<<

The University of New South Wales

Faculty of Science

School of Materials Science and Engineering

Carbonylation of and from Reduced Oxides and

Submitted in partial fulfilment of the requirements

for the degree of Doctor of Philosophy

by

Yongli Cui

December 2015 To my parents and wife COPYRIGHT STATEMENT

'I hereby grant the University of New South Wales or its agents the right to archive and to make available my thesis or dissertation in whole or part in the University libraries in all forms of media, now or here after known, subject to the provisions of the Copyright Act 1968. I retain all proprietary rights, such as patent rights. I also retain the right to use in future works (such as articles or books) all or part of this thesis or dissertation.

I also authorise University Microfilms to use the 350 word abstract of my thesis in Dissertation Abstract International (this is applicable to doctoral theses only).

I have either used no substantial portions of copyright material in my thesis or I have obtained permission to use copyright material; where permission has not been granted I have applied/will apply for a partial restriction of the digital copy of my thesis or dissertation.'

Date

AUTHENTICITY STATEMENT

'I certify that the Library deposit digital copy is a direct equivalent of the final officially approved version of my thesis. No emendation of content has occurred and if there are any minor variations in formatting, they are the result of the conversion to digital format.'

Signed ......

Date ...... ORIGINALITY STATEMENT

'I hereby declare that this submission is my own work and to the best of my knowledge it contains no materials previously published or written by another person, or substantial proportions of material which have been accepted for the award of any other degree or diploma at UNSW or any other educational institution, except where due acknowledgement is made in the thesis. Any contribution made to the research by others, with whom I have worked at UNSW or elsewhere, is explicitly acknowledged in the thesis. I also declare that the intellectual content of this thesis is the product of my own work, except to the extent that assistance from others in the project's design and conception or in style, presentation and linguistic expression is acknowledged.'

Signed

Date

II THE UNIVERSITY OF NEW SOUTH WALES Thesis/Dissertation Sheet

Surname or Family name: CUI

First name: YONGLI other name/s:

Abbreviation for degree as given In the University calendar: PhD

School: School of Materials Science and Engineering Faculty: Science

Title: of Nickel and Iron from Reduced Oxides and Laterite Ore

Abstract 350 words maximum: The PhD project was undertaken within the ARC Discovery Grant (Project No. DP1094880) which examined a novel approach to processing of laterite based on the selective reduction of the ore and extraction of nickel by carbonylation of the selectively reduced ore. The project undertook a systematic study of carbonylation of pure nickel and iron, and carbonylation of selectively reduced laterite ore to develop a further understanding of carbonylation reactions and feasibility of extraction of nickel from laterite ore by carbonylation. The project studied the effects of operational parameters on carbonylation of nickel and iron obtained by reduction of pure oxides and in the selectively reduced Australian laterite ores, including temperature, monoxide pressure, gas flow rate, particle size, and the effect of sulphur-containing catalysts. The non-catalytic carbonylation of nickel at I 00 °C and CO pressnre 27 atm was close to completion in about 5.5 hours. Carbonylation of pure iron was slow; the extent of iron carbonylation at 100 °C and CO gauge pressure up to 55 atm was less than 5.0%. The extent of carbonylation nnder these conditions was less than 0.5%. Sulphur containing catalysts accelerated the carbonylation reaction. The time for a complete carbonylation of nickel was shortened from 5.5 hours in the non- catalytic reaction to 2 hours in the catalytic carbonylation at 100 °C and CO pressure 14 atm. The extent ofcarbonylation of nickel from the selectively reduced laterite ore at 100 °C and CO pressure 41 atm was below 50% (ore particle size 53-200 µm). The use of catalysts in the carbonylation of selectively reduced ore was inefficient. The rate of reaction increased significantly with decreasing particle size; the carbonylation of nickel in the ore with the particle size 38-53 and 75-90 µm, was close to completion after 4 hours reaction. Results of a systematic study of the carbonylation of nickel and selectively reduced latcrite ores are significant for further understanding of carbonylation reactions. Promising results were obtained for further development of technology of extraction of nickel from late rite ores by the carbonylation process.

Declaration relating to disposition of project thesis/dissertation

I hereby grant to the University of New South Wales or its agents the right to archive and to make available my thesis or dissertation in whole or in part in the University libraries in all forms of media, now or here after known, subject to the provisions of the Copyright Act 1968. I retain all property rights, such as patent rights. I also retain the right to use in future works (such as articles or books) all or part of this thesis or dissertation.

I also authorise University Microfilms to use the 350 word abstract of my thesis in Dissertation Abstracts International (this is applicable to doctoral theses only) .

Signature Witness Date

The University recognises that there may be exceptional circumstances requiring restrictions on copying or conditions on use. Requests for restriction for a period of up to 2 years must be made in writing. Requests for a longer period of restriction may be considered in exceptional circumstances and reauire the approval of the Dean of Graduate Research.

FOR OFFICE USE ONLY Date of completion of requirements for Award:

THIS SHEET IS TO BE GLUED TO THE INSIDE FRONT COVER OF THE THESIS Acknowledgements

This project was made possible by funding from the Australian Research Council ( ARC )and the Commonwealth Scientific and Industrial Research Organisation ( CSIRO ). The project code is DP1094880.

Assistance during this project has come in many forms, both in technical and moral support. I am indebted to my supervisor Professor Oleg Ostrovski and co- supervisor Dr. Guangqing Zhang for believing in me and their advice and support during the project. I would like to thank Dr. Jianqiang Zhang and Sharif Jahanshahi for their advice during project reviews.

The technical staff was invaluable in their help and support for experimental work, Mr. John Sharp for equipment maintenance; Mr. Anthony Zhang and Dr. Rahmat Kartono for lab safety support; Dr. Yu Wang for XRD analysis; Dr. Deming Zhu and Dr. Leah Koloadin for SEM/EDS analysis; Ms. Rabeya Akter and Ms. Dorothy Yu for ICP-OES analysis, Dr. George Yang for metallography and lab assistant; Mr. Danny Kim and Ms. Jane Gao for I.T. support.

The administrative staff was generous with their time and kind works, helping with more than just paperwork, Mrs. Lana Strizhevsky, Ms. Qing Xia and Mrs. Judy Lim for their help with general administrative issues.

I have greatly enjoyed the opportunity to work with all the talented and dedicated people in our group as: Dr. Yan Li, Dr. Xiaohan Wan, Dr. Xing Xing, Mr. Le Yu, Mr. Jun Yang, and Mrs. Jing Zhang. Your encouragement and advice inspired this work.

Finally, I would like to thank all my family members especially my wife Mrs. Jieqing Gan for their mental support and constant encouragement throughout the whole course of my PhD study.

IV Abstract

Australia has abundant deposits of laterite ores which role in production of nickel is increasing with rising demand for nickel and depletion of sulphide reserves.

Laterite ores cannot be efficiently upgraded prior to pyrometallurgical or hydrometallurgical processing what to significant challenges in extraction of nickel. The PhD project was undertaken within the ARC (Australian Research Council) Discovery Grant (Project No. DP1094880) which examined a novel approach to processing of laterite ores based on the selective reduction of the ore and extraction of nickel by carbonylation of the selectively reduced ore. Selective reduction of the laterite ore was studied by J. Yang in his PhD project (Yang, 2014).

The ultimate aim of this project was to establish the feasibility of extraction of nickel by carbonylation of selectively reduced laterite ore. The project undertook a systematic study of carbonylation of pure nickel and iron, and carbonylation of selectively reduced laterite ore to develop a further understanding of carbonylation reactions and extraction of nickel from laterite ore by carbonylation.

Specific objectives of this study included:

1) to study the effects of reaction parameters on carbonylation of nickel and iron, including reaction temperature (80-100 oC), (gauge) pressure (0-

56atm), gas flow rate (0.14-0.50 L·min-1), nickel mass (0.8-3.2g) and particle sizes (0.29

– 2.67 µm), and the effect of sulphur-containing catalysts (sulphur, iron sulphide and sulphide);

2) to study the non-catalytic and catalytic carbonylation of laterite ores; the impacts of reduction conditions; and effects of reaction parameters;

3) to develop further understanding of kinetics and mechanisms of carbonylation processes.

Pure metals and nickel-iron mixture were prepared by the reduction of associated oxides by hydrogen at 500oC (gas flow rate 1.0 L·min-1); the degree of reduction of oxides was over 98%. Two types of Australian laterite ores supplied by CSIRO were examined (labelled by CSIRO): BCS ore containing 1.35 wt% Ni, 10.4 wt% Fe and 0.038 wt% Co, V particle size 53-200 µm; and MIN ore with particle sizes 38-53 µm, 75-90 µm, 140-200 µm, and 355-495 µm, containing 1.68-2.37 wt% Ni, 8.90-11.8 wt% Fe and 0.135-0.144 wt% Co). Selective reduction of laterite ores was conducted by CO-CO2 gas mixture (60 vol% CO) at 750 oC.

Carbonylation experiments were conducted in a flexible U – shaped reactor (max pressure 68 atm) immersed into the oil bath. Carbonylation was studied using CO at different pressures. Carbonyls were absorbed by in two Dreschel bottles. Samples were taken from the aqua regia solution at different times and analysed by ICP-OES. The extent of carbonylation was calculated using results of the ICP-OES analysis.

The non-catalytic carbonylation of nickel at 100 oC and CO pressure 27 atm was close to completion (extent of reaction was 98%) in about 5.5 hours. Carbonylation of pure iron was slow; the extent of iron carbonylation at 100 oC and CO pressure up to 55 atm (gauge) was less than 5.0%. The extent of cobalt carbonylation under these conditions was less than 0.5%.

Sulphur containing catalysts accelerated the carbonylation reaction. The time for a complete carbonylation of nickel was shortened from 5.5 hours in the non-catalytic reaction to 2 hours in the catalytic carbonylation at 100 oC and CO pressure 14 atm.

The extent of non-catalytic carbonylation of nickel from the selectively reduced BCS laterite ore at 100 oC and CO pressure 41 atm was below 50%. The use of catalysts in the carbonylation of selectively reduced ore was inefficient. The major parameter affecting the rate of ore carbonylation was the particle size. The rate of reaction increased significantly with decreasing particle size; the carbonylation of nickel in MIN ore with the particle size 38-53 and 75-90 µm, was close to completion after 4 hours reaction.

Results of a systematic study of the carbonylation of nickel and selectively reduced laterite ores are significant for further understanding of carbonylation reactions. Promising results were obtained for further development of technology of extraction of nickel from laterite ores by the carbonylation process.

VI Contents

COPYRIGHT STATEMENT ...... I

AUTHENTICITY STATEMENT ...... I

ORIGINALITY STATEMENT ...... II

Acknowledgements ...... IV

Abstract ...... V

Contents ...... VII

List of Figures ...... XI

List of Tables ...... XIX

Chapter 1 Introduction ...... 1

Chapter 2 Literature Review ...... 6

2.1 Nickel Deposits ...... 6

2.2 Extractive Metallurgical Processes for Laterite Ores ...... 10

2.2.1 Pyrometallurgical Processes ...... 11

2.2.2 Hydrometallurgical Processes ...... 12

2.2.3 Caron Process ...... 14

2.2.4 Bio-metallurgical Process ...... 16

2.3 Carbonylation ...... 16

2.3.1 Industrial Carbonylation Process ...... 16

2.3.2 Carbonylation of Nickel ...... 21

2.3.2 Carbonylation of Iron ...... 24

2.3.3 Carbonylation of Cobalt ...... 25

2.3.4 Factors Affecting the Carbonylation ...... 26 VII 2.4 Reduction of Metal oxides ...... 35

2.5 Summary and Project’s Objectives ...... 36

Chapter 3 Experimental ...... 37

3.1 Materials...... 37

3.2 Sample preparation ...... 39

3.3 Experimental Set-up and Procedures ...... 41

3.3.1 Reduction Experiments ...... 41

3.3.2 Carbonylation Experiments ...... 42

3.4 Analytical Equipment and Methodology ...... 44

3.4.1 X-Ray Diffraction (XRD) Analysis ...... 44

3.4.2 ICP-OES/MS Analysis ...... 45

3.4.3 Scanning Electron Microscopy (SEM)/Energy Dispersive X-ray Spectrometer (EDS) ...... 46

3.4.4 X-Ray Fluorescence (XRF) Spectroscopy ...... 46

3.5 Selection of the absorption reagent ...... 47

3.6 Data Analysis ...... 48

3.6.1 Definition of extent of carbonylation and the content of catalyst 48

3.6.2 Extent of metal carbonylation ...... 48

Chapter 4 Thermodynamic Analysis of Carbonylation ...... 51

4.1 Effect of CO Pressure ...... 53

4.2 Effect of Temperature ...... 55

Chapter 5 Experimental Results ...... 57

5.1 Reduction of Iron, Nickel and Cobalt Oxides ...... 57

VIII 5.2 Non-catalytic Carbonylation of Metal ...... 61

5.2.1 Non-Catalytic Carbonylation of Pure Nickel ...... 62

5.2.2 Non-Catalytic Carbonylation of Pure Iron ...... 74

5.2.3 Non-catalytic Carbonylation of Cobalt ...... 76

5.3 Catalytic Carbonylation of Nickel ...... 78

5.3.1 Effect of the Content of Iron Sulphide and Sulphur on the Carbonylation of Nickel ...... 79

5.3.2 Effect of Temperature ...... 81

5.3.3 Effect of Carbon Monoxide Pressure ...... 83

5.3.4 Effect of the Gas Flow Rate ...... 83

5.4 Carbonylation of Selectively Reduced Laterite Ores ...... 85

5.4.1 Non Catalytic Carbonylation of Selectively Reduced Laterite Ores ...... 86

5.4.2 Catalytic Carbonylation of Selectively Reduced Laterite Ore .... 93

Chapter 6 Discussion...... 107

6.1 Impurities in Nickel and Iron Oxides ...... 107

6.2 Reduction of Nickel, Iron and Cobalt Oxides ...... 108

6.3 Non-Catalytic Carbonylation of Nickel ...... 111

6.3.1 Effect of Temperature ...... 112

6.3.2 Effect of Carbon Monoxide Pressure ...... 116

6.3.3 Effect of Gas Flow Rate ...... 118

6.3.4 Effect of Nickel Mass ...... 120

6.3.5 Effect of Particle Size ...... 121

6.4 Non-catalytic Carbonylation of Iron and Cobalt ...... 123 IX 6.5 Non-catalytic Carbonylation of Nickel from Ni-Fe Mixture...... 124

6.6 Catalytic Carbonylation of Nickel ...... 125

6.6.1 Mechanism of catalytic carbonylation of nickel ...... 125

6.6.2 Catalytic Carbonylation of Nickel ...... 130

6.6.3 Effect of Gas Flow Rate ...... 134

6.6.4 Effect of CO Pressure ...... 135

6.7 Carbonylation of Selectively Reduced Laterite Ores ...... 136

Chapter 7 Conclusions and Recommendations for Further Work ...... 140

7.1 Conclusions ...... 140

7.2 Recommendations for Further Work ...... 142

Reference ...... 143

Appendix: ...... 159

Appendix A: Calibration of gas flow rate: ...... 159

Appendix B. Calibration of reduction furnace: ...... 160

Appendix C. Calibration of oil bath: ...... 161

Appendix D. Selection of absorption solution ...... 162

Appendix E. Statistics of nickel particles reduced at different temperature ...... 163

Appendix F: Reaction Rate Control in the Non-catalytic Carbonylation ...... 163

X List of Figures

Figure 1.1 World nickel mine production ...... 2 Figure 2.1 The trend in percentage of Ni produced from laterite ores ...... 9 Figure 2.2 Flow sheets of conventional metallurgical processes for laterite ores (Norgate et al. 2011) ...... 11 Figure 2.3 Schematic flow sheet of Caron process (Taylor 2013) ...... 16 Figure 2.4 Flow sheet of the extraction of Ni and Co using carbonylation process (Terekhov 2001) ...... 18 Figure 2.5 Flow sheet of nickel production using the atmospheric pressure carbonylation at INCO, Clydach (Teng-1 2006)...... 19 Figure 2.6 Flow sheet of nickel production using medium pressure carbonyl process at INCO, Clydach (Teng 2006) ...... 20

Figure 2.7 Effect of temperature on the Ni(CO)4 pressure at the reactor outlet at 3 -1 different CO flow rates (in cm s ): (a) 0.37, (b) 2.8, (c) 12.0. Broken line: Ni(CO)4 pressure at chemical equilibrium; Pco=1 atm (De Groot et al. 1980) ...... 22 Figure 2.8 Effect of CO pressure on the rate of nickel carbonylation at 20 oC (De Groot et al. 1980) ...... 23 Figure 2.9 Effect of temperature and pressure on synthesis of (Mond et al. 1922) ...... 25 Figure 2.10 Effect of CO pressure on catalytic carbonylation of nickel at 70 oC (2 wt% sulphur added before carbonylation); a: atmospheric pressure, b: 9.9 atm, c: 19.8 atm . 27 o Figure 2.11 The effect of H2S on carbonylation of nickel at 25 C and 1 atm (Heinicke et al. 1963) ...... 28 o Figure 2.12 Influence of H2S on iron and nickel carbonylation at 25 C and CO pressure 1.1 atm (Heinicke et al. 1970)...... 29

Figure 2.13 Chemical adsorption of H2S on nickel surface (Den Besten et al. 1962) ... 30 Figure 2.14 The effect of sulphur on the carbonylation of cobalt at 200 oC and CO pressure 200 atm for 10 hours (Kipnis et al. 1973) ...... 30 Figure 2.15 The effect of sulphur on iron carbonylation at 200 oC and CO pressure 200 atm (Hieber et al. 1950) ...... 31

XI Figure 2.16 Effect of sulphur on nickel carbonylation at 70oC and CO pressure of 10 atm (Wang et al., 2009) ...... 31 Figure 2.17 Effect of sulphur on extraction of iron from Nb-Fe-B via carbonylation ... 32 Figure 2.18 Effect of on the carbonylation of Ni (110) at 130oC. (a) 1ppm oxygen in carbon monoxide, (b) 3ppm in carbon monoxide (Lascelles et al. 1983) ...... 33 Figure 2.19 Carbonylation of nickel with pre-treatment by hydrogen at 150 and 250 oC (Wang et al. 2009) ...... 34 Figure 2.20 Effect of carbon content on cobalt carbonylation at 180 oC and CO gauge pressure 70 atm (Kozyrev et al. 2005) ...... 34 Figure 3.1 The standard Gibbs free energy ∆Go and equilibrium constant (presented as log(k)) for the reaction of iron sulphide with hydrogen ...... 40 Figure 3.2 Stages in the experimental study of carbonylation of Ni and Fe ...... 41 Figure 3.3 Schematic experimental set-up for reduction of Ni and Fe oxides and laterite ore ...... 41 Figure 3.4 Schematic setup for carbonylation experiment ...... 44 Figure 3.5 Extent of nickel carbonylation measured by the weight loss and by the ICP- OES analysis of nickel in aqua regia ...... 48 Figure 4.1 The standard Gibbs free energies ∆rGo and equilibrium constants (presented as log(K)) for reactions of carbonylation of nickel and iron ...... 52 Figure 4.2 Conversions of (a) nickel and (b) iron at different temperatures and CO pressures ...... 53

Figure 4.3 Effect of CO pressure on nickel conversion to Ni(CO)4 ...... 54

Figure 4.4 Effect of pressure on iron conversion to Fe(CO)5 ...... 54

Figure 4.5 Effect of temperature on nickel conversion to Ni(CO)4 ...... 55

Figure 4.6 Effect of temperature on iron conversion to Fe(CO)5...... 55 o Figure 5.1 XRD patterns of Fe2O3 samples reduced by hydrogen at 400 C for 1, 4 and 6 hours ...... 58 Figure 5.2 XRD patterns of NiO samples reduced by hydrogen at 500 oC ...... 59

Figure 5.3 XRD patterns of the NiO-Fe2O3 mixture before and after reduction by hydrogen at 500 oC for 6 hours ...... 60 o Figure 5.4 XRD patterns of Co3O4 sample reduced by hydrogen at 500 C for 6 hours ...... 61

XII Figure 5.5 Carbonylation of the Ni-Fe mixture (50 wt% Ni) at 80 oC and CO gauge pressures 14 and 54 atm ...... 64 Figure 5.6 Nickel carbonylation at 100 oC and CO gauge pressure 27 atm at different gas flow rates ...... 64 Figure 5.7 Effect of CO flow rate on nickel carbonylation at 100 oC and CO gauge pressure 27 atm ...... 65 Figure 5.8 Effect of temperature on the carbonylation of nickel at CO gauge pressure 27 atm ...... 66 Figure 5.9 Effect of temperature on nickel carbonylation at CO gauge pressure 0, 14 and 27 atm after 6.5 hours reaction ...... 66 Figure 5.10 Effect of CO pressure (gauge) on the carbonylation of nickel at 80, 90 and 100 oC...... 68 Figure 5.11 Extents of Ni carbonylation at 80, 90 and 100 oC at different CO gauge pressures after 6.5 hours reaction ...... 68 Figure 5.12 Effect of reduction temperature on nickel particle size ...... 69 Figure 5.13 Effect of particle size on the carbonylation of nickel at 100 oC and CO gauge pressure of 27 atm...... 70 Figure 5.14 Effect of particle size (reduction temperature) on the carbonylation of nickel at 100 oC and CO gauge pressure 27 atm after 4 hours reaction ...... 71 Figure 5.15 Carbonylation of nickel of different mass at 100 oC and CO gauge pressure of 27 atm...... 72 Figure 5.16 Carbonylation of nickel from the Ni-Fe mixtures of different compositions at 100 oC and CO gauge pressure 27 atm on mixture ...... 73 Figure 5.17 Carbonylation of nickel from the Ni-Fe mixtures of different compositions at 100 oC and CO gauge pressure 27 atm after 6.0 hour reaction...... 73 Figure 5.18 Carbonylation of iron at 80-100 oC and CO gauge pressure 0, 27 and 54 atm ...... 74 Figure 5.19 Carbonylation of iron at 80-100 oC and CO gauge pressure 0, 27 and 54 atm after 6.5-hour reaction ...... 75 Figure 5.20 Effect of CO gauge pressure on carbonylation of iron at 80-100 oC...... 75 Figure 5.21 Effect of CO pressure (gauge) on the carbonylation of iron at 80-100 oC after 6.5 hours reaction...... 76

XIII Figure 5.22 Effect of temperature on carbonylation of cobalt at CO gauge pressures 41 and 54 atm ...... 77 Figure 5.23 Effect of CO pressure on carbonylation of cobalt at 80, 90 and 100 oC ...... 78 Figure 5.24 Catalytic carbonylation of nickel at 100 oC and CO pressure (gauge) 14 atm with addition of different amounts of sulphur, introduced as S and FeS ...... 80 Figure 5.25 Catalytic carbonylation of nickel at 100 oC and CO pressure (gauge) 14 atm with addition of different amounts of sulphur in the form of S and FeS after 4 hours reaction ...... 81 Figure 5.26 Effect of temperature on the catalytic carbonylation of nickel ...... 82 Figure 5.27 Effect of temperature on the catalytic carbonylation of nickel after 4 hours reaction ...... 82 Figure 5.28 Catalytic nickel carbonylation at 100 oC and CO pressures (gauge) of 14 and 27 atm ...... 83 Figure 5.29 Effect of CO gas flow rate on the catalytic carbonylation of nickel at 100 oC and CO pressure (gauge) of 14 atm...... 84 Figure 5.30 Effect of gas flow rate on the catalytic nickel carbonylation at 100 oC and CO pressure (gauge) 14 atm after 3 hours reaction ...... 85 Figure 5.31 Effects of reduction conditions on carbonylation of nickel at 100 oC, CO gauge pressure 27 atm and gas flow rate of 0.5 L·min-1...... 87 Figure 5.32 Carbonylation of nickel and iron from the selectively reduced laterite ores at 80-100 oC and CO gauge pressure 27 atm ...... 88 Figure 5.33 Carbonylation of nickel and iron from laterite ore at 100 oC and CO pressure of 27 and 41 atm (gauge) ...... 89 Figure 5.34 Carbonylation of nickel and iron from selectively reduced laterite ore by CO-CO2 gas mixture (60 vol% CO) in comparison with carbonylation of ore o o reduced by H2 at 650 and 850 C; carbonylation temperature 100 C, CO gauge pressure 27 atm ...... 91 Figure 5.35 Carbonylation of nickel and iron from the selectively reduced laterite ore at 100 oC and CO pressure (gauge) 27 atm with flow rates 0.36 and 0.50 L·min-1 ...... 92 Figure 5.36 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 80, 90 and 100 oC; CO pressure (gauge) was 27 atm ...... 94 XIV Figure 5.37 Effect of temperature on the extents of catalytic carbonylation of Ni and Fe from the selectively reduced laterite ore at CO gauge pressure 27 atm after 4 hours reaction ...... 95 Figure 5.38 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 100 oC and CO gauge pressure 14, 27 and 41 atm ...... 96 Figure 5.39 Effect of CO gauge pressure on the extents of catalytic carbonylation of Ni and Fe from the selectively reduced laterite ores at 100 oC after 3 hours reaction ...... 96 Figure 5.40 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 100 oC and CO pressure (gauge) 27 atm at the gas flow rates 0.35 and 0.50 L·min-1...... 97 Figure 5.41 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore of different size at 100 oC and CO gauge pressure 27 atm ...... 99 Figure 5.42 Effect of particle size on the catalytic carbonylation of Ni and Fe from the selectively reduced laterite ore at 100 oC and CO gauge pressure 27 atm after 3.5 hours reaction...... 99 Figure 5.43 Reduction of Ni, Co and Fe oxides from the laterite ore with size (a) < 53 μm and (b) 53-200 μm in the gas atmosphere containing 60vol% CO and 40 vol%CO2, reduction time was 60 min (Yang, 2014) ...... 101 Figure 5.44 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 650, 750 and 850 oC; carbonylation temperature 100 oC, CO gauge pressure 27 atm ...... 101 Figure 5.45 Effect of the ore reduction temperature on the catalytic carbonylation of Ni and Fe at 100 oC and CO gauge pressure 27 atm ...... 101 Figure 5.46 Effect of CO partial pressure on reduction of Ni, Co and Fe oxides from 53-200 µm laterite ore at 740 oC for 60 min; gas flow rate was 700 ml∙min-1 ...... 102 Figure 5.47 Catalytic carbonylation of nickel and iron from laterite ore selectively reduced by CO-CO2 gas with different CO concentration. Carbonylation temperature was 100 oC, CO gauge pressure 27 atm ...... 103

Figure 5.48 Carbonylation of nickel and iron from laterite ore reduced by CO-CO2 gas o (60 vol% CO) with addition of 5 vol% H2S. Carbonylation temperature was 100 C; CO gauge pressure 27 atm; 1: Introduction of H2S during last 10 min of the reduction of XV laterite ore; 2: Introduction of H2S during last 30 min of the reduction of laterite ore; 3: catalytic carbonylation (with addition of 1 vol% H2S) of the selectively reduced laterite ore without catalyst; ...... 105 Figure 6.1 XRD patterns of reduced nickel and residual sample after carbonylation at 100 oC and CO pressure 27 atm at a gas flow rate of 0.5 L∙min-1 ...... 108 Figure 6.2 SEM images of nickel reduced at 600 and 700 oC for 6 and 12 hours ...... 110 Figure 6.3 SEM image of nickel reduced from nickel oxide by hydrogen at 500 oC .. 111 Figure 6.4 Schematic diagram of pure nickel carbonylation ...... 112 Figure 6.5 XRD patterns of residuals of samples after carbonylation at different temperatures at CO gauge pressure 27 atm (6.5 hours reaction) ...... 113 Figure 6.6 Effect of temperature on the carbonylation of nickel at CO pressure 56 atm ...... 114 Figure 6.7 Plots of ln k versus 1/T at temperatures between 80 and 100 oC and CO pressure 56 atm, gas flow rate 0.5 L∙min-1 ...... 115 Figure 6.8 Rate of non-catalytic carbonylation of nickel at 80, 90 and 100 oC and CO pressure 27 atm; gas flow rate 0.5 L∙min-1 ...... 116 Figure 6.9 XRD patterns of residual samples after non-catalytic carbonylation of nickel at 100oC at CO pressures of 14 and 27 atm ...... 117 Figure 6.10 Rate of non-catalytic carbonylation of nickel at 100oC and CO gauge pressures 14 and 27 atm (gas flow rate 0.5 L·min-1) ...... 118 Figure 6.11 XRD patterns of residuals of samples after non-catalytic carbonylation of nickel at 100 oC and CO pressure 27 atm at flow rates 0.14 and 0.50 L·min-1 after 6.5 hours reaction...... 119 Figure 6.12 Rates of nickel carbonylation at 100 oC and CO pressure 27 atm at gas flow rates of 0.23, 0.35 and 0.50 L∙min-1 ...... 120 Figure 6.13 Mass of carbonylated nickel at different concentrations of in gaseous phase at a gas flow rate of 0.5 L·min-1 ...... 121 Figure 6.14 Rate of carbonylation of nickel with different particle size at 100 oC and CO pressure of 27 atm (gas flow rate 0.5 L∙min-1)...... 123 Figure 6.15 Rate of nickel carbonylation from Ni-Fe mixtures at 100oC, CO pressure 27 atm and gas flow rate 0.5 L·min-1 ...... 125

XVI Figure 6.16 XRD patterns of residual samples of Ni catalyzed by S and FeS (content of sulfur 5.0 wt%) at temperature 100 oC and CO pressure 14 atm; experiments were terminated at 10, 20 and 30 minutes in reactions with addition of sulphur, and 20 and 30 in reactions with addition of FeS. Samples marked as 00 min were not subjected to carbonylation; they were reactivated by heating in hydrogen and then cooled down to room temperature...... 126 Figure 6.17 XRD patterns of residual samples of nickel after carbonylation catalyzed by sulphur (3.5 and 7.5 wt%) in the form of FeS at 100 oC and CO pressure 14 atm after 3-4 hours reaction ...... 127 Figure 6.18 XRD patterns of residual samples of Ni-Fe mixture catalyzed by sulphur and iron sulphide FeS (content of sulfur 5.0%) at temperature 100 oC and pressure 27 atm; experiments were terminated at 10, 20 and 30 minutes...... 129 Figure 6.19 Schematic of the carbonylation of nickel catalyzed using elemental sulphur...... 129 Figure 6.20 Catalytic carbonylation of Ni at 100 oC and CO pressure 14 atm in comparison with non-catalytic carbonylation at CO pressure 14 atm (at 100 oC) and 27 atm (80-100 oC) ...... 131 Figure 6.21 Rates of catalytic nickel carbonylation at 100oC and CO pressure 14 atm (flow rate 0.5 L·min-1) with different contents of sulphur in the form of elemental sulphur and iron sulphide ...... 132 Figure 6.22 XRD patterns of residual samples after carbonylation of nickel catalysed by elemental sulphur (0.1 - 7.0 wt%) at 100 oC and CO pressure 14 atm (reaction time 3 hours) ...... 134 Figure 6.23 Rate of catalytic nickel carbonylation with addition of FeS (0.1 wt% S) at 100 oC and CO pressure 14 atm, and non-catalytic carbonylation at 100 oC and CO pressure 27 atm with different gas flow rates ...... 135 Figure 6.24 Extent and rate of catalytic nickel carbonylation with addition of FeS (0.1 wt% S) and non-catalytic carbonylation at 100 oC and CO pressure 14 and 27 atm (gas flow rate 0.5 L·min-1) ...... 136 Figure 6.25 Schematic diagram of extraction of nickel from by carbonylation ...... 137 Figure 6.26 Extent and rate of non-catalytic and catalytic carbonylation of nickel in selectively reduced laterite ore at 100oC and CO pressure 27 atm ...... 138 XVII Appendix A. Calibration of gas flow meter (a & b gas flow meter for reduction. c. gas flow meter for carbonylation) ...... 160 Appendix B. Calibration of the isothermal zone and correlation between setup temperature and isothermal temperature ...... 161 Appendix C.2 Calibration of temperature of the oil bath ...... 162 Appendix D. Absorption of nickel and iron carbonyls in nitric and aqua regia (a and c: 50% - solution, b and d: aqua regia) ...... 162 Appendix E. Statistical chart of nickel particles reduced at different temperature163 Table F1 Plots of X for gas film control, 1 – (1 – X)1/3 for the intrinsic control and 1 - 3(1-X)2/3 + 2(1-X) for the diffusion control in the first 2.5 hours carbonylation of Ni at 80 - 100 oC and CO pressure 55 atm with gas flow rate 0.5 L∙min-1 ...... 166

XVIII List of Tables

Table 1-1 2014 Ni production and total reserves by country (Statista, 2015a, Statista, 2015b) ...... 2 Table 2–1 Nickel bearing minerals and nickel contents (Boldt 1957) ...... 6 Table 2-2 Compositions of typical laterite ores, wt% (Li et al. 2007, Wang et al. 2008) 8 Table 2-3 Chemical composition and processing technology of laterite ores (Liu 2002) ...... 10 Table 2-4 Extents of metals extraction from laterite ores by HPAL process at Moa Bay Nickel Company (Peng, 2005) ...... 13 Table 3-1 Chemicals and gases ...... 37 Table 3-2 Composition of MIN and BCS garnieritic laterite ores (wt%) ...... 38 Table 3-3 Facilities and glasswares ...... 39 Table 3-4 Parameters in the reduction process ...... 42 Table 3-5 Parameters in carbonylation experiments ...... 43 Table 3-6 ICP-OES analysis parameters ...... 46

Table 5-1 Degree of reduction of Fe2O3 by hydrogen ...... 58 Table 5-2 Degree of reduction of NiO by hydrogen at 400 and 500 oC ...... 59 o Table 5-3 Degree of reduction of Fe2O3 – NiO mixture at 500 C ...... 60 Table 5-4 Parameters in a study of non-catalytic carbonylation of Ni, Fe, Co and Ni-Fe mixture ...... 62 Table 5-5 Particle size of Ni reduced at different temperatures ...... 69 Table 5-6 Experiment parameters in the catalytic carbonylation of nickel ...... 79 Table 5-7 Reaction parameters in non-catalytic carbonylation of laterite ores ...... 86 Table 5-8 Extents of carbonylation of nickel at 100 oC and CO pressure 27 atm (5.5 h reaction), produced by reduction of NiO by H2 and CO-CO2 gas mixture (60 vol% CO) ...... 87 Table 5-9 Extents of nickel and iron carbonylation from the selectively reduced laterite ores at 80-100 oC after 3.5 hours reaction ...... 88 Table 5-10 Effect of CO pressure on non-catalytic carbonylation of nickel and iron from the selectively reduced laterite ores at 100 oC ...... 90 Table 5-11 Extents of carbonylation of nickel and iron from selectively reduced laterite ore by CO-CO2 gas mixture (60 vol% CO) and from the ore reduced by H2 XIX at 650 and 850 oC (carbonylation temperature 100 oC, CO gauge pressure 27 atm) after 3.5 hours reaction ...... 91 Table 5-12 The extents of nickel and iron catalytic carbonylation from the selectively reduced laterite ore at 100 oC and CO pressure 27 atm at the gas flow rates 0.35 and 0.50 L·min-1 after 3.5 hours reaction...... 98 Table 5-13 Extents of catalytic carbonylation of nickel and iron from ores reduced by CO-CO2 gas with different contents of CO ...... 103 Table 5-14 Extents of carbonylation of Ni and Fe from laterite ore reduced with o addition of H2S. Carbonylation temperature was 100 C; CO pressure 27 atm; reaction time 2.5 hours ...... 105 Table 6-1 BET surface area and pore size of samples produced by reduction at different temperatures and residual samples after carbonylation ...... 122 Table 6-3 BET surface area and pore size of MIN ore with different particle size ..... 139 Appendix C.1 Calibration of the temperature (oC) in oil bath ...... 161

XX Chapter 1 Introduction

Nickel, a silvery-white lustrous metal, belongs to the transition metals of group IVB. It has relatively low electrical and thermal conductivities, and readily alloys with other metals to increase their strength and resistance to corrosion as well as their electrical, magnetic and heat resistant properties. Nickel is widely used in , magnets, electroplating and other areas. About 65% of nickel is consumed in the manufacturing of stainless steel and heat-resisting steels for construction, military, marine, transport and aerospace industries. The great demand of stainless steel, especially in China, promotes the continuous increase of nickel production.

Nickel in ores exists in the form of sulphides and oxides. Laterites are classified as wet or tropical laterites (from Cuba and Indonesia) and dry laterites (Australia) Chen et al. (2004). Sulphide reserves are still the major resource of nickel production; they have nickel of high grade and can be easily upgraded by beneficiation. The demand and production of nickel has been increasing, but there are a few new sulphide deposits reported last decades. The increasing price of nickel and co-extraction of cobalt make it economically to produce nickel from laterites; the proportion of nickel production from laterite ores has been increasing. Figure 1.1 shows nickel mined from both laterite and sulphide ores from 2000 and the predicted production to 2030 by International Nickel Study Group (Miner 2015).

Australia is among the major suppliers and producers of nickel ore in the world. In 2014, Australia was the fifth largest nickel producer in the world (after Philippines, Russia, Indonesia and ). The economic reserve of nickel in Australia is about 19 million tonnes accounting for 35 per cent of the nickel resources in the world. Table 1-1 lists some statistical data in the year of 2014 (Statista-a 2015, Statista-b 2015).

Compared with nickel sulphide deposits that are found hundreds of metres under the ground, laterites are usually located nearer to or at the surface with abundant reserve, which renders it available to open mining. The major challenges in processing of laterite

1 ores include ores’ low and variable grades; besides, laterite ores cannot be concentrated by beneficiation.

Figure 1.1 World nickel mine production

Table 1-1 2014 Ni production and total reserves by country (Statista, 2015a, Statista, 2015b)

Country Production Country Reserves

Metric tons Million Metric tons

Philippines 440000 Australia 19

Russia 260,000 New Caledonia 12

Indonesia 240000 Brazil 9.1

Canada 233000 Russia 7.9

Australia 220000 Cuba 5.5

New Caledonia 165000 Indonesia 4.5

Brazil 126000 South Africa 3.7

China 100000 Philippines 3.1

Colombia 75000 China 3

2 Cuba 66000 Canada 2.9

South Aferica 54700 Madagascar 1.6

Madagascar 37800 Colombia 1.1

Dominican Republic 15800 Dominican Republic 0.93

United States 0.16

Other Countries 6.5

The conventional processes applied in extracting metals (nickel, iron and cobalt) from laterites are pyrometallurgical and hydrometallurgical routes. In the pyrometallurgical processing, ore is treated by drying, calcining and reducing at temperatures above 900 oC, followed by the nickel and iron over 1300 oC. This process has a high fuel consumption; it is currently used in extracting nickel from saprolite ores rich in nickel. Pyrometallurgical processing of the low grade limonitic ore with nickel content less than 1.8% is not economical.

In hydrometallurgical routes, nickel, iron and cobalt are extracted from laterite ores at temperatures over 200 oC and high pressures above 100 atm; strong concentrated sulphuric acid is used to solve metals from ores lattice to enhance the reaction rate and degree of extraction, then nickel, iron and cobalt are recovered by precipitation. The residual solution must be neutralised before discharging into the environment. Nevertheless, acid and alkaline solutions which are used in the process cause the environmental concern. The operation of hydrometallurgical process is also influenced by iron and magnesium, which contents are much higher than contents of nickel and cobalt.

Laterite ores cannot be efficiently upgraded prior to pyrometallurgical or hydrometallurgical processing what leads to significant challenges in extraction of nickel and cobalt. Heating the ore containing a significant amount of impurities to high temperatures in pyrometallurgical route causes high fuel consumption and emission of to the environment. High contents of iron and magnesium in the ores result in high consumption of acid applied in the hydrometallurgical route. Therefore application of conventional processes to extraction of nickel from laterites results in serious environment problems and high cost of nickel production. Therefore, new 3 technologies which to the reduction of the production cost and decrease environmental impact have been under development.

Promising results in extraction of nickel and cobalt from laterite ores were obtained using bio-technology and acid by (Coto et al. 2008, McDonald et al. 2008). However, development of these technologies has not reached the industrial scale. The rate of bioleaching is very low, extents of nickel and iron extraction are less than 50%. The hydrochloric acid is highly corrosive to facilities; the use of the lower acid strength decreases efficiency of nickel and cobalt extraction from laterites (Peng 2005, Zhou 2005).

A major nickel process is . Carbonyl process (based on the Mond process (Herrmann 1990)) is also broadly used for nickel refining. Nickel and iron form carbonyls by reactions (1-1) and (1-2), which are separated by distillation.

Ni + 4CO = Ni(CO)4 (1-1)

Fe + 5CO = Fe(CO)5 (1-2)

Then nickel and iron are recovered by thermal decomposition of carbonyls by reverse reactions of (1-1) and (1-2), and carbon monoxide is recycled.

Carbonylation technologies have been used in refining nickel and iron over 100 years; the carbonylation of nickel had been intensively studied in the first decades of 20th century. However, a study of carbonylation reactions was limited only in application to pure metals. No detailed investigation of kinetics and mechanisms of carbonylation of nickel and iron at elevated temperatures and medium pressures have been reported, especially in the carbonylation of reduced laterites.

Compared with pure nickel, the properties and chemical compositions of reduced laterites are more complex; their carbonylation needs a detailed investigation. This project will study a novel approach to the processing of laterite ores based on the selective reduction and carbonylation of reduced ore to extract nickel and iron. Reduction of laterite ore was examined in the PhD project “Selective Reduction of Laterite Ore” (Yang 2014).

4 In the project, prior to carbonylation, laterite ores are selectively reduced by CO-CO2 mixture; in selective reduction, most of nickel and only part of iron are reduced. Nickel and iron in the selectively reduced ore are extracted by carbonylation.

The ultimate aim of this project is to assess the feasibility of extraction of nickel and iron by carbonylation of selectively reduced laterite ore. The project undertakes a systematic study of carbonylation of pure nickel and iron, and carbonylation of selectively reduced laterite ore to develop a better understanding of carbonylation reactions and extraction of nickel from laterite ore by carbonylation. Specific objectives of this study are:

1) to study the effects of reaction parameters on carbonylation of nickel and iron, including reaction temperature, CO pressure, gas flow rate, particle size, sample mass, and the effect of sulphur-containing catalysts;

2) to study the non-catalytic and catalytic carbonylation of laterite ores; the impacts of reduction conditions; and effects of reaction parameters;

3) to develop further understanding of kinetics and mechanisms of carbonylation processes.

5 Chapter 2 Literature Review

Nickel is an important metal and alloying component applied in stainless steel, alloys, plating and numerous other applications in industry, among them stainless steel accounts for about two thirds of the annual nickel consumption (Zhao et al. 2012). Driven by the increasing production of stainless steel and superalloys, world nickel demand increased from 1.10 Mt in 2001 to 1.57 Mt in 2011, representing an annual growth rate of 4.2% on average (INSG-2 2013). The world primary nickel production reached 1.76 Mt in 2012, 1.96 Mt in 2013 and increased to 1.99 Mt in 2014 (INSG-1 2013, INSG-3 2015).

2.1 Nickel Deposits

Nickel is commonly present in two principal ore types - sulphide and laterite ores with a reserve of about 160 Mt, of which near 70% are contained in laterite ores (Valix et al. 2001, Zhou 2005, Norgate et al. 2011).

Table 2-1 lists the principal nickel-bearing minerals and their chemical compositions, in which pentlandite, (Ni, Fe)9S8 , accounts for about three-quarters of nickel mined in the past in the world (Boldt 1957).

Table 2–1 Nickel bearing minerals and nickel contents (Boldt 1957)

Formula Nickel Content, wt%

Sulfides

Pentlandite (Ni, Fe)9S8 34.2

Millerite NiS 64.7

Heazlewoodite Ni3S2 73.3

6 Linnaeite series (Fe, Co, Ni)3S4 variable

Polydymite Ni3S4 57.96

Violarite Ni2FeS4 38.9

Siegenite (Co, Ni)3S4 28.9

Gersdorffite NiAsS 35.4

Silicate and Oxide

Garnierite (Ni, Mg)6Si4O10(OH)8 Variable up to 47

Nickel-ferrous limonite (Fe, Ni)O(OH)·nH2O Variable

Sulphide deposits, with the primary ore mineral of pentlandite, are formed as the results of segregation and concentration of droplets of liquid sulphide from mafic or ultramafic magma (Naldrett 1999). This type ore can be beneficiated by flotation, gravity and magnetic separation to produce concentrates before extracting nickel from ores. This operation significantly reduces the capital and operation cost (Chen et al. 2004, Bo 2009). The sulphide deposits are mainly distributed in Canada, Russia, Australia, China, South Africa, Zimbabwe and Botswana (Cao 2005, Zhou 2005, Li et al. 2009, Liang et al. 2009).

Although over 50% of nickel is produced from sulphide minerals in the past, the major nickel bearing minerals are laterites; they are formed by chemical weathering of precursor ultramafic rocks and supergene enrichment of weathering products mostly under tropical climatic conditions (Oliveira et al. 2001, Chen et al. 2004, Norgate et al. 2011, Berger et al. 2013). Laterites can vary significantly in mineralogy according to location, climate and depth, for example, the clay-rich nickel laterites in Western Australia differ from Moa Bay (Cuba) laterites in mineralogy and have comparatively high silica and low chromium content. Nickel-bearing laterite deposits also contain 7 valuable cobalt minerals (Andersen et al. 2009). The main host minerals for nickel and cobalt can be either limonites (SiO2 ≈ 6%; MgO 3%; Ni 1.4%; Co 0.15% and Fe >40%) or saprolites (SiO2 ≈ 38%; MgO 25%; Ni 2.4%; Co 0.05% and Fe <15%) depending on the content of iron and magnesium (Whittington et al. 2000, Berger et al. 2013, Thubakgale et al. 2013). Large quantities of laterites are found in Australia, Cuba, Indonesia, New Caledonia and the Philippines (Moskalyk et al. 2002, USGS 2009). The typical laterite ores compositions are listed in Table 2-2 (Li et al. 2007, Wang et al. 2008).

Table 2-2 Compositions of typical laterite ores, wt% (Li et al. 2007, Wang et al. 2008)

Ni Co Fe MgO SiO2 CaO Al2O3 Cr

New Caledonia Ore 2.43 0.04 9.3 28.8 42.2

Indonesia Ore 2.6 0.1 14.47 25.5 36.4 0.75

International Nickel 2 19.3 17.4 33.3 Company (Canada)

2.3 0.07 21.57 15.77 35.9 1.24 Philippines Ore 1

Yuanjiang Ores (China) 1.24 0.08 24.6 19.4 31.8 0.34 6.9

Philippines Ore 2 1.15 0.09 38 0.6 10 1.5

Yabulu Ore (Australia) 1.57 0.1 30

Cawse Ore (Australia) 1.02 0.07 18 1.58 42.5

Bulong Ore (Australia) 1.9 0.08 18 1~2 42.5

Albania Ores I 0.89 0.06 42.9 2.67 13.9 2.31 5.9

Albania Ores II 0.96 0.06 50.4 1.33 6.48 2.46 3

In the recent past, the majority of nickel was produced from sulphide ores (Whittington et al. 2000), but with the declining sulphide ore reserves and the increasing demand for

8 nickel in the world, more nickel is extracted from laterite ores (Valix et al. 2001, Zhai et al. 2009, Norgate et al. 2011), which accounts for about 70% of nickel resources (Li et al. 2009). The proportion of nickel extracted from laterite ores is shown in Figure 2.1, the percentage has increased from about 15% in 1950 to about 50% in 2010 (Li et al. 2010). Besides the abundant reserves, nickel laterites are important sources of cobalt; they are of increasing significance for exploration and production (Andersen et al. 2009).

60

50

40

30

20

Proportion of laterite ores (%) of laterite Proportion 10

0 1950 1960 1970 1980 1990 2000 2010 Year

Figure 2.1 The trend in percentage of Ni produced from laterite ores

By now, more than 120 Ni-Co laterite deposits have been explored in the world. Different from the existing state of pentlandite in sulphide ores, nickel and cobalt in laterite ores present as cations substituted within manganese oxide, goethite and /or clays; the distribution of nickel throughout the lattice of the ore particles makes the beneficiation by flotation or gravity separation impractical, which results in intrinsically high capital and operating costs (Whittington et al. 2000, Norgate et al. 2011).

Nevertheless, there are many obvious benefits for mining and processing laterite ores. Firstly, nickel laterite ores are usually nearer to the surface and can be processed by open-cut mining (Chemlink 1997, Moskalyk et al. 2002), which reduces the high mining cost. Secondly, the reserve of laterite ores is abundant accounting for 70% of nickel reserve on land, which ensures the supply for nickel resource in future. Finally, cobalt, an important metal with eight-times higher value than nickel and representing up

9 to 40% of the total value of production, is often associated with nickel in laterite ores (Chemlink 1997, LI et al. 2009). The recovery of cobalt makes the laterite ores processing more practicable (Berger et al. 2013). For example, Anaconda Nickel NL commenced development of Ni laterite deposits for Ni and Co production in Murrin Murrin, WA (Wells et al. 2003).

2.2 Extractive Metallurgical Processes for Laterite Ores

The conventional metallurgical processes for extraction of nickel from laterite ores can be divided into three routes: pyrometallurgical route (ore smelting), hydrometallurgical route (ore leaching), and Caron-type processes (Berger et al. 2013). Because composition of laterites varies significantly in different deposits, and laterite ores cannot be effectively upgraded by beneficiation, the choice of the metallurgical technology depends on the ore composition and physical properties. Both pyro- and hydro- metallurgical processes have been operated in nickel metallurgy for many decades, and are being improved. Table 2-3 lists the typical laterite ore types and their processing technology (Liu 2002).

New technologies for processing of laterites are under development, such as chloridizing segregation and bioleaching (McDonald et al. 2008).

Table 2-3 Chemical composition and processing technology of laterite ores (Liu 2002)

Chemical composition % Ore type Processing technology Ni Co Fe Cr2O3 MgO

Limonite 0.8-1.5 0.1-0.2 40-50 2-5 0.5-5 Hydro

1.5-1.8 0.02-0.1 25-40 1-2 5-15 Pyro or Hydro

Saprolite 1.8-3 0.02-0.1 10-25 1-2 15-35 Pyro

Flow sheets for processing of laterite ores are schematically shown in Figure 2.2 (Norgate et al. 2011). 10 Figure 2.2 Flow sheets of conventional metallurgical processes for laterite ores (Norgate et al. 2011)

Generally, saprolite ores with relatively high nickel, magnesia and silica contents are mainly processed by pyrometallurgical processes. Goethite-rich limonite ore which has a low content of nickel and high content of iron and cobalt is mainly processed by hydrometallurgical technology. The transition layer of laterite ore (saprolite-limonite ore) can be processed by pyrometallurgical or hydrometallurgical processes

2.2.1 Pyrometallurgical Processes

Pyrometallurgical process is the primary technology used in processing of sulphide ores (Moskalyk et al. 2002, Mudd 2009, Norgate et al. 2011). In application to laterite ore, it was firstly used to produce ferronickel in New Caledonia in 1879 (Norgate et al. 2011, Oxley et al. 2013). About 70% of mined laterite ores have been treated by pyrometallurgical processes (Xu 2005, Warner et al. 2006).

Pyrometallurgical processing of laterite ores is based on reduction of nickel oxide, which is less stable than other oxides in the ore. However, iron oxides are also partially reduced. Two major process routes are used in processing of laterite ores: production of ferronickel by selective reduction, and production of nickel matte from the selectively reduced laterite ore followed by conversion of nickel sulphide to nickel. 11 Production of ferronickel starts with processing of laterite ore in Rotary Kiln-Electric Furnace (RKEF) process (Whittington et al. 2000, Guo et al. 2009, Zhai et al. 2009, Zhao et al. 2012). Initially, the ores are screened, crushed and blended to produce feed for a consistent process plant with defined ratios of iron and nickel. Then the feed is charged into a rotary kiln where laterite ore is calcined and prereduced by and powdered coke. Afterwards the calcine and residual coke are charged into the electric furnace to produce crude ferronickel with nickel contents varying from 13 to 25%. The other components (mainly FeO, SiO2, MgO) are removed with (Christian Redl 2013).

Nickel matte smelting process (Whittington et al. 2000, Guo et al. 2009, Zhai et al. 2009, Zhao et al. 2012) is a variation of a standard flow sheet, in which sulphur is added to the kiln. The crude metal/matte is further processed/refined to produce nickel (Xu 2005).

Although pyrometallurgical processes present a mature technology, their disadvantages are also obvious. Because laterites cannot be efficiently concentrated before drying, the production of nickel from oxide ores by pyrometallurgical technologies consumes two to three times energy as processing sulphide ores with significant environmental impact. This technology is suitable only for saprolite ores. Finally, the valuable cobalt cannot be recovered (F.Habashi 1997, Whittington et al. 2000, Berger et al. 2013, Thubakgale et al. 2013).

2.2.2 Hydrometallurgical Processes Compared with , hydrometallurgical processes are more efficient in extraction of nickel and cobalt from low nickel grade laterite ores (Oxley et al. 2013). Hydrometallurgical processes are applied to the goethite-rich limonite of relatively low nickel content but high iron and cobalt contents (Whittington et al. 2000). Several hydrometallurgical processes have been developed, which include Atmospheric Leaching (AL), Acid Heap Leaching (AHP), Chloride Leaching (CL), Enhanced Pressure Acid Leaching (EPAL) and High Pressure Acid Leaching (HPAL), among which HPAL and EPAL processes are the only two processes reported to be used in industry (Xu 2005, McDonald et al. 2008, Zhai et al. 2009, Norgate et al. 2011).

12 In the laboratory study, AL was tested using sulphuric acid for limonite, smectite (clay) and saprolite laterites with advantages of lower capital and operation costs (McDonald et al. 2008). The leaching rate increased with temperature, acid concentration and the activity of the acid (McDonald et al. 2008).

The first commercial HPAL plant, operated at temperatures between 245 and 270 oC, started production of nickel and cobalt in 1959 at Moa Bay in Cuba and still continues Ni production (Oxley et al. 2013). HPAL process is also been used to extract nickel and cobalt from laterite ores in Western Australia (Whittington et al. 2003).

Hydrometallurgical processes include acid leaching and leaching technologies. The ammonia leaching was applied in the Caron process, which will be considered in the next section.

The acid leaching is implemented in titanium-lined autoclaves (Australian projects) or acid brick and lead-lined autoclaves (Moa Bay project) at approximately 250℃. The nickel and cobalt bound with the goethite or clays are released to the acid solution in the pressure leach at high pressure (>200 atm), nickel and cobalt can be extracted with high recoveries at 99.7%, and 98% correspondingly (Whittington et al. 2000, Jiang et al. 2004). The extraction of metals by of HPAL in Cuban Moa Bay Nickel Company is shown in Table 2-4 (Peng 2005).

Table 2-4 Extents of metals extraction from laterite ores by HPAL process at Moa Bay Nickel Company (Peng, 2005)

Component Leaching solution g·L-1 Leaching residue % Leaching rate % Ni 5.95 0.08 96

Co 0.04 0.008 95

Cu 0.1 - 100

Zn 0.2 - 100

Fe 0.8 51 0.4

Mn 2.0 0.4 57

13 Cr 0.3 - 3.0

Cr2O3 - 0.3 -

The main chemical (leaching) reactions involved in the hydrometallurgical process are given below (Jiang et al. 2009).

2 NiO 2H  Ni  H2 O (2-1)

2 2Ni2 O 4  12H  6Ni  6H2 O  [O 2 ] (2-2)

3 FeOOH 3H  Fe  2H2 O (2-3)

+ 3+ Fe2 Si 4 O 10 (OH) 2 + 6H  2Fe + 4SiO2 + 4H 2 O (2-4)

+ 2+ Co2 O 3 + 6H 2Co + 3H2 O (2-5)

Currently, sulphuric acid is used in most hydrometalligical processes for leaching nickel laterites. The use of hydrochloric acid for leaching process has also been examined (McDonald et al. 2008).

The advantage of hydro-metallurgy is high recoveries of both nickel and cobalt, but there are many challenges. One big challenge is the high capital and operation costs of the high pressure and temperature process and the scale formation in autoclaves. The use of sulphuric acid and ammonia for leaching potentially causes serious environmental problems. This route is not economic for processing ores containing acid consuming compounds, such as magnesia; in which the leaching of nearly all iron and magnesia results in a high acid consumption (Whittington et al. 2000, McDonald et al. 2008, Zhai et al. 2009).

2.2.3 Caron Process Caron process was developed in the 1920s for processing low grade laterite ores (Berger et al. 2013). Caron process is a combined pyro-/hydro-metallurgical process (Graaf et al. 1979, Whittington et al. 2000, Zhai et al. 2010); its schematic flow sheet is shown in Figure 2.3 (Taylor 2013).

In the Caron process, the first step is the reduction of nickel and cobalt oxides to metallic state at temperature above 700 oC, then nickel and cobalt are leached by the 14 ammonia-ammonium solution. After metals are transferred to the solution, ammonia is removed by boiling, and basic nickel carbonate precipitates which, upon at 1200 oC, is converted to nickel oxide (Whittington et al. 2000, Nicol et al. 2004).

The main leaching and precipitation reactions in Caron process are listed as follows:

1 Ni O  6NH  CO  Ni(NH )22  CO (2-6) 2 2 3 2 3 6 3

22 Ni(NH3 ) 2  Ni 2NH3 (2-7)

22   6Ni 6OH  2CO3  3Ni(OH)2  2NiCO3 (2-8)

3Ni(OH)2  2NiCO3  5NiO  3H2 O  2CO2 (2-9)

3 2Co O  12NH  3CO  2Co(NH )32 3CO (2-10) 2 2 3 2 3 4 3

1 Fe O  nNH  CO  Fe(NH )22  CO (2-11) 2 2 3 2 3 3 3

1 2Fe(NH )2 5H O  O  2Fe(OH)   2(n  2)NH  4NH (2-12) 3 n 2 2 2 3 3 4

15 Figure 2.3 Schematic flow sheet of Caron process (Taylor 2013)

Caron process has disadvantages of the both pyro-and hydrometallurgical processes. Firstly, the initial ore drying step is characterised by high energy consumption; secondly, the cobalt recovery is low (usually less than 40%) and the extraction is particularly low from saprolite ores; finally it is sensitive to mineral composition and requires careful blending (Zhou 2005, Norgate et al. 2011).

2.2.4 Bio-metallurgical Process Bioleaching is a novel technology applicable to metal extraction from low-grade ores. During the bioleaching process, metals are dissolved from oxide and silicate ores into organic (citric acid, , gluconic acid, tartaric acid and pyruvic acid) produced from heterotrophic microorganisms. Aspergillus and Penicillium are two important filamentous fungi extensively studied for extraction of metals from oxide ores (Kar et al. 1996, Coto et al. 2008). In fungal bio-leaching of laterite ores for 21 days, nickel extraction reached only 36 %, cobalt 54% and iron 0.76 % (Valix et al. 2001). Bio-leaching is less efficient in comparison with extraction of metals by conventional technologies. There is a long way to go before bio-metallurgical extraction of Ni and CO can be commercially used in the industry.

2.3 Carbonylation

Carbonylation process was developed in application to nickel refining at the beginning of last century; it was also applied for production of ultra-pure iron (Mond et al. 1890, Herrmann 1990).

Nickel and iron carbonyls are formed at elevated temperatures and pressures, separated, isolated and thermally decomposed to yield pure metals and carbon monoxide gas. Cobalt which is often present with nickel and iron in laterite ores and tailings, forms nitrosyl carbonyls and can be extracted (Terekhov 2001).

2.3.1 Industrial Carbonylation Process The industrial application of carbonylation technology to nickel refining commenced in 1902 (Herrmann 1990).

16 The refining process was based on the different reaction rates and physical properties of metal carbonyls (different boiling points). Except for nickel, iron and cobalt, other metal carbonyls are not directly synthesized by reactions with carbon monoxide. The carbonylation rates of nickel, iron and cobalt vary greatly. Nickel reacts with CO rapidly and is completely converted into gaseous nickel tetracarbonyl, while iron and cobalt react with carbon monoxide relatively slowly (Rigin 1993). After carbonylation, nickel tetra-carbonyl and iron penta-carbonyl are separated by distillation from the carbonyls mixture. Cobalt carbonyl, having much lower vapour pressure than nickel and iron carbonyls, remains in the carbonyl reactor together with solid residues of the carbonylation reaction (Huang 1990, Terekhov 2001, Peng 2005).

The first patent (English patent No. 12626) on carbonylation was granted in 1890 (Herrmann 1990). In this patent, nickel ores were firstly roasted (calcined), then reduced at temperatures as low as possible using CO-H2 gas mixture, and finally nickel was converted to nickel tetra-carbonyl by reaction with carbon monoxide at 50 oC and 0.1 MPa. The mixture of the excess of CO and nickel tetracarbonyl was heated to 180 oC.

At this temperature, Ni(CO)4 was decomposed to pure metallic nickel and CO with nickel recovery of 92%. Based on this technology, Mond-nickel Co. Ltd produced 3000 tons of nickel in 1910, the Vale Limited started production of nickel carbonyl under pressure 7 MPa at 180 oC in 1973 (Herrmann 1990, Huang 1990).

In another patent (Terekhov 2001), cobalt and nickel carbonyls were formed at elevated temperatures and pressures, separated, isolated and thermally decomposed to yield pure metal pellets and carbon monoxide gas. Impure cobalt carbonyl reacted with a complex gaseous mixture of nitric oxide/carbon monoxide to produce cobalt nitrosyl carbonyl. The process was carried out at 60~100 oC, under 3 atm operation pressure. The reported recoveries of nickel and cobalt reached 97% and 94% correspondingly. Figure 2.4 shows the schematic diagram from this patent (Terekhov 2001).

17 Figure 2.4 Flow sheet of the extraction of Ni and Co using carbonylation process (Terekhov 2001)

Besides the plants mentioned above, carbonylation process to refine nickel or iron, has been operated by different companies. In Germany, BASF Aktiengesellschaft is o reported to annually produce approximately 10,000 tonnes of Fe(CO)5 at 200 C and CO pressure 204 atm (Herrmann 1990), which is used for matins iron and iron crude oxide pigments.

International Nickel Company (INCO, Canada) applied the medium pressure carbonylation process (7 MPa, 180 oC) to produce annually 45,000 tonnes of nickel (Peng 2005). Norilsk company (Russia), extracted nickel by high pressure carbonylation process (25 MPa, 280 oC) with a production of 5, 000 tons per year from 1963 (Peng 2005).

The carbonylation refining was operated under atmospheric, medium or high carbon monoxide pressure. Figure 2.5 shows the schematic diagram of atmospheric carbonylation at INCO of Clydach (Boldt 1957, Teng-1 2006, Teng 2006).

Under atmospheric pressure, the impurities in crude nickel metal did not enter the gaseous phase. Iron forms volatile iron pentacarbonyl with very low rate, cobalt can form tetracarbonyl, Co2(CO)8, and tricarbonyl, Co4(CO)12, but both carbonyls have low volatility. Therefore, the extraction of nickel as carbonyl from a crude feed is a highly selective process.

18 At INCO refinery, Clydash (Wales), the first step was the reduction of oxides by hydrogen-water gas at relatively low temperatures 343 - 427 oC, the degree of reduction at these temperatures was about 90%. Reduced nickel reduced was carbonylated at 38 – 60 oC by carbon monoxide; the extent of carbonylation was approximately 95% after a 4 days reaction.

NICKEL OXIDES GRANULES Water Gas

Combustion Gas Liquor To Atmosphere Water

Gas Water REDUCER H2 (3 in Series) CO ABSORBER CO 300 psig ABSORBER Producer 300 psig Gas Gas

CO to DUST COMPRESSER 2 SCRUBBER Atmosphere CO Air Gas SEPARATORS Impure Nickel Sludge to Granules Thickeners Air

Nickel Carbonyl Gas

Granules Make-up CO Gas

Cooling Gas Water Filter

Recycle Pellets

VOLATILIZER (6 in Series) Overflow Pellets SCREEN Hot Gases -5/16" PREHEATER

CO Gas DECOMPOSER

Residue to Sulfating Roast & Leach

NICKEL PELLETS TO MARKET

Figure 2.5 Flow sheet of nickel production using the atmospheric pressure carbonylation at INCO, Clydach (Teng-1 2006)

In order to improve the carbonylation process, the operation pressure was increased to enhance the reaction rate and stabilize the carbonyl at temperatures over 100 oC. Liquid nickel tetra-carbonyl produced under medium pressures was then vaporized and purified

19 by distillation. Finally, ultrapure nickel powder was produced from the decomposition of nickel tetra-carbonyl at atmospheric pressure and temperature above 180 oC.

Figure 2.6 shows the schematic flow sheet of nickel carbonylation under medium pressures at INCO (Teng 2006).

NICKEL OXIDES Hydrogen Enriched GRANULES Water Gas Combustion Gas To Atmosphere

REDUCER Water Producer Gas

Inert Gas DUST SCRUBBER Gas Air Sludge to STORAGE Thickeners HOPPERS

Inert Gas Reaction Gas Liquid Nickel (CO & Carbonyls) Carbonyl BATCH VOLATILIZER Cooling Hot Water Cooling (300 PSIG) Water Water VAPORIZER CO DISTILLATION Liquid co COLUMN BLOWER Carbonyls DECOMPOSER SEPARATOR Vapor Hot Gas Residue to Return CO Sulfating Roast COMPRESSOR & Leach Liquid Hot Water CO from Copper Liquor Plant REBOILER

CO

Iron Carbonyl Residue

NICKEL POWDER TO MARKET

Figure 2.6 Flow sheet of nickel production using medium pressure carbonyl process at INCO, Clydach (Teng 2006)

High pressure carbonylation was used by the Norilsk company (Teng 2006). The carbonylation reactions were carried out at temperatures of 150-250 oC and pressures of 225 atm. About 95% of nickel was carbonylated after 3 days operation. Iron and cobalt were also partially converted to carbonyls, nickel tetracarbonyl was purified by rectification. 20 2.3.2 Carbonylation of Nickel

As mentioned above, nickel tetracarbonyl, Ni(CO)4, was firstly synthesized in 1890 (Mond et al. 1890); fine metallic nickel reduced from nickel oxide reacted with CO at about 100 oC under atmospheric pressure (reaction (2-13)), then gaseous nickel carbonyl was collected in cold container. In order to recover nickel, nickel carbonyl was decomposed to Ni and CO at about 180 oC (reaction (2-14)) (Mond et al. 1891). In industry, nickel carbonylation is produced at elevated temperatures and pressures to improve the reaction rate and yield. Nickel is recovered by decomposition of Ni(CO)4 at temperatures from 150 to 316 oC (Peng 2005).

Ni + 4CO = Ni(CO)4 (2-13)

Ni(CO)4 = Ni + 4CO (2-14)

(Goldberger et al. 1963) studied the kinetics of nickel carbonyl formation from pure carbon monoxide and reduced nickel powder in a fixed bed at temperatures between 25 and 150 oC, CO pressures up to 4 atm, and gas velocities from 1.17 cm·s-1 to 1.92 cm·s-1.

The increase in CO pressure had a positive effect on the reaction; the maximum carbonylation rates were obtained at about 75 oC under all pressures. Surface was found to control the overall reaction rate. Experimental data were described by a modified first-order rate equation (2-22) o as ∗ In o = k (P − P )CO ∗ θ (2-15) as −x

Where, θ is reaction time, minute; P is partial pressure of CO, atm; P* is equilibrium

o partial pressure of CO, atm; x is conversion or fraction of nickel; a s is activity of nickel.

(Wang et al. 2009) discussed the effect of ore , particle size, CO partial pressure, reaction temperature, flow rate, sulphur and hydrogen sulphide on the carbonylation of nickel. Ferronickel re-activated with H2 is essential in the process, the optimum parameters for the carbonylation reaction were temperature 70oC and CO pressure of 9.9 atm; addition of sulphur and hydrogen sulphide greatly increased the carbonylation rate.

21 (Mazurek et al. 1982) established that the lowest CO pressure in production of Ni(CO)4 was 667 Pa, and increasing CO pressure could enhance the reaction rates. In order to check whether the reactivation by H2 was from the removal of O2 (the reactivation mechanism was supposed to remove oxygen adsorbed on the nickel surface before o carbonylation), samples in their work were heated at 400 C in H2 atmosphere for 4 hours prior to the carbonylation; once the carbonylation rate was steady, H2 was introduced to the system instead of CO for 30 minutes at carbonylation temperature to check whether the reaction rate is boosted again. It was suggested that the heat treatment of metal sample rather than the removal of oxygen from the surface promoted the carbonylation reaction.

(De Groot et al. 1980) in a study of nickel carbonylation at temperatures up to 160 oC observed the maximum reaction rate at 125oC. The effect of temperature on the formation of nickel carbonyl is shown in Figure 2.7; Figure 2.8 demonstrates the effect of CO pressure on the rate of formation of nickel carbonyl.

(111) SURFACE -3

-4 (a)

-5 (b) Log(Ni(CO)4 pressure) Log(Ni(CO)4

(c) -6

40 80 120 160 200 Temperature (oC)

Figure 2.7 Effect of temperature on the Ni(CO)4 pressure at the reactor outlet at 3 -1 different CO flow rates (in cm s ): (a) 0.37, (b) 2.8, (c) 12.0. Broken line: Ni(CO)4 pressure at chemical equilibrium; Pco=1 atm (De Groot et al. 1980)

The reaction rate increased with carbon monoxide pressure at an apparent order of 1.45 (Figure 2.8).

22 Figure 2.8 Effect of CO pressure on the rate of nickel carbonylation at 20 oC (De Groot et al. 1980)

A study of nickel carbonylation using chemisorption (Schmidt et al. 1982) showed that CO mainly adsorbs in the form of linearly and bridge-bonded molecules Ni(CO) and

Ni2(CO). The intermediate Ni(CO)2 complex was assumed to be formed by reaction (2-

16) when the coverage of nickel surface by CO, θ1 , was over 0.5, which represents the fraction of Ni surface atoms covered by linearly bonded CO.

3Ni(CO)  Ni(CO)22 + Ni (CO) (2-16)

 The adsorption equilibrium constant for the reaction Ni2 CO + CO 2Ni(CO) was

2 calculated as KP211 /CO (1 ) . A reaction order for this process was found equal to 1.5.

Further study confirmed that nickel carbonylation was more readily taken place on step atoms than perfect crystal surface atoms (Schmidt et al. 1982).

In a study of the carbonylation of Cu-Ni alloy (Teng et al. 2007) at temperatures between 120 and 150 oC under CO pressure in the range 7 ~ 18 MPa, the recovery of Ni after 42- hour reaction was about 95%; the reaction rate increased with the decrease of 23 particle size from 5 to 0.5 mm. The rate limited step was found to be the diffusion of nickel tetracarbonyl in the gas phase (Teng et al. 2007).

The maximum rate of the carbonylation reaction in work (Goldberger et al. 1963) was observed at 117 oC; the reaction rate was found to be a function of reaction time and surface composition (coverage by C, S and O).

In industry, the carbonylation is performed in a fixed bed reactor in vertical columns under a gauge pressure of 200-250 atm. For a complete recovery, the size of Ni grains should not exceed 10 ~ 15 µm (Kozyrev et al. 2005)

2.3.2 Carbonylation of Iron

Iron pentacarbonyl (Fe(CO)5) was synthesized by reaction (2-17) in 1891. Iron carbonyl decomposes to iron and CO at 180 oC by reaction (2-18) (Mond et al. 1891, Mond et al. 1891, Dewar et al. 1905).

Fe + 5CO = Fe(CO)5 (2-17)

Fe(CO)5 = Fe + 5CO (2-18)

Mond (Mond et al. 1891-1) targeted production of iron carbonyl by reaction of fine iron o reduced from iron oxide and reactivated by H2 at 300 C, and carbon monoxide, at a temperature of 180 ~ 220 oC under CO pressure of 150 ~ 250 atm for 24 hours. The reaction rate was very slow, only about 1% carbonyl was collected (Mond et al. 1891-1).

Effects of temperature (100-300 oC) and pressure (100 - 300 atm) on synthesis of iron pentacarbonyl studied by Mond and Wallis (1922) are shown in Figure 2.9. The o optimum temperature for production of Fe(CO)5 was found to be near 200 C, above which carbonyl decomposes to iron and CO (Mond et al. 1922).

24 15 14 13 12 11 10 9 300 8 7 200 6 5 150 4 3

Cubic centimetre of Fe(CO)5 100 2 1 0 100 150 200 250 300 Temperature Figure 2.9 Effect of temperature and pressure on synthesis of iron pentacarbonyl (Mond et al. 1922)

2.3.3 Carbonylation of Cobalt Cobalt carbonyl was synthesized by reaction (2-19) at 50 oC under CO pressure 150 atm in 1908 (Mond et al. 1908, Mond et al. 1908)

2Co 8CO Co28 (CO) (2-19) Cobalt carbonyls are large molecules, which are not readily vaporized under these experimental conditions; the carbonylation reaction was greatly enhanced by raising the pressure to 200 atm and the temperature to 200oC (Mond et al. 1908, Mond et al. 1908).

Compared with pure cobalt, the carbonylation rate of cobalt from the alloy with nickel (88-94 wt% Ni) increased 83 times. The extraction of cobalt was affected by nickel and iron carbonylation as a result of breakdown of the crystal lattice with removal of nickel and iron, or the catalytic and autocatalytic action of the formed carbonyls (Kipnis et al. 1973). Cobalt carbonyl nitrosyl, which is gaseous at room temperature, can be synthesized by reaction of cobalt tetra carbonyl with nitric oxide (reaction (2-20)) at 75-80 oC (Ivanova et al. 1999):

Co(CO)43 NO  Co(CO)  NO  CO (2-20)

25 Cobalt carbonyl nitrosyl can also be prepared by the method developed by Blanchard et al. (1934) in an alkaline suspension of cobaltous cyanide (Brockway et al. 1937) by the following reactions (Coleman et al. 1936, Ewens et al. 1939, Blanchard et al. 1940):

2CoCl2  11CO  12KOH  3K2 CO 3  2KCo(CO)4  4KCl (2-21)

1 KCo(CO) NO  H O  Co(Co) NO CO  KOH  H (2-22) 4 2 3 2 2

2.3.4 Factors Affecting the Carbonylation

The carbonylation processes are affected by temperature, carbon monoxide pressure, gas flow rate, particle/grain size. Sulphur containing compounds have a catalytic effect on the reaction. Carbon and reactivation of metals by hydrogen were also reported to influence the carbonylation reactions.

Effect of Temperature In the synthesis of Ni, Fe and Co carbonyls in Mond’s experiments, temperature was a critical parameter (Abel 1963). The formation of carbonyls was reported to be accelerated at elevated temperatures in many papers (Windsor et al. 1933, Hieber et al. 1950, Podall 1961, Abel 1963, Nametkin et al. 1975, Liang et al. 1983, Terekhov 2001, Kozyrev et al. 2005, Teng et al. 2007). Although the carbonylation of nickel, iron and cobalt could occur at temperatures below 100 oC, commercial carbonylation process is carried out at relatively high temperatures and pressures to increase the reaction rate and yield; in work by (Kipnis et al. 1973) carbonylation temperature was 200 oC under CO pressure 200 atm.

The effect of temperature on nickel carbonylation has been reported in many papers, the optimum temperature was found to be 70 oC at 10 atm (Wang et al. 2009), 75 oC (temperature investigated between 25 and 150 oC) at CO pressures up to 4 atm (Goldberger et al. 1963), 77 oC (temperature investigated between -3 and 147 oC) under pressures below 3*10-2 Pa(Liang et al. 1983), 117 oC(Goldberger et al. 1963) and 125 oC (in a study at temperatures up to 160 oC) (De Groot et al. 1980).

Effect of Carbon monoxide pressure

26 Carbon monoxide pressure was another important parameter in Mond’s experiments on synthesis of metal carbonyls (Abel 1963). The carbonylation reaction is favoured by the high CO gas pressure (Medvedev et al. 1998). High pressure of carbon monoxide was reported to increase the rate of carbonyl formation on a compact metal surface (Windsor et al. 1933). The increase of carbon monoxide pressure enhanced the reaction rate in work by Wang et al. (2009). Effect of CO pressure on catalytic carbonylation of nickel is shown in Figure 2.10 (Wang et al. 2009). Initial rate of carbonylation decreased with increasing CO pressure from 10 to 20 atm, although it was significantly higher at higher CO pressure after approximately 2.5 hours reaction. In industry, the carbonylation is performed under a gauge pressure of 200-250 atm (Kipnis et al. 1973, Kozyrev et al. 2005).

Figure 2.10 Effect of CO pressure on catalytic carbonylation of nickel at 70 oC (2 wt% sulphur added before carbonylation); a: atmospheric pressure, b: 9.9 atm, c: 19.8 atm

Catalytic effect of sulphur

Sulphur and inorganic sulphur compounds (FeS, NiS and H2S) can catalyse the carbonylation of nickel and iron (Heinicke et al. 1963, Heinicke et al. 1970).

Figure 2.11 shows the effect of hydrogen sulphide content in CO gas on the nickel carbonylation rate (Heinicke et al. 1963). The reaction rate increased sharply with increasing hydrogen sulphide content up to 1.0 vol%, further increase in the H2S concentration had a minor effect on the carbonylation rate.

27 100 ) -1 90

80 30 minutes 30  2 70 10  4 60

50

40

30

20

10 Reaction yield (mmol Ni(CO) (mmol yield Reaction 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0

% H2S o Figure 2.11 The effect of H2S on carbonylation of nickel at 25 C and 1 atm (Heinicke et al. 1963)

Effect of hydrogen sulphide on the rate of carbonylation of nickel and iron studied by (Heinicke et al. 1970) is shown in Figure 2.12.

(Windsor et al. 1933) studied nickel carbonylation with addition of iron sulphide. They suggested that the catalytic effect of sulphur is related to formation of nickel tetracarbonyl from nickel sulphide and carbon monoxide.

28 25 Ni 20

15 CO

CO + 3 % H2S 10

5

0 0 5 10 15 20 25 30 35 40 45 50 55

14 Fe 12

Carbonyl yield (mmol/h) yield Carbonyl 10

8 CO

6 CO + 3 % H2S 4 2 0 0 5 10 15 20 25 30 35 40 45 50 55 Processing time (hour)

o Figure 2.12 Influence of H2S on iron and nickel carbonylation at 25 C and CO pressure 1.1 atm (Heinicke et al. 1970)

(Den Besten et al. 1962) suggested the following mechanisms of catalytic effect of sulphur on the nickel carbonylation. Hydrogen sulphide is chemisorbed on nickel surface and dissociated to hydrogen and sulphur (Figure 2.13) Hydrogen is adsorbed on available surface sites, but at high surface coverage, hydrogen is forced out to the gas phase. Sulphur reacts with nickel forming Ni-S complexes as shown in Figure 2.13, which is easily converted to nickel carbonyl.

29 Figure 2.13 Chemical adsorption of H2S on nickel surface (Den Besten et al. 1962)

However, the increase of sulphur content results in a decrease of the rate and extent of carbonylation (Kipnis et al. 1973, Miura et al. 2008). Conversion of cobalt to carbonyl decreased with increasing content of sulphide phases as shown in Figure 2.14 (Kipnis et al. 1973), what was related to the binding of cobalt in sulphide form.

100 quenched at 550 oC cast 80 quenched at 900 oC Alloys (Cu:Ni= 1:9) 200 oC 200 atm 10 h 60

40

20 Degree of Co extraction (%)

0 0 2 4 6 8 10 12 S content (%)

Figure 2.14 The effect of sulphur on the carbonylation of cobalt at 200 oC and CO pressure 200 atm for 10 hours (Kipnis et al. 1973)

In the synthesis of iron carbonyl in work (Hieber et al. 1950) , the addition of a small amount of sulphur (approximately 1 wt%) had a remarkable positive effect on the formation of iron pentacarbonyl at 200 oC and CO pressure of 200 atm, which gradually diminished with increasing sulphur addition. The effect of sulphur content on the carbonylation of iron is shown in Figure 2.15.

30 100

80

60

40

20 Extraction proportion of Fe (%)

200 atm; 200 oC; 15 hour 0 0 10 20 30 40 50 60 70 80 90 100 Addition of S (Atom - %)

Figure 2.15 The effect of sulphur on iron carbonylation at 200 oC and CO pressure 200 atm (Hieber et al. 1950)

A similar effect of sulphur addition was observed in carbonylation of nickel in the Ni- Fe alloy, which is shown in Figure 2.16 (Wang et al. 2009).

100

80

60

40 Yield (%)

2% S, CO 20 5% S, CO 10% S, CO 0% S,CO + H2S 0% S,CO + H2S (added intermittently) 0 0 4 8 12 16 20 24 Time (hr) Figure 2.16 Effect of sulphur on nickel carbonylation at 70oC and CO pressure of 10 atm (Wang et al., 2009)

Miura et al. (2008) used sulphur as a catalyst in the extraction of iron from the scrap (powder) of Nb-Fe-B sintered magnets via the carbonylation; reactions were implemented at 200 oC under CO pressure of 30 MPa for 24 h. Effect of sulphur addition on the carbonylation of iron is shown in Figure 2.17. The yield of iron carbonyl Fe(CO)5 was evaluated to be ~ 56% without pre-treatment. In other

31 experiments, iron produced by the disproportionation of Sm2Fe17 in hydrogen (HD) (Sugimoto et al. 2002), reacted with carbon monoxide under the same conditions, a maximum yield of 92% was achieved. Both the results are shown in Figure 2.17.

100

80

60

40

20 Extent of carbonylation (%) Extent As-obtained HD 0 0 2 4 6 8 10 12 14 16 18 20 S/[S+Fe] (%)

Figure 2.17 Effect of sulphur on extraction of iron from Nb-Fe-B via carbonylation

Effect of Oxygen

The effect of oxygen on nickel carbonylation is conflicting (Lascelles et al. 1983) reported that oxygen had a positive effect on nickel carbonylation as shown in Figure 2.18. It was suggested that oxygen is chemisorbed on the nickel surface and incorporated into the nickel lattice, which activated nickel atoms in the carbonylation reaction. However, Mazurek et al. (1982) reported that reactivation of metals by hydrogen before carbonylation removed oxygen.

32 6.0E-9 ) -1  min -1 4.0E-9

 cm (b)

2.0E-9

(a) Reaction (mole rate

0.0 0 20 40 60 80 100 Time (minutes) Figure 2.18 Effect of oxygen on the carbonylation of Ni (110) at 130oC. (a) 1ppm oxygen in carbon monoxide, (b) 3ppm in carbon monoxide (Lascelles et al. 1983)

Reactivation of Metal Surface by Hydrogen Hydrogen was used to reactivate the metal before carbonylation by heating metal in hydrogen (Windsor et al. 1933). (Wang et al. 2009) annealed a metal sample in hydrogen atmosphere at 400oC. Ni-Fe alloy was reactivated by hydrogen at 150oC for two hours before the introduction of carbon monoxide. Figure 2.19 shows the effect of reactivation of nickel by hydrogen at 150 and 250 oC on the reaction of nickel carbonylation (Wang et al. 2009). Heating of nickel in a reducing atmosphere removed adsorbed oxygen, which inhibits nickel carbonylation (Germer et al. 1962).

100

80

60

Yield (%) 40

20

t : reactivation temperature ℃ ℃ r tr 150 , tc 70 , CO+H2S

t 250 ℃ , t 70 ℃ , CO+H S tc: carbonylation temperature r c 2 0 0 5 10 15 20 Time (hr)

33 Figure 2.19 Carbonylation of nickel with pre-treatment by hydrogen at 150 and 250 oC (Wang et al. 2009)

Effect of Particle size The decrease of nickel particle size has a positive effect on the carbonylation of nickel. The extent of carbonyl formation on a compact metal surface was extremely small; very fine subdivision of the metal increased the yield of carbonylation (Windsor et al. 1933). In Mond’s experiments (Mond et al. 1891-1), iron carbonyl was prepared using iron fines reduced from iron oxide. The decrease of particle size increased the specific surface area, the contacting area between carbon monoxide and metal.

Effect of Carbon In carbonylation of metals at high temperature, carbon can be deposited on the metal surface as a result of partial decomposition of carbon monoxide. (Kozyrev et al. 2005), added carbon to nickel-copper-cobalt alloy to study the effect of carbon on cobalt carbonylation. It was reported that carbon inhibited carbonylation of cobalt. The effect of carbon on carbonylation of cobalt is shown in Figure 2.20 Increase in carbon concentration led to a dramatic decrease in the cobalt recovery from the alloy (Kozyrev et al. 2005).

20

16

12

8

Mass recovery of Co (%) 4

0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 Mass content of carbon in alloy (%) Figure 2.20 Effect of carbon content on cobalt carbonylation at 180 oC and CO gauge pressure 70 atm (Kozyrev et al. 2005)

34 2.4 Reduction of Metal oxides

Metal oxides are reduced to the metallic state prior to carbonylation. This project included reduction of nickel, iron and cobalt oxides by hydrogen and selective reduction of laterite ore by the carbon monoxide - carbon dioxide mixture. However, a study of the reduction processes was outside the scope of this project; they were investigated in the PhD project “Selective reduction of laterite ore” by Jun Yang (Yang, 2014). This section presents a brief overview of literature on the reduction of nickel, iron and cobalt oxides by hydrogen, which was studied in numerous papers.

Reduction of oxides at low temperature was suggested by Mond et al. (1890)(Mond et al. 1890) to minimise sintering of reduced particles.

Complete reduction of iron oxide by hydrogen in work by (Kawasaki et al. 1962) was achieved within 60 minutes at 700 oC; at 345 oC, reduction took 120 min (Sastri et al. 1982). Reduction of at temperatures below 570 oC occurred by the subsequent steps Fe2 O 3 Fe 3 O 4 Fe (Jozwiak et al. 2007). Hematite with grains smaller than 5μm were completely reduced in 10 minutes at temperatures above 550oC. Rising temperature accelerated the reaction rate up to 900oC (Wagner et al. 2008). Lin et al. (2003) first reduced hematite to magnetite at 297 oC, which was further converted to iron at 497 oC. Reduction of nickel oxide was achieved at temperatures from 250 to 350 oC (Rodriguez et al. 2002, Janković et al. 2008). Nickel oxide was completely reduced at 300 oC by hydrogen with gas flow rate of 3.0 L·min-1 for 24 hours (Parravano 1952);

Rashed and Rao (1997) reduced nickel oxide by hydrogen at temperatures over 340 oC. Jeangros et al. (2013) reported NiO reduction at temperatures between 350 and 400 oC; nickel grains size increased when the temperature exceeded 600 oC. The reduction rate increased with increasing temperature from 400 to 600 oC (Utigard et al. 2005), and decreased noticeably with further increase in the temperature; reduction became very fast at temperature above 950 oC.

35 2.5 Summary and Project’s Objectives

Industrial production of nickel from laterite ores is gaining momentum. Processing of laterite ores, started in Australia in 1998, has significantly enhanced Australian role on the international nickel market.

New technology development in processing of laterite ores has been mainly in the direction of hydrometallurgical processes. However, in general, hydrometallurgical processing of laterite ores is complex, involves high capital investment, consumes huge amount of hazardous chemical materials such as ammonia and sulphuric acid, and causes significant environmental concerns. Hydrometallurgical processing of laterite ores is economically sustainable only when cobalt value of the ore is also recovered.

An attractive approach to processing the low nickel grade laterite ores was undertaken in the ARC Discovery project DP1094880 “A Novel Approach to Processing of Australian Laterite Ores through Selective Reduction and Carbonylation”.

This PhD project studied carbonylation of selectively reduced laterite ore.

Carbonylation technologies have been used in the refinery of nickel for many decades. However, literature data on kinetics and mechanisms of carbonylation reactions, effects of temperature, CO pressure and other parameters, catalytic effect of sulphur are limited and often inconsistent. The aim of this project is a comprehensive study of carbonylation of nickel and iron reduced from pure oxides and selectively reduced laterite ore.

The project includes a study of non-catalytic carbonylation of pure Ni and Fe, and nickel-iron mixture; catalytic carbonylation of these metals with addition of FeS, S and

H2S; non-catalytic and catalytic carbonylation of reduced laterite ores.

36 Chapter 3 Experimental

The project involved an experimental study of reduction of nickel and iron oxides, carbonylation of nickel, iron and reduced laterite ores, measurements of metals concentrations in aqua regia by ICP-OES, and characterization of reduced samples and residue of samples after carbonylation by XRD, XRF and SEM/EDS.

3.1 Materials

Chemicals and gases used in the project are listed in Table 3-1.

Table 3-1 Chemicals and gases

Chemicals Specification Application

Carbonylation after reduction Nickel oxide (NiO) Powder, - 45μm, 99% by H2 Carbonylation after reduction Iron oxide (Fe2O3) Powder, < 5μm, > 99% by H2

Sulfur (S8) Powder, 99.9% Catalyst used in carbonylation Iron (FeS) Powder, - 150μm Catalyst used in carbonylation Ultra high purity, Reaction gas for reduction; Hydrogen (H2) 99.999% activation gas for carbonylation Ultra high purity, Nitrogen (N2) Purging gas 99.999% Reactant gas for carbonylation Carbon monoxide (CO) High purity, 99.99% reducing gas for laterite ores

A gas component for selective Carbon dioxide (CO2) High purity, 99.99% reduction of laterite ores Hydrochloric acid Reagent for preparing aqua Concentration 32% wt% (HCl) regia

37 Reagent for preparing aqua Nitric acid (HNO3)) Concentration 70% wt% regia Sodium Absorbent of corrosive gas from Pellets, lab research grade (NaOH) the outlet

-1 Deionized water (H2O) Resistivity 18 Ω∙cm Diluting reagent

Glycerol Analytical grade Fluid for the oil bath

Nickel nitrate ICP-OES standard 2% Standard in the ICP-OES

(Ni(NO3)2) HNO3 analysis ICP-OES standard 2% Standard in the ICP-OES Iron nitrate (Fe(NO3)3) HNO3 analysis Cobalt nitrate ICP-OES standard 2% Standard in the ICP-OES

(Co(NO3)3) HNO3 analysis ICP-OES standard 2% Standard in the ICP-OES (H2SO4) HNO3 analysis

BCS and MIN garnieritic laterite ores were provided by CSIRO Australia. The composition of the ores analysed by XRF and ICP-OES at the Analytical Center of UNSW is shown in Table 3-2. Effect of the particle size on carbonylation reactions was studied for the selectively reduced MIN ore; effects of other parameters (temperature, CO pressure and flow rate) on carbonylation of Ni and Fe was examined for the selectively reduced BCS ore.

Table 3-2 Composition of MIN and BCS garnieritic laterite ores (wt%)

BCS Ore MIN ore Element 53 ~ 200 μm <53 μm 38 ~ 53* 75 ~ 90* 140 ~ 200* 355 ~ 495* Ni 1.35 1.66 2.37 2.12 2.094 1.684 Mg 12.6 13 8.59 6.72 7.14 7.99 Si 20.5 19.4 - - - - Fe 10.4 11.7 11.83 10.68 10.57 8.90 Co 0.038 0.045 0.14 0.14 0.14 0.13

38 Al 2.2 2.1 - - - - Cr 0.5 0.5 - - - - Mn 0.2 0.2 - - - -

The experimental facilities and glassware used in the study of reduction and carbonylation processes are listed in Table 3-3.

Table 3-3 Facilities and glasswares

Facilities/ glassware Specification Application

Horizontal furnace Max. 1000 oC A furnace for reduction experiments Oil bath Max. 300 oC Carbonylation temperature control Dreschel bottle Vol. 125 ml, 250 ml Absorbent container Volumetric flask Vol. 1000 ml Diluting liquid sample

Measuring cylinder Vol. 200 ml Prepare aqua regia

Powder drill - Stirrer in reaction Swagelok flex tube SS-12BHT-36 Carbonylation reactor

Concentrate nitric acid solution and aqua regia were tested as the absorbent for nickel and iron carbonyls in preliminary experiments. Nitric acid solutions, 30 wt% and 50 wt%, were prepared by diluting concentrate nitric acid (70 wt%) by deionized water; aqua regia was obtained by mixing nitric acid (70 wt%) with hydrochloric acid (32 wt%) with the volume ratio of 1 to 3. Absorption solution (20 wt% sodium hydroxide solution) for the outlet gas, which corroded the brass connector, was made by dissolving 40- g pellet of sodium hydroxide into 160 ml of water.

3.2 Sample preparation

The project studied carbonylation of pure Ni and Fe, Ni-Fe mixture, Ni and Fe mixed with catalyst FeS and S, and reduced laterite ore.

Pure Ni and Fe were produced by reduction of metal oxides NiO and Fe2O3 by hydrogen. Nickel-iron mixture was obtained by mixing of Ni and Fe. Nickel oxide (NiO)

39 and iron oxide (Fe2O3) were reduced separately by hydrogen and then mixed in an agate mortar for about 20 minutes.

Iron sulphide was mixed with nickel and iron oxides in an agate mortar for about 15 minutes before reduction, degree of reduction of iron sulphide by hydrogen under experimental conditions is expected to be very low. Equilibrium constant for the reaction of iron sulphide with hydrogen, shown in Figure 3.1 is below 1.82*10-3 at 800 oC.

Sulphur has low melting and boiling temperatures, 115.21 oC and 444.6 oC respectively, therefore it could not be added to metal oxides before reduction. Sulphur was grinded and mixed with samples of metallic nickel or iron, or Ni-Fe mixture in an agate mortar for about 15 minutes.

68 -2

66 -4 64

62 -6 (kJ) o

G 60  Log(k) 58 -8

56 -10 54

52 -12 0 150 300 450 600 750 900 Temperature (oC)

Figure 3.1 The standard Gibbs free energy ∆Go and equilibrium constant (presented as log(k)) for the reaction of iron sulphide with hydrogen

Concentration of Ni and Fe in aqua regia was too high, more than 3000 mg·L-1, to be used in the ICP-OES analyses. In order to be analysed, the solution samples taken from the Dreschel bottles were diluted 10 to 50 times to 10 ml by 2 wt% aqua regia in a plastic tube. After carbonylation, the residual solution of aqua regia was diluted to 1000 ml in a volumetric flask and analysed by ICP-OES.

40 3.3 Experimental Set-up and Procedures

Stages in the experimental study of carbonylation of Ni and Fe are schematically shown in Figure 3.2. This section describes experimental set-ups and procedures for reduction and carbonylation processes.

Figure 3.2 Stages in the experimental study of carbonylation of Ni and Fe

3.3.1 Reduction Experiments Metal oxides were reduced by hydrogen in a horizontal furnace (Ceramic Engineering,

Sydney, Australia) at atmospheric pressure. In the reduction of laterite ore, the CO-CO2 gas mixture was used for selective reduction of metal oxides. Experimental setup is schematically shown in Figure 3.3.

Outlet

Flow meter Sample

Furnace

N2 H2

Figure 3.3 SchematicFig. experimental 1 Flow diagram set-up for for the reduction reduction of of Ni metal and oxides Fe oxides and laterite ore

The gas flow meter and thermocouple of the horizontal furnace were calibrated prior to the experiment study; the calibration results are shown in Appendixes A and B. Reduction was examined under conditions presented in Table 3-4.

41 Table 3-4 Parameters in the reduction process

Variable Unit Value

Temperature oC 400~900

Flow rate L·min-1 1

Content of CO in the CO- Vol% 60 CO2 gas mixture

6h in reduction of Ni and Fe Time hour oxides 4h in reduction of laterite ore

In the reduction, gas was supplied from gas cylinders of nitrogen, hydrogen, carbon monoxide and carbon dioxide. The gas flow rate was controlled by a needle valve and monitored by a gas flow meter. Oxides and ores were reduced in a horizontal furnace, which isothermal zone was determined in preliminary experiments by measuring the furnace temperature profile.

Samples in quartz boat were placed in the isothermal zone of the furnace, and heated in nitrogen (flow rate 0.5 L· min-1) at a rate of 5 oC · min-1. When the experimental temperature was reached, nitrogen was switched off and hydrogen or CO- CO2 gas mixture was introduced to the system. The reduction time was 6 hours in the reduction of Ni and Fe oxides, and 4 hours in the reduction of laterite ore. Samples were cooled to the room temperature in nitrogen.

3.3.2 Carbonylation Experiments Carbonylation experiments were conducted in a flexible U – shaped reactor (max pressure 68 atm) immersed into the oil bath. The schematic setup for carbonylation experiments is shown in Figure 3.4. Carbonylation was studied using CO at different pressures. The reaction temperature was controlled by the oil bath and monitored by three thermocouples placed in the centre of the reactor, outside the reactor and at the bottom of the oil bath. Calibration of the thermocouple for measurement of temperature

42 of the oil bath is shown in Appendix C. The outlet gas flow rate was adjusted by a needle valve and monitored by a gas flow meter. Pressure was regulated by clockwise rotating the adjustment screw/handle to increase the CO pressure to its desired value. Carbonyls were absorbed using two Dreschel bottles filled with aqua regia. The third absorption bottle filled with 20 wt% sodium hydroxide solution was used to treat the corrosive gas (Princeton 2014) from Dreschel bottles.

The carbonylation of metals was examined under conditions listed in Table 3-5.

Table 3-5 Parameters in carbonylation experiments

Variable Unit Value Temperature oC 80 ~ 100 Pressure atm 0 ~ 56

Gas flow rate L·min-1 0.1 ~ 0.5

Sample weight gram 0.8 ~ 3.2 Particle size μm 38 ~ 495

The flexible U-shape tube (internal layer: Teflon; outer layer: stainless steel) was purged by compressed air to remove residue powder and then by nitrogen to remove air prior to loading samples. Samples were inserted into the reactor, which was placed in the oil bath and connected to the gas system. A sample after reduction was exposed to air when it was moved to the carbonylation reactor. To reduce oxides which could be formed on the surface of reduced samples, they were heated to 150 oC in hydrogen for 2 hours. Cold glycerol (room temperature) was added into the oil bath to decrease the temperature to the experimental value for the carbonylation reaction. Afterwards, carbon monoxide was introduced into the reactor; CO pressure was raised to the experimental value at a specific flow rate. During the experiment, samples were taken from the solutions in Dreschel bottles at different times and diluted to 10 ml to be analysed by ICP-OES. When the experiment was finished, the CO gas was switched off

-1 and the gas flow rate of H2 was increased to 1L·min , hydrogen was introduced to the reactor to remove residue CO and carbonyls; the system was purged with hydrogen for

43 20 minutes at a flow rate of 1 L·min-1 at atmospheric pressure. Aqua regia in Dreschel bottles was diluted with ionised water to 1000±2 ml and examined by ICP-OES.

I II III

CO

V b

a c IX VI N2 IV

VII Outlet e d

a: Oil Bath b: Stirrer c: Flow Meter d: Absorption Bottles e: Absorption bottle I ~ VIII: Valves I: CO Gas Cylinder Valve II: CO Gas Cylinder Valve Regulator III: CO Break Valve IV: N2 Break Valve V: Three Way Valve VI: Needle Valve VII: Three Way Valve

Figure 3.4 Schematic setup for carbonylation experiment

3.4 Analytical Equipment and Methodology

3.4.1 X-Ray Diffraction (XRD) Analysis The XRD analysis was conducted using the PANalytical Xpert Multipurpose X-ray Diffraction System (MPD), which is a theta to theta goniometer system used for powder samples.

The MPD system is equipped with the following components:

1) Incident beam path–Programmable divergence slit, which can be configured as a fixed slit or a variable slit, masks.

2) Diffracted beam path–Programmable anti-scattering slit and receiving slit, curved monochromators and proportional detector.

3) Automatic multi-sample changer (15 samples) can increase productivity and a sample spinner is to eliminate the preferred orientation.

44 The sample analysed by MPD needed to be grounded into powder of size less than 10μm in order to obtain a random orientation of material. The powdered sample was closely packed into the sample holder (10mm diameter) with flat surface, and then loaded into an automatic sample changer for XRD analysis.

The XRD analysis in this project was conducted with the following parameters:

 X-ray source: copper Kα radiation with λ=1.54 Å

 Voltage=45 kV; Current=40 mA

 Scan range (°): 10-100

 Scan step (°): 0.0131303

 Time per step (s): 25.50

 Net time per step (s): 23.97

 Scan speed (°/s): 0.131303

 No. of steps: 6855

 Total scanning time: 11 min 53 sec

 Divergence slit (°): 0.5

 Diffraction beam (°): 1

3.4.2 ICP-OES/MS Analysis An Inductively Coupled Plasma (ICP) is generated by heating a flow of argon gas through a torch fitted with a radio frequency induction coil. A dissolved sample is nebulised and fed into the plasma (at up to 10,000 K) where the elements in the sample ionise and emit characteristic radiation. The quantity of elements, Ni and Fe, can be simultaneously measured by detecting the characteristic emission of the element (ICP- AES).

Inductively Coupled Plasma - Optical Emission Spectroscopy (ICP-OES) (also known as ICP-AES Inductively Coupled Plasma - Atomic Emission Spectroscopy) is one of the most commonly used methods for the determination of trace concentrations of elements in samples. The detection limits are generally in the ppb range. The upper limit for a particular emission line is usually 104 to 106 times the detection limits. The precision of

45 analysis is usually in the 1-2% RSD (Relative Standard Deviation) range. Better precision can be obtained with trade-offs in speed. It is capable of multi-element, qualitative and quantitative analyses.

During the analysis, a work line was calibrated from standard solution in the range from

10-2 to 102 mg·L-1. Concentration of elements in the analysed sample was calculated based on the work line.

The operation parameters are listed in Table 3-6.

Table 3-6 ICP-OES analysis parameters

Nebulizer Auxiliary Plasma RF power Sample pump flow Plasma Nebulizer flow flow flow viewing W L·min-1 L·min-1 L·min-1 mL·min-1 Ar 1300 0.8 0.2 15 0.50 Axial

Background Processing Auto Read delay Rinse Spray Replicates correction mode integration chamber s s points Area Area 30 15 3 7 3

3.4.3 Scanning Electron Microscopy (SEM)/Energy Dispersive X-ray Spectrometer (EDS) The morphology of samples was examined by the HITACHI S3400 Field Emission Scanning Electron Microscope (FESEM). This instrument is fitted with secondary and backscatter electron detectors that allow for topographic and compositional (atomic number contrast) surface imaging of samples. It is used for imaging up to 3000X magnification. Elements in the samples can be identified by the linked X-ray Energy Dispersive Spectrometer (EDS).

Because the metal samples were conductive, they were adhered directly to the carbon tape and analysed. The samples of laterites were coated with gold and then tested.

3.4.4 X-Ray Fluorescence (XRF) Spectroscopy X-ray fluorescence (XRF) spectroscopy uses an X-ray source to eject core-shell electrons from an atom to create an excited state. The resulting cascade of electrons to 46 fill the holes results in emission of X-radiation from the atom (fluorescence) that has a characteristic wavelength/energy specific to each element. The fluorescence can be quantified to enable elemental analysis from the ppm level.

In this project, laterite ores samples were grinded to fine particles (< 300 mesh) and then dried at 200 oC for 24 hour; after the preparation, samples were submitted to the XRF Lab of the UNSW Analytical Centre for analysis.

3.5 Selection of the absorption reagent

The extent of nickel and iron carbonylation was calculated from the concentration of these metals in the absorbent measured by ICP – OES. Metal carbonyls are readily absorbed by an oxidative acid, such as concentrated sulphuric acid and nitric acids (Mond et al. 1891, Mond et al. 1891-1). Sulphuric acid was excluded in this project because sulphur containing catalysts were used in the carbonylation process. The alternative absorbents tested in this project were concentrated nitric acid and aqua regia. Results of the test presented in Appendix D, show that iron pentacarbonyl was completely absorbed by the concentrated nitric acid (50% wt) and aqua regia, while the absorption of nickel tetracarbonyl was more difficult. Two Dreschel bottles with aqua regia were needed for absorption of nickel carbonyl. To confirm this conclusion, the extent of nickel carbonylation was calculated from the weight loss and ICP – OES analysis. Results compared in Figure 3.5 demonstrate good agreement between two methods, what confirms that aqua regia is an appropriate absorbent for both, nickel and iron carbonyls.

47 100

80

60

Recovery (%) 40

20 From ICP-OES From weight loss

0 80 84 88 92 96 100 Temperature (oC)

Figure 3.5 Extent of nickel carbonylation measured by the weight loss and by the ICP- OES analysis of nickel in aqua regia

3.6 Data Analysis

3.6.1 Definition of extent of carbonylation and the content of catalyst Extent of metal carbonylation, y, is defined as the extent of metal converted to metal carbonyls (percentage)t, it was calculated using equation (3–1).

mass of converted metal y  * 100 (3–1) initial mass of metal

Another important factor in the catalytic carbonylation is the content of catalysts, it is defined as follows.

mass of sulfur element in catalyst Content = * 100 (3–2) mass of a sample

3.6.2 Extent of metal carbonylation As described, metal carbonyls were absorbed in two tandom Dreschel bottles, and about 7 ml of solution were taken from each bottle (I 125 ml, II 250 ml) during the experiment.

48 Metal in the absorption bottles and sample tubes taken out for ICP-OES were calculated, and the latter one should be accumulated. Because samples used in ICP-OES were all diluted to 10 ml, hence a dilution factor should be considered to get the concentration of metal in Dreschel bottles.

The extent of metal carbonylation was calculated from results of the ICP-OES analysis as follows:

i

(**ciI,,,,,, x iI V iI c iII **) x iII V iII  ( c,I  c ,II )*10  0 yMi,  *100 (3–3) mM

Where, (****)ciI,,,,,, x iI V iI c iII x iII V iII is the mass of metal in aqua regia in Dreschel

i th bottles when the i sample is taken; and (CC,I  ,II )*10 is the mass of metal in  0 sample tubes in ICP-OES.

M represents Ni, Fe or Co, mM is the initial mass of metal M, g, i is the serial number of samples taken at different times, 0 ~ 6,

I, II represent the tandem absorption bottles,

th -1 ci is the concentration of metal M in i sample for ICP-OES, g·ml , xi is the dilution factor in analysis of sample i, 10 ~ 50,

Vi is the volume of aqua regia in absorption bottles when sample i is taken, ml

th ci,,, I** x i I V i I is the mass of metal M absorbed in the first bottle when the i sample was taken, g

th ci,,, II** x i II V i II is the mass of metal M absorbed in the second bottle when the i sample was taken, g

49 i

 c ,I 10 is the mass of metal M in ICP-OES samples from bottle I  0

th accumulated to the i sample (  =0 means the sample was taken before carbon monoxide was introduced, actually this concentration was detected to be 0 or minus near 0); 10 is the volume of solution after dilution in sample tube, g

i

c ,II 10 is the mass of metal M in ICP-OES samples from bottle II  0 accumulated to the ith sample ( =0 means the sample was taken before carbon monoxide was introduced, actually this concentration was detected to be 0 or minus near 0); 10 is the volume of solution after dilution in sample tube, g yM,i is the extant of carbonylation of metal M when sample i is collected, %.

Degree of reduction of metal oxides:

Extent of reduction of nickel and iron oxides in the thesis was defined as the weight loss of a sample during the reduction process divided by the oxygen mass in Ni or Fe oxide:

푊푒𝑖푔ℎ푡 푙표푠푠 표푓 표푥𝑖푑푒푠 푠푎푚푝푙푒 퐷푒푔푟푒푒 표푓 푟푒푑푢푐푡푖표푛 = ∗ 100 퐶표푛푡푒푛푡 표푓 표푥푦푔푒푛 𝑖푛 표푥𝑖푑푒푠 푠푎푚푝푙푒 (3–4)

50 Chapter 4 Thermodynamic Analysis of Carbonylation

Carbonylation of nickel and iron can be presented by reactions (4-1) and (4-2):

Ni(s) + 4 CO(g) = Ni(CO)4(g) (4-1)

Fe(s) + 5 CO(g) = Fe(CO)5(g) (4-2)

Carbonylation reactions were analysed in Chapter 2. This chapter presents the thermodynamic analysis of above reactions to assess conditions for carbonylation of selectively reduced laterite ore.

The standard Gibbs free energies of reactions (4-1) and (4-2) were calculated by equations (4-3) and (4-4) using data from the HSC Chemistry database:

o o o o ∆G 4-1 = G Ni(CO)4 - 4G CO -G Ni (4-3)

o o o o ∆G 4-2 = G Fe(CO)5 - 5G CO -G Fe (4-4)

Calculated standard Gibbs free energies and equilibrium constants (log (K)) for reactions (4-1) and (4-2) are presented in Figure 4.1. Both the standard Gibbs free energies increase with temperature. The standard Gibbs free energy for nickel carbonylation is negative below 110 oC; while it is positive for iron carbonylation in the examined temperature range. This means that only formation of nickel carbonyl from CO and Ni is spontaneous under standard conditions; the equilibrium constant is high for the reaction of nickel carbonyl formation and low for the reaction of formation of iron carbonyl.

51 40 8

30 Ni + 4CO(g) = Ni(CO) (g) 4 6 Fe + 5CO(g) = Fe(CO)5(g) 20 4

10 2

(KJ) 0 0  G r  -10 -2 Log(K)

-4 -20 -6 -30 Ni + 4CO(g) = Ni(CO) (g) 4 -8 Fe + 5CO(g) = Fe(CO) (g) -40 5 40 50 60 70 80 90 100 110 Temperature (oC) Figure 4.1 The standard Gibbs free energies ∆rGo and equilibrium constants (presented as log(K)) for reactions of carbonylation of nickel and iron

Equilibrium constant for the reaction of nickel carbonyl formation decreases from 2.57*105 at 40 oC to 3.31 at 110 oC; the equilibrium constant for the reaction of iron carbonyl formation varies from 0.047 at 40 oC to 3.16 * 10-7 at 110 oC.

Conversions of nickel and iron to carbonyls by reactions (4-1) and (4-2) in a stoichiometric system were further calculated to examine the effects of temperature and CO pressure. Conversion was defined as a fraction of Ni or Fe in the form of carbonyl in the gas phase. Molar fractions of Ni carbonyl in the Ni(CO)4(g) –CO gas formed in reaction (4-1) and Fe carbonyl in the Fe(CO)5(g) –CO gas formed in reaction (4-2) were calculated as follows :

XNi(CO)4= 4PNi(CO)4/(4PNi(CO)4+PCO) (4-5)

XFe(CO)5= 5PFe(CO)5/(5PFe(CO)5+PCO) (4-6)

4 P Ni(CO)4 = K4-1*PCO (4-7)

5 P Fe(CO)5 = K4-2*PCO (4-8) where K4-1 and K4-2 are equilibrium constants for reactions (4-1) and (4-2).

52 Calculated conversions of Ni and Fe to carbonyls at different temperatures and CO pressures are presented in Figure 4.2. Both conversions increase with CO pressure and decrease with the increase of temperature. The carbonylation of iron (Figure 4.2 (b)), is more sensitive to reaction conditions. More than 90% of nickel is converted to nickel tetracarbonyl below 150 oC when the CO pressure is over 5 atm; at temperatures below 110oC, conversion of nickel to carbonyl is thermodynamically feasible at CO pressure below 1 atm. However, only about 10% of iron is converted to carbonyl at 100 oC when CO pressure is 10 atm.

150 0.10 150 0.0 ( a. ) ( b ) 0.19 0.10 135 135 0.28 0.20 120 0.37 0.30

120 ) ) ℃ ℃ 0.46 105 0.40 105 0.55 0.50 90 90 0.64 0.60 Temperature ( Temperature ( 75 0.73 0.70 75 0.82 60 0.80

60 0.91 0.90 Conversion (1.0 for 100%, 0.1 for 10%) 0.1 for 100%, Conversion (1.0 for

Conversion (1.0 for 100%, 0.1 for 10%) 0.1 for 100%, Conversion (1.0 for 45 1.0 1.0 0 10 20 30 40 50 60 10 20 30 40 50 60 CO partial pressure (atm) CO partial pressure (atm)

Figure 4.2 Conversions of (a) nickel and (b) iron at different temperatures and CO pressures

4.1 Effect of CO Pressure

Effect of CO pressure on conversion of nickel and iron to carbonyls is shown in Figures 4.3 and 4.4. More than 95% nickel is converted into carbonyls at temperatures below 105oC under CO pressure below 20 atm.

53 1

0.9

0.8 1

o 0.7 45 C 0.98 65 oC o 0.96

Conversion 85 C o 0.6 105 C 0.94 125 oC 0.92 145 oC 165 oC 0.9 0.5 0 20 40 60 80 0 15 30 45 60 75 Pressure (atm)

Figure 4.3 Effect of CO pressure on nickel conversion to Ni(CO)4

Conversion of iron under the same conditions is much lower, it is less than 50% at 100oC under CO pressure of 20 atm

1.0

0.8 50oC 100oC o 0.6 150 C 200oC 250oC 0.4 Conversion

0.2

0.0 0 10 20 30 40 50 60 70 80 Pressure (atm)

Figure 4.4 Effect of pressure on iron conversion to Fe(CO)5

54 4.2 Effect of Temperature

Figures 4.5 and 4.6 show the effect of temperature on conversions of nickel and iron to carbonyls. Conversions of both Ni and Fe decrease with the increasing temperature.

1.0

0.8 1 atm 0.6 5 atm 10 atm 20 atm 30 atm Conversion 0.4 40 atm 50 atm 0.2 60 atm 70 atm 80 atm 0.0 40 80 120 160 200 240 280 320 Temperature (oC)

Figure 4.5 Effect of temperature on nickel conversion to Ni(CO)4

Conversion of iron to Fe(CO)5 drops much faster with the increase of temperature; slightly over 80% of iron carbonyl can be obtained at 100oC at CO pressure 34 atm.

1.0 1.00 0.98 0.96 0.8 0.94 0.92

0.90 20 40 60 80 100 0.6 1 atm 7 atm

16 atm 0.4 25 atm

Conversion 34 atm 40 atm 50 atm 0.2 60 atm 70 atm

0.0 25 50 75 100 125 150 175 200 Temperature (oC)

Figure 4.6 Effect of temperature on iron conversion to Fe(CO)5 55 It can be concluded from the thermodynamic analysis that high extent of nickel carbonylation can be achieved at temperatures below 100 oC and CO pressure above 20 atm; carbonylation of iron at 100 oC requires higher pressure, above 30 atm. Kinetics of carbonylation reactions is affected by the thermodynamic driving force, which is much stronger in carbonylation of nickel in comparison with iron. The project will focus on carbonylation of Ni and Fe and selectively reduced laterite ore at 80-100 oC and CO pressure up to 55 atm.

56 Chapter 5 Experimental Results

As summarized in Chapter 2, the reaction rate and yield of nickel, iron and cobalt carbonylation, are affected by reaction temperature, carbon monoxide pressure, gas flow rate, particle size, and catalysts. This chapter focuses on the study of the effects of these parameters on the carbonylation of nickel, iron, nickel-iron mixture and laterite ores. The reaction of cobalt with carbon monoxide was also investigated to find out the extent of cobalt carbonylation under experimental conditions, which is expected to be low. The carbonylation process was proposed to be used in the application to reduced laterites, therefore, the project studied carbonylation of metals which were also reduced from associated oxides. The structure of this chapter is as follows: (1) The reduction of iron, nickel and cobalt oxides at different temperatures and various reaction times; (2) non- catalytic carbonylation of pure metals (nickel, cobalt and iron) and nickel-iron mixture; (3) catalytic carbonylation of nickel; and (4) non-catalytic and catalytic carbonylation of laterite ores (BCS laterite ore and MIN laterite ore).

5.1 Reduction of Iron, Nickel and Cobalt Oxides

In the reduction process, iron oxide (Fe2O3) was reduced by hydrogen at temperatures from 300 to 500 oC. The flow rate of hydrogen was 1.0 L·min-1. Degree of reduction of iron oxide measured by weight loss as a result of oxygen removal in the reduction experiments is listed in Table 5–1.

A reduction degree of 97.5% was achieved in a 6-hour experiment at 400 oC. Figure 5.1 shows the XRD patterns of samples reduced at 400 oC for 1, 4 and 6 hours. Iron oxide was not identified in the XRD pattern of the sample after 6-h reduction; hence it can be concluded that reduction of Fe2O3 to metallic iron was close to completion. The degree of reduction of iron oxides was enhanced to 98.8% after a 6-hour reaction at 500 oC.

57 Table 5-1 Degree of reduction of Fe2O3 by hydrogen

Degree of reduction of Temperature Time Fe2O3 oC Hour %

1 54.8

300 4 55.4

4 67.7 400 6 97.5

1 69.5 500 6 98.8

Fe 16384 Fe2O3 Fe3O4

8192

4096 6 hour 7500

6000 unt(cts)

Co 4500 4 hour 8000

6000

4000 1 hour 2000 20 30 40 50 60 70 80 Position (2)

o Figure 5.1 XRD patterns of Fe2O3 samples reduced by hydrogen at 400 C for 1, 4 and 6 hours

Based on the results obtained in the reduction of iron oxide, the reduction of nickel oxide was studied from 400 oC. Reduction of NiO was incomplete after 6-hour reaction at this temperature, therefore, the reduction temperature was increased to 500 oC. The reduction degree of NiO by hydrogen at 400 and 500 oC is presented in Table 5-2. The

58 degree of reduction of NiO after 6 hours reaction at 500 oC reached 97.4 %. The XRD analysis shown in Figure 5.2, did not detect nickel oxide residue in the sample.

Table 5-2 Degree of reduction of NiO by hydrogen at 400 and 500 oC

Temperature Time Degree of reduction of NiO

oC Hour %

4 86.1 400 6 91.8

500 6 97.4

2E5 NiO

1E5

5E4 NiO

3E5 t (cts)

2E5 Coun Ni

2E5

8E4 6 hour

20 30 40 50 60 70 80 90 Position (2) Figure 5.2 XRD patterns of NiO samples reduced by hydrogen at 500 oC

Reduction of Fe2O3 – NiO mixture by hydrogen was also tested. Nickel oxide and iron oxide were mixed in agate mortar for 30 minutes before reduction and then reacted with

59 hydrogen at 500oC for 4 and 6 hours. Degree of reduction calculated from the weight loss is given in Table 5-3.

o Table 5-3 Degree of reduction of Fe2O3 – NiO mixture at 500 C

Temperature Time Degree of reduction

oC Hour %

4 97.6 500 6 97.5

XRD analysis presented in Figure 5.3, confirmed that 6-hour reduction of oxide mixture at 500 oC was close to completion; nickel and iron oxides were not detected by the XRD analysis of the Fe2O3 – NiO mixture after reduction. Reduced samples contained nickel, iron and ferronickel. Some sintering of the metallic powder was observed after the reduction process.

NiO Fe2O3 10000 Ni Fe NiFe

8000

Reduced sample

6000 unt (cts)

Co 4000

2000 Raw mixture

0 20 30 40 50 60 70 80 90 Position (2)

Figure 5.3 XRD patterns of the NiO-Fe2O3 mixture before and after reduction by hydrogen at 500 oC for 6 hours

o Cobalt oxide (Co3O4) also was reduced by hydrogen at 500 C for 6 hours; the degree of reduction was 99.6%, indicating that reduction of cobalt oxide was close to completion.

60 The XRD patterns of Co3O4 and reduced sample are shown in Figure 5.4. Only metallic cobalt was detected in the reduced sample.

Co O 16384 3 4 Co

8192

Reduced Sample

4096 Count (cts)

2048 Raw oxide

1024 20 40 60 80 100 Position (2)

o Figure 5.4 XRD patterns of Co3O4 sample reduced by hydrogen at 500 C for 6 hours

The experimental results demonstrated that 6-hour reduction at 500 oC was sufficient for the reduction of oxides and oxides mixture. Therefore, all the metal samples (apart from samples in the study of the effect of particle size on the reduction of nickel oxide and laterite samples) were reduced by hydrogen at 500 oC for 6 hours.

5.2 Non-catalytic Carbonylation of Metal

This section presents results of a comprehensive study of the effect of reaction time, gas flow rate, reaction temperature, carbon monoxide pressure, particle size and other parameters on the non-catalytic carbonylation of nickel, iron, cobalt and Ni-Fe mixture. The reaction parameters are listed in Table 5-4.

61 Table 5-4 Parameters in a study of non-catalytic carbonylation of Ni, Fe, Co and Ni-Fe mixture

Ni-Fe Parameter/Metal Nickel Iron and Cobalt mixture Carbonylation 80, 90, 100 80, 90, 100 100 Temperature, oC 1, 28 55 for iron Gauge Pressure, atm 14, 27, 41, 55 27 44, 55 for cobalt

Gas Flow rate, L∙min-1 0.14, 0.23, 0.36, 0.50 0.50 0.50

Reduction temperature, 500, 600,700, 800,900 500 500 oC

Ni Sample mass, g 0.803, 1.52, 3.36 ~ 1.3 ~ 1.3

Nickel was also carbonylated at 100 oC and atmospheric pressure to evaluate the effect of high temperature and pressure.

5.2.1 Non-Catalytic Carbonylation of Pure Nickel The reported in literature reaction time needed for carbonylation of nickel and ferronickel alloy varied from seconds to several days (Liang et al. 1983, Herrmann 1990, Wang et al. 2009). The reaction time needed for carbonylation of metals under experimental conditions of this project (temperature range 80-100 oC) was assessed in a study of carbonylation of two samples of nickel-iron mixture (50 wt% Ni).

5.2.1.1 Assessment of Reaction Time for Carbonylation

Carbonylation of Ni-Fe mixture (mass 2 g; 50 wt% Ni) was examined at 80 oC (the lowest temperature in the experimental temperature range) at CO gauge pressures of 14 and 54 atm, with CO flow rate at 0.5 L•min-1. The Ni-Fe mixture was obtained by reduction of mixed nickel and iron oxides; the reduced samples were grinded in an agate mortar before being transferred to the reactor.

62 Carbonylation curves for nickel and iron from Ni-Fe mixtures are plotted in Figure 5.5. The rate of carbonylation of both, Ni and Fe increased with increasing CO pressure. Carbonylation of nickel had a much higher rate than that of iron, which is consistent with the literature (Mond et al. 1890, Mond et al. 1891).

The progress of carbonylation reaction became slow after 6 hours. The extent of carbonylation of Ni in experiments at 14 atm CO gauge pressure practically did not change after 6 hours of reaction; increase in the reaction time from 6 to 10 hours increased the extent of formation of nickel tetra-carbonyl at CO gauge pressure 54 atm from 50.1 to 55.9 %, Therefore, 6-hour reaction time was sufficient to study effects of temperature, CO pressure and other parameters on carbonylation of Ni, iron and their mixture.

14 atm 60 54 atm

50

40

30

20

10 Extent of Ni carbonylation (%)

0 0 2 4 6 8 10 Time (hour)

14 atm 60 54 atm

50

40

30

20

ent of iron carbonylation (%) 10 Ext 0 0 2 4 6 8 10 Time (hour)

63 Figure 5.5 Carbonylation of the Ni-Fe mixture (50 wt% Ni) at 80 oC and CO gauge pressures 14 and 54 atm

5.2.1.2 Effect of Gas Flow Rate

The effect of the gas flow rate on nickel carbonylation was studied at 100 oC and CO gauge pressure of 27 atm. Gas flow rate was changed from 0.14 to 0.50 L∙min-1; carbonylation curves are shown in Figures 5.6 and 5.7. The gas flow rate had a significant effect on the Ni carbonylation. The rate of carbonylation was constant when the gas flow rate was 0.14 L·min-1. Increase in the gas flow rate enhanced the reaction rate; carbonylation was close to completion in about 6 hours at the gas flow rate 0.5 L· min-1.

100

80

60

40

0.50 Lmin-1 -1 Extentcarbonylation Ni of (%) 20 0.35 Lmin -1

0.23 L? min 0.14 Lmin-1 0 0 123456 Time (Hour) Figure 5.6 Nickel carbonylation at 100 oC and CO gauge pressure 27 atm at different gas flow rates

Figure 5.7 shows the effect of gas flow rate on nickel carbonylation after 6-hours reaction. The carbonylation of pure nickel was greatly influenced by the gas flow rate; but the effect decreased with the increase of the CO flow rate. About 99% of nickel was converted to carbonyl in 6.0 hours when the gas flow rate was 0.5 L•min-1. A study of

64 the effects of other parameters on carbonylation of metals was operated at CO flow rate of 0.5 L•min-1.

100

80

60

40 tent of Ni carbonylation (%) 20 Ex

0 0.0 0.1 0.2 0.3 0.4 0.5 Gas flow rate (Lmin-1)

Figure 5.7 Effect of CO flow rate on nickel carbonylation at 100 oC and CO gauge pressure 27 atm

5.2.1.3 Effect of Temperature

Temperature is a significant factor in the nickel carbonylation (Abel 1963). The effect of temperature on carbonylation of nickel was examined in the range of 80 - 100 oC at CO gauge pressure 27 atm, the gas flow rate was 0.5 L•min-1.

Nickel carbonylation curves at 80, 90 and 100 oC are plotted in Figure 5.8. These plots include experimental data obtained by sampling the absorption solution (aqua regia, Chapter 3) at different times. These plots also include “ultimate” points, which were obtained after the completion of the experiment, and purging the reactor with hydrogen to remove Ni(CO)4 left in the reactor. The reaction system had a “dead” volume, which was estimated at about 0.2 L; as a result, a small fraction of Ni(CO)4 was trapped in the system.

65 100 80 oC 80 ultimate 90 oC 80 90 ultimate 100 oC 100 ultimate

60

40 Extent of Ni carbonylation (%) carbonylation of Ni Extent 20

0 0 1 2 3 4 5 6 7 Time (hour)

Figure 5.8 Effect of temperature on the carbonylation of nickel at CO gauge pressure 27 atm

The extents of nickel carbonylation after 6.5 hours reaction (“ultimate” points) at CO gauge pressure 0, 14 and 27 atm are presented in Figure 5.9. The extent of nickel carbonylation increased with temperature at constant CO pressure. Carbonylation of nickel at 100 oC and CO pressure 27 atm after 6.5 hours reaction was close to completion (extent of carbonylation was 99%).

100 0 atm 14 atm 80 27 atm

60

40

20 Extent of Ni carbonylation (%)

0 80 85 90 95 100 Temperature (oC) Figure 5.9 Effect of temperature on nickel carbonylation at CO gauge pressure 0, 14 and 27 atm after 6.5 hours reaction

66 5.2.1.4 Effect of Carbon Monoxide Pressure

The effect of CO pressure on nickel carbonylation was tested in experiments at 80-100 oC with carbon monoxide pressure (gauge pressure) up to 54 atm at gas flow rate 0.5 L·min-1.

Carbonylation curves are plotted in Figure 5.10. Figure 5.11 shows the effect of CO partial pressure on the extents of nickel carbonylation at 80, 90 and 100 oC after 6.5 hours reaction. Effect of CO pressure on nickel carbonylation depended on temperature. At all temperatures, the effect of CO pressure was very strong when the gauge pressure was increased from 0 to 14 atm, and weak when CO gauge pressure changed from 27 to 54 atm (within the experimental error, which was estimated to be 4.5%). The increase in CO pressure from 14 to 27 atm had a strong effect on the rate and extent of carbonylation at 80 and 100 oC and weaker at 90 oC.

The extents of Ni carbonylation at 80 and 90 oC were 47% and 75% at pressures over 27 atm after 6.5 hours reaction. A significant increase in the extent of nickel carbonylation to over 99% was observed when the temperature was increased to 100 oC. The extent of Ni carbonylation at 100 oC at CO pressure 14 atm after 6.5 hours reaction was 82.3%, which is higher than results at 80 and 90 oC at CO pressure above 27 atm.

100 100 80 oC 90 oC 80 80

60 60

40 40

20 20 Extent of(%) NiExtent carbonylation of(%) NiExtent carbonylation 0 0 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 Time (hour) Time (hour)

100 100 oC

80 00 atm 60 14 atm

40 27 atm 43 atm 20 54 atm Extent of(%) NiExtent carbonylation 0 0 1 2 3 4 5 6 7 Time (hour)

67 Figure 5.10 Effect of CO pressure (gauge) on the carbonylation of nickel at 80, 90 and 100 oC.

100

90

80

70

60

50

40

30

20 80 oC Extent of nickel carbonylation (%) 90 oC 10 100 oC 0 0 5 10 15 20 25 30 35 40 45 50 55 CO gauge Pressure (atm) Figure 5.11 Extents of Ni carbonylation at 80, 90 and 100 oC at different CO gauge pressures after 6.5 hours reaction

5.2.1.5 Effect of Particle Size

Particle size of metallic nickel formed in the reduction of NiO depended on the reduction temperature. The particle size of metallic nickel subjected to carbonylation was adjusted in the reduction experiment by the change of the reduction temperature from 500 to 900 oC in 6 hour reduction.

Figure 5.12 shows the SEM images of nickel samples reduced at temperatures from 500 to 900 oC. Sintering at the elevated reduction temperatures coarsened the nickel particles; particle size increased with increasing reduction temperature. The mean particle sizes of samples produced at different temperatures are listed in Table 5-5; the statistical data are presented in Appendix E. The mean particle sizes of nickel were 0.29, 0.51, 1.10, 2.07 and 2.67 μm in samples reduced at 500, 600, 700, 800 and 900 oC respectively.

68 Figure 5.12 Effect of reduction temperature on nickel particle size

Effect of the nickel particle size on carbonylation was studied at 100 oC and CO gauge pressure of 27 atm, the gas flow rate was 0.5 L·min-1.

Table 5-5 Particle size of Ni reduced at different temperatures

Temperature, Number of counted Particle size oC particles Max. Min. Mean μm

500 61 0.46 0.17 0.29

600 86 0.89 0.02 0.51

700 86 2.22 0.61 1.10

800 63 3.58 1.05 2.07

69 900 77 4.35 1.45 2.67

The carbonylation curves obtained in experiments with nickel of different particle size are plotted in Figure 5.13. Nickel particles produced in the reduction at 500 and 600 oC were of sub-micron size, what made the nickel powder highly reactive. It can be suggested that after reduction, these particles were partially re-oxidised (from the surface), what retarded the carbonylation reaction. Figure 5.14 shows the effect of the reduction temperature (particle size) on the extents of nickel carbonylation after 6.5 hours reaction. Experimental data obtained in the first 3-hour carbonylation are rather scattered. However, they are consistent after 4 hours of reaction: the extent of carbonylation decreased with increasing particle size (Figure 5.14).

100

80

60

40 0.29 m 0.51 m 20 1.10 m

Extent of nickelExtent carbonylation (%) 2.07 m 2.67 m

0 0 1 2 3 4 5 6 Time (Hours)

Figure 5.13 Effect of particle size on the carbonylation of nickel at 100 oC and CO gauge pressure of 27 atm.

The effect of the particle size on the nickel carbonylation was strong when the mean particle size was above 1.1 μm (the reduction temperature increased over 700 oC).

70 100 o 500 C o o 700 C 600 C 800 oC 80

900 oC 60

40

20 Extent of nickel carbonylation (%) of nickel carbonylation Extent

0 0.5 1.0 1.5 2.0 2.5 3.0 Particle size (m)

Figure 5.14 Effect of particle size (reduction temperature) on the carbonylation of nickel at 100 oC and CO gauge pressure 27 atm after 4 hours reaction

5.2.1.6 Effect of Nickel Mass The effect of a sample mass on the nickel carbonylation was tested at 100 oC under CO gauge pressure 27 atm. The mass of nickel samples was 0.803, 1.52 and 3.36 g.

Carbonylation curves obtained in experiments with samples of different mass are shown in Figure 5.15. Effect of the sample mass on the nickel carbonylation was quite minor; the nickel carbonylation after 4-hour reaction was close to completion in all experiments.

71 100

80

60

40 0.803g 1.52g 20 3.36g Extent of Ni carbonylation (%) carbonylation Ni of Extent

0 0 1 2 3 4 5 6 Time (Hour)

Figure 5.15 Carbonylation of nickel of different mass at 100 oC and CO gauge pressure of 27 atm.

5.2.1.7 Effect of Iron on Nickel Carbonylation

The mixture of Ni and Fe was obtained by mixing nickel and iron reduced from Ni and iron oxides at 500.oC by hydrogen. Carbonylation of nickel in the Ni-Fe mixtures of different compositions was examined at 100 oC and CO gauge pressure of 27 atm. The proportion of iron in the nickel-iron mixture was 11, 20, 50, 80 and 89 wt%, the sample mass was 1.7 g.

The carbonylation curves are shown in Figure 5.16. The extents of the carbonylation reactions did not change after 3 hours for samples containing over 20 wt% of Fe, and within 4 hours for the sample with 11 wt% of Fe.

72 100 11% Fe 90 20% Fe 80 Pure nickel 50% Fe 80% Fe 70 89% Fe 60 50 40 30 20 10 Extent of Ni carbonylation (%) 0 0 1 2 3 4 5 6 7 Time (hour) Figure 5.16 Carbonylation of nickel from the Ni-Fe mixtures of different compositions at 100 oC and CO gauge pressure 27 atm on mixture

The extent of Ni carbonylation from mixtures with different iron contents at 6.0 hours reaction is shown in Figure 5.17; the extent carbonylation of pure nickel isalso drawn in the figure. The extent of nickel carbonylation decreased with the increase of iron content. The extent of nickel carbonylation dropped sharply from over 98% to 55.4% when 11 wt% of iron was added to nickel; it declined further to only 6.2% when iron content in the mixture increased to 89 wt%.

100 Pure Ni Ni

80

60

40

Extents of Extents Ni (%) carbonylation 20

0 0 20 40 60 80 100 Content of Fe in Ni-Fe mixture (%)

Figure 5.17 Carbonylation of nickel from the Ni-Fe mixtures of different compositions at 100 oC and CO gauge pressure 27 atm after 6.0 hour reaction.

73 5.2.2 Non-Catalytic Carbonylation of Pure Iron

The carbonylation of pure iron was studied at 80, 90 and 100 oC at CO gauge pressure of 0, 27 and 54 atm. The gas flow rate was 0.5 L·min-1, sample mass 1.3 g.

Carbonylation curves obtained in experiments at 80-100 oC and CO gauge pressure 0, 27 and 54 atm are shown in Figure 5.18.

10 10 0 atm 27 atm

8 8

6 6

4 4

2 2

Extent of Fe carbonylation (%) 0 Extent of Fe carbonylation (%) 0 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 Time (hour) Time (hour) 10 54 atm o 8 80 C 90 oC o 6 100 C

4

2

Extent of Fe carbonylation (%) 0 0 1 2 3 4 5 6 7 Time (hour) Figure 5.18 Carbonylation of iron at 80-100 oC and CO gauge pressure 0, 27 and 54 atm

Extents of Fe carbonylation after 6.5-hour reaction are given in Figure 5.19. The rate and extents of iron carbonylation were much lower in comparison with carbonylation of pure nickel. The extent of iron carbonylation under atmospheric pressure increased only from 1.5% at 80 oC to 3.3% at 100 oC. The extent of carbonylation of pure iron at 80 oC increased with increasing CO pressure; however, the effect of temperature on the carbonylation of iron at CO gauge pressure 27 and 53 atm was insignificant. The extent of iron carbonylation at 100 oC under pressures 27 and 53 atm was 3.2-3.7%.

74 10 00 atm 27 atm 54 atm 8

6

4

2 Extent of iron carbonylation (%) of iron Extent

0 80 85 90 95 100 Temperature (oC)

Figure 5.19 Carbonylation of iron at 80-100 oC and CO gauge pressure 0, 27 and 54 atm after 6.5-hour reaction

The effect of CO pressure on the iron carbonylation at 80-100 oC is shown in Figures 5.20 and 5.21. The rate and extent of the iron carbonylation increased with increasing CO pressure at 80 and 90 oC; however, increase in CO pressure had no effect on the carbonylation of iron at 100 oC under experimental conditions.

10 10 o 80 C 90 oC

8 8

6 6

4 4

2 2 xtent of Fe carbonylation (%) Extent of Fe carbonylation (%) E 0 0 0 1 2 3 4 5 6 0 1 2 3 4 5 6 Time (hour) Time (hour) 10 100 oC

8 00 atm 27 atm 6 54 atm

4

2

Extent of Fe carbonylation (%) 0 0 1 2 3 4 5 6 Time (hour)

Figure 5.20 Effect of CO gauge pressure on carbonylation of iron at 80-100 oC.

75 20 o 18 80 C 90oC 16 100oC 14 12 10 8 6 4 Extent of Fe carbonylation (%) Extent 2 0 0 5 10 15 20 25 30 35 40 45 50 55 CO pressure (atm)

Figure 5.21 Effect of CO pressure (gauge) on the carbonylation of iron at 80-100 oC after 6.5 hours reaction

5.2.3 Non-catalytic Carbonylation of Cobalt

Formation of cobalt carbonyl requires high temperature and high CO pressure which effect has been reported to be more critical. The experimental CO pressure in the project was up to 55 atm (gauge pressure 54 atm) and the temperature up to 100oC. Although formation of cobalt carbonyl under these experimental conditions is expected to be low, it is important to find the extent of Co carbonylation in carbonylation of reduced nickel laterites, which contain cobalt. The carbonylation of cobalt obtained from reduction of cobalt oxide was studied at temperatures 80, 90 and 100 oC and CO pressure (gauge) 43 and 54 atm; gas flow rate was 0.5 L·min-1.

Effects of temperature and CO pressure on the cobalt carbonylation are shown in Figures 5.22 and 5.23 respectively. The carbonylation of cobalt under given experimental conditions was very slow; less than 0.5% of cobalt was adsorbed in aqua regia. Change of reaction temperature did not affect the carbonylation of cobalt (Figure 5.22). The increase of carbon monoxide pressure from 43 to 56 atm slightly reduced the yield of cobalt carbonylation at 80 and 90 oC (Figure 5.23).

76 5 41 atm 80 oC 90 oC 4 o 100 C

3

2

1 Extent of Co carbonylation (%) of Co carbonylation Extent

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5

Time (Hour) 5 55 atm 80 oC 90 oC 4 o 100 C

3

2

1 Extent of Co carbonylation (%) of Co carbonylation Extent

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (Hour)

Figure 5.22 Effect of temperature on carbonylation of cobalt at CO gauge pressures 41 and 54 atm

77 3 100 oC 43 atm 56 atm

2

1 Extent of Co carbonylation (%) of Co carbonylation Extent

30 90 oC

2

1 Extent of Co carbonylation (%) of Co carbonylation Extent

03 80 oC

2

1 Extent of Co carbonylation (%) of Co carbonylation Extent

0 0 1 2 3 4 Time (Hour)

Figure 5.23 Effect of CO pressure on carbonylation of cobalt at 80, 90 and 100 oC

In the non-catalytic carbonylation of pure metals, nickel was readily converted to carbonyl at 100oC under elevated CO pressure; but the carbonylation of iron and cobalt was much slower, less than 4% and 0.5% of iron and cobalt carbonyls were detected in the adsorption solution under the same conditions.

5.3 Catalytic Carbonylation of Nickel

In the catalytic carbonylation, sulphur, iron sulphide (FeS) and hydrogen sulphide (H2S) are used as catalysts (Teng et al. 2007, Wang et al. 2009). In this project, sulphur and iron sulphide were used in the carbonylation of nickel.

Nickel (about 1.3 g) was mixed with a catalyst before the carbonylation reaction; operational parameters in the catalytic carbonylation experiments are listed in Table 5-6.

78 Table 5-6 Experiment parameters in the catalytic carbonylation of nickel

Parameter Unit Value

Temperature oC 80, 90 100

CO gauge pressure atm 14, 27 43

-1 Gas flow rate L·min 0.29, 0.36, 0.50

Catalyst content in the wt % 0.1 ~ 7.0 mixture with nickel

5.3.1 Effect of the Content of Iron Sulphide and Sulphur on the Carbonylation of Nickel Effect of the content of catalysts on the carbonylation of nickel was studied at 100oC and CO gauge pressure of 14 atm at a flow rate of 0.5 L·min-1. Carbonylation curves obtained in experiments with different contents of sulphur introduced as iron sulphide and sulphur are shown in Figure 5.24. Figure 5.25 shows effects of addition of catalysts on carbonylation of nickel after 4 hours reaction. Compared with the non- catalytic carbonylation of nickel, sulphur containing catalysts greatly accelerated the carbonylation reaction. The reaction time for the extent of carbonylation of 60% was shortened from 4.2 hours in the non-catalytic carbonylation to less than 40 minutes in the catalytic reaction. Catalytic effects of sulphur and iron sulphide were close to one another.

79 100 S

80

60

40 0.0% 0.1% 0.5% 20 2.0% Extent of Ni carbonylation (%) carbonylation of Ni Extent 5.0% 7.0% 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 Time (Hour)

100 FeS

80

60

40 0.0 % 0.1% 0.4% 0.5% 20 2.0% Extent of Ni carbonylation (%) 3.5% 5.5% 7.0% 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 Time (Hour) Figure 5.24 Catalytic carbonylation of nickel at 100 oC and CO pressure (gauge) 14 atm with addition of different amounts of sulphur, introduced as S and FeS

The extent of nickel carbonylation increased from about 65% to over 95% when 0.1 wt% of sulphur was added in the form of FeS; and over 90% with addition of 0.1 wt% of sulphur. Further increase of the catalyst content to 7.0% S resulted in the decrease of extent of nickel carbonylation to 80%. Therefore, 0.1% of sulphur in the form of FeS 80 was added to nickel in the study of the effects of the other parameters on the catalytic carbonylation of nickel.

100 S FeS

90

80

70

60 Extent of Ni Extent (%) carbonylation

50 0 1 2 3 4 5 6 7 8 Content of sulphur (%)

Figure 5.25 Catalytic carbonylation of nickel at 100 oC and CO pressure (gauge) 14 atm with addition of different amounts of sulphur in the form of S and FeS after 4 hours reaction

5.3.2 Effect of Temperature

The effect of temperature on the catalytic carbonylation of nickel was examined in the range of 80 – 100 oC at CO gauge pressure of 14 atm with gas flow rate of 0.5 L·min-1. The sample (1.3g) contained 0.1 wt% sulphur in the form of FeS. Carbonylation curves obtained in experiments at 80, 90 and 100 oC are shown in Figure 5.26. Figure 5.27 shows the extents of the nickel carbonylation after 3 hours reaction. Nickel was carbonylated at a lower rate at 80 oC, which increased with increasing temperature to 90 oC; however nickel carbonylation was not influenced by the increase of temperature from 90 to 100 oC.

81 100

80

60

40

o 20 80 C 90 oC Extent of Ni carbonylation (%) 100 oC 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Time (Hour) Figure 5.26 Effect of temperature on the catalytic carbonylation of nickel

After 3 hours reaction, the increase in temperature from 80oC to 90-100 oC increased the extent of nickel carbonylation from 74.9 to 97.3 %.

100

80

60

40

Extent of Ni Extent (%) carbonylation 20

0 75 80 85 90 95 100 105 Temperature (oC)

Figure 5.27 Effect of temperature on the catalytic carbonylation of nickel after 4 hours reaction

82 5.3.3 Effect of Carbon Monoxide Pressure The effect of carbon monoxide pressure on the catalytic carbonylation of nickel was studied at 100 oC and CO pressure of 14 and 27 atm. The gas flow rate was 0.5 L·min- 1; 0.1% of FeS was added to 1.3 g of nickel.

Catalytic nickel carbonylation at CO pressures (gauge) of 14 and 27 atm is shown in Figure 5.28. The rate of nickel carbonylation at CO pressure of 27 atm was lower than at 14 atm within the first two hours of reaction; the extents of nickel carbonylation after 3 hours reaction at both CO pressures were very close; 96.9% at CO gauge pressure of 14 atm and 94.9% at 27 atm.

100

80

60

40

20 14 atm Extent of Ni carbonylation (%) Extent 27 atm

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Time (Hour)

Figure 5.28 Catalytic nickel carbonylation at 100 oC and CO pressures (gauge) of 14 and 27 atm

5.3.4 Effect of the Gas Flow Rate

The effect of gas flow rate on the catalytic carbonylation of nickel was studied at 100 oC and CO gauge pressure of 14 atm at 0.23, 0.36 and 0.50 L·min-1; samples contained 0.1% sulphur in the form of FeS.

83 The carbonylation curves are plotted in Figure 5.29. Figure 5.30 presents the extents of nickel carbonylation after 3 hours reaction. The increase in the carbon monoxide flow rate increased the reaction rate. The effect of the CO flow rate was particularly strong when the gas flow rate was increased from 0.36 to 0.5 L·min-1.

100

80

60

40

20 0.23 Lmin-1 Extent of Ni carbonylation (%) carbonylation Ni of Extent 0.36 Lmin-1 0.50 Lmin-1 0 0.0 0.51.01.52.02.53.0 Time (Hour)

Figure 5.29 Effect of CO gas flow rate on the catalytic carbonylation of nickel at 100 oC and CO pressure (gauge) of 14 atm.

The extent of carbonylation after 3 hours reaction was 66% at CO flow rate 0.23 L·min-1, it increased to 70 % when the gas flow rate was 0.36 L·min-1, and reached 96.5% when gas flow rate was further increased to 0.5 L·min-1.

84 100

80

60

40

20 Extent of Ni carbonylation (%) Extent

0 0.20 0.25 0.30 0.35 0.40 0.45 0.50 Gas flow rate (Lmin-1) Figure 5.30 Effect of gas flow rate on the catalytic nickel carbonylation at 100 oC and CO pressure (gauge) 14 atm after 3 hours reaction

5.4 Carbonylation of Selectively Reduced Laterite Ores

Selective reduction of laterite ore has been studied by J. Yang in his PhD project (Yang 2014). He examined the reduction of laterite ores, sized to <53 µm and in the range 53- 200 µm. Chemical compositions of these samples are shown in Table 3-2; nickel content was 1.66% and 1.35%, iron content was 11.7% and 10.4%, cobalt content was 0.045% and 0.038% correspondingly. The extent of selective reduction of nickel oxides from the ore <53 µm reached 91% and 85% from the ore with the size range 53-200 µm. The extent of cobalt oxide reduction from these samples was 85% and 99% respectively; however less than 19% of iron oxide was reduced. Ni, Co and Fe oxides were reduced to , which fine (submicron) particles were homogeneously distributed in silicate matrix.

In the thesis, the selective reduction of laterites was based on Yang’s work (Yang 2014), o which was carried out by the reduction of laterite ore at 750 C by the CO-CO2 mixture with 60 vol% CO. Carbonylation of the selectively reduced laterite ores was studied at different temperatures, CO pressures and gas flow rates, and ore particle sizes. As introduced in Chapter 3, two types of laterite ores were supplied, labelled ‘MIN’ and

85 ‘BCS’ ores. They were analysed by XRF and ICP-OES, their composition is presented in Chapter 3. The reduction behaviour of these ores was very similar. Effects of temperature, CO pressure and flow rate on carbonylation of Ni and Fe were studied using BCS laterite ore; the effect of particle size was examined using MIN laterite ore.

Carbonylation of the selectively reduced ores by CO-CO2 gas mixture was compared with carbonylation of the ore reduced by hydrogen; in this case, the reduction was close to completion. Catalytic carbonylation of the selectively reduced laterite ore was also studied.

5.4.1 Non Catalytic Carbonylation of Selectively Reduced Laterite Ores

Non-catalytic carbonylation of selectively reduced laterite ores was studied under conditions listed in Table 5-7.

Table 5-7 Reaction parameters in non-catalytic carbonylation of laterite ores

Parameters Unit Value

Temperature oC 80, 90, 100

CO pressure atm 0, 14, 27, 41 (gauge)

Gas flow rate L·min-1 0.23, 0.35, 0.50

In a study of carbonylation of Ni, Fe and Co, metal oxides were reduced using hydrogen while selective reduction of laterite ores was implements using CO-CO2 gas mixture, containing 60 vol% CO. Effect of reduction of NiO under different conditions on nickel carbonylation was tested at 100 oC and CO pressure 27 atm. The extents of carbonylation of pure nickel reduced by hydrogen at 700 and 800 oC (Section 5.2.1.5) are compared with carbonylation of nickel produced by the reduction o of NiO at 750 C using CO-CO2 gas mixture with 60 vol% CO, in Table 5-8.

86 Table 5-8 Extents of carbonylation of nickel at 100 oC and CO pressure 27 atm (5.5 h reaction), produced by reduction of NiO by H2 and CO-CO2 gas mixture (60 vol% CO)

Reductant and temperature Extent of nickel carbonylation

o H2, 700 C 95.8

o H2, 800 C 90.3

o CO-CO2 (60 vol% CO), 750 C 91.3

o 91.3% of nickel produced by reduction of NiO at 750 C using CO-CO2 gas mixture (60 vol% CO) was converted to nickel tetracarbonyl in 5.5 hours, which was near to the o extent of carbonylation of metallic nickel produced by reduction of NiO by H2 at 800 C. The carbonylation curves are shown in Figure 5.31.

100

80

60

40

CO + CO gas (60 vol% CO) Extent of nickel carbonylation (%) 20 2 o H2 800 C o H2 700 C 0 0 1 2 3 4 5 6 Time (Hour) Figure 5.31 Effects of reduction conditions on carbonylation of nickel at 100 oC, CO gauge pressure 27 atm and gas flow rate of 0.5 L·min-1.

5.4.1.1 Effect of Temperature The effect of temperature on the carbonylation of selectively reduced laterite ore (a sample mass was 1.5 g) was studied at 80, 90 and 100 oC under CO pressure 27 atm at a gas flow rate of 0.5 L·min-1. Carbonylation of nickel and iron at 80-100 oC is

87 shown in Figure 5.32. The extents of nickel and iron carbonylation after 3.5 hours reaction are listed in Table 5-10. Carbonylation of both nickel and iron at 80 and 90 oC was very slow; the carbonylation rate increased significantly when the reaction temperature increased from 90 to 100 oC.

50 o o 80 C 90 oC 100 C

40

30

20 f Ni carbonylation (%)

10 Extent o

0 o o 80 C 90 oC 100 C

30

20 f Fe carbonylation (%) 10 Extent o

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (hour)

Figure 5.32 Carbonylation of nickel and iron from the selectively reduced laterite ores at 80-100 oC and CO gauge pressure 27 atm

The extent of nickel carbonylation after 3.5 hours reaction increased from 0.36% at 80 oC to 1.71 at 90 oC and further to 34.13% at 100 oC. However, the extent of carbonylation of iron at 100oC reached only 3.12%.

Table 5-9 Extents of nickel and iron carbonylation from the selectively reduced laterite ores at 80-100 oC after 3.5 hours reaction

Temperature, oC Fe, % Ni, %

80 0.20 0.36

88 90 0.50 1.71

100 3.12 34.1

5.4.1.2 Effect of Carbon Monoxide Pressure The effect of CO pressure on carbonylation of selectively reduced laterite ore was studied at 100 oC and CO gauge pressure 27 and 41 atm at gas flow rate 0.5 L·min-1. Carbonylation curves are plotted in Figure 5.33; the extents of Ni and Fe carbonylation after 5 hours reaction are presented in Table 5-10. Increase in CO pressure (gauge) from 27 to 41 atm increased the rate of carbonylation of nickel in the first hour of reaction and had no effect on the carbonylation of iron.

50 Ni 40

30

20 f Ni carbonylation (%)

10 27 atm 41 atm Extent o

500

Fe 27 atm 41 atm 40

30

20 f Fe carbonylation (%)

10 Extent o

0 0 1 2 3 4 5 Reaction time (Hour)

Figure 5.33 Carbonylation of nickel and iron from laterite ore at 100 oC and CO pressure of 27 and 41 atm (gauge)

89 Table 5-10 Effect of CO pressure on non-catalytic carbonylation of nickel and iron from the selectively reduced laterite ores at 100 oC

CO pressure, atm Fe, % Ni, %

27 4.05 41.6

41 4.61 48.3

5.4.1.3 Carbonylation of Laterite Ore Reduced by Hydrogen

Carbonylation of selectively reduced laterite ore was compared with carbonylation of ore reduced by hydrogen at 650 and 850 oC. Carbonylation was studied at 100 oC and

CO gauge pressure 27 atm, the gas flow rate was 0.5 L·min-1; the samples of reduced ores were about 1.5 g.

Carbonylation of ores reduced by hydrogen at 650 and 850 oC in comparison with o carbonylation of selectively reduced ore by CO-CO2 gas (60 vol% CO) at 750 C is shown in Figure 5.34. The extents of carbonylation of nickel and iron after 3.5 hours reaction are presented in Table 5-11. The increase of the temperature of reduction by hydrogen enhanced the rate of iron carbonylation, but had a slight negative effect on the carbonylation of nickel, which was slow for ores reduced at both temperatures 650 and 850 oC. The rates of carbonylation of nickel and iron from the selectively reduced ore were higher in comparison with the rates of carbonylation of ores reduced by hydrogen.

90 50 o o Ni CO + CO2 750 C H2 850 C o 40 H2 650 C

30

20

tent of Ni carbonylation (%) of Ni carbonylation tent 10 Ex

0 Fe o o CO + CO2 750 C H 850 C 16 2 o H2 650 C

12

8

tent of Fe carbonylation (%) of Fe carbonylation tent 4 Ex

0 0 1 2 3 4 Time (hour) Figure 5.34 Carbonylation of nickel and iron from selectively reduced laterite ore by CO-CO2 gas mixture (60 vol% CO) in comparison with carbonylation of ore o o reduced by H2 at 650 and 850 C; carbonylation temperature 100 C, CO gauge pressure 27 atm

Table 5-11 Extents of carbonylation of nickel and iron from selectively reduced laterite ore by CO-CO2 gas mixture (60 vol% CO) and from the ore reduced by H2 at 650 and 850 oC (carbonylation temperature 100 oC, CO gauge pressure 27 atm) after 3.5 hours reaction

Reductant Reduction temperature, oC Fe, % Ni, %

CO +CO2 (60 vol% CO) 750 7.48 29.11

850 5.32 3.68 H2 650 1.61 5.40

91 5.4.1.4 Effect of Gas Flow Rate

Selectively reduced laterite ore was carbonylated at 100 oC and CO gauge pressure 27 atm at gas flow rates of 0.36 and 0.50 L·min-1 to study the effect of gas flow rate on the non-catalytic carbonylation of nickel and iron. Carbonylation curves for nickel and iron are shown in Figure 5.35. Change in the CO flow rate from 0.36 to 0.50 L·min-1 had no effect on the carbonylation of nickel. The rate and extent of carbonylation of iron were low at both gas flow rates; they increased with increasing gas flow rate.

50 0.35 Lmin-1 0.50 Lmin-1

40

30

20 f Ni carbonylation (%)

10 Extent o

0

0.35 Lmin-1 0.50 Lmin-1

6

3 f Fe carbonylation (%) Extent o

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (hour)

Figure 5.35 Carbonylation of nickel and iron from the selectively reduced laterite ore at 100 oC and CO pressure (gauge) 27 atm with flow rates 0.36 and 0.50 L·min-1

92 5.4.2 Catalytic Carbonylation of Selectively Reduced Laterite Ore

Catalytic carbonylation of selectively reduced laterite ore was studied with the use of hydrogen sulphide H2S, which was introduced to the system with carbon monoxide with a concentration of 1.0 vol%.

5.4.2.1 Effect of Temperature

Temperature in the catalytic carbonylation of selectively reduced laterite ore was in the range of 80-100 oC; CO pressure (gauge) was 27 atm, the gas flow rate was 0.5 L·min-1.

Carbonylation curves for nickel and iron at 80, 90 and 100 oC are plotted in Figure 5.36. The extents of carbonylation after 4 hours reaction are shown in Figure 5.37. Increase in temperature from 80 to 90 oC increased the rate of carbonylation of nickel and had no effect on the carbonylation of iron. Further increase in temperature to 100 oC led to a marginal increase in the extent of carbonylation of nickel (less than 2% after 4.5 hours reaction), and increased the rate of iron carbonylation; although the extent of iron carbonylation at 100 oC after 4 hours reaction remained low, below 3%.

93 50 100oC 80 oC 90 oC

40

30

20 f Ni carbonylation (%)

10 Extent o

0 100oC 80 oC 90 oC 15

10 f Fe carbonylation (%)

5 Extent o

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (Hour)

Figure 5.36 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 80, 90 and 100 oC; CO pressure (gauge) was 27 atm

94 20 50

16 40

12 30

8 20

4 10 Extent Ni of Extent (%) carbonylation Extent Fe of Extent (%) carbonylation

0 0 80 85 90 95 100 Temperature (oC)

Figure 5.37 Effect of temperature on the extents of catalytic carbonylation of Ni and Fe from the selectively reduced laterite ore at CO gauge pressure 27 atm after 4 hours reaction

5.4.2.2 Effect of Carbon Monoxide Pressure

The effect of CO pressure on the catalytic carbonylation of selectively reduced laterite ore was studied at 100 oC; CO gauge pressure in these experiments was 14, 27 and 41 atm, gas flow rate was 0.5 L·min-1.

Carbonylation curves for nickel and iron are plotted in Figure 5.38; Figure 5.39 presents the extents of carbonylation of nickel and iron after 3 hours reaction. Increasing CO pressure had a strong enhancing effect on the carbonylation of nickel; the extent of Ni carbonylation after 3 hours reaction increased from 15.4 to 49.0% when CO pressure (gauge) increased from 14 to 41 atm. However, the change of CO pressure had a minor effect on the carbonylation of iron.

95 60 41 atm 27 atm 14 atm

50

40

30

20

10 Extent of Ni carbonylation (%)

0 50 41 atm 27 atm 14 atm

40

30

20

10 Extent of Fe carbonylation (%)

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (Hour) Figure 5.38 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 100 oC and CO gauge pressure 14, 27 and 41 atm

60

Ni Fe 50

40

30

Extents (%) Extents 20

10

0 10 15 20 25 30 35 40 45 CO pressure Figure 5.39 Effect of CO gauge pressure on the extents of catalytic carbonylation of Ni and Fe from the selectively reduced laterite ores at 100 oC after 3 hours reaction

96 5.4.2.3 Effect of Gas Flow Rate

The effect of the gas flow rate on the catalytic carbonylation of the selectively reduced laterite ore was studied at 100 oC and CO pressure (gauge) 27 atm. The gas flow rate was set at 0.36 and 0.5 L·min-1.

The carbonylation curves for nickel and iron are plotted in Figure 5.40. The extents of carbonylation of nickel and iron after 3.5 hours reaction are presented in Table 5-12.

The increase in the gas flow rate from 0.35 to 0.50 L · min-1 promoted catalytic carbonylation of both nickel and iron.

50 0.35 Lmin-1 0.50 Lmin-1

40

30

20 f Ni carbonylation (%)

10 Extent o

0

25 0.35 Lmin-1 0.50 Lmin-1

20

15

10 f Fe carbonylation (%)

5 Extent o

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (Hour) Figure 5.40 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 100 oC and CO pressure (gauge) 27 atm at the gas flow rates 0.35 and

0.50 L·min-1.

97 The extents of nickel and iron carbonylation after 3.5 hours reaction increased from 29.3% to 33.8%; and from 0.33 to 3.10% respectively, when the gas flow rate increased from 0.35 to 0.50 L·min-1.

Table 5-12 The extents of nickel and iron catalytic carbonylation from the selectively reduced laterite ore at 100 oC and CO pressure 27 atm at the gas flow rates 0.35 and 0.50 L·min-1 after 3.5 hours reaction.

Extent of carbonylation Extent of carbonylation -1 Gas flow rate, L·min of Fe, % of Ni, %

0.35 0.33 29.3

0.50 3.10 33.4

5.4.2.4 Effect of Particle Size

The effect of particle size was studied in the carbonylation of selectively reduced MIN laterite ore, which chemical composition was presented in Table 3-2. The particle size varied from 38 to 495 µm by sieving the crushed ore before reduction, and according to the particle size, particles were divided to four groups: 38-53, 75-90, 140-200 and 355- 495 µm. Catalytic carbonylation was studied at 100 oC and CO gauge pressure 27 atm. The carbonylation curves for nickel and iron are plotted in Figure 5.41. The extents of carbonylation after 3.5 hours reaction are presented in Figure 5.42. The particle size had a strong effect on the carbonylation of nickel, which rate and extent increased with decreasing particle size. The extent of nickel carbonylation after 3.5 hours reaction reached over 90% when the particle size decreased below 90 μm.

The change in the particle size from the range 75-90 to 355-495 μm had a weak and inconsistent effect on the carbonylation of iron; however, decrease in the particle size to the range 38-53 μm increased the rate of iron carbonylation; the extent of iron carbonylation with the particle size in this range reached 6% after 3.5 hours reaction.

98 100

80

60

40 f Ni carbonylation (%)

20 Extent o 38-53 um 75-90 um 140-200 um 355-495 um 0

15

10 38-53 um 75-90 um 140-200 um 355-495 um f Fe carbonylation (%)

5 Extent o

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Time (Hour)

Figure 5.41 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore of different size at 100 oC and CO gauge pressure 27 atm

100 38-53 m 75-90 m 80 140-200 m 355-495 m

60

40 Exttent of carbonylation (%) 20

0 Fe Ni Metal

Figure 5.42 Effect of particle size on the catalytic carbonylation of Ni and Fe from the selectively reduced laterite ore at 100 oC and CO gauge pressure 27 atm after 3.5 hours reaction

99 5.4.2.5 Effect of the Reduction Temperature on the Catalytic Carbonylation of Laterite Ore

Temperature in the reduction of laterite ore affected both the degree of reduction and the physical state of the reduced ore. This section examines carbonylation of selectively laterite ore (BCS ore 53 ~ 200 µm) reduced at 650, 750 and 850 oC. Carbonylation was studied at 100 oC and CO gauge pressure 27 atm; the gas flow rate was 0.5 L·min-1; the gas contained 1.0 vol% H2S.

Figure 5.43 shows the effect of temperature on the reduction of Ni, Fe and Co oxides in the laterite ore in the range 700-900 oC (Yang 2014). Extents of reduction of Ni and Fe oxide were the highest in the reduction at 750 oC. Carbonylation curves for Ni and Fe are plotted in Figure 5.44. The extents of Ni and Fe carbonylation after 2.75 hours reaction are presented in Figure 5.45. The rate and extent of Ni carbonylation increased with increasing reduction temperature. Carbonylation of iron was not affected by the reduction temperature.

100 Figure 5.43 Reduction of Ni, Co and Fe oxides from the laterite ore with size (a) < 53 μm and (b) 53-200 μm in the gas atmosphere containing 60vol% CO and 40 vol%CO2, reduction time was 60 min (Yang, 2014)

100 Ni-650 Ni-750 Ni-850

80

60

40 f Ni carbonylation (%)

20 Extent o

0

40

30

20 f Fe carbonylation (%)

10 Extent o

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (Hour)

Figure 5.44 Catalytic carbonylation of nickel and iron from the selectively reduced laterite ore at 650, 750 and 850 oC; carbonylation temperature 100 oC, CO gauge pressure 27 atm

100 Fe Ni

80

60

40

20 Extent of carbonylation (%) carbonylation of Extent

0 650 700750800850 Reduction temperature (oC)

Figure 5.45 Effect of the ore reduction temperature on the catalytic carbonylation of Ni and Fe at 100 oC and CO gauge pressure 27 atm 101 5.4.2.6 Catalytic Carbonylation of Laterite Ore Reduced by CO-CO2 gas with different CO Content

Composition of CO-CO2 gas mixture affects the reduction of laterite ore. The extent of the reduction of Ni, Fe and Co oxides from laterite ore (53-200 μm) at 750 oC using

CO-CO2 gas mixture with CO partial pressure in the range 0.2 to 0.7 atm is shown in

Figure 5.46 (Yang, 2014). When CO concentration in CO-CO2 gas was 70 vol%, the reduction was not selective; the degree of reduction of iron oxides was above 50%. .

o In the study of the carbonylation, laterite ore was reduced at 750 C by CO-CO2 gas with 49, 60 and 68 vol% CO.

Figure 5.46 Effect of CO partial pressure on reduction of Ni, Co and Fe oxides from 53-200 µm laterite ore at 740 oC for 60 min; gas flow rate was 700 ml∙min-1

Catalytic carbonylation of the reduced ore was studied at 100 oC and CO pressure 27 atm. Carbonylation curves for Ni and Fe are plotted in Figure 5.47. Effect of CO content in the CO-CO2 gas used in reduction of laterite ore on the catalytic carbonylation of Ni and Fe after 3.5 hours reaction is shown in Table 5-13. The rate and extent of carbonylation of Ni and Fe in samples reduced by the CO-CO2 gas with 40 vol% CO were lower than in samples obtained in reduction by gas containing 60 and 68 vol% CO. The extent of Ni carbonylation after 3.5 hours reaction was 20.3% in the carbonylation of a sample reduced by gas with 49 vol% CO, and 33-34% when

102 reduction was in the gas with 60-68 vol% CO. Carbonylation of iron in all cases had a low extent; the highest extent of iron carbonylation (3.1 %) was reached in carbonylation of a sample produced by reduction with gas containing 60 vol% CO.

50

49 vol% CO 68 vol% CO 60% vol% CO 40

30

20 f Ni carbonylation (%)

10 Extent o

0

49 vol% CO 30 68 vol% CO

60 vol% CO 20 f Fe carbonylation (%) 10 Extent o

0 Time (Hour) Figure 5.47 Catalytic carbonylation of nickel and iron from laterite ore selectively reduced by CO-CO2 gas with different CO concentration. Carbonylation temperature was 100 oC, CO gauge pressure 27 atm

Table 5-13 Extents of catalytic carbonylation of nickel and iron from ores reduced by CO-CO2 gas with different contents of CO

Concentration of CO (vol %) in Carbonylation of Ni Carbonylation of Fe (%) CO-CO2 gas (%)

49 0.55 20.27

60 3.12 34.13

68 1.02 32.85

103 5.4.2.7 Carbonylation of Laterite Ore Selectively Reduced with Addition of Hydrogen Sulphide

It was shown that sulphur has a catalytic effect on carbonylation of selectively reduced laterite ore. It can be expected that incorporation of sulphur into the sample in the process of reduction of laterite ore also catalyses carbonylation of the reduced ore.

Hydrogen sulphide (as 5 vol% H2S-CO gas mixture) was added to the CO-CO2 gas (60 vol% CO) in the reduction of laterite ore at 750 oC.

In one experiment, H2S was introduced during last 10 min of reduction; in another, H2S was added to the gas at the end of the reduction experiment during 30 min. Carbonylation of selectively reduced laterite ore was studied at 100 oC in pure carbon monoxide at 27 atm. The sample’s mass was about 1.5g, and the gas flow rate was 0.5 L

·min-1.

Carbonylation curves for nickel and iron are plotted in Figure 5.48. The extents of carbonylation of nickel and iron after 2.5 hours reaction are presented in Table 5-14.

Addition of 5.0 vol% H2S to the CO-CO2 gas during 10 min at the end of reduction of laterite ore increased the rate and extent of carbonylation of nickel; the extent of Ni carbonylation after 2.5 hours reaction increased from 32.0. to 49.1%. However, extension of introduction of H2S to 30 min had a negative effect on Ni carbonylation; the extent of Ni carbonylation after 2.5 hours reaction decreased to 28.5%.

Carbonylation of iron from laterite ore reduced with addition of H2S was very slow; the extent of Fe carbonylation after 2.5 hours reaction was 0.4-0.6%.

104 60

50

40

30

1

f Ni carbonylation Ni f (%) 20 2 10 3 Extent o Extent

0 1

30 2 3

20 f Fe carbonylation Fe f (%) 10 Extent o Extent

0 Time (hour)

Figure 5.48 Carbonylation of nickel and iron from laterite ore reduced by CO-CO2 gas o (60 vol% CO) with addition of 5 vol% H2S. Carbonylation temperature was 100 C; CO gauge pressure 27 atm; 1: Introduction of H2S during last 10 min of the reduction of laterite ore; 2: Introduction of H2S during last 30 min of the reduction of laterite ore; 3: catalytic carbonylation (with addition of 1 vol% H2S) of the selectively reduced laterite ore without catalyst;

Table 5-14 Extents of carbonylation of Ni and Fe from laterite ore reduced with o addition of H2S. Carbonylation temperature was 100 C; CO pressure 27 atm; reaction time 2.5 hours

Time of introduction of H2S Carbonylation of Fe, % Carbonylation of Ni, %

10 minutes 0.42 49.08

30 minutes 0.58 28.51

105 0* 2.81 31.99

*: Catalytic carbonylation (with addition of 1 vol% H2S) of Ni and Fe from laterite ore reduced by CO-CO2 gas with 60 vol% CO without addition H2S.

106 Chapter 6 Discussion

The results presented in Chapter 5 demonstrated the complexity of the carbonylation processes, which are affected by many factors (or operation parameters). This chapter presents analysis of experimental data.

6.1 Impurities in Nickel and Iron Oxides

The purity of nickel, iron and cobalt oxides was over 99% (Table 3-1). The impurities in raw oxides and reduced samples were under the detection limit in XRD measurements. After carbonylation of metallic nickel, the content of impurities in residual samples was high enough for the XRD analysis. When over 98% of nickel was carbonylated from a sample containing over 99% of nickel oxide (before reduction), the content of impurity in the residual sample was more than 30%. Figure 6.1 shows the XRD patterns of a reduced nickel sample and residual sample after carbonylation; only silicon dioxide was detected in the sample besides uncarbonylated metallic nickel and nickel oxide.

Because iron carbonylation was very slow, less than 5.0% of iron was converted to iron carbonyl in the experiments within 6.5 hours reaction. The impurity content was below the detection level by XRD analysis, but SEM/EDS analysis detected silicon dioxide in the residual sample.

Based on the XRD and SEM/EDS analysis, the impurities identified in nickel samples used in carbonylation was only silicon dioxide. Both oxides are stable under experimental conditions and do not influence the carbonylation of nickel and iron.

107 280000

Ni NiO SiO2 240000

200000

160000 ount (cts) ount

C 120000

80000

40000

Reduced sample 3000

2000 ounts (cts) ounts C 1000

Residue after carbonylation 0 10 20 30 40 50 60 70 80 90 100

Position 2 [o] Figure 6.1 XRD patterns of reduced nickel and residual sample after carbonylation at 100 oC and CO pressure 27 atm at a gas flow rate of 0.5 L∙min-1

6.2 Reduction of Nickel, Iron and Cobalt Oxides

In the early studies (Mond et al. (1890), Mond et al. 1891, Mond et al. 1908), reduction temperature was suggested to be as low as possible to obtain fine metal particles and avoid sintering. Based on literature (Kawasaki et al. 1962, Turkdogan et al. 1971, Sastri et al. 1982, Qiu et al. 1992, Rashed et al. 1997, Rodriguez et al. 2002, Lin et al. 2003, Utigard et al. 2005, Pineau et al. 2006, Janković et al. 2008, Wagner et al. 2008, Jeangros et al. 2013), the reduction of iron oxide (which is the strongest oxide in comparison with nickel and cobalt oxides) in the thesis was examined from as low temperature as 300oC to find the lowest reduction temperature for nickel, iron and cobalt oxides.

Reduction of oxides was promoted by increasing temperature (Section 5.1) as a result of the effect of temperature on the thermodynamics and kinetics of the reduction reactions.

108 Iron oxide was completely reduced by hydrogen at 400oC for 4 hours. However reduction of nickel oxide was incomplete under the same reduction conditions, because of the slower reduction kinetics. The complete reduction of nickel oxide was achieved at 500oC after 6 hours reaction (Figure 5.2). Further experiments confirmed that cobalt oxide and mixture of nickel and iron oxides were completely reduced at 500 oC (for 6 hours) as shown in Table 5-3, Figures 5.3 and 5.4.

When nickel oxide was mixed with iron oxide prior to reduction, the reduction product was ferronickel alloy. However, it was impossible to obtain fine alloy particles (in µm) in the laboratory for the carbonylation experiments. Therefore, the nickel - iron mixture was prepared from nickel and iron reduced from associated oxides (apart of the mixture for the analysis of the carbonylation time).

Sintering of fine metallic particles took place during reduction. SEM images of nickel samples reduced at 600 and 700 oC for 6 and 12 hours (Figure 6.2) demonstrate that nickel particles sintered and grew up at high temperature, which is consistent with literature (Igharo et al. 1985, Sehested et al. 2001). However, metallic particle sizes were not affected by the variation of reduction time from 6 to 12 hours. Based on this phenomenon, nickel particles of different sizes used in the study of the effect of particle size in Section 5.2.1.5 were obtained from reduction of nickel oxide at temperatures varied from 500 to 900 oC.

109 Figure 6.2 SEM images of nickel reduced at 600 and 700 oC for 6 and 12 hours

The carbonylation process was used in extracting metals from selectively reduced laterite ores. The reduction of laterite ores had been studied by J. Yang in his PhD thesis (Yang, 2014). The ore reduction was examined in a horizontal tube furnace at 700 – o 1000 C in the CO – CO2 gas mixture containing 20, 40, 60 and 70 vol% CO; the effect of gas flow rate on the ore reduction was examined in the range of 350 – 1050 mL∙min-1. Two sizes of ore, 53-200 µm and <53 µm were examined; the main outcomes of Yang’s thesis include:

The maximum degree of reduction of nickel and cobalt oxides was obtained at 740 oC in a CO –CO2 gas mixture with 60 vol% CO; extent of nickel oxide reduction reached 91% for ore in the size range <53 µm, and 85% in the size range 53 – 200 µm, the extents of reduction of cobalt oxide were 94 and 99% respectively. The effects of temperature and CO partial pressure on the reduction of nickel and iron oxides were shown in Figures 5.43 and 5.46. Ni, Co and Fe oxides were reduced to ferronickel, which fine (submicron) particles were distributed in the silicate matrix.

110 Optimum conditions for selective reduction of laterite ores identified by Yang (2014) were used in this thesis for preparation of samples for investigation of carbonylation of Ni and Fe from the selectively reduced laterite ores.

6.3 Non-Catalytic Carbonylation of Nickel

This section discusses the non-catalytic carbonylation of pure nickel at different temperatures, carbon monoxide pressures, gas flow rates, nickel weights and particle sizes. The experimental results were presented in Section 5.2.1.

Main stages in the solid state-gas reaction include mass transfer in the gas phase, diffusion through the ash (reacted zone), and chemical reaction. Metallic nickel carbonylation of which was experimentally studied was reduced from nickel oxide at 500 oC. SEM image of the reduced nickel is shown in Figure 6.3; the sintering of nickel particles was not observed.

Figure 6.3 SEM image of nickel reduced from nickel oxide by hydrogen at 500 oC

In the process of carbonylation, nickel was vaporised into the gas phase, and the impurities in nickel oxide formed an ash layer. Because the content of silicon dioxide was very low (<<1%), the ash layer at the initial stage of carbonylation was fluffy and 111 the influence of diffusion can be ignored. When the most of nickel was carbonylated and flushed off by carbon monoxide, the core of nickel particle became small, and the role of internal diffusion through the ash layer in the rate control of carbonylation became more significant. These processes are schematically shown in Figure 6.4. The ash layer was observed in all carbonylation reactions in which the extent of carbonylation was over 98%.

Figure 6.4 Schematic diagram of pure nickel carbonylation

The ash contained silicon dioxide and nickel oxide. These oxides were stable in both reduction and carbonylation reactions, but the diffusion through the ash layer influenced the carbonylation process when most of nickel was extracted.

6.3.1 Effect of Temperature

Thermodynamic analysis in Chapter 4 showed that carbonylation of nickel and iron is favoured by decreasing temperature. However, reaction kinetics requires high temperatures, as the reaction rate constant exponentially increases with temperature. Temperature interval 80-100 oC presents a compromise between thermodynamic and kinetic conditions for the carbonylation reactions.

Experiments conducted at 80-100 oC showed (Figure 5.8) that an increase in temperature enhanced the reaction rate and yield of carbonylation reaction. A linear correlation was found between the extent of nickel carbonylation after 6.5 hours reaction and temperature (Figure 5.9). The increase of temperature from 80 to 100 oC made the reaction complete in 6.5 hours under CO pressure over 27 atm. Figure 6.5 displays the XRD patterns of residuals of samples after carbonylation at 100 oC and CO

112 pressure 27 atm. The XRD spectra of residuals include peaks of NiO, which was not reduced and carbonylated. Only trace of nickel was identified in the residual of carbonylation at 100 oC; while XRD spectra of residuals of carbonylation at 80 and 90 oC contained strong nickel peaks. The content of impurities in residuals greatly increased after metallic nickel was removed.

NiO SiO 131072 Ni 2

65536

32768 ounts (cts) ounts

C 16384

8192 131072 80 oC 65536

32768

ounts (cts) ounts 16384 C

8192

4096 90 oC 3000

2000 ounts (cts) ounts

C 1000

100 oC 0 10 20 30 40 50 60 70 80 90 100

Position 2 [o]

Figure 6.5 XRD patterns of residuals of samples after carbonylation at different temperatures at CO gauge pressure 27 atm (6.5 hours reaction)

The rate constant as a function of temperature is described by on Arrhenius equation (6- 1)(Laidler 1984);

Ea  k Ae RT (6-1)

Where, k is reaction rate constant s-1; A is a frequency factor or pre-exponential factor; -1 -1 -1 Ea is activation energy, kJ•mol ; R is the gas constant, kJ•mol •K ; T is temperature, K.

Equation (6-1) can be presented in the following form:

113 E ln(kA ) ln( ) a (6-2) RT

The activation energy of nickel carbonylation can be obtained from the plot of ln k vs 1/T.

It follows from the carbonylation curves obtained at 80 - 100 oC and CO pressure 56 atm (Figure 6.6) that the reaction rates were constant in the first two hours. The rate constants calculated from the experimental data of Figure 6.6 in the temperature range 80 to 100 oC presented as ln k are plotted vs 1/T in Figure 6.7.

100 80 oC 85 oC o 80 90 C 95 oC 100 oC 60

40

20 Extent of nickel carbonylation (%)

0 0 1 2 3 4 5 6 Time (hour)

Figure 6.6 Effect of temperature on the carbonylation of nickel at CO pressure 56 atm

The liner plot indicates that the chemical reaction kinetics (intrinsic rate control) contributed to the control of the rate of the non-catalytic carbonylation of nickel. The Arrhenius activation energy determined from the experimental data at temperatures varied from 80 to 100 oC was found 114 kJ∙mol-1. The activation energy value is reasonably close to the Arrhenius activation of 128 kJ∙mol-1 reported by (Redmon 1980).

114 Temperature (oC) 80 85 90 95 100 -1.5

-2.0

-2.5

-3.0

Ln(k) -3.5 Equation y = a + b*x Weight No Weighting Residual Sum 0.20513 -4.0 of Squares Pearson's r -0.96392 Adj. R-Square 0.90552 -4.5 Value Standard Err Intercept 34.43167 6.0046 Ln(k) Slope -13668.722 2179.32952 -5.0 0.00267 0.00270 0.00273 0.00276 0.00279 0.00282 0.00285 1/T (K-1) Figure 6.7 Plots of ln k versus 1/T at temperatures between 80 and 100 oC and CO pressure 56 atm, gas flow rate 0.5 L∙min-1

Figure 6.8 shows the change in the rate of nickel carbonylation with time at 80, 90 and 100 oC in the non-catalytic carbonylation of nickel at CO pressure 27 atm. It can be seen that the reaction rate at 100 oC gradually decreased with the reaction time and only slightly decreased when the experiments were conducted at 80 and 90 oC (the final extents of nickel were less than 80% and 50% after 6.5 hours reaction respectively). The decrease in the reaction rate at 100 oC can be attributed to the decreasing surface area of Ni particles, and diffusion through the ash layer which started to influence the process when most of nickel was carbonylated.

Figure 6.8 also shows that the increase of temperature from 80 to 90 oC had a slight effect on the nickel carbonylation, but the reaction rate was enhanced by 1.5 times when the reaction temperature was increased from 90 to 100 oC.

115 Figure 6.8 Rate of non-catalytic carbonylation of nickel at 80, 90 and 100 oC and CO pressure 27 atm; gas flow rate 0.5 L∙min-1

6.3.2 Effect of Carbon Monoxide Pressure

Carbon monoxide pressure is another critical parameter in the carbonylation reaction (Mond et al. 1891-1, Mond et al. 1908, Wang et al. 2009). Experimental results in Figures 5.10 and 5.11 demonstrate the importance of selecting an appropriate operation pressure.

Increasing CO pressure in the non-catalytic carbonylation enhanced the reaction rate when it was less than 27 atm (gauge pressure). Complete carbonylation was reached at 100oC and CO pressure over 27 atm after 5 hours reaction. Figure 6.9 shows XRD patterns of residual samples after carbonylation at 100 oC and CO pressures of 14 and 27 atm. The content of nickel was high in samples carbnonylated at 14 atm, when the extent of carbonylation was about 50 % nickel; only a small amount of nickel was detected in the sample after carbonylation at 27 atm.

116 NiO Ni SiO2

131072

65536

32768 ounts (cts) ounts C 16384

8192 100 oC 14 atm 3000

2000 ounts (cts) ounts

C 1000

100 oC 27 atm 0

Position 2 [o]

Figure 6.9 XRD patterns of residual samples after non-catalytic carbonylation of nickel at 100oC at CO pressures of 14 and 27 atm

The following should be taken into account in the consideration of the effect of carbon monoxide pressure on the carbonylation reaction. Firstly, increasing carbon monoxide pressure shifts the reaction towards the formation of nickel carbonyl as discussed in Chapter 4; secondly, it increases carbon monoxide concentration on the interface, and enhances the rate of the first order reaction of direct synthesis of nickel tetra-carbonyl (Day et al. 1968); finally, the gas velocity in the reactor decreases with the increase of carbon monoxide pressure when the outlet gas flow rate (under atmosphere pressure) is constant. The decrease of the gas flow rate at the outlet at constant CO pressure was found to have a negative effect on the nickel carbonylation (Figure 5.6). It can be concluded that positive effects of the increasing CO pressure on the carbonylation rate dominate over the negative effect.

The carbonylation rates of non-catalytic carbonylation of nickel at 100 oC and CO gauge pressures 14 and 27 atm (gas flow rate 0.5 L·min-1) are plotted in Figure 6.10. The carbonylation rates in the first 2 hours increased from 15% hour-1 to about 25%

117 100

75

50

-1 of nickel carbonylation (%) hour (6725 % increase) when the carbon monoxide gauge pressure raised from 14 atm to

27 atmExtent (86 % increase). This means, that the decrease in the gas velocity (53.6%) slowed down the increase in the carbonylation rate. 0 100 )

-1 14 atm 27 atm hour

 80

60

40

20 of nickel carbonylation (% carbonylation of nickel Rate 0 0 1 2 3 4 5 Time (Hour)

Figure 6.10 Rate of non-catalytic carbonylation of nickel at 100oC and CO gauge pressures 14 and 27 atm (gas flow rate 0.5 L·min-1)

6.3.3 Effect of Gas Flow Rate

The effect of gas flow rate on the non-catalytic nickel carbonylation was studied with the outlet gas flow rate 0.14, 0.23, 0.35 and 0.50 L·min-1 (at room temperature and atmospheric pressure)

The results in Figure 5.6 demonstrated that the increasing outlet gas flow rate enhanced the reaction rate. XRD patterns of residual samples are shown in Figure 6.11. Nickel was detected in the residual sample after 6.5 hour experiment at 0.14 L·min-1 ( about 56% extent of nickel carbonylation), while in the residual sample in the experiment with gas flow rate 0.5 L·min-1 only unreduced NiO and impurities were detected (about 99% extent of nickel carbonylation). This analysis confirms results presented in Figure 5.6.

118 262144 Ni NiO SiO2

131072

65536

ount (cts) 32768 C

16384

-1 8192 0.14 Lmin 16384

8192 ount (cts) C 4096

0.50 Lmin-1 2048 10 20 30 40 50 60 70 80 90 100 Position (2o]) Figure 6.11 XRD patterns of residuals of samples after non-catalytic carbonylation of nickel at 100 oC and CO pressure 27 atm at flow rates 0.14 and 0.50 L·min-1 after 6.5 hours reaction.

The rates of non-catalytic carbonylation of nickel, found from the experimental data in Figure 5.6, are shown in Figure 6.12. Carbonylation rates increased in the first half hour, and then gradually decreased. The rates at the beginning of the nickel carbonylation were about 17, 21.3 and 24.5 %∙hour-1 for gas flow rates of 0.23, 0.35 and 0.50 L∙min-1 correspondingly.

119 100 0.50 Lmin-1 0.35 Lmin-1

) -1 -1 0.23 Lmin 80 hour 

60

40

20 Rate of nickel carbonylation (%

0 0 1 2 3 4 5 6 Time (Hour)

Figure 6.12 Rates of nickel carbonylation at 100 oC and CO pressure 27 atm at gas flow rates of 0.23, 0.35 and 0.50 L∙min-1

However, the decline in the carbonylation rate also increased with the increasing gas flow rate. The rates of carbonylation after 5-hour reaction were about the same at all studied gas flow rates. The carbonylation of nickel was close to completion after 6 hours reaction.

6.3.4 Effect of Nickel Mass

Results in Figure 5.15 demonstrate that the carbonylation of pure nickel was not affected by the nickel sample weight below 3.2g; the rates of carbonylation of samples with mass 0.8, 1.3 and 3.2 g were the same.

Thermodynamic analysis of the carbonylation reaction in Chapter 4, showed an over o 99% conversion of nickel to Ni(CO)4 at 100 C and CO pressure 43 atm. According to the carbonylation curves in Figure 5.15, the concentration of nickel carbonyl in the gas phase was constant during the first two hours. The mass of the carbonylated nickel calculated for different concentrations of nickel tetracarbonyl (from 0.5 to 10 vol%) at a gas flow rate of 0.5 L∙min-1 (constant carbonylation rate from Figure 5.15) is presented in Figure 6.13.

120 25 6 10 % 5 20

2% 4 15

3

10 2 Nickel weight (g) 1% Nickel weight (g)

5 1 0.5%

0 0 0 20 40 60 80 100 120 140 160 180 Time (minutes)

Figure 6.13 Mass of carbonylated nickel at different concentrations of nickel tetracarbonyl in gaseous phase at a gas flow rate of 0.5 L·min-1

In the experiments, the extent of nickel carbonylation of all three samples in 180 minutes was about 85%, what means that the amount of carbonylated nickel was from 0.68 g to 2.72 g. According to Figure 6.13, the concentration of nickel carbonyl in the reactor was less than 1.0 vol%. The sample mass used in the experiments was 1.3g, therefore the maximum concentration of nickel carbonyl was less than 0.5 vol% during the carbonylation process.

6.3.5 Effect of Particle Size

Mond (Mond et al. 1890, Mond et al. 1891-1, Mond et al. 1908) recommended fine metal particles for carbonylation reaction to increase specific surface area and rate of gas-solid reaction. However, the specific surface area and reaction rate are also affected by porosity and pore structure.

SEM images in Figure 5.12 show that nickel particles sintered and increased in size with the increase of the reduction temperature (Table 5-5). BET specific surface area and BJH pore size of samples produced by reduction at different temperatures are

121 presented in Table 6-1. Samples varied in size from 0.29 to 2.67 µm; their BET surface area and BJH pore size are also listed in this table.

Table 6-1 BET surface area and pore size of samples produced by reduction at different temperatures and residual samples after carbonylation

BET (Brunauer-Emmett- Particle size (reduction Teller) specific surface BJH (Barrett-Joyner- temperature) µm Halenda) pore size, nm area, m2·g-1

0.51 (600 oC) 0.470±0.022 24.2

1.10 (700 oC) 0.171±0.019 -

2.07 (800 oC) 0.0208±0.0081 -

2.67 (900 oC) 0.0093±0.0040 -

Residue of sample reduced 25.6±0.67 18.8 at 500oC

Residue of sample reduced 10.1±1.31 25.2 700 oC

Residue sample 500 oC* 33.2±1.20 20.8

* This sample was from the catalytic carbonylation reaction (0.1 wt% S in the form of FeS)

The BET surface areas of the reduced samples decreased with the increase of the particle size. The BJH desorption average pore size was calculated only for four samples; for other samples, it was over the measurement range. It can be expected that the pore size increased with the increasing particle size (Figure 5.12). The decrease in the specific surface area had a negative effect on the carbonylation rate, what explains the decrease in the rate of carbonylation of nickel with the increase of nickel particle size (Figure 6.14).

122 120 0.29 m ) -1 0.51 m 100 1.10 m 2.07 m 2.67 m 80

60

40

20

Rate of nickel carbonylation - hour (% 0 0 1 2 3 4 5 6 7 Time (hour) Figure 6.14 Rate of carbonylation of nickel with different particle size at 100 oC and CO pressure of 27 atm (gas flow rate 0.5 L∙min-1).

6.4 Non-catalytic Carbonylation of Iron and Cobalt

Experimental study of carbonylation of iron (Section 5.2.2) and cobalt (Section 5.2.3), demonstrated that carbonylation of these metal under given experimental conditions was very slow; less than 5% iron was carbonylated and less than 0.5% cobalt was converted to cobalt carbonyl at 100 oC and CO pressure of 41 and 55 atm. Carbonylation of cobalt was not studied further.

Slow non-catalytic carbonylation of iron observed in this project (Figure 5.18) was consistent with results obtained by other researchers (Mond et al. 1891-1, Dewar et al. 1907, Herrmann 1990). The increase of temperature from 80 to 100 oC had a minor effect on the reaction rate; effect of carbon monoxide pressure was also weak.

o The of Fe(CO)5 is 103 C under atmospheric pressure. However, partial pressure of Fe(CO)5 in the gas phase (CO-Fe(CO)5 gas mixture) was very low. Therefore, gaseous iron carbonyl was formed with pressure well below the saturation vapour point. The rate and extent of iron carbonylation at 80 oC were higher than at 100 oC, what confirms that iron carbonyl was not in the liquid state. Moreover, iron carbonylation was catalysed by sulphur addition, what also indicates that no liquid 123 phase was present in the reactor; otherwise, liquid Fe(CO)5 would block contact gas- solid contacts. In the early study of iron carbonylation by Mond (Mond et al. 1891-1, Mond et al. 1908), the reaction temperature was increased over 150 oC to reach acceptable rate of iron carbonylation, which is higher than the temperature range examined in this project.

Compared with iron, the carbonylation of cobalt was much lower in experiments, which make it impossible to extract cobalt during nickel carbonylation in the experiments. Thence, cobalt carbonylation was not further studied in the thesis.

6.5 Non-catalytic Carbonylation of Nickel from Ni-Fe Mixture

Iron is a main component of laterite ores (Oliveira et al. 2001, Moskalyk et al. 2002, Wells et al. 2003, Peng 2005, Xu 2005); it is partially converted to iron penta-carbonyl in the process of nickel carbonylation (Mond et al. 1891, Mond et al. 1891-1) although with a low rate. The nickel carbonylation was much faster than iron carbonylation under experimental conditions of this study. Figure 6.15 shows the rate of nickel carbonylation from the Ni-Fe mixtures of different compositions at 100 oC and CO pressure 27 atm.

50 11% Fe

) 20% Fe -1 50% Fe 40 80% Fe 89% Fe Pure nickel 30

20

10 Rate of nickel carbonylation (% - hour

0 0 1 2 3 4 5 6 Time (Hour)

124 Figure 6.15 Rate of nickel carbonylation from Ni-Fe mixtures at 100oC, CO pressure 27 atm and gas flow rate 0.5 L·min-1

Iron carbonylation was very slow in comparison with carbonylation of nickel. Iron remained in the sample in the process of nickel carbonylation what affected mass transfer of CO to nickel particles with a negative effect on the Ni carbonylation rate.

6.6 Catalytic Carbonylation of Nickel

The experimental results in Chapter 5 have showed the catalytic effects of sulfur containing substance on the carbonylation of pure nickel, iron and nickel-iron mixture.

6.6.1 Mechanism of catalytic carbonylation of nickel

In order to determine what happened to the catalysts in the process of carbonylation, the catalytic carbonylation experiments were terminated at 10, 20 and 30 minutes, and residual samples were analyzed by XRD. Samples reactivated with hydrogen at 130 oC before carbonylation were also analysed. XRD spectra of samples carbonylated for different time are shown in Figure 6.16.

When elemental sulphur was used as a catalyst, , Ni3S2, was detected in the samples before the carbonylation started and in the process of carbonylation.

Iron sulphide used as a catalyst, was observed in the sample reactivated with hydrogen before the carbonylation started. In the process of carbonylation, iron sulfide was gradually converted to nickel , as nickel has higher affinity for sulphur than iron.

XRD analysis of a residual sample after carbonylation of nickel catalyzed by FeS for 3 - 4 hours (Figure 6.17) detected NiO and FeS.

125 32768 Ni - S - 30 min Ni

Ni3S2 8192

2048

32768 Ni - S - 20 min

) 8192

2048

Counts (cts 32768 Ni - S - 10 min

8192

2048

32768 Ni - S - 00 min

8192

2048

10 20 30 40 50 60 70 80 90 100 Position (o2 )

32768 Ni- FeS - 30 min Ni S Ni NiS2 3 2 FeS Fe0.96S 8192

2048

32768 Ni - FeS - 20 min )

8192 Counts (cts 2048

32768 Ni - FeS - 00 min

8192

2048

10 20 30 40 50 60 70 80 90 100 Position (o2)

Figure 6.16 XRD patterns of residual samples of Ni catalyzed by S and FeS (content of sulfur 5.0 wt%) at temperature 100 oC and CO pressure 14 atm; experiments were terminated at 10, 20 and 30 minutes in reactions with addition of sulphur, and 20 and 30 in reactions with addition of FeS. Samples marked as 00 min were not subjected to carbonylation; they were reactivated by heating in hydrogen and then cooled down to room temperature.

126 1800 7% 3.5 % NiO 1500 FeS

1200

900 Counts

600

300

20 30 40 50 60 70 80 90 100 Position (o2 )

Figure 6.17 XRD patterns of residual samples of nickel after carbonylation catalyzed by sulphur (3.5 and 7.5 wt%) in the form of FeS at 100 oC and CO pressure 14 atm after 3-4 hours reaction

It indicates that nickel sulphide was consumed in the process of carbonylation; while unconverted FeS was left in the residue. Therefore, when FeS was used as a catalyst, two processes occurred; conversion of iron sulphide to nickel sulphide and of nickel carbonylation by nickel sulphide. Carbonylation of nickel made iron sulphide “visible” in samples after 3-4 hours reaction.

Similar results were obtained in the catalytic carbonylation of nickel-iron mixture. XRD patterns of samples withdrawn in the progress of carbolynation (10, 20 and 30 min reaction) are shown in Figure 6.18. Nickel sulfide was observed in samples after 10 minutes experiments; it was formed by reaction of nickel with elemental sulphur and with FeS added to nickel oxide as a catalyst. Complex sulphide (Fe,Ni)Sx was also observed in samples after 20-30 min carbonylation. Nickel was converted to Ni(CO)4 in the process of carbonylation, leaving FeS in the residue.

127 Ni : Fe = 1:1 5% S 30 min Fe FeS2 Ni Ni3S2 4096 Fe4.005Ni4.995S8

2048

7000 Ni : Fe = 1:1 5% S 20 min 6300

2048

15000 Ni : Fe = 1:1 5% S 10 min ts) 4096 Counts (c 2048 15000 Ni : Fe = 1:1 5% S 00 min

2048

140000 Ni : Fe = 1:1 5% S S 120000

32768

16384 10 20 30 40 50 60 70 80 90 100 Position (o2 )

10000 Fe 9000 Ni : Fe = 1 : 1 5 % FeS 30min FeS Fe S Ni 8000 1-x

2048

10500 Ni : Fe = 1 : 1 5 % FeS 20min Fe9Ni9S16 4096

2048 12000 10500 Ni : Fe = 1 : 1 5 % FeS 10min Ni3S2 ts) 4096 Counts (c 2048 16000 Ni : Fe = 1 : 1 5 % FeS 00min Fe S 14000 0.94

4096

2048 18000 Ni : Fe = 1 : 1 5 % FeS

4096

2048 10 20 30 40 50 60 70 80 90 100 Position (o2)

128 Figure 6.18 XRD patterns of residual samples of Ni-Fe mixture catalyzed by sulphur and iron sulphide FeS (content of sulfur 5.0%) at temperature 100 oC and pressure 27 atm; experiments were terminated at 10, 20 and 30 minutes.

Therefore, the actual catalyst in the carbonylation reaction was nickel sulfide in both cases when sulfur or iron sulphide were used as catalysts.

Catalytic nickel carbonylation with the use of sulphur is illustrated in Figure 6.19.

Figure 6.19 Schematic of the carbonylation of nickel catalyzed using elemental sulphur.

The following reactions occurred when iron sulphide was used as a catalyst:

2FeS + 3Ni  Ni3S2 + 2Fe (6-7)

Ni3S2 + 14CO  3Ni(CO)4 + 2COS (6-8)

tNi3S2 + xFe +(y-3t)Ni  FexNiyS2t (6-9)

FexNiySz + 4yCO  yNi(CO)4 + zFeS +(x-z)Fe (6-10)

129 6.6.2 Catalytic Carbonylation of Nickel

Addition of sulphur to nickel had a strong catalytic effect on nickel carbonylation which decreased with the increase in the catalysts content (Figures 5.24). These results are consistent with data reported in literature (Hieber et al. 1950, Kipnis et al. 1973, Miura et al. 2008, Wang et al. 2009)

Figure 6.20 compares catalytic carbonylation of nickel at 100 oC and CO pressure 14 atm with non-catalytic carbonylation at 80-100 oC and CO pressure 14 and 27 atm. Effect of sulphur on carbonylation of nickel was much stronger than effects of CO pressure or temperature. Catalytic carbonylation at 100 oC and CO pressure 14 atm was faster than non-catalytic carbonylation at CO pressure 27 atm. Compared with increasing temperature or carbon monoxide pressure (Figure 6.20), the addition of catalyst showed a better effect on nickel carbonylation in Figure 6.20. The addition of 0.1 – 0.5 wt% S in the form of FeS greatly increased the reaction rates (3.2-3.6 times) and shortened the reaction time to about 2 hours (it was near 5 hours in the non-catalytic carbonylation of nickel).

100

80

60 0.1 % 0.5 % 40 27 atm 14 atm

20 Extent of nickel carbonylation (%) 0 0 1 2 3 4 5 6 Time (hour)

130 100 0.1 % 0.5 % o 80 100 C 90 oC 80 oC 60

40

20 Extent of nickel carbonylation (%) 0 0 1 2 3 4 5 6 Time (hour)

Figure 6.20 Catalytic carbonylation of Ni at 100 oC and CO pressure 14 atm in comparison with non-catalytic carbonylation at CO pressure 14 atm (at 100 oC) and 27 atm (80-100 oC)

The increase in sulphur content from 0.1 to 7.0 wt% has a negative influence on the carbonylation reaction (Figure 5.24) what is consistent with data reported (Hieber et al. 1950, Kipnis et al. 1973).

Figure 6.21 shows the rates of nickel carbonylation catalyzed by different contents of sulphur introduced as elemental sulphur and iron sulphide. About 80% of nickel was carbonylated in the first hour of reaction, then the rate of carbonylation decreased as the mass of the sample decreased; in addition, diffusion through the ash layer started to influence the reaction rate by the end of the carbonylation reaction.

131 120 )

-1 0.1% 0.5% 100 hour  3.5% 7.0% 80 0.0%

60

40

20

Rate of nickel carbonylation (% (% carbonylation nickel of Rate FeS 0 )

-1 120 0.1% 0.5% hour  100 2.0% 7.0% 0.0% 80

60

40

20

Rate of nickel carbonylation (% (% carbonylation nickel of Rate S 0 0 1 2 3 4 5 6 Time (Hour)

Figure 6.21 Rates of catalytic nickel carbonylation at 100oC and CO pressure 14 atm

(flow rate 0.5 L·min-1) with different contents of sulphur in the form of elemental sulphur and iron sulphide

The catalytic effect depends on the concentration of the catalyst on the gas-solid interface rather than the content in the system. Concentration of sulphur at which the nickel-gas interface is covered by sulphur can be estimated by the following simplified analysis. The radius of non-bonded sulphur atom is 0.18nm (RSC 2012). Then the weight of a single saturated sulphur layer on 1 g sample can be calculated as follows (6- 11):

S m = BET *32.06 (6-11) s rN2 o

132 where, ms is the weight of sulphur used to saturate 1 g nickel sample surface; SBET is the

BET desorption surface area, m2·g-1; π is a mathematical constant, 3.1415926; r is the atom radius of sulphur; and No is Avogadro’s number, 6.023*1023.

The BET surface area of nickel reduced at 600 oC was 0.470 m2·g-1. After the reaction, the BET surface area of residual was 25.6 m2·g-1. The calculated weight of sulphur with

2 -1 -4 SBET = 25.6 m ·g is 2.45*10 g, hence nickel surface is saturated with sulphur at the

2.45*104 g following concentration: *100 0.0245% . 1g

The amount of sulphur added to nickel was more than 4 times of this theoretical value. With the increase of the catalysts contents, excess of sulphur is not involved into the catalytic reaction but is accumulated on the surface, decreasing the reaction rate.

XRD patterns of residual samples after catalytic carbonylation of nickel with different contents of elemental sulphur at 100 oC and CO pressure 14atm (3hours reaction) are shown in Figure 6.22 shows the XRD patterns of residual sample in sulphur catalysing reaction. Residues contained nickel oxide and nickel sulphides (NiS and Ni2S3), which were formed by reactions (6-12) and (6-13).

Ni + S = NiS (6-12)

2Ni + 3S = Ni23 S (6-13)

This XRD analysis confirmed that the increase in the content of sulphur and, therefore, nickel sulphide decreased the extent of nickel carbonylation.

133 NiO SiO2 12.0k NiS Ni2S3 Al(Al0.83Si1.08O4.85)

10.0k

--- 0.1%

8.0k

---0.5% 6.0k Counts (cts) Counts ---2.0%

4.0k ---5.0%

2.0k

---7.0% 10 20 30 40 50 60 70 80 90 100 Position (2 [o]) Figure 6.22 XRD patterns of residual samples after carbonylation of nickel catalysed by elemental sulphur (0.1 - 7.0 wt%) at 100 oC and CO pressure 14 atm (reaction time 3 hours)

6.6.3 Effect of Gas Flow Rate

The increase in the gas flow rate had a more profound effect on the rate of the catalytic carbonylation of nickel (Figure 5.29) in comparison with non-catalytic carbonylation (Figure 5.6). The effect of the gas flow rate on the rate of catalytic and non-catalytic carbonylation of nickel is shown in Figure 6.23. These data indicate the higher role of mass transfer of Ni(CO)4 in the gas phase in the rate control in catalytic carbonylation. Catalysis of the carbonylation process significantly increases the rate of chemical reaction, what increases role of diffusion in the rate control.

134 120

Non-catalytic Catalytic

) -1 -1 -1 100 0.50 Lmin 0.50 Lmin 0.35 Lmin-1 0.36 Lmin-1

hour -1 -1  0.23 Lmin 0.23 Lmin 80

60

40

20 Rate of nickel carbonylation (%

0 0 1 2 3 4 5 6 Time (Hour)

Figure 6.23 Rate of catalytic nickel carbonylation with addition of FeS (0.1 wt% S) at 100 oC and CO pressure 14 atm, and non-catalytic carbonylation at 100 oC and CO pressure 27 atm with different gas flow rates

6.6.4 Effect of CO Pressure

Increasing carbon monoxide pressure increased the rate of non-catalytic nickel carbonylation (Figures 5.10 and 6.10). However, the increase in CO gauge pressure from 14 to 27 atm decreased the rate of catalytic carbonylation of nickel (Figure 5.28). Effects of CO pressure on the extent and rate of non-catalytic and catalytic carbonylation of nickel at 100 oC are shown in Figure 6.24. In spite of a negative impact of increasing pressure on the catalytic carbonylation, the rate of the catalytic carbonylation at 27 atm was much faster than those at 14 and 27 atm in the non- catalytic carbonylation. When the carbon monoxide gauge pressure was increased from 14 atm to 27 atm, the gas velocity in the reactor reduced by about 46%; what is equivalent to the outlet gas flow rate of 0.26 L∙min-1 at 14 atm; The decrease in the gas velocity in the reactor decreased mass transfer coefficient in the gas phase with a negative effect on the reaction rate. As it was shown above, the mass transfer of

Ni(CO)4 in the gas phase can control or contribute to the rate control in the catalytic carbonylation of nickel.

135 100

75

50 of nickel carbonylation (%) 25 Extent

0

) 120

-1 14 atm Non-catalytic 27 atm Non-catalytic

 hour 100 14 atm Catalytic 27 atm Catalytic 80

60

40 nickel carbonylation (%

20 Rate of

0 0 1 2 3 4 5 Time (Hour)

Figure 6.24 Extent and rate of catalytic nickel carbonylation with addition of FeS (0.1 wt% S) and non-catalytic carbonylation at 100 oC and CO pressure 14 and 27 atm (gas flow rate 0.5 L·min-1)

6.7 Carbonylation of Selectively Reduced Laterite Ores

Laterite ore studied in the thesis had a particle size in the range from 53 to 200 µm (except of ore samples used in the study of the effect of laterite ore size), the average pore size in laterite ores was 7.65 nm (Table 6-1). The particle size of laterites was over 200 times of size of pure nickel particle, and the pore size was about four tenth of that of nickel oxide. The extraction of metal from laterite ore is much more complex. The following stages can be identified in the process of nickel carbonylation from the reduced laterite ore which are schematically shown in Figure 6.25: 1. Diffusion of

136 carbon monoxide from gas phase into the pore in ore particles. 2. Adsorption of carbon monoxide on metal surface. 3. Reaction of nickel with carbon monoxide. 4. Desorption of nickel carbonyl from metal surface. 5. Diffusion of nickel carbonyl to gas phase.

Figure 6.25 Schematic diagram of extraction of nickel from laterites by carbonylation

Considering the large particle size and small pore size, it can be suggested that the internal diffusion makes a significant contribution to the rate control of the carbonylation of nickel in laterite ores. Decrease in the ore size to < 53 µm resulted in the increase in the reduction rate (Yang 2014).

The nickel content of laterite is low (1.3 ~ 2.3 wt%); nickel is distributed in the particle interior. Addition of solid catalysts is inefficient (no contacts between nickel and sulphur); therefore in the catalytic carbonylation of selectively reduced laterite ore, sulphur was introduced in the form of hydrogen sulphide to carbon monoxide.

Effects of temperature on the non-catalytic and catalytic carbonylation of nickel in selectively reduced laterite ore is shown in Figure 6.26. Addition of catalysts had a minor effect on the rate of the carbonylation reaction. This confirms the intrinsic kinetics which was affected by the addition of the catalyst played a minor role in the reaction rate control.

The increase in the reaction rate in the non-catalytic (Figure 5.40) and catalytic (Figure 5.44) carbonylation with increasing temperature can be attributed to the increase in the diffusion of CO through the ore body to nickel. 137 50 Non-catalytic Catalytic 40

30

20

10 xtent of nickel carbonylation (%) E

0 ) -1 25

20

15

10

5 Rate of nickel carbonylation (%-hour

0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Time (Hour)

Figure 6.26 Extent and rate of non-catalytic and catalytic carbonylation of nickel in selectively reduced laterite ore at 100oC and CO pressure 27 atm

The effect of carbon monoxide flow rate on the non-catalytic and catalytic carbonylation of nickel and iron in selectively reduced laterite ores was not significant (Figures 5.35 and 5.40). This indicates that the external mass transfer did not contribute to the rate control. It can also be suggested that the increase of the outlet gas flow rate from 0.36 to 0.50 L·min-1 had a minor effect on the internal mass transfer, which played a significant role in the rate control.

Particle size of laterite ore had a significant effect on the nickel and iron carbonylation. The increase in laterite ore size resulted in the decrease of carbonylation rate of both nickel and iron (Figure 5.41). Results of BET analysis of MIN laterite ore which was

138 used in the study of the effect of ore size on the carbonylation process are shown in Table 6- 3. The pore size was in the range from 7.62 to 7.97 nm, with small difference between different size ranges. The BET surface area was 31.7 m2·g-1 for the ore with size 355-495 mμ, and about 40 m2·g-1 for other sizes. The pore length is proportional to the particle size, and the diffusion resistant increases with the pore length, which make the rate of carbonylation of both nickel and iron decreased with the increase of ore size.

Table 6-3 BET surface area and pore size of MIN ore with different particle size

Particle size, um BET surface area, m2·g-1 pore size, nm

38 - 53 42.44±0.22 7.65

75 - 90 38.02±0.19 7.62

140-200 40.74±0.23 7.87

355-495 31.73±0.13 7.97

139 Chapter 7 Conclusions and Recommendations for Further Work

7.1 Conclusions

Australia is among major suppliers and producers of nickel ore in the world. Processing of laterite ores, started in Australia in 1998, has significantly enhanced Australian role on the international nickel market. However, the production of nickel from oxide ores by established technologies consumes two to three times energy as processing of sulphide ores with significant environmental impact. This underlines the importance of development of more energy efficient processes for oxide ores. The project’s aim was to establish scientific fundamentals and technology feasibility of extraction of nickel from Australian laterite ore by carbonylation of selectively reduced nickel oxides.

The project studied carbonylation of Ni, Fe and Co reduced from oxides, and carbonylation of nickel and iron in the selectively reduced laterite ores.

Major findings in the study of non-catalytic and catalytic carbonylation reactions at different temperatures, carbon monoxide pressures, gas flow rates and particle sizes are summarized as follows.

Non-catalytic carbonylation of pure nickel was strongly affected by temperature, CO pressure, gas flow rate and particle size.

The extent of nickel carbonylation increased with temperature at constant CO pressure. Carbonylation of nickel at 100 oC and CO pressure 27 atm after 5.5 hours reaction was close to completion (extent of carbonylation was 99%). Effect of CO pressure on nickel carbonylation depended on temperature. At all temperatures, the effect of CO pressure was very strong when the gauge pressure was increased from 0 to 14 atm, and weak when CO gauge pressure changed from 27 to 54 atm.

The effect of particle size which was adjusted by changing temperature in reduction of nickel oxide from 500 to 900 oC, was strong when the mean particle size was above 1.1 140 μm (the reduction temperature was over 700 oC). The carbonylation rate increased with decreasing particle size.

Increase in the CO flow rate from 0.14 to 0.5 L·min-1 accelerated nickel carbonylation. The mass of nickel sample had no effect on the nickel carbonylation.

Carbonylation of pure iron was slow; the extent of iron carbonylation at 100 oC and CO pressure up to 55 atm (gauge) was less than 5.0%. The extent of cobalt carbonylation under these conditions was less than 0.5%.

Kinetic analysis of non-catalytic carbonylation of pure nickel indicated that the rate of the carbonylation had an intrinsic control at 80-89 oC. External mass transfer contributed to the rate control in carbonylation at 100 oC. Diffusion through the reacted (ash) layer also affected the reaction rate by the end of the carbonylation process.

Sulphur containing catalysts in the form of elemental sulphur and iron sulphide had a strong catalytic effect on carbonylation of pure nickel and iron. The extent of catalytic nickel carbonylation with addition of 0.1 wt% of sulphur reached over 95% in 3 hours; the extent of non - catalytic carbonylation under the same conditions was only 62% The rate of catalytic carbonylation decreased with the increasing content of catalyst.

Catalysts significantly accelerated the intrinsic kinetics, what weakened the effects of temperature and CO pressure on the rate of carbonylation, while the effect of CO flow rate became stronger in comparison with the non-catalytic carbonylation.

The extent of non-catalytic carbonylation of nickel from the selectively reduced BCS laterite ore at 100 oC and CO pressure 41 atm was below 50%. The use of catalysts in the carbonylation of selectively reduced ore was inefficient. Laterite ores had a higher particle size and smaller porosity in comparison with nickel oxides studied in the project. Concentration of nickel in the MIN ore was 1.7-2.4 wt% (depending on the ore size). The major parameter affecting the rate of ore carbonylation was the particle size. The rate of reaction increased significantly with decreasing particle size; the carbonylation of nickel in MIN ore with the particle size 38-53 and 75-90 µm, was close to completion after 4 hours reaction. This indicates that the major rate controlling stage in the carbonylation of selectively reduced laterite ores was diffusion of CO to nickel dispersed in the ore body.

141 Based on the experimental results, the extraction of nickel from selectively reduced fine laterite ores by carbonylation was over 90%, which shows a feasibility of the proposed technology for nickel extraction from laterite ores.

Results of a systematic study of the carbonylation of nickel and selectively reduced laterite ores are significant for further understanding of carbonylation reactions. Promising results were obtained for further development of technology of extraction of nickel from laterite ores by the carbonylation process.

7.2 Recommendations for Further Work

Kinetic analysis of carbonylation reaction included consideration of separate reaction stages in the non-catalytic carbonylation of nickel. Comprehensive analysis based on the kinetic modelling was beyond the scope of this project. Kinetic modelling of non- catalytic and catalytic carbonylation of nickel can be recommended for further work.

The project convincingly demonstrated a strong effect of the ore particle size on the nickel carbonylation. Carbonylation of nickel from the selectively reduced laterite ore with size 38-53 and 75-90 µm was close to completion. A more detailed study is recommended to develop a technology for extraction of nickel by carbonylation of selectively reduced fine laterite ores, including fluid bed reactor.

The project has confirmed that under given experimental conditions carbonylation of iron and cobalt is slow. Thermodynamic analysis showed that cobalt which is a highly valuable metal can be extracted in the form of volatile cobalt nitrosyl carbonyl. A study of formation of cobalt nitrosyl carbonyl is suggested for the further work.

Carbonylation of iron requires higher temperatures and CO pressures than those employed in the project. A study of carbonylation of iron under these conditions is also recommended for the further work.

142 Reference

Abel, E. W. (1963). "The metal carbonyls." Quarterly Reviews, Chemical Society 17(2): 133-159.

Andersen, J. C., G. K. Rollinson, B. Snook, R. Herrington and R. J. Fairhurst (2009). "Use of QEMSCAN® for the characterization of Ni-rich and Ni-poor goethite in laterite ores." Minerals Engineering 22(13): 1119- 1129.

Berger, V. I., D. A. Singer, J. D. Bliss and B. C. Moring (2013). Ni-Co Laterite Deposits of the World: Database and Grade and Tonnage Models, BiblioGov.

Blanchard, A. A. and P. Gilmont (1940). "Preparation of Cobalt Carbonyl, Cobalt Nitrosyl Carbonyl and Cobalt Carbonyl Hydride by the Cyanide Method." Journal of the American Chemical Society 62(5): 1192-1193.

Bo, L. D. Y. (2009). "Research Progress in Beneficiation of Copper-nickel Sulfide Ore [J]." Modern Mining 8: 006.

Boldt, J. R. (1957). The winning of nickel, Taylor & Francis.

Brockway, L. O. and J. S. Anderson (1937). "The molecular structures of iron nitrosocarbonyl Fe (NO)(2) (CO)(2) and cobalt nitrosocarbonyl CO (NO)(CO)(3)." Transactions of the Faraday Society 33(2): 1233-1238.

Cao, Y. (2005). "Status quo and prospects of nickel industry at home and abroad." World Nonferrous Metals(10): 5.

143 Chemlink, P.-L. (1997). "Nickel in Australia." 1997, from http://www.chemlink.com.au/nickel.htm.

Chen, T., J. Dutrizac, E. Krause and R. Osborne (2004). Mineralogical characterization of nickel laterites from New Caledonia and Indonesia. International Laterite Nickel Symposium 2004(as held during the 2004 TMS Annual Meeting).

Christian Redl, M. P., Roland Kristl, Siegfried Fritsch, Heinz Muller (2013). REFINING OF FERRONICKEL. The thirteenth International Ferroalloys Congress Efficient technologies in ferroalloy industry. Almaty, Kazakhstan: 229-236.

Coleman, G. W. and A. A. Blanchard (1936). "Preparation and properties of cobalt nitrosyl carbonyl and of cobalt carbonyl hydride." Journal of the American Chemical Society 58: 2160-2163.

Coto, O., F. Galizia, I. Hernández, J. Marrero and E. Donati (2008). "Cobalt and nickel recoveries from laterite tailings by organic and inorganic bio-acids." 94(1–4): 18-22.

Day, J. P., F. Basolo and R. G. Pearson (1968). "Kinetics and mechanism of nucleophilic substitution in nickel tetracarbonyl." Journal of the American Chemical Society 90(25): 6927-6933.

De Groot, P., M. Coulon and K. Dransfeld (1980). "Ni (CO) 4 formation on single Ni crystals: Reaction kinetics and observation of surface facetting induced by the reaction." Surface Science 94(1): 204-220.

Den Besten, I. E. and P. W. Selwood (1962). "The chemisorption of , methyl sulfide, and cyclohexene on supported nickel catalysts." Journal of Catalysis 1(2): 93-102. 144 Dewar, J. and H. O. Jones (1905). "The physical and chemical properties of iron carbonyl." Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 76(513): 558- 577.

Dewar, J. and H. O. Jones (1907). "On a New Iron Carbonyl, and on the Action of Light and of Heat on the Iron Carbonyls." Proceedings of the Royal Society of London. Series A 79(527): 66-80.

Ewens, R. V. G. and M. W. Lister (1939). "The structure of iron pentacarbonyl, and of iron and cobalt carbonyl hydrides." Transactions of the Faraday Society 35(1): 0681-0690.

F.Habashi (1997). Handbook of , Wiley-VCH.

Germer, L. and A. MacRae (1962). "Oxygen‐Nickel Structures on the (110) Face of Clean Nickel." Journal of Applied Physics 33(10): 2923- 2932.

Goldberger, W. M. and D. F. Othmer (1963). "Kinetics of Nickel Carbonyl Formation." Industrial & Engineering Chemistry Process Design and Development 2(3): 202-209.

Graaf, D. and JE (1979). "The treatment of lateritic nickel ores—a further study of the Caron process and other possible improvements. Part I. Effect of reduction conditions." Hydrometallurgy 5(1): 47-65.

Guo, X.-y., Z. WU and D. LI (2009). "The Status Quo and Prospect of the Technologies of Nickel Laterites [J]." Metal Materials and Metallurgy Engineering 2: 002.

145 Habashi, F. (1969). Principles of extractive metallurgy, CRC Press. 1: 140- 165.

Heinicke, G., N. Bock and H. Harenz (1970). "Chemische Beeinflussung mechanisch aktivierter Reaktionen. III. Zum Mechanismus der tribomechanisch aktivierten Metallcarbonylbildung unter Einfluß schwefelhaltiger Substanzen." Zeitschrift für anorganische und allgemeine Chemie 372(2): 162-170.

Heinicke, G. and H. Harenz (1963). "Chemische Beeinflussung mechanisch aktivierter Reaktionen. II. Chemische Aktivierung der mechanisch angeregten Nickel- und Eisencarbonylbildung." Zeitschrift für anorganische und allgemeine Chemie 324(3-4): 185-196.

Herrmann, W. A. (1990). "100 Years of Metal-Carbonyls - a Serendipitous Chemical Discovery of Major Scientific and Industrial-Impact." Journal of 383(1-3): 21-44.

Herrmann, W. A. (1990). "100 years of metal carbonyls: a serendipitous chemical discovery of major scientific and industrial impact." Journal of Organometallic Chemistry 383(1): 21-44.

Hieber, W. and O. Geisenberger (1950). "Über Metallcarbonyle. XLVII. Über den Einfluß von Chalkogenen auf die Entstehung von Eisenpentacarbonyl aus den Komponenten." Zeitschrift für anorganische Chemie 262(1‐5): 15-24.

Huang, Q. (1990). Gasification Metallurgy of Nickel. Metallurgy of nickel, China Science and Techn Press: 174-185.

146 Igharo, M. and J. Wood (1985). "Compaction and Sintering Phenomena in Titanium—Nickel Shape Memory Alloys." Powder metallurgy 28(3): 131- 139.

INSG-1. (2013). "INSG April 2013 Meetings Press Release." from http://www.insg.org/docs/INSG_Press_Release_April_13.pdf.

INSG-2. (2013). "About nickel - Production, usage and prices." from http://www.insg.org/prodnickel.aspx.

INSG-3. (2015). "Press release INSG April 2015 Meeting." from http://www.insg.org/%5Cdocs%5CINSG_Press_Release_April_2015.pdf.

Ivanova, A. R., G. Nuesca, X. Chen, C. Goldberg, A. E. Kaloyeros, B. Arkles and J. J. Sullivan (1999). "The effects of processing parameters in the chemical vapor deposition of cobalt from cobalt tricarbonyl nitrosyl." Journal of The Electrochemical Society 146(6): 2139-2145.

Janković, B., B. Adnađević and S. Mentus (2008). "The kinetic study of temperature-programmed reduction of nickel oxide in hydrogen atmosphere." Chemical Engineering Science 63(3): 567-575.

Jeangros, Q., T. W. Hansen, J. B. Wagner, C. D. Damsgaard, R. E. Dunin- Borkowski, C. Hébert and A. Hessler-Wyser (2013). "Reduction of nickel oxide particles by hydrogen studied in an environmental TEM." Journal of Materials Science 48(7): 2893-2907.

Jiang, J.-b. and J.-k. WANG (2009). "Review on Progresses for Hydrometallurgy of Laterite-nickel Ore " Hydrometallurgy of China 1.

147 Jiang, J. and H. Wang (2004). "Pressure Hydrometallurgy: Processing and Utilization of Resources for Sustainable Development Technology." CHINA VENTURE CAPITAL & HIGH-TECH (12): 73-75.

Jozwiak, W., E. Kaczmarek, T. Maniecki, W. Ignaczak and W. Maniukiewicz (2007). "Reduction behavior of iron oxides in hydrogen and carbon monoxide atmospheres." Applied Catalysis A: General 326(1): 17- 27.

Kar, R. N., L. B. Sukla, K. M. Swamy, V. V. Panchanadikar and K. L. Narayana (1996). "Bioleaching of lateritic nickel ore by ultrasound." Metallurgical and Materials Transactions B 27(3): 351-354.

Kawasaki, E., J. Sanscrainte and T. J. Walsh (1962). "Kinetics of reduction of iron oxide with carbon monoxide and hydrogen." AIChE Journal 8(1): 48-52.

Kipnis, A. Y., N. F. Mikhailova and V. G. Romu (1973). "Influence of the composition and structure of materials on the behavior of subordinate amounts of cobalt during carbonylation of nickel-containing materials." Zhurnal Prikladnoi Khimii 47(7): 1566-1570.

Kozyrev, V. F., R. A. Shvartsman and A. S. Mnukhin (2005). "On the Mechanism of Nickel Recovery from the Sulfide Component in the Ni-Cu- S System in Carbonylation." Russian Journal of Applied Chemistry 78(12): 1921-1924.

Kozyrev, V. F., R. A. Shvartsman and A. S. Mnukhin (2005). "A Study of the Behavior of Cobalt in Carbonylation of Copper-Nickel Metallurgical Half-Products at Medium Pressure of the Reaction Gas." Russian Journal of Applied Chemistry 78(12): 1925-1929.

148 Laidler, K. J. (1984). "The development of the Arrhenius equation." Journal of Chemical Education 61(6): 494.

Lascelles, K. and L. Renny (1983). "The rate of formation of nickel carbonyl from carbon monoxide and nickel single crystals." Surface Science Letters 125(1): L67-L72.

Levenspiel, O. (1972). Chemical reaction engineering, Wiley New York etc. 2: 357-372.

LI, Q.-h., W. Juan and L. Zhi-hong (2009). "Development of Exploitation and Hydrometallurgical Processes of Nickel Laterite Resources in the World." Hunan Nonferrous Metals(2): 4.

Li, Q.-h., J. Wang and Z.-h. Liu (2009). "Development of Exploitation and Hydrometallurgical Processes of Nickel Laterite Resources in the World [J]." Hunan Nonferrous Metals 2: 005.

Li, Y., H. Yu, D. Wang, W. Yin and Y. Bai (2010). "The Current Status of Laterite Nickel Ore Resources and its Processing Technology." Metal Mine(413): 5-9.

Li, Z.-m., Z. Tong and W. Jia-zheng (2009). "Using of the nickel resource and development of the ferronickel industry." China Nonferrous Metallurgy(1): 4.

Li, Z., Z. Youping, Z. Yusheng and L. Weiguo (2007). "CHARACTERISTICS OF LATERITE RESOURCE AND ANALYSIS ON ITS PYROMETALLURGY PROCESS [J]." Ferro-Alloys 6: 010.

149 Liang, D., G. Abend, J. Block and N. Kruse (1983). "Formation of nickel subcarbonyls from nickel and carbon monoxide." Surface Science 126(1): 392-396.

Liang, D. and Y. Bo (2009). "Research Progress in Beneficiation of Copper-nickel Sulfide Ore." Modern Mining(8): 3.

Liang, D. B., G. Abend, J. H. Block and N. Kruse (1983). "Formation of nickel subcarbonyls from nickel and carbon monoxide." Surface Science 126(1–3): 392-396.

Lin, H.-Y., Y.-W. Chen and C. Li (2003). "The mechanism of reduction of iron oxide by hydrogen." Thermochimica Acta 400(1): 61-67.

Liu, D. (2002). "Recent development in nickel and cobalt recovery technologies from laterite." Nonferrous Metals (Extractive Metallurgy) 3: 6-10.

Mazurek, H., R. Mehta, M. Dresselhaus, G. Dresselhaus and H. Zeiger (1982). "The nature of the initial transient in the rate of formation of nickel carbonyl." Surface Science 118(3): 530-534.

McDonald, R. and B. Whittington (2008). "Atmospheric acid leaching of nickel laterites review. Part II. Chloride and bio-technologies." Hydrometallurgy 91(1): 56-69.

McDonald, R. and B. Whittington (2008). "Atmospheric acid leaching of nickel laterites review: Part I. Sulphuric acid technologies." Hydrometallurgy 91(1): 35-55.

150 Medvedev, V., R. Börner and N. Kruse (1998). "Nickeltetracarbonyl formation on non-equilibrium Ni surfaces." Surface science 401(1): L371- L374.

Miner, T. E. P. R. (2015). "Operations update." from http://www.axiom- mining.com/IRM/Company/ShowPage.aspx/PDFs/2200- 10000000/AxiomMining2015AGMPresentation.

Miura, K., M. Itoh and K.-I. Machida (2008). "Extraction and recovery characteristics of Fe element from Nd–Fe–B sintered magnet powder scrap by carbonylation." Journal of Alloys and Compounds 466(1): 228-232.

Mond, L., H. Hirtz and D. Cowap (1908). "Note on A Volatile Compound of Cobalt with Carbon Monoxide." Chemical Reviews 98: 2.

Mond, L., H. Hirtz and D. Cowap (1908). "Note on a Volatile Compound of Cobalt with Carbon Monoxide." Chem. News 98: 165.

Mond, L. and C. Langer (1891). "XCIII. - On iron carbonyls." Journal of the Chemical Society, Transactions 59: 1090.

Mond, L. and C. Langer (1891-1). "XCIII.-On iron carbonyls." Journal of the Chemical Society, Transactions 59(0): 1090-1093.

Mond, L., C. Langer and F. Quincke (1890). "L.- Action of carbon monoxide on nickel." Journal of the Chemical Society, Transactions 57: 749.

Mond, L., C. Langer and F. Quincke (1890). "L.—Action of carbon monoxide on nickel." Journal of the Chemical Society, Transactions 57: 749-753.

151 Mond, L. and F. Quincke (1891). "LV.- Note on a volatile compound of iron with carbonic oxide." Journal of the Chemical Society, Transactions 59: 604.

Mond, L. and F. Quincke (1891). "LV.-Note on a volatile compound of iron with carbonic oxide." Journal of the Chemical Society, Transactions 59(0): 604-607.

Mond, R. L. and A. E. Wallis (1922). "V.-Researches on the metallic carbonyls." Journal of the Chemical Society, Transactions 121(0): 29-32.

Moskalyk, R. and A. Alfantazi (2002). "Nickel laterite processing and electrowinning practice." Minerals Engineering 15(8): 593-605.

Mudd, G. M. (2009). Nickel sulfide versus laterite: the hard sustainability challenge remains. 48th Conference of Metallurgists and Nickel & Cobalt.

Naldrett, A. (1999). "World-class Ni-Cu-PGE deposits: key factors in their genesis." Mineralium deposita 34(3): 227-240.

Nametkin, N. S., V. D. Tyurin, A. I. Nekhaev, M. Mavlonov and M. A. Kukina (1975). "Direct replacement of carbonyl in iron carbonyls by sulfur." Bulletin of the Academy of Sciences of the USSR, Division of chemical science 24(12): 2739-2739.

Nicol, M., A. Nikoloski and J. Fittock (2004). "A fundamental study of the leaching reactions involved in the Caron process."

Norgate, T. and S. Jahanshahi (2011). "Assessing the energy and greenhouse gas footprints of nickel laterite processing." Minerals Engineering 24(7): 698-707.

152 Oliveira, S. M. B. d., C. S. d. M. Partiti and J. Enzweiler (2001). "Ochreous laterite: a nickel ore from Punta Gorda, Cuba." Journal of South American Earth Sciences 14(3): 307-317.

Oxley, A. and N. Barcza (2013). "Hydro–pyro integration in the processing of nickel laterites." Minerals Engineering.

Parravano, G. (1952). "The reduction of nickel oxide by hydrogen." Journal of the American Chemical Society 74(5): 1194-1198.

Peng, R. (2005). Nickel Metallurge, Central South University Press

Peng, R. (2005). Nickel Metallurgy, Central South University Press.

Pineau, A., N. Kanari and I. Gaballah (2006). "Kinetics of reduction of iron oxides by H 2: Part I: Low temperature reduction of hematite." Thermochimica acta 447(1): 89-100.

Podall, H. E. (1961). "Recent developments in synthesis." Journal of Chemical Education 38(4): 187-null.

Princeton. (2014). "Aqua Regia ", from http://web.princeton.edu/sites/ehs/labsafetymanual/cheminfo/aquaregia.htm.

Qiu, S., R. Ohnishi and M. Ichikawa (1992). "Novel preparation of gold (I) carbonyls and nitrosyls in NaY zeolite and their catalytic activity for NO reduction with CO." J. Chem. Soc., Chem. Commun.(19): 1425-1427.

Rashed, A. and Y. Rao (1997). "Kinetics of Reduction of Nickel Oxide with Hydrogen Gas in the 230-452 C Range." Chemical Engineering Communications 156(1): 1-30.

Redmon, L. (1980). "kinetics of Ni(CO)4 formation."

153 Rigin, V. (1993). "Simultaneous atomic fluorescence spectrometric determination of traces of iron, cobalt and nickel after conversion to their carbonyls and gas-phase atomization by microwave-induced plasma." Analytica Chimica Acta 283(2): 895-901.

Rodriguez, J. A., J. C. Hanson, A. I. Frenkel, J. Y. Kim and M. Pérez (2002). "Experimental and theoretical studies on the reaction of H2 with NiO: role of O vacancies and mechanism for oxide reduction." Journal of the American Chemical Society 124(2): 346-354.

RSC. (2012). "Periodic Table - sulphur." from http://www.rsc.org/periodic- table/element/16/sulfur.

Sastri, M., R. Viswanath and B. Viswanathan (1982). "Studies on the reduction of iron oxide with hydrogen." International Journal of Hydrogen Energy 7(12): 951-955.

Schmidt, W. A., J. H. Block and K. A. Becker (1982). "The Topography of Nickel Crystals during the Reaction Towards Ni(Co)4 - a Field- Microscope Study." Surface Science 122(3): 409-421.

Sehested, J., A. Carlsson, T. V. W. Janssens, P. L. Hansen and A. K. Datye (2001). "Sintering of Nickel Steam-Reforming Catalysts on MgAl2O4 Spinel Supports." Journal of Catalysis 197(1): 200-209.

Statista-a. (2015). "Major countries in worldwide nickel mine production from 2010 to 2014 (in metric tons) ", from http://www.statista.com/statistics/264642/nickel-mine-production-by- country/.

Statista-b (2015). "World mine reserves of gold as of 2014, by country (in metric tons)." 154 Sugimoto, S., T. Maeda, D. Book, T. Kagotani, K. Inomata, M. Homma, H. Ota, Y. Houjou and R. Sato (2002). "GHz microwave absorption of a fine α-Fe structure produced by the disproportionation of Sm2Fe17 in hydrogen." Journal of Alloys and Compounds 330–332(0): 301-306.

Taylor, A. (2013). "Laterites - Still A Frontier of Nickel Process Development." from http://www.altamet.com.au/wp- content/uploads/2013/04/Laterites-Still-a-Frontier-of-Nickel-Process- Developoment1.pdf.

Teng-1, R. (2006). "Development trend of carbonyl nickel refining technology in China [J]." China Nonferrous Metallurgy 3: 005.

Teng, R.-h., C. Qu-shan, L. Si-lin and L. Xue-quan (2007). "INFLUENCE OF SULPHIDE ON CARBONYLATION OF CU-NI ALLOY." Powder Metallurgy Industry 17(7): 6.

Teng, R. (2006). "Nickel Refining by Optimized Carbonization in China." POWDER METALLURGY INDUSTRY 16(3): 30.

Terekhov, D. (2001). CARBONYL PROCESS FOR RECOVERY OF PURIFIED COBALT, WO Patent 2,001,055,463.

Terekhov, D. (2001). CARBONYL PROCESS FOR RECOVERY OF PURIFIED COBALT, WO Patent 2,001,055,463.

Terekhov, D. S. (2001). Carbonyl process for recovery of purified cobalt, Google Patents.

Thubakgale, C., R. Mbaya and K. Kabongo (2013). "A study of atmospheric acid leaching of a South African nickel laterite." Minerals Engineering.

155 Turkdogan, E. and J. Vinters (1971). "Gaseous reduction of iron oxides: Part I. Reduction of hematite in hydrogen." Metallurgical and Materials Transactions B 2(11): 3175-3188.

USGS. (2009). "2009 Minerals Yearbook of Cobalt." from http://minerals.usgs.gov/minerals/pubs/commodity/cobalt/myb1-2009- cobal.pdf.

Utigard, T., M. Wu, G. Plascencia and T. Marin (2005). "Reduction kinetics of Goro nickel oxide using hydrogen." Chemical Engineering Science 60(7): 2061-2068.

Valix, M., J. Tang and W. Cheung (2001). "The effects of mineralogy on the biological leaching of nickel laterite ores." Minerals Engineering 14(12): 1629-1635.

Valix, M., F. Usai and R. Malik (2001). "Fungal bio-leaching of low grade laterite ores." Minerals Engineering 14(2): 197-203.

Wagner, D., O. Devisme, F. Patisson and D. Ablitzer (2008). "A laboratory study of the reduction of iron oxides by hydrogen." arXiv preprint arXiv:0803.2831.

Wang, C., Y. Fei, C. Yongqiang, W. Zhong and W. Jun (2008). "Worldwide processing technologies and progress of nickel laterites." The Chinese Journal of Nonferrous Metals 18(1): s1-s8.

Wang, D., A. Chen, Z. Zhao and G. Huo (2009). "Extraction of nickel from nickel alloy by carbonylation." Transition Metal Chemistry 34(3): 313-316.

156 Warner, A., C. Díaz, A. Dalvi, P. Mackey and A. Tarasov (2006). "JOM world nonferrous smelter survey, Part III: Nickel: Laterite." JOM 58(4): 11-20.

Wells, M., C. Butt, I. Robertson, K. Scott and M. Cornelius (2003). "Murrin Murrin nickel laterite deposit, WA." CRC LEME, Canberra, Australia.

Whittington, B., J. Johnson, L. Quan, R. McDonald and D. Muir (2003). "Pressure acid leaching of arid-region nickel laterite ore: Part II. Effect of ore type." Hydrometallurgy 70(1): 47-62.

Whittington, B. I. and D. Muir (2000). "Pressure Acid Leaching of Nickel Laterites: A Review." Mineral Processing and Extractive Metallurgy Review 21(6): 527-599.

Windsor, M. M. and A. A. Blanchard (1933). "Nickel Carbonyl. A Study of the Mechanism of its Formation from Nickel Sulfide and Carbon Monoxide1." Journal of the American Chemical Society 55(5): 1877-1883.

Xu, Q.-x. (2005). "The past and the future of nickel laterites." China Nonferrous Metallurgy(12): 8.

Yang, J. (2014). Selective Reduction of Laterite Ore. Ph.D, University of New South Wales.

Zhai, X., Y. Fu, X. Zhang, L. Ma and F. Xie (2009). "Intensification of sulphation and pressure acid leaching of nickel laterite by microwave radiation." Hydrometallurgy 99(3): 189-193.

157 Zhai, Y.-C., W.-N. Mu, Y. Liu and Q. Xu (2010). "A green process for recovering nickel from nickeliferous laterite ores." Transactions of Nonferrous Metals Society of China 20: s65-s70.

Zhao, J., Z. Sun and P. Zheng (2012). "Review of Nickel Laterite Ore Processing Method." NON-FERROUS MINING AND METALLURGY 28(6): 39-45.

Zhou, Q. (2005). "Current Situation and Development Direction of Technology for Oxidized Nickel Ore Treatment." YUNNAN METALLURGY 34(6): 33-36.

Zhou, Q. x. (2005). "Current Situation and Development Direction of Technology for Oxidized Nickel Ore Treatment." Yunnan Metallurgy 34(6): 4.

158 Appendix:

Appendix A: Calibration of gas flow rate:

2500 N2 ( a )

H2

) 2000 -1 y = 38.98x + 82.69

min Adj. R-Square: 0.9953  1500

1000

Flow rate (mL Flow rate y = 11.54x + 134.81 500 Adj. R-Square: 0.9978

0 0 20 40 60 80 100 Scale 1400 CO ( b ) 1200 CO2

) y = 11.46x + 131.12 -1 1000 Adj. R-Square: 0.99688 min  800

600

400 Flow rate (mL Flow rate y=32.33 + 6.24x - 0.0068x2 200 Adj. R-Square: 0.9991

0 0 20 40 60 80 100 120 140 160 Scale

159 1800 ( C )

1500 y = 37.60195x + 82.70857

) Adj. R-Square: 0.98839 -1 1200 min 

900 y = 11.5265x + 128.9925 Adj. R-Square: 0.9956 600 Flow rate (mL Flow rate CO 300 y = 11.17216x + 145.34224 Adj. R-Square: 0.99684 N2 H2 0 0 15 30 45 60 75 90 105 Scale Appendix A. Calibration of gas flow meter (a & b gas flow meter for reduction. c. gas flow meter for carbonylation)

Appendix B. Calibration of reduction furnace:

900

(19, 768.7) (25, 777.8) (31, 769.6) 750 800℃

(20, 573.6) (26, 581.8) ) 600 (32, 572.5) ℃

600℃ 450 (19, 375.8) (26, 385) (33, 374.6)

400℃

Temperature ( 300 (20, 272.7) (27, 281.5) (33, 273.5)

300℃ 150

0 10 20 30 40 50 Distance (cm)

160 800 28 cm

C) 700 o

600

500

400

T = 0.9876 ts - 12.9186 Temperature measured ( 300 Adj. R-Squar: 0.99988 200 300 400 500 600 700 800 Setup temperature (oC)

800 24 cm

700 C) o 600

500

400

T = 0.99153 ts - 14.8508 300 Adj. R-Squar: 0.99985 Temperature measured (

200 300 400 500 600 700 800 Setup temperature (oC)

Appendix B. Calibration of the isothermal zone and correlation between setup temperature and isothermal temperature

Appendix C. Calibration of oil bath:

Appendix C.1 Calibration of the temperature (oC) in oil bath

Set point 48.0 55.0 60.0 65.0 70.0 75.0 80.0 85.0 90.0

Oil 47.8 54.5 59.8 64.8 69.9 75.3 80.1 85.3 90.3 Reaction 48 53 58 62 66 70 74 77 81 Tube

161 50

40 )

C Oil Bath Display o Oil Tube 30 Set Point Temperature ( 20

10 0 15 30 45 60 75 90 Operation time (Minutes)

Appendix C.2 Calibration of temperature of the oil bath

Appendix D. Selection of absorption solution

0.06 0.25 Bottle I a. Bottle I b. Bottle II Bottle II 0.20 Bottle III 0.04 0.15

0.10 0.02 Ni Recovery (g)

Ni Recovery (g) 0.05

0.00 0.00 0 2 4 6 8 10 0 2 4 6 8 10

Time (Hour) Time (Hour) 0.5 0.6 Bottle I c. Bottle I d. 0.4 Bottle II 0.5 Bottle II Bottle III Bottle III 0.4 0.3 0.3 0.2 0.2 Fe Recovery (g) 0.1 Fe Recovery (g) 0.1

0.0 0.0 0 2 4 6 8 10 0 2 4 6 8 10 Time (Hour) Time (Hour)

Appendix D. Absorption of nickel and iron carbonyls in nitric acid and aqua regia (a and c: 50% - nitric acid solution, b and d: aqua regia)

162 Appendix E. Statistics of nickel particles reduced at different temperature

100 100 o 500 C 600oC 80 80

60 60

40 40 Percentage (%) 20 Percentage (%) 20

0 0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 Particle size (m) Particle size (m)

100 100 700oC 800oC 80 80

60 60

40 40 Percentage (%) Percentage (%) 20 20

0 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Particle size (m) Particle size (m) 100 900oC 80

60

40

Percentage (%) 20

0 0 1 2 3 4 5 Particle size (m) Appendix E. Statistical chart of nickel particles reduced at different temperature

Appendix F: Reaction Rate Control in the Non- catalytic Carbonylation

The analysis of the effect of temperature on the rate of carbonylation in Section 6.2.1 has demonstrated that the intrinsic kinetics was the controlling stage or contributed to 163 the carbonylation rate control (the rate constant was described by the Arrhenius equation). Nickel oxide used in experiments was of high purity (> 99%), and non- porous solid. Nickel particles can be considered as spheres. In the process of carbonylation, gaseous nickel carbonyl was formed what led to the decreasing size of unreacted nickel. For kinetic analysis of nickel carbonylation, the shrinking core model can be applied.

The rate of nickel carbonylation can be controlled by one of the following three stages or have a mixed control by two or all three stages: (1) mass transfer in the gas phase which can be described by diffusion through the gaseous film (boundary layer) surrounding solid particles, (2) diffusion through the reacted (ash) solid layer and (3) chemical reaction on the surface of the unreacted core (Habashi 1969, Levenspiel 1972).

When the reaction rate is controlled by the diffusion through the boundary layer (film), the extent of reaction as a function of time is described by the equation (F-1). Rate controls by the intrinsic kinetics and diffusion through the ash layer are described by Equation (F-2) and (F-3) respectively.

X=kt (F-1)

kt = 1 – (1 – X)1/3 (F-2)

kt = 1 - 3(1-X)2/3 + 2(1-X) (F-3) where k = reaction rate constant (min-1); t = time in minutes;

X = extent of nickel carbonylation.

Functions X, 1 – (1 – X)1/3 and 1 - 3(1-X)2/3 + 2(1-X) at different temperatures in the first 2.5 hours carbonylation at CO pressure 55 atm are drawn in Table F1. Table F2 in this appendix presents the plots of these functions in the last 2.5 hours of nickel carbonylation, from 4 to 6.5 hours. Plots of 1 – (1 – X)1/3 vs time and X vs time are close to linear functions at all temperatures but 100 oC. Departure of plots of 1 - 3(1-X)2/3 + 2(1-X) vs time from a linear function was more significant in the first 2.5 hours of carbonylation reaction. It can be concluded that the carbonylation of nickel at

164 temperatures 80 to 95 oC was controlled by the intrinsic kinetics, and the diffusion through the gas film. The diffusion through the ash layer started to influence the carbonylation with the increase of the extent of nickel carbonylation, and therefore, increasing thickness of the ash layer.

Significant deviations of plots for the diffusion through the gas film and chemical reaction control were observed for carbonylation at the last stage at 100 oC. The best fit to the linear function was observed for the diffusion through the ash layer. This simplified analysis is not sufficient to establish the rate controlling stages for carbonylation of nickel at 100 oC after 2.5 hours reaction. A more comprehensive kinetic analysis can be done on the basis of kinetic modelling which was beyond the scope of this project.

100 R2 100 oC 1.0000 80 95 oC 0.9975 90 oC 0.9972 85 oC 0.9724 60 80 oC 0.9993 (%)

X 40

20

0 0.0 0.5 1.0 1.5 2.0 2.5 Time (Hour)

165 0.20 2 80 oC R o 85 oC 80 C 0.8691 o 90 oC 85 C 0.9411 o o 90 C 0.8683 0.15 95 C o o 95 C 0.8571 100 C 100 oC 0.7912

0.10

1-3(1-X)^(2/3)+2(1-X) 0.05

0.00 0.0 0.5 1.0 1.5 2.0 2.5 Time (Hour)

0.30 2 o R 80 C o o 80 C 0.9998 85 C o 0.25 o 85 C 0.9809 90 C 90 oC 0.9995 o 95 C 95 oC 1.0000 o o 0.20 100 C 100 C 0.9858

0.15

1-(1-X)^(1/3) 0.10

0.05

0.00 0.0 0.5 1.0 1.5 2.0 2.5 Time (Hour)

Table F1 Plots of X for gas film control, 1 – (1 – X)1/3 for the intrinsic control and 1 - 3(1-X)2/3 + 2(1-X) for the diffusion control in the first 2.5 hours carbonylation of Ni at 80 - 100 oC and CO pressure 55 atm with gas flow rate 0.5 L∙min-1

166 100 R2 95oC 0.9776 o 80 95 C 0.9969 95oC 0.9975 95oC 0.9862 100 oC 0.8815 60 X 40

20

0 0 1 2 3 4 5 6 7 Time (Hour)

0.7 R2 0.6 80 oC 0.7263 85 oC 0.8928 0.5 90 oC 0.8897 95 oC 0.9217 0.4 100 oC 0.9692

0.3

0.2 1-3(1-X)^(2/3)+2(1-X) 0.1

0.0 0 1 2 3 4 5 6 7 Time (Hour)

0.7 R2 0.6 80 oC 0.9648 85 oC 0.9972 0.5 90 oC 0.9942 95 oC 0.9984 0.4 100 oC 0.9375

0.3

1-(1-X)^(1/3) 0.2

0.1

0.0 0 1 2 3 4 5 6 7 Time (Hour)

Table F2 Plots of 1 – (1 – X)1/3 for the intrinsic control and 1 - 3(1-X)2/3 + 2(1-X) for the diffusion control in the last 2.5 hours carbonylation of Ni at 80 - 100 oC and CO pressure 55 atm with gas flow rate 0.5 L∙min-1

167