Sie sind auf Seite 1von 291

A KINETIC MODEL OF THE

PEIRCE-SMITH CONVERTER

By

ANDREW KEVIN KYLLO


B.A.Sc., The University of British Columbia, 1986
M.A.Sc., The University of British Columbia, 1989

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE


REQUIREMENTS FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY
in
THE FACULTY OF GRADUATE STUDIES
METALS AND MATERIALS ENG]NEERING

We accept this thesis as conforming


to the required standard

THE UMVERSITY OF BRITISH COLUMBIA


August 1994
Andrew Kevin Kyllo, 1994
______________

In presenting this thesis in partial fulfilment of the requirements for an advanced


degree at the University of British Columbia, I agree that the Library shall make it
freely available for reference and study. I further agree that permission for extensive
copying of this thesis for scholarly purposes may be granted by the head of my
department or by his or her representatives, It is understood that copying or
publication of this thesis for financial gain shall not be allowed without my written
permission.

(Signature)

Department of NeIz 1
L
2 i, (4tJev:alc
t
J

The University of British Columbia


Vancouver, Canada

Date

DE-6 (2/88)
ABSTRACT

The Peirce-Smith converter, as used for copper and nickel converting, has changed

little in the eighty years since its introduction. Over this time other metal production

techniques have been developed, including considerable improvements to non-ferrous

smelting. These improvements have had only a small effect on non-ferrous converting.

The tenacity of the Peirce-Smith converter can be attributed to its simplicity of operation,

however, it is not an efficient process. While the converter itself is limited primarily by

its overall heat balance, process improvement has been limited by the belief that it

operated at thermodynamic equilibrium.

A kinetic model has been developed to gain a better knowledge of the operation of

the Peirce-Smith converter. The model consists of two parts; a model of the gas flow in

the bath, and an overall model considering both the heat and mass flows around the

converter. The gas flow model calculates the bubble growth on the tuyere to detachment,

and its subsequent rise through the bath. A combination of Kelvin-Helmolz and

Rayleigh-Taylor instability theories is used to determine the stability of the bubble, both

during growth and while rising through the bath. This allows the calculation of bubble

breakup, which can be used to determine the total gas/liquid interfacial area.

The gas flow model calculates the amount of oxygen reacting within each phase, as

well as the heat lost to the gas and the total interfacial area. These values are applied to

the overall model which then calculates the heat and mass balances within the converter.

Material flows are based on mass-transfer considerations, with each phase being

considered separately. Both mass and heat-transfer occur between all phases present, and

each phase is assumed to be in internal equilibrium. As well as calculating the behaviour

11
of the more abundant elements within the converter, the behaviour of the more important

minor elements is considered. The model is validated using published physical

modelling results, as well as measurements made on operating converters.

The model results indicate that the efficiency of the Peirce-Smith converter may be

improved by a number of methods, provided that some means of controlling the bath

temperature is available. Increasing the tuyere submergence and decreasing the tuyere

size are predicted to provide a substantial improvement in operation efficiency, without

adversely affecting the minor element removal. The use of low levels of oxygen

enrichment also improves efficiency, but tends to reduce the extent of minor element

removal. Higher levels of oxygen enrichment are predicted to alter the overall process

chemistry, with the amount of iron reacting being controlled by liquid-phase

mass-transfer. This improves both the overall process efficiency and the extent of minor

element removal.

111
TABLE OF CONTENTS

Abstract ii
Table of Contents iv
Table of Tables viii
Table of Figures xii
Acknowledgments xx
1 INTRODUCTION 1
1.1 Development of the Copper Converter 1
1.2 Development of Ferrous Converting 4
1.3 Comparison of Ferrous and Non-Ferrous Converting 5
2 LITERATURE REVIEW 8
2.1 Introduction 8
2.2 Converter Modelling 8
2.2.1 Converter Operation 8
2.2.2 Impurity Distribution 10
2.3 Process Kinetics 12
2.3.1 Copper Converting Kinetics 13
2.3.2 Kinetics of Gas/Liquid Reactions 14
2.3.3 Kinetics of Liquid/Liquid Reactions 15
2.4 Process Thermodynamics 16
2.4.1 Matte Thermodynamics 16
2.4.2 Slag Thermodynamics 18
2.4.3 Internal Phase Equilibrium 18
3 OBJECTIVES AND SCOPE 19
4 INDUSTRIAL DATA 21
4.1 Copper Converter Trials 21
4.2 Nickel Converter Trials 26
4.3 Analysis 33
4.3.1 Oxygen Efficiency 33
4.3.2 Material Deportment 36
4.3.2.1 Major components 36
4.3.2.2 Minor elements 40

iv
4.3.3 Converter Dusts . 44
4.3.3.1 Flue dust 45
4.3.3.2 Cottrell dust 52
4.3.3.3 Baghouse dust 57

5 GASFLOWINTHEBATH 59
5.1 Basic Bubble Growth 59
5.2 Bubble Detachment Criterion 67
5.3 Bubble Break-up During Growth 71
5.4 Bubble Rise 77
5.5 Bubble Recombination 79

6 MODEL DEVELOPMENT 81
6.1 Introduction 81
6.2 Preliminary Considerations 81
6.2.1 Bath Velocity and Gas Holdup 81
6.2.2 Converter Geometry 84
6.3 Mass Balance 90
6.3.1 Equilibrium Within Phases 90
6.3.1.1 Blister copper 90
6.3.1.2 Matte 90
6.3.1.3 White metal 91
6.3.1.4 Slag 91
6.3.1.5 Gas 91
6.3.1.6 Calculation technique 92
6.3.2 Gas-Liquid Mass-transfer 92
6.3.2.1 Gas flow 92
6.3.2.2 Mass transfer from liquid to gas bubbles 95
6.3.3 Mass Transfer Between Liquid Phases 98
6.4 Energy Balance 100
6.4.1 Energy Loss 100
6.4.2 Energy Consumption 102
6.4.3 Energy Generation 104
6.4.4 Interphase Heat-transfer 105

v
6.5 Data. 105
6.5.1 Physical and Thermal Properties 105
6.5.2 Thermodynamic Data 107
6.5.2.1 Free energy 107
6.5.2.2 Enthalpy of reaction 111
6.5.2.3 Activity coefficients 112
6.5.2.4 Equilibrium vapour pressures 118
6.5.3 KineticData 119
6.5.3.1 Diffusivities 119
6.5.3.2 Mass-Transfer coefficients 121
6.5.3.3 Heat-Transfer coefficients 122
7 MODEL VALIDATION 123
7.1 Bubble Model Validation 123
7.1.1 Bubble Growth 123
7.1.2 Bubble Rise 124
7.2 Copper Converter Validation 128
7.2.1 Bath Temperature 128
7.2.2 Slag Composition 130
7.2.3 Matte Composition 133
7.3 Nickel Converter Validation 136
7.3.1 Bath Temperature 137
7.3.2 Slag Composition 140
7.3.3 Matte Composition 143
7.4 Discussion 146
7.4.1 Phase Compositions 146
7.4.1.1 Error in assays 146
7.4.1.2 Sulphur in slag 147
7.4.1.3 Oxygen in matte 148
7.4.1.4 Minor element distribution 149
7.4.2 Copper Blow 153

8 MODEL PREDICTIONS AND DISCUSSION 157


8.1 Gas Flow Model 157

vi
8.2 ConverterModel. 166
8.2.1 Introduction 166
8.2.2 Copper Converter Charge 167
8.2.3 Sensitivity Analysis 171
8.2.4 Converter Operation 186
8.2.4.1 Gasflowrate 186
8.2.4.2 Oxygen enrichment 188
8.2.4.3 Tuyere submergence and diameter 197
8.2.4.4 Slag skimming procedure 202
8.2.5 Minor Element Removal 204
8.2.5.1 Introduction 204
8.2.5.2 Minor element behaviour in the base charge 205
8.2.5.3 Model predictions 213
8.2.5.3.1 Gasflowrate 213
8.2.5.3.2 Oxygen enrichment 214
8.2.5.3.3 Tuyere submergence and diameter 229
8.2.5.4 Summary 233
8.2.6 Comparison with Previous Work 234
8.2.6.1 Overall model 234
8.2.6.2 Minor element distribution 235

9 MODEL APPLICATION 238


9.1 Introduction 238
9.2 Converter Optimization 238
9.3 New Process Development 242

10 CONCLUSIONS AND FURTHER WORK 247

11 NOMENCLATURE 250

12 REFERENCES 255

13 APPENDIX 266

vi
TABLE OF TABLES

Table IV-I Details of copper charges followed 22


Table IV-ll Material assays, copper converter charge 586, Feb. 1, 1994.
Nunbers in parentheses refer to the blow number in Table TV-I 23
Table TV-HI Material assays, copper converter charge 588, Feb. 2, 1994.
Nunbers in parentheses refer to the blow number in Table TV-I 24
Table TV-TV Material assays, copper converter charge 595, Feb. 4, 1994.
Nunbers in parentheses refer to the blow number in Table TV-I 25

Table TV-V Flux assay, copper converter plant trials, Feb., 1994 25
Table TV-VT Methods of detennination for each assay 26

Table IV-Vll Details of nickel converter charges 29


Table IV-VllI Material assays, nickel converter charge 105, May 2, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Nunbers in
parentheses refer to the blow number in Table TV-VU 31
Table IV-IX Material assays, nickel converter charge 106, May 24, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Nunbers in
parentheses refer to the blow number in Table TV-VU 31
Table IV-X Material assays, nickel converter charge 107, May 25, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Nunbers in
parentheses refer to the blow number in Table IV-Vll 32
Table IV-XT Material assays, nickel converter charge 108, May 26, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Nunbers in
parentheses refer to the blow number in Table TV-VU 32

Table IV-Xll Overall oxygen efficiencies for copper converter charges


followed in plant trials 34

Table IV-XllI Oxygen efficiencies calculated from matte samples 34

Table IV-XTV Calculated nickel converter oxygen efficiencies 35

Table TV-XV Distribution of major components between output streams of the


copper converter 37

Table IV-XVT Distribution of major components in nickel converting 38

Table IV-XVTI Compositions of converter dust samples 44

viii
Table V-I Tpical copper converting conditions used for bubble size
calculations 60

Table V-TI Bubble size at detachment calculated from different models 61

Table VT-I. Reacting components of the matte 90

Table VT-fl Reactions required to calculate the equilibrium composition of the


matte 91

Table VT-HI Integrated values of C. for use in equation 6.73. 103

Table VT-TV Enthalpies of reaction required for energy generation


36
calculation. 104

Table VT-V Physical and thermal properties of the gas and condensed phases. .. 106

Table VT-VT Coefficients for gas viscosity equation (air).


42 107

Table VI-VIl Free energy of formation of matte compounds.


142 107

Table VT-VIIT Free energy of reaction for matte-slag reactions.


35 108

Table VT-TX Free energy of reaction for slag and gas reactions.
135 108

Table VI-X Activity coefficients of the major constituents of the matte and
slag 112

Table VT-XT Minor element activity coefficients used in the model 117

Table VI-Xll Equilibrium vapour pressures of the minor elements 119

Table VT-XIIT Atomic/ionic/molecular radii of matte, slag and bullion


36
constituents. 121

Table VTT-T Comparison of measured and predicted gas penetration in


mercury 123

Table VTT-IT Comparison of measured and predicted gas fraction, bath velocity,
and spout height in vertical injection systems 126

Table VTT-TTT Comparison of model predicted slag compositions with assays,


(weight percent), #1 converter charge 586, Feb., 1994 131

Table VH-TV Comparison of model predicted slag compositions with assays,


(weight percent), #1 converter charge 588, Feb., 1994 131

Table VTI-V Comparison of model predicted slag compositions with assays,


(weight percent), #1 converter charge 595, Feb., 1994 132

x
Table Vu-VT Comparison of model predicted matte and blister copper
compositions with assays, (weight percent), #1 converter charge 586, Feb.,
1994 134
Table Vu-Vu Comparison of model predicted matte and blister copper
compositions with assays, (weight percent), #1 converter charge 588, Feb.,
1994. Asterisk indicates combined matte and slag to bring the matte silica
content up to assayed value 134
Table VIT-VITI Comparison of model predicted matte compositions with
assays, (weight percent), #1 converter charge 595, Feb., 1994. Asterisk
indicates combined matte and slag to bring the matte silica content up to assayed
value 135
Table Vu-TX Comparison of model predicted slag compositions with assays
taken at the end of the blow, (weight percent), #3 converter charge 105, May
1988 141
Table VuT-X Comparison of model predicted slag compositions with assays
taken at the end of the blow, (weight percent), #3 converter charge 106, May
1988 141

Table VIT-XI Comparison of model predicted slag compositions with assays


taken at the end of the blow, (weight percent), #3 converter charge 107, May
1988 142

Table VII-XII Comparison of model predicted slag compositions with assays


taken at the end of the blow, (weight percent), #3 converter charge 108, May
1988 142
Table VIT-XITI Comparison of model predicted matte compositions with
assays, (weight percent), #3 converter charge 105, May 1988 144

Table VII-XTV Comparison of model predicted matte compositions with


assays, (weight percent), #3 converter charge 106, May 1988 144

Table VTT-XV Comparison of model predicted matte compositions with assays,


(weight percent), #3 converter charge 107, May 1988 144

Table VuT-XVT Comparison of model predicted matte compositions with


assays, (weight percent), #3 converter charge 108, May 1988 145

Table VII-XVII Comparison of slag assays obtained using TCP and wet assay
techniques 147

Table VuuT-T Parameters tested in the sensitivity analysis 172

Table Vu-TI Predicted distribution of minor elements to the blister copper and
dust after 400 minutes of charge 586 (72 tonnes of blister copper produced) 205

x
Table VIJI-ifi Comparison of equilibrium and kinetic model predicted minor
element distributions with commercially observed ranges 237
Table IX-I Conditions used in the improved charge 240

Table IX-il Conditions used in the new process 244


Table TX-HI Composition of matte added in the new process 244

xi
TABLE OF FIGURES

Figure 1.1 Schematic of the Peirce-Smith converter 4


Figure 2.1 Comparison of reported values of FeS activity coefficient in copper
59

9
mattes. 17
Figure 4.1 Variation in measured bath temperature and air rate in a nickel
converter, (#3 converter, charge 105, May 23, 1988) 28
Figure 4.2 Variation of total copper, nickel, and cobalt in slag with nickel
converter matte grade 39
Figure 4.3 Variation of total copper, nickel, and cobalt in slag with slag oxygen
potential 40

Figure 4.4 Minor element distribution in Hudsons Bay Mining and Smelting
copper converters, charges 586, 588, and 595, February 1994 41

Figure 4.5 Antimony and bismuth distribution in Hudsons Bay Mining and
Smelting copper converter charge 595, February 1994 42

Figure 4.6 Minor element distribution in Inco nickel converter (#3 converter,
charges 105, 107, and 108, May 1988) 43

Figure 4.7 Schematic of the dust collection system at Hudsons Bay Mining
and Smeltings Fun Flon smelter. Numbers indicate approximate position from
which Cottrell dust samples were taken 45

Figure 4.8 Copper converter flue dust, x25 47

Figure 4.9 Copper converter flue dust, a. coated particle, x150, b. particle with
pore, x300 48

Figure 4.10 Copper converter flue dust, mounted and sectioned, x25 49

Figure 4.11 Copper converter flue dust, mounted and sectioned, a. particle
containing entrained slag and copper droplets, xl 25, b. particle containing
copper droplets, xlOO, c. fractured particle containing copper, x125, d. slag
particle containing copper droplet, x80 51

Figure 4.12 Copper converter Cottrell dust, a. sample 1, xl 00, b. sample 2,


xlOO, c. sample 4, x160. Numbers refer to sample positions shown in
Figure 4.7 54

Figure 4.13 Copper converter Cottrell dust with matrix removed, a. sample 1,
xlOO, b. sample 2, x160, c. sample 3, x160. Numbers refer to sample positions
shown in Figure 4.7 55

xii
Figure 4.14 Copper converter Cottrell dust, a. sample 2, x300, b. sample 1,
agglomerated fume, x2000. Numbers refer to sample positions shown in
Figure 4.7 56
Figure 4.15 Smelter baghouse dust, x1500 58
Figure 5.1 Schematic of high Froude number injection of air into water.

11 62
Figure 5.2 Ellipsoidal bubble growth 64
Figure 5.3 Comparison of horizontal and vertical injection conditions 65
Figure 5.4 Bubble detachment process 68
Figure 5.5 Bubble detachment geometry 69
Figure 5.6 Variation of the height of the bubble centre above the tuyere
centreline (h) and the distance of the bubble centre from the tuyere centre (s)
with bubble radius at detachment 70
Figure 5.7 Variation of bubble position at detachment with bubble diameter 71

Figure 5.8 Schematic illustration of flow around a bubble.


18 73

Figure 5.9 Variation of maximum bubble diameter and wavenumber with bath
velocity 77

Figure 6.1 Schematic cross-section of an idle converter 85


Figure 6.2 Schematic cross-sections of an operating converter: a. no gas flow
through the slag, b. gas flow through matte and slag, c. gas flow through slag
only 87
Figure 6.3 Schematic of emulsion formation during submerged injection
27 89

Figure 6.4 Comparison of reported values of free energy of reaction 109

Figure 6.5 Comparison of reported values of free energy of formation of SbS


and SbO 110

Figure 6.6 Comparison of reported values of free energy of formation of AsS


andAsO 111

37 and predicted spout height


Figure 7.1. Comparison of measured 125

Figure 7.2 Variation of fraction sulphur dioxide reacted with gas flow rate
including the fitted effect of surface reaction 127

Figure 7.3 Variation of fraction sulphur dioxide reacted with tuyere


submergence including the fitted effect of surface reaction 128

xiii
Figure 7.4 Comparison of measured and predicted copper converter
temperatures charge 586, February, 1994 129
Figure 7.5 Comparison of measured and predicted copper converter
temperatures charge 588, February, 1994 129
Figure 7.6 Comparison of measured and predicted copper converter
temperatures charge 595, February, 1994 130
Figure 7.7 Comparison of measured and predicted iron and silica contents in
copper converter slags, charges 586, 588, and 595, Feb. 1994 132
Figure 7.8 Comparison of measured and predicted a. iron and b. copper
contents in copper converter mattes, charges 586, 588, and 595, Feb. 1994 136
Figure 7.9 Comparison of model predicted matte and slag temperatures with
plant data, #3 converter charge 105, May 1988 138
Figure 7.10 Comparison of model predicted matte and slag temperatures with
plant data, #3 converter charge 106, May 1988 138
Figure 7.11 Comparison of model predicted matte and slag temperatures with
plant data, #3 converter charge 107, May 1988 139

Figure 7.12 Comparison of model predicted matte and slag temperatures with
plant data, #3 converter charge 108, May 1988 139
Figure 7.13 Comparison of measured and predicted iron and silica contents in
nickel converter slags, #3 converter, charges 105, 106, 107, and 108, May 1988.
143
Figure 7.14 Comparison of measured and predicted iron, nickel, and copper
contents in nickel converter mattes, #3 converter, charges 105, 106, 107, and
108, May 1988 145

Figure 7.15 Calculated variation of mole fractions of Fe


4 and FeO in matte.
0
3 ... 149

Figure 7.16 Comparison of model predicted lead, zinc and arsenic distributions
with measured values, copper converter charges 586, 588, and 595 Flin Flon,
February 1994. Solid lines indicate commercially observed range. 1 151

Figure 7.17 Comparison of model predicted antimony and bismuth


distributions with measured values, copper converter charge 595, Flin Flon,
February 1994 152

Figure 7.18 Comparison of model predicted lead, zinc and arsenic distributions
with measured values, nickel converter charges 105, 107, and 108, Copper Cliff,
May 1988 153

xiv
Figure 7.19 Predominance area diagrams for the iron-copper-sulphur-oxygen
system; a. 1300 K, b. 1500 K. Dashed lines indicate partial pressure of sulphur
dioxide 155
Figure 8.1 Variation of primary bubble temperature with time for injection
conditions given in Table V-I 158
Figure 8.2 Effect of ellipse eccentricity on surface area at constant volume,
shaded area indicates eccentricity range predicted by the model 160
Figure 8.3 Variation of primary bubble volume with time for injection
conditions given in Table V-I 160

Figure 8.4 Variation of average oxygen partial pressure with tuyere


submergence for injection conditions given in Table V-I 161

Figure 8.5 Variation of average oxygen partial pressure with tuyere diameter,
all other conditions given in Table V-I 161

Figure 8.6 Variation of primary bubble volume with tuyere diameter, all other
conditions given in Table V-I 162
Figure 8.7 Variation of total oxygen reacted with inlet gas oxygen content, all
other conditions given in Table V-I 164
Figure 8.8 Variation of primary bubble temperature with bath material, all other
conditions given in Table V-I 164

Figure 8.9 Variation of number of bubbles formed with bath material, all other
conditions given in Table V-I 165

Figure 8.10 Effect of tuyere interaction on average oxygen partial pressure all ,

other conditions given in Table V-I 165

Figure 8.11 Effect of tuyere interaction on primary bubble volume, all other
conditions given in Table V-I 166

Figure 8.12 Predicted matte temperature, charge 586. F-flux addition, M-matte
addition, I-idle period start, I*idle period end, S-slag skimmed, Sc-scrap added,
C-copper blow start 168

Figure 8.13 Predicted variation of oxygen use with time 170

Figure 8.14 Predicted variation of tuyere submergence (including spout) with


time 170

Figure 8.15 Predicted variation of oxygen use with tuyere submergence


(including spout), charge 586, Feb. 1994 171

Figure 8.16 Effect of model time step on the predicted variation of matte
temperature and iron content (weight percent) with time 173

xv
Figure 8.17 Effect of gas flow calculation frequency on the predicted variation
of matte temperature and iron content (weight percent) with time 174
Figure 8.18 Effect of initial bath temperature on the predicted variation of
matte temperature and iron content (weight percent) with time 175
Figure 8.19 Effect of initial bath temperature on the predicted variation of
matte oxygen content (weight percent) with time 176
Figure 8.20 Effect of slag emissivity on the predicted variation of slag
temperature with time 177
Figure 8.21 Effect of liquid-phase diffusivity on the predicted variation of
matte temperature and iron content (weight percent) with time 178
Figure 8.22 Effect of liquid-phase diffusivity on the predicted variation of
matte zinc content (weight percent) with time 179
Figure 8.23 Effect of liquid-phase diffusivity on the predicted variation of slag
oxygen content (weight percent) with time 180

Figure 8.24 Effect of liquid-phase diffusivity on the predicted variation of


oxygen use with time 180

Figure 8.25 Effect of gas-phase mass-transfer coefficient on the predicted


variation of matte temperature and iron content (weight percent) with time 182

Figure 8.26 Effect of gas-phase mass-transfer coefficient on the predicted


variation of slag oxygen content (weight percent) with time 183

Figure 8.27 Effect of weight of matte in a ladle on the predicted variation of


matte temperature and iron content (weight percent) with time 185

Figure 8.28 Effect of gas flow rate on the predicted variation of matte
temperature and iron content (weight percent) with time 187

Figure 8.29 Effect of gas oxygen content, with total oxygen input constant, on
the predicted variation of matte temperature and iron content (weight percent)
with time 190

Figure 8.30 Effect of gas oxygen content, with total oxygen input constant, on
the predicted variation of matte iron content with iron as magnetite removed
(weight percent) with time 191

Figure 8.31 Effect of gas oxygen content, with increased oxygen input, on the
predicted variation of matte temperature and iron content (weight percent) with
time 192

Figure 8.32 Effect of gas oxygen content on the predicted variation of matte
iron content (weight percent) with time at 1400 K 195

xvi
Figure 8.33 Effect of gas oxygen content on the predicted variation of oxygen
use with time at 1400 K 195
Figure 8.34 Variation of the predicted amount of iron and copper suiphides
reacting, as a percentage of the total reaction in the matte, with time at 1400 K
and 84% oxygen in the injected gas 196
Figure 8.35 Effect of gas oxygen content on the predicted variation of copper
produced with time at 1400 K 196
Figure 8.36 Effect of gas oxygen content on the predicted variation of sulphur
potential with time at 1400 K 197
Figure 8.37 Effect of tuyere submergence on the predicted variation of matte
temperature and iron content (weight percent) with time 199
Figure 8.38 Effect of tuyere diameter on the predicted variation of matte
temperature and iron content (weight percent) with time 200
Figure 8.39 Effect of a combined reduction in tuyere diameter and increase in
tuyere submergence on the predicted variation of matte temperature and iron
content (weight percent) with time 201
Figure 8.40 Effect of skimming procedure on the predicted variation of matte
temperature and iron content (weight percent) with time 203
Figure 8.41 Effect of skimming procedure on the predicted variation of oxygen
use with time 204
Figure 8.42 Predicted variation of minor element concentrations in blister
copper with time, a. lead and zinc, b. arsenic, antimony, and bismuth 207
Figure 8.43 Variation of surface area-to-volume ratio of blister copper with
depth 208
Figure 8.44 Predicted variation of minor element concentrations in matte with
time, a. lead and zinc, b. arsenic, antimony, and bismuth 210

Figure 8.45 Predicted variation of minor element concentrations in slag with


time, a. lead and zinc, b. arsenic, antimony, and bismuth 211

Figure 8.46 Predicted variation of minor element partial pressures in gas with
time, a. lead and zinc, b. arsenic, antimony, and bismuth 212

Figure 8.47 Predicted variation of minor element distribution to the dust with
gas flow rate, a. lead b. zinc, c. arsenic, d. antimony, and e. bismuth 214

Figure 8.48 Predicted variation of minor element distribution to the dust with
gas oxygen content with constant total oxygen input, a. lead b. zinc, c. arsenic,
d. antimony, and e. bismuth 216

xvii
Figure 8.49 Predicted variation of minor element distribution to the blister
copper with gas oxygen content with constant total oxygen input, a. lead b. zinc,
c. arsenic, d. antimony, and e. bismuth 217
Figure 8.50 Predicted variation of minor element distribution to the dust with
gas oxygen content, with increased total oxygen input, a. lead b. zinc, c. arsenic,
d. antimony, and e. bismuth 218
Figure 8.51 Effect of gas oxygen content, with increased total oxygen input, on
the predicted variation of arsenic partial pressure in the off gas with time 219
Figure 8.52 Predicted variation of minor element distribution to the blister
copper with gas oxygen content, with increased total oxygen input, a. lead
b. zinc, c. arsenic, d. antimony, and e. bismuth 220
Figure 8.53 Effect of gas oxygen content, with increased total oxygen input, on
the predicted variation of arsenic content in blister copper with time 221
Figure 8.54 Effect of gas oxygen content, with increased total oxygen input, on
the predicted variation of antimony content in blister copper with time 222
Figure 8.55 Predicted variation of minor element distribution to the blister
copper with gas oxygen content, with constant total oxygen input and
temperature, a. lead b. zinc, c. arsenic, d. antimony, and e. bismuth 225
Figure 8.56 Effect of gas oxygen content, with constant total oxygen input, on
the predicted variation of arsenic content in blister copper with time at 1400 K.
226
Figure 8.57 Predicted variation of minor element distribution to the dust with
gas oxygen content, with constant total oxygen input and temperature, a. lead
b. zinc, c. arsenic, d. antimony, and e. bismuth 227
Figure 8.58 Effect of gas oxygen content, with constant total oxygen input, on
the predicted variation of arsenic partial pressure in the off gas with time at
1400 K 228
Figure 8.59 Effect of gas oxygen content, with constant total oxygen input, on
the predicted variation of moles of arsenic and mole fraction of copper sulphide
in matte with time at 1400 K 228

Figure 8.60 Predicted variation of minor element distribution to the blister


copper with tuyere submergence, a. lead b. zinc, c. arsenic, d. antimony, and
e. bismuth 231

Figure 8.61 Predicted variation of minor element distribution to the dust with
tuyere submergence, a. lead b. zinc, c. arsenic, d. antimony, and e. bismuth 232

Figure 8.62 Effect of tuyere submergence on the predicted variation of arsenic


content in blister copper with time 233

xviii
Figure 9.1 Effect of modifications given in Table DC-I on the predicted
variation of matte temperature and iron content (weight percent) with time 241
Figure 9.2 Effect of modifications given in Table DC-I on the predicted
variation of oxygen use with time 242
Figure 9.3 Predicted variation of matte temperature and iron content (weight
percent) with time for the new process 245
Figure 9.4 Predicted variation of oxygen use with time for the new process. ... 246
Figure 9.5 Predicted variation of weight of blister copper produced with time
for the new process 246
Figure 13.1 Simplified flow chart of the overall model 267
Figure 13.2 Flow chart of the gas flow calculation 268
Figure 13.3 Flow chart of the mass balance calculation 269
Figure 13.4 Flow chart of the heat balance calculation 270

xix
ACKNOWLEDGMENTS

I would like to acknowledge the financial support for this project from The National
Science and Engineering Council and Hudson Bay Mining and Smelting Ltd. I would
also like to thank Kevin Scott and Dominic Verheist for their assistance during the plant
trials, and my supervisor, Dr. Greg Richards. Finally, I should thank my wife for her
continued patience and assistance and my children for allowing me to work.

xx
1.1 Development of the Copper Converter

1 INTRODUCTION

The pyrometallurgical production of copper from sulphide concentrates is a

complex procedure which at present requires a number of steps before the refining stage.

Although some modernization of the process has been carried out, most smelters use

techniques which have changed little in eighty years. In the last forty years important

innovations have been made in copper smelting technology, but only a small amount of

research has been carried out on copper converting. Thus the Peirce-Smith converter has

changed little in this period, and has not taken advantage of many innovations which

have benefitted other industries. In order to demonstrate the extent of the problem the

development of the copper converter can be compared with that of the ferrous converter.

1.1 Development of the Copper Converter

Copper converting originated at the same time as steel converting with Bessemers

introduction of the pneumatic converter. However, there are problems associated with

copper mattes which do not arise in the converting of pig iron. The most important of

these are the relative densities of three phases involved, (slag, matte and blister copper)

and the volume of slag produced.

The blister copper is the most dense of the phases produced, and so forms the

bottom layer in the converter. This leads to two operating difficulties for the

Bessemer-type converter: the heat generated by the reaction of air with copper is

insufficient to prevent tuyere blockage and the air oxidizes the copper rather than the

matte. Although both of these difficulties are experienced only after the formation of

blister copper, they proved to be a serious impediment to the development of the copper

converter. The corrosive nature of the matte and slag on the acidic refractories also posed

1
1.1 Development of the Copper Converter

a problem. The large volume of slag produced required intermediate slag skimming and

matte charging before the finishing blow. This was not a serious problem, but it did

require extra cranes, longer overall charge times, and frequent relining.

The problems relating to bottom blowing were overcome in 1880 by Pierre Manhs.

His solution was to move the tuyeres to the side of the converter and place them a few

inches above the base to give a quiescent zone for the copper to collect in. This form of

converter was introduced into the United States four years later, where it was developed

further, first as the Parrot converter, and then as the Great Falls converter. It is no longer

in use.

Manhs had also worked with the idea of using a horizontal converter, although he

gave this up in favour of the Bessemer style upright converter. However, others in the

industry preferred the use of the barrel converter, apparently due to the lower injection

pressures which could be used with it.


2 Both types of converters used a sacrificial lining

of siicious material, which had to be replaced after only a few charges. It was not until

1909 that this problem was solved by Peirce and Smith who developed the basic

magnesite lining which allowed, with a change in fluxing, a much longer refractory life.
3

The horizontal, basic lined converter then became known as the Peirce-Smith converter.

Over the last 80 years the size of the Peirce-Smith converter has increased, the

quality of the refractory has improved and automatic tuyere punching has been added.

Some smelters have introduced oxygen enrichment of up to 30% in the injected air.

Otherwise there has been little change.

The present form of the Peirce-Smith converter is shown in Figure 1.1. It is a

horizontal cylinder with lengths from seven to eleven metres. Chrome-magnesite

refractory is used in most operations, but the proportions of the components varies. A

2
1.1 Development of the Copper Converter

copper converter charge consists of a number of slag blows followed by the copper or

finish blow. Before each slag blow molten matte is charged to the converter. Air is then

blown through the matte causing the iron to be removed by the reaction

3 ...[l.1]
FeS +O2 = FeO+ SO
2

The iron oxide is removed to the slag, where it reacts with the silica in the flux to form

fayalite,

=iO
2FeO-i-Si0
S
2
4 Fe
Blowing then continues until almost all of the iron has been removed. The slag is then

skimmed, more matte is added, and blowing continues. After the last slag blow, slag is

skimmed, but no more matte is added. During the copper blow the oxygen in the injected

gas reacts with the copper sulphide to produce copper,

S
Cu
=
2 2Cu+S0
+0
Nickel converting is similar to copper converting, except that the copper blow is

replaced with miss and dry-up blows. The procedure for these is the same as for the

copper blow, in that no matte is added, but the purpose is to ensure that the iron levels in

the matte are sufficiently low for further processing. The slag formed in the dry-up

blow is very viscous, and contains a large amount of nickel. It is generally termed

mush, and remains in the converter to be cleaned and to provide the flux for the first

slag blow of the following charge.

Recently new methods for the pyrometallurgical processing of copper suiphides

have been developed. Although the majority of these are new smelting techniques, there

has been some advance in converting technology. Flash smelting has been found to be

able to produce a high grade matte or blister copper, but the slag losses under the

3
1.2 Development of Ferrous Converting

BUSTLE
PIPE

Figure 1.1 Schematic of the Peirce-Smith converter.

conditions required for direct blister production are quite high.


4 The Noranda Process, a

side blown bath smelting technology can also be used to produce copper directly, but this

mode of operation is avoided due to impurity and refractory problems.


5 The Mitsubishi

continuous converter uses a top lancing system in which oxygen enriched air is generally

6 Most recently Inco has developed a modified Peirce-Smith converter which


employed.

allows oxygen top lancing.


7

1.2 Development of Ferrous Converting

Shortly after the introduction of the Bessemer converter a basic lining for ferrous

converting was developed by Gilchrist and Thomas in 1877, primarily to allow the

removal of phosphorous from the steel. The use of tonnage oxygen was first introduced

in Austria in 1948. This was in a top blown Bessemer converter, and has become known

as the L-D process.

4
1.3 Comparison of Ferrous and Non-Ferrous Converting

Since then further innovations include the combined-blowing converter, which uses

oxygen lancing with inert gas stirring, and the introduction of the shrouded tuyere which

allows oxygen to be introduced either through the bottom or the side of the converter.

These recent improvements all make use of high pressure injection, either in the lance or

through the tuyeres.

1.3 Comparison of Ferrous and Non-Ferrous Converting

From the outlines of converter development given above it is evident that

technologically the non-ferrous industries are considerably behind the ferrous. While the

use of high pressure gas has been tested,


8 it is not yet used in any installation. The main

concern which is preventing its use is that the increased energy input to the bath will be

more likely to cause bath slopping and increased splashing. Physical modelling studies

of slopping in the Peirce-Smith converter have shown that it is directly related to the

9 These studies also indicate that increasing


bouyant power per unit mass of the bath.

tuyere submergence allows a greater gas flow rate to be used,


9 and that for an

intermediate range of tuyere submergences there is a range of gas flow rates, in some

cases quite high, at which no waves are formed.


This suggests that higher flow rates
1

could be used. In the steel converter, the vessel shape and tuyere arrangement allows a

considerably higher flow rate to be used. The vessel shape also allows an increased

tuyere submergence.

The batch nature of the copper converter can lead to it being idle for over 50% of

The steel converter is also a batch process, but the differences


the total charge time.
1

between the processes allow the steel converter to operate more efficiently. The main

difference between the two processes is the final form of the material being removed

from the bath. In steel converting carbon is removed as a mixture of CO and CO


2 gas,

5
1.3 Comparison of Ferrous and Non-Ferrous Converting

while in the slag blows of a copper converter iron must be removed as an oxide contained

in a liquid slag. This results in a large volume of slag which must be removed from the

copper converter. It is interesting to note that there is a considerable amount of research

being carried out to develop a continuous steel making process.


4 It is possible that the

1
development of a continuous copper converting process would be preferable to altering

the present process.

Oxygen lancing has recently been introduced in a modified converter,


7 and oxygen
enrichment of the blast air is a fairly common practice.
The use of oxygen gives a
1

higher converting rate and gives an off gas high in SO


, which is more economical to
2

clean than the low concentrations presently obtained. However, the use of tonnage

oxygen, as in steel converting has not been used. There are three main reasons for

limiting the oxygen content of the blast air. The first of these is that at high oxygen

contents the tuyere life is reduced dramatically. The solution to this problem is already

available, in the form of shrouded tuyeres, as used in bottom blown steel converters. The

second problem relates to the heat balance. It has been determined that the off gas

accounts for the majority of heat removal from the converter.


5 Reducing the nitrogen

content and, therefore, total gas throughput will cause a large increase in bath

temperature and a related decrease in refractory life. However, improved refractory and

changes in slagging practices may allow higher temperatures to be used. Also, the higher

matte grades available from the new smelting furnaces produce considerably less heat

due to their lower iron contents, and any increase in available heat could be used to

increase scrap recycling. The other potential problem relating to the use of tonnage

oxygen in the copper converter is the removal of minor elements. It has been determined

that a large proportion of lead, arsenic and bismuth are removed from the converter by

6
1.3 Comparison of Ferrous and Non-Ferrous Converting

volatilization, and it has been suggested that the use of oxygen will result in a much
16 This problem, along with obtaining a better overall
lower removal of these elements.

understanding of the Peirce-Smith converter through the use of a kinetic model, will be

the focus of this study.

7
2.2.1 Converter Operation

2 LITERATURE REVIEW

2.1 Introduction

The development of a kinetic model of the Peirce-Smith converter requires a wide

range of background information. This chapter will cover previous models of the

converter and related technology, as well as the general kinetics and overall

thermodynamics required. The literature relating to other aspects of the modelling will

be introduced where appropriate.

2.2 Converter Modelling

The majority of 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.

2.2.1 Converter Operation

A model of the copper converter, developed by Goto, has been gradually improved

over the last fifteen years, most recently being applied to the copper flash smelter.

72

The first part of the model to be developed considered only the mass balance.
7

Assuming that the converter is in thermodynamic equilibrium, a set of simultaneous

equations was formed and solved using a modified form of the Newton-Raphson

22 The activity coefficients required in these equations


technique formulated by Brinkley.

were derived from published data and experimental work by the authors. The model

considered nine elements and nineteen compounds, to give a representation of the

distribution of most of the important elements present in copper concentrates in Japanese

smelters. Of the minor elements of interest here only lead was considered.

8
2.2.1 Converter Operation

The second development was the addition of a heat balance to complete the

8 The model applied a heat balance to the


representation of the entire converter.

converter, using the mass balance calculations to derive a heat generation term. An

iterative technique was used to calculate the temperature change caused by any net heat

production during a given interval. Calculations were carried out over a two minute time

step throughout the converting cycle, allowing for charging and skimming. It was

claimed that the model was able to predict temperature variations fairly well,
18 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

9 and the addition of a calculation of oxygen


extension to cover the copper flash furnace,

consumption using kinetic considerations.



2

A model of the heat and mass balances in the nickel converter has recently been

It is based on the work of Goto et al., and has been found to be able to
14
produced.

predict both the bath temperature and composition fairly accurately.

A model of the Noranda continuous converting process has been developed which

concentrates primarily on the minor element distribution. This model calculates the

equilibrium phase compositions at a given temperature, partial pressure of sulphur

dioxide, magnetite activity, and, in the case of matte-making, matte grade. It does not

attempt to model the entire process. The most important findings of these papers with

respect to the present project relate to impurity removal, particularly removal to the gas

phase. The majority of volatilization occurs during matte-making and the removal rates

of arsenic and antimony are very slow during copper-making. Increasing temperature

was found to have only a slight effect on the removal rates, but calculations were not

done above 1250C. Oxygen enrichment was found to be detrimental to impurity

9
2.2.2 Impurity Distribution

volatilization due to the reduction in gas throughput. The slag composition was found to

have no significant effect. More recently the model has been extended to the flash

furnace and converting with high oxygen enrichment and calcium ferrite slags by Sohn et

al. 16,29-3 1 The results of these calculations will be discussed below. The modelling

technique used in these papers appears to have some problems. In particular, the

prediction that oxygen partial pressure is independent of the 0/Fe ratio in the slag is

suspect.

The other model which assumes a constant temperature appears to be more a

demonstration of a possible use for the Solgasmix program than a bone fide model.
32 A

more recent extension of this model using a modification of Solgasmix has been written

to simulate a concentrate injection smelting technique.


32 It does not include a heat

balance as the process is assumed to be isothermal and the important minor elements

were not considered.

2.2.2 Impurity Distribution

The more recent papers on modelling of the Noranda Process and other new

processes are primarily concerned with the distribution and volatilization of impurities,
263 The distribution of
which are of particular importance in direct smelting
16
systems.

minor elements between copper, matte and slag is calculated assuming that there is only

monatomic dissolution of arsenic, antimony and bismuth in the slag, and that the oxides

25 However, the model significantly underpredicts the amount


and sulphides are unstable.

of each of these elements reporting to the slag. Although the value of L (% As in

copper/% As in slag) is brought close to the industrially reported value using mechanical

10
2.2.2 Impurity Distribution

suspension indices, the calculated values of 14 and 14 remain very high.


25 This indicates

that the oxide form of these minor elements may be an important factor in their

dissolution in slags.

A technique to calculate the relative amount of impurities lost to the off-gas is also

26 The technique calculates the partial pressure of the minor elements and
formulated.

their compounds from various thermodynamic variables, and in particular the partial

pressures of oxygen and sulphur. It should be reiterated that this is an equilibrium model

calculating steady-state volatilization. The overall minor element removal is directly

related to the volume of gas passing through the bath and the final equilibrium

composition of the constituent phases. As such it is not valid for a batch process with a

continuously changing composition. In fact its validity can be questioned for any

submerged injection process, including the Noranda process for which it was written, as

it is not likely that the gas leaving the bath close to the charging end of the reactor will be

in equilibrium with the final products. The equilibrium calculations do give an indication

of which compounds are likely to be the most important with respect to minor element

removal to the gas. However, when this model is compared to industrial observations

from the Noranda process it significantly under predicts the amounts of arsenic and

antimony reporting to the off-gas during copper-making, although it is closer for

27 The under prediction during copper-making is the result of the


slag-making.

equilibrium assumption. Both arsenic and antimony have very low activity coefficients

in copper. Under equilibrium this causes the activities of these elements to drop

significantly in all phases when copper is present in the system. This lower activity

causes a reduction in the volatilization rate and, hence, an under prediction of the

amounts of these elements reporting to the off gas.

11
2.3 Process Kinetics
The same model has been applied to copper converting by considering the process

as a series of microsteps Overall it is a simplistic model which does not allow for

changing temperatures or charging to and skimming from the bath. The results of the

model agree with industrial data to a certain extent, but the amount of data available is

very limited. The effects of increasing the oxygen enrichment were also studied,
and
3

found to have a very detrimental effect on minor element removal. This result is

expected due to the reduced gas throughput. The effects of oxygen enrichment and the

use of a calcium ferrite slag on the overall minor element distribution have also been

predicted with this model.


31 In general it was predicted that increasing oxygen

6

enrichment reduced the amount of volatilization due to the reduced volume of off gas

produced. However, the amount of minor elements reporting to the slag was increased.

The overall result was an increased minor element content in the copper.
The use of a
3

calcium ferrite slag was predicted to have little effect on the elimination of lead and

bismuth, but to be extremely detrimental in the removal of arsenic and antimony due to a

large reduction in the extent of volatilization.


6

An attempt to theoretically model the distribution of impurities between matte and

blister copper in the copper converter using the Temkin model has been carried out.
34
The model gives good results for some of the impurity elements, but is quite poor for

predicting arsenic and antimony distributions.

2.3 Process Kinetics

The kinetics of some of the processes involved are important, particularly with

respect to the minor element removal. Therefore, a review of a variety of aspects relating

to the process kinetics is required.

12
2.3.1 Copper Converting Kinetics

2.3.1 Copper Converting Kinetics

The kinetics of copper converting have undergone little study. This is due to the

apparently high oxygen efficiency which suggests that it operates close to equilibrium.

Calculations by Ashman et al.


35 have shown that the reactions in both the slag and copper

blows are under gas-phase mass-transfer control. The high oxygen efficiencies

suggest,therefore, that the gas residence time in the bath is sufficient to allow the

reactions to come close to or attain equilibrium.

A more recent study of the oxidation of molten iron sulphide assumed that the rate

was under mixed control through the gas and liquid boundary layers around a bubble.
36

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

significant portion of the reaction occurs during bubble growth.


37 Unfortunately, this

study neglected the effect of mass-transfer at the bath surface, which can be considerable,

especially at low tuyere submergences.


38

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,


or
4

39

concern the controlling stage of the chemical reaction,


42 which is not relevant, since the

41

reaction is considerably faster than the mass-transfer in the system.

Although it does not concern the converter directly, a study of the removal of

bismuth and lead from copper during vacuum induction melting does highlight a factor

which could be of considerable importance in converting.


43 It was determined that the

removal rates of these elements were controlled by liquid-phase mass-transfer and

evaporation. This suggests that the removal of the minor elements during converting will

13
2.3.2 Kinetics of Gas/Liquid Reactions

also be controlled by liquid-phase mass-transfer and evaporation and will not be well

represented by equilibrium calculations. This conclusion is supported by a study of

impurity removal from copper mattes.

2.3.2 Kinetics of Gas/Liquid Reactions

There have been some studies relating to the mass-transfer around submerged gas

3
37 8 These were primarily low temperature physical models, a majority dealing
jets.
4
55

with the reaction of CO


2 with an NaOH solution.
4648 Two studies have considered the
kinetics of the deoxidation of copper by carbon monoxide.
The first determined that
5

49

the deoxidation rate was controlled by gas-phase mass-transfer down to an oxygen

concentration of 0.1%, below which it was under liquid-film control.


48 The second study

found that liquid-phase control remained in effect down to 0.005% oxygen, below which

the reaction rate became significant.

A model of mass-transfer between a horizontal jet and a liquid was combined with a

physical model injecting a mixture of air and SO


2 into a bath containing 2
H
4
.
0 5 The
experimental work was designed to ensure that the process was limited by gas-phase

mass transfer. For this work it was assumed that the gas flow could be represented by an

expanding jet, so the results are given as k, where x is the interfacial area per unit

length of the jet. Unfortunately, the lowest modified 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 modelling used the absorption

2 in water during vertical injection. In this case the rate of absorption was
of CO

controlled by liquid-phase mass transfer, which was represented in the model using

Higbie s penetration theory. One of the more interesting results of this study was the

14
2.3.3 Kinetics of Liquid/Liquid Reactions

determination of the effect of surface reactions. In particular, it was found that at low

flow rates and low tuyere submergences surface reactions could account for well over

50% of the total CO


2 absorbed.
38

2.3.3 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

51 All of these are studies related to bottom injection, particularly ladle


reviewed by Mon.

refining, and have determined that the rate of mass-transfer is dependent upon the gas

flow rate. There are three mass-transfer regimes, depending upon the gas flow rate:

I. at low flow rates there is a slow increase in the mass-transfer coefficient with

increasing gas flow, probably related to the increase in bath mixing velocity,

II. above a critical gas flow rate there is a rapid increase in mass-transfer rates, due

to the formation of droplets of slag in the metal, and

Ill, with further increases in gas flow the increase in mass-transfer rate is reduced,

since the entire slag has been emulsified.


5254

The values of the critical flow rates are dependent upon the slag depth, reactor

dimensions, and tuyere position, as well as the physical properties of the phases

54 As such, the values obtained in physical modelling studies are not


involved.

transferable directly to the copper converter, however the three mass-transfer regimes are

likely to be found in the converter. Also, the observation that off-centre tuyeres have a

considerably greater critical velocity for emulsion formation than central tuyeres,

suggests that the conditions in a Peirce-Smith converter are far from ideal for matte-slag

15
2.4.1 Matte Thermodynamics

mass-transfer. A large stagnant region of slag was also observed in a two-phase

physical model with an off-centre tuyere,


55 again suggesting that the mass-transfer

conditions in the Peirce-Smith converter are far from ideal.

2.4 Process Thermodynamics

2.4.1 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 use of a Cu-Fe-S-O quaternary

system appears to be an unnecessary complication, since the oxygen appears to be present

as dissolved iron oxides. The first reported thermodynamic measurements on this system

were carried out by Krivsky and Schuhmann in 1957.56 They determined that the

S-FeS pseudobinary deviated negatively from ideality. Later researchers determined


2
Cu

that the Temkin


57 or Flood
58 models could be used to calculate activities within the

psuedobinary system, but that the addition of oxygen invalidated their use.
58 Work

carried out by Sinha and Nagamori determined the activity coefficients of cobalt, iron,

and their suiphides in copper-saturated mattes.


59 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 et al. 17-20 derived a set of

4 in the matte. These were


equations for the activity coefficients of FeS, FeO, and Fe
0
3

Rosenqvist and Hartvig,


derived from the data of Korakas,
6 and Kuxmann and Bor,
6 62

7 The relationship obtained for the activity


and relate to mattes saturated with magnetite)

coefficient of FeS

16
2.4.1 Matte Thermodynamics

= exp((
18
J ln(O.54 +1 4XFeS logX + 0.52XFeS)J

is plotted in Figure 2.1. The same figure shows the experimental points of Sinha and

59 in copper saturated matte. It is evident that the agreement is very poor,


Nagamori

demonstrating the extreme difficulty which exists in characterizing the solution

thermodynamics of mattes.

1.1

Kemori et al. [20]


1400 K
- 1500 K
0.9 Sinha and Nagamori [59]
1400K
x 1500K
0.8
*

U) 0.7
cj
U
xx
0.6

0.5

0.4

0.3
0 0.2 0.4 0.6 0.8

X FeS

Figure 2.1 Comparison of reported values of FeS activity coefficient in copper


mattes.

Attempts to represent copper and nickel mattes using the associated solution model

6377 These models appear to have a good predictive capability in


have been carried out.

binary and ternary systems, but require a large number of fitted parameters, and have not

been extended sufficiently to allow their use for industrial mattes.

17
2.4.2 Slag Thermodynamics

2.4.2 Slag Thermodynamics

There has been a considerable amount written regarding slag thermodynamics, and

an extensive review was produced by Mackey in 1980.78 More recent studies appear to

have concentrated on ferrite slags, as are used in the Mitsubishi process.


7983 The activity

coefficients for the major constituents in fayalite slags have been calculated by Goto
17

from the data of Muan,


84 Michal and Schuhmann,
85 and Korakas.
An associated
6

solution model of the Fe-O-Si0


-CaO system has also been produced.
2 86

85

2.4.3 Internal Phase Equilibrium

The distribution of various elements between the matte and the slag at equilibrium

is an important topic, since it 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 large body of papers written about the subject.
8895

The form of the copper in the slag appears to depend primarily upon the matte

88 At low matte grades dissolution as a sulphide predominates, while oxidic


grade.

dissolution becomes important as the matte grade increases. It would probably be more

correct to consider the form of dissolution as a function of oxygen potential, with oxidic

dissolution being favoured by the high oxygen potentials which exist with high matte

96 In fact, measurements made in ferrite slags show a large variation in


grades. with

82 In order to deal with this the Temkin or Herasymeriko models


oxygen partial pressure.
95 The Herasymenko model has been shown to represent
have been applied to the slag.

copper slags well, including the transition between sulphidic and oxidic dissolution of

copper.

18
3 OBJECTIVES AND SCOPE
3 OBJECTIVES AND SCOPE

The persistence of the copper converter indicates that it fulfils its function

satisfactorily. However, it has a number of drawbacks. All the elements required to

develop a new, efficient process are in existence, yet little has been done. One of the

most important problems which has to be overcome before oxygen can be effectively

utilized in a coppermaking process is that of minor element removal. In the Peirce-Smith

converter most of the arsenic, antimony, bismuth and lead are removed under present

practices, however it has been found that under direct coppermaldng conditions there is

insufficient removal of these elements. Also, previous models of minor element removal

indicate that using oxygen injection will exacerbate the problem.

The main objective of this work is to develop an improved understanding of the

operation of copper and nickel converters. In particular, a better understanding of the

mechanism by which minor elements are removed from the converter is to be obtained,

and so aid in the development of an improved overall process. The emphasis is to be

placed on the behaviour of the more important minor elements: arsenic, antimony,

bismuth, lead, and zinc. To aid in attaining this goal a kinetic model of Peirce-Smith

converter operation will be developed and validated. The model may then be used to

predict the effects of altering process variables on overall converter operation, as well as

minor element behaviour.

Equilibrium models of both the copper and nickel converters have been previously

developed, and appear to successfully represent converter operation. However, oxygen

potential measurements carried out in an operating converter indicated that the three

condensed phases were not in equilibrium.


96 It has been determined that the rates of the

primary converting reactions are controlled by gas-phase mass-transfer. This is dealt

19
3 OBJECTWES AND SCOPE

with in the thermodynamic models by assuming that a certain proportion of the oxygen

reacts with the bath. However, the actual amount of oxygen reacting may vary

considerably within the course of a charge, and possibly even within a blow, so assuming

a single value may lead to incorrect conclusions. Finally, when considering the

behaviour of the minor elements, the kinetics are likely to become more important due to

the relatively low concentrations in the liquid phases.

Validation of the model will be carried out using data obtained from operating

converters. Validation will cover nickel converting operation, using data obtained during

plant trials at Incos Copper Cliff smelter, and copper converting using data obtained

from plant trials at Hudsons Bay Mining and Smeltings Fun Flon smelter. Following

validation, the model will be applied to determining the effects of process variables on

converter operation. These will concentrate primarily upon factors which can be changed

directely in converting practice, as well as simulating conditions which are not normally

obtainable in a standard converter, but may result in a significant improvements in overall

operation. The effects of these variables on the distribution of the minor elements will

also be considered. A comparison with a previous equilibrium model of the nickel

converter will also be made.

20
4.1 Copper Converter Trials

4 INDUSTRIAL DATA

4.1 Copper Converter Trials

To obtain further information regarding converter operation, and data for use in

model verification, in-plant sampling was carried out at Hudsons Bay Mining and

Smeltings Fun Flon smelter between January 31 and February 5, 1994. Two complete

charges and the slag blows of a third charge were followed in detail, taking samples of

most materials added to and removed from the converters. Occasional tuyere samples

were also taken throughout each charge. The times of all sampling, as well as blowing

and idle periods were recorded, along with the amount of air blown. Since this particular

installation does not routinely monitor the bath temperature, temperature measurements

were also carried out, using disposable thermocouples inserted into a tuyere. Off gas

compositions were not measured, as facilities for this were not present, and the prescence

of leakage air makes accurate interpretation of such measurements difficult.

Details of the charges followed are given in Table IV-I, and the assays of the

samples obtained are given in Tables IV-II to IV-V. The assay techniques used to obtain

these values are given in Table IV-VI. Approximate magnetite contents of all samples

were measured using an Inspiration Consolidated Copper Company Magnetite Meter,

which is based on magnetic inductance.The magnetite meter was calibrated using

standard samples. None of the charges followed can be considered normal. This is

because the other operating converter was in its final charges before relining, so the

white metal was transferred to the test converter before the copper blow. This provided

an unusual input stream for the two full charges followed, leading to shorter overall

charge lengths. The final charge followed, charge 595, was the second charge in a newly

relined converter. While no material was transferred to the converter during this charge,

21
4.1 Copper Converter Trials

charge 586 588 595


nitial Contents Matte 132+80(1) vlatte 48 Vlatte 148
tonnes) Slag 12 Slag 8 Slag 8
Scrap 4 Scrap 4 Scrap 4
ilow 1 2 C 1 2 C 1 2 3 4 5
otal Time 113 100 241 118 128 254 126 92 64 164 84
(mm)
Eime Idle (mm) 50 34 28 44 66 22 78 60 46 56 14
fldle Periods 7 5 3 5 5 5 4 4 1 9 3
vlatte Added 16 32 - 16 32 - 16 48 32 32 32
tonnes) 72(1)
1ux Added 5.5 7.7 - 6.6 9.9 - 6.6 4.4 - 9.9 8.25
tonnes)
Scrap Added - - 15 - - 15 12 25(2) - - -

tonnes)
Slag Skimmed 31.25 37.5 25 31.25 43.75 18.75 0 18.75 12.5 62.5 25
tonnes)
(1) Transferred matte
(2) Copper slag

Table TV-I Details of copper charges followed.

the amount of flux added in the first blows was insufficient, as can be seen from the first

two slag assays in Table TV-TV. The low silica and corresponding high magnetite content

of the slag caused it to become very viscous. This can also be seen in the high copper

and sulphur contents of the slag samples, indicating a large amount of matte entrainment.

The fluxing problems are also the main cause of the large amount of slag in the

first matte samples and their correspondingly high magnetite contents. While the matte

samples taken in the first blow of the other charges followed do contain a higher silica

content than the other mattes, they are still considerably lower than the values from

charge 595, and are caused by the lower volume of material in the bath, which drops the

slag level close to the tuyereline. It should be noted that the average converter slag

22
4.1 Copper Converter Trials
composition at the Fun Ron smelter is 26% silica and 38% iron, with magnetite ranging

between 20 and 30%. Only one of the slag samples taken during the trials is close to this

(charge 595, slag 3), but even this slag had a high magnetite content.

Sample Cu Fe Zn Pb S 2
Si0 4
0
3
Fe As
(Blow)
Initial Matte 42.8 25.2 2.78 1.08 24.5 .21 9.0 0.05
Matte 1 (1) 70.9 4.1 0.83 2.89 20.2 2.33 2.5 0.03
Matte 2 (2) 71.0 3.09 0.59 2.63 21.6 0.28 1.5 0.02
Matte 3 (2) 74.7 1.1 0.3 0.65 19.8 0.5 1.2 0.01
Transfer Matte 65.6 7.3 1.01 1.13 17.9 2.25 11.0 0.03
Metal (C) 90.1 0.49 0.13 0.23 5.55 0.22 0.0 0.04
Blister Copper 99.04 0.08 0.002 1 0.0072 - - - 0.0073
Slag 1 (1) 7.13 41.6 3.97 0.7 2.2 17.6 66.0 0.04
Slag 2 (2) 6.46 39.2 3.89 0.73 2.07 22.0 53.0 0.04
Slag 3 (2) 6.13 37.4 4.00 1.28 1.66 22.8 52.0 0.04
Copper Slag 38.2 21.8 2.16 2.84 0.6 12.1 21.5 0.04
Table TV-il Material assays, copper converter charge 586, Feb. 1, 1994. Numbers
in parentheses refer to the blow number in Table IV-I.

23
4.1 Copper Converter Trials

Sample Cu Fe Zn Pb S 2
Si0 4
0
3
Fe As
(Blow)
Initial Matte 42.2 23.7 2.81 0.91 25.1 0.29 8.0 0.05
Matte 1 (1) 23.2 38.6 4.82 0.47 7.7 14.3 24.2 0.03
Matte 2 (2) 70.3 2.87 0.66 0.51 20.8 0.19 1.4 0.03
Matte 3 (2) 74.0 1.43 0.28 0.4 19.7 0.25 1.5 0.02
Transfer Matte 71.2 2.88 0.54 0.67 20.0 0.31 1.9 0.03
Blister Copper 97.89 0.67 0.0044 0.0072 - - - 0.0096
Slag 1 (1) 5.35 45.1 4.51 0.67 1.54 21.8 44.0 0.05
Slag 2 (2) 7.04 36.5 3.82 0.95 2.39 23.2 37.0 0.05
Copper Slag 1 31.4 22.3 2.48 3.21 0.64 18.3 44.5 0.05
Copper Slag 2 41.5 20.1 2.16 2.55 0.33 14.0 47.0 0.04
Table IV-ffl Material assays, copper converter charge 588, Feb. 2, 1994. Numbers
in parentheses refer to the blow number in Table TV-I.

24
4.1 Copper Converter Trials

Sample Cu Fe Zn Pb S 2 Fe
SiC) 4
0
3 As Sb Bi
(Blow)
Initial Matte 42.2 20.8 2.59 1.35 22.5 0.53 7.5 0.034 .0045 .0032
Matte 1 (1) 37.7 23.3 2.75 1.00 15.1 7.43 23.2 0.022
Matte 2 (1) 31.2 25.6 2.95 0.8 11.1 9.62 28.5 0.019
Matte 3 (1) 57.1 11.8 1.6 1.17 17.9 4.33 10.0 0.014 .0028 .0013
Matte 4 (2) 49.8 14.5 1.83 1.01 17.1 4.24 12.0 0.02
Matte 5 (2) 58.4 9.01 1.31 1.13 18.4 2.96 6.0 0.014
Matte 6 (3) 63.2 7.43 1.33 1.21 20.1 0.83 3.8 0.018 .0022 .0014
Matte 7 (4) 58.5 8.89 1.48 1.18 20.0 0.72 3.0 0.016
Matte 8 (4) 59.0 8.45 1.45 1.32 21.2 0.45 4.0 0.017
Matte 9 (4) 62.7 7.87 1.3 1.15 20.3 1.03 3.2 0.015 .0019 .0002
Matte 10 (5) 65.7 4.15 0.75 0.89 19.9 1.06 2.2 0.012
Matte 11 (5) 62.8 4.39 0.82 0.73 17.3 2.55 3.6 0.01
Slag 1 (2) 8.1 39.8 3.93 0.8 2.54 18.8 47.0 0.015
Slag 2 (3) 12.4 38.3 3.8 0.7 4.71 17.8 41.5 0.022 .0058 .0007
Slag 3 (4) 4.16 38.2 4.38 1.45 1.19 25.5 34 0.012
Slag 4 (5) 3.52 40.6 4.44 2.06 0.67 24.3 48.5 0.016 .0057 .0002
Table IV-IV Material assays, copper converter charge 595, Feb. 4, 1994. Numbers
in parentheses refer to the blow number in Table IV-I.

2
Si0 Fe CaO 3
0
2
A1 0
2
H
Flux I 75.6 2.2 1.4 8.2 I 7.0
Table TV-V Flux assay, copper converter plant trials, Feb., 1994.

25
4.2 Nickel Converter Trials

Sample Element Assay Method


Matte, White Metal, Cu, Zn, Fe, -NaCO fusion with LC.P. finish
0
2
Na
3
Slag, Dusts S, Si0
2
As 4 digestion with AAS finish
HC1O
Sb, Bi High: Distillation and I.C.P.
Low: Digestion with graphite furnace
analysis
Blister Copper Cu Digestion, interference removal and
electroplating finish
Zn Acid digestion with AAS finish
Pb, As, Fe, Digestion, hydroxide precipitation,
Bi, Sb separation, redigestion with AAS finish
Flux 2
Si0 0 fusion with Si0
2
NaOH-Na 2 dehydration
Fe NaOH-Na fusion, Fe
0
2 3 precipitation and
concentrated 4
SO titration
NH
CaO, A1
3
0
2 HF, HC1O
4 and HC1 digestion with AAS
Table IV-VI Methods of determination for each assay.

4.2 Nickel Converter Trials

Data similar to that obtained from the copper converters is also available for a total

of four complete charges of a nickel converter. These data were obtained during a

previous study at the Copper Cliff smelter of Inco Ltd.


97 In this case both the air rate and

bath surface temperature were measured constantly, and recorded on the smelter

computer at one minute intervals. These two variables are plotted for one charge in

Figure 4.1. This figure shows that there is considerable variation of both temperature and

air rate throughout a charge. Through a single blow there is usually an initial temperature

drop corresponding to flux addition, followed by a relatively continuous rise up to the

end of the blow. Interruptions to the steady temperature increase may be due to matte or

26
4.2 Nickel Converter Trials

scrap additions, or enforced idle times required by environmental considerations. The air

rate usually follows a similar pattern through a blow, with a rapid initial drop related to

flux addition, followed by a steady increase.

Details of the charges followed are given in Table IV-Vll, and assays of the primary

components are given in Tables IV-VIll to IV-XI. It can be seen from a comparison of

these tables with Tables TV-IT to IV-IV that there is more composition data available from

the copper converter, while the continuous temperature measurements in the nickel

converter gives considerably more information regarding the heat balance.

Table TV-VU shows that there is considerable variation between charges, with a

number of different sources of converter inputs. The relatively large number of operating

converters at Copper Cliff increases the potential for inter-converter transfers of both

matte and slag, as well as transfers of material with a high copper content from the

copper converters which are situated in the same converter aisle. It is also evident that

considerably more scrap is used in nickel converting than in copper converting. This is

due to the larger amount of iron present in the nickel reverberatory furnace matte, which

causes an increased amount of heat generation.

27
4.2 Nickel Converter Trials

1550

1500

w
D

1450

Ui
I

1400

700

600

z
Ui

500

400
0 200 400 600 800
TIME (mm)

Figure 4.1 Variation in measured bath temperature and air rate in a nickel
converter, (#3 converter, charge 105, May 23, 1988).

28
H

- D
c.
.-.- C)
!
CD CD
CD C
0 a CD
I-i) 0
E
a CD c C,
C,) s _- C,, -
1
CD
u IIIIIiH
CD
0

CD

C i

z

CD
,,
---
00 C
I I I I I I

I-
%__

1:3

I I
00 i
Ji C C
cc)

%

I
4.2 Nickel Converter Trials

Charge 105 106 107 108


5 total time (mm) 126 74
time idle (mm) 30 30
# of idle periods 3 2
matte added (tonnes) 20+40(1) 36(2)
flux added (tonnes) 18.2 18.2
scrap added (tonnes) 10 4.5
slag skimmed (tonnes) 36 36
6 total time (mm) 64 48
time idle (mm) 22 18
# of idle periods 1 1
matte added (tonnes) - -

flux added (tonnes) 18.2 7.3


scrap added (tonnes) - -

slag skimmed (tonnes) 36 18


7 total time (mm) 18 26
time idle (mm) - -

# of idle periods - -

matte added (tonnes) - -

flux added (tonnes) 20 20


scrap added (tonnes) 15 10
slag skimmed (tonnes) - -

(1) Transferred Matte


(2) Copper converter washout

Table IV-Vll Details of nickel converter charges (continued).

30
4.2 Nickel Converter Trials

Sample Cu Ni Co Fe S 2
Si0 Pb Zn As
(Blow)
F. Matte 1 8.32 25.6 1.16 33.9 28.0 0.0894 0.104 0.0272 0.0202
F. Matte 2 8.47 21.8 0.85 37.5 27.9 0.0638 0.11 0.0414 0.0139
C. Matte 1 (1) 13.8 38.8 1.6 16.9 25.2 0.181 0.080 0.0116 0.0202
C. Matte 2 (4) 15.4 42.5 1.45 11.4 24.9 0.0618 0.071 0.0043 0.0218
Bess. Matte 23.5 48.3 0.949 0.792 22.4 <.0079 0.0625 0.003 0.0252
Slag 1 0.568 0.972 0.591 51.1 1.95 24.4 0.0419 0.0555 <.0115
Slag 2 0.487 0.77 0.536 51.7 1.75 24.1 0.0352 0.0574 <.0113
Slag 3 0.496 1.02 0.605 50.6 1.39 26.7 0.0414 0.0552 <.0116
Slag 4 0.444 0.90 0.675 49.7 1.01 29.5 0.0497 0.052 <.0114
Slag 5 1.43 3.39 1.14 48.6 1.39 26.0 0.0585 0.0467 <.0115
Slag 6 1.07 3.35 1.55 44.6 0.619 30.0 0.07770.0401 <.0119
Table TV-Vu Material assays, nickel converter charge 105, May 2, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Numbers in
parentheses refer to the blow number in Table IV-Vll.

Sample Cu Ni Co Fe S 2
Si0 Pb Zn As
F. Matte 1 8.52 27.7 1.19 33.8 28.4 0.0519 0.0957 0.0242 0.0133
F. Matte 2 7.25 27.5 1.04 33.2 27.4 0.0918 0.114 0.0224 0.0181
C. Matte 1 (2) 18.4 42.5 1.27 9.83 24.1 0.165 0.0693 <.0024 0.0165
C. Matte 2 (3) 21.7 46.8 1.03 5.79 22.5 0.0591 0.0853 <.0025 0.0206
Bess. Matte 24.1 51.4 0.789 0.956 21.8 0.0247 0.058 <.0023 0.0225
Slag 1 1.1 2.28 0.788 49.8 1.56 24.2 0.0494 0.054 <.0118
Slag 2 0.863 2.6 1.37 51.1 .55 26.9 0.0607 0.0442 <.0118
Slag 3 0.712 1.17 0.515 55.6 2.15 20.4 0.0363 0.0792 <.012
Slag 4 1.23 3.95 1.56 46.0 0.581 26.8 0.0833 0.0418 <.0111
Table IV-IX Material assays, nickel converter charge 106, May 24, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Numbers in
parentheses refer to the blow number in Table IV-Vll.

31
4.2 Nickel Converter Trials

Sample Cu Ni Co Fe S 2
Si0 Pb Zn As
F. Matte 1 6.6 25.9 1.06 34.6 27.6 0.0994 0.0976 0.0246 0.0168
F. Matte 2 8.34 24.1 0.943 34.5 27.4 0.136 0.0922 0.0337 0.0154
C. Matte (2) 14.7 42.8 1.53 14.6 25.5 0.0726 0.085 0.0040 0.02
Bess. Matte 21.9 51.7 1.1 1.17 23.0 0.13 0.0634 <.0024 0.0352
Slag 1 0.759 1.39 0.545 55.2 2.62 21.1 0.0334 0.0611 <.0105
Slag 2 0.54 0.99 0.602 52.6 1.57 24.8 0.0346 0.0505 <.0113
Slag 3 1.72 4.37 0.754 48.6 3.02 23.9 0.0375 0.0442 <.012
Slag 4 1.32 3.46 0.787 48.0 1.97 26.2 0.042 0.047 <.012
Slag 5 1.85 4.79 1.09 42.7 2.23 29.9 0.0637 0.0362 <.012
Slag 6 2.22 6.00 1.36 39.3 1.96 31.9 0.0671 0.0328 <.0116
Table IV-X Material assays, nickel converter charge 107, May 25, 1988. Slag
samples numbers correspond to the blow they were skimmed after. Numbers in
parentheses refer to the blow number in Table IV-Vfl.

Sample Cu Ni Co Fe S 2
Si0 Pb Zn As
F. Matte 1 6.62 27.4 1.27 32.8 27.8 0.108 0.101 0.0176 0.0176
F. Matte 2 9.48 24.2 0.833 34.1 27.8 0.08 14 0.138 0.0435 0.0148
C. Matte 1 (1) 14.8 45.5 1.53 12.5 25.5 0.083 0.0916 0.005 0.0263
C. Matte 2 (3) 14.9 44.9 1.39 12.4 25.2 0.253 0.0813 0.0074 0.0243
Bess. Matte 19.5 55.6 1.05 1.1 23.0 0.0606 0.0613 <.0025 0.0326
Slag 1 0.952 2.27 0.663 50.4 1.72 24.1 0.0464 0.0501 <.0119
Slag 2 0.637 1.54 0.778 49.2 1.07 28.1 0.0543 0.0469 <.0119
Slag 3 0.862 2.82 1.31 46.8 0.716 28.8 0.064 0.0426 <.0118
Table IV-XI Material assays, nickel converter charge 108, May 26, 1988. Slag
samples numbers correspond to the blow they were skimmed after.

32
4.3.1 Oxygen Efficiency
4.3 Analysis

4.3.1 Oxygen Efficiency

The oxygen efficiency of non-ferrous converters is usually thought to be very high,

but reported values show a wide variation. In fact, it has been found that there is a

considerable variation within a single operation.


98 Also, there is no way to calculate an

accurate value of oxygen efficiency. Johnson et al. give the equation

% 0 efficiency = 100 x
Theoretical 02 Required to Convert Matte to Blister .. .[4. 1]
02 Blown into Converter

which appears straightforward, but is far from accurate. There are a number of

complicating factors, including an imprecise knowledge of the amount of matte and other

materials which are added to the converter.

Values of oxygen efficiency can be derived from the data collected in the plant

trials by calculating the total amounts of iron removed between matte samples and

comparing it with the amount of air blown during that period. While this procedure may

still suffer from the same problems, especially for cases where the matte samples were

separated by material charging, better values should be obtained where the samples were

taken close together.

Overall oxygen efficiencies calculated from the assays for the copper converter

charges followed are given in Table IV-Xll. Separate values are also given for the

combined slag blows and the copper blow. It is evident that oxygen efficiencies are

considerably lower than generally thought, and that efficiencies are higher in the copper

blow. This has been reported previously,


98 but no explanation has been given.

The oxygen efficiencies calculated from the matte assays are shown in

Table IV-XllI. Not all of the matte assays could be used, due to the high silica contents,

33
4.3.1 Oxygen Efficiency

Charge Oxygen Efficiency (%)


Overall Slag Blows Copper Blow
586 76 55 82
588 69 63 72
595 - 68 -

Table IV-Xll Overall oxygen efficiencies for copper converter charges followed in
plant trials.

indicating a large slag content in the samples. The first oxygen efficiency reported for

each charge in the table is for the period between charge up and the given matte sample.

In all cases the first matte sample taken contained too much slag to allow it to be used.

Charge Matte Oxygen Efficiency (%)


586 2 94
3 48
588 2 75
3 59
595 6 70
7 50
8 67
9 65
10 78
Table IV-XllI Oxygen efficiencies calculated from matte samples.

It can be seen in charges 586 and 588, that there is considerable variation

throughout the charges, with a generally lower efficiency when the iron contents are low.

The lower oxygen efficiencies at low iron contents suggest that the iron removal rate may

come under liquid phase mass-transfer control, with the oxygen beginning to react with

copper sulphide. This does not reduce the overall efficiency, and may cause the

34
4.3.1 Oxygen Efficiency

calculated efficiency in the copper blow to be too high. This is not apparent in charge

595; however, the iron content in these mattes did not drop below 4%, so the reduction in

efficiency is not likely to be seen.

As mentioned above, the values calculated for oxygen efficiency in the nickel

converter can only be approximate, since weights of the mattes at the time of sampling

are not known. Also, both the weights and compositions of other materials added

between sampling were not measured accurately. To determine approximate weights for

the sampled mattes, copper and nickel balances can be carried out using the required

weights as unknowns. The balances provide two equations which can be solved

simultaneously to give the required weights. An iron balance can then give the amount of

iron reacted. Table IV-X1V gives the values of oxygen efficiency calculated in this

manner as well as the overall efficiencies for the charges where inter-sample values could

not be determined.

Charge Matte Oxygen Efficiency (%)


105 - 93
106 1 66
2 98
107 - 90
108 1 88
2 55
Table IV-XIV Calculated nickel converter oxygen efficiencies.

Generally, the calculated oxygen efficiency is higher than in the copper converter.

The most likely explanation for this is the greater tuyere submergence in this nickel

converter, due to its larger size. However, the amount of magnetite in the charged

materials is not considered in the calculations. This increases the calculated values of

35
4.3.2 Material Deportment
iron oxidized, and hence the calculated oxygen efficiencies. The magnitude of this effect

could be large, since the magnetite content in the reverberatory matte can be over 8%.

Unfortunately, insufficient magnetite assays were carried out at the time of the testing to

allow the inclusion of this in the oxygen efficiency calculations.

4.3.2 Material Deportment

4.3.2.1 Major components

There are four major components in the copper converter; copper, iron, sulphur, and

silica. These materials are distributed between the slag, blister copper, dust, and gas in

the proportions given in Table IV-XV. From the table it can be seen that the removal of

iron and sulphur from the matte is almost complete, with the majority of the iron being

removed in the slag, and the sulphur reporting to the gas. It appears that almost all of the

silica added is also removed in the slag, but the amount of flux added is not known with

any accuracy, so the input silica was calculated based on the outputs to the slag and dust.

This does not include the flux which is collected in the dusts beyond the flue, or is blown

out to the atmosphere as the converter turns into stack after flux addition.

Between 80 and 85% of the copper fed to the converters reports to the blister

copper, with the majority of the remainder being removed in the slag. Of the copper in

the slag, about 60% is in the slag removed at the end of the copper blow. This material is

recycled directly to another converter, while the remaining slags and dust are returned to

the reverberatory furnace. The amount of copper in the skimmed slag is relatively

constant, due to the practice of blowing to copper high before skimming. Since the

copper content in the matte is approximately the same at each skim, the copper in the slag

is also. However, the sulphur content of all slags sampled, with the exception of the

copper slags, was more than sufficient to allow all the copper to be present as suiphide,

36
4.3.2 Material Deportment

Charge Stream Cu Fe S 2
Si0
Blister Copper 85.2 0.2 - -

Slag Total 12.0 99.1 4.1 99.4


586 Slag Blows 3.9 82.6 3.7 81.7
Copper Blow 8.1 16.5 0.4 17.7
Dust 2.8 0.6 2.5 0.6
Gas - - 93.4 -

Blister Copper 83.8 1.5 - -

Slag Total 12.3 97.9 4.8 99.8


588 Slag Blows 5.2 86.3 4.5 84.1
CopperBiow 7.1 11.6 0.3 15.7
Dust 3.9 0.6 2.9 0.2
Gas - - 92.3 -

Blister Copper 79.6 0.12 - -

Slag Total 17.8 98.7 4.4 99.0


595 Slag Blows 6.3 86.0 4.0 85.6
Copper Blow 11.5 12.7 0.4 13.4
Dust 2.6 1.2 1.9 1.0
Gas - - 93.7 -

Table IV-XV Distribution of major components between output streams of the


copper converter.

either dissolved or entrained. An indication of the amount of entrainment can be seen

from the slags with a low silica content in charge 595. In these slags the copper content

is close to double the usual amount. This will be almost entirely due to entrainment, and

is caused by the high slag viscosity.

In nickel converting there are two major output streams, the slag and the Bessemer

matte, and one minor stream, the dust. Ideally, all the copper, nickel, and cobalt will

report to the Bessemer matte, while the iron will be removed to the slag. The results of

37
4.3.2 Material Deportment
an analysis of the data collected from the nickel converter are shown in Table IV-XVI.

This shows that the recovery of nickel and copper to the matte is over 90%, while over

95% of the iron reports to the slag.

Charge Phase Cu Ni Co Fe
Slag 4.9 5.3 58.2 97.7
105 Bessemer Matte 91.7 91.1 39.2 0.9
Dust 3.4 3.6 2.7 1.4
Slag 2.8 3.2 48.6 95.4
106 Bessemer Matte 94.4 93.9 48.2 2.2
Dust 2.8 2.9 3.2 2.4
Slag 6.0 6.0 44.3 96.9
107 Bessemmer Matte 91.4 91.2 53.2 1.9
Dust 2.6 2.8 2.5 1.5
Slag 4.8 4.6 49.4 96.0
108 Bessemmer Matte 91.2 92.1 47.6 1.8
Dust 4.0 3.3 3.0 2.2
Table IV-XVI Distribution of major components in nickel converting.

Figure 4.2 shows that the amount of valuable metals lost to the slag increases with

matte grade. The usual explanation for this relates to increased oxidic dissolution of the

valuable metals at higher oxygen partial pressures, which is seen in equilibrium

experiments. However, when the total mole fraction of copper, nickel and cobalt in the

slag is plotted against the oxygen potential of the slag calculated from the 3
/Fe ratio,
2
Fe

the result is Figure 4.3, which does not show any evidence that oxygen potential has an

effect. This suggests that there is another explanation for the increased metal losses. It is

possible that, at lower concentrations of iron in the matte, liquid-phase mass-transfer

becomes important. As the amount of iron decreases there will be a point at which the

mass-transfer rate of iron to the gas-liquid interface becomes less than the rate of oxygen

38
4.3.2 Material Deportment
mass-transfer to the interface. Below this concentration cobalt and nickel will begin to

react, and since both cobalt and nickel oxides are thermodynamically favoured over the

metals, they will report to the slag. Table IV-XVI also shows that close to half of the

cobalt is lost to the slag. However, the relatively small amounts of cobalt present may

lead to considerable errors in these numbers.


.08

.07
z
0
I
.06
U
I
I
o .05
Ui

0
.04
Li)
z

.01
.5 .55 6 .65 .7 .75

Cu+Ni+Co IN MATtE (WEIGHT FRACTION)

Figure 4.2 Variation of total copper, nickel, and cobalt in slag with nickel
converter matte grade.

39
4.3.2 Material Deportment

0.13
0.12
j 0.11
W 01
.
0.09
1
0.08
0.07
LI.
o 0.06
z
Q 0.05
.
I. I
0.04
0.03
0.02
0.01
0 ,. . .,

11 -9 -7 -5
LOG(PARTIAL PRESSURE OF OXYGEN)

Figure 4.3 Variation of total copper, nickel, and cobalt in slag with slag oxygen
potential.

4.3.2.2 Minor elements

The most important minor elements present at the Fun Flon smelter are lead and

zinc. Of the other minor elements usually associated with copper ores, arsenic is the only

one which is present in any appreciable amount. While antimony and bismuth are

present, their concentrations are not sufficient to cause any problems.

The distributions of lead, zinc, and arsenic between the three output streams are

given in Figure 4.4. These values are calculated from the overall mass balances of the

separate charges. For the minor elements, it is assumed that all of the input materials

which are not accounted for in the blister copper or the slag, report to the dusts. In this

case the dusts include those collected in the cottrell precipitators and the baghouse. Both

lead and zinc are removed effectively, with less than 0.5% of the lead and 0.1% of the

zinc reporting to the blister copper. Approximately 10% of the zinc and up to 35% of the

40
_____

4.3.2 Material Deportment


lead are removed in the dusts, with the remainder reporting to the slag. The arsenic

distribution is notably different, with about 10% of the input arsenic reporting to the

blister copper, and 60% being concentrated in the dusts. A similar distribution can be

seen for bismuth in Figure 4.5, while the same figure shows that the proportion of

antimony reporting to the gas is considerably lower, with a correspondingly larger

proportion reporting to the slag.

100
Zn Pb As
90 CHARG
586
+
80 + 588
Lii 0 595
70 +

:i:

40
0.
Ui

30

20
1-

10 10
0 I I I I I I

SLAG BLISTER DUST SLAG BLISTER DUST SLAG BLISTER DUST

Figure 4.4 Minor element distribution in Hudsons Bay Mining and Smelting
copper converters, charges 586, 588, and 595, February 1994.

41
4.3.2 Material Deportment

100

90 ELEMENT
+ Bi +
80 Sb
uJ
70

20
0
10 +

0
SLAG BLISTER DUST

Figure 4.5 Antimony and bismuth distribution in Hudsons Bay Mining and
Smelting copper converter charge 595, February 1994.

As is the case of copper converting at Flin Flon, the only minor elements present in

appreciable quantities in the Copper Cliff nickel converters were lead, zinc, and arsenic.

Of these, lead is the most abundant, but in many cases the concentration of the minor

elements is close to, or below, the lower threshold of the assay technique. The

distributions of these elements between the output streams are given in Figure 4.6. Only

three of the four charges could be used in this case, due to problems with the mass

balance on charge 106. These problems relate to a complete lack of knowledge of the

weight and composition of the mush which is initially present in the converter.

Normally the mass balances can be carried out by assuming that the weight and

composition of the mush is essentially the same after each charge. However, for

charge 106 the amount of every element removed from the converter was considerably

42
4.3.2 Material Deportment
higher than the amount charged, leading to the conclusion that an unusually large

quantity of mush was present at the beginning, or an unusually small quantity remained

at the end.

90
- Pb CHARGE Zn As
80

70 -
+ 107
1
Lu 0 108 +
0
4:
I
0.
o .
1 50
0
z
40
0
.
30 +

20 -

10 .
0
I
0 I I I I

SLAG MATTE DUST SLAG MATTE DUST SLAG MATTE DUST

Figure 4.6 Minor element distribution in Inco nickel converter (#3 converter,
charges 105, 107, and 108, May 1988).

Figure 4.6 shows that the distribution of the lead is relatively even between the

three output streams, with a slightly higher proportion reporting to the dust and a slightly

lower proportion reporting to the slag. This is similar to the distribution in the copper

converter at the end of the slag blows. The zinc distribution is also similar to the copper

converter, with the majority of the zinc being removed in the slag. The arsenic

distribution, however, is quite different, with between 50 and 70% reporting to the matte,

while under 20% is removed with the dust. This is the reverse of the copper converter

distribution, and may suggest that arsenic has a greater affinity for nickel mattes than for

copper mattes. Unfortunately, the accuracy of the assays is not sufficient to allow any

firm conclusions to be drawn.

43
4.3.3 Converter Dusts
4.3.3 Converter Dusts

The analysis of the minor element distribution in the copper converter indicates that

they are concentrated to a certain extent in the converter dusts. During the plant trials a

set of dust samples was obtained. Flue dust samples were taken after each charge, and

during charge 590 four samples were taken from different points along the length of the

bank of Cottrell precipitators, and a sample of baghouse dust was also obtained.

Figure 4.7 shows a schematic of the dust collection system at Flin Flon, and indicates the

approximate location where the Cottrell samples were taken. The assays of these

materials are given in Table IV-XVII. Some interesting trends can be seen in this table.

The concentration of copper drops continuously with distance from the converter, while

those of zinc, lead, arsenic, antimony, and bismuth all increase. The silica content of the

dusts increases initially, with a peak in the two samples taken from the middle of the

Cottrells.

Sample Cu Fe S 2
5i0 Zn Pb As Sb Bi
Flue Dust (c586) 66.7 4.13 18.6 2.19 1.18 1.72 0.03 - -

Flue Dust (c588) 70.5 3.9 20.0 0.97 1.63 0.78 0.05 - -

Flue Dust (c595) 53.6 12.4 19.1 5.12 2.18 1.94 0.019 - -

Flue Dust 62.6 4.81 16.8 5.1 0.95 1.16 0.023 .0081 .0026
Cottrell Dust 1 51.1 4.43 13.1 17.7 1.68 2.0 0.059 .0039 .0027
Cottrell Dust 2 35.5 4.17 10.4 28.7 2.23 5.4 0.12 - -

CottrellDust3 22.2 3.55 8.4 31.8 3.78 5.37 0.145 .0108 .0141
CottrellDust4 20.5 5.52 11.3 15.0 8.72 5.33 0.33
Baghouse Dust 8.28 4.26 14.2 4.0 16.4 12.3 1.01 .0294 .064
Table IV-XVII Compositions of converter dust samples.

44
4.3.3 Converter Dusts
Since a large proportion of some of the minor elements report to the dusts, it is

important to gain a better understanding of how the dusts are formed. It has been

reported that there are four mechanisms by which dusts can be formed in a steelmaking

99 vapourization, metal and slag ejection due to bubbling, and solids entrainment.
vessel;

To this should be added mechanical ejection due to slopping in the converter. Different

mechanisms will be responsible for the removal of different elements. The dust samples

have been analyzed using optical and scanning electron microscopy to aid in determining

which mechanisms are responsible for the removal of the various elements.

FLUE DUST COTTRELL DUST BAGHOUSE DUST

CONVERTER

Figure 4.7 Schematic of the dust collection system at Hudsons Bay Mining and
Smeltings Flin Flon smelter. Numbers indicate approximate position from which
Cottrell dust samples were taken.

4.3.3.1 Flue dust

The wide variation in the flue dust assays shown in Table IV-XVII is related to the

stage of the charge they were formed at. The large amount of flue dust generated does

45
4.3.3 Converter Dusts

not allow the collection of a sample representative of the entire charge, but the separate

samples show the differences which occur during the charge. Generally, samples with

low iron and silica contents will have been produced during the copper blow, while those

with higher silica contents will have been produced during the slag blows.

The converter flue dust is primarily composed of spherical particles mixed with a

few grains of flux. Figure 4.8 is a micrograph of the flue dust and shows a number of

interesting features. It is evident that there is a wide range of particle sizes present, but

most of the particles have diameters between 0.25 and 1 mm. Also, some of the particles

can be seen to have a coating of a white powder, while another shows evidence of a pore.

These features become more evident at higher magnifications. Figure 4.9a shows a

particle coated with what appears to be a layer of condensed fume, while Figure 4.9b

shows a pore in a lightly coated particle. If the particles are mounted and sectioned,

(Figure 4.10) it can be seen that the majority of the spheres are, in fact, hollow. These

pores are most likely caused by dissolved gas coming out of solution during freezing, but

may also be a coating held on small bubbles by surface tension.

46
4.3.3 Converter Dusts

Figure 4.8 Copper converter flue dust, x25.

47
4.3.3 Converter Dusts

a.

b.
Figure 4.9 Copper converter flue dust, a. coated particle, x150, b. particle with
pore, x300.

48
4.3.3 Converter Dusts

Figure 4.10 Copper converter flue dust, mounted and sectioned, x25.

At higher magnifications, the polished sections show some other features.

Figure 4.11 a shows a non-porous particle which contains entrained particles of flux and

what appears to be copper droplets which may have separated out as the material cooled.

Figures 4.1 lb and 4.1 ic also show particles which contain copper. In Figure 4.1 lb. the

copper forms a single droplet at the edge of a larger particle. This is probably a particle

ejected during the copper blow. Figure 4.1 ic shows a slightly porous particle which

appears to be fractured. The central pore and the fractures contain copper, which

probably separated out during solidification. The primary material in this particle also

appears to be a different colour from the particles around it. Wavelength dispersive x-ray

(WDX) analysis of a similar particle (Figure 4.11 d) shows that the darker phase is slag

49
4.3.3 Converter Dusts
with an iron-copper ratio of 3.91. This is considerably higher than is found in the slags

from the slag blows, and lower than slags from the copper blows. However, the

contained particle is almost pure copper, with trace amounts of lead, zinc, and sulphur,

and no measurable iron. This suggests that the material is slag ejected during the copper

blow.

In general, this analysis indicates that the flue dusts are predominantly material

ejected from the bath. The assays show that it is primarily matte, with smaller amounts

of slag, copper and white metal. The differences in the compositions shown in

Table IV-XVII are caused by variations in the proportions of each of these materials

present. The relatively low zinc assays and the lead assays which are slightly higher than

found in matte, suggest that the fume seen is lead based. X-ray diffraction analysis of the

flue dusts showed the presence of lead sulphate, which thermodynamics predicts to be the

predominant phase.

50
4.3.3 Converter Dusts

- .4

s.
.,

..e ,&.

S., -

1
SI .1

a. b.

c. d.
Figure 4.11 Copper converter flue dust, mounted and sectioned, a. particle
containing entrained slag and copper droplets, x125, b. particle containing copper
droplets, xlOO, c. fractured particle containing copper, x125, d. slag particle containing
copper droplet, x80.

51
4.3.3 Converter Dusts

4.3.3.2 Cottrell dust

The Cottrell dust assays indicate that there is an increase in minor element content

with distance from the converter. Figure 4.12 shows samples of the dust from three

positions along the bank of Cottrells. The dust appears to be a mixture of spherical

particles and angular flux particles surrounded by a matrix of fine, powdery material.

The samples taken farther away from the converter show an increase in the proportion of

fine material present, as well as a decrease in the size of the larger particles. This last

point can be seen more clearly in Figure 4.13, which shows the samples with the fine

material removed. While all samples contain small diameter particles, the maximum

particle diameter decreases from 0.5 to 0.1 mm with distance from the converter. The

larger number of angular flux particles in the second Cottrell sample, corresponding to

the increased silica assay, can also be seen in Figure 4. 13b. X-ray diffraction analysis of

the Cottrell dusts shows a predominance of silica and lead sulphate, with a small amount

of As
3 possibly present.
0
2

A previous analysis of Cottrell dusts from the Kidd Creek Mines copper smelter

showed a much higher proportion of lead and zinc than were found in the present case.

These dusts were produced by the Mitsubishi process, so may be considerably different

from converter dusts, however there may still be some similarities, particularly in the

4 was the primary form of lead present,


mineralization. It was determined that PbSO

while zinc was present as oxide, sulphide, and a ferrite. Copper was present in oxide,

ferrite, and arsenide forms. It is possible that the ferrites are present in the copper

converter dusts, although the measured magnetite content of the dusts was below 4% in

all cases. The iron content of the dusts is also low, and, since the zinc and copper ferrites

are also magnetic, they would increase the value of magnetite measured. The sulphur

52
4.3.3 Converter Dusts
contents of the converter Cottrell dusts also indicates that the copper is primarily

combined with sulphur, and x-ray diffraction indicates that it is present as copper

sulphide, which was not found in the Mitsubishi process dusts. It is probable that the

converter dusts contain a much higher proportion of ejected and entrained material than

the Mitsubishi process dusts, which contain more condensed fume leading to higher

minor element contents.

Scanning electron microscopy of the converter Cottrell dusts shows that all of the

particles are coated with the fine material, (Figure 4.14a). As well as the spherical and

angular particles, other small, irregular particles are present. Figure 4. 14b shows one of

these at high magnification. It appears to be an agglomeration of the fine material. WDX

dot-mapping of the Cottrell dusts shows that the majority of the dusts are covered with

zinc oxide, with some areas also showing high concentrations of lead and sulphur. This

confirms that the matrix material is condensed vapour, which has formed on the larger,

ejected particles. The increased amount of fine material with distance from the converter

is caused by a combination of the reduction in gas temperature, causing more material to

condense, and the reduced number of ejected particles which are small enough to remain

entrained in the gas. This conclusion is supported by the assays, and indicates that

vapourization is the primary mechanism for removal of both lead and zinc. The similar

trends in arsenic, antimony, and bismuth concentrations suggests that they are removed

by the same mechanism.

53
4.3.3 Converter Dusts

C.
Figure 4.12 Copper converter Cottrell dust, a. sample 1, xlOO, b. sample 2, xlOO,
c. sample 4, xl 60. Numbers refer to sample positions shown in Figure 4.7.

54
4.3.3 Converter Dusts

C.
Figure 4.13 Copper converter Cottrell dust with matrix removed, a. sample 1,
xlOO, b. sample 2, x160, c. sample 3, x160. Numbers refer to sample positions shown in
Figure 4.7.

55
4.3.3 Converter Dusts

b.
Figure 4.14 Copper converter Cottrell dust, a. sample 2, x300, b. sample 1,
agglomerated fume, x2000. Numbers refer to sample positions shown in Figure 4.7.

56
4.3.3 Converter Dusts

The form of the silica present in the Cottrell dusts indicates that it is removed by the

solids entrainment mechanism. This is also evident from observations of converter

operation. At Fun Flon, flux is added by ladle through the mouth, and a portion of it can

be seen to be blown out of the converter as it turns back into the blowing position. The

distribution of the silica along the Cottrell bank can be attributed to the variation in flux

size fraction.

4.3.3.3 Baghouse dust

The baghouse dust is very similar to the Cottrell dusts in appearance, and follows

the same pattern as was seen in the Cottrells. In fact, the baghouse dust is almost entirely

made up of the fine condensed vapour, with only a very small proportion of silica and

ejected bath material. This can be seen in Figure 4.15. It should be noted that the dust

collected in the baghouse comes from the reverberatory furnace as well as the converters.

However, there does not appear to be a considerable difference between the dusts

reaching the baghouse. Both x-ray diffraction and WDX indicate that the predominant

materials present in the baghouse dust are zinc oxide and lead sulphate, with some silica

stifi present.

57
4.3.3 Converter Dusts

Figure 4.15 Smelter baghouse dust, x1500.

58
5.1 Basic Bubble Growth

5 GAS FLOW j THE BATH

In order to properly model the gas/liquid reaction and mass-transfer it is essential to

have a reasonable representation of the average bubble surface area present at any time.

Unfortunately, little information is available on this subject, particularly with respect to

horizontal injection. While some experimental work has been carried out, in all cases the

lumped parameter kA is the end result. This value is not transferable between systems,

so can only be used to give an idea of trends. However, there is a large body of

theoretical work available concerning vertical injection which can be modified to

represent horizontal injection. This chapter will develop the theory required to give an

adequate representation of the bubble growth and rise in the bath.

5.1 Basic Bubble Growth

In past mathematical models, the gas has been assumed to form either a jet or

bubbles at the tuyere tip. The simplest modelling technique is to assume the gas flows in

a specific, calculable trajectory, and from that derive an average surface area.
2

101

Unfortunately, the assumption upon which it is based is not valid at the gas flow rates in

copper converters. It also neglects the period of bubble growth on the tuyere which

accounts for a considerable portion of the


103 and does not deal well

37
mass-transfer,
4

with jet break-up.

A number of theoretical equations have been developed to represent bubble growth

and
105 and the basic principles of these can be transferred to the situation
35
detachment,

1

of the copper converter. In almost every case the models are developed for vertical

injection, so a direct transfer to this process may not be valid. These models consider

only isothermal and non-reactive conditions so the validity of their direct use would be

59
5.1 Basic Bubble Growth
questionable. A modification of a force-balance type model to the copper converter has

been developed by Ashman et al.


35 which does include both the reaction and heating

effects.

A comparison of the predicted bubble size at detachment and bubble frequency

from a number of different models is given in Table V-TI. The bubble sizes are

calculated for the typical copper converter conditions used by Ashman et al.
35 and are
given in Table V-I.

Air flow rate per tuyere (m


sj
3 0.25
Inlet air temperature (K) 298
Tuyere diameter (m) 0.04
Tuyere submergence (m) 0.4
Bath velocity (m s) 1.8
Bath temperature (K) 1450
Matte density (kg m
)
3 4200
Matte surface tension (N m
)
1 0.4
Gas velocity (m s) 198.94
Fr(-) 2.83

Table V-I Typical copper converting conditions used for bubble size calculations.
35

Physical modelling, however, has shown that for the conditions prevalent in the

converter neither the jet nor the single bubble assumptions are valid. Instead, a

combination of the two appears to be more correct. A schematic of the process observed

for the injection of air into water, with a high Froude number, is shown in Figure 5.1

The figure indicates that there is a short jet formed, from which a bubble breaks off
periodically, as well as bubbles forming above the jet at the tuyere tip. Physical

60
5.1 Basic Bubble Growth

Bubble Bubble Bubble Ref.


volume diameter frequency
(ms) (m) (s)
Ashman et al. (with heating) 0.045 0.441 10.78 35
Ashman et al. (without heating) 0.019 0.331 13.16 35
Davidson et al. 0.066 0.502 3.77 105
Mersmann (K=10) 0.026 0.370 9.46 107
Mersmann (K=26) 0.042 0.433 5.87 107
Wraith 0.052 0.465 4.76 108
Wraith et al. 0.074 0.52 1 3.38 109
Calderbank 0.0125 0.288 20 110

Table V-TI Bubble size at detachment calculated from different models.

modelling studies at lower Froude numbers also show an elliptical gas volume fraction

distribution in the horizontal plane. To allow the modelling of this process the simple

bubbling assumption must be modified. At low Froude numbers it is unlikely that the gas

will have sufficient momentum to form a significant jet. However, there should be

sufficient to distort the bubble to an ellipsoidal shape.


112 Thus in the following

development it will be assumed that the bubble is an ellipsoid, with the axes

perpendicular to the tuyere being equal.

To calculate the penetration of the bubble into the bath (2b) a simple force balance

may be carried out:

61
5.1 Basic Bubble Growth

1=0 O.013s.

0.026s. 0.039s.

0.078s. O.091s.

Figure 5.1 Schematic of high Froude number injection of air into water.

11

d db
PgQo =(Mv)+6ivjir--

where

db
V
dt

2
and

+3a
2
a(8b
) ...[5.31
M PBQt
(3b a)
2
8b

Combining these results in the second order differential equation

62
5.1 Basic Bubble Growth

t+(KpBQ+6i_P-_o ...[5.4]
2 dt1
dt KPBQ ) PBK
where

+3a
2
a(8b
)
K
(3b a)
2
8b

is assumed to be constant. If it is also assumed that the viscosity term is negligible, the

general solution

pvt
b lnt+k
1
k
2
= .

is obtained, where k
1 and k
2 are integration constants. Unfortunately this solution is not

valid at t=O, so is not very useful.

To provide a useful solution to equation 5.4 a numerical approach may be used.

Figure 5.2 shows the calculated growth of a bubble under the conditions given in

Table V-I, neglecting the effects of temperature and buoyancy.

To include temperature and buoyancy, the model of Ashman et al.


35 provides a

good basis. Although the model was developed assuming spherical bubbles it may be

altered to reflect the variation in bubble shape. Other modifications will also be required

to give a more accurate representation of bubble growth.

In the original model it was assumed that the drag coefficient was a constant with

the value 0.4. This was based on correlations for a solid sphere at a high Reynolds

number. In the case of the copper converter this choice of correlation is not valid.

However, the sensitivity analysis did indicate that the drag coefficient had little effect on

the bubble size at detachment, so the assumption may not be important. What is more

63
5.1 Basic Bubble Growth

5 cm

Figure 5.2 Ellipsoidal bubble growth.

important is the main implication of this; that the drag term in the overall equation is

negligible. This result is probably caused by the high value of bath circulation velocity

used in the model. In the equation used by Ashman et 35


al.

Vg + CD VB) = K- [V(vb 1
yB)

the overall drag term is a function of (yb-yB)


2 which will be very small at high bath

velocities. Measurements of fluid flow in a 1:4 scale water model of the Peirce-Smith

converter using Laser Doppler Velocimetry have determined that the maximum flow rate

occurs at the bath surface, with the flow rate close to the tuyere being approximately half

of the maximum value.


113 In the water model, the highest reported velocity was

.163 ms, with a gas velocity at the tuyere of 79.2 ms


. This indicates that, while the
1

bath velocity in the actual converter will be higher than the measured value, it is unlikely

to reach the 3.6 ms required to give a velocity of 1.8 ms at the tuyereline.

64
5.1 Basic Bubble Growth
Other aspects of the model which must be modified are the apparent lack of

consideration of horizontal motion, and the bubble detachment criterion. In particular,

the absence of horizontal motion results in the erroneous equation

ds

where s is the distance from the centre of the tuyere tip to the centre of the bubble. This

equation is valid for vertical injection, where the bubble rises directly above the centre of

the tuyere, but this is not the case in horizontal injection, particularly with an ellipsoidal

bubble. Figure 5.3 compares the two cases, and it is evident that for a rapidly expanding

bubble equation 5.8 cannot be used, unless it is only considering the vertical component

of the velocity in which case s should be replaced by h, the height of the bubble centre

above the tuyere centreline. While this differentiation may appear trivial, it may cause

considerable errors in the determination of the bubble height, which will affect the bubble

detachment calculation.

A. B.

Figure 5.3 Comparison of horizontal and vertical injection conditions.

65
5.1 Basic Bubble Growth
If the above considerations are taken into account the model of Ashman et at.
35 can
be used as a basis for a more complex model. In doing so a number of the assumptions

made in the original model must be carried through to the new one, however, some may

be discarded or altered. Those which are still required are:


4

35
1) the bubbles are not affected by those forming on adjacent tuyeres;

2) there is no coalescence of bubbles from the same tuyere;

3) the gas flow through the tuyeres is constant;

4) at detachment a volume of gas equal to a hemisphere of diameter d remains on the

tuyere;

5) the bath viscosity and turbulence result in a drag force on the bubble;

6) gas surface tension effects are negligible;

7) the gas is ideal;

8) the bubble wall is a black body and

9) only hydrostatic pressure is exerted on the bubble.

The assumption of no interaction between adjacent tuyeres is obviously incorrect, as it

has been ascertained that there is tuyere interaction during converting,


4 but this
interaction appears to take place primarily during the copper blow and rarely involves

5 The effect of interaction on bubble growth and detachment are


more than two tuyeres.

not well known, however observations of interacting tuyeres in both physical models and

operating converters indicate that interaction has little effect on bubble frequency. The

tuyeres have a common gas feed, so the pressures in the bubbles will be the same

regardless of whether they are coalesced or not. With the only force working to alter the

bubble shape being the surface tension, which is very small in comparison to the other

forces involved, it is not unreasonable to assume that bubble coalescence during growth

66
5.2 Bubble Detachment Criterion

has no noticeable effect on the growth and detachment process.

The assumptions which will be replaced are:

1) the gas inside the bubble is backmixed;

2) the bubble detaches when its base reaches the top of the tuyere and

3) the drag coefficient has a constant value of 0.4.

The replacement for the first assumption will be discussed later. The original drag

coefficient will be replaced by a value calculated from

4tpgdz,
D
- PB Vj,

This correlation has been developed for spherical drops and bubbles in pure systems.
116 It

should be noted that surface active components within the bath will cause this to be an

underestimate particularly at high Reynold s numbers.


7 The bubble detachment

criterion will be discussed in the following section.

5.2 Bubble Detachment Criterion

In previous models, the point at which the bubble detaches from the tuyere is

generally calculated as a simple relationship between the bubble radius and the distance

of the bubble centre from the tuyere. These simple equations can be used for vertical

injection, but not for horizontal injection. Ashman et at.


35 assumed that the bubble

detached when its base reached the top of the tuyere. This is similar in general form to

the vertical injection correlations. However, using the assumption that a volume of gas

equal to a hemisphere of diameter d, remains on the tuyere at detachment, a better

criterion can be determined. This assumption was introduced by Davidson and 5


SchUler

based on observations of vertical injection systems. The exact volume of gas remaining

on the tuyere is not known, so an assumption must be made.

67
5.2 Bubble Detachment Criterion
Figure 5.4 shows the assumed process of detachment and Figure 5.5 shows the

geometry of the system at detachment. For this purpose it is assumed that the bubble

reverts to a spherical shape once necking has occurred, since gas momentum no longer

affects it. The geometry can be used to determine the position of the bubble when

detachment occurs as follows.

B.

C. D

Figure 5.4 Bubble detachment process.

The gas remaining on the tuyere at detachment forms part of a sphere as shown in

Figure 5.5. If it is assumed that the detaching bubble is touching the wall directely

above the tuyere, then at detachment

68
5.2 Bubble Detachment Criterion

...[5.10]
h =(s2_r)2

Figure 5.5 Bubble detachment geometry.

Values of h and s for a range of bubble sizes at a tuyere diameter of 0.05 m are

plotted in Figure 5.6. It can be seen from the figure that, especially for larger bubble

sizes, detachment takes place considerably before the base of the bubble reaches the top

of the tuyere. This point can be seen more clearly from Figure 5.7. In this case it is

possible that the new bubble growing on the tuyere may punch through into the detached

bubble, causing further growth.

69
5.2 Bubble Detachment Criterion

0.45

0.4

0.35

0.3
E
w 0.25
C)
z
0.2

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4
BUBBLE RADIUS (m)

Figure 5.6 Variation of the height of the bubble centre above the tuyere centreline
(h) and the distance of the bubble centre from the tuyere centre (s) with bubble radius at
detachment.

70
5.3 Bubble Break-up During Growth

bath

10.1

Figure 5.7 Variation of bubble position at detachment with bubble diameter.

5.3 Bubble Break-up During Growth

The growth and detachment equations outlined above are still not sufficient to

completely model the bubbling phenomenon. In particular, the predicted size of the

bubbles at detachment is much larger than can be considered reasonable. Therefore,

there must be some other mechanism which reduces bubble size at work. It has been

noted that large bubbles rising in a liquid are unstable and have a tendency to break up,

and this is likely to be the case for the bubbles growing on the tuyere. Thus it is proposed

71
5.3 Bubble Break-up During Growth

that as a bubble grows on the tuyere it reaches a size where it becomes unstable and

breaks into two bubbles, one of which is still attached to the tuyere while the other is free

to rise.

To calculate the minimum bubble diameter at which breakup will occur a

combination of Kelvin-Helmolz and Rayleigh-Taylor instability theory is used.


8 This

theory has recently been applied to the bubbling/jetting transition in submerged

19 but is more applicable to the present application. It assumes that an initial


injection,

perturbation with amplitude q, produces a travelling wave with the equation

1 =q,exp[kc,t+ik(x Crt)] ...[5.1l]

Which can be rewritten as

11 Th exp[kctJ cos(k(x Crt)) ...[5. 12]

This equation indicates that c represents the disturbance propagation speed, and kc, is the

amplitude growth factor. The model of Kitscha and Kocamustafaogullari predicts the
8 For the
breakup of fluid particles using a relationship between these two quantities.

case of a bubble rising in a fluid (Figure 5.8) the propagation speed is given by

_( PBVBSifl8O ...[5.13]
cI
2pB+pbcoth(khb)

and the growth factor is given by


! ...[5.14]
k f PBPb 2
( l.
coth(khb)k VB Sfl
5
9)2
3
ak g I Ap I k
J

1.

(PB + Pb coth(khb))
2 PB + Pb coth(khb)
From Figure 5.8 it can be shown that for spherical bubbles

72
5.3 Bubble Break-up During Growth

B.

VB

VB

Figure 5.8 Schematic illustration of flow around a bubble.


118

db ...[5.15]
hb 0
=--(cos8 cos8)

while for an ellipsoidal bubble

...[5.16]
hb 0
+
=(a

)
2
)cos9
(ba cosO
02 +
(a

) cose)
(ba cos9

where e0 is the point where the disturbance originates and e is the wake angle. Both of

these forms of the equation will be required since the cross-section perpendicular to the

tuyere axis will be circular.

From the above equations the breakup criterion may be derived as

...[5.17]
EpB+pbcoth(khb)ld 1 Eitan()i I2
(1.5vsinO)
ppcoth(khb)k g
3
ak IApIk l>
[ PBVB
3 Jb
Itan()I 1 (PB+PbCOthQthb))
2 PB+PbcothQthb)S

and for rising bubbles Cg is given by

73
5.3 Bubble Break-up During Growth

...[5.18]
+ Pb coth(khb)J2
Cg =

To calculate the minimum bubble diameter at which breakup will occur values are

required for the wake angle and the wavenumber. The wake angle is given by the

empirical correlation

12

5it l9it ...[5.19]


e =+-jj-exp(O.62Re.)
04

To determine the wavenumber which will cause the most unstable wave growth,

equation 5.14 is differentiated with respect to k and set to zero. This results in the

equation

apbhb csch
3
k (khb)
2

jJ
0
[
2
(khb)coth(khb)
2
2hbpbcsch
+k2{3a[pB + p coth(kh)]
PBPJ1.51B h csch2(khb)]}
[5 20]
_k{2PBPb1.5vB sin[ coth(khb) +gpph csch2(khb)} gpB(pR + Pb coth(khb)) = 0

If the value of khb is large it can be assumed that coth(khb) is equal to one, and that terms

(khb) are negligible. Using these simplifications equation 5.20 can be


involving csch
2
reduced to

( ( 2 ...[5.21]
_2Pbl.5vBsinJJk_PB=0
2
3ak

This indicates that, for a given bath composition, the critical wavenumber is a function of

the velocity alone, and that it has a minimum value when the velocity is equal to zero. If

these assumptions are applied to equation 5.17 the maximum bubble diameter can be

calculated as

74
5.3 Bubble Break-up During Growth

3vB ,..[5.22]
db
= c{*2252k2i23k}

where
...[5.23]
[Itan)I
[Itan)I
and

* Pb ...[5.241
P
PB

However, at low velocities, application of the above equations results in values of

the wavenumber which are considerably smaller than the minimum wavenumber which

could cause breakup:

2it ...[5.25]
kmill

C ,i

Also, at low velocities the simplifying assumptions are not valid. For this case, when the

calculated values of wavenumber which should give the most rapid breakup are less than

k., it follows that k should be used in equation 5.17. After considerable manipulation

this gives the equation

...[5.26]
I (1 + p) 2 ) J
*
PB 8W Ow) PB w

75
5.3 Bubble Break-up During Growth

for the maximum bubble size. Equation 5.26 can be solved using the standard procedure

for quadratic equations. lii all of these calculations the bath velocity used must be the

velocity of the top surface of the bubble relative to the bath. It is important to note that

all of these derivations are for spherical bubbles. For the elliptical cross-section of the

bubble db must be replaced with 2b(l-e), where


! ...[5.27]
(b2 a2
2

8=1

i b2
is the eccentricity of the ellipse.

Figure 5.9 shows the variation of predicted maximum bubble sizes using

equations 5.22 and 5.26. This shows that equation 5.26 is only valid at very low

velocities (less than .33 ms), which are not likely to occur under injection conditions.

Using equation 5.22 for the conditions given in Table V-I gives a maximum bubble

diameter of approximately .2 m. This is considerably smaller than the diameter predicted

by the growth models. Figure 5.9 shows that there is a strong dependence on the bath

velocity, with a higher velocity allowing a larger bubble. This result is expected, as the

criterion by which the occurrence of breakup is determined, is based on a comparison of

the rate of disturbance growth with the rate the disturbance passes over the bubble

surface. If the disturbance reaches the side of the bubble before it is large enough to

cause breakup then it is swept off the bubble without any effect. Equation 5.13 indicates

that if the bath velocity is faster, the disturbance propagation speed (Cr) is also faster,

allowing a larger bubble to form.

76
5.4 Bubble Rise

1 350

0.9
300
0.8

0.7 250
E
LU 0.8
I- 200
LU LU
0.5
U D
LU z
-J LU
>
D
0.3 100

0.2
50
0.1

0 0
0 0.5 1 1.5 2

BATH VELOCITY (me)

Figure 5.9 Variation of maximum bubble diameter and wavenumber with bath
velocity.

5.4 Bubble Rise

After the bubbles detach, either from the tuyere or the growing bubble, they will

rise due to both the buoyancy force and the bath motion. For modelling, the assumption

that the bubbles rise at their terminal velocity should not introduce a significant error.

Thus the bubble rise velocity is given by



2
! ...[5.28]
(2(g lAp Idb
VT)L
P8
2

77
5.4 Bubble Rise

The assumption that the bubble rise velocity relative to the bath can be represented by the

rise velocity of a bubble in a stagnant bath has recently been shown to be valid under gas

injection conditions.
122

Further breakup of the bubbles may occur as they rise, and the same theory as

detailed in the previous section may be applied. In this case the velocity to be considered

is the bubble rise velocity with respect to the bath. From equation 5.28 it can be

determined that any bubble with a diameter less than 0.05 m will have a tenninal velocity

under 0.33 ms, and so will have a maximum bubble size as determined from

equation 5.26. For bubbles in this size range combining the bubble breakup criterion
with the terminal velocity expression gives

f a (2w3l ...[5.29]

2
d
b
sin2J_22}
2
{+)

For a bubble rising at its terminal velocity with a high Reynolds number

-. 5 ...[5.301
18

and for the large density difference in bubbling systems it can be assumed that

lAp ...[5.31]
PB

and that the term involving p is negligible. Using these assumptions the equilibrium

bubble size can be calculated as


...[5.32]
d,, = 3.258951--
i PB

78
5.5 Bubble Recombination

For larger bubbles the introduction of the terminal velocity equation does not result

in the same degree of simplification. Equation 5.21 becomes

0.l786gpdkgp=0
2
3ak ...[5.33]

The assumption that the term involving the bubble diameter is negligible introduces a

maximum error of under 0.1%, and leads to the simplified equation


1
[534]
k=1.8083J

Applying equations 5.28 and 5.34 to equation 5.22 results in

...[5.351
0.96868pbd+ 1.333d4PiJ 4a=0

The actual minimum bubble size will be determined by equation 5.35 if it is greater

than 0.05 m. Otherwise it will be calculated using equation 5.32. The bubbles will break

up progressively until the minimum diameter is reached. It is important to note that the

secondary and lower order bubbles formed by this procedure will also undergo breakup,

leading to the formation of a very large number of bubbles.

5.5 Bubble Recombination

With the large number of bubbles formed within the relatively small plume area,

bubble coalescence is inevitable. Unfortunately, it is also very difficult to model

theoretically. Instead an semi-empirical measure is required to determine whether

recombination occurs. The work of Sano and Mon


23 regarding the size of bubbles in a

plume gives the following equation for bubble size.


0.5 ...[5.36]
dvs=6.903J v
0
1

79
5.5 Bubble Recombination

This shows that the effect of the surface tension/density ratio, as discussed above, is

modified by the superficial gas velocity. The surface tension/density ratio represents the

equilibrium bubble size, while the effect of superficial gas velocity is due to bubble

coalescence. It is evident from equation 5.36, that at high superficial gas velocities

coalescence occurs more frequently, resulting in larger bubbles.

The effect of the superficial gas velocity appears to be important, but it has received

little attention, with only low values being used in physical modelling; ranging from
120
0.0065 ms
1 to 0.4 ms. The superficial gas velocity in the copper converter can be

over 3 ms, so it is probable that there will be considerably more coalescence than

observed in the physical models. In order for the gas volume flow through the bath to be

sufficient to give the required superficial velocity, the average bubble rise velocity must

be greater than the superficial velocity. Since the bubbles have been determined to be

rising at their terminal velocity relative to the bath, it follows that a high superficial gas

velocity will require larger bubble sizes to achieve the required volume throughput.

Therefore, in the model it is assumed that bubbles will coalesce until the average bubble

velocity is greater than the gas superficial velocity.

80
6.2.1 Bath Velocity and Gas Holdup

6 MODEL DEVELOPMENT

6.1 Introduction

In the past, models of copper and nickel converters have assumed that thermal and

thermodynamic equilibrium prevailed throughout the system. However, one study raised

some doubts as to whether the matte, slag and gas were at the same temperature.
97 This
would imply that interphase equilibrium was not attained, and this implication is

supported by gas temperature measurements carried out in a nickel converter.


25 A
further analysis of composition data obtained from a nickel converter
97 confirms that the
matte and slag are not in equilibrium. Therefore, to properly model the converter a

kinetic approach is required.

6.2 Preliminary Considerations

6.2.1 Bath Velocity and Gas Holdup

At least an approximate knowledge of the bath velocity is required, both for the gas

flow calculations, and for the calculation of mass and heat-transfer coefficients. The only

mathematical model of fluid flow in the copper converter determined that the maximum

bath velocity was 0.022 ms


126 However, this model did not consider the effect of the
.
1

slag layer, or calculate the velocity of the liquid in the plume. The calculated velocity

appears to be quite low, but it is explained by the large bath/refractory contact area and

low tuyere submergences, which are dictated by the form of the converter.
126 Physical

modelling of the Peirce-Smith converter has reported bath velocities of up to 0.168 ms

in a 1:4 scale water model.


113

An approximate estimate of the maximum bath velocity in the bubble column can

be obtained from the spout height. Applying an energy balance to the spout, considering

energy per unit volume, gives

81
6.2.1 Bath Velocity and Gas Holdup

Energy input = bath kinetic energy

PBV(1 P)

and

Energy output = bath potential energy .. . [6.2]

= pBgh
(
8 l

where is the gas fraction in the plume and , is the gas fraction in the spout. Equating

input to output gives

.[6.3]
PBVB(l
2 Pr) = pgh,(1

or

...[6.4]
(l_p)12
_(
2g hgp
VB
(1 ))

The gas holdup values can also be calculated from the spout height, by assuming

that the overall bath level is unaffected by gas injection. This implies that the volume of

gas in the plume is equal to the volume of liquid in the spout. With the assumption that

the basal radius of the spout is double the plume radius, based on time lapse photography

of the spout in a water model of a Peirce-Smith converter,

Ah4 =Ah(l )

or

3h
h
8(1 )

Substituting equation 6.6 into equation 6.4 gives

82
_____

6.2.1 Bath Velocity and Gas Holdup

2
[67]
yE 1
=h(
)J

This equation shows that the maximum bath velocity is independent of the value chosen

for the spout gas fraction, so only the spout height will be affected by.

An estimate of the gas holdup and liquid velocity in the plume can be obtained by

considering the volume of gas within the plume,

Vgp=Qtr
where tr is the average residence time of a bubble in the plume. The total plume volume,

is given by

V=Ah,
so

VgpQtr ...[6.10]
V Ah,

With

...[6.ll]

and


...[6.12]
= + VT
r

equation 6.10 becomes


...[6.13]
VB + VT

The average height reached by the liquid in the spout will be the centre of mass, so

...[6.14]

IB
( 4
2

cp))

83
6.2.2 Converter Geometry

Combining equations 6.13 and 6.14 and rearranging gives

...[6.15]
4Ijht
+,J 4=0

+2
vupv) + +2
vSUPvT)

If it is assumed that the average bubble is the equilibrium size calculated from

equation 5.33, then equation 6.15 may be solved numerically for q. It should be

emphasized that this is very approximate because the validity of some of the assumptions

is uncertain.

6.2.2 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. Figures 6.1 and 6.2 show schematic

cross-sections of a converter, idle and blowing respectively.

The idle case is the easiest to deal with, so will be considered first. The matte/slag

interfacial area is given by

22
! ...[6.16]
1 2
AM/s = 2(r (r hM) )L

Similarly, the slag/gas interfacial area is given by

22
...[6.17]
1 2
As,G = 2(r
(r hM ) )L
5
h

Unfortunately, values for hM and h are not directly calculable. Given the mass of matte,

MM, and slag, M, present in the converter, hM can be determined from

...[6. 18]
MM (7r _j r JiM (r hM)
LL i
2 2

r sin rChM
2
(


=
hM)
2

and h is calculated from

84
6.2.2 Converter Geometry

2 2
_l(rchs(rchs)( 221
...[6.19]
r sin rh )
5
h
Ps 2

PM

Figure 6.1 Schematic cross-section of an idle converter.

The matte/refractory contact area is given by

_/rhM](rchM)
1
.[6.20]
2
sin I h)2]
.

AM,R (2rChM -

r 2
-

where the second term in the equation represents the end-wall contact area. Similarly, the

slag/refractory contact area is given by


2
(r h
)
5 ...[6.21]
Lr[1c 2siif 11+2 3 h)2]_AM/R
(2rh

ASIR =
),)

r rc

85
6.2.2 Converter Geometry

When air is being blown into the converter a spout is formed. The possible

variations in spout form are shown schematically in Figure 6.2. If the spout height is

greater than the slag thickness (Figure 6.2a), only the slag thickness and matte/slag

interfacial area are affected. The value of h must be modified to account for the spout

volume

r _iis] (r h
)
5 ...[6.221
L[rc2sin-{ 5 hJ ntabh
(2rh 8

Values of a and b can be determined from the bubble formation calculations outlined in

Section 5. If the modified slag thickness is greater than the spout height, then the

configuration conforms to Figure 6.2b. To calculate the slag/matte interfacial area the

extra area due to the spouts must be included. 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, so

I N ...[6.23]
= 2(r (r ntab 2
a s sin EJ
where

b+h ...[6.24]
b=

If the matte spout height is less than the slag thickness (Figure 6.2b) the slag will

form part of the spout. The slag spout height can be calculated in a similar manner to the

matte spout height, using the value (hshspMhM) as the submergence. In this calculation

h will be given by equation 6.22 to account for the matte spout volume. The matte/slag
interfacial area is still given by equation 6.23, but the slag/gas interfacial area must be

86
6.2.2 Converter Geometry

sp

h,

Figure 6.2 Schematic cross-sections of an operating converter: a. no gas flow


through the slag, b. gas flow through matte and slag, c. gas flow through slag only.

87
6.2.2 Converter Geometry
modified to give
! ( ...[6.25]
= 2(r (r h + hM+ h,,j
L
)
2 n1t; ab a2__.sin_1 c

A third possibility is shown in Figure 6.2c. 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

equation 6.20, and the slag/gas interfacial area will be given by equation 6.25. When

blister copper is present it is also possible that the gas will pass through it as well. In this

case the above equations must be modified.

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 formation, with the

extreme case being slopping, when the bath surface can be approximated by a standing

wave. In this case it is likely that the extent of inter-phase 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.
27
This model uses a force balance approach to determine the radius of the slag droplets

formed, and gives

1 ...[6.26]
r
2 !
rd_l6g(pM_ps)cos31l_[l__ 2
l28Gg(pMpS)Co5
sV.
2
P j
1 is the velocity at the matte/slag interface and
where v is the angle shown in Figure 6.3.

To calculate the interfacial velocity, the equation

11627
2 1 -2

N ( VMXS (VMXS V
..

(PM
K =0.13671 II II V I [(1 K)(0.11080.0693K)]
5
.-

Ps} V ) )

was derived, where

88
6.2.2 Converter Geometry

V. ...[6.281
K=

and x is the slag thickness.


127

Finally, an energy balance is carried out to determine the rate of droplet generation.

For the present case it is assumed that the combined spouts produce a linear interface for

droplet formation, stretching the length of the converter. This gives the number of

droplets formed per second as


27

...[6.291
= v
2
O.83O8L(pvx)
Ps)COS f

00
0 0
o 0
DROPLETS
000
0
0
QL
r0
0
0
0
MATTE
00
0
0
0
0
0

GAS

Figure 6.3 Schematic of emulsion formation during submerged injection.


27

89
6.3.1 Equilibrium Within Phases

6.3 Mass Balance

6.3.1 Equilibrium Within Phases

It is assumed for the calculations that each phase is in internal equilibrium. As

such, each must be considered separately.

6.3.1.1 Blister copper

The copper phase is the easiest to deal with, since it only contains components in

their elemental forms. Thus, there are no reactions and the composition will be governed

only by mass-transfer considerations.

6.3.1.2 Matte

The matte is considerably more complex, since it contains a mixture of sulphides,

oxides and elements. The equilibrium composition within the matte will be governed by

the sulphur 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. The reacting components of the

matte are given in Table VT-I, and the equations required to calculate the equilibrium

composition are given in Table VT-TI. All other components are considered non-reacting,

so are not subject to equilibrium calculations. The equilibrium calculations are combined

with mass balances for each element to give the final composition.

Elements Cu, Ni, Fe, Co, Pb, Zn, S


Compounds Cu
S
2 , NiS, Ni
S, FeS,
2
CoS, PbS, ZnS
Table VI-I. Reacting components of the matte.

90
6.3.1 Equilibrium Within Phases

Number Reaction
1 2Cu + 2
1/2S S
2
Cu
2 =NiS
2
Ni+112S
3 2Ni + 2 = Ni
1/2S S
2
4 =FeS
2
Fe+1/2S
5 Co + 2 = CoS
1/2S
6 =PbS
2
Pb+112S
7 =ZnS
2
Zn+112S
Table VT-il Reactions required to calculate the equilibrium composition of the
matte.

6.3.1.3 White metal

The white metal is a special case of the matte phase, in which Cu


S is the only
2

remaining sulphide component. As such its sulphur potential will be governed by

reaction 1 in Table VT-TI.

6.3.1.4 Slag

While the slag is a complex mixture of oxides, it can be assumed that iron is the

only component which has more than one oxide form present. As such, the oxygen

potential of the slag will be controlled by the reaction

3FeO + 1/202 = Fe
4
0
3 ...[6.30]

and the composition is determined by the iron and oxygen mass balances.

6.3.1.5 Gas

Like the slag, the equilibrium composition within the gas is dominated by a single

reaction

+
112S
=
2 S0
0 ...[6.31]

91
6.3.2 Gas-Liquid Mass-transfer

Combining this with sulphur and oxygen mass balances and setting the total pressure to

atmospheric, allows the composition to be cakulated.

6.3.1.6 Calculation Technique

For each of the phases, with the exception of the blister copper, there is a system of

simultaneous equations which 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

non-linear equation solving routine. As in a previous work, a Quasi-Newton technique is

28 Model flow charts are given in the Appendix.

97
used.

6.3.2 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 gas-matte reaction is controlled by

gas-phase mass transfer. This suggests that the reaction will be

=FeO+S0
FeS+3/20
2 ...[6.32J

because the partial pressure of oxygen at the bubble surface will be very low, favouring

the least oxidized iron compound. Minor element removal may be controlled by

gas-phase, or liquid-phase mass-transfer, or by evaporation at the gas/liquid interface. To

calculate these rates a representation of the gas flow within the bath is required.

6.3.2.1 Gas flow

The basic theory regarding the gas flow within the bath has been developed in

Section 5. What remains is its implementation in a computer model. Two stages of the

model are required: bubble growth on the tuyeres and bubble rise. It is necessary that

these models also consider heat transfer to the bubble, so all aspects will be considered

92
6.3.2 Gas-Liquid Mass-transfer

together.

For the bubble growth calculations, the basic model of Ashman et al.
35 is followed,
with the modifications noted in Section 5. Assuming that the minor elements are present

in the gas in very small quantities, the mole balance within the bubble is given by

0
QP ...[6.33]
dn (x 1 2

where Ab is the bubble surface area given by

2 2tab ...[6.34]
ta +
2
Ab= sin e
. -

e is the bubble eccentricity, and x is the ratio of 02 reacted to SO


2 produced in the
reaction. It has a value of 1.5 for reaction with matte, 1 for reaction with white metal,

and will be infinite for reaction with slag and blister copper. In the latter case, to avoid

the numerical problems introduced by the us of infinity in the model, 10,000 is used so

that the term (x-1)/x in equation 6.33 is approximately equal to one.

The heat balance is given by

dTb QP dn ...[6.35]
nC---=CT +hbAb(TB -Tb)+Ab(cLT-61)-CPTl,---
3

This accounts for heating by convection and radiation. The heat of reaction is assumed to

report entirely to the bath. The bubble volume depends on the moles of gas and gas

temperature, according to

dVRTidn vT ...[6.36]
dtP dt+ dt

The upwards motion of the bubble is given by

93
6.3.2 Gas-Liquid Mass-transfer

dw dV ...[6.37]
V-j--=2V+CDitabw 2 -w-

where

WVbVB ...[6.38]

The change in oxygen content of the bubble can be calculated from

2
dC
J QP
0
k A b C02 C2
dV
...[6.39]
dt RTO) dt

and the sulphur dioxide content is given by

0
dC [6.40]
dV
dt

These equations determine the size and position of the growing bubble. In the

model the equation governing the bubble shape (equation 5.4) is solved by a fmite

difference technique, while the above equations are solved simultaneously using a

Runge-Kutta solution technique. At the end of each time step the bubble stability is

determined according to equation 5.31 or 5.35, depending on the interfacial velocity, If

the bubble is unstable then it is assumed to break up into two bubbles, one of which is

still attached to the tuyere. The size of the free (secondary) bubble will depend on the

size of the initial bubble at breakup, and must be used as a fitting parameter.

The equations relating to the motion and breakup of the bubble as it rises through

the bath have been developed in Section 5. What must be considered here is the heat and

mass balance within the rising bubble. The same basic equations can be used as in the

previous analysis, modified to account for the different conditions. Thus the mole

balance and heat balance become

94
6.3.2 Gas-Liquid Mass-transfer

p ...[6.41]
dri (x1,
dtl YQAbRT

and

dT,, dn ...[6.42]

respectively.

The equations governing the volume variation and sulphur dioxide concentration

(6.36 and 6.40) remain unchanged, while the change in oxygen concentration is given by

dC2_ [6.43]
dV
.

i A ,-b (-b ..

b-O
O
2
d
d

The modified equations can then be solved by the same technique. By calculating the

passage of a primary bubble and all its related secondary bubbles through the bath, a

number of different parameters may be determined. The gas composition and

temperature at the bath surface are calculated, along with the amount of oxygen reacted

and, therefore, the moles of FeO formed. From the amount of oxygen reacted, the heat

generated by the reaction can be calculated for use in the heat balance. Also, by

combining the gas flow model with a mass-transfer model for the minor elements, the

rate of minor element volatilization may be calculated. This will by dealt with in the

following section.

6.3.2.2 Mass transfer from liquid to gas bubbles

The mass transfer of the minor elements from any of the liquid phases to the gas

bubbles may be controlled by one of three transport steps, or by a combination of the

steps. These are the liquid phase transport from the bulk of the bath to the bubble

95
_________

6.3.2 Gas-Liquid Mass-transfer

surface, vapourization at the bubble surface, and gas phase transport from the bubble

surface to the bulk of the gas. The mass-transfer rate for each step may be detennined

separately, and the three combined to give an overall rate equation.

By assuming quasi steady state conditions, mass-transfer in the liquid phase can be

characterized by the equation

1 =
1i kA(Cf *)
1
C ...[6.44]
or, replacing concentrations with mole fractions to give compatibility with the

thermodynamic model

nI=kLAbp(Xr_X,*) ...[6.45]
The vapourization rate can be calculated using the Langmuir-Knudsen equation.
129

aAb *
...[6.46]
(PtP.)
2itRTM,

It is important to note that this equation was derived from the kinetic theory of gases for

vapourization of a solid into a vacuum,


and so some error may be introduced due to the
13

gas pressure within the bubble. This equation is also modified to allow for integration

with the equilibrium model using

p+po. ...[6.47]

which gives

...[6.48]
)
*
(Py.X. P.
-J2iR7M

Mass-transfer within the bubble can be characterized by the equation

kQAb * B
...[6.49]
n =--(P P
)
1

96
______

6.3.2 Gas-Liquid Mass-transfer

An overall equation for the mass transport rate to the gas bubbles can be obtained

by combining equations 6.45, 6.48, and 6.49. Rearranging equations 6.48 and 6.49 gives

...[6.50]

and

* (RT B
...[6.51]
1
P ni1jAJ+Pi

respectively. These equations can then be combined to give

= .(I2RTMI +z+PB ...[6.52]


Ac kGA)

Rearranging equation 6.45 gives

...[6.53]
X. =XB
*

, m
Lb PB

which, when combined with equation 6.52 and rearranged gives

,i. 1 1 pB ...[6.54]
-=I i 1 1 1x7
1 j:Oy
Ab kkkpm

where

...[6.55]
k=
-j2iuRTM,

and

y,
0
1
kP ...[6.561
kg
RT

97
6.3.3 Mass Transfer Between Liquid Phases

This equation is very similar to that derived for vacuum refining models, with differences

32 The one exception to this


being due to the gas pressure within the bubble.

131

formulation is the behaviour of the zinc in matte. Zinc oxide is thermodynamically more

stable than iron oxide, so the zinc reaching the gas/liquid interface will react. Thus, the

vapourizing zinc species will be zinc oxide, and any oxide which does not vapourize will

report to the slag.

6.3.3 Mass Transfer Between Liquid Phases

Within any phase, I, the mass-transfer rate of component i to an interface is given

by

ri *
...[6.57]
= k[(C[ C[)

or, converting to mole fractions

...[6.58]

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)

=
K(X*)n .. .[6.59]

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

ri *
...[6.601
j=kfp(Xi X!)

98
6.3.3 Mass Transfer Between Liquid Phases

or, rearranging

ri1 ...[6.61]
Akp

Combining equations 6.58, 6.59 and 6.61, and rearranging, assuming n=1, gives

( 1 ...[6.62]

1
kp ?p

This is a general equation which 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.

With no reaction, n=1 and

...[6.63]
K=L

Where there is a reaction at the interface, for example

1 1 ...[6.641
S) +
2
Cu = Cu
0
2 (s) + S

the equilibrium constant is calculated as

...[6.65]
oxcu
Ycu
o
2 P.
KR
SXCU
CU
S
2 P

so n=1 and

...[6.66]
oPs
ycu
2

99
6.4.1 Energy Loss
In these equations the partial pressures of sulphur and oxygen represent the sulphur and

oxygen potentials of the matte and slag respectively, and do not imply the presence of

gaseous species. The equilibrium constant is calculated at the average temperature of the

two phases.

6.4 Energy Balance

Without the assumption of thermal equilibrium each of the component phases

within the converter may be at different temperatures. While this adds to the complexity

of the model, the temperature variation may be significant. Heat transfer from the bath to

the gas is calculated as part of the overall gas flow model, so does not need to be

considered further. However, the cooling effect of the gas on the bath is important.

To calculate the temperature of each phase an energy balance must be carried out.

That is

energy accumulated = energy input + energy generated .. . [6.67]

- energy loss energy consumed


-

Each of the terms on the right hand side of equation 6.67 will be considered separately.

6.4.1 Energy Loss

Energy losses come from three sources, materials leaving the converter, conduction

through the walls, and radiation through the mouth. A thermodynamic model of the

nickel converter determined that, during blowing, energy leaving the converter as the

sensible heat of the gas accounted for almost 90% of the total energy losses, while

radiation accounted for the majority of the remaining losses.


5

100
6.4.1 Energy Loss

Energy lost with the gas leaving the converter is calculated as part of the energy

consumption term, so will be considered there. The sensible heat of skimmed or cast

liquids does not affect the temperature of the remaining materials, so is not required for

the model.

Radiation losses from the slag are assumed to be only to the mouth of the converter.

As such they will be dependent on either the hood temperature during blowing, or the

ambient temperature (no mouth cover) while the converter is idle. The rate of energy loss

will be given by

...[6.68]

The emissivity of the slag must be assumed.

Allowing the different phases to be at different temperatures adds complexity to the

calculation of energy loss through the walls. The losses are not uniform over the entire

converter, and the outer wall of the converter is not at a uniform temperature. This was

determined by radiation pyrometer measurements carried out on an operating nickel

converter, which indicated that there was a considerable temperature variation over the

converter shell.
33

However, since the wall losses are very small compared to the energy removed with

the gas and the radiation losses, a relatively simple calculation technique can be used.

Thus, for the end walls


97

k, ...[6.69]
4ew = 2ACf (TB
T)+--(T
T)]

and for the barrel section

L
1
2k r 1 .. .[6.70]
q cb = R I (TB
T) +
2
(T T)
ln(JL

101
6.4.2 Energy Consumption

where the constants k


1 and k
2 are due to the variation of refractory thermal conductivity

with temperature according to


134

kR =k
(1 +k
1 T)= 1.08(1 +4.5(10)T)
2 ...[6.71]

for chrome-magnesite refractory.

6.4.2 Energy Consumption

In the model the consumed energy is that which is required to bring all charged

materials up to the temperature of the phase to which it reports. This value is calculated

as

=
J
T
CP(T)dT+Ae12]
...[6.72]

The integral in equation 6.72 has been pre-calculated for use in the model, and gives a

final equation
I I d. i ...[6.73]
0 =
q Eai+biTB+ciT+_!JJ
TB

Values of the constants in this equation are given in Table VT-HI.


97

102
6.4.2 Energy Consumption

jCin J mor

i a b, 3
cx10 x10
1
d
5 Notes
Fe 7841.02 24.476 4.226
Ni -6529.9 25.104 3.766
Cu -141.3 31.38
Co -3940.3 13.807 12.259
FeS -8925.9 51.045 4.979 T<1468K
FeS 4664.09 71.128 T>1468K
S
2
Ni -25182.5 20.318
NiS -12784.5 38.911 13.389
S
2
Cu -6958.62 84.935
CoS -13682.7 44.35 5.251
4
0
3
Fe -52247.3 200.832
FeO -32218.8 68.199
NiO -9589.16 46.777 4.226
0
2
Cu -19636.7 62.342 11.924 T.<1509K
0
2
Cu -38545.1 96.274 .293 -1.925 T>1509K
CoO -14205.8 48.283 4.268 -1.674
2
S -12165.7 36.484 .355 3.766
02 -9507.5 29.957 2.092 1.674
2
SO -15407.7 43.43 5.314 5.941
2
N -8493.394 27.865 2.134
2
Si0 -29666.3 58.911 5.021

Table VT-HI Integrated values of C for use in equation 6.73.135

103
6.4.3 Energy Generation

6.4.3 Energy Generation

The only sources of generated energy in the converter are the reactions. Since the

rates of the main reactions are controlled by oxygen mass-transfer within the gas, the

total energy generated within a time step is approximated by

qgen = A1RAt
2
flO ...[6.741

The value of AHR will depend on the phase in which the reaction occurs. The reactions

which may occur are: in blister copper

2Cu + 1/202 = Cu
0
2 ...[6.751

in white metal

S
2
Cu + 02= 2Cu + SO
2 ...[6.76]

in matte

FeS + 3/202 = FeO + SO


2 .. .[6.77]

and in slag

3FeO + 1/202 4
0
3
Fe ...[6.78]

Values of the enthalpies of reaction for each of these are given in Table VI-TV. The

actual reaction heat is also dependent on the temperature, so the values in Table VT-TV

are modified accordingly. Some energy generation or consumption may also result from

enthalpies of solution, but this will be small in comparison to the main reaction enthalpy.

Reaction AHR
(klmor
)
1
6.75 -166.7
6.76 -190.4
6.77 -463.4
6.78 -317.6
136
Table VI-IV Enthalpies of reaction required for energy generation calculation.

104
6.5.1 Physical and Thermal Properties
6.4.4 Interphase Heat-transfer

Energy will be transferred between phases by convection and mass-transfer. The

convective heat-transfer will be given by

1
4
(
A!,
11
=h

1 T
1
T) ...[6.79]
and the heat transferred with materials passing between phases is given by

= ...[6.80]
It should be noted that in cases where there is a reaction at the interface the heat

generated by the reaction is assumed to be contained in the metal-containing product, and

will report to its corresponding phase.

6.5 Data

In a model of this sort a wide range of data is required. Unfortunately, much of this

data is unavailable in the literature, or only present for a narrow range of temperatures

and/or compositions. Therefore, it is necessary to assume that whatever data is available

is valid over the entire range of conditions present in the model and, if nothing is

available, required values must be assumed or fitted.

6.5.1 Physical and Thermal Properties

Data concerning the physical and thermal properties is required for the gas and the

three condensed phases. Most of these properties vary considerably with composition

and temperature, but it is unusual to find a study which considers both together. The

required properties and the equations used to calculate them are given in Table VT-V.

The equation for the slag viscosity given in Table VT-V is derived directly from

measured viscosities,
40 while the equations for matte and blister copper viscosities are

estimated from measured diffusivities. The constants used in the gas viscosity equation

in Table VT-V are given in Table VT-VT.

105
6.5.1 Physical and Thermal Properties

Phase Property Ref.


Density_(kg m
)
3
Slag 8 0.1 1454T[SiO
3297 + 128.45[FeO] 1
2 137

Matte 3880 + 403.6[Cu]M + 4589[Cu] 3750[Cu] 138

Blister Copper 7800 139


Gas G
3.469(103)_J
-

Viscosity_(kg_rn_i s
)
1
Slag [Fej V 140
2.063(10 0.7l43[s.QJ
,JJ
Matte 139
J
5000
3.36(loiexp(
TM

Blister Copper 4866 141


7.688(10
)
5 exPL\ TB

Gas 2
Co+ClTG+C
T
3
7
4 +C 142

Surface Tension (N m
)
1
Slag 0.71483.17(10)(T273) 138

Matte ( 4
X (2.2711.673x ( X, (1.149+1.188Xg 143
XF+XCJL 53X JXF+XcuJ1 XN
2
+
3 J
Blister Copper 1.136 )TB
5
1.6(10 144

Thermal_Conductivity_(W rn_i K
)
1
Slag 2.09 77
Matte 13.399 17.875[Cu]M+25.551[CU1 13.74[Cu] 138

Blister Copper 133.9 139


Table VT-V Physical and thermal properties of the gas and condensed phases.

106
6.5.2 Thermodynamic Data

Co C, 2
C 3
C 4
C
(x10
)
5 (x10
)
7 (x10
)
9 (xlO
)
2 (xlO
)
6
7.164 -5.083 1.578 -1.785 6.667
Table VT-VT Coefficients for gas viscosity equation (air).
42
6.5.2 Thermodynamic Data

6.5.2.1 Free energy

To calculate the equilibrium composition of the matte the free energies of the

reactions given in Table VT-il must be known. Values of these are given in Table VI-Vil.

Reaction AG
)
1
(Jmo1
2Cu + 2 = Cu
1/2S S
2 -147 100+42.05T
Ni + 2 = NiS
112S -l29300+54.06T
2Ni + 2 = Ni
112S S
2 -152500+46.06T
Fe+ 1/2S
=FeS
2 -115300+30.63T
Co + 2 = CoS
1/2S -125000+48.12T
Pb+ 112S
= PbS
2 -138400+64.14T
Zn + 1/2S
2 = ZnS -264800+98.20T
Table VT-Vil Free energy of formation of matte compounds.
42

In addition, free energy data is required to calculate the equilibrium constants used

to determine rates of interphase mass-transfer. The reactions and their corresponding free

energies are given in Table VI-VITI.

Finally, to calculate the compositions of the slag and gas phases, free energies of

two further reactions are required. These are given in Table VT-TX.

107
6.5.2 Thermodynamic Data

Reaction AG
(J_mor)
S
2
Cu + 1/202 = Cu
0
2 + 2
1/2S 28360-2.59T
NiS + 1/20 = NiO + 2
112S -122800+41.67T
FeS + 1/202 = FeO + 2
112S -128200+21.13T
CoS + 1/202 = CoO + 2
1/2S -107800+22.6 iT
PbS + 1/202 = PbO + 2
1/2S -57950+21.38T
ZnS + 1/202 = ZnO + 1/2S
2 -80330-3.35T
Table VT-Vu Free energy of reaction for matte-slag reactions.
35

Reaction AG
(Jmol)
3FeO + 1/202 = Fe
4
0
3 -402000+169.77T
2 + 02 = SO
1/2S 2 -363000+72.42T
Table VI-IX Free energy of reaction for slag and gas reactions.
35

The data available regarding the free energies the minor elements presents some

problems. While the agreement for reactions involving single elements in the gas phase

is not bad (Figure 6.4), the free energy of formation of the oxides and suiphides of arsenic

and antimony do not appear to be consistent. Figure 6.5 shows the variation of AG with

temperature for the formation of SbO and SbS, and Figure 6.6 shows the values for AsO

and AsS. It is obvious that a problem exists, particularly in the case of SbO, where two

papers, published in the same year and having one author in common,
3 use vastly

30

different equations for the free energy of the reaction

1 1 ...[6.811
(g) + 2
2
Sb 0
( g) = SbO (g)

108
6.5.2 Thermodynamic Data

50

40

30

20

C)

10

-10

-20
1000 1200 1400 1600 1800 2000
TEMPERATURE (K)

Figure 6.4 Comparison of reported values of free energy of reaction.

109
6.5.2 Thermodynamic Data

5
SbS

Chaubal et al. [291


Seo and Sohn [30]
-10
Nagamori and Chaubal [26]
-

Hino et al. [144]


Kim and Sohn [31]
.----

-20 -

-25 I I

1000 1200 1400 1600 1800 2000


TEMPERATURE (K)

20

SbO
10

-10

-20 I

1000 1200 1400 1600 1800 2000


TEMPERATURE (K)

Figure 6.5 Comparison of reported values of free energy of formation of SbS and
SbO.

110
6.5.2 Thermodynamic Data

5
AsS
0

Chaubal et al. [29]


-5 Sea and Sohn [30]
Nagamori and Chaubal [261

-10
0

-15

-20

-25
1400 1600 1800 2000
TEMPERATURE (K)

-5

AsO

-20

-25

-30

1000 1200 1400 1600 1800 2000


TEMPERATURE (K)

Figure 6.6 Comparison of reported values of free energy of formation of AsS and
AsO.

6.5.2.2 Enthalpy of Reaction

Values required to calculate the heat produced by the reactions are given in

Table VI-IV.

111
6.5.2 Thermodynamic Data
6.5.2.3 Activity coefficients

The equilibrium calculations within each phase and the mass-transfer calculations

between phases, require values for activity coefficients. Equations for calculating these

for most of the major constituents can be found from the literature, and those used are

given in Table VI-X.

Phase i Ref.

Slag FeO exp()log(l.42Xpo .2)) 20

4
0
3
Fe 0.69 + 56
XFeO
8
. + 5.45X
0 17

NiO exp(() 1.62) 146

0
2
Cu 9 147
CoO 0.66 148
Matte Fe 40 149
Ni 15 97
Cu 14 17
Co 25 149

FeS exp((8)ln(.54 + + .52XFeS)) 20

(1 1840 I
I I j I.6
S
2
Ni 10 M)
150
NiS 1 97
S
2
Cu 1 17
CoS .4 151
4
0
3
Fe (1573\ 20
expq
TM J
(4.96 + 9.9 logX + 7.43(logX)
2

))
3
)
2
+2.55(logX
Table VI-X Activity coefficients of the major constituents of the matte and slag.

112
6.5.2 Thermodynamic Data
Lead is a common contaminant in copper mattes, and is often present in relatively

large concentrations. Copper is also an important contaminant in lead, so there is

considerable information available.


15261 Azuma et al. have determined that the activity

coefficient of lead in copper mattes is equal to one at 1473 K.


52 However, the data of

Eric and Timuin suggest that there is a considerable negative deviation from ideality,

particularly at high matte grades.


58 The same study indicates that the activity coefficient

of PbS at infinite dilution in white metal is 0.035,158 and in dilute solutions

lny = 6.970Xpbs 3.344 ...[6.82]


at 1473 K. Rome and Jailcanen calculated a value of 0.155 for the same parameter, and

suggest that it is relatively constant down to X


2 = 0.7.160 Their results also indicate that

the activity coefficient of lead in white metal is approximately 4, but is quite sensitive to

the sulphur activity.

The activity coefficient of lead in copper metal has been calculated by a number of

162468 and there is fairly good agreement between the various studies. A value
researchers,

of 5.7 is used in the model. The activity coefficient of PbO in slag in zinc slag-fuming is

given by
169

(3926 ...[6.83]
T J
and this agrees well with measurements in copper smelting slags.
This value appears
7

to be unaffected by the oxygen partial pressure.

There is considerably less information available regarding zinc. However, there is

sufficient data to give values for activity coefficients in the three phases. In copper, zinc

is reported to be a regular solution with


63

113
6.5.2 Thermodynamic Data

r64O0X1 ...[6.84]
Yzn=exP[
RT j
For dilute solutions of ZnS in matte, the activity coefficient has been calculated as
25

= 6.8 O.02(T 1523) [Cu]M ...[6.85]


while at 1523 K it is reported to be represented by the polynomial
33

= 14.21 136.6X + 437.3X 405X ...[6.861


In slag the ZnO activity coefficient can be calculated from
26

E9201 ...[6.87]

Unlike lead and zinc, antimony is generally thought to dissolve in atomic form in all

three condensed phases present in copper converting. The activity coefficient of

antimony in copper is very low, with reported values ranging from 3.7x10
3 to

1
.
2

60
1 2 The higher of these values are considered to be the more accurate.
.4x10
6 71
72

The activity coefficient of antimony in matte and white metal is considerably higher

than for copper, with the value for white metal being 0.44 and the value in matte

calculated from
173

=exp[4.8584 l.936X] ...[6.88]

at 1473 K. However, both of these values are very sensitive to the sulphur activity when

the matte is sulphur deficient. The activity coefficient of antimony in matte is also

sensitive to the oxygen potential of the system, indicating that some oxide may be formed

at high oxygen partial pressures.


93

There is relatively little data concerning antimony dissolution in slags.


174 What is

170

available is given in the form of distribution ratios, which must be converted to activity

coefficients. An analysis of the data of Nagamori et al. indicates that there is a

114
6.5.2 Thermodynamic Data
considerable scatter in the experimental results.
The distribution ratio (Lsb) was given
7
as 30, with the oxygen potential having no apparent effect. This results in an activity

coefficient of 0.4 for antimony in slag.

The behavior of arsenic in copper converting systems is similar to antimony. As in

the case of antimony there is a wide variety of information available concerning arsenic

dissolution in
160
matte,
1 61 and 1
73
75 27
copper,
1
75181
62 with much less available
63

regarding the slag.


182 There is a very wide range of reported values of ; from

70
167
7
5x10 to
18 Some variation with temperature is seen, but it is not consistent
.
2
1x10

between experimenters. It is probable that the lower values of arsenic activity coefficient

are incorrect due to the values of free energy and gas vapour pressures used in their

181
calculation.

The activity coefficient of arsenic in white metal is reported to be in the range 0.2 to

0.12, dropping to 2
l.2x10 as sulphur is removed.
176 Variation of the iron content of the

matte has been found to cause a considerable change in the arsenic activity

160
coefficient.
1
61 At 1473 K
73

log = 1.35 0.848X


2 ...[6.89]

but sulphur deficiency in the matte causes a considerable reduction in the value

calculated from this equation.


73

There is significant scatter in measured data regarding arsenic in slags, although

there is agreement that the dissolved form is monatomic As. Lynch and Schwartze

measured values of varying from 1 to 10 depending on slag composition.


82 Nagamori

et al. reported an average value of LAS of 300, indicating that dissolution of arsenic in

slag equilibrated with molten copper is very small.



7

115
6.5.2 Thermodynamic Data
Like antimony and arsenic, it is generally accepted that bismuth dissolves in

monatomic form in all phases concerned in copper converting, although there is some

evidence of sulphide formation at high sulphur activities.


60 However, its behaviour does

not follow the same trends. The activity of bismuth in copper shows a positive deviation

from ideality,
83 with an activity coefficient given by

63
7

c 43751 ...[6.91]
exp[ 2.04
j
=
+ T

51001 ...[6.92]
YB=exp 2.45+
T

Both of these equations give values which agree well with the results of other

Information on bismuth in white metal and matte is limited, and the studies

available show very poor However, the difference can be explained

by the value of bismuth vapour pressure used.


85 Oxygen has no noticeable effect on the

activity of bismuth in matte, and BiS vapour is formed in preference to BiO, which is

closer to the behavior of lead than arsenic and antimony.



6

A number of studies have been carried out concerning the distribution of bismuth

between slag and copper or


7
161
matte.
8 192
0 Most of these calculate distribution

coefficients, and there is considerable disagreement regarding the form of bismuth in the

92 report that oxygen partial pressure has an


slag. However, the most recent studies

6

effect on the amount of bismuth reporting to slag, and one gives activity coefficients in

slag of 1500 and 1.1(108) for Bi and BiO respectively at 1458K and 92
). At
4
PBI=7.5(10

116
6.5.2 Thermodynamic Data
higher temperatures the activity coefficient of Bi is increased and that of BiO is reduced.

However, the magnitude of these numbers indicates that very little bismuth will report to

the slag.

Table VI-XI summarizes the activity coefficients of the minor elements used in the

model.

Matte
Pb 23
PbS exp(.2008 2.3245X)

Zn 25
ZnS 6.8 (TM
02
O. 1523) [CuJM

As (1.35_ o.848x)
1
Sb exp(4.8584 l.936X
)
5

Bi 13.6
White Metal
Pb 4
PbS exp(6.97XPbS 3.344)

Zn 3
ZnS 6.8 O.02(TM lS
)[CuJM
23

As 0.12
Sb 0.44
Bi 6.1

Table VI-XI Minor element activity coefficients used in the model.

117
6.5.2 Thermodynamic Data

Slag
PbO (3926
expi

ZnO (920
expiT
8

As 5
Sb .4
Bi 1500
Blister Copper
Pb 5.7
Zn
expi
( 6400X
RTB
As 1(102)
Sb 1.4(102)
Bi exp_2.04+4?5J

Table VI-XI Minor element activity coefficients used in the model (continued).

6.5.2.4 Equilibrium vapour pressures

Equilibrium vapour pressures of the minor elements are required to calculate

liquid-gas mass-transfer rates. The values used are given in Table V-Xll.

118
6.5.3 Kinetic Data

Specie Vapour Pressure Ref.

As exp((_)+ 10.62) 31

Bi exp((_22)+ 11.75) 31

Sb exp((_)+12.336) 29

( 23325
Pb exp_)+l9.O65_0.985lnT) 29

PbO exp(()+13.843) 29

PbS exp((_6)+12.977) 29

I( 15243
Zn exp(_--)+21.782_1.255lnT) 26

ZnO exp((_2)+4.651) 26

Table VI-Xll Equilibrium vapour pressures of the minor elements.

6.5.3 Kinetic Data

6.5.3.1 Diffusivities

The diffusivity of each species in each phase is required for the mass-transfer

calculations. Unfortunately, very little data is available for the system under

consideration. Therefore, a theoretical approach must be used. Assuming spherical

particles following Stokes law in a viscous liquid, the Nernst-Einstein relation for the

diffusivity is
193

kT ...[6.92]
6iitr

where r is the atomic/molecular radius of the particle and k is the B oltzman constant.

119
6.5.3 Kinetic Data
This equation requires atomic or molecular radii of the diffusing species, which

poses another problem: while the 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% ionic, while that in slags is closer to 50%.136

It is important to note that these numbers are for solid compound, and there may be some

differences for liquids. In fact, it has been suggested that the molten matte is completely

94 but there is no direct evidence of this. The relatively low amount of ionic
ionic,

bonding in mattes suggests that the diffusing species will be covalent molecules, while in

slags diffusion of ions is likely to predominate. Based on this assumption the required

radii are given in Table VI-XllI. Values of diffusivity caluculated using equation 6.92

for Fe
2 diffusion in slag and matte can be compared with measured values. A diffusivity

of 1 .4( 10h1) m
c is calculated for Fe
2 2 diffusion in a slag with an iron to silica ratio of

0.25 at 1400 K, while a value of 1.13(1 O) m


s is obtained for Pe
2 2 diffusion in matte at

the same temperature. These are lower than reported values, which range between
5(10h1) s and 5(108) m
2
m 98 and between 2.9(10) m
s in slag
2
95 s and 1.4(10) m
2 s
2

in

In the gas phase oxygen is the primary diffusing species and its diffusivity is given

4
by

120
6.5.3 Kinetic Data

Specie/ion radius Specie/ion radius


(nm) (nm)
As 0.125
4
Cu 0.096
Bi 0.17 24
Pb 0.12
Sb 0.145
24
Zn 0.074
Cu 0.128 24
Fe 0.076
Pb 0.175 3
Fe 0.064
Zn 0.133 2
Ni 0.072
FeS 0.241
24
Co 0.074
S
2
Cu 0.35 1 NiS 0.237
CoS 0.239 S
2
Ni 0.303
PbS 0.285 2
5i0 0.299
ZnS 0.239
Table VI-XllI Atomic/ionic/molecular radii of matte, slag and bullion
constituents.
36
= 3.754(10
)T2.31(10j
7 ...[6.93]
This equation is actually for oxygen diffusivity in argon, but calculations based on an

equation derived from the kinetic theory of gases indicate that the error involved in using

this equation is less than five percent.



20

6.5.3.2 Mass-Transfer Coefficients

The mass-transfer coefficient between the liquid and gas for rising bubbles is

calculated from
42

...[6.941
d,.
1
k ( dbw
= 1+1 l+
q3i

121
6.5.3 Kinetic Data
Unfortunately, there are no correlations available for mass-transfer on the gas side during

bubble rise, so it is assumed that equation 6.94 holds for both sides of the gas-liquid

interface. However, an equation for mass-transfer during drop formation has been

20
derived,

15
...[6.95]
D )rb
0
(v
2
kg =2.31
Ub

which can be applied to the bubble formation section of the model.

The liquid-liquid mass-transfer coefficient can be given as

...[6.961
k, =

where 8 is the boundary layer thickness. While values of boundary layer thickness are

not known, they can be approximately calculated if it is assumed that they are

comparable to the velocity boundary layer.

6.5.3.3 Heat-Transfer Coefficients

The gas-liquid heat-transfer coefficient is calculated from


143

Nu =2+ O.6Rep) . [6.97]


and the liquid-liquid heat-transfer coefficient between phases I and J is calculated from

h
1 ...[6.98]
h
= 6
1+ 6
h h
1

122
7.1.1 Bubble Growth
7 MODEL VALIDATION

7.1 Bubble Model Validation

7.1.1 Bubble Growth

There is a limited amount of data available concerning bubble growth during

horizontal injection for comparing with the model predictions, however some validation

can be carried out. Oryall and Brimacombe have reported the gas fraction distribution in

horizontal injection of air into mercury.


From this an approximate maximum bubble
1

penetration can be determined for comparison with the model. Table Vu-I compares the

model predicted maximum bubble penetration for four different injection conditions with

the 1% gas fraction line reported by Oryall and Brimacombe.


1 The maximum

difference is 6 mm, indicating a relatively good fit.

Gas Flow Tuyere Predicted Measured Predicted Measured


Rate diameter Penetration Penetration Width Width
)
1
s
3
(Nm (m) (m) (m) (m) (m)
0.002676 .00325 .076 .07 .047 .046
0.001615 .00325 .058 .064 .039 -

0.001844 .00476 .054 .055 .042 -

0.000710 .00325 .037 .043 .029 -

Table Vll-I Comparison of measured


and predicted gas penetration in mercury.
1

Further validation can be obtained from the bubble frequency. Hoefele and

Brimacombe have reported a bubble formation frequency of between 10 and 12 s_i in a

nickel converter.
202 The model predicts a variation in bubble formation frequency, with a

range of 9.5 - 10.5 s. Other studies have determined that the bubble frequency in copper

converters is affected by the extent of refractory wear.


203 The bubble frequency in a

5

newly relined converter was found to be between 8 and 14 s_I, but this value dropped

rapidly to a steady value close to 4 The rapid decrease in bubble frequency is

123
7.1.2 Bubble Rise
explained by the accelerated refractory wear around the tuyereline, which usually results

in the formation of a notch. There are two main mechanisms by which the notch reduces

bubble frequency; by reducing the bath flow rate at the tuyere tip, and by preventing the

section of the bubble close to the tuyere from rising. The notch may also increase tuyere

interaction, but this does not appear to affect the bubble frequency.
15 The model

assumes that there is no notch, so the agreement is good when compared to the newly

relined converter.

7.1.2 Bubble Rise

Verification of the bubble rise portion of the model is even more difficult. Only a

small amount of information relating to horizontal injection is available. The physical

modelling carried out by Adjei gives some data which can be used, but it can only give a

rough validation due to its imprecise nature.


37

The formula derived for calculating the bath velocity within the plume, gas holdup,

and spout height can be tested by comparing measured and predicted spout heights and

gas fractions. Figure 7.1 shows a comparison of calculated spout heights with those

. It can be seen that the model predicts a shorter spout height than
measured by Adjei
37

was measured. However, this is to be expected, as the spout heights were measured from

time lapse photographs, which tend to include the liquid ejected from the spout in the

spout itself, thus increasing the apparent size. This suggests that the predicted values are

at least approximately correct.

Physical modelling in bottom injection systems also provides some data which may

be used for verification. Table Vil-il shows a comparison of measured values of gas

fraction, bath velocity in the plume, and spout height with those predicted by the model.

It can be seen that the predicted spout heights are slightly higher than those measured,

124
7.1.2 Bubble Rise

018 -

V
0.17- TUYERE SUBMERGENCE (m)
0.16 -

086 A MEASURED
0.15 PREDICTED
v MEASURED
-

.117
0.14 -
PREDICTED
0.13 -

I
I 0.12 -
0
0.11 -

0 V
A
----------------
V
V
0.08 -

0.07

0.06 -

0.05

0.04- I I
0.001 0.003 0.005 0.007 0.009

GAS FLOW RATE m


1s

Figure 7.1. Comparison of measured


37 and predicted spout height.

while the predicted gas fractions are high at the low flow rate and low at the high flow

rate. However, the values of superficial gas velocity are not precise, due to uncertainties

in the cross-sectional area of the plume. Also, the data of Castillejos and

205 indicate that the average gas fraction varies with height above the

204
Brimacombe

tuyere in bottom injection, while Oryall and Brimacombe report that the gas fraction is

relatively constant in the plume region for horizontal injection.



11

The predicted bubble surface area and gas-phase mass-transfer can be tested by a

comparison of predicted and measured gas utilization. However, there is a problem with

the measured data, since Adjei


37 did not consider the effect of a surface reaction. In this

case the model calculates a considerably lower gas utilization than was measured.

Unfortunately there is not sufficient information available to calculate the extent of the

surface reaction. In particular, the diffusion boundary layer thickness on the bath surface

125
7.1.2 Bubble Rise

Gas Approximate Approximate Predicted Approximate Predicted Approximate Predicted Ref.


flow superficial gas fraction gas bath velocity bath spout height spout
rate gas velocity fraction )
1
(ms velocity (cm) height
s)
3
(cm (ms) )
1
(ms (cm)

371 .11 .12 .16 - - - - 204


1257 .18 .25 .22 - - - - 204
371 .11 - .16 4 4.7 206
876 .17 - .21 - - 6 6.9 206
1257 .18 - .22 - - 7 7.4 206
1630 .21 - .25 - - 8 9 206
167(1) .11 .26 .2 207
500(1) .32 .54 .39 207
1000(1) .66 .61 .57 207
167(2) .059 .15 .13 V 207
500(2) .17 .26 .27
V

207
1000(2) .45 .42 .41 207
100 .03 .05-.3 .08 .2-.3 .235 - - 122
(1) nozzle diameter = .28 cm
(2) nozzle diameter = .5 cm

Table VU-TI Comparison of measured and predicted gas fraction, bath velocity,
and spout height in vertical injection systems.

and the total bath surface area are required, but are not calculable. The diffusion

boundary layer thickness wifi be inversly proportional to the square root of the gas

velocity over the bath surface


, so will be reduced by increasing both the gas flow rate
208

and the tuyere submergence. Both of these factors will also increase the bath surface

area, and so have a large effect on the surface mass-transfer. If it is assumed that the bath

surface is flat and that the surface boundary layer thickness is a function of the gas flow

rate and tuyere submergence, then the predicted overall gas utilization can be fitted to the

measured values using

o = (.2374 20.06Q
+ .00256

J (1.621 (l0)
+ 4.023(10-5)]

126
7.1.2 Bubble Rise
and Figures 7.2 and 7.3 show the fit obtained. While these curves are fitted, they show
that the model is at least able to predict the general trends in gas utilization. It should

also be noted that some of the errors in gas utilization may be caused by the uncertainty

in the gas-phase masstransfer coefficient.

099 -

o
+
LU
0.95 -
.&. __

+ -

a: ---p
LU

Q o. ---.
o -

a:
D -

TUYERE SUBMERGENCE (m)


X MEASURED
0.85 - .057 PREDICTED

0.001 0.003 0.005 0.007 0.009

GAS FLOW RATE (ms)


Figure 7.2 Variation of fraction sulphur dioxide reacted with gas flow rate
including the fitted effect of surface reaction.

127
___________

7.2.1 Bath Temperature

0.95
C
uJ
0.9
w
a: 0.85
(ii
C
0.8
+
C
a: 0.75
D

j 0.7
D
Cl) MEASURED
0.65 PREDICTED
F
0
a:
LI
0.55

0.5 I I I I I I
0.01 0.03 0.05 0.07 0.09 0.11 0.13
BATH DEPTH (m)
Figure 7.3 Variation of fraction sulphur dioxide reacted with tuyere submergence
including the fitted effect of surface reaction.

7.2 Copper Converter Validation

The charges followed in the plant trials were simulated using the model. The

validity of the model can be tested by comparing predicted temperatures and

compositions.

7.2.1 Bath Temperature

The temperature variation for the three charges are given in Figures 7.4-7.6.

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 fits

the measured temperatures well.

128
7.2.1 Bath Temperature

1700

1600

1500

LU
a: 1400
D
I
a:
LU 1300
0
LU
I-
1200

1100

1000
0 100 200 300 400 500

TIME (mm)
Figure 7.4 Comparison of measured and predicted copper converter temperatures
charge 586, February, 1994.
1600

1500

LU 1400
D

UJ
a 1300
LU
I-

1200

1100
0 100 200 300 400 500

TIME (mm)
Figure 7.5 Comparison of measured and predicted copper converter temperatures
charge 588, February, 1994.

129
7.2.2 Slag Composition
1600

1550

1500
Ui
a:
D
1450
a:
Ui
a
LU
IU 1400

1350

1300
0 100 200 300 400 500

TIME (mm)
Figure 7.6 Comparison of measured and predicted copper converter temperatures
charge 595, February, 1994.

7.2.2 Slag Composition

A comparison of the measured and predicted slag compositions are given in

Tables Vil-ifi VII-V, and Figure 7.7 shows a comparison of the measured and predicted
-

slag iron and silica contents. The iron contents of the slags are predicted well for charges

586 and 595, but not as well for charge 588. Tn most cases, the amount of silica in the

slag is overpredicted. There are two factors which contribute to this overprediction: the

model includes all inert materials with the silica and a large, but unknown, quantity of

flux is lost as the converter begins blowing after a flux addition. This latter reason will

also explain some of the discrepancies in the iron assays.

The amount of copper in the slags is underpredicted in almost every case. The

exceptions being in charge 595, following the addition of copper slag to the converter. In

130
7.2.2 Slag Composition
one of these slags the copper content is overpredicted, while in the other it is very close to

the measured value. Zinc in slags is generally overpredicted, while lead is usually
overpredicted for the slags early in the charge, and underpredicted for later slags. The

arsenic content of the slags is underpredicted for charges 586 and 588, but is quite close

for charge 595, for which more accurate assays were available.

Slag Fe Cu Zn Pb 2
Si0 As
1 Model 45.4 2.81 4.62 .55 27.9 .02
Assay 41.6 7.13 3.97 0.7 17.6 .04
2 Model 40.0 0.85 4.55 0.18 38.0 .010
Assay 39.2 6.46 3.89 .73 22.0 .04
3 Model 38.7 0.78 4.29 0.16 39.5 .01
Assay 37.4 6.13 4.00 1.28 22.8 .04
Table Vil-ifi Comparison of model predicted slag compositions with assays,
(weight percent), #1 converter charge 586, Feb., 1994.

Slag Fe Cu Zn Pb 2
Si0 As
[ 1 Model 40.9 3.41 6.18 1.54 29.8 .0252
L_________ Assay 45.1 5.35 4.51 .67 21.8 .05
2 Model 49.7 2.19 5.38 .372 21.2 .021
Assay 36.5 7.04 3.82 .95 23.2 .05
Table Vll-IV Comparison of model predicted slag compositions with assays,
(weight percent), #1 converter charge 588, Feb., 1994.

131
7.2.2 Slag Composition

Slag Fe Cu Zn Pb 2
Si0 As Sb Bi
1 Model 36.9 13.9 4.19 1.41 26.3 .0241
Assay 39.8 8.1 3.93 0.8 18.8 .015
2 Model 39.5 12.6 4.61 1.27 23.8 .0218 .0019 .0002
Assay 38.3 12.4 3.80 0.7 17.8 .022 .0058 .0007
3 Model 44.1 6.44 3.72 0.645 26.2 .0113
Assay 38.2 4.16 4.38 1.45 25.5 .012
4 Model 48.7 2.43 3.68 0.239 24.9 .0044 .0014 .0003
Assay 40.6 3.52_- 4.44 2.06 24.3 .016 .0057 .0002
Table Vll-V Comparison of model predicted slag compositions with assays,
(weight percent), #1 converter charge 595, Feb., 1994.

60

I
Z 40
uJ
-

C.)
Lu
0
I
z
(!3
ij 30- A
A
A
Lu
IR0N
A A SILICA
20 -

10 I
10 20 30 40 50

MEASURED (WEIGHT PERCENT)

Figure 7.7 Comparison of measured and predicted iron and silica contents in
copper converter slags, charges 586, 588, and 595, Feb. 1994.

132
7.2.3 Matte Composition
7.2.3 Matte Composition
Matte and blister copper compositions are compared in Tables Vil-VI Vil-VIll,
-

and Figure 7.8 shows a comparison of the measured and predicted iron and copper

contents of all sampled mattes. The tables and the figure show that the predicted iron,

copper and sulphur in the matte are close to the measured values, up until the formation

of white metal. The one sample of metal taken during the copper blow of charge 586

(sample 4, Table Vil-VI) has a much higher copper content and correspondingly lower

sulphur content than is predicted by the model, this sample is not included in Figure 7.8.

This is most likely explained by entrainment of blister copper in the matte. The predicted

compositions of three mattes reported in the tables, one in Table Vil-Vil and two in

Table Vil-Vifi, are actually a combination of matte and slag in the correct proportions to

give the assayed silica content of the mattes. Even with this correction, however, the

model predictions are not good. Some later mattes in charge 595 also had a relatively

high silica content, (see Table IV-TV), but the model predicted silica contents close to the

assayed values so no correction was made.

Zinc is predicted reasonably well for charges 586 and 588, but is consistently

underpredicted for charge 595. The lead content of the mattes is overpredicted for almost

every sample. This, combined with the underprediction of lead content in the slags,

indicates that there is more oxidation of lead from the matte, particularly at higher matte

grades. Matte arsenic contents are predicted well for all charges, and the antimony and

bismuth predictions for charge 595 are also reasonable, with the exception of the bismuth

in sample 9 which is too high.

133
7.2.3 Matte Composition

Matte Fe Cu Zn Pb S As
1 Model 6.24 68.6 1.1 2.21 19.3 .032
Assay 4.1 70.9 .83 2.89 20.2 .03
2 Model 4.18 71.4 .69 2.38 19.0 .031
Assay 3.09 71.0 .59 2.63 21.6 .02
3 Model 1.13 74.9 .093 3.24 18.1 .033
Assay 1.1 74.7 .3 .65 19.8 .01
4 Model .62 73.9 .125 4.38 17.7 .045
Assay .49 90.1 .13 .23 5.55 .04
Blister Model - 99.707 .1449 .145 1 - .0020
Assay - 99.04 .002 1 .0072 - .0073
Table Vil-VI Comparison of model predicted matte and blister copper
compositions with assays, (weight percent), #1 converter charge 586, Feb., 1994.

__Matte Fe Cu Zn Pb S As
1 Model* 22.7 40.2 1.41 3.0 12.3 .0351
Assay 38.6 23.2 4.82 .47 7.7 .03
2 Model 4.03 71.5 .42 1.21 18.9 .0365
Assay 2.87 70.3 .66 .51 20.8 .03
3 Model 1.11 73.9 .19 1.11 18.3 .0346
Assay 1.43 74.0 .28 .4 19.7 .02
Blister Model - 99.68 .1729 .1731 - .0006
Assay - 97.89 .0044 .0072 - .0096
Table Vu-Vu Comparison of model predicted matte and blister copper
compositions with assays, (weight percent), #1 converter charge 588, Feb., 1994.
Asterisk indicates combined matte and slag to bring the matte silica content up to assayed
value.

134
7.2.3 Matte Composition

Matte Fe Cu Zn Pb S As Sb Bi
1 Model* 17.1 52.2 1.81 1.38 16.8 .0248
Assay 23.3 37.3 2.75 1.00 15.1 .022
2 Model* 18.4 49.3 1.9 1.43 14.8 .0221
Assay 25.6 31.2 2.95 .8 11.1 .019
3 Model 9.29 62.0 .973 1.5 18.2 .0202 .0026 .0016
Assay 11.8 57.1 1.6 1.17 17.9 .014 .0028 .0013
4 Model 10.9 60.8 .747 1.56 18.4 .0188
Assay 14.5 49.8 1.83 1.01 17.1 .02
5 Model 10.3 62.0 .451 1.62 18.4 .0147
Assay 9.01 58.4 1.31 1.13 18.4 .014
6 Model 10.6 62.1 .292 1.7 18.9 .0156 .0021 .0014
Assay 7.43 63.2 1.33 1.21 20.1 .018 .0022 .0014
7 Model 12.0 60.9 .395 1.7 19.5 .0172
Assay 8.89 58.5 1.48 1.18 20.0 .016
8 Model 10.7 62.6 .22 1.69 19.4 .0171
Assay 8.45 59.0 1.45 1.32 21.2 .015
9 Model 8.2 65.6 .0658 1.79 19.2 .0167 .0022 .0015
Assay 7.87 62.7 1.3 1.15 20.3 .015 .0019 .0002
10 Model 6.28 68.1 .172 1.84 19.3 .0182
Assay 4.15 65.7 .75 .89 19.9 .012
11 Model 2.59 72.5 .0315 1.98 18.7 .0175
Assay 4.39 62.8 .82 .73 17.3 .01
Table Vil-Vill Comparison of model predicted matte compositions with assays,
(weight percent), #1 converter charge 595, Feb., 1994. Asterisk indicates combined
matte and slag to bring the matte silica content up to assayed value.

135
7.3 Nickel Converter Validation

a. 15

I
z
w

5 10 15
MEASURED (WEIGHT PERCENT)

b.

70

2
uJ
II 65
uJ
I
I
0
ij 60

45
45 50 55 60 65 70 75
MEASURED (WEIGHT PERCENT)

Figure 7.8 Comparison of measured and predicted a. iron and b. copper contents in
copper converter mattes, charges 586, 588, and 595, Feb. 1994.

7.3 Nickel Converter Validation

As with the copper converter, there is phase composition and temperature data

available for the nickel converter which may be compared to the model predictions. In

this case, however, there is considerably more temperature data available, but the

136
7.3.1 Bath Temperature
compositions are less accurate.

7.3.1 Bath Temperature

A comparison of the measured and predicted bath temperatures are given in

Figures 7.9-7.12. The measured temperatures were obtained using a two colour

pyrometer mounted in the hood and aimed at the bath surface, so in most cases should

give the slag temperature. However, the figures 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 figures that the temperature fit is 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 are 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 first blow of the following

charge, but its composition, weight, and temperature are all unknown. It is interesting to

note that 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 sufficient 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.

137
7.3.1 Bath Temperature

1600

1550

Ui
1500
D
I
ct
Ui
0.
LU 1450
I-

1400

1350
0 100 200 300 400 500 600 700 800

TIME (miii)

Figure 7.9 Comparison of model predicted matte and slag temperatures with plant
data, #3 converter charge 105, May 1988.

1550

1600

Ui
a
1450
LU
I-

1400

1350
100 200 300 400

TIME (mm)

Figure 7.10 Comparison of model predicted matte and slag temperatures with plant
data, #3 converter charge 106, May 1988.

138
7.3.1 Bath Temperature

1550

1500

w
D
1450
w
0
w
I-

1400

1350
0 100 200 300 400 500 600 700
TIME (mm)
Figure 7.11 Comparison of model predicted matte and slag temperatures with plant
data, #3 converter charge 107, May 1988.

1550

1500

Lii
D
I
1450
Lii
a
w
I

1400

1350
0 100 200 300 400 500

TIME (mm)
Figure 7.12 Comparison of model predicted matte and slag temperatures with plant
data, #3 converter charge 108, May 1988.

139
7.3.2 Slag Composition
7.3.2 Slag Composition

A comparison of measured and predicted slag compositions are given in

Tables Vil-IX VII-XII, and Figure 7.13 shows a comparison of the measured and
-

predicted iron and silica contents of the slags. Of the minor elements, only lead and zinc

are included in the tables, because the measured concentrations of arsenic, antimony, and

bismuth were below the lower threshold of the assay technique. Also, the Si0
includes
2
the other inert materials, such as alumina and lime. In all charges the Si0
values
2

predicted are close to the 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 are a number of possible reasons for this,

including insufficient 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 slag, which is not accounted for in the model. Tables IV-Vffl IV-XI-

indicate that there is a relatively large amount of sulphur in the early slags, which

indicates a correspondingly large amount of entrainment. The amounts of lead and zinc

in the slag are predicted fairly well, but the zinc predictions are high for the last slags.

The reason for this is most likely for the same as for the increased nickel and cobalt.

140
7.3.2 Slag Composition

Slag Fe Ni Cu Co 0 Pb Zn
2
Si0
1 Model 54.5 .58 .37 .35 21.0 .052 .061 22.7
Assay 51.1 .97 .57 .59 22.2 .042 .055 24.4
2 Model 52.4 .24 .09 .05 19.8 .021 .061 26.9
Assay 51.7 .77 .48 .53 17.6 .035 .057 28.6
3 Model 49.3 .3 .08 .04 18.7 .021 .05 31.1
Assay 50.6 1.02 .49 .6 16.3 .041 .055 31.0
4 Model 47.5 .29 .09 .04 16.3 .025 .15 35.0
Assay 49.7 .9 .44 .67 14.7 .05 .052 33.6
5 Model 33.6 10.7 .08 5.23 17.0 .022 .54 32.3
Assay 48.6 3.39 1.43 1.14 15.3 .059 .047 30.1
6 Model 21.0 17.4 .11 3.12 13.6 .032 .26 43.8
Assay 44.6 3.35 1.07 1.55 15.5 .078 .04 34.0
Table Vil-IX Comparison of model predicted slag compositions with assays taken
at the end of the blow, (weight percent), #3 converter charge 105, May 1988.

Slag Fe Ni Cu Co 0 Pb Zn
2
Si0
1 Model 49.0 .82 .64 .51 18.7 .071 .052 29.9
Assay 49.8 2.28 1.1 .79 17.6 .049 .054 28.4
2 Model 51.1 .78 .59 .46 19.7 .064 .052 26.9
Assay 51.1 2.6 .86 1.37 13.6 .061 .044 30.5
3 Model 47.2 1.76 1.62 2.69 19.3 .017 .286 26.2
Assay 55.6 1.17 .71 .51 17.4 .036 .079 24.6
4 Model 25.9 21.1 2.33 2.13 16.4 .014 .153 30.7
Assay 46.0 3.95 1.23 1.56 16.7 .0833 .0418 30.56
Table Vll-X Comparison of model predicted slag compositions with assays taken
at the end of the blow, (weight percent), #3 converter charge 106, May 1988.

141
7.3.2 Slag Composition

Slag Fe Ni Cu Co 0 Pb Zn
2
Si0
1 Model 55.5 .84 .32 .22 20.8 .041 .044 21.9
Assay 55.2 1.39 .75 .54 16.2 .033 .061 25.9
2 Model 51.9 .83 .42 .06 19.5 .014 .04 25.8
Assay 52.6 .99 .54 .60 16.3 .035 .051 29.0
3 Model 49.5 .88 .45 .038 18.7 .01 .036 28.7
Assay 48.6 4.37 1.72 .75 15.0 .038 .044 29.6
4 Model 46.4 .61 .29 3.45 18.7 .011 .29 29.2
Assay 48.0 3.46 1.32 .78 15.3 .042 .047 31.1
5 Model 30.5 12.1 .47 3.93 16.0 .014 .15 35.1
Assay 42.7 4.79 1.85 1.09 14.3 .064 .036 35.3
6 Model 22.6 21.9 .33 2.35 15.1 .015 .096 36.2
Assay 39.3 6.0 2.22 1.36 13.9 .067 .033 37.2
Table VII-XI Comparison of model predicted slag compositions with assays taken
at the end of the blow, (weight percent), #3 converter charge 107, May 1988.

Slag Fe Ni Cu Co 0 Pb Zn
2
Si0
1 Model 56.4 1.56 .47 .46 21.8 .033 .057 18.5
Assay 50.4 2.27 .95 .66 15.9 .046 .05 26.5
2 Model 51.4 .73 .33 .654 19.4 .011 .046 27.4
Assay 49.2 1.54 .63 .78 15.2 .054 .047 32.6
3 Model 37.5 7.85 .16 2.87 17.1 .013 .095 33.9
Assay 46.8 2.82 .86 1.31 14.8 .064 .043 33.4
Table Vll-XII Comparison of model predicted slag compositions with assays
taken at the end of the blow, (weight percent), #3 converter charge 108, May 1988.

142
7.3.3 Matte Composition

60

I
Z 50
w
C-)
cc
w
0
I
C3
i:i 40

w
I
0
0
w
cc
0

20
20 30 40 50 60

MEASURED (WEIGHT PERCENT)


Figure 7.13 Comparison of measured and predicted iron and silica contents in
nickel converter slags, #3 converter, charges 105, 106, 107, and 108, May 1988.

7.3.3 Matte Composition

A comparison of predicted and assayed matte compositions are given in

Tables VII-XIll Vll-XVI, and a comparison of measured and predicted iron, nickel, and
-

copper contents of all matte samples is given in Figure 7.14. The tables and the figure

indicate that the model predictions are relatively close for the major elements, although

the relative amounts of nickel and copper do not always correspond. This will be due to

uncertainties in the feed compositions. Of the minor elements, zinc is generally predicted

well, but lead and arsenic are consistently overpredicted. This could be the result of

uncertainties in the diffusion coefficients, or an underprediction of the matte/slag

interfacial area.

143
7.3.3 Matte Composition

Matte Fe Ni Cu Co 0 S Pb Zn As
1 Model 16.2 38.2 14.6 1.61 1.52 27.3 .1 .016 .02
Assay 16.9 38.8 13.8 1.60 3.54 25.2 .08 .012 .02
2 Model 9.76 43.9 15.9 1.82 .88 26.9 .52 .005 .049
Assay 11.4 42.5 15.4 1.45 3.97 24.9 .071 .004 .022
F Model .82 50.8 22.4 .15 .29 24.1 .7 .001 .137
Assay .792 48.3 23.5 .949 3.34 22.4 .0625 .003 .025
Table VII-XIII Comparison of model predicted matte compositions with assays,
(weight percent), #3 converter charge 105, May 1988.

_Matte Fe Ni Cu Co 0 S Pb Zn As
1 Model 9.89 41.9 20.9 1.37 .715 24.35 .082 .011 .044
Assay 9.83 42.5 18.4 1.27 3.64 24.1 .069 - .017
2 Model 2.42 48.1 24.1 .919 .834 22.3 1.1 .001 .044
Assay 5.79 46.8 21.7 1.03 1.91 22.5 .085 - .017
F Model .8 50.7 27.6 .28 .28 21.1 .686 .0002 .133
Assay .96 51.4 24.1 .79 .83 21.8 .058 - .023
Table VII-XIV Comparison of model predicted matte compositions with assays,
(weight percent), #3 converter charge 106, May 1988.

Matte Fe Ni Cu Co 0 S Pb Zn As
1 Model 16.4 37.7 14.6 1.33 1.72 27.5 .39 .007 .02
Assay 14.6 42.8 14.7 1.53 .66 25.5 .085 .004 .02
F Model .78 50.2 22.8 .134 .293 24.1 .4 .0004 .33
Assay 1.17 51.7 21.9 1.10 .8 23.0 .063 - .035
Table VII-XV Comparison of model predicted matte compositions with assays,
(weight percent), #3 converter charge 107, May 1988.

144
7.3.3 Matte Composition

Matte Fe Ni Cu Co 0 S Pb Zn As
1 Model 12.0 41.9 13.7 1.58 1.02 27.1 .19 .01 .041
Assay 12.5 45.5 14.8 1.53 .09 25.5 .0916 .005 .0263
2 Model 8.02 45.3 14.4 1.78 .796 27.0 .89 .003 .061
Assay 12.4 44.9 14.9 1.39 .71 24.5 .0813 .0074 .0243
F Model 1.15 50.7 18.7 .366 .431 25.1 1.38 .0001 .105
Assay 1.10 55.6 19.5 1.05 - 23.0 .061 - .033
Table VII-XVI Comparison of model predicted matte compositions with assays,
(weight percent), #3 converter charge 108, May 1988.

60

60- A A
I
z
w A
C) A
40-
w
a A
I-.
(5
w 30
x
-

0
w -
20 -
x-
0
w _
IRON
a:
a. 4__ A NICKEL
10 X COPPER
-

0 w_______ I I

0 10 20 30 40 50 60

MEASURED (WEIGHT PERCENT)

Figure 7.14 Comparison of measured and predicted iron, nickel and copper
contents in nickel converter mattes, #3 converter, charges 105, 106, 107, and 108, May
1988.

145
7.4.1 Phase Compositions
7.4 Discussion

7.4.1 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 coefficients, diffusivities, interfacial areas, and assayed

compositions. Certain assumptions made in the modelling may also add to the errors.

7.4.1.1 Error in assays

Errors in assayed compositions are a particular problem in the nickel converter

analysis, because of the assay technique used. This point is best illustrated by comparing

two sets of assays for the same material. These are available for slag samples obtained

from charge 110 during the plant trials at Copper Cliff. Table VII-XVII shows a

comparison of assay results for the major components obtained using the standard ICP

technique, by which all other assays of the nickel converter samples were obtained, with

wet assay techniques. It is evident that there are some large differences between the

values, so the model predicted values are, for the most part, within the accuracy of the

assays.

146
7.4.1 Phase Compositions

Slag Assay technique Cu Ni Co Fe 2


Si0
1 ICP 1.34 3.10 0.718 54.2 21.9
wet 0.9 2.3 0.74 52.1 22.7
2 ICP 0.54 1.1 0.586 55.4 23.8
wet 0.53 1.32 0.62 50.9 24.4
3 ICP 0.673 1.58 0.712 53.4 25.6
wet 0.59 1.58 0.75 50.1 25.9
4 ICP 0.784 2.12 0.839 51.6 27.8
wet 1.41 4.33 0.95 45.5 29.0
5 ICP 0.911 2.99 1.37 48.7 29.6
wet 0.92 3.23 1.6 45.8 28.7
Table VII-XVII Comparison of slag assays obtained using ICP and wet assay
techniques.

7.4.1.2 Sulphur in slag

Neglecting the presence of sulphur in the slag will have some effect on the

predicted slag compositions. The sulphur 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 sulphur content under 0.9%.89 There are two main

mechanisms by which matte entrainment may occur; emulsification of matte in slag and

splashing from the spout. The former mechanism is not usually reported in physical

modelling studies, but has been reported in one experimental system,


209 and is found in

top blowing and combined blowing systems.


21 This suggests that the majority of

210

entrained matte is produced by splashing from the spout, that is, matte which is either

carried out of the spout by its own momentum, or ejected from the spout by collapsing

bubb1es. 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

147
7.4.1 Phase Compositions
single value of a suspension index as used by Nagamori and Mackey for the Noranda

reactor will not be valid for this system. While it may be possible to produce a formula

for suspension indices to give a better fit between the predicted and assayed slag

compositions, it would not be particularly meaningful considering the other inaccuracies

involved in the plant trials.

7.4.1.3 Oxygen in matte

For the modelling it was assumed that all of the oxygen in the matte was in the form

of Fe
, however, it is probable that there is also some FeO present. The equilibrium
4
0
3

mole fraction of FeO in matte may be obtained from the reaction

1 ...[7.3]
3FeO+O = Fe
2 4
0
3

The oxygen potential of the matte may be obtained from the magnetite formation

reaction, and the activity coefficient of FeO in matte is given by



2

YFeO = exp{
J (5.1 + 6.2 logX + 6.41 (1ogX )
2
+ 2.8(logXcus)3)}

The resulting variation of Fe


4 and FeO with time, calculated by the model for charge
0
3

586, are given in Figure 7.15. While the mole fraction of FeO is as high as 50% of the

magnetite mole fraction, the relative proportions of iron and oxygen present as FeO will

be considerably less. The effect of including FeO will be to increase the iron content in

the matte slightly.

148
7.4.1 Phase Compositions
0.035

0.03

0.025
z
0
0.02

0.015
0

0.01

0.005

0 100 200 300 400 500


TIME (mm)

Figure 7.15 Calculated variation of mole fractions of Fe


4 and FeO in matte.
0
3

7.4.1.4 Minor element distribution

The magnitude of the errors in the assays, along with the uncertainties inherent in

industrial scale tests raises some questions regarding the validity of any analysis of minor

element behaviour. However, for the purpose of the analysis of the copper converter,

precise values are not as important as overall distributions. If the model predicts the

measured distributions correctly, then the transport mechanisms assumed are likely to be

correct. If this is the case, then the model should still predict the effects of changes in

operation on the minor element distribution correctly.

Figure 7.16 shows a comparison of the measured and model predicted distributions

of lead, zinc and arsenic in the copper converter. From the figure it can be seen that the

model predicts the distribution of zinc and arsenic quite well. The amount of lead

reporting to the matte/blister copper is also predicted well, but the lead distribution

between the slag and gas is not close. The most likely cause of this is the behaviour of

lead at the gas/liquid interface. The model assumes that it vapourizes, in a manner

149
7.4.1 Phase Compositions

similar to arsenic, while it appears from the distributions that at least part of the lead

reaching the interface is oxidized and reports to the slag. The comparisons of predicted

and assayed slag compositions suggest that the majority of this oxidation takes place

towards the end of a charge. The free energies of formation of lead and copper oxides are

very close over the temperature range seen in converters, so it is likely that some of the

lead will oxidize towards the end of the slag blows and during the copper blow.

A comparison of the measured and predicted antimony and bismuth distributions

for charge 595 is given in Figure 7.17. This shows that the model can predict these

distributions well, particularly the proportion reporting to the matte. The difference

between the predicted and actual antimony in the slag indicates that some oxidation may

be occurring.

The predicted minor element distributions in the nickel converter are not as close as

for the copper converter. Figure 7.18 shows that, while the arsenic distribution is

predicted reasonably, the amount of lead and zinc reporting to the slag is underpredicted.

The most likely explanation for this is that the activity coefficients of the minor elements

used in the model are for copper mattes, and may be considerably different for nickel

mattes.

150
_____ ______________

7.4.1 Phase Compositions

100
Pb Zn As

e x
80
MEASURED +
586
588
A595
w
C/) +
< 60
PREDICTED
0 x586
+588
i595
40
0
0
w
+
o
- A
20

0 -I
SLAG MATTE GAS SLAG MATTE GAS SLAG MATTE GAS

Figure 7.16 Comparison of model predicted lead, zinc and arsenic distributions
with measured values, copper converter charges 586, 588, and 595, Fun Flon, February
1994. Solid lines indicate commercially observed range.
31

151
7.4.1 Phase Compositions

100

MEASURED +
Sb Bi
80
PREDICTED
w XSb Bi
(I)

I
60
0

I
(D
z
I
o 40
x
x

20 -

+
0 I

SLAG MATTE GAS

Figure 7.17 Comparison of model predicted antimony and bismuth distributions


with measured values, copper converter charge 595, Fun Flon, February 1994.

152
7.4.2 Copper Blow

100
Pb MEASURED As
0 105
107
108
80
PREDICTED
x 105
Iii + 107
(I)
.. 108
:i: +
0 60
0
I

z x
I
a::
0
a 40
w
cr
+ x
A *

20
x
+
+
-I. -I.

0
SLAG MATTE GAS SLAG MATTE GAS SLAG MATTE GAS

Figure 7.18 Comparison of model predicted lead, zinc and arsenic distributions
with measured values, nickel converter charges 105, 107, and 108, Copper Cliff, May
1988.

7.4.2 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. Towards 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 sulphur content of the matte significantly.

153
7.4.2 Copper Blow
To determine the reaction products of the matte-slag reaction, the most stable

species can be found from a predominance area diagram. Figure 7.19 shows

predominance area diagrams for the iron-copper-sulphur-oxygen system at 1300 K and

1500 K. These are equilibrium diagrams which assume an ideal solution, but should give

a good idea of which species will be produced by the reaction. It is interesting to note

that Cu
0 is not stable at either temperature, with the oxidized copper being in a copper
2

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, so the product of the matte-slag

reaction is most likely to be 2


0.Fe
Cu
.
3
0

This suggests that the main reaction between the matte and the slag at this stage is

likely to be

S
2
Cu + 4=2
O
3
4Fe 0.Fe
Cu
3
0 + lOFeO + SO
2

This reaction will reduce both the oxygen and sulphur 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 or not any reaction occurs beyond this point will depend on the

sulphur 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

insufficient 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.
211
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

154
7.4.2 Copper Blow

-2

-4
a:
D
-6 (1)
U)
Lii
a:
-8 0
-J

-10
a-
-12
:i:
a-
-14
U)
(!3
- 0
-j

-18

-20
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2 0
LOG(OXYGEN PARTIAL PRESSURE)
0

-2

-4 ii;
a:
-6 U)
(I)
w
a:
-8 a-
-J

-10
a-
-12
a-
-14
U)

-IQ
0
-j

-18

-20
-20 -18 -16 -14 -12 -10 -8 -6 -4 -2 0

LOG(OXYGEN PARTIAL PRESSURE)

Figure 7.19 Predominance area diagrams for the iron-copper-sulphur-oxygen


system; a. 1300 K, b. 1500 K. Dashed lines indicate partial pressure of sulphur dioxide.

155
7.4.2 Copper Blow

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.

156
8.1 Gas Flow Model

8 MODEL PREDICTIONS AND DISCUSSION


8.1 Gas Flow Model

The predictions of the gas flow section of the model are of interest by themselves,

and so are considered separately here. The following discussion will consider a single

bubbling event, and only follow the gas phase. However, the results of this portion of the

model will be useful in discussing the predictions of the complete model.

The model predicted gas temperature is shown in Figure 8.1, for the standard

conditions shown in Table V-I. All reaction heat is assumed to report to the liquid phase,

so before detachment the gas temperature is controlled by a balance between the heat

input from the bath and the cooling effect of the injected gas. Initially, the heat input

predominates, but as the bubble grows the two factors become more equal. This causes

the gas temperature to level out below the bath temperature. In some cases, as the bubble

grows the gas temperature starts to decrease until breakup occurs and the temperature

starts to rise again. This effect is small, and does not show on Figure 8.1. After

detachment at 0.107 s, the gas temperature rises quickly to the bath temperature, and the

primary bubble exits the bath 0.113 s after detachment.

A comparison of the temperatures predicted by the present model with those

35 shows that the changes made in the initial assumptions


predicted by Ashman et al.

cause considerable differences. The present model predicts a much faster temperature

increase, due both to the bubble shape and the heat-transfer coefficient. An elliptical

bubble has more surface area than a spherical one of equal volume (Figure 8.2) giving an

increased area available for heat transfer. Also, Ashman et al.


35 used a constant

1 which is considerably lower than that predicted


heat-transfer coefficient of 290 Wm
K
2

by equation 6.97, and does not account for variations with bubble size. Bubble breakup

157
8.1 Gas Flow Model

1400

w
1200

Lu

E:
1000

400

0 0.05 0.1 0.15 0.2

TIME (s)
Figure 8.1 Variation of primary bubble temperature with time for injection
conditions given in Table V-I.

during growth will also affect the gas temperature, because larger bubbles heat up more

slowly than small bubbles. Figure 8.3 shows that four bubbles break off from the

primary bubble before detachment, and that the primary bubble breaks up one more time

before leaving the bath. At the bath surface the primary bubble has a volume of 0.05 m
3

and represents approximately 34% of the gas input during a single bubbling event.

Figure 8.4 shows that there is an initial rapid decrease in the total bubble oxygen

content during bubble growth, but the rate of decrease drops off before detachment. This

is caused by the reduced mass-transfer rate due to increased bubble size. After

detachment the oxygen content of the gas decreases more rapidly, but does not reach the

initial rate, since there is no longer enhancement of the mass-transfer rate due to the gas

injection. After 0.22 s the oxygen partial pressure becomes essentially constant. This is

because only small bubbles are remaining in the bath. These bubbles contain a relatively

158
8.1 Gas Flow Model
small proportion of the total gas and, due to their size, already have a low oxygen partial

pressure. Therefore, the amount of oxygen removed from these bubbles is insignificant

when compared to the oxygen which has already exited the bath in the larger bubbles. It

is evident that a relatively large proportion of the oxygen is not reacted (11.5%), which

appears to contradict the high oxygen efficiencies reported for converters. If the tuyere

submergence is increased to 0.7 m the partial pressure of oxygen remaining decreases

slightly to 0.015 atm. This value is approximately half of that reported by Rodoiff and

98 for a copper converter with a tuyere submergence of 0.76 m and a Froude number
Rana

of 10.92. This difference may be related to the amount of slag present in the actual

converter, or could indicate that the gas-phase mass-transfer coefficient used in the model

is too high.

The effect of changing the tuyere diameter on the primary bubble temperature is

small, with the main difference being caused by slightly different detachment times.

However, decreasing the tuyere diameter causes a significant increase in oxygen use, as

shown in Figure 8.5. This is caused by the increased gas velocity which increases the

rate of mass transfer during bubble formation. A similar trend was reported by Rodolff

and Rana
98 and is also predicted for changes in flow rate. The effect of tuyere diameter

on the extent of bubble break up during growth is large. Figure 8.6 shows that increasing

the tuyere diameter significantly increases the number of times the bubble breaks up.

This is caused by the reduced gas velocity which reduces the equilibrium bubble size as

explained in Section 5.3. The reduced slope of the oxygen in gas curves after

detachment at lower tuyere diameters (Figure 8.5) is a result of the reduced break up as

well as the lower oxygen concentration. Mass-transfer of oxygen from the bubbles is

reduced both by the larger bubble size and the lower driving force.

159
8.1 Gas Flow Model
2.5

1.5

0 0.2 0.4 0.6 0.8


ECCENTRICITY

Figure 8.2 Effect of ellipse eccentricity on surface area at constant volume, shaded
area indicates range of eccentricity predicted by the model.

0.07

0.06

0.05

0.04
D

0.03

-
0.02

0.01

0
0 0.05 0.1 0.15 0.2

TIME (s)
Figure 8.3 Variation of primary bubble volume with time for injection conditions
given in Table V-I.

160
8.1 Gas Flow Model

0.2

z
UI
>-
><
0 0.15
0
UI
D
Cl)
C)
UI
a: 0.1
a-
-J

I-
a:
0.
Ui 0.05
a:
UI

0
0 0.1 0.2 0.3 0.4

TIME (s)

Figure 8.4 Variation of average oxygen partial pressure with tuyere submergence
for injection conditions given in Table V-I.

0.2

z
UI
(
>-
x
0 0.15
U
0
UI
a:
D
Cl)
Cl)
UI
a: 0.1
a-
-J

I
a:
a-
UI
(3 0.05
a:
UI
>

0
0 0.1 0.2 0.3

TIME (s)

Figure 8.5 Variation of average oxygen partial pressure with tuyere diameter, all
other conditions given in Table V-I.

161
8.1 Gas Flow Model

0.08

0.07

0.06
w
D
_i 0.05
0
>
w
0.04

D
>. 0.03

0.02
a-

0.01

0
0 0.05 0.1 0.15 0.2
TIME (s)
Figure 8.6 Variation of primary bubble volume with tuyere diameter, all other
conditions given in Table V-I.

Increasing the oxygen content of the gas results in an increase in the oxygen partial

pressure of the off gas. However, Figure 8.7 shows that the actual amount of oxygen

reacted is also increased. There is also a corresponding increase in the partial pressure of

sulphur dioxide in the off gas. These are the result of a higher oxygen mass-transfer rate

caused by the higher partial pressure of oxygen in the bulk of the gas.

The model predicts a slightly higher oxygen utilization in the copper blow than in

the slag blow, which was also found by Rodoiff and Rana.
98 However, the lowest value

of oxygen remaining was attained for the case of blowing directly into slag. The primary

reasons for this can be seen from Figures 8.8 and 8.9. When blowing into slag the gas

temperature rises slowly, due to the low slag thermal conductivity, Also, the higher

162
8.1 Gas Flow Model

viscosity of the slag reduces the bubble velocity, allowing increased bubble breakup.

These result in the formation of a larger number of smaller bubbles, which increases the

amount of mass-transfer, as well as the total bubble residence time.

The effect of tuyere interaction can be approximated by artificially doubling the gas

input to the bubble, while keeping the gas flow rate through the tuyere constant. The

model predicts that tuyere interaction will have a small effect on the primary bubble

temperature. The initial heating rate is lower than with no interaction, due to the larger

bubble size, but the bath temperature is still attained shortly after detachment. The

oxygen utilization, however, is reduced considerably, as shown in Figure 8.10. This

effect is also caused by the increased bubble size during the formation stage,

(Figure 8.11) which reduces the mass-transfer rate. This indicates that heat and mass

transfer effects are not exactly analogous. Figure 8.11 also shows that even with the

increased bubble size, the primary bubble still breaks up only four times before

detachment. This effect is again caused by the increased growth velocity which allows a

larger bubble to form.

163
8.1 Gas Flow Model

0.25

Cl)
w 0.2
-J
0

0
Ui
I 0.15
0

a:
z
Ui
0 0.1
>-
><
0

0.05

0
0 0.05 0.1 0.15 0.2 0.25 0.3
TIME (s)
Figure 8.7 Variation of total oxygen reacted with inlet gas oxygen content, all other
conditions given in Table V-I.

1600

1400

1200

Ui
1000

Ui
0
Ui
F-
600

400

200
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
TIME (s)
Figure 8.8 Variation of primary bubble temperature with bath material, all other
conditions given in Table V-I.

164
8.1 Gas Flow Model

25

20

15
Ui
-J
LU
LU
D
LU
10

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
TIME (s)
Figure 8.9 Variation of number of bubbles formed with bath material, all other
conditions given in Table V-I.

0.2

z
Ui
ci,
0 0.15
U
0
Ui
cc
D
CO
CO
Ui
cc 0.1
0
-J

I-
cc

Ui
0.05

Ui
>

0
0 0.05 0.1 0.15 0.2 0.25 0.3
TIME (s)
Figure 8.10 Effect of tuyere interaction on average oxygen partial pressure all ,

other conditions given in Table V-I.

165
8.2.1 Introduction

0.15

COS

w
D
0
>
w
-J
XI
D
m
>_
0.05

ci

0
0 0.05 0.1 0.15 0.2
TIME (s)
Figure 8.11 Effect of tuyere interaction on primary bubble volume, all other
conditions given in Table V-I.

8.2 Converter Model

8.2.1 Introduction

For the purposes of this discussion, a single copper converter charge is used to test

the effects of variations of model and operating parameters. While none of the industrial

charges followed were normal, charge 586 can be considered as representing the final

two slag blows and the copper blow of a normal charge, because the transferred matte

was an initial charge to the converter. As in a previous work, the discussion will focus on

the matte temperature and iron content,


98 but other properties will be discussed as

required.

166
8.2.2 Copper Converter Charge

8.2.2 Copper Converter Charge

The predicted matte temperature from charge 586 is shown in Figure 8.12.

Although there are some similarities between the temperature variations of copper and

nickel converter charges, the temperature of the copper charge does not follow the regular

pattern exhibited by the nickel converter temperature (Figure 4.1). While the figure does

show the same rapid temperature increase during blowing, there is not the initial

temperature decrease that is seen in the nickel converter. This difference is caused by the

fluxing methods used. At Copper Cliff, the flux is added through the mouth of the

converter while it is blowing, usually over the first ten to fifteen minutes of each blow.

At Flin Ron, flux is added from ladles as it is needed throughout the charge. Thus,

instead of one initial, large temperature drop, there are a few small drops throughout the

blow (marked F on Figure 8.12). The weight of flux required is also much lower at Flin

Flon, due to the smaller converter size and the higher initial matte grade.

The matte temperature falls only slightly during idle periods. This is caused by the

insulating effect of the slag, which prevents radiation losses. The only heat losses from

the matte are through the walls and to the slag, both of which are small. Some probable

causes of the low predicted temperatures during the slag blow have been discussed in

Section 7.4, but a further examination is required. A temperature decrease, as predicted

during the copper blow, can be caused by material additions or reduced heat inputs to the

bath. Since extra material is only added at the times indicated in Figure 8.12, it follows

that the heat input has been reduced. There are two further reasons for this decline in

heat production. In the copper blow, copper suiphide is being oxidized rather than iron

sulphide, so the principle reaction is

167
8.2.2 Copper Converter Charge

1600

1500

1400
D

1300

1200

1100
0 100 200 300 400 500

TIME (mm)
Figure 8.12 Predicted matte temperature, charge 586. F-flux addition, M-matte
addition, I-idle period start, I*idle period end, S-slag skimmed, Sc-scrap added,
C-copper blow start.

S + O = 2Cu + SO
2
Cu 2 AHR = 190.4kJ moF
whereas, in the slag blows, the reaction is

=FeO+SO
FeS+O
2 AIIR =308.9kJmoF
1

Thus the heat produced per mole of oxygen reacting is reduced by 38 percent.

The second reason can be seen in Figure 8.13. This shows that the oxygen use is

considerably lower in the copper blow than in the slag blows. It is important to note that

the graph shows oxygen use, not oxygen efficiency. For the purpose of this discussion,

oxygen use is defined as

168
8.2.2 Copper Converter Charge

oxygen use =
(I 1 ----s
,
x 100% =
(I 1
pp
2 2
x 100%
...[8.3]

L 2 i
The nitrogen partial pressure terms in equation 8.3 are to account for the change in total

moles of gas present due to reaction, the total moles of nitrogen being unchanged. There

are two main reasons for the reduction in oxygen use during the copper blow; low slag

volume and reduced tuyere submergence. The rapid heating rate of the slag during slag

blows (Figures 7.7-7.9) indicates that there is a considerable amount of slag oxidation

occurring, and this accounts for a large proportion of the total heat input to the converter.

However, slag is routinely skimmed just prior to the copper blow, to reduce copper

losses. This leaves approximately five tonnes of slag, which is barely sufficient to cover

the surface of the bath. If the slag remains as a separate phase when a spout is formed

there will be very little contact between the gas and the slag, but it is likely that the slag is

completely emulsified in the matte. Under either of these conditions there will be little

reaction of the slag with the gas, causing a reduction in the total heat generated.

Also, as copper is produced, the total volume of material in the converter is

reduced. The resulting reduction in tuyere submergence is shown in Figure 8.14. For

example, if 120 tomies of white metal is reacted to produce blister copper, there is a 45

percent reduction in volume, which results in a 0.5 m decrease in tuyere submergence.

At Flin Ron the tuyere submergence can not be increased during the copper blow due to

the requirement for almost continuous tuyere punching, which can only be carried out at

a specific converter position. Figure 8.4 shows that decreasing the tuyere submergence

reduces oxygen use, leading to the considerably lower heat generation. This point is

emphasized in Figure 8.15, which shows that there is a definite relationship between

tuyere submergence and oxygen use. However, the scatter in Figure 8.15 indicates that

169
8.2.2 Copper Converter Charge

100

90

80
w
Cl)
D
z
w
(3
70
0

60

50
0 100 200 300 400 500

TIME (mm)

Figure 8.13 Predicted variation of oxygen use with time.

other variables also contribute to determining the total oxygen use.

0.9

0.8

E 0.7
w
0
z
w
(3 0.6
Iii
m
D 0.5
Cl)
w
w
>- 0.4
D
I

0.3

0.2
0 100 200 300 400 500
TIME (mm)
Figure 8.14 Predicted variation of tuyere submergence (including spout) with time.

170
8.2.3 Sensitivity Analysis

100

H NH
90


N N

H N
Lii H
U,

z
Lii
0
70
0

60
IlN

I I
50
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

TUVERE SUBMERGENCE (m)

Figure 8.15 Predicted variation of oxygen use with tuyere submergence (including
spout), charge 586, Feb. 1994.

8.2.3 Sensitivity Analysis

In the development and operation of the model there are a number of parameters for

which accurate values or equations are not available. It is important to determine the

effects of these parameters on the model predictions. This process can identify areas

where further work is required to obtain better values, and may also aid in understanding

what is happening in the converter. Table V1ll-I gives the parameters analyzed and the

range of values used.

A shorter time step should increase the accuracy of the model predictions, but

increases the total running time. Figure 8.16 shows that reducing the time step to one

minute only has a small effect on the matte temperature and iron content. While running

time is not a serious problem for the model, the inaccuracies involved in the plant trials

171
8.2.3 Sensitivity Analysis

Parameter low base high


Time step (mm) 1 2 -

Gas flow calculation frequency (time steps) 1 5 -

Initial bath temperature (K) 1350 1400 1450


Hood temperature (K) 623 823 1023
Slag emissivity .5 .7
Liquid-phase diffusivity base/lO eq. 6.92 basexl0
Gas-phase diffusivity base-l0% eq. 6.93 base+10%
Gas-phase mass-transfer coefficient base-10% eq. 6.94 base+l0%
Matte ladle size (tonnes) 14.4 16 17.6
Water in flux (%) 2.5 5 7.5

Spout gas fraction multiplier


-J 1.1 1.5
. . .

Table Vu-I Parameters tested in the sensitivity analysis.

and other parameters are sufficiently large to make the difference essentially

insignificant. The most time consuming section of the model is the gas flow calculation.

To avoid long running times this calculation was only carried out every five time steps

(10 minutes), or immediately following an idle period. This could represent a significant

source of error for the model, particularly when the temperature is changing rapidly.

However, Figure 8.17 shows that the difference between carrying out the gas flow

calculation every time step and every five time steps is small.

172
8.2.3 Sensitivity Analysis

1600

1500
w
a:
D

a: 1400
w
0

1300

1200

0
0 100 200 300 400 500
TIME (ml,,)
Figure 8.16 Effect of model time step on the predicted variation of matte
temperature and iron content (weight percent) with time.

The initial bath temperature is not known, but can be estimated based on measured

matte temperatures at the end of a slag blow. This basis is used because the initial matte

for charge 586 was transferred from another converter, rather than from the reverberatory

furnace. Figure 8.18 shows that an increase in the initial temperature only has an effect

on the first blow temperature and iron content, with the difference at later stages of the

173
8.2.3 Sensitivity Analysis

1600

1500
w
D 1400

1300

w
I
w 1200
I

1100

1000

6
I
5

4
z
0 3

0
0 100 200 300 400 500
TIME (mm)
Figure 8.17 Effect of gas flow calculation frequency on the predicted variation of
matte temperature and iron content (weight percent) with time.

charge being negligible. The temperature difference is reduced throughout the first blow,

with the difference in the higher initial temperature case being reduced faster due to

higher heat losses. The difference in iron content is primarily due to the temperature

dependance of the magnetite activity coefficient. A lower temperature increases the

activity coefficient and so reduces the iron content. This can be seen from the variation
of matte oxygen content given in Figure 8.19.

174
8.2.3 Sensitivity Analysis

1600

1500

LU 1400
D

LU
0 1300
LU
I
LU
I
1200

1100

0.08

0.07

0.06
LU
I
0.05

0.04
z
0
0.03

0.02

0.01

0
0 100 200 300 400 500
TIME (mm)
Figure 8.18 Effect of initial bath temperature on the predicted variation of matte
temperature and iron content (weight percent) with time.

Variations in the hood temperature and slag emissivity have only a small effect on

the matte temperature, and even less on the matte iron content. The slag emissivity,

175
8.2.3 Sensitivity Analysis

1.5

0.5

0
0 100 200 300 400 500

TIME (mm)

Figure 8.19 Effect of initial bath temperature on the predicted variation of matte
oxygen content (weight percent) with time.

however, has a relatively large effect on the slag temperature (Figure 8.20). An increase

in the hood temperature has a larger effect on slag temperature than a decrease, but it is

small compared to the effect of emissivity. Neither of these parameters has a significant

effect on the slag composition.

The liquid-phase diffusivities are calculated from a theoretical equation and, as

shown in Section 6.5.3.1, may be too low. However, if the system is under gas-phase

diffusion mass-transfer control, diffusivities in the liquid should not have a significant

effect on the model predictions. Figure 8.21, however, indicates that there is an effect,

both to the matte temperature and iron content. The effect of an increase in diffusivity is

caused by an increased amount of zinc reacting. Figure 8.22 shows that with the higher

diffusivity the zinc removal from the matte is very rapid. This is because the zinc

concentration is low enough to be under liquid-phase diffusion control, so an increase in

176
8.2.3 Sensitivity Analysis

1700

1600

1500

UI
0 1400
D

1300
UI
F-
1200

1100

1000
0 100 200 300 400 500
TIME (mm)
Figure 8.20 Effect of slag emissivity on the predicted variation of slag temperature
with time.

the diffusivity allows more zinc to reach the gas/liquid interface. All of the zinc arriving

at the interface should react, because ZnO is thermodynamically more stable than FeO.

This reduces the amount of iron oxidized and, to a lesser extent, the heat generated.

The increased iron content of the matte during the slag blows caused by a decrease

in liquid-phase diffusivity indicates that the reaction of iron has also become controlled

by diffusion in the liquid. That is, there is insufficient iron reaching the gas/liquid

interface to react with all the oxygen. This should not affect the amount of oxygen

reacting, so there is an increase in the amount of copper reacting. This causes the copper

phase to be formed at an early stage of the process. This reduced iron removal and

increased copper removal combine to produce the large increase in matte iron content

seen in Figure 8.21, particularly at the end of the first blow. After this point the effect of

an increase in diffusivity is reduced, due to the lower levels of zinc in the matte. The

177
8.2.3 Sensitivity Analysis

1800

1700

1600

1500

1400
I
w
1300

1200

1100

0 100 200 300 400 500


TIME (mm)

Figure 8.21 Effect of liquid-phase diffusivity on the predicted variation of matte


temperature and iron content (weight percent) with time.

rapid increase in matte temperature seen at the end of the low diffusivity case in

Figure 8.21 corresponds to an increase in the matte magnetite content. The heat

generated to produce the magnetite combined with the small amount of matte remaining

cause the very rapid temperature increase. The effect of changing the liquid-phase

178
8.2.3 Sensitivity Analysis
diffusivity on the oxygen content of the slag (Figure 8.23) reveals another interesting

point. Reducing the diffusivity has a considerable effect on the slag oxygen content,

again indicating that the oxidation reaction in the slag is under liquid-phase diffusion

control. However, there is no other component in the slag to react, so there is a reduction

in the overall oxygen use (Figure 8.24). Increasing the liquid-phase diffusivity had no

effect on the oxygen content of the slag over the first ten minutes of the first blow.

However, after this time there is an increase in the slag oxygen content, indicating that in

the base case there is a change in the reaction controlling mechanism from gas-phase

diffusion to liquid-phase diffusion. This same pattern is repeated in the second blow, and

the decreased oxygen use is evident in Figure 8.24. This figure also shows that

increasing the liquid-phase diffusivity has a slight effect on the oxygen use, suggesting

that the gas/slag reaction is under mixed control throughout the charge.

1.5

0.5

0
0 100 200 300 400 500
TIME (mm)
Figure 8.22 Effect of liquid-phase diffusivity on the predicted variation of matte
zinc content (weight percent) with time.

179
8.2.3 Sensitivity Analysis

20

19

18

17
(3

a) 16
z
z
w
(3 15
>-
><
0
14

13

12
0 100 200 300 400 500

TIME (mm)
Figure 8.23 Effect of liquid-phase diffusivity on the predicted variation of slag
oxygen content (weight percent) with time.

100

90

80

w
C,)
D
z 70
LLI
(3
>-
x
60

50

40
0 100 200 300 400 500
TIME (mm)

Figure 8.24 Effect of liquid-phase diffusivity on the predicted variation of oxygen


use with time.

180
8.2.3 Sensitivity Analysis
The effects of changing the gas-phase diffusivity and the gas-phase mass-transfer

coefficient are essentially the same. Figure 8.25 shows that increasing the gas-phase

mass-transfer coefficient increases the matte temperature and reduces its iron content.

These results are what would be expected for a system controlled by gas-phase diffusion.

The increased mass-transfer coefficient results in more oxygen reaching the gas/liquid

interface, allowing more reaction to occur which produces more heat and removes more

iron from the matte. The magnitude of the effect on iron content increases with time

during the slag blows. In the initial stages of the first blow, the difference between the

cases is too small to show in Figure 8.25, but becomes evident 46 minutes into the blow.

At the beginning of the copper blow the case with the higher gas-phase mass-transfer

coefficient has a higher matte iron content than the other cases. This is caused by the

reaction between the gas and iron suiphide coming under liquid-phase control at a higher

iron content in the matte, which is caused by the increased amount of oxygen reaching

the reaction site. It is important to note that the model is very sensitive to the gas-phase

diffusivity, with a very high value causing extreme temperatures. This is indicative of a

process under gas phase mass transfer control. However, since the model appears to

predict the converter temperatures reasonably well, it can be argued that either the value

used is close to the actual, or the gas surface area is under or over predicted and

compensates for an error in the mass transfer coefficient.

The effect of increasing the gas-phase mass-transfer coefficient on the slag,

however, is very small. Figure 8.26 shows that the oxygen content of the slag is changed

very little by the increase, again indicating that the reaction in the slag is controlled by

liquid-phase diffusion. The fact that a reduction in the gas-phase mass-transfer

coefficient does reduce the slag oxygen content again suggests that the gas/slag reaction

181
8.2.3 Sensitivity Analysis

1600

w
Ix
D 1500

w
1400
w
I

1300

1200

1100

0 100 200 300 400 500


TIME (mm)
Figure 8.25 Effect of gas-phase mass-transfer coefficient on the predicted variation
of matte temperature and iron content (weight percent) with time.

in the base case is under mixed gas-phase and liquid-phase mass-transfer control.

182
8.2.3 Sensitivity Analysis

19

18

17

C!,
16
-J
Cl)
z
z 15
w
>-
x
0 14

13

12
0 100 200 300 400 500

TIME (mm)

Figure 8.26 Effect of gas-phase mass-transfer coefficient on the predicted variation


of slag oxygen content (weight percent) with time.

The amount of weight added in a ladle can be calculated approximately based on

the ladle volume. However, the shape of the ladle is such that a large proportion of the

weight is in the top section. This means that a small difference in the height to which it is

filled will have a large effect on the actual amount of matte added. Generally, the weight

of matte added will average out to calculated amount. Figure 8.27 shows that an increase

in the amount of matte added results in an increase in matte temperature and iron content,

although the iron content is not increased by as much as may be expected. The increase

in matte temperature is caused by the increased tuyere submergence due to the greater

total volume of matte in the converter. The increased amount of reaction occurring

because of the extra tuyere submergence is the reason for the lower than expected iron

content; more iron is being added, but more is also being removed.

183
8.2.3 Sensitivity Analysis
The flux used at Fun Hon is sand taken from a local sand pit and brought to the

smelter by truck. The water content of the flux is quite variable, and will often depend on

the weather conditions. A variation of the water content of the flux, however, has no

noticeable effect on the matte, and only a slight effect on the slag temperature and silica

content. The effect of the assumed spout gas fraction-plume gas fraction ratio is also

small. This value only affects the calculated spout height, and so represents a small

portion of the total tuyere submergence. As such, an increase in the ratio does increase

the matte temperature slightly due to the slightly increased extent of reaction. This also

reduces the iron content of the matte by a small amount. The size of these variations is

insignificant when compared to the other uncertainties in the system.

184
8.2.3 Sensitivity Analysis

1600

1500

w
1400

cc
Lii
1300

1200

1100

0
0 100 200 300 400 500

TIME (mm)
Figure 8.27 Effect of weight of matte in a ladle on the predicted variation of matte
temperature and iron content (weight percent) with time.

185
8.2.4 Converter Operation

8.2.4 Converter Operation

There are very few operating parameters which can be changed on a copper

converter. This is probably one of the reasons it has continued, relatively unchanged, for

so long: a copper converter will run by itself with very little control and only intermittent

operator intervention required. In the past, there have been suggestions made of ways to

improve operation, and these almost always relate to gas injection. Unfortunately,

changes to the gas injection can seriously affect the overall heat balance. To determine

the effect of changes on the length of a charge, the time taken to produce 72 tonnes of

blister copper will be compared. This approach is required because the actual end point

of a charge depends on the extent of the matte/slag reaction during the copper blow,

which can not be predicted by the model. The particular weight chosen is the amount

predicted to have been produced after 400 minutes of the base charge.

8.2.4.1 Gas flow rate

Perhaps the most direct method to improve converter operation is to increase the

gas flow rate. This should shorten the total charge time without altering the heat balance,

but is limited by the onset of slopping and causes increased splash from the spout. The

model predicts that increasing the gas flow rate does increase the rate of iron removal

from the bath, but Figure 8.28 also shows that it will reduce the matte temperature. This

effect is caused by a decrease in the total oxygen use. The amount of heat obtained from

the reaction per unit volume of gas is decreased, while the heat lost per unit volume of

gas is unchanged. Since the overall heat balance of the converter is essentially controlled

by these two factors, decreasing the heat input without changing the heat removed will

result in a reduction in the heat accumulated in the matte. Thus, even though there is

more oxygen reacting, the matte temperature is reduced by the increased heat lost to the

186
8.2.4 Converter Operation

gas.

1600

1500

w 1400
a:
D

1300
0

1200

1100

1000

0
0 100 200 300 400 500
TIME (mm)
Figure 8.28 Effect of gas flow rate on the predicted variation of matte temperature
and iron content (weight percent) with time.

187
8.2.4 Converter Operation
8.2.4.2 Oxygen enrichment

Oxygen enrichment of the injected gas is used in some installations to reduce

charge times and allow an increased amount of cold charge addition. The extra scrap

burning capability is available because the use of oxygen enrichment affects the heat

balance. The effects of oxygen enrichment can be analyzed in two different ways;

keeping the total moles of oxygen added constant by reducing the number of tuyeres

used, or increasing the total amount of oxygen added. The latter is what is carried out in

industry, but the former may prove interesting.

Figure 8.29 shows the results of using oxygen enrichment while keeping the total

oxygen input constant using a reduced number of tuyeres. The matte temperature is

considerably higher than without oxygen enrichment, but the iron content of the matte is

also increased. However, the oxygen content of the matte is higher, indicating that the

increased iron content is mainly caused by the reduction of the magnetite activity

coefficient in the matte at the higher temperatures. If the weight percent of iron as

magnetite is subtracted from the total iron content, (Figure 8.30) it can be seen that there

is a small improvement in iron removal with increasing oxygen enrichment, which results

in a two minute reduction in charge time at 30% oxygen in the gas. This is the result of

an increase in mass-transfer rate within the bubble, which is just sufficient to overcome

the reduction in the number of tuyeres used.

If the total oxygen input to the converter is increased the temperature rise is even

larger than in the previous case. The increased mass-transfer rate through the regular

number of tuyeres increases the iron removal, and, hence, the temperature. Figure 8.31

shows that, even with the large increase in temperature, which will increase the matte

magnetite content, the total iron in the matte is reduced. This figure also shows that the

188
8.2.4 Converter Operation
case with 30% oxygen in the gas took about 80 minutes less to complete the charge. The

improvement was only 10 minutes for oxygen enrichment to 25%, and in both cases is

due to the increased rate of mass-transfer to the gas/liquid interface. This effect is

enhanced by higher temperatures, which is the explanation for the large differences in

charge times. The higher temperatures are also the cause of the high iron content in the

matte during the copper blow, as explained in Section 8.2.2.

189
8.2.4 Converter Operation

1800

- 1700
w
1600

1500

Lii 1400

1300

1200

1100

0
0 100 200 300 400 500
TIME (mm)
Figure 8.29 Effect of gas oxygen content, with total oxygen input constant, on the
predicted variation of matte temperature and iron content (weight percent) with time.

190
8.2.4 Converter Operation

0
0 100 200 300 400 500
TIME (mm)
Figure 8.30 Effect of gas oxygen content, with total oxygen input constant, on the
predicted variation of matte iron content with iron as magnetite removed (weight percent)
with time.

To further analyze the effects of oxygen enrichment, the model can be modified to

keep the temperatures of each phase constant. This allows higher gas oxygen contents to

be studied, while removing the complications introduced by the high temperatures

normally produced. For this analysis the total oxygen input to the converter was also

kept constant, with the number of tuyeres used being reduced in proportion to the

increases in oxygen content. This technique significantly reduces the total gas/liquid

interfacial area, but the liquid-phase kinetics around a single tuyere are not altered.

Figure 8.32 shows the predicted effect of increasing the oxygen enrichment on the

matte iron content. It is evident that there is an initial increase in iron removal rate with

increasing oxygen enrichment, but at some point between oxygen contents of 35% and

42% the rate of iron removal begins to decrease, and at high oxygen enrichments the

191
8.2.4 Converter Operation

2200

2000

1800
w
ft
D
1600
ft
w
0
1400
I

1200

1000

0
0 100 200 300 400 500
TIME (mm)
Figure 8.31 Effect of gas oxygen content, with increased oxygen input, on the
predicted variation of matte temperature and iron content (weight percent) with time.

192
8.2.4 Converter Operation

matte iron content remains higher than the base case (21% oxygen). A possible

explanation for this is that the reduced interfacial area may be reducing the total amount

of reaction occurring. However, Figure 8.33 shows that, while the total oxygen use is

reduced during the slag blows at higher oxygen enrichments, the magnitude of the

reduction is not large, indicating that the overall amount of reaction occurring in the

matte has increased. This suggests that at some point during the charge the reaction of

iron is coming under liquid-phase mass-transfer control, and copper suiphide is also

reacting.

Figure 8.34 shows that there is some reaction with copper sulphide occurring during

the slag blows. The amount of copper suiphide reacting, however, is not sufficient to

account for all of the increase in iron content. The effect of the oxygen enrichment on the

weight of blister copper produced is shown in Figure 8.35. This figure shows that there is

copper production throughout the charge, even when there is no direct reaction of the gas

with copper sulphide. The prediction that there is copper produced early in the charge,

even for the base case, is unexpected, and can be explained by the sulphur potential of the

matte. Figure 8.36 shows that the sulphur potential of the matte varies between iO
7 and
10.6.8, but is not significantly affected by increasing the oxygen enrichment. It can be

seen from a potential area diagram of the system (Figure 7.13) that these values are close

to the boundary between copper and copper sulphide, which is why some copper is being

produced. This effect will be greater at higher temperatures, where the copper phase is

stable at higher sulphur potentials.

The reduction in charge time achieved with oxygen enrichment to 84% at 1400 K is

28 minutes. However, the total amount of copper produced after the full 450 minutes is

193
8.2.4 Converter Operation

increased by 20 tonnes, while the weight and copper content of the matte remaining is

significantly reduced. This is not likely to occur in the actual process due to the

matte/slag reaction, which forms the copper slag.

This analysis indicates that, in general, increasing the oxygen content of the gas will

reduce the overall charge time by a combination of mechanisms. Increasing the total

amount of oxygen supplied to the bath allows more reaction to take place, and increases

the rate of mass-transfer to the gas-liquid interface, however, the percent of the oxygen

used is reduced. At very high oxygen enrichments, or with the use of tonnage oxygen,

rates of liquid-phase mass-transport begin to become important, and some reaction with

copper sulphide is possible at lower matte grades. However, at the oxygen enrichment

levels usable under the temperature constraints of the converter, the reduction in charge

time achieved is relatively small, but could be improved if the temperature is controlled

using high grade scrap. It should be noted that, while this increased copper oxidation

would be advantageous in copper converting, it represents a limit to oxygen enrichment

in the nickel converter. In nickel converting, this form of liquid-phase mass-transport

control would cause increased cobalt and nickel oxidation, and result in higher slag

losses.

194
8.2.4 Converter Operation

10

0
0 100 200 300 400 500
TIME (mm)
Figure 8.32 Effect of gas oxygen content on the predicted variation of matte iron
content (weight percent) with time at 1400 K.

100

90

w 80
U)
D
z
w
0 70

0
60

50

40
0 100 200 300 400 500

TIME (mm)
Figure 8.33 Effect of gas oxygen content on the predicted variation of oxygen use
with time at 1400 K.

195
8.2.4 Converter Operation

100

z 80
0
I
0
w
a: 60
-J
I
0
I.
LL
0 40
I
z
w
0
a:
w
0
20

0
0 100 200 300 400 500
TIME (mm)
Figure 8.34 Variation of the predicted amount of iron and copper suiphides
reacting, as a percentage of the total reaction in the matte, with time at 1400 K and 84%
oxygen in the injected gas.

120

100

80
Lii
z
z
0
b 60
F

w
40

20

0
0 100 200 300 400 500
TIME (mm)
Figure 8.35 Effect of gas oxygen content on the predicted variation of copper
produced with time at 1400 K.

196
8.2.4 Converter Operation

-6.6

-6.7

-6.8

-6.9
-J

I.-_.-. -7
ZE
-7.1

-7.2

j_J 7.3
D
Cl)
-7.4

-7.5

-7.6

-7.7
0 100 200 300 400 500
TIME (ml,,)
Figure 8.36 Effect of gas oxygen content on the predicted variation of sulphur
potential with time at 1400 K.

8.2.4.3 Tuyere submergence and diameter

The tuyere submergence in a Peirce-Smith converter is usually limited both by the

design of the converter and by the requirement for tuyere punching. However, in a

system which is controlled by gas-phase mass-transfer, increasing the total contact area

between the gas and liquid-phases, by increasing the tuyere submergence, should

improve the overall process efficiency. Figure 8.37 shows that it does increase the

amount of reaction and, since there is no additional cooling, the matte temperature.

Increasing the tuyere submergence by 0.1 m reduces the time required to produce

72 tonnes of blister copper by 20 minutes, while a similar decrease in tuyere

submergence increases the time by 12 minutes. It should be noted that the tuyere

submergence varies throughout the charge (Figure 8.14), based on the volume of material

197
8.2.4 Converter Operation
in the bath, so the base submergence can not be given as a number.

The use of a reduced tuyere diameter, to increase the velocity of the gas leaving the

tuyere, has been tested. This removes the need to punch the tuyeres, and reduce

tuyereline refractory wear. Figure 8.38 indicates that there is another advantage to using

smaller diameter tuyeres. The increased tuyere velocity produces a more elongated

bubble, and enhances mixing in the bubble during growth. Both of these lead to more

efficient mass-transfer during bubble growth, and an overall increase in the iron removal,

such that charge time is reduced by 18 minutes.

Both reducing the tuyere diameter and increasing the tuyere submergence improve

the kinetics of converting, so combining the two offers the possibility of making a

considerable improvement to the operation. There is, however, one possible drawback.

Figure 8.39 shows that, while the rapid decrease in iron is as desired, the increase in

matte temperature is considerable. If a reliable source of high grade scrap material is

available, then this may be an asset rather than a drawback, but if the scrap is not

available the increased temperatures may seriously degrade the refractory.

198
8.2.4 Converter Operation

1700

1600

w 1500

1400
w
0
w
I 1300
LU
I

1200

1100

6
LU
I
5

z
z
0
3

0
0 100 200 300 400 500

TIME (mm)
Figure 8.37 Effect of tuyere submergence on the predicted variation of matte
temperature and iron content (weight percent) with time.

199
8.2.4 Converter Operation

1900

1800

1700

W 1600
cc
D
1500
cc
1400
w
I 1300

1200

1100

1000

0
0 100 200 300 400 500
TIME (mm)
Figure 8.38 Effect of tuyere diameter on the predicted variation of matte
temperature and iron content (weight percent) with time.

200
8.2.4 Converter Operation

1900

1800
g
w 1700
D
I 1600
Ui
0 1500
UI
I
1400

1300

1200

1100

0
0 100 200 300 400 500
TIME (mm)
Figure 8.39 Effect of a combined reduction in tuyere diameter and increase in
tuyere submergence on the predicted variation of matte temperature and iron content
(weight percent) with time.

201
8.2.4 Converter Operation
8.2.4.4 Slag skimming procedure

Slag skimming is usually carried out at the end of each blow, before the addition of

matte. This results in a low initial slag cover, which increases throughout the blow. The

effect of altering this practice to provide either an increase or decrease in the slag

thickness during blowing is shown in Figure 8.40. In the low slag cover case, after the

initial skim at 38 minutes, slag is skimmed as soon as 12.5 toimes (one ladle) is present,

while in the high slag cover case, the base amount of slag skimmed is reduced by one

half. In all cases the slag cover in the copper blow is the same.

It is evident from Figure 8.40 that a low slag cover reduces the matte temperature.

This is primarily caused by a large reduction in the slag temperature due to the absence of

a gas/slag reaction, and the low thermal mass of the slag. The absence of the gas/slag
reaction can be clearly seen from the reduction in oxygen use shown in Figure 8.41. A

small part of this reduction is also caused by the lower temperatures, which reduce the

oxygen mass-transfer rate and, hence, the extent of the gas/matte reaction. This results in

a 12 minute increase in the charge time, and is the cause of the higher amount of iron in

matte during the slag blows.

202
8.2.4 Converter Operation

.-. 1600

w
1500
I
1400
0

1300

1200

1100

1000

0
0 100 200 300 400 500
TUYERE SUBMERGENCE (m)
Figure 8.40 Effect of skimming procedure on the predicted variation of matte
temperature and iron content (weight percent) with time.

203
8.2.5 Minor Element Removal

100

90

80
C,)
D
z
w
C!3
70
0

60

50
0 100 200 300 400 500
TIME (mm)
Figure 8.41 Effect of skimming procedure on the predicted variation of oxygen use
with time.

8.2.5 Minor Element Removal

8.2.5.1 Introduction

The removal of minor elements from the matte is an important factor which must be

considered when operating variables are to be changed. The effects of changing the same

operating variables as in the previous section will be considered. For this purpose the

distribution of the minor elements between the phases will be considered, although there

are occasions when the actual predicted concentrations are required. In particular, if

changing a variable results in the formation of a greater amount of any particular phase

the percentage of the total input reporting to that phase may increase, even though the

actual concentrations within the phase are reduced.

204
8.2.5 Minor Element Removal

For the purposes of calculating distributions, the copper slag is considered as being

a mixture of the remaining matte and slag at the end of the charge. In an actual converter

the two phases will be well mixed, and there will be a considerable reaction between

them, removing the majority of the sulphur from the matte. This reaction has not been

included in the model due to insufficient data concerning the mixing and reaction. This is

particularly important for lead and, to a lesser extent, zinc, as a relatively large proportion

of these are removed in the copper slag. For this analysis, distributions will be calculated

at a constant weight of blister produced. This provides a constant basis for the

comparisons.

8.2.5.2 Minor element behaviour in the base charge

While the minor element distributions give an overall indication of the minor

element behaviour, the variation of the minor element content of each phase with time

can give a better idea of what is actually happening. The discussion of distributions will

focus on the proportions of the minor elements reporting to the blister copper and the

dust. The base distributions to these phases are given in Table VIlI-Il

Element Distribution (%)


Blister Dust
Pb 3.93 36.7
Zn 3.61 9.90
As 2.28 61.5
Sb 0.568 40.1
Bi 1.20 37.7

Table VIlI-Il Predicted distribution of minor elements to the blister copper and
dust after 400 minutes of charge 586 (72 tonnes of blister copper produced).

205
8.2.5 Minor Element Removal

Figures 8.42, 8.44, and 8.45 show the variation with time of the concentration of

minor elements in the blister copper, matte and slag respectively, while Figure 8.46

shows the variation of their partial pressures in the off gas. The concentrations of all of

the minor elements in the blister copper follow the same pattern. There is an initial rapid

increase in minor element content, caused by the large surface area to volume ratio of the

blister copper (Figure 8.43). The rate of change of minor element concentration in the

blister copper is directly proportional to this ratio, so it can be seen that at deeper depths

of blister copper the rate of change will be reduced. At higher volumes of copper the rate

of minor element transfer does not keep pace with the rate of copper production, causing

the levelling off of the curve in Figure 8.42. When the copper blow begins, the minor

element content in the blister copper begins to drop, because the rate of copper

production is considerably higher than the minor element mass-transfer rate. The sudden

increases in minor element content during the copper blow correspond to the addition of

scrap, which has a higher minor element content than the blister copper. Otherwise, the

minor element content decreases during the copper blow. However, this does not imply

that the minor elements are being transferred from the blister to the matte. In fact, the

total weight of the minor elements in the copper is increasing slightly, but not as quickly

as the total weight of copper. The magnitude of the effect of the cold charge addition is
related to the concentration of the particular minor element in the cold charge.

Concentrations of lead, zinc, and arsenic are relatively large, so the cold charge addition

has a larger effect on their concentrations. However, for antimony and bismuth, the

concentrations in the cold charge are not much higher than their concentrations in the

blister copper, so the effect is smaller.

206
8.2.5 Minor Element Removal

a. 200 200
ci
ci
Ct
w 150 .4n
I.Jv
0 0
0 0
0 0
C) C)
100 100 Ct
F-
w
I
Cl) C/)
-J -J
50 50
U C)
w z
-j N
0 0
0 100 200 300 400 500
TIME (mm)

ci
h ci 30
It
w
0
0
0
C)
Ct 20
w
I
C/)
:i 15
z
z 10
0

a:
I
z
w
o 0
z 0 100 200 300 400 500
0
O TIME (mm)
Figure 8.42 Predicted variation of minor element concentrations in blister copper
with time, a. lead and zinc, b. arsenic, antimony, and bismuth.

207
8.2.5 Minor Element Removal

2.3

2.2

2.1

1.9

1.8

1.7

1.6
0 0.1 0.2 0.3 0.4
DEPTH (m)
Figure 8.43 Variation of surface area-to-volume ratio of blister copper with depth.

The variation of the minor element content of the matte (Figure 8.44) shows that

arsenic, antimony, and bismuth follow the same basic pattern, while the lead and zinc

behave differently. Figure 8.44a shows that the zinc content in matte decreases during

blowing, with sudden increases when matte is added. The rate of the decrease of zinc

content becomes much smaller during the copper blow. The behaviour of the lead in

matte is the opposite of this. The lead content increases continuously throughout the

charge, except for the abrupt decreases caused by matte addition. The increase continues

throughout the copper blow, as lead is concentrated in the matte. The concentrations of

the other minor elements in the matte are relatively constant throughout the slag blows

(Figure 8.44b), but increase slightly during the copper blow. This concentrating effect in

the matte occurs because the rate of copper removal from the matte is considerably faster

than the rate of minor element removal.

208
8.2.5 Minor Element Removal

The behaviour of the minor elements in the slag is shown in Figure 8.45. This

shows that all of the minor elements except for zinc behave in a similar manner;

decreasing during the slag blows and remaining relatively constant during the copper

blow. The zinc behaviour is quite different. During the slag blows the zinc content tends

to increase, but is subject to rapid drops when flux is added. The increasing zinc content

is caused by the reaction of ZnS in the matte with oxygen to form ZnO. The extent of

this reaction becomes very small during the copper blow, when the high concentration of

zinc oxide in the slag causes a fairly rapid rate of mass-transfer to the matte/slag

interface. The oxide reacts with sulphur at the interface, and the zinc is returned to the

matte. This causes the reduction of the zinc content during the copper blow.

The variations of the minor element partial pressures in the gas, shown in

Figure 8.46, all follow the same basic pattern; increasing during the slag blows and

decreasing during the copper blow. The extent of the decrease at the beginning of the

copper blow is related to the difference in activity coefficients between matte and white

metal. Thus, arsenic and antimony partial pressures drop significantly, while the partial

pressures of lead, zinc, and bismuth are only slightly decreased. The increase in partial

pressure during the slag blows is caused by the increasing bath temperatures, and

likewise, the decrease during the copper blow is caused by the decreasing temperatures.

209
__

8.2.5 Minor Element Removal

4 1.4
a.
.3.5 1.2

UI
2.5 0.8

0.6
z z
0.4
UI
-J N
1 0.2

0.5 0
0 100 200 300 400 500
TIME (mm)

900
b.
800
F-
F
600
Z 500
400
300
1t 200
I
Z 100
Lu
00
Z 0 100 200 300 400 500
0
0 TIME (mm)
Figure 8.44 Predicted variation of minor element concentrations in matte with
time, a. lead and zinc, b. arsenic, antimony, and bismuth.

210
8.2.5 Minor Element Removal

2 5
a.
A
+.c: o
1.5 0

CD 4
-J 1 -J
C,) Cl)
z 3.5z
0 0
LU 0.5 z
-J 3Ri

0
0 100 200 300 400
TIME (mm)
-5OO

b.
CD 400
-J
Cl)
300
z
0
200
F
z
LU 100
0
z
0
00
0 100 200 300 400 500
TIME (mm)
Figure 8.45 Predicted variation of minor element concentrations in slag with time,
a. lead and zinc, b. arsenic, antimony, and bismuth.

211
8.2.5 Minor Element Removal

(/ -2
<
-

-10
0 100 200 300 400 500

TIME (mm)

-3

-4

-5 -f
C),
-6
z
Ui -7

-8

-9
_1

-10 I
-J

11
0
-J
-12
0 100 200 300 400 500

TIME (mm)
Figure 8.46 Predicted variation of minor element partial pressures in gas with
time, a. lead and zinc, b. arsenic, antimony, and bismuth.

212
8.2.5 Minor Element Removal

8.2.5.3 Model predictions

8.2.5.3.1 Gas flow rate

Increasing the gas flow rate should increase the total gas/liquid interfacial area, as

well as the total gas throughput. Both of these factors should improve the removal

kinetics of the minor elements to the gas, and this is predicted by the model. Figure 8.47

shows that there is an increase in the distribution of all of the minor elements to the dust.

This is particularly evident for lead, the proportion of which reporting to the dust is

increased by 33%. The increase in arsenic, antimony, and bismuth in the low gas flow

rate case is unexpected, but can be explained by the increased time required to reach the

end point and the increased tuyere submergences due to the reduced iron removal rate.

These combine to increase the overall surface area available for mass-transfer. The

distribution of minor elements to the blister is not significantly affected by the gas flow

rate.

213
8.2.5 Minor Element Removal

::-
a.

b.

61.55

::
61.45

d.

-10% BASE +10%

GAS FLOW RATE

Figure 8.47 Predicted variation of minor element distribution to the dust with gas
flow rate, a. lead b. zinc, c. arsenic, d. antimony, and e. bismuth.

8.2.5.3.2 Oxygen enrichment

Using oxygen enrichment, while keeping the total volume of oxygen constant, will

reduce the total gas flow rate, and so may be expected to reduce removal to the dust.

This effect is predicted by the model, as shown in Figure 8.48. There is also an increase

in the distribution to the slag by all materials except zinc. The reduction of zinc reporting

to the slag is caused by the reduced gas/liquid interfacial area. The increases in the other

minor elements in the slag are caused by the higher concentrations in the matte, which

214
8.2.5 Minor Element Removal

increase the driving force for mass-transfer between the liquid phases. This effect is also

seen in the increased concentration of minor elements in the blister copper. In particular,

increasing the oxygen content to 30% causes the concentration of antimony in blister

copper to increase by 27%, while arsenic and bismuth are increased by 10%

(Figure 8.49). The differences in the magnitude of these increases is related to the

activity coefficients in the matte and copper.

If the total gas flow rate is not changed while oxygen enrichment is used there are

some significant differences. Figure 8.50 shows that there is an increase in the proportion

of minor elements reporting to the dust. The primary explanation for these differences is

the bath temperature. At higher gas oxygen contents the bath temperature is dramatically

increased, which increases the rate of mass-transfer to the gas/liquid interface. The

resulting increase in vapourization (Figure 8.51) more than makes up for the slight

decrease in off-gas volume caused by the increased oxygen content.

While there is a slight increase in the proportion of minor elements reporting to the

blister copper at 24% oxygen, Figure 8.52 shows that only the antimony concentration is

increased at 30% oxygen in the gas. The explanation of this is more complicated. At
24% oxygen enrichment the increase is caused by the higher mass-transfer rates at

increased temperatures and this is also the case for the antimony at 30% oxygen.

However, for the other minor elements the composition of the cold charge becomes

important.

215
8.2.5 Minor Element Removal

a.

C.

61.2-

d.

::
40.1 -Sb

e. - Bi

37.65 -

37.55

21 25 30

GAS OXYGEN CONTENT (%)


Figure 8.48 Predicted variation of minor element distribution to the dust with gas
oxygen content with constant total oxygen input, a. lead b. zinc, c. arsenic, d. antimony,
and e. bismuth.

216
8.2.5 Minor Element Removal

a. Pb
3.96

3.92

b. Zn
3.64

3.62 -

3.6 -

2.5 As

2 2.4

2.3

d.
Sb
0.7

0.6-

1.32 Bi
e.
.

1.28 -

1.24 -

1.2 -

2530

GAS OXYGEN CONTENT (%)


Figure 8.49 Predicted variation of minor element distribution to the blister copper
with gas oxygen content with constant total oxygen input, a. lead b. zinc, c. arsenic,
d. antimony, and e. bismuth.

217
8.2.5 Minor Element Removal

E
C.

. 61.4

d. 41.2
Sb

e.

37.5

21 25 30

GAS OXYGEN CONTENT (%)

Figure 8.50 Predicted variation of minor element distribution to the dust with gas
oxygen content with increased total oxygen input, a. lead b. zinc, c. arsenic, d. antimony,
and e. bismuth.

218
8.2.5 Minor Element Removal

-3

-4
/.-
LU
II
-5
D
U) L
r /
U)
w
U: -6
0
-J

I-
-7
U:
0
-8 GAS OXYGEN CONTENT (%)
0 21 (BASE)
0
-J
-9 24
30
-10

11
0 100 200 300 400 500
TIME (mm)
Figure 8.51 Effect of gas oxygen content with increased total oxygen input, on
predicted variation of arsenic partial pressure in the off gas with time.

At an oxygen enrichment of 30%, there are some significant differences in the

minor element behaviour in the blister copper, as shown in Figure 8.53. The increased

temperatures cause the arsenic content in the copper to rise more quickly, but there is a

rapid drop close to the end of the first blow. This is because essentially all of the iron

suiphide has been removed from the matte, causing the production of copper, as well as a

considerable drop in the arsenic activity coefficient. In this case the direction of

mass-transfer is reversed and, when combined with the initial small volume of copper

present and the production of additional copper, the effect is quite large. Following the

addition of matte, the arsenic content begins to increase again, until copper begins to be

produced. After this point the curve follows the same pattern as the base curve.

However, copper is produced faster at the higher oxygen content, so when the scrap is

added it has a smaller effect on the minor element concentrations, due to the larger total

219
8.2.5 Minor Element Removal

C. 2.4 As

d.

0.56

2 25 30

GAS OXYGEN CONTENT (%)


Figure 8.52 Predicted variation of minor element distribution to the blister copper
with gas oxygen content with increased total oxygen input, a. lead b. zinc, c. arsenic,
d. antimony, and e. bismuth.

mass of copper present. Thus the arsenic concentration in the blister copper remains

lower at the higher oxygen enrichment. If the effect of the scrap is overestimated,

however, the final arsenic content of the blister copper in the oxygen enriched case may

actually be higher than the base case.

220
8.2.5 Minor Element Removal

This is actually seen for the antimony, as shown in Figure 8.54. While the same

pattern is followed for the behaviour of the antimony concentration in the base charge,

there is considerably less antimony in the scrap. In this case, the increased initial

antimony content caused by the higher temperatures is not reduced to the level of the

base charge, and the scrap additions to the base charge do not bring its antimony content

above the oxygen enriched case. This emphasizes the importance of obtaining good

expressions for transport properties.

25

20

15

10

0
0 100 200 300 400 500

TIME (mm)
Figure 8.53 Effect of gas oxygen content, with increased total oxygen input, on the
predicted variation of arsenic content in blister copper with time.

If the model is run using oxygen enrichment, but assuming that the temperature is

constant, some interesting results are obtained. Increasing the oxygen content under

these conditions reduces the distribution of minor elements to the blister copper, as

shown in Figure 8.55. The lower concentrations are primarily caused by the increased

rate of copper production early in the charge (Figure 8.56). The increased amount of

221
8.2.5 Minor Element Removal

20

15

IQ
S

1:

0
0 100 200 300 400 500
TIME (mm)
Figure 8.54 Effect of gas oxygen content, with increased total oxygen input, on the
predicted variation of antimony content in blister copper with time.

copper reduces the interfacial area to volume ratio, so while the mass-transfer rate is

essentially unchanged, the actual concentrations in the copper remain lower. The

increased weight of copper also reduces the effect of the cold charge addition as

explained above.

While increasing the oxygen enrichment to 42% causes a small decrease in the

proportions of arsenic, antimony, and bismuth reporting to the dust, Figure 8.57 shows

that increasing the oxygen content to 84% increases the proportions of all of the minor

elements reporting to the dust. This result is not expected, and requires some

explanation. The off gas volume, at 84% oxygen in the injected gas, will be reduced by

considerably more than a factor of four, so to increase the distribution to the dust, the

partial pressures of these minor elements must increase by even more. Figure 8.58 shows

that the arsenic partial pressure is increased at high oxygen contents, and remains

222
8.2.5 Minor Element Removal

relatively high, even through the copper blow. Most of the increase in arsenic partial

pressure will be directly related to the reduced off-gas volume. It is important to note,

however, that the volume of gas injected through a single tuyere is unchanged by the

oxygen enrichment. This means that, for the bubble growth and early in the bubble rise,

the surface area available for mass-transfer per tuyere is only slightly reduced.

Therefore, there is little change in the initial mass-transfer of arsenic to the gas around a

single tuyere. As the oxygen is removed from the gas by reaction there is a concentrating

effect which causes the higher partial pressure.

This effect alone will not increase the total amount of minor elements removed to

the dust, so some other mechanism must be involved. The explanation of this can be seen

in Figure 8.59. This figure shows that, for the slag blows, there is a very slight increase

in the arsenic content of the matte at 84% oxygen in the gas. During the slag blows the

arsenic content of the matte drops quite quickly, but during the copper blow the removal

rate is very low. This is because the activity coefficients of the minor elements are much

lower in white metal, reducing the driving force for mass transfer. The composition

dependence of the activity coefficients in the matte is also the reason the matte arsenic

content in the case with a high gas oxygen content remains close to the base case during

the slag blows. The difference between the two cases occurs at the end of the second slag

blow. Throughout the charge, the mole fraction of copper suiphide is lower at the high

gas oxygen content, and this effectively prolongs the slag blow, allowing the increased

removal rate to continue for a longer time. This results in the lower final arsenic content

in the matte seen in Figure 8.59. The reason the effect is not seen for lead is that its

activity coefficient is not as sensitive to matte grade, so its mass-transfer rate is not

increased significantly over the base charge. It should be noted that the formation of

223
8.2.5 Minor Element Removal

white metal is determined by the iron suiphide content of the matte. Theoretically, white

S, but it is actually a matte with a very low iron sulphide content. The
metal is Cu
2

presence of lead and zinc suiphides cause the white metal to have a Cu
S mole fraction
2
between 0.8 and 0.9.

224
8.2.5 Minor Element Removal

a.

::
2.1
z
0
-

1.9 -

I
D
m 1.7-

1.5-
0
d. Sb

::

214284

GAS OXYGEN CONTENT (%)


Figure 8.55 Predicted variation of minor element distribution to the blister copper
with gas oxygen content with constant total oxygen input and temperature, a. lead b. zinc,
c. arsenic, d. antimony, and e. bismuth.

225
8.2.5 Minor Element Removal

25

20

z Is
0

I
z
uJ 10
()
z
0
C-)

0
0 100 200 300 400 500
TIME (mm)
Figure 8.56 Effect of gas oxygen content, with constant total oxygen input, on the
predicted variation of arsenic content in blister copper with time at 1400 K.

226
8.2.5 Minor Element Removal

a. 33-Pb
32.8 -

32.6-
32.4 -

32.2 -

32 -

b. Zn
16

14 -

12-

10 -

As
65

Q63-
D
61-
C)
d. Sb

41-

39-

Bi
e. 41

39

37

21 42 84

GAS OXYGEN CONTENT (%)


Figure 8.57 Predicted variation of minor element distribution to the dust with gas
oxygen content with constant total oxygen input and temperature, a. lead b. zinc,
c. arsenic, d. antimony, and e. bismuth.

227
8.2.5 Minor Element Removal

.4 GAS OXYGEN CONTENT (%)


_________ 21
84

2 -5
(U

2
(U
-6

.7

I I I I
-8
0 100 200 300 400 500

TIME (mm)
Figure 8.58 Effect of gas oxygen content, with constant total oxygen input, on the
predicted variation of arsenic partial pressure in the off gas with time at 1400 K.

550 0.9

i
C
0
0.8
(a
500
w
a)
0
E
w
0.7

450 z
w
0
0.6
0.
-J
D
C)
400 Ui
0
0.5
C)

350 0.4
100 200 300 400 500

TIME (mm)

Figure 8.59 Effect of gas oxygen content, with constant total oxygen input, on the
predicted variation of moles of arsenic and mole fraction of copper sulphide in matte with
time at 1400 K.

228
8.2.5 Minor Element Removal

8.2.5.3.3 Tuyere submergence and diameter

The predicted distributions of the minor elements to the dust and copper for changes

in tuyere submergence are shown in Figures 8.60 and 8.61 respectively. Increasing the

tuyere submergence is predicted to increase the amount of minor elements reporting to

the dusts and reduce the amount of minor elements (with the exception of zinc) reporting

to the slag. The increased concentration in the dust is caused by an increase in the

gas/liquid interfacial area, which allows more vapourization to occur. The increased

interfacial area also increases the amount of zinc oxidized and, hence, the zinc content of

the slag.

During the copper blow the rate of removal of the minor elements from the matte is

considerably slower than the rate of copper fonnation. This causes them to be

concentrated in the matte remaining at the end of the charge, which forms the copper

slag. Figure 8.61 shows that the use of tuyere submergences less than the base value

has little effect on the lead and zinc reporting to the blister copper. However, the

distributions of all of the minor elements to the copper is reduced by an increase in the

tuyere submergence. Reducing the tuyere submergence also decreases the proportions of

arsenic, antimony, and bismuth in the blister. Figure 8.62 shows the variation of arsenic

concentration in the blister copper up to the point where 72 tonnes of copper have been

produced. Although the higher bath temperature causes a more rapid increase in the

arsenic content of the copper, the reduction in total concentration begins earlier, and the

rate of copper production is greater. The larger weight of copper present at scrap

additions reduces the effect of the added arsenic, so the arsenic content at deeper tuyere

submergences is lower throughout the copper blow. The reduced distribution to the

copper seen for arsenic, antimony, and bismuth is caused by the length of time between

229
8.2.5 Minor Element Removal

the last scrap addition and the end point. Figure 8.62 shows that the increased time

between the addition of the scrap and the end point at the reduced tuyere submergence

allows the arsenic content of the blister copper to drop below the final arsenic content of

the base case, causing the reduced distribution. This indicates that scrap should not be

added close to the end of a charge.

The effect of reducing the tuyere diameter follows the same basic pattern as

increasing the tuyere submergence. The increased temperatures and gas/liquid interfacial

area tend to increase the removal of zinc to the slag and the other minor elements to the

dust. As might be expected, combining a high tuyere submergence with a low tuyere

diameter has the same effect as each of the separate changes. The overall effect,

however, is slightly larger, with a lower final concentration of minor elements in the

blister copper, and a higher percentage reporting to the dust.

230
8.2.5 Minor Element Removal

a. Pb

b Zn
3.5 -

3-

2.5 -

c. - As
2.2-

z 2-

D
m 1.8 -

1 1.6 -

- Sb
0.56 -

0.54 -

0:::

- Bi
1.2 -

0.9

.BSE+.m

Figure 8.60 Predicted variation of minor element distribution to the blister copper
with tuyere submergence, a. lead b. zinc, c. arsenic, d. antimony, and e. bismuth.

231
8.2.5 Minor Element Removal

a.

z 2-
F =

1.8-
ci:
1.6-

d.

-.lm BASE +.1m

Figure 8.61 Predicted variation of minor element distribution to the dust with
tuyere submergence, a. lead b. zinc, c. arsenic, d. antimony, and e. bismuth.

232
8.2.5 Minor Element Removal

30

25
E

cc
w 20
0
0
0
0
cc
w 15
I
Cl)
-J

z
0 10
z
w
C,)
cc
5

0
0 100 200 300 400 500
TIME (mm)
Figure 8.62 Effect of tuyere submergence on the predicted variation of arsenic
content in blister copper with time.

8.2.5.4 Summary

In general, factors which increase the extent of oxygen use or the total gas flow

through the bath will increase the minor element removal to the dust and reduce the

distribution to the blister copper. The same factors which have the greatest effect on

oxygen use, gas/liquid interfacial area, temperature, and, to a lesser extent, gas injection

velocity, also affect the rate of minor element transport to the gas. Low levels of oxygen

enrichment are detrimental to minor element removal, due to the lower gas volume

produced, but this may be overcome by the effect of temperature. Very high levels of

oxygen enrichment have a large effect on the distributions to the dust and blister copper,

increasing the former and decreasing the latter. These effects are directly related to the

process kinetics. The timing of scrap additions is also predicted to have a large effect on

233
8.2.6 Comparison with Previous Work
the amount of minor elements reporting to the blister copper. The scrap provides a large

proportion of the total amount of minor elements in the blister copper, but its effect can

be reduced by increasing the time between the final scrap addition and the end of the

charge.

8.2.6 Comparison with Previous Work

8.2.6.1 Overall model

While the publications regarding most previous models do not contain sufficient

data to allow a reasonable comparison, a comparison may be made with the results of an

equilibrium model of the nickel 97


converter. In fact, the data obtained from the previous

study has been used in the verification of the present model, so there is a good basis for

comparison. The equilibrium model assumed that the two condensed phases were at the

same temperature, and that they were in equilibrium with the gas exiting the bath. An

overall oxygen efficiency of 95% was used.

A direct comparison of the temperatures and compositions predicted by both

models for the Copper Cliff nickel charges, shows that the predictive ability of the two

models is similar. The temperatures in the equilibrium model, however, did require some

fitting at the beginning of each charge to account for the shape of the temperature curve.

This fitting was carried out by assuming that the mush dissolved slowly and removed

heat continuously throughout the first part of the first blow. Fitting of this nature is not

required in the kinetic model, because the abrupt change in the slope of the measured

temperatures matched the crossover between matte and slag temperatures in most cases

(see Section 7.3.1). This also occurred in later blows, where the equilibrium model again
failed to match the lower initial temperatures.

234
8.2.6 Comparison with Previous Work

The compositions predicted by both models are also comparable, and generally

within the error of the measurements. The equilibrium model is closer for slags at the

ends of the charges, but the kinetic model is better for most slags at other times. The

kinetic model is also better able to predict the Bessemer matte composition. In general,

however, both models are able to predict the measured temperatures and compositions to

within the accuracy of the measurements, although minor elements were not included in

the equilibrium model, so no comparison of them may be made.

It is in the predictions regarding operating parameters that the models differ. Both

models do predict the same trends for increases in oxygen enrichment and gas flow rate,

but such variables as tuyere submergence and tuyere diameter have no effect on the

equilibrium model. These are parameters which affect the process kinetics, so have no

bearing on the system if equilibrium is assumed. Perhaps the most important difference

between the two models is that, using the equilibrium model, it was determined that there

was little which could be done to improve the chemistry of the converter.
97 However, the

kinetic model indicates that there are some variables which may be altered to improve

converter kinetics.

8.2.6.2 Minor element distribution

Previous modelling work on minor element distribution has been limited to a

number of versions of a single model. These concentrate on determining the effects of

oxygen use and initial matte


2
6 6 In all cases it has also been assumed that
grade.
3
2 13
1

the three condensed phases were in equilibrium with each other and the gas. The

distribution of minor elements between phases was calculated using distribution

coefficients and the volatilization of minor elements was also based on equilibrium

235
8.2.6 Comparison with Previous Work

considerations. It was assumed that the oxygen efficiency was 100%, with all of the

oxygen reacting with FeS until white metal was formed, at which point reaction with

S began. The models also assumed isothermal conditions.


2
Cu

The results of these models, for converting conditions, indicate that the amount of

minor element removal is almost entirely determined by the total volume of off gas

produced during smelting and/or converting. As such, increasing oxygen enrichment and

initial matte grade both reduce the amount of minor element removal to the dust. The

oxygen dependence of the distribution coefficients does suggest that there will be

increased slagging of minor elements at higher oxygen enrichments, but the extent of this

is much smaller than the reduction in volatilization. One of the papers did note that the

increase in temperature accompanying oxygen enrichment would counteract the reduced

gas volume.
213

The large differences in the basic assumptions between the previous models and the

present case make a comparison difficult. While some of the same trends are predicted in

both cases, in particular, the basic effect of total off gas volume and the effect of

increasing temperature, the effects caused by general kinetic considerations and

especially variations in matte composition due to kinetic effects, are completely

overlooked by the equilibrium models. With respect to the comparative validity of the

two models, the present case appears to able to calculate the minor element distributions

in a copper converter charge quite well, whereas Table VIll-ilI shows that the

volatilization models have some difficulties. While the present model does appear to

have difficulty predicting the bismuth distribution, this arises only during the copper

blow. Predictions up to the end of the slag blows are close to measured values. The

equilibrium model predictions of lead and arsenic are outside of the observed range, and

236
8.2.6 Comparison with Previous Work

the distributions of antimony and bismuth to the gas are very close to the top of the range.

The fact that all of the predicted distributions to the gas, with the exception of zinc,

appear to be high is an indication that the removal kinetics are controlling the

volatilization rates, rather than the thermodynamics.

Phase I Pb I Zn As I Sb Bi
Commercially Observed
16
slag 40-80 70-90 10-50 30-70 5-20
gas 20-55 10-30 20-70 5-50 70-90
Equilibrium Model
6
slag 29 82 8 33 6
gas 64 18 80 48 85
Kinetic Model
slag 59 86 36 59 45
gas 37 10 62 40 54
Table Vu-Ill Comparison of equilibrium and kinetic model predicted minor
element distributions with commercially observed ranges.

237
9.1 Introduction

9 MODEL APPLICATION

9.1 Introduction

The kinetic model developed here can be applied to determine improved operating

conditions for the Peirce-Smith converter in its present form, or to indicate a direction for

the development of a new process.

9.2 Converter Optimization

There are three primary factors to be considered when trying to optimize the

performance of the Peirce-Smith converter: process time, oxygen efficiency, and minor

element deportment. Any optimization must also be carried out within the constraints set

by bath temperature and bath motion considerations. The variables which may be used

are tuyere submergence and diameter, gas flow rate, gas oxygen content, slag skimming

practice, and number of tuyeres used. Each of these will affect the primary factors in

different ways, both beneficial and detrimental. It is important to note that if only oxygen

efficiency and the process time are considered, equilibrium operation should be optimal.

However, under equilibrium conditions the amount of minor elements reporting to the

blister copper will be high, so the kinetics become important.

Oxygen enrichment is not available in many installations, and its effect on the bath

temperature generally limits its use to situations where extra scrap is available. At low

levels, it has a detrimental effect on minor element removal, and the extent of oxygen

enrichment usable is limited by its effect on the tuyere-line refractory. All of these

factors suggest that oxygen enrichment contributes little to the process, and so should not

be considered further.

238
9.2 Converter Optimization

Of the remaining variables, a reduced tuyere diameter and an increased tuyere

submergence both increase the oxygen use and reduce the total charge time, as well as

having a beneficial effect on the minor element removal to the dust. Increasing the tuyere

submergence also allows a higher gas flow rate to be used without initiating bath

9 Reducing the tuyere diameter increases the gas exit velocity, which has been
slopping.

shown to remove the need for tuyere punching due to the formation of stable,

8 The increased gas velocity also reduces the


non-blocking accretions on the tuyeres.

115 and does not affect the bath motion, which is related to the
extent of tuyere interaction,

buoyant power per unit mass of the bath.


9 All of these factors suggest that a small tuyere

diameter in a deep bath will provide the optimum operation. However, both of these

variables also cause an increase in the bath temperature, so there is a limit to their use.

Increasing the gas flow rate was also found to reduce the total charge time and the

distribution of the minor elements to the blister copper. Unlike the changes in tuyere

submergence and diameter, an increased gas flow reduces the oxygen use and the bath

temperature. Increasing the gas flow rate may also cause bath slopping, but this may be
9 Reducing the slag cover increases the
controlled by an increased tuyere submergence.

total charge time slightly and reduces both the oxygen use and the matte temperature.

These factors indicate that present converter operations can be improved

considerably by using a smaller tuyere diameter at a deeper tuyere submergence and a

higher gas flow rate. The increased gas flow rate will aid in controlling the temperature

increase, and may be aided by altering the skimming practice. The reduced tuyere size

should remove the need for tuyere punching, and so extend refractory life. The increased

tuyere submergence will reduce the likelihood of bath slopping, even at the higher gas

flow rate. In addition to this a reduced number of tuyeres could be used to further

239
9.2 Converter Optimization

increase the gas exit velocity and reduce the amount of tuyere interaction.

To illustrate this, Figure 9.1 shows the results of an improved version of the base

charge used in Section 8. The conditions used for this prediction are given in Table IX-I.

the lengths of the slag blows are reduced, but all idle times are unchanged. Figure 9.1

shows that the total charge time is reduced by 70 minutes, while the bath temperature is

kept within the range of the base charge. Figure 9.2 shows that the oxygen use in the

improved charge is lower than the base charge during the slag blows, but is higher

during the copper blow. This suggests that there is still room for improvements, but they

may be limited by the bath temperature. In fact, it is quite possible that 100% oxygen use

is not attainable because of the temperature constraints.

Variable Value
Tuyere submergence base+0.05 m
Tuyere diameter 0.03 8 m
Gas flow rate base+10%
Slag cover low

Table IX-I Conditions used in the improved charge.

240
9.2 Converter Optimization

1700

1600
w
1500

1400

1300

1200

1100

1000

Lii 6
I
5

4
z
0 3
Cl
2

0
0 100 200 300 400 500
TIME (miri)

Figure 9.1 Effect of modifications given in Table IX-I on the predicted variation of
matte temperature and iron content (weight percent) with time.

241
9.3 New Process Development

100

90

80
C/)
D
z
w
(3
70
0

60

50
0 100 200 300 400 500

TIME (mm)

Figure 9.2 Effect of modifications given in Table IX-I on the predicted variation of
oxygen use with time.

9.3 New Process Development

The results of the model runs using a very high oxygen content in the gas at a

constant temperature indicate a possible direction for process development. In the

standard converter, equilibrium represents the limit to operating efficiency, which can be

approached by altering variables to improve the kinetics. At high levels of oxygen

enrichment the kinetics of the system can be used to give considerable improvements in

operation. Unfortunately, high levels of oxygen in the gas also cause a large increase in

the bath temperature.

242
9.3 New Process Development

The Peirce-Smith converter, like most other pyrometallurgical reactors, is designed

to keep heat in. This is the opposite of what is required to allow process improvement. If

the reactor walls are designed to remove heat, rather than retain it, then higher levels of

oxygen could be used. For example, the use of water cooled panels would increase the

heat lost through the walls considerably, and give a certain degree of control over the

bath temperature. The bath temperature can also be controlled by the addition of cold

materials such as flux, concentrate, or solidified matte.

The primary improvement available using a high gas oxygen content is the direct

production of copper from matte. This is possible if the extent of the reaction of iron

sulphide becomes controlled by liquid-phase mass-transfer. Under these circumstances

the excess oxygen at the gas/liquid interface will react with copper suiphide to form

copper. A process designed to utilize this would have the added advantages of low gas

2 content.
volumes and an off gas with a high SO

To provide an example of such a process, the model has been modified to increase

the heat flux through the walls. The initial conditions of the base charge are used subject

to the modifications given in Table DC-il. The composition of the matte added is given in

Table TX-ill. Both solidified matte and flux are added continuously throughout the run

and 15 tonnes of slag is skimmed every 10 minutes. Figure 9.3 shows the predicted

variation of matte temperature and iron content for this process. Following an initial

decrease, the bath temperature increases up to the end of the run. This could be

controlled by increasing the heat removal through the walls, to provide an essentially

isothermal system. The weight percent of iron in the matte drops initially, but begins to

level out towards the end of the run. This could also be controlled by altering the iron

content of the added matte, the rate of matte addition, the gas flow rate, or the gas oxygen

243
9.3 New Process Development

content, to provide a constant value.

Figure 9.4 shows that the oxygen use in this system is very high. Even with the

high oxygen content in the gas, the oxygen use is higher than the model preducts for any

of the regular converter charges. Finally, Figure 9.5 indicates that the rate of blister

copper production is almost one tonne per minute. This is considerably higher than the

Peirce-Smith converter, which can produce about 10 tonnes per hour.


98

Variable Value
Tuyere submergence base+0.3 m
Tuyere diameter 0.03 m
Gas flow rate/tuyere 0.2325 Nm
s
3
Gas oxygen content 95%
Number of tuyeres 20
Flux addition rate 0.5 tonnes min
1
Matte addition rate 2.75 tonnes min
1

Table TX-il Conditions used in the new process.

Element Fe Cu 0 S Pb Zn
Content (wt.%) 24.7 43.8 3.4 24.1 1.16 2.66

Table TX-ill Composition of matte added in the new process.

244
9.3 New Process Development

1650

w
1600

w
1550
w
I

1500

1450

TIME (mm)

Figure 9.3 Predicted variation of matte temperature and iron content (weight
percent) with time for the new process.

245
9.3 New Process Development

100

99

98
UI
(I,
D
z
UI
0
97
0

96

95
0 10 20 30 40 60
TIME (mm)

Figure 9.4 Predicted variation of oxygen use with time for the new process.

50

(I)
a>
C
C
.2
40
Ui
0
D

0
0 30
UI
0
0
0
0
cx 20
UI

10

0
0 10 20 30 40 50
TIME (mm)

Figure 9,3 Predicted variation of weight of blister copper produced with time for
the new process

246
10 CONCLUSIONS AND FURTHER WORK

10 CONCLUSIONS AND FURTHER WORK

The kinetic model of the Peirce-Smith converter developed here, has been shown to

be able to predict both temperatures and compositions of the condensed phases to within

the accuracy of the measurements. A number of conclusions may be drawn based on the

predictions of both the gas flow model and the overall model, but it must be recognized

that there are still considerable inaccuracies involved with some of the data used.

The apparent success of the gas flow model indicates that it is possible to

theoretically model the formation and rise of gas bubbles through a liquid bath. While

the fluid flow portion of the model is far from rigorous, it is effective, and able to predict

gas fractions within the plume and spout heights with reasonable accuracy. Perhaps more

important is that gas/liquid interfacial areas are predicted well enough to use for kinetic

modelling purposes, however the almost complete lack of information regarding

gas-phase mass-transfer in bubbles does add some doubt to the accuracies of the

predictions.

Using the gas flow model, it can be concluded that increases in the tuyere

submergence and the gas exit velocity would increase the amount of oxygen available for

reaction. While increasing the oxygen content of the gas produces more reaction, it also

leaves a larger amount of unreacted oxygen because the increase in the mass-transfer rate

is not sufficient to react all of the extra oxygen. The material properties of the liquid

phase also have an effect on the rate of bubble growth, and tuyere interaction

significantly reduces the amount of oxygen available for reaction due to the larger

bubbles produced.

The increased oxygen use with increased tuyere submergence and gas exit velocity
98 can only be explained if the converter is kinetically
reported in an operating converter

247
10 CONCLUSIONS AND FURTHER WORK

controlled. These phenomena were predicted by the gas flow model, and are directly

transferred to the overall model. Thus, to properly model the converting process, the

kinetics of the system must be considered.

The effects of many of the process variables on the major components in the

converter follow directly from the results of the gas flow model. Any variable which

increases the oxygen use will increase the rates of iron removal and temperature increase

in the matte. What cannot be predicted by the gas flow model are the relative amounts of

reaction in the matte and slag, and the behaviour of the minor elements. The separation

of the reactions with the matte and slag are an important part of the kinetic model, which

is not possible with the assumption that all of the phases are in equilibrium.

A number of important conclusions may be drawn from the overall model. The

effects of tuyere submergence and gas exit velocity (controlled by the tuyere diameter

and the gas flow rate) have already been mentioned. As well, it has been determined that

the reaction between the gas and slag is at least partially controlled by liquid-phase

mass-transfer. As such it may be possible to alter the amount of reaction occurring by

changing the physical properties of the slag. In particular, the slag viscosity will have a

direct effect on the liquid-phase diffusivities, and can be altered by changing the iron to

silica ratio. It has also been determined that small amounts of copper are produced

throughout the converting cycle.

Oxygen enrichment does not necessarily cause an increase in the amount of minor

elements reporting to the blister copper. The higher temperatures associated with oxygen

enrichment increase the mass-transfer rates within the liquid phases as well as the

equilibrium vapour pressures, so more of the minor elements are vapourized. At very

high oxygen enrichments it is predicted that white metal is formed later in the converting

248
10 CONCLUSIONS AND FURTHER WORK

cycle, with copper being produced directly from a high grade matte. This could have a

bearing on future process design, as it also increases the amount of minor element

vapourization.

There is still insufficient data regarding these, particularly with respect to

mass-transport, so the predictions regarding them must be considered to be tentative. The

inaccuracies involved in the overall model may also fairly large. Measurements made on

operating equipment are far from precise, and make any validation uncertain. To

improve on this carefully controlled cold modelling could be used, or, ideally a small

scale test converter could be used. Some tests have been carried out, but insufficient data

has been reported to allow their use in the present case.

Some other physical modelling studies could provide information which would be

useful in improving the kinetic model. These would primarily involve injection into two

and three phase systems, to obtain a better understanding of the behaviour of the

interfaces involved. Ideally these would also involve mass-transfer measurements.

Perhaps most important is the development of an expression relating to gas-phase

mass-transfer coefficient in rising bubbles. Such an expression is essential to improving

our understanding of all gas injection processes.

249
11 NOMENCLATURE
11 NOMENCLATURE

A Area (m
)
2
a Minor axis of ellipse (m)
b Maj or axis of effipse (m)
C Concentration (mol m
)
3
CD Drag coefficient
Heat capacity (J kg Kj
Disturbance propagation speed (eq. 4.22)
Diffusivity (m
2 s)
d Diameter (m)
F Force (N)
Fr Modified Froude number
f Fraction 02 in gas
2)
g Acceleration due to gravity (9.81 m
h Height (m)
h Heat-transfer coefficient (W m
2K)
1
hB Height shown in Figure. 4.7
K Constant
k Mass-transfer coefficient (m sj
k Wave number (ma) (2t

kc
1 Growth factor (eq. 4.23)
L Converter length (m)

1
L Distribution coefficient between copper and slag
1 Length (m)
M Molecular weight (g mor)
M Mass (kg)
[M] Weight percent of element M
Nu Nusselt number
n Number, Moles
ii Molar flux (mol sj
250
11 NOMENCLATURE
P Pressure (Pa)
P Equilibrium vapor pressure of pure substance (Pa)
Pr Prandtl number
p Partial pressure (Pa)
Q Volume flow rate (m
3s )
1
4 Heat flux (J s)
R Gas Constant
Re Reynolds number
r Radius (m)
s Distance from tuyere centre to bubble centre (m)
T Temperature (K)
t Time (s)
V Volume (m)
3
v Velocity (m s)
w Relative velocity of bubble with respect to the bath (m 1
)
X Mole fraction
x Thickness (m)
x Reaction constant (equation 6.34)
Greek letters

Accommodation coefficient in Langmuir-Knudsen equation


Thermal diffusivity 2
(m s)
Absorptivity of gas
Angle (Figure 5.4 or Figure 6.3)
Activity coefficient
boundary layer thickness (m)
8 Eccentricity of ellipse
8 Emissivity

ln[(
[Itan(I
Perterbation amplitude (m)

251
11 NOMENCLATURE

00 Disturbance initiation angle

Wake angle

Wavelength (m)
Dynamic viscosity (kg m
11 s
)
v Kinematic viscosity (m
2 s)
p Density (kg m
)
3
,rn Molar density (mol m
)
3

Dimensionless density

a Surface tension (N m
)
1
a Stefan-Boltzman constant, 5.67(108) W m
2K4
Gas holdup

252
11 NOMENCLATURE

Subscripts
amb Ambient
B Bath
Bu Buoyancy
b Bubble
c converter
cb converter barrel
con Consumed
D Drag
d droplet
e Evaporation
ew End wall
ext External
G Gas-phase
g Gas
gp Gas in plume
gen Generated
H Hood
i Specie, Initial, Interfacial
mt Internal
L Liquid-phase
M Matte-phase
m Mouth
mm Minimum
mt Mass-transfer
o Oriface
p Plume
R Reaction, Refractory
r Remaining, Residence
rad Radiation
S Slag-phase, Sphere
s Surface tension
sp Spout
253
11 NOMENCLATURE

sup superficial
T Temina1
t Tuyere
w Wall
x Cross-section
Superscripts
B Bulk
C Copper
I Interfacial
I,J Phase
M Matte
S Slag
W White metal
* Interfacial (actual)
+ Interfacial (equilibrium)

254
12 REFERENCES

12 REFERENCES

1. A. Dutton and S.W. Marcusen, CIMBulletin, voL 87, No. 978, pp. 156-159.

2. E.D. Peters, The Practice of Copper Smelting, McGraw-Hill, New York, 1911.

3. D.V. Browne, Trans. Canadian Mining Inst., 15, 1912, pp. 115-122.

4. A. Warczok, T. Utigard, W. Mroz, 3. Kpwalczyk, M. Warmuz and St. Musial,


Pyrometallurgy of Copper, (eds. C. Diaz, C. Landolt, A. Luraschi and
C.J. Newman), Pergamon Press, New York, 1991, pp. 637-650.

5. Y. Prevost and D. Verheist, Pyrometallurgy of Copper, (eds. C. Diaz, C. Landolt,


A. Luraschi and C.J. Newman), Pergamon Press, New York, 1991, pp. 583-598.

6. T. Shibasaki and M. Hayashi, J. Metals, 44, (9), 1991, pp. 20-26.

7. C.A. Landolt, A. Fritz, S.W. Marcuson, R.B. Cowx and 3. Miszczak, Pyrometallurgy
of Copper, (eds. C. Diaz, C. Landolt, A. Luraschi and C.J. Newman), Pergamon
Press, New York, 1991, pp. 15-29.

8. A.A. Bustos, J.K. Brimacombe, G.G. Richards, A. Vahed and A. Pelletier,


Pyrometallurgy of Copper, (eds. C. Diaz, C. Landolt and A. Luraschi), University
of Chile, 1987, pp. 347-373.

9. G.G. Richards, K.J. Legard, A.A. Bustos, J.K. Brimacombe, and D. Jorgensen, The
Reinhardt Schuhmann mt. Symp. on Innovative Technology and Reactor Design
in Extraction Metallurgy, (ed. D.R. Gaskell), Colorado Springs, CO. TMS-AIME,
WArrendale, PA, 1986, pp. 385-403.

10. J-L. Liow and N.B. Gray, Metall. Trans. B, 21B, 1990, pp 987-996.

11. R.E. Johnson, N.J. Themelis and G.A. Eltringham, Copper and Nickel Converters,
(ed. Johnson, R.E.), TMS-AIME, 1979, pp. 1-32.

12. E. Aukrust, Iron and Steel Eng., May, 1990, pp. 23-25.

13. T.P McAloon, Iron Steelmaker, January, 1991, pp. 22-23.

14. G. Nabi, Iron Steelmaker, March, 1993, pp. 7 1-77.

15. A. Kyllo, G.G. Richards and S.W. Marcuson, Metall. Trans. B, 23B, 1992,
pp. 573-582.
16. H.G. Kim and H.Y. Sohn, Pyrometallurgy of Copper, (eds. C. Diaz, C. Landolt,
A. Luraschi and C.J. Newman), Pergamon Press, New York, 1991, pp. 6 17-636.

17. S. Goto, Copper-Metallurgy Practice and Theory, ilvIM, London, 1974, pp. 23-34.
-

255
12 REFERENCES
18. S. Goto, Copper and Nickel Converters, (ed. Johnson, R.E.), TMS-AIIvIE, 1979, pp.
33-54.

19. R. Shimpo, S. Watanabe, S. Goto and 0. Ogawa, Advances in Suiphide Smelting vol.
1, TMS-AIME, 1983, pp. 295-3 16.

20. N. Kemori, T. Kimura, Y. Mori and S. Goto, Pyrometallurgy 87, 1MM, London,
1987, pp. 647-666.

21. M. Bustos and M. Sanchez, Pyrometallurgy of Copper, (eds. C. Diaz, C. Landolt and
A. Luraschi), University of Chile, 1987, pp. 473-487.

22. S.R. Brinkley, J. Chem. Phys., 15, 1947, pp. 107-1 10.

23. A. Kyllo and G.G. Richards Metall. Trans. B, 22B, 1991, pp. 153-161.

24. M. Nagamori and P.J. Mackey, Metall. Trans. B, 9B, 1978, pp. 255-265.

25. M. Nagamori and P.J. Mackey, Metall. Trans. B, 9B, 1978, pp. 567-579.

26. M. Nagamori and P.C. Chaubal, Metall. Trans. B, 13B, 1982, pp. 3 19-329.

27. M. Nagamori and P.C. Chaubal, Metall. Trans. B, 13B, 1982, pp. 33 1-338.

28. P.C. Chaubal and M. Nagamori, Metall. Trans. B, 13B, 1982, pp. 339-348.

29. P.C. Chaubal, H.Y. Sohn, D.B. George and L.K. Bailey, Metall. Trans B, 20B, 1989,
pp. 39-51.
30. K.W. Seo and H.Y. Sohn, Metall. Trans. B, 22B, 1991, pp. 791-799.

31. H.G. Kim and H.Y. Sohn, EPD Congress 91, (ed. D.R. Gaskell), TMS, 1991,
pp. 437-467.
32. B. Bjrkman and G. Eriksson, Can. Metall. Quart., 21, 1982, pp. 329-337.

33. H.E. Flynn and A.E. Morris, Mathematical Modelling ofMaterials Processing
Operations, (eds. Szekelly, J., et. al.), TIVIS-AIME, 1987, pp. 767-797.

34. J.R. Taylor, Advances in Sulphide Smelting vol. 1, TMS-AIME, 1983, pp. 2 17-229.

35. D.W. Ashman, J.W. McKelliget and J.K. Brimacombe, Can. Metal. Quart., 20, 1981,
pp. 387-395.
36, Y. Fukunaka, T. Nishihara, Z. Asaki and Y. Kondo, Todays technology for the
mining and metallurgical industries, MMIJ/IMM, London, 1989, pp. 579-588.

37. E. Adjei, M.A.Sc. Thesis, University of British Columbia, 1989.

256
12 REFERENCES
38. S. Taniguchi, A. Kikuchi, H. Matsuzaki and N. Bessho, Trans. ISIJ, Vol. 28, 1988,
pp. 262-270.
39. F. Ajersch and J.M. Toguri, Metall. Trans., vol.3, 1972, pp. 2187-2193.

40. Z. Asaki, S. Ando and Y. Kondo, Metall. Trans. B, 19B, 1988, pp. 47-52.

41. J.J. Byerley, G.L. Rempel and N. Takebe, Metall. Trans., vol.5, 1974,
pp. 2501-1506.
42. P.T. Morland, S.P. Matthew and P.C. Hayes, Metall. Trans. B, 22B,1991,
pp. 211-217.
43. R. Ohno, Metall. Trans. B, 22B, 1991, pp. 447-465.

44. A. Rome, Metall. Trans. B, 18B, 1987, pp. 213-223.

45. J.K. Brimacombe, E.S. Stratigakos and P. Tarasoff, Metall. Trans. 5, 1974,
pp. 763-771.
46. S. Inada and T. Watanabe, Trans. ISIJ, Vol. 17, 1977, pp. 21-27.

47. R.J. Fruehan and L.J. Martonik, Third International Iron and Steel Congress, ASM,
1978, pp. 229-23 8.

48. T. Stapurewicz and N.J. Themelis, Can. Metal. Quart., 26, 1987, pp. 123-128.

49. N.J. Themelis and P.R. Schmidt, Trans AIME, 239, 1967, pp. 1313-1318.

50. C.R. Nanda and G.H. Geiger,Metall. Trans., 2, 1971, pp. 1101-1106.

51. K. Mori, Trans. ISIJ, Vol. 28, 1988, pp. 246-261.

52. S.-H. Kim, R.J. Fruehan and R.I.L. Guthrie, Scaninject V. MEFOS, 1989,
41 8.
-
385
pp.
53. M. Hirasawa, K. Mori, M. Sano, A. Hatanaka, Y. Shimatami and Y. Okazaki, Trans.
ISIJ, Vol 27, 1987, pp. 277-282.

54. M. Hirasawa, K. Mori, M. Sano, Y. Shimatami and Y. Okazaki, Trans. ISIJ, Vol. 27,
1987, pp. 283-290.

55. S.-H. Kim and R.J. Fruehan, Metall. Trans. B, 18B, 1987, pp. 381-390.

56. W.A. Krivsky and R. Schuhmann, Jr., Trans. AIME, 209, 1957, pp. 98 1-988.

57. C.W. Bale and J.M. Toguri, Can. Metal. Quart., 15, 1976, pp. 305-3 18.

257
12 REFERENCES

58. J. Lumsden, Metal-Slag-Gas, Reactions and Processes, (ed. F.A. Foroulis and W.W.
Smeltzer), The electrochemical Society, Princeton, New Jersey, 1975,
pp. 155-169.

59. S.N. Sinha and M. Nagamori, Metall. Trans. B, 13B, 1982, pp. 46 1-470.

60. N. Korakas, Trans. 1MM, 72, 1962-63, pp. 35-53.

61. T. Rosenqvist and T. Hartvig, Report No. 12 from the Metallurgical Committee,
Trondheim, Norway, 1958.

62. U. Kuxmann and F.Y. Bor, Erzmetall, 18, 1965, pp. 441-450.

63. H.H. Kellogg, Physical Chemistry in Metallurgy; Proceedings of the Darken


Conference, (eds. R.M. Fisher, R.A. Oriani and E.T. Turkdogan), U.S. Steel
Research Lab., Monroeville, PA., 1976, pp. 49-68.

64. S.L. Lee and J.M. Laffain, Can. Metal. Quart., 19, 1980, pp. 183-190.

65. D.R. Fosnacht, R.P. Goel and J.M. Larrain, Metal. Trans. B, 1 1B, 1980, pp. 69-7 1.

66. R.C. Sharma and Y.A. Chang, Metal. Trans. B, lOB, 1979, pp. 103-108.

67. R.C. Sharma and Y.A. Chang, Z Metallkde., 70, H.2, 1979, pp. 104-108.

68. R.C. Sharma and Y.A. Chang, Metal. Trans. B, 1 1B, 1980, pp. 139-146.

69. R.C. Sharma and Y.A. Chang, Metal. Trans. B, 1 1B, 1980, pp. 575-583.

70. Y.-Y. Chuang, K.-C. Hsieh and Y.A. Chang, CALPHAD, 5, no.4, 1981,
pp. 277-289.
71. Y.-Y. Chuang and Y.A. Chang, Metal. Trans. B, 13B, 1982, pp. 379-385.

72. Y.-Y. Chuang and Y.A. Chang, Proceedings of the First Synposium ofMolten Salt
Chemistry and Technology, Aprill 20-22, 1988, Kyoto, Japan, pp.201-208.

73. Y.-Y. Chuang and Y.A. Chang, Second International Symposium on Metallurgical
Stags and Fluxes, TMS-AIME, 1984, pp. 73-79.

74. Y.-Y. Chuang, K.-C. Hsieh and Y.A. Chang, Metal. Trans. B, 16B, 1985,
pp. 277-285.
75. R.C. Sharma, J.-C. Lin and Y.A. Chang, Metal. Trans. B, 18B, 1987, pp. 237-244.

76. Y.A. Chang and K.-C. Hsieh, Can. Metall. Quart., 26, 1987, pp. 311-327.

77. J.M. Larrain, S.L. Lee and H.H. Kellogg, Can. Metall. Quart., 18, 1979, pp. 395-400.

258
12 REFERENCES
78. P.J. Mackey, Can. Metall. Quart., 21, 1982, pp. 221-260.

79. A. Yazawa, Y. Takeda and Y. Waseda, Can. Metall. Quart., 20, 1981, pp. 129-134.

80. Y. Takeda, S. Kanesaka and A. Yazawa, Proc. 25 Conference ofMetallurgists,


1986, CIM, Toronto, Ontario, 1986, PP. 185-202.

81. A. Yazawa and Y. Takeda, Met. Rev. MMIJ, 4(1), 1987, pp. 53-65.

82. M. Nagamori, Y. Takeda and A. Yazawa, Met. Rev. MMIJ, 6(1), 1989, pp. 6-21.

83. M. Nagamori, K. Itagaki and A. Yazawa, Met. Rev. MMIJ, 6(1), 1989, pp. 22-37.
84. A. Muan, Trans. AIME, 203, 1955, pp. 965-976.

85. E.J. Michal and R. Schuhmann, Jr., J. Metals, 4, 1952, pp. 723-728.

86. R.P. Goel, H.H. Kellogg and J.M. Larrain, Metall. Trans. B, 1 1B, 1980, pp. 107-117.

87. R.P. Goel and H.H. Kellogg, Metallurgical Slags and Fluxes, (eds. H.A. Fine and
D.R. Gaskell), TMS-AIME, Warrendale, PA., 1984, pp. 347-355.

88. F. Shenalek and I. Imris, Advances in Extractive Metallurgy and Refining, 1MM,
London, 1972, pp. 39-62.

89. M. Nagamori, Metall. Trans., 5, 1974, pp. 53 1-538.

90. J.M. Toguri and N.H. Santander, Can. Metall. Quart., 8, 1969, pp. 167-17 1.

91. R. Altmann and H.H. Kellogg, Trans. 1MM, 81, 1972, pp. C163-C175.

92. J.R. Taylor and J.H.E. Jeffes, Trans. 1MM, 84, 1975, pp. C18-C24.

93. B.J. Elliot, J.B. See and W.J. Rankin, Trans. 1MM, 87, 1978, p. C204-C211.

94. T. Oishi, M. Kamuo, K. Ono and J. Mariyama, Metall. Trans. B, 14B, 1983,
pp. 101-104.
95. C.C. Acholonu and R.G. Reddy, paper presented at the TMS-AIME meeting,
Atlanta,Ga. March, 1983.

96. H. Kurokawa, Y. Kondo, K. Baba, T. Inami, and N. Kemori, International


Symposium on Injection in Process Metallurgy, (ed. T. Lehner, P.J. Koros, and V.
Ramachandran), TMS, 1991, pp. 253-264.

97. A.K. Kyllo, M.A.Sc. Thesis, University of British Columbia, 1989.

98. D.W. Rodoiff and l.A. Rana, Copper and Nickel Converters, (ed. R.E. Johnson),
TMS-AIME, 1979, pp. 8 1-109.

259
12 REFERENCES
99. C. Delhaes, A. Hauck, and D. Neuschutz, Steel Research, 64, (1), 1993, pp. 22-27.

100. R. Lastra-Quintero, N. Rowlands, S.R. Rao and J.A. Finch, Can. Metall. Quart., 26,
1987, pp. 85-90.

101. N.J. Themelis, P. Tarassoff and J. Szekely, Trans. A.I.M.E., Vol. 245, 1969,
pp. 2425-2433.
102. T.A. Engh and J. Bertheussen, Scand. J. Metall., Vol. 4, 1975, pp. 241-249.

103. J.O. Hirschfelder, R.B. Bird and E.L. Spotz, Chem. Revs., 44, 1949, pp. 205-23 1.

104. F.A.N. Rocha and J.R.F. Guedes de Carvaiho, Chem. Eng. Res. Des., Vol. 62, 1984,
pp. 303-3 14.
105. J.F. Davidson and B.O.G. Schtiler, Trans. Instn. Chem. Engrs., Vol. 38, 1960,
pp. 145-154.

106. J.F. Davidson and B.O.G. Schler, Trans. Instn. Chem. Engrs., Vol. 38, 1960,
pp. 335-342.
107. A. Mersmann, VDI-Forschungsheft 491, Ausgabe B, Band 28, 1962.

108. A.E. Wraith, Chem. Eng. Sci., 26, 1971, pp. 1659-.

109. A.E. Wraith and T. Kakutani, Chem. Eng. Sci., 29, 1974, pp. 1-.

110. P.H. Calderbank, Trans. Instn. Chem. Engrs., 34, 1956, pp. 79-.

111. G.N. Oryall and J.K. Brimacombe, Metall. Trans. B, 7B, 1976, pp. 39 1-403.

112. T. Abel Engh and M. Nilmani, Metall. Trans. B, 19B., 1988, pp. 83-94.

113. K. Hem and J. Schmidt, Erzmetall, 43, 1990, pp. 494-497.

114. M. Nilmani, J.G. Malone and J.M. Floyd, Advances in Transport Processes in
Metallurgical Systems, (eds. Y. Sahai and G.R. StPierre), Elsevier Science
Publishers B.V., 1992, pp. 207-258. -

115. C.R. Fountain, Ph.D. Thesis, University of Melbourne, 1988.

116. V.Y. Rivkind and G.M. Ryskin, Fluid Dynamics, 1, 1976, pp. 5-12.

117. J.R. Grace and T. Wairegi, Encyclopedia of Fluid Mechanics, vol. 3,


(N.P. Cheremisinoff, ed.), Gulf Publishing, Houston, Texas, 1986, pp. 43-57.

118. J. Kitscha and G. Kocamustafaogullari, mt. J. Multiphase Flow, Vol. 15, 4, 1989,
pp. 573-588.

260
12 REFERENCES
119. Y.-F. Zhao and G.A. lions, Metall. Trans. B, 21B, 1990, pp. 997-1003.

120. R. Clift, J.R. Grace and M.E. Weber, Bubbles, Drops and Particles, Academic
Press, New York, 1978.

121. J.R. Grace, T. Wairegi and T.H. Nguyen, Trans. Instn. chem. Engrs., 54, 1976,
pp. 167-173.
122. Y.Y. Sheng and G.A. Irons, Metall. Trans. B, 23B, 1992, pp. 779-788.

123. M. Sano and K. Mori, Trans. ISIJ, 20, 1980, pp. 675-68 1.

124. F. Yamashita and H. Inoue, Kagaku Kogaku Ronbunshu, 2, 1976, p. 250.

125. R. Orr, personal communication, Thompson, Manitoba, 1988.

126. A.A. Shook, M.A.Sc. Thesis, University of British Columbia, 1986.

127. T. Wei and F. Oeters, International Symposium on Injection in Process Metallurgy,


(ed. T. Lehner, P.J. Koros, and V. Ramachandran), TMS, 1991, pp. 143-164.

128. C. Moore, UBC NLE: Zeros ofNonlinear Equations, Computing Centre, University
of British Columbia, 1984.

129. R.D. Pehlke, Unit Processes ofExtractive Metallurgy, American Elsevier Pub. Co.,
New York, 1973.

130. I. Langmuir, Jour. Amer. Chem. Soc., vol. 35, No. 2, 1913, pp. 105-127.

131. R. Harris and W.G. Davenport, Metall. Trans. B, 13B, 1982, pp. 58 1-588.

132. E. Ozberk and R.I.L. Guthrie, Metall. Trans. B, 17B, 1986, pp. 87-103.

133. A.K. Kyllo, unpublished research, Copper Cliff, 1988.

134. J. Szekely and N.J. Themelis, Rate Phenomena in Process Metallurgy,


Wiley-Interscience, New York, NY, 1971.

135. 0. Kubaschewski and C.B. Alcockm Metallurgical Thermochemistry, Pergamon


Press, 1979.

136. R.D. Harrison (ed.), Book of data, The Nuffield Foundation, Longman Group Ltd.,
London, 1978.

137. M. Sun and Y. Luo, Nonferrous Metals, 45, No. 3, 1993, pp. 53-59.

138. J.F. Elliot and M. Mounier, Can. Metall. Quart., 21, 1982, pp. 415-428.

261
12 REFERENCES
139. A.K. Biswas and W.G. Davenport, Extractive Metallurgy of Copper, 2nd Edn.,
Pergamon Press, Oxford, 1980.

140. M. Kucharski, N.M. Stubina and J.M. Toguri, Can. Metall. Quart., 28, 1989,
pp.7-il.

141. T. Ejima and T. Yamamura, Conference on the Properties ofLiquid Metals,


Proceedings of the 2nd International Conference, Tokyo, Japan, Sept. 3-8, 1972,
Taylor and Francis Ltd., London, 1973, pp. 537-541.

142. R.B. Bird, W.E. Stewart and E.N. Lightfoot, Transport Phenomena, John Wiley &
Sons, New York, NY, 1960.
143. S.W. Ip and J.M. Toguri, Metall. Trans. B, 24B, 1993, pp. 657-668.

144. M. Hino, M. Nagamori and J.M. Toguri, Metall. Trans. B, 17B, 1986, pp. 913-914.

145. S.W. Ip and J.M. Toguri, Metall. Trans. B, 23B, 1992, pp. 301-311.

146. E.J. Grimsey, Metall. Trans. B, 19B, 1988, pp. 243-247.

147. S.S. Wang, A.J. Kurtis and J.M. Toguri, Can. Metall. Quart., 12, 1973, pp. 383-390.

148. R.S. Celmer and J.M. Toguri, Proceedings 25th Annual Conference ofMetallurgists,
CIM, 1986, pp. 147-163.

149. M. Nagamori, Metall. Trans., 5, 1974, pp. 53 1-538.

150. M. Nagarnori, Metall. Trans., 5, 1974, pp. 539-548.

151. S.N. Sinha and M. Nagamori, Metall. Trans. B, 13B, 1982, pp. 461-470.

152. K. Azuma, S. Goto and N. Takebe, J. Mm. Met. Inst. Japan, 86, 1970, pp. 35-40.

153. S. Goto, 0. Ogawa, I.Asakura and T.C. Nae, J. ofMMIJ, 98, no. 1132, 1982,
pp. 37-40.
154. S. Goto, 0. Ogawa and K. Kuriyama, J. ofMMIJ, 99, no. 1148, 1983, pp. 41-44.

155. S. Goto and 0. Ogawa, J. ofMMIJ, 99, no. 1150, 1983, pp. 39-45.

156. U.V. Choudary, Y.E. Lee and Y.A. Chang, Metall. Trans. B, 8B, 1977, pp. 541-546.

157. U.V. Choudary and Y.A. Chang, Metall. Trans. B, 7B, 1986, pp. 655-660.

158. R.H. Eric and M. Timuin, Metall. Trans. B, 12B, 1981, pp. 493-500.

159. R.H. Eric and M. Timucin, J. S.A.I.M.M., 89, 1989, pp. 141-146.

262
12 REFERENCES
160. A. Rome and H. Jalkanen, Metall. Trans. B, 16B, 1985, pp. 129-141.

161. A. Rome, Metall. Trans. B, 18B, 1987, pp. 203-212.

162. J. Bode, J. Gerlach and F. Pawlek, Erzmetal, 24, 1971 pp. 480-485.

163. T. Azakami and A. Yazawa, Can. Metall. Quart., 15, 1976, pp. 111-122

164. T. Azakami, J. Mm. Met. Inst. Japan, 86, 1970, pp. 865-869.

165. A. Yazawa, T. Azakami and T. Kawashima, J. Mm. Met. Inst. Japan, 82, 1966,
pp. 5 19-524.
166. G.V. Kim and M.A. Abdeev, Zh. Neorg. Khim., 8, 1963, p. 1408.

167. M. Hayashi, T. Azakami and M. Kameda, J. Mm. Met. Inst. Japan, 90, 1974,
pp. 51-56.
168. J.D. Esdaile and J.C.H. McAdam, Proc. Australas. Inst. Mm. Metall., 239, 1971,
pp. 71-79.
169. H.H. Kellogg, Trans. AIME, 239, 1967, pp. 1439-1449.

170. M. Nagamori, P.J. Mackey and P. Tarassoff, Metall. Trans. B, 6B, 1975,
pp. 295-301.
171. S. Itoh and T. Azakami, J. Japan. Inst. Met., 48, 1984, pp. 405-413.

172. M. Hino and J.M. Toguri, Metall. Trans. B, 18B, 1987, pp. 189-194.

173. D.C. Lynch, S. Akagi and W.G. Davenport, Metall. Trans. B, 22B, 1991,
pp. 677-688.
174. A. Yazawa andY. Takeda, Met. Rev. MMIJ, 3, 1986, pp. 117-130.

175. K.H. Lau, R.H. Lamoreaux and D.L. Hildenbrand, Metall. Trans. B, 14B, 1983,
pp. 253-258.
176. M. Hino and J.M. Toguri, Metall. Trans. B, 17B, 1986, pp. 755-76 1.

177. D.C. Lynch, Metall. Trans. B, 11B, 1980, pp. 623-629.

178. D.M. Dabbs and D.C. Lynch, Metall. Trans. B, 14B, 1983, pp. 502-504.

179. T. Azakami and A. Yazawa, J. Mm. Met. Inst. Japan, 85, 1969, pp. 97-102.

180. D.G. Jones and D.H. Philipp, Trans. 1MM, 88, 1979, pp. C7-C10.
181. M. Hino, J. Mm. Met. Inst. Japan, 101, 1985, pp. 543-548.
263
12 REFERENCES
182. D.C. Lynch and K.W. Schwartze, Can. Metall. Quart., 20, 1981, pp. 269-278.

183. P. Taskinen and J. Niemal, Scand. J. Metall., 10, 1981, pp. 195-200.

184. S. Arac and G.H. Geiger, Metall. Trans. B, 12B, 1981, pp. 519-578.

185. V.S. Sibanda and E.H. Bakker, Trans. 1MM, 88, 1979, pp. C129-C130.

186. G.K. Sigworth and J.F. Effiot, Can. Metall. Quart., 13, 1974, pp. 455-461.

187. B. Predall and A. Emam, Z. Metallkd., 64, 1973, pp. 496-501.

188. J.B. See and W.J. Rankin, National Institute for Metallurgy, Report 2099, 1981.

189. E.A. Johnson, P.E. Sanker, L.L. Oden and J.B. See, United States Bureau of Mines,
RI 8655, 1982.

190. S. Goto, 0. Ogawa and I. Jimbo, Australia-Japan Extractive Metallurgy Symp.


Proc., Aus. Inst. Metall., Sydney, Australia, 1980, pp. 127-132.

191. I. Jimbo, S. Goto and 0. Ogawa, Metall. Trans. B, 15B, 1984, pp. 535-541.

192. S. C. Marschman and D.C. Lynch, Metall. Trans. B, 19B, 1988, pp. 627-641.

193. W. Jost, Diffusion in Solids, Liquids, Gases, Academic Press Inc., New York, NY,
1960.

194. A.H. Alyaser and J.K. Brimacombe, Metall. Trans. B, 25B, 1994, in press.

195. M.P Borom and J.A. Pask, J. Amer. Ceram. Soc., 51(9), 1968, pp. 490-490.

196. K. Mon and K Suzuki, Trans. ISIJ, 9, 1969, pp. 409-412.

197. F.D. Richardson, Physical Chemistry ofMelts in Metallurgy, Academic Press,


London, 1974.

198. Y. Ukyo and K.S. Goto, Metall. Trans B, 12B, 1981, pp. 449-454.

199. S.V. Sapunov and S.E. Vaisburd, Tsvetn. Metall., 26, (2), 1985, pp. 10-12.

200. A. Ghosh and H.S. Ray, Principles of Extractive Metallurgy, John Wiley & Sons,
New York,.

201. A.S. Zheleznyak, J. Applied Chem. USSR, 40, 1967, pp. 834-837.

202. E.0. Hoefele and J.K. Brimacombe, Metall. Trans. B, lOB, 1979, pp. 631-648.

264
12 REFERENCES
203. J.K. Brimacombe, A.A. Bustos, D. Jorgensen, and G.G. Richards, Physical
Chemistry ofExtractive Metallurgy, (eds. V. Kudryk and Y.K. Rao), TMS,
AIIv1E, New York, 1985, pp. 327-351.

204. A.H. Castilejos and J.K. Brimacombe, Metall. Trans. B, 18B, 1987, pp. 659-671.

205. A.H. Castillejos and J.K. Brimacombe, Metall. Trans. B, 20B, 1989, pp. 595-601.

206. V. Sahajwalla, A.H. Castillejos, and J.K. Brimacombe, Metall. Trans. B, 21B, 1990,
pp. 71-80.
207. H.K Park and J.K. Yoon, Metall. Trans. B, 21B, 1990, pp. 665-675.

208. V.L Streeter and E.B. Wylie, Fluid Mechanics, McGraw-Hill, New York, 1979.

209. D. Poggi, R. Minto, and W.G. Davenport, J. Metals, 21, (11), 1969, pp.40-45.

210. J. Schoop, W. Resch, and G. Mahn, Ironmaking Steelmaking, 5, (2), 1978,


pp. 72-79.
211. G. Turner and S. Jahanshahi, Trans. ISIJ, 27, 1987, pp. 734-739.

212. T. Kimura, S. Tsuyuguchi, Y. Ojima, Y. Mori, and Y. Ishii, J. Metals, 38, (9), 1986,
pp. 38-42.
213. P.C. Chaubal, Trans. Inst. Mi Metall., (sec. C), 98, 1989, pp. C83-84.

265
13 APPENDIX
13 APPENDIX

Model Calculation Technique

A simplified flow chart of the overall model calculation technique is given in

Figure 13.1, and more detailed flow charts of the bubble, mass balance, and heat balance

calculations are shown in Figures 13.2-13.4. Figure 13.1 shows that if the change in the

temperature of any phase is greater than a preset tolerence (10 K), then the mass balance

calculation is carried out again using the average temperatures. The heat balance is then

recalculated based on the modified compositions. With this exception, it is assumed that

all interphase heat and mass-transfer can be calculated based on the temperatures and

compositions at the end of the previous time step.

266
13 APPENDIX

INPUT OPERATING PARAMETERS

CALCULATE PHASE DEPTHS

GAS FLOW CALCULATION

MASS BALANCE CALCULATION

*
HEAT BALANCE CALCULATION
I MODIFY
I_TEMPERATURE

YES
AT>r)

NO

NO
t=t+At t>t

YES

END

Figure 13.1 Simplified flow chart of the overall model.

267
13 APPENDIX

INPUT BLOWING PARAMETERS

BUBBLE GROWTH CALCUlATION

STABILITY CALCULATION

NO
DETACHMENT?
YBS

t=t+At

4
PLUMB CHARACTERISTICS

BUBBLE RISE CALCULATION

STABILITY Cl LCULATION

OUThUT

Figure 13.2 Flow chart of the gas flow calculation.

268
13 APPENDIX

LUT

MAJOR COMPONENTS REACTION CALCULATION

MINOR COMPONENTS MASS-TRANSFER CALCULATION

LIQUID/LIQUID MASS-TRANSFER

EQUILIBRIUM CALCULATIONS

OUTPUT

Figure 13.3 Flow chart of the mass balance calculation.

269
13 APPENDIX

INPUT

MATTE HEAT BALANCE CALCULATION

SLAG HEAT BALANCE CALCULATION

BLISTER

YES

BLISTER HEAT BALANCE CALCULATION

Figure 13.4 Flow chart of the heat balance calculation.

270

Das könnte Ihnen auch gefallen