82. M. Hirasawa, K. Mori, M. Sano, Y. Shimatami, and Y. Okazaki: Trans. Iron Steel Inst. Jpn., 1987, vol. 27, pp. 283-90. S.-H. Kim and R.J. Fruehan: Metall. Trans. B, 1987, vol. 18B, pp. 381-90. W.A. Krivsky and R. Schuhmann, Jr.: Trans. AIME, 1957, vol. 209, pp. 981-88. C.W. Bale and J.M. Toguri: Can. Metall. Q., 1976, vol. 15, pp. 30518. J. Lumsden: in Metal-Slag-Gas, Reactions and Processes, F.A. Foroulis and W.W. Smeltzer, eds., The Electrochemical Society, Princeton, NJ, 1975, pp. 155-69. S.N. Sinha and M. Nagamori: Metall. Trans. B, 1982, vol. 13B, pp. 461-70. P.J. Mackey: Can Metall. Q., 1982, vol. 21, pp. 221-60. A. Yazawa, Y. Takeda, and Y. Waseda: Can. Metall. Q., 1981, vol. 20, pp. 129-34. Y. Takeda, S. Kanesaka, and A. Yazawa: Proc. 25th Conf. of Metallurgists, 1986, CIM, Toronto, Canada, 1986, pp. 185-202. A. Yazawa and Y. Takeda: Met. Rev. MMIJ, 1987, vol. 4 (1), pp. 5365. M. Nagamori, Y. Takeda, and A. Yazawa: Met. Rev. MMIJ, 1989, vol. 6 (1), pp. 6-21. M. Nagamori, K. Itagaki, and A. Yazawa: Met. Rev. MMIJ, 1989, vol. 6 (1), pp. 22-37. F. Shenalek and I. Imris: Advances in Extractive Metallurgy and Rening, IMM, London, 1972, pp. 39-62. M. Nagamori: Metall. Trans., 1974, vol. 5, pp. 531-38. J.M. Toguri and N.H. Santander: Can. Metall. Q., 1969, vol. 8, pp. 167-71.

51. R. Altmann and H.H. Kellogg: Trans. IMM, 1972, vol. 81, pp. C163C175. 52. J.R. Taylor and J.H.E. Jeffes: Trans. IMM, 1975, vol. 84, pp. C18C24. 53. B.J. Elliot, J.B. See, and W.J. Rankin: Trans. IMM, 1978, vol. 87, pp. C204-C211. 54. T. Oishi, M. Kamuo, K. Ono, and J. Mariyama: Metall. Trans. B, 1983, vol. 14B, pp. 101-04. 55. C.C. Acholonu and R.G. Reddy: Paper presented at the TMS-AIME Meeting, Atlanta, GA, Mar. 1983. 56. H. Kurokawa, Y. Kondo, K. Baba, T. Inami, and N. Kemori: Int. Symp. on Injection in Process Metallurgy, T. Lehner, P.J. Koros, and V. Ramachandran, eds., TMS, Warrendale, PA, 1991, pp. 253-64. 57. T. Wei and F. Oeters: Int. Symp. on Injection in Process Metallurgy, T. Lehner, P.J. Koros, and V. Ramachandran, eds., TMS, Warrendale, PA, 1991, pp. 143-64. 58. O. Kubaschewski and C.B. Alcock: Metallurgical Thermochemistry, Pergamon Press, Elmsford, NY, 1979. 59. A.K. Kyllo: Masters Thesis, University of British Columbia, Vancouver, 1989. 60. C. Moore: UBC NLE: Zeros of Nonlinear Equations, Computing Centre, University of British Columbia, Vancouver, 1984. 61. M. Sun and Y. Luo: Nonferrous Metals, 1993, vol. 45 (3), pp. 53-59. 62. J.F. Elliot and M. Mounier: Can. Metall. Q., 1982, vol. 21, pp. 41528. 63. A.K. Biswas and W.G. Davenport: Extractive Metallurgy of Copper, 2nd ed., Pergamon Press, Oxford, United Kingdom, 1980. 64. M. Kucharski, N.M. Stubina, and J.M. Toguri: Can. Metall. Q., 1989, vol. 28, pp. 7-11. 65. T. Ejima and T. Yamamura: Conf. on the Properties of Liquid Metals, Proc. 2nd Int. Conf., Tokyo, Sept. 38, 1972, Taylor and Francis Ltd., London, 1973, pp. 537-41. 66. S.W. Ip and J.M. Toguri: Metall. Trans. B, 1993, vol. 24B, pp. 65768. 67. M. Hino, M. Nagamori, and J.M. Toguri: Metall. Trans. B, 1986, vol. 17B, pp. 913-14. 68. A.A. Bustos, J.K. Brimacombe, and G.G. Richards: Can. Metall. Q., vol. 27, 1988, pp. 7-21. 69. Book of Data, R.D. Harrison, The Nufeld Foundation, Longman Group Ltd., London, 1978. 70. E.J. Grimsey: Metall. Trans. B, 1988, vol. 19B, pp. 243-47. 71. S.S. Wang, A.J. Kurtis, and J.M. Toguri: Can. Metall. Q., 1973, vol. 12, pp. 383-90. 72. R.S. Celmer and J.M. Toguri: Proc. 25th Ann. Conf. of Metallurgists, CIM, Toronto, 1986, pp. 147-63. 73. M. Nagamori: Metall. Trans., 1974, vol. 5, pp. 539-48. 74. W. Jost: Diffusion in Solids, Liquids, Gases, Academic Press Inc., New York, NY, 1960. 75. M.P. Borom and J.A. Pask: J. Am. Ceram. Soc., 1968, vol. 51 (9), pp. 490-90. 76. K. Mori and K. Suzuki: Trans. Iron Steel Inst. Jpn., 1969, vol. 9, pp. 409-12. 77. F.D. Richardson: Physical Chemistry of Melts in Metallurgy, Academic Press, London, 1974. 78. Y. Ukyo and K.S. Goto: Metall. Trans. B, 1981, vol. 12B, pp. 44954. 79. S.V. Sapunov and S.E. Vaisburd: Tsv. Metall., 1985, vol. 26 (2), pp. 10-12. 80. D. Poggi, R. Minto, and W.G. Davenport: J. Met., 1969, vol. 21 (11), pp. 40-45. 81. J. Schoop, W. Resch, and G. Mahn: Ironmaking and Steelmaking, 1978, vol. 5 (2), pp. 72-79. 82. G. Turner and S. Jahanshahi: Trans. Iron Steel Inst. Jpn., 1987, vol. 27, pp. 734-39. 83. C. Delhaes, A. Hauck, and D. Neuschutz: Steel Res., 1993, vol. 64 (1), pp. 22-27. 84. T. Kimura, S. Tsuyuguchi, Y. Ojima, Y. Mori, and Y. Ishii: J. Met., 1986, Sept., pp. 38-42. 85. A.K. Kyllo and G.G. Richards: Metall. Mater. Trans. B, 1998, vol. 29B, pp. 261-68.

METALLURGICAL AND MATERIALS TRANSACTIONS B

VOLUME 29B, FEBRUARY 1998249

Potrebbero piacerti anche