Download as pdf or txt
Download as pdf or txt
You are on page 1of 376

CHEMICAL AND ELECTROCHEMICAL

LEACHING STUDIES OF SYNTHETIC AND


NATURAL ILMENITE IN HYDROCHLORIC
ACID SOLUTIONS

By
NURUL AIN JABIT
B. Eng. (Mineral Resources Engineering)

2017

This thesis is presented for the degree of Doctor of Philosophy


at
Murdoch University
I declare that this thesis is my own account of my research and contains as its

main content work which has not previously been submitted for a degree at any

tertiary education institution.

Nurul Ain Jabit


ABSTRACT

The research interest for upgrading ilmenite to synthetic rutile has increased

over the years due to the increasing demand for titanium dioxide white pigment and

titanium metal. Sulfate and chloride leaching processes in the absence or presence of

reducing agents are the most commonly tested leaching routes. However, chloride

leaching is more economical and hence the preferred option. It allows relatively easy

treatment of waste solutions and therefore the regeneration and recycling of

hydrochloric acid. Despite the many attempts to develop chloride processes to

upgrade ilmenite to synthetic rutile, systematic kinetic and mechanistic studies based

on chemical and electrochemical dissolution are lacking in the literature. In addition,

no comparison has been made between flat natural or synthetic ilmenite surfaces and

particles of ilmenite concentrates of different origin and composition in chloride

solutions. The main objective of this thesis is to bridge this gap and improve the

understanding of the kinetics and reactions for the dissolution of synthetic ilmenite

and natural ilmenite concentrates of different compositions, under non-reducing and

reducing conditions.

The measured dissolution rates of flat surfaces of synthetic ilmenite were

found to decrease with time and obey a parabolic rate law, due to the blockage of the

surface by an insoluble product. The dissolution rates of both iron and titanium were

higher in the presence of tin(II) chloride as a reducing agent and remained unaffected

over time. The reductive role of tin(II) ions appears to be that of inhibiting the

formation of surface blocking solids, which can be used to propose a reaction

sequence for ilmenite leaching.

i
In electrochemical studies the measured rest potential of an ilmenite electrode

in hydrochloric acid solutions was found to be lower compared to that measured in

sulfuric acid solutions, indicating higher dissolution rates in the former case, due to

reducing conditions. The measured potentials and surface characterisation studies

indicated the reductive dissolution of titanium(IV) in the solid phase to titanium(III)

in the aqueous phase, which facilitates the reductive leaching process by removing

the insoluble product layers. Hematite mineral, which is found associated with

weathered ilmenite, was also found to have a higher dissolution rate under the

cathodic conditions which facilitated leaching.

The leaching of three ilmenite samples of different compositions collected

from local producers over the years, labelled as North Capel, Iluka and Tiwest, gave

different results with regards to the iron and titanium dissolution as a result of

different degrees of alteration. The leaching efficiency appeared to be in the

descending order of North Capel > Iluka > Tiwest, which appeared to be inversely

proportional to the degree of alteration. The highest extraction in hydrochloric acid

alone was obtained from North Capel ilmenite (99% Ti and ~100% Fe) under

conditions of 11 M HCl, pulp density of 4 g/L, 80 -110 °C and 2 h of leaching time.

However, in the presence of a reducing agent, dissolution rates of iron and titanium

increased in solutions of low acid concentrations and temperature.

The initial rate of ilmenite dissolution in hydrochloric acid under reducing

conditions was found to be half order with respect to the concentration of reducing

agent, which suggests the involvement of an electrochemical reaction. However, the

chemical/electrochemical leaching studies of rotating flat surfaces of ilmenite, based

on the parabolic rate law and kinetic study revealed the same diffusivity of protons

through the product layer.

ii
The batch leaching of ilmenite particles obeyed a shrinking core kinetic model,

supporting the view that the proton diffusion through a product layer is the rate

controlling step. The magnitude of proton diffusivity obtained using chemical

leaching studies of rotating flat surfaces of synthetic or natural ilmenite or batch

leaching studies of different ilmenite concentrates agree reasonably well despite the

difference in mineral composition.

iii
ACKNOWLEDGMENTS

First and foremost words cannot express my gratitude towards my mentor

and principal supervisor Associate Professor Gamini Senanayake throughout my

PhD study and successful completion of my thesis. Nevertheless this is extended to

my second supervisor Professor Mike Nicol for his undivided dedication in showing

me the path to complete my thesis. Thank You.

To the laboratory staff especially Ken Seymour, Murray Lindau and Stewart

Kelly your technical support are a blessing in my quest to reach my goals. To Dr

Anna d’Aloya de Pinilla and Chanika Dharmasiri from Chemical and Metallurgical

Engineering, School of Information Technology and Engineering, thank you for

proof reading the thesis.

To all my dear colleagues, Dr Nur Shahidah Ali, Dr Dmitry Pugaev, Dr Nur

Saidatul Zam Zam, Dr Siti Aida, Rajat Hans, Noresma Jahya, Sariati Azman, Nor

Wanis, Dr Suhaina Ismail and Nadira Batool your experience and knowledge sharing

completes my studies and this thesis is a testimony for all your kind assistance.

To my beloved family; husband (Yuzran Azam), children (Akmal Syahmi &

Arissa Sofiya), parents (Zainab Othman & Jabit Chik), brother (Abdul Rahman) and

in-laws (Azam Hussein & Che Maheran Maina) thank you for all the love, prayers

and support throughout my entire journey.

Last but not least I am truly grateful for the PhD scholarship from Ministry of

Higher Education, Malaysia and Universiti Sains Malaysia (USM) which provided

the opportunity for me to pursue my study in Perth.

iv
PUBLICATIONS AND CONFERENCE
PRESENTATION

1. Jabit, N.A, Senanayake, G., Nicol, M. (2012). An electrochemical and leaching

study of ilmenite in hydrochloric acid solution. 10th Spring Meeting of the

International Society of Electrochemistry, 15-18 April 2012, Perth, Australia.

2. Jabit, N.A, Senanayake, G., Nicol, M. (2012) . Electrochemical study of ilmenite

using carbon paste electrode under reducing condition. Pacific RIM Meeting on

Electrochemical and Solid-State Science, 7-12 October 2012, Honolulu, Hawaii.

3. Jabit, N.A, Senanayake, G. (2014). Reductive leaching of Ilmenite in

hydrochloric Acid: A-preliminary study using a rotating synthetic ilmenite

dissolution. Proceeding of 7th Symposium of Hydrometallurgy 2014, 22 to 24

June 2014, Volume II, pp. 837-848, Victoria, BC, Canada.

v
TABLE OF CONTENTS

TABLE OF CONTENTS
DECLARATION
ABSTRACT i
ACKNOWLEDGEMENT iv
PUBLICATIONS AND CONFERENCE PRESENTATION v
TABLE OF CONTENT vi
LIST OF FIGURES xiii
LIST OF TABLES xxvii

CHAPTER 1 INTRODUCTION 1

1.1. The ilmenite and rutile market 1

1.2. Titanium and its applications 5

1.3. Treatments methods 7

1.4. Hydrometallurgical processes 11

1.5. Dissolution studies of ilmenite in hydrochloric acid 14

1.6. Objectives 17

CHAPTER 2 LITERATURE REVIEW 18

2.1. Mineralogy and occurrence 18

2.1.1. Titanium mineralogy 18

2.1.2. Natural weathering of ilmenite 23

2.1.3. Crystal structure 29

2.2. Production of synthetic rutile and TiO2 pigment 32

2.2.1. Sulfate process 32

2.2.2. Chloride process 34

2.2.3. Upgrading ilmenite to synthetic rutile 37

vi
TABLE OF CONTENTS

2.3. Direct leaching of ilmenite 39

2.3.1. Eh-pH and speciation diagrams 39

2.3.2. Kinetic models 42

2.3.3. Kinetics sulfuric acid leaching 44

2.3.4. Factors affecting sulfuric acid leaching 46

2.3.5. Kinetics of hydrochloric acid leaching 49

2.3.6. Factors affecting hydrochloric acid leaching 53

2.3.7. Mixed brine leaching 60

2.3.8. Reductive leaching of ilmenite in HCl 62

2.4. Rotating disc study of ilmenite 70

2.4.1. Background 70

2.4.2. Effect of disc rotation speed 71

2.4.3. Effect of temperature 74

2.4.4. Investigations relevant to this study 75

2.5. Electrochemical study of Ilmenite 76

2.5.1. Open circuit potential (OCP) 79

2.5.2. Mixed potential theory 83

2.5.3. Voltammetry studies 87

2.6. Conclusions 98

CHAPTER 3 MATERIALS AND METHODS 100

3.1. Materials 100

3.1.1. Concentrates 100

3.1.2. Materials and reagents 101

vii
TABLE OF CONTENTS

3.1.3. Synthetic ilmenite preparation 102

3.1.4. Rotating disc dissolution experiment 103

3.2. Electrochemical measurements 105

3.2.1. Equipment, electrodes and electrochemical cell 105

3.2.2. Electrode materials and electrolytes 108

3.2.3. Batch leaching experiments 109

3.2.4. Solution potential measurement 110

3.3. Analytical measurement 111

3.3.1. Atomic absorption spectroscopy 111

3.3.2. UV/Visible spectrophotometry 112

3.3.3. Inductive coupled plasma optical emission 112


spectrometry (ICP-OES)

3.3.4. Assay analysis 112

3.3.5. X-ray diffraction 113

3.3.6. Particle size analysis 113

3.3.7. Optical microscopy 114

3.3.8. Scanning electron microscopy 114

3.3.9. Brunauer-Emmet-Teller (BET) surface area analysis 114

CHAPTER 4 CHARACTERIZATION OF SYNTHETIC AND 115


NATURAL ILMENITE

4.1. Introduction 115

4.2. Description of minerals 115

4.2.1. Synthetic ilmenite 116

4.2.2. Iluka concentrates 119

viii
TABLE OF CONTENTS

4.2.3. North Capel concentrate 125

4.2.4. Tiwest concentrate 131

4.3. Summary and conclusion 138

CHAPTER 5 DISSOLUTION OF SYNTHETIC ILMENITE 139


ROTATING DISC IN HYDROCHLORIC ACID
SOLUTIONS
5.1. Introduction 139

5.2. Dissolution of synthetic ilmenite in hydrochloric acid 141


using rotating disc

5.2.1. Effect of temperature 142

5.2.2. Effect of disc rotation speed 146

5.2.3. Effect of hydrochloric acid concentration 151

5.2.4. Parabolic rate law 159

5.3. Dissolution of ilmenite under reducing condition 162

5.4. Proposed reaction mechanism 164

5.5. Summary and conclusions 172

CHAPTER 6 ELECTROCHEMICAL STUDY OF ILMENITE 173


DISSOLUTION IN HYDROCHLORIC ACID

6.1. Introduction 173

6.2. Open circuit potential 175

6.2.1. Comparison of massive ilmenite and platinum 175


electrodes

6.2.2. Comparison of OCP in HCl and H2SO4 182

6.2.3. General trends and applicability of Nernst equation 184

6.2.4. Comparison with HSC data 188

ix
TABLE OF CONTENTS

6.2.5. Effect of Fe(II), Fe(III) and Sn(II) 192

6.3. Cyclic voltammetric study 196

6.3.1. Cyclic voltammetry of synthetic ilmenite in HCl 197


solution

6.3.2. Cyclic voltammetry of massive ilmenite 202


electrode

6.3.3. Cyclic voltammetry of natural ilmenite powder in 203


sulfuric acid solution

6.3.4. Cyclic voltammetry of natural ilmenite in HCl 207


solution

6.3.5. Effect of reducing agent 213

6.4. Limiting current on Sn(II) and Fe(II) oxidation 215

6.5. Coulometric measurement and dissolution study 220

6.5.1. Massive ilmenite electrode 220

6.5.2. Iluka ilmenite carbon paste electrode 224

6.5.3. Comparison of rates of iron and titanium 227


dissolution

6.6. Conclusion 228

CHAPTER 7 REDUCTIVE AND NON-REDUCTIVE LEACHING 229


OF NATURAL ILMENITE IN HYDROCHLORIC
ACID SOLUTION

7.1. Introduction 229

7.2. Effect of particle size and mineralogy 232

7.2.1. Leaching efficiency 232

7.2.2. Effect of alteration 235

7.2.3. Fe-Ti correlation 237

7.3. Effect of acid concentration 239

x
TABLE OF CONTENTS

7.3.1. Leaching efficiency 239

7.3.2. Effect of alteration 244

7.3.3. Fe-Ti correlation 245

7.4. Effect of temperature 248

7.4.1. Leaching efficiency 248

7.4.2. Predicted equilibrium constant 251

7.4.3. Fe-Ti correlation 253

7.5. Effect of reducing agent 256

7.5.1. Effect of iron metal powder 256

7.5.2. Effect of tin metal powder 257

7.5.3. Electrode potentials and leaching reactions 263

7.6. Residue analysis 269

7.7. Reaction rates and orders 274

7.7.1. General equations 274

7.7.2. Reaction orders under non-reducing conditions 277

7.7.3. Reaction orders under reducing conditions 279

7.8. Activation energy 282

7.9. Kinetic models 287

7.9.1. Shrinking sphere and core models 287

7.9.2. Factors affecting apparent rate constants 289

7.9.3. Effect of particle size 292

7.9.4. Diffusion coefficient of H+ 292

7.10. Summary and conclusion 296

xi
TABLE OF CONTENTS

CHAPTER 8 SUMMARY, CONCLUSION AND FUTURE WORK 298

8.1. Summary and conclusion 298

8.2. Recommendations for future work 303

REFERENCES 304

APPENDIX 315
Appendix A1 315
Appendix A2 319
Appendix A3 330
Appendix A4 333

xii
LIST OF FIGURES

LIST OF FIGURES

Fig. 1.1. World production of ilmenite in years 2013-2015 (Bedinger, 2


2015; Bedinger, 2016).

Fig. 1.2. Location of major heavy minerals sands provinces in Australia 4


(Pownceby et al., 2008).

Fig. 1.3. Percentage of TiO2 used in various products (Egerton, 2000) 6

Fig. 1.4. A schematic diagram of processes for the production of Ti metal 8


and TiO2 pigment (Zhang et al., 2011a)

Fig. 1.5. Summary of existing and proposed processes for upgrading 9


ilmenite to synthetic rutile

Fig. 2.1. Titanium minerals (Barksdale, 1966; Zhang et al., 2011a) 19

Fig. 2.2. Theoretical TiO2 content (%, w/w) in titaniferous oxide phases 25
(Reyneke and Wallmach, 2007).

Fig. 2.3. Backscattered image of ilmenite grains. Progressive alteration 26


of ilmenite depicting (a) unaltered ilmenite, (b) incipient
alteration of ilmenite along grain boundaries and fractures, (c)
extensive alteration throughout the grain and (d) ‘spotty’
alteration of ilmenite involving the formation of hematite,
pseudorutile, ferropseudobrookite and rutile/anatase (Reyneke
and Wallmach, 2007)

Fig. 2.4. Backscattered image of (a) altered ilmenite with containing 27


various Ti-bearing phases, including remnant ilmenite,
pseudorutile, ferropseudobrookite and rutile/anatase; (b) porous
secondary particle containing ilmenite, pseudorutile,
ferropseudobrookite and rutile/anatase and associated with the
Si-contaminants (Reyneke and Wallmach, 2007)

Fig. 2.5. Optical micrograph of ilmenite grains (a) exsolved ilmenite- 27


hematite with curvature of the lamelle suggest that the grain was
subjected to physical strain after its crystallization; (b) ilmenite
altered to pseudorutile and ferropseudobrookite along cracks and
grain boundaries (Reyneke and Wallmach, 2007)

xiii
LIST OF FIGURES

Fig. 2.6. Unit cell model of (a) ilmenite and (b) hematite. Titanium is 30
shown in green, iron in red and oxygen in grey
.(http://www.geocities.jp/ohba_lab_ob_page/structure6.html

Fig. 2.7. A schematic sulfate flowsheet recently developed by BHP 33


Billiton researchers (Stuart et al., 2010)

Fig. 2.8. A simplified flowsheet of the chloride process (Dimitrios and 35


Guillaume, 2009)

Fig. 2.9. Eh-pH diagram of the Ti-Fe-H2O system at 25°C, (excluding 40


Ti2+) considering TiO2 (c, hydrated), Ti2O3 (c,hydrated) and
Fe2O3 as the solid titanium oxide and Fe(III) phases, and for 0.1
and 0.01 dissolved iron and titanium activities, respectively
(Kelsall and Robbins, 1990)

Fig. 2.10. Ti-Cl-H2O system (10-6 M Ti, 0.1 M Cl-, 25 oC) (Vaughan and 41
Alfantazi, 2006)

Fig. 2.11. Ti(IV) chloride speciation diagram for 0.1 Ti(IV) activity at 298 52
K (van Dyk et al., 2002)

Fig. 2.12. Flowsheet developed for production of TiO2 (Lakshmanan et al., 60


2008)

Fig. 2.13. A proposed flowsheet of synthetic rutile production by reductive 63


leaching process (Mahmoud et al., 2004)

Fig. 2.14. Leaching of ilmenite ore with HCl only (A) and in presence of 67
iron powder. (B) [HCl] = 20%, ilmenite /acid ratio = 1/7.3,
temperature = 110 oC, Fe powder = 0.075 g/g ore, time of
addition = 20 min (Mahmoud et al., 2004)

Fig. 2.15. A comparison of the dissolution of the rotating discs of Fe3O4, 71


Fe2O3 and copper, nickel, and zinc ferrites in 1 M HCl (based on
iron dissolved) at 25 oC, 400 rpm (Barton and McConnel, 1979)

Fig. 2.16. (A) Dissolution rate, dc/dt, at 25 oC of calcium ion from 73


superphosphate as a function of square root of rotation velocity.
(B)(▲) Reduced ilmenite disc, 25 oC, 0.1 M FeCl3; (●) Pure iron
disc, 25 oC, 0.02 M Fe2+/0.1 M HCl (Barton and McConnel,
1979 (A); Ward, 1990 (B)

xiv
LIST OF FIGURES

Fig. 2.17. Open circuit potential, Eo as functions of time for magnetite 81


(ISH) and ilmenite electrodes reacted in anoxic solutions at
different pH (White et al., 1994)

Fig. 2.18. Open circuit potentials of rotating ilmenite and platinum 83


electrodes in 450 g L-1 H2SO4 solutions containing various
concentration of iron at 60 °C. (a) platinum [20 g L-1 Fe2+], (b)
ilmenite [ 20 g L-1 Fe2+], (c) ilmenite [H2SO4 only], (d) platinum
[10 g L-1 Fe2+ and 10 g L-1 Fe3+], (e) ilmenite [10 g L-1 Fe2+ and
10 g L-1 Fe3+] (Zhang and Nicol, 2009).

Fig. 2.19. Schematic representation of a mixed potential in solution (Rand 85


and Woods, 1984)

Fig. 2.20. Mixed potential diagram 86

Fig. 2.21. Schematic Evan diagram for the leaching of Fe3O4 in acidified 86
Cu(II)-Cu(I) solutions (Lu and Muir, 1985)

Fig. 2.22. Voltammogram of Ilmenite (2 mg of Australian ilmenite ore) in 89


H2SO4 (1M); scan rate = 10-3 V s-1; arrow indicates that the start
of scan in the anodic direction (Adriamanana et al., 1984)

Fig. 2.23. Linear sweep voltammogram (5 mV/s) of massive ilmenite 90


electrode (A); carbon paste electrode (B) in 450 g/L H2SO4
solution at 60 ⁰C. Negative sweep from OCP, reversed at -0.4 V
(Zhang and Nicol, 2009)

Fig. 2.24. Potentiodynamic scans showing measured current plotted as 91


functions of applied potential E for ilmenite electrode in anoxic
pH 3 solutions. Vertical dashed lines correspond to potentials
defining the stability of water (White et al., 1994)

Fig. 2.25. Photomicrograph of the surface of a massive ilmenite electrode 94


after application of a potential 0.1 V for 4 h in a 450 g/L H2SO4
solution at 60 oC. Areas of A, B, C and D indicate corrosion of
hematite lamellae (Zhang and Nicol, 2009)

Fig. 2.26. Potentiostatic dissolution of massive (A) and carbon paste (B) 96
ilmenite electrode in 450 g/L H2SO4 at various potential at 60 oC
(▲) Ti, (◊) Fe, (■) Charge (Zhang and Nicol, 2009)

xv
LIST OF FIGURES

Fig. 2.27. Effect of temperature on the rate of the dissolution of (A) 97


massive ilmenite and (B) carbon paste ilmenite at 0 V potential
in 450 g/L H2SO4 (▲) Ti, () Fe, (■) Charge (Zhang and Nicol,
2009)
Fig. 3.1. An illustrated diagram of rotating disc dissolution experiment. 104
A: rotating motor; B: rotating stand; C: shaft; D: platinum
electrode, E: reaction vessel; F: synthetic ilmenite disc; G: tube;
H: jack; I: temperature controller; J: water bath

Fig. 3.2. A schematic diagram of the electrochemical cell and experiment 107
set up (A) Metrohm glass cell; (B) water jacket (C) luggin
capilarry (D) reference electrode ; (E) counter electrode; (F)
working electrode; (G) water bath; (H) N2 cylinder; (I) rotater;
(J) potentiostat; (K) computer; (L) printer, (M) rotater controller;
(N) stand ; (O) temperature controller; (P) platinum wire

Fig. 3.3. A schematic diagram of special electrode used for carbon paste 108
ilmenite electrode

Fig. 3.4. A schematic diagram for dissolution experiment. (i) retort stand; 111
(ii) temperature regulator; (iii) condenser; (iv) reactor; (v)
floating magnetic stirrer; ((vi) hotplate with temperature and
magnetic controller; (vii) sample drawn vial.

Fig. 4.1. XRD patterns for synthetic ilmenite (SI) (KαCo, λ = 1.788965) 117

Fig. 4.2. (a) SEM surface morphology of synthetic ilmenite disc and (b) 118
EDX analysis

Fig. 4.3. Particle size distribution of Iluka ilmenite 119

Fig. 4.4. XRD patterns for Iluka ilmenite (KαCo, λ = 1.788965) 120

Fig. 4.5. Surface morphology of Iluka ilmenite concentrate 121

Fig. 4.6. (a) Surface morphology of polished section of Iluka ilmenite 122
sample (b) EDX analysis at low magnification

Fig. 4.7. (a) Surface morphology of polished section of Iluka ilmenite 123
concentrate and (b) EDX analysis on selected area, at higher
magnification

Fig. 4.8. Optical micrograph showing (a) the weathering effect (b) altered 124
particles with patchy pattern on Iluka sample
xvi
LIST OF FIGURES

Fig. 4.9. Particle size distribution of North Capel ilmenite 126

Fig. 4.10. Surface morphology North Capel ilmenite particles 126

Fig. 4.11. XRD patterns for North Capel ilmenite (KαCo, λ = 1.788965) 127

Fig. 4.12. (a) Surface morphology of polished section of North Capel 128
ilmenite sample and (b) EDX analysis at low magnification

Fig. 4.13. (a) Surface morphology of polished section of North Capel 129
ilmenite sample and (b) EDX analysis at higher magnification

Fig. 4.14. Optical micrograph showing the weathering effect on North 130
Capel sample (a) altered particles (b) particles with patchy
pattern

Fig. 4.15. Particle size distribution of Tiwest sample 132

Fig. 4.16. XRD patterns for Tiwest concentrate (KαCo, λ = 1.788965). 133

Fig. 4.17. Surface morphology of Tiwest ilmenite sample 134

Fig. 4.18. (a) Surface morphology of polished section of Tiwest ilmenite 135
sample and (b) EDX analysis at low magnification

Fig. 4.19. (a) Surface morphology of polished section of Tiwest ilmenite 136
sample and (b) EDX analysis at higher magnification

Fig. 4.20. Optical micrograph showing the weathering effect on Tiwest 137
ilmenite sample

Fig. 5.1. Effect of temperature on iron dissolution from synthetic ilmenite 143
disc in 7 M HCl solution at 800 rpm

Fig. 5.2. Effect of temperature on titanium dissolution from synthetic 143


ilmenite disc in 7 M HCl solution at 800 rpm

Fig. 5.3. Plot of iron dissolution as a function of titanium dissolution in 144


7 M HCl at 800 rpm at different temperatures.

Fig. 5.4. Arrhenius plot for iron and titanium dissolution (T = 50, 60, 70, 145
80 oC , [HCl] = 7 M, stirring speed = 800 rpm)

xvii
LIST OF FIGURES

Fig. 5.5. Extent of dissolution of synthetic ilmenite given by iron 148


concentration as a function of time at various stirring speed in
7 M HCl at 60 oC

Fig. 5.6. Extent of dissolution of synthetic ilmenite given by titanium 148


concentration as a function of time at various stirring speed in
7 M HCl at 60 oC

Fig. 5.7. Dissolution rate of titanium at 60 oC in 7 M HCl as a function of 150


square root of stirring rate.

Fig. 5.8. Plot iron dissolution as a function of titanium dissolution in 7 M 150


HCl at 60 oC at different stirring rates.

Fig. 5.9. Effect of HCl concentration on iron dissolution at 60 oC and 800 151
rpm.

Fig. 5.10. Effect of HCl concentration on titanium dissolution at 60 oC and 152


800 rpm.

Fig. 5.11. Plot of iron dissolution as a function of titanium dissolution at 153


60 oC, 800 rpm at different acid concentration.

Fig. 5.12. Speciation diagram of Ti(IV) (total 0.001 mol/L) species in 154
chloride media based on HSC database version 7.1 (Roine, 2011)

Fig. 5.13. Eh-pH diagram of Ti-Cl-H2O at 25 oC (Vaughan and Alfantanzi, 154


2004)

Fig. 5.14. SEM images of synthetic ilmenite disc surface before dissolution 156
(A) and after dissolution in 7 M HCl at 80 oC (B) with EDX
spectrum analysis before (C) and after (D) dissolution.

Fig. 5.15. SEM images of synthetic ilmenite disc surface after dissolution 156
in 7 M HCl at 80 oC and EDX analysis of white precipitate
(marked by white arrow).

Fig. 5.16. Logarithmic plot of initial rate of iron and titanium dissolution as 158
a function of [H+] at 60 oC and 800 rpm

Fig. 5.17. Change in square of dissolved amount of Fe and Ti from 161


synthetic ilmenite disc at 60 oC, 800 rpm.

xviii
LIST OF FIGURES

Fig. 5.18. Effect of tin(II) chloride on the rate of iron dissolution from 163
synthetic ilmenite disc in HCl (7 M) solution at 60 oC, 800 rpm.
Fig. 5.19. Effect of tin(II) chloride on the rate of titanium dissolution from 163
synthetic ilmenite disc in HCl (7 M) solution at 60 oC, 800 rpm.

Fig. 5.20. Logarithmic plot of initial rate of iron and titanium dissolution as 164
a function of Sn(II) concentration in 7 M HCl, 60 oC, 800 rpm.

Fig. 5.21. Eh-pH diagrams for Fe-Ti-H2O system at 25 oC, (excluding Ti2+ 166
ions) considering TiO2 (c,hydrated), Ti2O3 (c, hydrated) and
Fe2O3 as the solid titanium oxide and Fe(III) phases, and for 0.1
and 0.01 dissolved iron and titanium activities, respectively
(Kelsall and Robbins, 1990).

Fig. 5.22. Eh-pH diagram for the Sn/H2O-Cl system at 298 K, at a dissolve 167
tin activity = 10-3 , aCl = 5, p (SnH4) = 1 (House and Kelsall,
1984).

Fig. 5.23. SEM images of synthetic ilmenite disc surface after dissolution 169
in 7 M HCl in the presence of 2.0 g/l SnCl2 at 60 oC.

Fig. 5.24. (A)-(B) EDX spectrum and SEM analysis of bulk synthetic 169
ilmenite disc surface and (C)–(D) EDX spectrum and SEM
analysis of white spot mark by red arrow (Full scale of EDX
spectrum are presented in Appendix A2.3).

Fig. 6.1. Open circuit potential of rotating massive ilmenite (MI) and 176
platinum electrodes in 7 M HCl solutions at 50 oC, rotated at 200
rpm.

Fig. 6.2. Open circuit potential of rotating massive ilmenite (MI) and 177
platinum electrodes in 7 M HCl in the presence of 10 g/L Fe(II)
and 10 g/L Fe(III) solutions at 50 oC, rotated at 200 rpm.

Fig. 6.3. Open circuit potential of rotating massive ilmenite (MI) and 177
platinum electrodes in 7 M HCl solutions at 50 oC in the
presence of 20 g/L Fe(II), rotated at 200 rpm.

Fig. 6.4. Open circuit potential of rotating massive ilmenite (MI) and 178
platinum electrodes in 7 M HCl solutions at 50 oC in the
presence of 20 g/L Fe(III), rotated at 200 rpm.

Fig. 6.5. Open circuit potential of rotating massive ilmenite and platinum 178
electrodes in 7 M HCl solutions at 50 oC in the presence of
20 g/L Sn(II) solutions, rotated at 200 rpm.

xix
LIST OF FIGURES

Fig. 6.6. Open circuit potentials of rotating massive ilmenite (MI) 179
electrode in 7 M HCl solution containing various concentration
of Fe(II) and Fe(III) ions at 50 °C.

Fig. 6.7. Open circuit potentials of platinum electrode in 7 M HCl solution 179
containing various concentration of Fe(II) and Fe(III) ions at
50 °C.

Fig. 6.8. Measured OCP in (a) H2SO4 (4.6 M) (Zhang and Nicol, 2009), 183
(b) HCl (7 M).

Fig. 6.9. Effect of concentration ratio of Fe(II)/Fe(III) on OCP of MI and 187


Pt electrodes at 50oC (data from previous figures).

Fig. 6.10. Open circuit potentials of platinum electrode in H2SO4 or HCl 189
(Conditions: 4.6 M H2SO4 at 60 oC (Zhang and Nicol, 2009) or
4 M HCl at 50 oC (this work), 200 rpm under nitrogen (measured
data points from Table 6.1). Note the lines are reduction
potentials from HSC 7.1(Roine, 2011) and the data points with
symbols are from OCP measurements.

Fig. 6.11. Open circuit potentials of massive ilmenite electrode in H2SO4 191
or HCl (Conditions: 4.6 M H2SO4 at 60 oC (Zhang and Nicol,
2009) or 4 M HCl at 50 oC (this work), 200 rpm under nitrogen.
Note the lines are reduction potentials from HSC 7.1 and the
data points with symbols are from OCP measurements.

Fig. 6.12. Effect of Fe(II) and Sn(II) on open circuit potentials of (a) 193
massive ilmenite and (b) platinum electrodes in 7 M HCl in the
absence of added Fe(III), (Fe(II) = 10 g/L, Sn(II) = 20 g/L (T =
50 oC, stirring rate = 200 rpm).

Fig. 6.13. Effect of temperature on equilibrium constants (Roine, 2011) 195

Fig. 6.14. Cyclic voltammogram (CV) of synthetic ilmenite carbon paste 198
(SI-CP) electrodes recorded at 60 oC in 4 M HCl at a scan rate of
5 mV/s (0 rpm).

Fig. 6.15. Peak potentials of cyclic voltammogram of synthetic ilmenite 199


carbon paste (SI-CP) electrode (Conditions: 4 M HCl, 60 oC,
0 rpm). Note the lines are reduction potentials from HSC 7.1
and the data points with symbols are from CV peak potential
measurements at 60 oC.

Fig. 6.16. CV of a stationary massive ilmenite (MI) electrode recorded at 202


60 oC in 4 M HCl at a scan rate of 5 mV/s.

xx
LIST OF FIGURES

Fig. 6.17. (a) CV of ilmenite carbon paste electrode in sulfuric acid (4 M) 205
at 60 ºC, scan rate: 5 mV/s by Zhang and Nicol (2009); (b) CV
of ilmenite (Iluka) carbon paste electrode obtained in this study.
Sweep initiated in a negative direction from open circuit
potential and reversed at -0.400 V.

Fig. 6.18. CV peak potentials of Iluka ilmenite/carbonpaste electrodes in 206


4.0 or 4.6 M H2SO4 or 4 M HCl at 60 oC. Note the lines are
reduction potentials from HSC 7.1 and the data points with
symbols are from CV peak potential measurements.

Fig. 6.19. CV of stationary ilmenite (Iluka) carbon paste electrode in (I) 207
H2SO4 (4 M) solution and (II) HCl (4 M) at 60oC, scan rate = 5
mV/s, (III) carbon paste only , sweep was initiated in negative
direction from the open circuit potential and reversed at -0.400 V
(see arrow).

Fig. 6.20 The Eh-pH diagram for (A) Ti-HCl-H2O system at 25 oC 209
(Vaughan and Alfantazi, 2006).

Fig. 6.21. CV of ilmenite carbon paste electrode (TW & NC) in 4 M HCl 211
solution at 60 oC, scan rate = 5 mV/s, sweep initiated in negative
direction from open circuit potential and reversed at - 0.400 V.

Fig. 6.22. CV peak potentials of Tiwest/North Capel ilmenite/carbon paste 212


electrodes in 4 M HCl at 60 oC. Note the lines are reduction
potentials from HSC 7.1 (Roine, 2011) and the data points with
symbols are from CV peak potential measurements.

Fig. 6.23. CV of (A) ilmenite (Iluka) in 4 M HCl alone at 60 oC; (B) in the 214
presence of SnCl2 (20 g/L) in 4 M HCl at 60 oC; (C) carbon paste
electrode with no ilmenite, under the same conditions as (B).

Fig. 6.24. Anodic limiting current curves measured at a rotating carbon 216
paste disc electrode in 1.0 M HCl solution in the presence of 0.5
g/L Sn(II) at 25 oC. The potential was scanned from OCP to 1.2
V (Scan rate: 5 mV/s).

Fig. 6.25. Linear relationship of anodic limiting current for Sn(II) with 217
square root of rotation rate.

Fig. 6.26. Anodic limiting current curves measured at a rotating carbon 218
paste disc electrode in 1.0 mol/L HCl solution in the presence of
0.25 g/L Fe(II) at different electrode rotation speeds. The
potential was scanned from the OCP to 1.2 V; scan rate =
20 mV/s.
xxi
LIST OF FIGURES

Fig. 6.27. Linear relationship of anodic limiting current for Fe(II) oxidation 218
with square root of rotation rate ().

Fig. 6.28. Current- time transients during reduction of massive ilmenite 221
electrode in 4 M HCl at 60 oC and different potentials (lower
than the rest potential) in 1 hour.

Fig. 6.29. Current-time transient of massive ilmenite electrode in 4 M HCl 221


at 60 oC and different potentials (lower than the rest potential) in
45 seconds.

Fig. 6.30. Potentiostatic dissolution of massive ilmenite electrode in 4 M 223


HCl at 60 oC (Charge passed in 10-4 faraday cm-2 s-1).

Fig. 6.31. Current-time transient during reduction of Iluka ilmenite carbon 224
paste electrode in 4 M HCl at 60 oC at different potentials (lower
than the rest potential) in 4 hours.

Fig. 6.32. Current-time transient in of Iluka ilmenite carbon paste electrode 225
in 4 M HCl at 60 oC at different potentials (lower than the rest
potential) in 60 seconds.

Fig. 6.33. Potentiostatic dissolution of ilmenite (Iluka) carbon paste 226


electrode in 4 M HCl at 60 oC (Charge passed is in faraday
m-2 s-1).

Fig. 7.1. Different composition of North Capel (NC), Tiwest (TW) and 231
Iluka (IL) ilmenite concentrate.

Fig. 7.2. Effect of particle size on dissolution efficiency of Fe and Ti 234


from North Capel ilmenite sample ([HCl] = 7 M, pulp density =
4 g/L ,T = 80 oC, t = 5 h).

Fig. 7.3. Effect of particle size on dissolution efficiency of (a) Fe and (b) 234
Ti from Iluka ilmenite sample ([HCl] = 7M, pulp density = 4 g/L
,T = 80 oC, t = 5 h).

Fig. 7.4. Effect of particle size on dissolution efficiency of (a) Fe and (b) 235
Ti from Tiwest ilmenite sample ([HCl] = 7 M, pulp density = 4
g/L ,T = 80 oC, t = 5 h).

Fig. 7.5. Correlation between Fe and Ti extraction from ilmenite samples 238
for different size fractions (a) North Capel (b) Iluka (c) Tiwest
([HCl] = 7 M, pulp density = 4 g/L ,T = 80 oC, t = 5 h).
xxii
LIST OF FIGURES

Fig. 7.6. Effect of acid concentration on Fe and Ti dissolution from North 240
Capel ilmenite sample. ([HCl] = 4 M -11 M, pulp density = 4 g/L
,T = 80 oC, t = 5 h, size fraction = +53-63 µm).

Fig. 7.7. Effect of acid concentration on Fe and Ti dissolution of Iluka 242


ilmenite sample. ([HCl] = 4 M-11 M, pulp density = 4 g/L ,T =
80 oC, t = 5 h, size fraction= +53 -63µm).

Fig. 7.8. Effect of acid concentration on dissolution of Fe and Ti of 243


Tiwest ilmenite sample. ([HCl] = 4 M-11 M, pulp density = 4
g/L ,T = 80 oC, t = 5 h, size fraction = +53 -63µm).

Fig. 7.9. Correlation between extraction of Fe and Ti from ilmenite 246


samples at different HCl concentrations (a) North Capel (b)
Iluka (c) Tiwest. ([HCl] = 4 M - 11 M, pulp density = 4 g/L, T
= 80 oC, t = 5 h, size fraction = +53-63 µm).

Fig. 7.10. Effect of temperature on dissolution of Fe and Ti for North Capel 249
ilmenite sample ([HCl] = 7 M, pulp density = 4 g/L, size fraction
= +53-63 µm, leaching time = 5 h).

Fig. 7.11. Effect of temperature on dissolution of Fe and Ti for Iluka 250


ilmenite sample ([HCl] = 7 M, pulp density = 4 g/L, size fraction
= +53-63 µm, leaching time = 5 h).

Fig. 7.12. Effect of temperature on dissolution of Fe and Ti for Tiwest 251


ilmenite sample ([HCl] = 7 M, pulp density = 4 g/L, size fraction
= +53-63 µm, leaching time = 5 h).

Fig. 7.13. Effect of temperature on equilibrium constants for oxide 252


dissolution in acid-chloride solutions (Roine, 2011).

Fig. 7.14. Correlation of Fe – Ti dissolution from ilmenite samples at 254


different temperatures (a) North Capel (b) Iluka (c) Tiwest.

Fig. 7.15. Effect of reducing agent (RA) on dissolution of Fe and Ti for 257
North Capel (NC), Iluka (IL), Tiwest (TW) ilmenite sample
([HCl] = 7 M, T = 80 oC, pulp density = 4 g/L, size fraction =
+53-63 µm, time = 5 h).

Fig. 7.16. Effect of Sn metal on Fe and Ti leaching efficiency for Iluka 260
ilmenite sample. ([HCl] = 7 M, = 80 oC, pulp density = 4.0 g/L,
size fraction = +53-63 µm, t = 5 h).

Fig. 7.17. Effect of Sn metal on leaching efficiency of Fe and Ti for Tiwest 260
sample ([HCl] = 7 M; T = 80 oC, pulp density = 4.0 g/L, size
fraction = +53-63µm, t = 5 h).
xxiii
LIST OF FIGURES

Fig. 7.18. Correlation of Fe – Ti dissolution from ilmenite samples at 262


different Sn dosage (a) North Capel, (b) Iluka (c) Tiwest.

Fig. 7.19. Measured potential of platinum electrode (Eh or Ept) during 264
leaching of Iluka ilmenite sample in 7 M HCl solution at 60 oC
with and without the presence of reducing agent (Sn).

Fig. 7.20. Measured potential of carbon paste electrode (Em) during 264
leaching of Iluka ilmenite sample in 7 M HCl solution at 60 oC
with and without the presence of reducing agent (Sn).

Fig. 7.21. Eh-pH diagrams for Fe-Ti-H2O system at 25 oC, (excluding Ti2+ 266
ions) considering TiO2 (c, hydrated), Ti2O3 (c, hydrated) and
Fe2O3 as the solid titanium oxide and Fe(III) phases, and for 0.1
and 0.01 dissolved iron and titanium activities, respectively
(Kelsall and Robbins, 1990).

Fig. 7.22. XRD pattern of Iluka sample and Iluka leach residues after 270
leaching in 7 M HCl, at 110 oC in the presence of tin (2 g/L),
pulp density = 40.6 g/L. I = ilmenite; R = rutile; P= pseudorutile;
H = hematite.

Fig. 7.23. SEM images of particles obtained before and after leaching in 7 272
M HCl at 40.6 g/L pulp density, 2 g/L of tin metal powder and
110 oC (a) before leaching, (b) after leaching (2 h), (c) after
leaching (5 h), (d) after leaching (5 h) at higher magnification.

Fig. 7.24. Fraction (X) of Fe dissolution with respect to time at different 276
acid concentration for polynomial fitting (sample = Iluka, [HCl]
= 7 M – 11 M HCl, T = 80 oC, pulp density = 4 g/L, size
fraction = +53-63 µm).

Fig. 7.25. Logarithmic plot of initial rates of Iluka samples (a) [dX/dt]Fe 277
and (b) [dX/dt]Ti as a function of acid concentration sample
([HCl] = 7, 9, 10, 11 M, pulp density = 4 g/L, T = 80 oC, t = 5 h,
time = 5 h, size fraction = +53-63 µm).

Fig. 7.26. Logarithmic plot of initial rates of North Capel sample (a) 278
(dX/dt)Fe and (b) (dX/dt)Ti as a function of acid concentration
([HCl] = 7, 9, 10, 11 M, pulp density = 4 g/L ,T = 80 oC, time =
5 h, size fraction = +53-63 µm).

xxiv
LIST OF FIGURES

Fig. 7.27. Logarithmic plot of initial rates of Tiwest sample (a) (dX/dt)Fe 274
and (b) (dX/dt)Ti as a function of acid concentration ([HCl] = 7,
9, 10, 11 M, pulp density = 4 g/L ,T = 80 oC, t = 5 h, size fraction
= +53-63 µm).

Fig. 7.28. Initial rate of dissolution of Iluka ilmenite sample (a) Fe and (b) 280
Ti as a function of Sn concentration in HCl solutions ([Sn] = 0.5
g/L -4.0 g/L); time = 5 h, [HCl] = 7 M, T = 80 oC, time = 5 h,
size fraction = +53-63 µm).

Fig. 7.29. Initial rate of dissolution of Tiwest ilmenite of (a) Fe and (b) Ti 280
as a function of Sn concentrations in HCl solutions. ([Sn] = 0.5
g/L -4.0 g/L), [HCl] = 7 M, pulp density = 4 g/L, T = 80 oC, time
= 5 h, size fraction = +53-63 µm).

Fig. 7.30. Arrhenius plot for Fe and Ti dissolution on ilmenite samples 283
(a) - (b) North Capel , (c) - (d) Iluka, (e) - (f) Tiwest ([HCl] = 7
M, pulp density = 4 g/L,T = 80 oC, 90 oC, 100 oC, 110 oC, t = 5 h,
size fraction = +53-63 µm).

Fig. 7.31. Effect of alteration factor on (a) oxide composition (% w/w) in 286
ilmenite (b) mol/100 g ilmenite (c) activation energy for the
leaching of four samples (SI, NC, IL, and TW) (Data from
previous figures).

Fig. 7.32. Validity of shrinking core model for Fe (A) and Ti (B) for the 288
leaching results (Iluka ilmenite sample, in 7 M HCl at 80 oC
with pulp density of 4 g/L).

Fig. 7.33. Effect of leaching parameters on relative kpl values (Data from 291
Table 7.20) (a) North Capel, (b) Iluka, (c) Tiwest.

Fig. A2.1. Effect of temperature on iron dissolution from synthetic ilmenite 328
disc in 7 M HCl solution at 800 rpm.

Fig. A2.2. Effect of temperature on iron dissolution from synthetic ilmenite 328
disc in 7 M HCl solution at 800 rpm.

Fig. A2.3. Figure A2.3 EDX spectrum for bulk synthetic ilmenite disc for 329
SEM image in Figure 5.24 (B).

Fig.A2.4 EDX spectrum on white spot mark for SEM image in Figure 5.24 329
(D).

Fig. A4.1. Effect of the presence of tin powder on e and Ti leaching 333
efficiency for North Capel sample. ([HCl] = 7 M, T = 80 oC,
[Sn] = 1.0 g/L, pulp density = 4.0 g/L, size fraction = +53-63

xxv
LIST OF FIGURES

µm, t = 5 h).

Fig. A4.2. Fitting of shrinking core model for North Capel ilmenite in HCl 334
solution from 7 M – 11 M at 80 oC, (a) Fe and (b) Ti.

Fig. A4.3. Fitting of shrinking core model for Iluka ilmenite in HCl 335
solution from 7 M – 11 M at 80 oC, (a) Fe and (b) Ti.

Fig. A4.4. Fitting of shrinking core model for Tiwest ilmenite in HCl 336
solution from 7 M – 11 M at 80 oC, (a) Fe and (b) Ti.

Fig. A4.5. Fitting of shrinking core model for Iluka ilmenite in the presence 337
of tin metal powder (Sn) in 7 M HCl at 80 oC, (a) Fe and (b) Ti.

Fig. A4.6. Fitting of shrinking core model for Tiwest ilmenite in the 338
presence of tin metal powder (Sn) in 7 M HCl at 80 oC, (a) Fe
and (b) Ti.

Fig. A4.7. Fitting of shrinking core model for North Capel ilmenite in 7 M 339
HCl solution at temperature from 80 oC to 110 oC, (a) Fe and (b)
Ti.

Fig. A4.8. Fitting of shrinking core model for Iluka ilmenite in 7 M HCl 340
solution at temperature from 80 oC to 110 oC, (a) Fe and (b) Ti.

Fig. A4.9. Fitting of shrinking core model for Tiwest ilmenite in 7 M HCl 341
solution at temperature from 80 oC to 110 oC, (a) Fe and (b) Ti.

Fig. A4.10. Variation of kpl as a function of initial particle radius, 1/r2 (North 342
Capel, size fraction of +45-53 µm, +53-63 µm, +63-75 µm, +75-
90 µm, 80 oC, solid liquid ratio = 4 g/L in 7 M HCl) Data from
Fig. 7.2.

Fig. A4.11. Variation of kpl as a function of initial particle radius, 1/r2 (Iluka, 342
size fraction of +45-53 µm, +53-63 µm, +63-75 µm, +75-90 µm,
80 oC, solid liquid ratio = 4 g/L in 7 M HCl) Data from Fig. 7.3.

Fig. A4.12. Variation of kpl as a function of initial particle radius, 1/r2 343
(Tiwest, size fraction of +45-53 µm, +53-63 µm, +63-75 µm,
+75-90 µm, 80 oC, solid liquid ratio = 4 g/L in 7 M HCl) Data
from Fig. 7.4.

xxvi
LIST OF TABLES

LIST OF TABLES

Table 1.1 Summary of processes for upgrading ilmenite to synthetic rutile 10

Table 1.2 Extraction of Ti and Fe from ilmenite ores using different 13


lixiviants.

Table 1.3 Comparison of sulfuric acid and hydrochloric acid leaching 15

Table 2.1 List of minerals commonly associated with heavy mineral sand 21
deposits

Table 2.2 Chemical composition of ilmenite from different origin 28

Table 2.3 Summary of crystal structures of ilmenite and related oxides 31

Table 2.4 Comparison of sulfate processes with chloride processes 36

Table 2.5 Processes of upgrading ilmenite to synthetic rutile (SR) 38

Table 2.6 Published activation energy data derived from leaching 46


experiments in sulfuric acid

Table 2.7 Kinetic parameters for ilmenite leaching in concentrated HCl 50


solutions

Table 2.8 Effect of ore particle size on ilmenite leaching 53

Table 2.9 Effect of HCl concentration on ilmenite leaching 55

Table 2.10 Effect of temperature on ilmenite leaching 57

Table 2.11 Effect of solid liquid ratio (S/L) on ilmenite leaching 59

Table 2.12 Study on oxide minerals by electrochemical method 78

Table 2.13 Steady state potential values of magnetite and ilmenite 81

Table 3.1 Lists of materials and reagents 101

Table 4.1 Chemical composition of representative samples of ilmenite 116


ores

xxvii
LIST OF TABLES

Table 4.2 Summary of certain properties of ilmenite samples 138

Table 5.1 Rate data and activation energy 146

Table 5.2 Reported and current values of activation energy 146

Table 5.3 EDX analysis of white precipitate in Fig. 5.15 157

Table 5.4 Effect of HCl concentration on initial rates 158

Table 5.5 Potential data for synthetic ilmenite disc dissolution in 165
7 M hydrochloric acid at 60 oC

Table 5.6 EDX analysis on bulk surface of SEM (Fig. 5.24. (A)) and white 170
spot mark by red arrow (Fig. 5.24. (B))

Table 6.1 Summary of OCP of different electrodes under different 180


conditions

Table 6.2 Standard reduction potentials 183

Table 6.3 Stability constants of iron (III) and (II) complexes in aqueous 187
media

Table 6.4 Summary of peak potentials from cyclic voltammogram curves 197
under different conditions.

Table 6.5 Cathodic peak potential and current density value for cyclic 201
voltammogram under different conditions

Table 6.6 Diffusion coefficients of H+, Sn(II) and Fe(II) 219

Table 6.7 Effect of potential on the mean dissolution rates of Fe and Ti and 223
electrical charge in 4 M HCl at 60 oC for massive ilmenite
electrode

Table 6.8 Effect of potential on the mean dissolution rates of Fe and Ti from 226
the Iluka carbon paste electrode and charge passed

Table 6.9 Dissolution rates of different ilmenite electrodes 227

Table 7.1 Parameters used in ilmenite leaching 231

Table 7.2 Chemical analysis of each ilmenite sample 232

xxviii
LIST OF TABLES

Table 7.3 Alteration factor for ilmenite samples 236

Table 7.4 Correlation data between Fe and Ti for ilmenite samples of 237
different size fractions

Table 7.5 Effect of HCl concentration on efficiency of Fe and Ti leaching 241


from natural ilmenite

Table 7.6 Slopes of linear correlation between Fe and Ti for ilmenite 247
dissolution at different HCl concentrations

Table 7.7 Effect of temperature on efficiency of Fe and Ti leaching from 253


natural ilmenite

Table 7.8 Slopes of linear correlation data between Fe and Ti dissolution 255
from ilmenite samples at different temperatures

Table 7.9 Effect of reducing agents on iron and titanium dissolution 257

Table 7.10 Effect of Sn as reducing agent on Fe and Ti leaching from natural 259
ilmenite

Table 7.11 Slopes of linear correlation data between Fe and Ti dissolution 261
from ilmenite samples at different tin dosages

Table 7.12 Potential data measured using platinum electrode and ilmenite 267
electrode under non-reductive and reductive conditions

Table 7.13 Chemical analysis of feed and leach residues 271

Table 7.14 EDX analysis of Iluka ilmenite particles before and after leaching 273

Table 7.15 Initial rate data (dX/dt) at various HCl concentrations 279

Table 7.16 Reaction orders and degree of alteration 279

Table 7.17 Initial reaction rates (dX/dt) and reaction orders (n) with respect to 281
Sn(II)

Table 7.18 Initial rate data at different temperatures 284

Table 7.19 Calculated activation energy values 284

Table 7.20 Summary of effect of variables on apparent rate constants and 290
diffusivity in HCl leaching of ilmenite samples
xxix
LIST OF TABLES

Table 7.21 Dependence of proton diffusion coefficient, DH+ on medium and 295
materials

Table 7.22 Maximum leaching efficiency of iron and titanium from ilmenite 297
under different experimental conditions.

Table A1.1 Thermodynamic data for possible reaction in the dissolution of 315
ilmenite at 25 or 60 °C

Table A1.2 Equilibrium constants 316

Table A1.3 Standard Gibbs free energy for titanium and iron species (Kelsall 317
and Robbins, 1990).

Table A1.4 Thermodynamic data for possible reaction in the dissolution of 318
ilmenite

Table A2.1 Effect of stirring rate 319

Table A2.2 Effect of acid concentration 321

Table A2.3 Effect of temperature 323

Table A2.4 Effect of reducing agent, SnCl2 325

Table A3.1 Effect of potential on dissolution rate of massive ilmenite (Iluka) 331
electrode

Table A3.2 Effect of potential on dissolution rate of Iluka carbon paste 332
electrode (CP IL)

Table A4.1 Results of measured potentials during leaching 333

Table A4.2 Initial radius and kpl data of North Capel, Iluka and Tiwest for 343
effect of particle sizes

xxx
Chapter 1: Introduction

1 CHAPTER 1

INTRODUCTION

1.1. The ilmenite and rutile market

The future of the titanium and white titanium dioxide pigment market rests on

the development of new and economical processes. Therefore, the recent research

has been focused on the development of new technology for the conversion of the

abundant ilmenite (FeTiO3) ore to a product that is chemically equivalent to rutile

(TiO2), which can be further refined to white pigment. Due to the shortage of rutile

resources, ilmenite has been a substitute for rutile as feedstock for production of

titanium metal or titanium dioxide (TiO2) pigment. Several successful methods have

been developed for the conversion of ilmenite to rutile, namely the Benilite process,

the Becher process, the Murso process and the Altair process. However, the high

energy cost of these processes makes it necessary to conduct further research in

order to develop cost effectives process and increase the production of rutile (Zhang

et al., 2011a).

Fig. 1.1 shows the world’s ilmenite production from 2013 until 2015.

According to the United States Geological Survey (USGS) data for 2013 - 2015,

South Africa became the largest producer of ilmenite from year 2013 until 2014

(Bedinger, 2015). China has become the second largest producer of ilmenite from

year 2013 to 2014 and the largest producer in 2015 (Bedinger, 2016). During the

same period, Australia was the third largest ilmenite producer. However, in 2015

Australia became the second largest ilmenite producer after China (Bedinger, 2016).

1
Chapter 1: Introduction

Other major producers of Ilmenite include United States, Brazil, India, Vietnam and

Canada. Australia is also the largest producer of rutile, with about 52% of world

production, followed by South Africa and Sierra Leone.

Ilmenite supplies about 90% of the world’s demand for titanium minerals and

world resources of ilmenite and rutile total more than 2 billion tons (Bedinger,

2015). The demand for high grade titanium ores and white titanium dioxide pigments

has greatly encouraged ilmenite upgrading by removal of iron and other impurities

from its grain lattice. In 2016, Ilmenite (in bulk, minimum 54% TiO2) was worth

US$91 per metric tonne and rutile around US$795 per metric tonne (Bedinger,

2015). The ilmenite prices and productions are forecast to increase in the next two

years in 2019 by 13-15 % (Hope, 2016).

Fig. 1.1. World production of ilmenite in the years 2013-2015 (Bedinger, 2015;
Bedinger, 2016).

2
Chapter 1: Introduction

The weathering over millions of years of igneous and metamorphic rocks,

containing minerals such as rutile, ilmenite, zircon (ZrO2) and leucoxene (TiO2),

results in sand grains with high densities. The grains are generally accessory

minerals in quartz sands. They are concentrated by coastal, alluvial, and eolian

processes that carry away the less dense quartz grains, leaving behind the denser

heavy mineral sands. Most heavy mineral deposits around the world are found in

ancient beach deposits that are now located far from today's oceans (Pownceby et al.,

2008).

The ilmenite component in many heavy mineral deposits can present

significant mineral concentration and processing problems due to their complex

chemistry and mineralogy. Generally, the heavy mineral concentrates are sent to

'dry’ mills and the individual minerals are separated using their different magnetic

and electrical properties at various elevated temperatures. Separation equipment

includes high-tension (electrical), high-intensity magnetic and electrostatic plate

separators (Premaratne and Rawson, 2003).

Many of the present and former coastal regions of Australia are endowed

with heavy mineral deposits that contain ilmenite and much of this ilmenite is, by

virtue of its chemical composition and the mineralogy of its impurities, particularly

suited to processing by the chloride route. Fig. 1.2 shows the distribution of heavy

mineral sand deposits in Australia which covers along east and southwest coastlines,

as well as the Murray Basin inland region. Minerals of high specific gravity (SG) are

concentrated in these deposits by intense wind and waves with the primary economic

minerals being ilmenite, altered ilmenite (leucoxene, TiO2) and pseudorutile

(Fe2Ti3O9).

3
Chapter 1: Introduction

Fig. 1.2. Location of major heavy minerals sand provinces in Australia (Pownceby et
al., 2008).

Australia has been the largest producer of ilmenite concentrate, but only

contributes to 4% of the world’s production of titanium dioxide pigment. The

ilmenite processing plants in Western Australia typically convert less than one-half

of the ilmenite to synthetic rutile, of which one-third is supplied to Western

Australia’s two pigment plants (Tiwest and Millenium inorganic chemicals), with the

remainder exported. The ilmenite which is not converted to synthetic rutile is often

of a lower grade (52 to 57% TiO2) and is largely suitable for sulfate based plants.

The Becher process developed in Western Australia is suitable for local

ilmenite that contains low concentrations of magnesium and chromium. The

concentrations of these impurities in ilmenite from most other places are usually too

high for successful use of the Becher process (Becher et al., 1965).

4
Chapter 1: Introduction

Currently, the Becher process faces problems with the grade of the final

product in the form of synthetic rutile. This is due to high amounts of pseudorutile

and leucoxene, which are present as inclusions within the ilmenite grains in the TiO2

grade region. Alternatively, slagging, which is practiced in South Africa and needs

to be conducted on a large scale to be economical, results in a low grade synthetic

rutile of 85% TiO2. Direct leaching is potentially another substitute to the current

process.

1.2. Titanium and its applications

Titanium is the ninth most common element and is widely distributed in the

earth crust. Titanium metal and white titanium dioxide pigment have been the most

significant materials in various applications because of their properties. Titanium

metal is light in weight, non-toxic, durable and has important uses in the aerospace

industry (aero-engine and airframes). Titanium exhibits exceptional resistance to

corrosive attack by salt water or marine atmosphere and also to a broad range of

acids and industrial chemicals. For these reasons, titanium has special applications in

the construction of water desalination and chemical plants. Titanium also has very

important applications in medical science such as surgical implants into the human

body, including heart pacemakers and artificial limbs and joints.

The growth of the titanium dioxide industry started in the early 1900’s. The

worldwide TiO2 pigment production is valued in excess of US$10 billion per year,

making it one of the world’s most important inorganic chemical industries

(Gambogi, 2009). The most important commercial use of titanium dioxide is as a

white pigment in a wide range of products, including paint, plastics, paper and inks

(Wang and Yuan, 2006).

5
Chapter 1: Introduction

In addition, white titanium dioxide is also an excellent opacifier, which means that

the pigment can beautify the outer surface of an object covering the underlying

surface from view. Therefore, TiO2 pigment has widespread use in general painting

applications for commercial and private dwellings. Moreover, unusual optical

properties appear when the average particle size of TiO2 is reduced to <100 nm,

including high transparency to visible light and high UV absorption (Zhang et al.,

2011a). Nanoparticles of TiO2 also cause some components of visible light to be

reflected and refracted differentially, leading to the phenomenon of iridescence.

Nanoparticles of TiO2 have found applications in the cosmetics, porcelains and

ceramics industries, as coating material and additives. Nanoparticles of TiO2 have

received great attention recently for their potential applications in catalysis and as

photo-electrochemical materials. The use of titanium dioxide in photocatalysis, by

irradiation of visible light has been rapidly developed in recent years (Zhang et al.,

2011a). Fig. 1.3 shows the percentages of white pigment used in various applications

around the world.

Fig. 1.3. Percentage of TiO2 used in various products (Egerton, 2000).

6
Chapter 1: Introduction

1.3. Treatments methods

The processes for production of pigment grade titanium dioxide and titanium

metal are schematically presented in Fig. 1.4. White titanium dioxide pigment is

produced by two processes, namely the sulfate process and the chloride process. The

two processes differ in both their chemistry and raw material requirement. The

sulfate process, which utilizes ilmenite as a raw material, is well known and widely

used but unfavorable. This is due to environmental concerns and the production of a

large volume of by-products (ferrous sulfate) that require high capital cost for

treatment. Alternatively, the chloride process, which utilizes rutile as a raw material,

offers a more economical flowsheet and generates less waste materials. Nowadays,

about 60% of the world’s TiO2 pigments are manufactured using the chloride

process, in which natural or synthetic rutile (SR), or titanium-rich slag are used as

feedstocks (Xue et al., 2009). The shortage of rutile feedstock has prompted the

upgrading of ilmenite to synthetic rutile. The schematic diagram of existing and

proposed processes of upgrading ilmenite is summarized in Fig. 1.5. The use of ore

with greater TiO2 content allows a pigment producer to reduce the amount of iron

sulfate by-product produced in the process (Egerton, 2000).

Sulfate pigment plants require sulfuric acid, ilmenite or titania slag. Low

levels of vanadium and chromium in feed material could affect the colour of the final

product. Chloride pigment plants usually use a high grade feed of natural or synthetic

rutile with a low content of iron, which consumes chlorine. Several processes have

been applied for ilmenite upgrading, but most are uneconomical, they involve

thermal oxidation and reduction by roasting, leaching and physical separation steps.

7
Chapter 1: Introduction

Iron is converted to elemental iron by reduction at high temperatures followed by

acid leaching and aeration to obtain synthetic rutile and iron oxide via soluble

iron(II) in 1% NH4Cl.

Fig. 1.4. A Schematic diagram of processes for the production of Ti metal and TiO2
pigment (Zhang et al., 2011a).

The growing titanium metal industry also relies on high grade rutile. All these

factors make the upgrading of ilmenite to synthetic rutile more and more important.

However, the upgrading processes are generally expensive due to the involvement of

multiple steps of energy sensitive thermo reductive conversion and leaching to

remove iron impurities. Titanium metal is commercially produced by thermo

chemical reduction processes using TiCl4 as the feed material. The low efficiency

and high energy consumption in batch operations make the thermo chemical

processes rather expensive (Zhang et al., 2011a).

8
Chapter 1: Introduction

Electrochemical and direct chemical reduction of TiO2 in molten CaCl2 to produce

pure titanium metal and its alloy has been developed in recent years as a potential

alternative to the conventional commercial processes.

Fig. 1.5. Summary of existing and proposed processes for upgrading ilmenite to
synthetic rutile (Zhang et al., 2011a)

Most commercial flow sheets involve processes which convert the ilmenite to

synthetic rutile without smelting, via solid state oxidation or reduction at solid state

and leaching in acid media. Two commercial processes have been developed during

the 1960s and 1970s. They are the Becher process and the Benilite process in

Western Australia and the United States of America, respectively (Egerton, 2000;

Zhang et al., 2011a). There is an increasing interest in developing hydrometallurgical

processes for ilmenite ores to avoid high energy consuming processes such as

smelting and slagging.

9
Chapter 1: Introduction

Some researches aim to produce TiO2 pigment directly from ilmenite ore by

acid leaching. Based on several reviews, a few main processes that successfully

upgraded ilmenite into high grade feedstock are summarized in Table 1.1 (Dimitrios

and Guillaume, 2009; Egerton, 2000; McConnel, 1978; Ward, 1990; Zhang et al.,

2011a). The details of these processes will be discussed in the next chapter.

Table 1.1
Summary of processes for upgrading ilmenite to synthetic rutile.

Process Pyro-treatment Leaching Advantage Disadvantage

Smelting or Iron is reduced HCl or H2SO4 Separation of iron Need large scale
slagging and melted to at an early stage operation to be
process separate the Fe simplifies the economical and
from Ti leaching process produce low grade
synthetic rutile (85%
TiO2)

The Dunn Selective N/A Cl2 recycle by Produce highly


process chlorination of oxidation of FeCl2 corrosive chlorine gas
iron in ilmenite to Fe2O3
with Cl2

The Becher Iron oxidized to (a) NH4Cl/O2 Environmentally Should have low
process Fe2O3 and friendly concentration of Mg
reduced to (b) 0.5 M and Cr in feedstock
metallic Fe at H2SO4
1200°C

The Benelite Convert Fe(III) to 18-20% HCl Involve only one Limited ilmenite type
process Fe(II) by carbo- step of iron as feed material
thermic reduction conversion

The Austpac Magnetisation of 25% (w/w) HCl Produces higher Require high
process ilmenite at 800- grade TiO2 (97%) concentration of acid
1000°C
(Zhang et al., 2011a)

10
Chapter 1: Introduction

1.4. Hydrometallurgical processes

In the past three decades, several new concepts have been developed to

overcome some economic and environmental problems that are inherent to the

industrial routes for the production of TiO2 pigment or titanium metal. Most of these

new processes were aimed to produce TiO2 pigment directly from low grade ilmenite

concentrates at a cost and environmental impact significantly lower than those of the

sulfate and the chloride process. The new processes offer a few advantages such as

low energy cost and fewer by-products. There are two processes that are still under

development and they are the Auspact and Altair processes (Walpole, 1999). The

details of these processes will be explained in Chapter 2.

The Altair process is a new alternative route to the sulfate and chloride

processes for producing white pigment from ilmenite by hydrochloric acid leaching

and solvent extraction (Verhulst et al., 2003). These processes do not involve

roasting of ilmenite, which is a high energy consuming stage. The

hydrometallurgical route has the potential to bring major cost benefits to the industry

(Mahmoud et al., 2004). However, none of these processes were developed beyond

the pilot stage (Dimitrios and Guillaume, 2009). To overcome the shortcomings

mentioned above, several new attempts have been made to increase the recovery of

titanium dioxide and minimize the operating and capital costs of current processes.

11
Chapter 1: Introduction

As a result, previous researchers investigated a few possible lixiviants which

can be divided into 3 different categories according to different leaching conditions,

ligands or reductants.

I. Acid leaching with sulfuric acid, hydrochloric acid, oxalic acid, ammonium

chloride or mixed chloride solution (e.g. NaCl, MgCl2) at atmospheric

conditions.

II. Reductive leaching using metallic iron as a reductant.

III. Alkaline leaching using lixiviants such as KOH and NaOH.

Some of the ilmenite ores are from massive rocks, sand minerals or have undergone

pre-treatment processes as outlined in Table 1.2. The table also depicts the extraction

efficiency of titanium (Ti) and iron (Fe) in different lixiviants, along with the reagent

concentrations and references.

Most research has been directed to improve the overall efficiency of the

dissolution reaction either by increasing the rate of dissolution or increasing the

extent of retention of titanium in the ore. This has been done on an empirical basis,

by noting the effect of changing the overall conditions of temperature, acid

concentration, acid to solid surface area ratio and various mineral pretreatment

methods. Despite the research interest in direct leaching or reductive leaching of

ilmenite, the proposed new processes are still at the early stage of development. The

fundamental understanding of the processes is still poor and further work should be

conducted in order to understand the reaction mechanism(s).

12
Chapter 1: Introduction

Table 1.2
Extraction of Ti and Fe from ilmenite ores using different lixiviants.

Extraction(%)
Lixiviants Concentration Other
Material References
(mol dm-3) reagents
Ti Fe
H2SO4 Imahashi and
SO 0.03 - 1 - NR NR
Takamatsu, 1976
MA H2SO4 9.2 - 65 90 Sasikumar et al., 2007
Ti(III)
SO H2SO4 4 -6 (4 g/L), NR NR Zhang and Nicol, 2010
SO2
SO HCl 11.3-11.6 - 85 NR Tsuchida et al., 1982

MI HCl 7.2 – 9.6 - 80 NR Olanipekun, 1999


Fe
MI HCl 7 (0.075g/g 90 0.8b Mahmoud et al., 2004
ore)
Fe
SO HCl 2-5 89 97 Lasheen, 2005
(5g -20g)

Ox & Re HCl 2-5 - NR 49 Sarker et al., 2006


Fe
HCl 98 81 El-Hazek et al., 2007
SO (0.1g/g ore)
MA HCl 3.4 – 8.6 - 92a 2.1b Li et al., 2008
Ox & Re,
HCl 7 - 90.5a 1.37b Zhang et al., 2011b
MA
MgCl2 Lakshamanan et al.,
SO HCl 7 96.9 84.8
( 300 g/L) 2008

Re HCl 10%-30% Gypsum NR 95 Gireesh et al., 2015

SO HCl 32% - 90 90 Haverkamp et al., 2016

NH4Cl, 1.5%(w/w)
Rtl Acetic acid 93 4.8 Jha et al., 2008
methanol 0.5%(v/v)

HCl, 5-7.5
SO CaCl2 (500 g/L Cl-) - 96 97 Das et al., 2013

a
calculated as TiO2 in residue sample
b
calculated as total iron in residue sample
SO : Sand ore
MI : Massive ilmenite
MA: Ilmenite undergone mechanical activation
Rtl: Rutile
Ox&Re: Undergone oxidation and reduction
NR : Not reported

13
Chapter 1: Introduction

1.5. Dissolution studies of ilmenite in hydrochloric acid

The interest in the dissolution study of ilmenite in hydrochloric acid, compared

to that in sulfuric acid leaching has increased during the last two decades

(Haverkamp et al., 2016; Gireesh et al., 2015; Sasikumar, 2007; El-Hazek et al.,

2007; Ibrahim et al., 2003; van Dyk et al., 2002; Lanyon et al., 1999). Further studies

on hydrochloric acid leaching are important for several reasons:

1) Lack of hydrochloric acid leaching studies on various types of Australian

ilmenite.

2) Lack of published data on the kinetics and mechanisms of the dissolution of

ilmenite in hydrochloric acid.

3) Hydrochloric acid is more favorable than sulfuric acid due to environmental

concerns.

4) Recent technology has been developed for recycling the acid by hydrolysis and

pyrohydrolysis that can reduce operating costs and protect the environment.

Hydrochloric acid leaching gives a high quality product of up to 96% TiO2, and

hydrochloric acid offers a few advantages compared to sulfuric acid. Table 1.3 lists

the advantages and disadvantages of chloride and sulfate leaching.

14
Chapter 1: Introduction

Table 1.3
Comparison of sulfuric acid and hydrochloric acid leaching.

Process Advantages Disadvantages References

Hydrochloric acid 1. Allow easier 1. Corrosive. Zhang et al.,


regeneration of free HCl 2011a
from its waste solution 2. Requires high grade
by hydrolysis or ilmenite concentrate.
pyrohydrolysis
processes.

2. Allows separation of
metal from waste
solution.

3. Low environmental
impact due to zero
effluent disposals.

Sulfuric acid 1. Use of low grade 1. Produce large volume Dimitrios and
ilmenite concentrates. of waste (8-10 ton of 20% Guillaume, 2009
H2SO4 and FeSO4 waste
for one ton TiO2).

2. Requires high capital


cost (for waste water
treatment)

3. Slow leaching rate


(demand more energy)

Recent advances in acid regeneration technology have made HCl leaching

processes more attractive to the point where two processes have been proposed as

options for treating ilmenite in Australia. These are the Murso process (Sinha, 1973)

and the ERMS process (Walpole, 1999). These processes are just several variations

of the acid leach option. All the variations use a pre-leach treatment that converts

most of the iron content to the ferrous state and use HCl as the leaching agent.

Leaching is normally carried out at temperatures near the boiling point, but in some

cases elevated pressure is used to allow higher leaching temperatures. Jackson and

Wadsworth (1976) found that the acid concentration and temperature can

significantly affect the dissolution behaviour of ilmenite.

15
Chapter 1: Introduction

Tsuchida et al. (1982) observed that stirring speed had no significant effect

on titanium and iron extraction. Some researchers also investigated the effect of

adding a reducing agent such as iron to the hydrochloric acid leach medium (El-

Hazek et al., 2007; Lasheen, 2005; Mahmoud et al., 2004). They found that the

addition of a reducing agent could selectively leach iron from ilmenite and also

increase the rate of iron dissolution. Furthermore, Lakshmanan et al. (2008) studied

the effect of mixed brine in hydrochloric acid solution (6 M) and by residue analysis.

The X-Ray diffraction pattern showed a high intensity peak of rutile and about 95%

TiO2 recovery from the ore.

Electrochemical techniques also have been used by various researchers

(Adriamanana et al., 1984; McConnel, 1978; Zhang and Nicol, 2009) to investigate

the reduction and dissolution of ilmenite in sulfuric acid. White et al. (1994)

performed an electrochemical study of an ilmenite sample high in Fe(III) in 0.1 M

NaCl solution in the pH range from 1 to 7. The electrochemical studies of ilmenite in

H2SO4 have been well documented. However, electrochemical studies in

hydrochloric acid solution are scarce. A previous study of reduced ilmenite in the

Becher process (Kumari et al., 2002) focused on electrochemical aspects in

ammonium chloride leaching of reduced ilmenite.

It is important to extend the electrochemical studies to rationalize the

dissolution behaviour of ilmenite in hydrochloric acid solutions. It is also essential to

study a wide range of different types of ilmenite feed material to establish the

mechanism of dissolution reactions, in order to minimize the reagent consumption

and develop further processing routes for ilmenite to produce titanium dioxide

pigment.

16
Chapter 1: Introduction

1.6. Objectives

The main objectives of this thesis are:

(i) To briefly review the literature relevant to ilmenite processing with special

attention to proposed methods for leaching.

(ii) To conduct experimental studies on the effect of physico-chemical factors

that will improve the efficiency of batch leaching of ilmenite.

(iii)To investigate the heterogeneous dissolution of synthetic ilmenite disc under

controlled and well defined hydrodynamic conditions.

(iv) To study the kinetics and mechanism of dissolution for particulate ore and

flat surfaces of synthetic ilmenite.

(v) To investigate the electrochemistry of ilmenite dissolution in hydrochloric

acid solutions.

(vi) To compare results based on different types of ilmenite and different

techniques.

17
Chapter 2: Literature Review

2 CHAPTER 2

LITERATURE REVIEW

2.1. Mineralogy and occurrence

2.1.1. Titanium mineralogy

Titanium was discovered for the first time by William Gregor in 1791 in

ilmenite from Manaccan Valley, England (McConnel, 1978). Only after late 1946,

the metal started to be recognized for its uses and produced commercially. Titanium

can be found in igneous and metamorphic rocks and their sediments. The main

mineralogical occurrences are oxides, titanates and silicatitanates depending on the

silicon, iron, calcium and magnesium contents. Fig. 2.1 shows the classification of

titanium minerals and their components. Rock magmas rich in silica and poor in the

base-forming elements (iron and magnesium) contain titanium in the oxide form.

They consist of titanium dioxide (TiO2) in the three different mineralogical forms

(anatase, brookite and rutile) as shown in Fig. 2.1. Relatively high proportions of

calcium and silicon yield calcium titanate (CaTiO3, perovskite), while higher iron

and lower acid forming oxide (silica) content result in iron titanates such as ilmenite

(FeTiO3) and arizonite (FeTi2O5) (Barksdale, 1966).

18
Chapter 2: Literature Review

Fig. 2.1. Titanium minerals (Barksdale, 1966; Zhang et al., 2011a).

Among all the titanium minerals, rutile is the most beneficial mineral to

mine for the extraction of titanium metal because of the higher percentage of

titanium in its formula. However, it is not present at high concentration in igneous

deposits like ilmenite. The shortage of rutile deposits has resulted in ilmenite ore

being the main resource for the production of titanium metal and titanium dioxide

white pigment.

Ilmenite occurs naturally in three forms : (i) as massive ilmenite or rock

deposits, often in association with other minerals such as magnetite or hematite, (ii)

as a constituent of beach sands; (iii) as an accessory mineral in rocks (McConnel,

1978). The two forms that are commercially important are the relatively rare massive

deposits and the more widely occurring beach sand deposits.

19
Chapter 2: Literature Review

Most of the world’s reserves of ilmenite are in the form of rock deposits such

as Allard Lake, Canada, and Adirondack, New York, deposits. However, producers

of ilmenite ore have always favored sand deposits because they are the easiest and

cheapest to process (Barksdale, 1966). Essentially, the world’s production of rutile is

obtained from sand deposits. A survey of the world’s reserves of titanium ores and

their occurrence is given by Barksdale (1966).

In sedimentary deposits, known as heavy minerals sands (HMS) or placers,

both rutile and ilmenite can be concentrated into useable ores and there are enough

of these deposits around the world (Pownceby et al., 2008). In Australia, HMS is one

of the major economic concentrations occurring along the east and southwest coasts

as well as inland (Pownceby et al., 2008). During the formation of ancient shoreline,

the titaniferous oxide minerals were originally concentrated within a variety of

metamorphic and mafic igneous rock. These Mesozoic formations appear to have

been eroded and transported from streams and rivers and re-deposited in the ocean

(Baxter, 1977). Ocean currents and wave action separated the finer sand fractions,

resulting in the concentration of heavy minerals along the coastal ridge. Due to the

shore-line gradually receding from the coastal ridge, the heavy mineral sand deposits

were left close to the surface and inland from the present coastline. The primary

economic minerals consists of ilmenite (FeTiO3), leucoxene (FeTiO3), pseudorutile

(Fe(III)2Ti3O9), rutile (TiO2), zircon (ZrSiO4), and minor monazite (CePO4). Other

associated minerals may include anatase (TiO2), spinel (AB2O4), garnet

(Fe(II)3Al2Si3O12), staurolite (Fe(II)2Al9Si4PO23(OH)), quartz (SiO2), goethite

(FeOOH) and tourmaline (NaFe(II)3Al6Si6B3O27(OH)4). A list of minerals

commonly associated with HMS deposits is shown in Table 2.1 (Pownceby et al.,

2008).

20
Chapter 2: Literature Review

Table 2.1
List of minerals commonly associated with heavy mineral sand deposits.

Mineral Ideal formula Major impurities Specific Gravity

Ilmenite Fe(II)TiO3 Mg, Mn, V, Nb, 4.7-4.8


Fe(III)

Rutile(Anatase-low TiO2 Nb, Ta, Sn, Fe 4.2-5.5


polymorph)
(3.8 -4.0)

Pseudorutile Fe(III)2Ti3O9 Mg, Mn, Fe(III), Cr, 3.3-3.8


OH,(H2O,Al2O3,SiO2)a

Leucoxene An alteration product and Mg, Mn, Fe(III), Cr, 4.3-4.6


mixture of Fe-oxides, Most V, Nb, Ca, P, (H2O,
‘leucoxene’ is anatase or Al2O3, SiO2)a
rutile

Zircon ZrSiO4 Hf, Fe, Al, U, Th 4.6-4.7

Monazite CePO4 La, Y, Ca, Th, U, Al, 5.0-5.3


Fe(III)

Spinel (General Fe(II)Cr2O4-chromite Mn, Zn, Fe(III) 5.1


name for mineral
group with the Fe(II)Al2O4-hercynite
formula –AB2O4)
MgCr2O4-
magnesiochromite

MgAl2O4-spinel

Fe(II)2TiO4-ulvospinel

Fe(II)Fe(III)2O4-Magnetite

Garnet Fe(II)3Al2Si3O12- almandine Ca, Mn, Mg, Ti, Cr, 3.5-4.3


Fe(III), OH-

Staurolite Fe(II)2Al9Si4PO23(OH) Fe(III), Mg, OH- 3.7-3.8

Tourmaline NaFe(II)3Al6Si6B3O27(OH)4- Mg, Mn, Li, F- 3.0-3.2


schorl

Sillimanite Al2SiO5 Fe(III) 3.2-3.3(3.5-3.6)


(kyanite-high
P polymorph)

Goethite FeO.OH Si, Mn(III), Al ~4.3

Quartz SiO2 Al, Fe(III), Ti 2.6-2.7


a
These species are commonly co-precipitated with, or adsorbed onto the alteration product during weathering
(Pownceby et al., 2008)

21
Chapter 2: Literature Review

During the separation/concentration process the ore is subjected to gravity,

magnetic and electrostatic separation processes to produce an ilmenite concentrate.

With hard rock ilmenite the ore is crushed and any coexisting magnetite and/or

hematite removed. The material is then treated in a similar fashion to beach sand to

remove gangue minerals such as quartz, feldspar, garnet and recover other ores

(McConnel, 1978). The heavy mineral concentrate is treated via a physical

separation circuit to selectively separate ilmenite (which is sent to synthetic rutile

plant) from the remaining leucoxene, rutile, zircon, staurolite, monazite and quartz

which are all produced as separate products (von Horn, 1993). The HMS is screened

to remove remaining aluminosilicates from the mineral surface. The HMS is then

separated to conducting (ilmenite/rutile) and non conducting (zircon/staurolite, etc.)

streams in the primary high tension circuit (von Horn, 1993). The flowsheet of the

separation plant has been described in Tiwest Joint Venture report (von Horn, 1993).

Many of the world’s heavy mineral beach sand deposits contain minerals that

are formed as a result of the natural weathering and alteration of ilmenite (Gambogi,

2009). This process is very important since it not only influences the titanium

content of the ore but also appears to change its chemical behaviour. Ilmenite

concentrates that contain significant levels of impurities, either internal or external to

the grains, have a lower market value and therefore, there is considerable incentive

for producers to carefully characterize the nature and mode of occurrence of these

impurities.

22
Chapter 2: Literature Review

2.1.2. Natural weathering of ilmenite

Ilmenite and altered ilmenite from sand deposits are the major (or abundant)

titanium resources in Western Australia (Deysel, 2007). As a result of oxidation,

exsolution and hydrothermal processes, ilmenite forms an assortment of intergrowths

with other titaniferrous oxides (Deysel, 2007). Prolonged exposure to oxidizing

and/or acidic environments may also cause changes to the chemistry of ilmenite

grains, with elements such as Fe and Mn being significantly leached. During this

process, the ilmenite may slowly weather through two main phases, pseudorutile and

leucoxene, to produce rutile. However, before the weathering process is described, it

is necessary to define the nomenclature of the alteration products (Reyneke and

Wallmach, 2007).

Ideally, ilmenite has a composition of 52.7% TiO2 and 47.3% FeO, but usually

includes some iron in the form of iron(III) (Egerton, 2000). Fig. 2.2 shows the TiO2

content %(w/w) range in titaniferous oxide phases. It can be seen that ilmenite

theoretically averages approximately 52% TiO2, whilst pseudorutile is classified as

being approximately 60% TiO2 and leucoxene approximately 68% to 90% TiO2

(Deysel, 2007).

As stated previously, the alteration process involves the removal of iron from

the ilmenite lattice through leaching. The two stage continuous alteration process of

ilmenite through pseudorutile to leucoxene may be described by the Grey and Reid

(1975) model. Pseudorutile has also been identified as one of the products of the

higher temperature oxidation of ilmenite (McConnel, 1978).

23
Chapter 2: Literature Review

During the first stage, ilmenite is converted to pseudorutile by oxidation of

all of the Fe(II) to Fe(III) and one third of the iron is lost through leaching (Eq. 2.1)

(Deysel, 2007).

6Fe(II)TiO3 + 3H2O + 1.5O2→ 2Fe(III)2Ti3O9 + 2Fe(OH)3 (2.1)

During the second stage, dissolution and re-precipitation occurs simultaneously. That

is, the titanium and iron are both dissolved, but the titanium is redeposited

immediately after leaching of the iron (Deysel, 2007). This can be described by Eq.

2.2:

Fe(III)2Ti3O9 + 3H2O → Fe(III)Ti3O6(OH)3 + Fe(OH)3 (2.2)

The continued removal of iron from the pseudorutile grain converts it to rutile. This

stage involves disruption of the oxygen lattice with the removal of both iron and

titanium according to the equation below:

Fe(III)2Ti3O9→ 3TiO2 + Fe2O3 (2.3)

24
Chapter 2: Literature Review

Fig. 2.2. Theoretical TiO2 content (%, w/w) in titaniferous oxide phases (Reyneke
and Wallmach, 2007).

A detailed study of ilmenite concentrates using combined techniques of

chemical analysis, X-ray, magnetic concentration and optical microscopy was

conducted for the first time in 1954 (McConnel, 1978). Since then, ilmenite

concentrates have been analyzed by various technical methods such as X-ray

diffraction (XRD), X-ray fluorescence (XRF), scanning electron microscopy

(SEM), energy dispersive X-ray analysis (EDX) and automated analytical methods

(QemSCAN and XRD) (Reyneke and Wallmach, 2007). These methods have been

successfully used by researchers to study the mineralogical variations that reflect

various stages and processes of alteration. Figs. 2.3-2.5 show the SEM micrograph of

ilmenite concentrates. The figures illustrate the extent of alteration from unaltered

ilmenite to different stages of alteration.

25
Chapter 2: Literature Review

In addition to the compositional variation of the ilmenite grains themselves,

the HMS deposits can also vary significantly in the abundance and types of gangue

mineral phases present. For example, Table 2.2 shows the reported chemical

composition of various ilmenite ores of different origin, which significantly affects

the reactivity of titanium and iron towards different types of lixiviants (Chernet,

1999).

Fig. 2.3. Backscattered image of ilmenite grains. Progressive alteration of ilmenite


depicting (a) unaltered ilmenite, (b) incipient alteration of ilmenite along grain
boundaries and fractures, (c) extensive alteration throughout the grain and (d)
‘spotty’ alteration of ilmenite involving the formation of hematite, pseudorutile,
ferropseudobrookite and rutile/anatase (Reyneke and Wallmach, 2007).

26
Chapter 2: Literature Review

(a) (b)
Fig. 2.4. Backscattered image of (a) altered ilmenite containing various Ti-bearing
phases, including remnant ilmenite, pseudorutile, ferropseudobrookite and
rutile/anatase; (b) porous secondary particle containing ilmenite, pseudorutile,
ferropseudobrookite and rutile/anatase and associated with the Si-contaminants
(Reyneke and Wallmach, 2007).

(a) (b)

Fig. 2.5. Optical micrograph of ilmenite grains (a) exsolved ilmenite-hematite with
curvature of the lamelle suggest that the grain was subjected to physical strain after
its crystallization; (b) ilmenite altered to pseudorutile and ferropseudobrookite along
cracks and grain boundaries (Reyneke and Wallmach, 2007).

27
Chapter 2: Literature Review

Table 2.2
Chemical compositionof ilmenite from different origin.
References Oxide and metal composition % (w/w)
TiO2 Fe2O3 FeO SiO2 MgO V2O5 MnO Cr2O3 CaO Al2O3 P2O5 ZrO2 Ti Fe Mg
Tsuchida et al., 1982 52.0 0.4 3.1 0.04 2.30 31.2 27.2
Olanipekun, 1999 26.4 35.2 1.8
Lanyon et al., 1999 59.1 36.5 0.6 0.2 0.04 0.01 0.66 0.04 0.16 35.5
Ogasawara et al., 2000 47.0 41.3 6.8 0.3 4.53 28.2 32.1
van Dyk et al., 2002 43.3 47.8 0.9 0.6 0.3 0.8 0.07 26.0
Sarker et al., 2006 39.5 33.0 24.2 23.7 41.8
Li et al., 2008 47.3 5.6 34.2 2.6 6.2 0.1 0.002 1.08 1.49 28.3 30.5
Mahmoud et al., 2004 41.4 28.6 24.4 2.4 0.6 0.4 0.4 0.36 0.15 0.63 0.02 24.8 39.0
Lasheen,2005 46.7 20.9 27.5 0.6 0.8 0.2 1.2 0.29 0.44 0.80 0.26 26.4 37.0
El-Hazek et al., 2007 44.0 21.4 28.5 0.8 0.8 0.2 1.2 0.29 0.44 0.80 0.26 26.3 35.9
Lakshmanan et al., 2008 4.7 2.8 0.13 22.8 38.0
Sasikumar et al., 2007 55.1 19.9 20.3 1.5 1.0 0.2 0.10 0.80 0.10 33.0 22.7
Imahashi and Takamatsu, 1976 52.9 46.1 1.6 31.7

28
Chapter 2: Literature Review

2.1.3. Crystal structure

The ilmenite crystal structure was first determined by Barth and Ponsjak

(1934). The magnetic structure at low and high temperature and hematite-ilmenite

solid solution was later elucidated by several other researchers (Ishikawa, 1958b;

Ishikawa, 1962; Raymond and Wenk,1971; Shirane et al., 1959). Preliminary studies

predicted that, on the basis of the close similarity of its powder pattern with hematite,

ilmenite would have essentially the same structure as hematite (Rumble, 1976). This

is because hematite and ilmenite form a solid solution series that is complete at high

temperatures (above 960 °C) but interrupted by a miscibility gap at lower

temperature (Rumble 1976).

Fig. 2.6 shows the crystal structure of ilmenite and hematite. Ilmenite is a

heavy mineral of metallic to submetallic luster, crystallizing in the trigonal system

with specific gravity of 4.5 to 5.0, hardness of 5 to 6 and molar volume of

30.4 cm3 mol-1. There is controversy in determining the valence of metal ions in

ilmenite, whether the formula is Fe2+Ti4+O3 or Fe3+Ti3+O3. The X-ray adsorption

edge of ilmenite, showed that it was much closer to Mg2+Ti4+O3 and Ti4+O2 than

(Ti3+)2O3 and this was confirmed by Mössbauer-effect studies. Thus, the formula for

ilmenite is considered to be Fe2+Ti4+O3 (Rumble, 1976). The general formula for the

group is ATiO3, where A can be iron, magnesium, zinc and/or manganese (Lindsley,

1976). The ilmenite group members differ from the other members of the hematite

group in that the structure is more ordered, with the titanium and A ions occupying

alternating layers between the oxygen layers.

29
Chapter 2: Literature Review

(a) (b)

Fig. 2.6. Unit cell model of (a) ilmenite and (b) hematite. Titanium is shown in
green, iron in red and oxygen in grey.
(http://www.geocities.jp/ohba_lab_ob_page/structure6.html).

The oxygen layers are hexagonally packed. Each metal ion is bonded to three

oxygens in the layer above and three oxygens in the layer below (Lindsley, 1976).

Table 2.3 depicts a summary of the crystal system and cell constants for the different

oxides that can be related to ilmenite. However, the morphology of natural ilmenite

crystal has shown that the space group must have a lower symmetry than that of

hematite; R3с (a subgroup of R3) was chosen for ilmenite on this basis. The

structure and space group was confirmed through single crystal X-ray studies

(Lindsley, 1976).

30
Chapter 2: Literature Review

Table 2.3
Summary of crystal structures of ilmenite and related oxides.

Mineral References
Formula System Cell constants (Å)
name

a =5.0881± 0.0005 Lindsley, 1976


Ilmenite FeTiO3 Rhombohedral
c = 14.29± 0.0005

α = 5.035 Burkin, 2001


Hematite α-Fe2O3 Rhombohedral
c = 13.72

Cubic (fcc-inverse Burkin, 2001


Magnetite Fe3O4 α = 8.33-8.38
spinel)

a = 5.137 Lindsley, 1976


Pyrophanite MnTiO3 Rhombohedral
c = 14.29

a = 5.054 Lindsley, 1976


Geikielite MgTiO3 Rhombohedral
c = 13.898

Ilmenite solid solution is a p-type semiconductor with resistivity decreasing

from 1 to 0.1 ohm m as the hematite component increases from 5 to 50% (Ishikawa,

1958a). This high resistivity, relative to magnetite, is due to higher band gap energy

for ilmenite (2.8 eV) (White et al., 1994). The electrical conductivity occurs within

iron layers of the basal plane of ilmenite (White et al., 1994). At room temperature

pure ilmenite is essentially paramagnetic, but at sufficiently low temperature it

becomes anti-ferromagnetic. In ferromagnetic ilmenite, all Fe2+ ions in a given layer

have parallel spin, but alternate iron layers have antiparallel spin (Ishikawa, 1958a).

31
Chapter 2: Literature Review

2.2. Production of synthetic rutile and TiO2 pigment

2.2.1. Sulfate process

The sulfate process for manufacturing titanium dioxide pigments was the

earliest process used, having been first practiced in 1916 in Niagara Falls, New York

and in 1915 in Norway (Egerton, 2000). All pigment producers who supplied

pigment prior to the late 1950s used the sulfate process (Egerton, 2000). At present,

43% of all TiO2 pigments are produced via the sulfate process. Normally, the

feedstocks for the conventional sulfate process consists of either low grade ilmenite

(44% TiO2) or Ti slag (>70% TiO2). However, the direct leaching of ilmenite in

sulfuric acid does not require a specific grade of TiO2, which means that it can

eliminate the ilmenite upgrading process to produce synthetic rutile (SR). Thus,

different kinds of ilmenite ores have been processed using sulfuric acid lixiviants in

many countries since 1976 until recently (Imahashi and Takamatsu, 1976; Li et al.,

2007; Liang et al., 2005; McConnel, 1978; Sasikumar et al., 2004; Vaughan and

Alfantazi, 2006; Zhang and Nicol, 2010).

The sulfate process is a batch process which utilizes a large number of relatively

simple unit operations. The primary unit operations are:

1. Digestion: The ground ilmenite or titanium slag is digested in concentrated

sulfuric acid at 150-180°C to dissolve the titanium and iron according to Eq.2.4

(Liang et al., 2005).

FeTiO3 + 2H2SO4 → TiOSO4 + FeSO4 + 2H2O (2.4)

32
Chapter 2: Literature Review

2. Precipitation: the ferric ions in solution are reduced with scrap iron to ferrous

sulfate, which is then cooled and precipitated as FeSO4.7H2O.

3. Calcination: thermal formation of pure TiO2 crystals.

The environmental problems and economic consideration caused by the

sulfate route make this process unfavorable. A flowsheet of the sulfate process

recently developed by the BHP Billiton researchers is shown in Fig. 2.7. The kinetics

of dissolution of ilmenite in sulfuric acid are slow (10 to 12 hours per batch), costly

and the ferrous sulfate by-product is not marketable and poses an environmental

hazard.

Fig. 2.7. A schematic sulfate flowsheet recently developed by BHP Billiton


researchers (Stuart et al., 2010).

33
Chapter 2: Literature Review

2.2.2. Chloride process

The chloride process for producing TiO2 pigments, invented in 1950,

separates titanium from the host mineral in the form of titanium tetrachloride (TiCl4),

purifies it, converts it to TiO2 pigment in an oxidation reactor and recycles the

chlorine. The principal unit operations are:

1. Chlorination

Reaction of chlorine and a titanium containing mineral under reducing conditions

typically takes place between 800 – 1000 °C according to the reaction shown in Eq.

2.5 (Ward, 1990).

TiO2 + 2Cl2 + C → TiCl4 + CO2 (2.5)

2. Condensation and purification

The titanium tetrachloride is purified by fractional distillation to remove the other

chlorides since practically everything present in the mineral is chlorinated.

3. Oxidation

After purification the titanium tetrachloride is heated to 1000°C in the presence of

oxygen to produce pure TiO2 (McConnel, 1978; Ward, 1990). The chlorine gas

regenerated in the reaction (Eq. 2.6) is recycled for use in the initial chlorination

step. The flowsheet for the chloride process is shown in Fig. 2.8.

TiCl4 + O2 → TiO2 + 2Cl2 (2.6)

34
Chapter 2: Literature Review

Fig. 2.8. A simplified flowsheet of the chloride process (Dimitrios and Guillaume,
2009).

The chloride process for producing titanium dioxide was commercialized in

1958 in the United States (Egerton, 2000). Because of its complexity and demanding

requirements for materials of construction, successful large scale commercial

operation of the chloride process was established after 4 decades. Since then, the

chloride process has been a major process for producing titanium dioxide (Egerton,

2000). However, there are some advantages and disadvantages of these two

processes to be compared. The process features are compared in Table 2.4.

35
Chapter 2: Literature Review

Table 2.4
Comparison of sulfate processes with chloride processes.

Process Feed Product Advantage Disadvantage

Traditional Natural ilmenite Anatase Processing low High H2SO4


sulfate leaching grade ilmenite; consumption;
process (>44 TiO2) (>98% TiO2)
Low capital cost; Large iron sulfate
Ti slag (for papers, waste and dilute
ceramics and inks) Low energy H2SO4
(78% TiO2) consumption; production.

Simple technology

BHP Billiton Diverse ilmenite >99% TiO2 with Reduced waste Increased process
improved sulfate ores; Fe scrap solvent extraction; complexity
process reductant Producing clean
>97% with gypsum; Recycling large
crystallization volume of dilute
Better selectivity acid solution
by SX to produce
purer products

Thermo chloride 90% TiO2 Pigment grade Recycle of HCl Need for higher
processes natural or products (> 99.5% reagent grade feed;
synthetic rutile TiO2);
Suitable for larger More corrosive
High grade Ti Ti Metal scale
slag High energy
Purer products consumption;

Environmentally Complicated
acceptable waste technology

Chloride Natural ilmenite >99.5% TiO2 More complete Higher capital


leaching process & Ti Slag HCl acid recycle; costs for
equipment and
Lower cost for construction;
waste and more
environmentally Higher
acceptable; requirement for
operation and
Higher purity maintenance skill
products

Caustic process Natural ilmenite >99.3% TiO2 Higher leaching Need for
& Ti slag selectivity for Ti transformation of
over Fe; titanate to
hydrous TiO2 on
Mild leach acidic solution;
conditions at low
temperature and Recycling of large
atmosphere amount of KOH
solution and
energy
consumption
(Zhang et al., 2011a)

36
Chapter 2: Literature Review

2.2.3. Upgrading ilmenite to synthetic rutile

Titanium dioxide pigment has traditionally been produced by the sulfate

process and the chloride process as described previously. In the chloride process,

rutile (TiO2) is utilized as a raw material. However, due to the shortage of natural

rutile, extensive research has been carried out to convert ilmenite to high titanium

slag or synthetic rutile, which actually supplies the major feedstock for the

chlorination process (Becher et al., 1965; Egerton, 2000). Pyrometallurgical

production of this feedstock includes mainly ilmenite smelting and roasting.

Smelting produces a high titanium slag and low manganese iron byproduct and this

is mainly conducted in South Africa, Canada and Norway. On the other hand, solid

state reduction involves several processes: the Becher Process (developed in

Australia), the Benilite Process used in USA, the Murso Process, the Dunn process,

the Kataoka process (Japan) and the Austpac process (Dimitrios and Guillaume,

2009; Zhang et al., 2011a). Table 2.5 shows a summary of processes used for

upgrading ilmenite to synthetic rutile.

37
Chapter 2: Literature Review

Table 2.5
Processes of upgrading ilmenite to synthetic rutile (SR).

Process/Country Feed (% TiO2) Pyro-treatment Leaching SR (% grade) Advantages Disadvantages References

The Becher process/ 40-65 (the rest being Iron oxidized to (a) 1% NH4Cl/O2 90 Allowing diverse Multistep Becher at al., 1965;
Western Australia iron oxide) Fe2O3 and reduced to ilmenite ores as feed conversions and
metallic Fe by coal at (b) 0.5M H2SO4 leaching; high Zhang et al., 2011a
1200oC energy consumption;
emission of CO2

The Benilite 55-62 Carbon-thermo Pressure leaching 95 Simple one step Limited ilmenite Dimitrios and
process/USA, reduction with 18-20% HCl conversion type as the feed Guillaume, 2009
Malaysia

The Murso process Pre-oxidation step; 20% HCl 95-96 Improving efficiency Similar to Becher Sinha, 1973;
hydrogen rich by using fluidized process
reductant beds.; Easier HCl Zhang et al., 2011a
recycle than sulfate
system

The Auspact process Magnetisation of the 25% (w/w) HCl >97 Magnetic separation Higher acidity Dimitrios and
ilmenite at 800- for higher>97% TiO2 needed for leaching Guillaume, 2009
1000 oC remaining magnetic
iron form

The Dunn process Selective - Cl2 recycle by Handling highly Zhang et al., 2011a
chlorination of iron oxidation of FeCl2 to corrosive Cl2
in ilmenite with Cl2 Fe2O3

The Kataoka process Conversion to H2SO4 95 Less corrosive using Large quantities of Zhang et al., 2011a
/Japan ferrous form H2SO4 than HCl ; sulfate wastes
Low leaching
temperature

38
Chapter 2: Literature Review

A number of patents and recent publications have highlighted the importance

of direct hydrometallurgical treatment of ilmenite to obtain TiO2 pigment without

production of intermediate synthetic rutile. Some of these processes, still at a bench

or pilot stage, would eliminate many intermediate steps such as pre-reduction and

make the processes more economical (Lakshmanan et al., 2008; Zhang et al., 2011b;

Walpole, 1999). These new processes will also involve other hydrometallurgical unit

operations such as precipitation and solvent extraction (Lakshmanan et al., 2008).

The common lixiviants suitable for direct leaching of ilmenite are sulfuric acid,

hydrochloric acid, mixed brine and alkali such as KOH and NaOH. Thermodynamic,

kinetic and electrochemical aspects related to sulfate and chloride leach systems are

described in the following sections.

2.3. Direct leaching of ilmenite

2.3.1. Eh-pH and speciation diagrams

Appendix A1 summarizes the relevant thermodynamic data for the Fe-Ti-

H2O system, including balanced chemical reactions, standard free energy and

entropy data for species, equilibrium constants and standard reduction potentials.

The Eh-pH diagrams are important in the identification process of the predominant

species present at equilibrium with respect to reduction potential and the hydrogen

ion concentration.

39
Chapter 2: Literature Review

The diagrams represent the potential (Eh) and acidity (pH) regions in which

the elements and their compounds are stable and provide insights into possible

decomposition products. In general, the thermodynamics of the dissolution of

ilmenite can be summarized by the Eh-pH diagram of Ti-Fe-H2O system in Fig. 2.9.

Fig. 2.9. Eh-pH diagram of the Ti-Fe-H2O system at 25°C, (excluding Ti2+)
considering TiO2 (c, hydrated), Ti2O3 (c, hydrated) and Fe2O3 as the solid titanium
oxide and Fe(III) phases, and for 0.1 and 0.01 dissolved iron and titanium activities,
respectively (Solid line refers to the stability area of FeTiO3) (Kelsall and Robbins,
1990).

40
Chapter 2: Literature Review

The reductive dissolution of ilmenite to Fe2+ and Ti3+ ions is predicted to be

occurring at low Eh. Reductive dissolution of Fe(OH)3/Fe2O3 is also predicted at low

Eh which may have kinetic benefits for FeTiO3 dissolution, since ex-solution

hematite is often found in ilmenite, presumably as a result of aerial oxidation of an

ore body (Zhang and Nicol, 2009), described in the previous section.

At sufficiently low Eh, reductive decomposition of FeTiO3 is predicted to

form Ti2O3. The relative leaching behavior of titanium in HCl can also be

rationalized using the Eh-pH diagram for the Ti-Cl-H2O system at 25oC using 10-6 M

Ti and 0.1 M Cl-shown in Fig. 2.10 (Vaughan and Alfantazi, 2006). The figure

indicates that Ti dissolves in HCl as TiOCl42-only under oxidizing conditions and in

a very limited range of pH (up to ~1.5).

Fig. 2.10. Ti-Cl-H2O system (10-6 M Ti, 0.1 M Cl-, 25oC) (Vaughan and Alfantazi,
2006).

41
Chapter 2: Literature Review

At lower Eh, titanium(III) hydrolyses to form Ti(OH)2+. At higher pH it

precipitates as TiO2.H2O or Ti(OH)3 depending on Eh. This implies that in

hydrochloric acid solution titanium can be precipitated under reducing conditions

and at lower acid concentration even at ambient temperature. The phase stability

boundaries in the Eh-pH diagram are expected to be different at higher temperatures.

Free energy data at higher temperatures have not been published (Sasikumar, 2007).

2.3.2. Kinetic models

In hydrometallurgical processes, leaching is carried out either by agitating

particulate solid in a solution containing the reagents required or by causing the

solution to flow through a porous solid phase, which may be a porous rock or a heap

of broken material. The most important practical requirement when designing a

leaching process is to understand what chemical reactions take place and what

factors control their rates. The kinetic information can only be derived from

experimentation and observation and it is influenced by a number of factors such as

mineralogy, surface area, reactant concentrations, product layer formation and

temperature (Velasquez-Yevenas et al., 2010). The most convenient way of

determining this is to use samples of ground ore or mineral of narrows size range, if

the behaviour of a single substance is being studied (Burkin, 2001). For the purpose

of process control it is convenient to fit a mathematical equation to the curve

showing recovery of the metal values by leaching as a function of time.

42
Chapter 2: Literature Review

The most widely used models describing fluid-solid reaction kinetics of dense

particles are shrinking sphere/core models (Burkin, 2001). A shrinking sphere model

describes a system in which either the rate of the reaction taking place on the surface

of the particle or mass transport of the reactant(s) to the surface is limiting the

overall rate of reaction. The rate of the process is independent of hydrodynamic

conditions where the order of the reaction with respect to the lixiviants can vary in

the range 0 to 2. Activation energy is larger than 40 kJ mol-1 for the chemically

controlled reactions. This model can be expressed in the form of Eq. 2.7 (Burkin,

2001), for the reaction A(a) + bB(s) = products, in which X = the fraction of metal

dissolved at time t, kss = apparent rate constant, [H+] = concentration of acid in bulk

solution (mol cm-3), r = initial particle radius (cm), ρ = molar density of the

dissolving metal in the initial particle (mol cm-3) and b = stoichiometric factor. Note

that kss is inversely proportional to the initial particle radius.

1-(1-X)1/3 = b [H+]bulkkr-1ρ-1t = ksst (2.7)

In contrast, a shrinking core model describes the system in which the leaching rate is

limited by diffusion of the reagent to the surface of the mineral through a porous

layer of reaction product. The activation energy is low, usually up to 25 kJ mol-1, for

the diffusion of ions and molecules in porous media filled with liquid. For solid state

diffusion processes the activation energy could be much higher (Barton and

McConnel, 1979).

43
Chapter 2: Literature Review

This model can be expressed in the form of Eq. 2.8 a, b (Levenspiel, 1972;

Senanayake and Das, 2004; Senanayake et al., 2015) in which X = the fraction

dissolved at time t, kpl = apparent rate constant, [H+] = concentration of acid in bulk

solution (mol cm-3), D = diffusion coefficient of H+ through the product layer

(cm2 s-1), r = initial particle radius (cm), ρ = molar density of the dissolving metal in

the initial particle (mol cm-3), b = stoichiometric factor and ε = particle porosity.

Note that kpl is inversely proportional to the square of the initial particle radius.

Although the particles were assumed to be spherical, the final equation is generally

applicable to particles of any isometric shape.

1 - 3(1-X)2/3 + 2(1-X) =6 b [H+]bulk DH+ r-2 ρ-1t = kplt (2.8a)

1 - 3(1-X)2/3 + 2(1-X) =6 b [H+]bulk DH+ r-2 {(1-ε) ρ}-1t = kplt (2.8b)

2.3.3. Kinetics of sulfuric acid leaching

The rate of leaching of ilmenite in H2SO4 was found to be controlled by a

chemical reaction at the ilmenite surface (Barton and McConnel, 1979). Recently, a

process has been described for the reductive leaching of ilmenite minerals in sulfate

media (Zhang and Nicol, 2010).

Barton and McConnel (1979) dissolved ilmenite with 4.7–12.5 M sulfuric

acid solutions at a rather lower temperature compared to temperature used by other

researchers, as listed in Table 2.6. They suggested that the reaction was in chemical

controlled kinetic regime, and the activation energy was estimated to be 90 kJ/mol.

44
Chapter 2: Literature Review

Han et al. (1987) used 4.7–18.8 M sulfuric acid to leach ilmenite from Southern

Thailand at 88–115 °C with an initial acid/ilmenite ratio of 500:3. The reaction

followed chemically controlled kinetics as found by a higher activation energy of

64.4 kJ mol-1. Xu and Huang (1993) tested Panzhihua ilmenite with an acid/ilmenite

ratio of 400:3 at 95 °C –118 °C and determined an activation energy of 44.7 kJ mol-


1
. These reports focused on low temperature, low acid concentration and low pulp

density which are far from the industrial scale operation conditions.

Jablonski and Przepiera (2001) studied the kinetics of reaction of

Norwegian ilmenite with sulfuric acid at a concentration of 14.2 M and an initial

acid/ilmenite ratio of 400:100 in a non-isothermal and non-adiabatic calorimeter.

Their reaction conditions were much more close to those of commercial operations.

As a result of vigorous reaction, the temperature of the reactant mixture rose from

80 °C to about 175 °C. They obtained apparent activation energies of 41.4 kJ mol-1

and 49.5 kJ mol-1, respectively, from the data fitted to two different experiential

kinetic models. The summary of activation energy data derived by the researchers

discussed above is presented in Table 2.6. Kinetic studies by previous researchers

(Imahashi and Takamatsu, 1976; Li et al., 2007; Liang et al., 2005; McConnel 1978;

Sasikumar et al., 2004; Vaughan and Alfantazi, 2006; Zhang and Nicol, 2010) deal

with the effect of acid concentration, temperature, stirring, particle sizes and pulp

density on the rates of leaching of ilmenite. Selected examples are described below.

45
Chapter 2: Literature Review

Table 2.6
Published activation energy data derived from leaching experiments in sulfuric acid.

EA
Acid to (kJ mol-1)
Ore pre- Temp. [H2SO4]
Ore / Origin ore ratio References
treatment (°C) (M)
(w/w) Fe Ti

Barton and
New Zealand None 65-85 4.7 – 12.5 90 McConnel,
1979
Panzhizua, Liang et al.,
None 100-198 15.4 1.5:1 72.6
China 2005
Iluka, Zhang and
None 85-100 4.6 4.5:1 75
Australia Nicol, 2010
a
15 Welham and
Mechanically
Australia 80-120 4.6 50:1 70b Llewellyn,
activated
1998
Manavalkuchi, Mechanically Sasikumar et
70-100 9.2 10:1 68 70
India activated al., 2007
Panzhizua, Mechanically Li et al.,
50-120 4.6 38
China activated 2007
Southern Han et al.,
None 88-115 4.7-18.8 500:3 64.4
Thailand 1987
Jablonski
41.4 and
Norway None 80-175 14.2 400:100
49.5 Przepiera,
2001
a
First stage; bSecond stage

2.3.4. Factors affecting sulfuric acid leaching

i. Acid concentration

The overall reaction for ilmenite leaching in strong sulfuric acid can be

represented by Eq. 2.9. The acid strength significantly influences the dissolution

behaviour of both titanium and iron. When the sulfuric acid concentration is less than

14 M and the temperature is lower than 200 °C, the products TiOSO4 and FeSO4 are

soluble and thus the dissolution can be expressed by Eq. 2.10 (Liang et al., 2005).

46
Chapter 2: Literature Review

FeTiO3 + 2H2SO4→ TiOSO4 + FeSO4 + 2H2O (2.9)

FeTiO3(s) + 4H+(aq) → TiO2+ + Fe2+ (aq) + 2H2O (2.10)

Imahashi and Takamatsu (1976) studied the leaching of ilmenite in sulfuric

acid (0.3 – 1.0 M) and dissolved the iron and titanium to a nearly stoichiometric

ratio. Li et al. (2007) found that the dissolved titanium slowly hydrolysed in 20%

(w/w) sulfuric acid to form a very compact product which covered the surface of

unreacted ilmenite ore hindering further dissolution after 30 minutes. In more dilute

acid, most of the dissolved titanium was hydrolysed within the mass transfer

boundary layer rather than in the bulk of the solution. Li et al. (2007) also found that

it was very difficult to obtain high titanium extraction with a sulfuric acid solution of

concentration less than 50% (w/w). However, Zhang and Nicol (2010) found that, in

the range of acid concentration 450 g/L to 600 g/L, the rate of dissolution only

slightly increased with increasing acid concentration.

ii. Temperature

Imahashi and Takamatsu (1976) investigated the leaching of ilmenite in

sulfuric acid at temperatures in the range of 40 °C to 90 °C. The dissolution rate was

influenced by both the temperature and the concentration of the acid solution. They

concluded that the titanium and iron ions were transported to the solution when

hydrated protons diffused into the solid, causing an increase in the leaching rate of

the ions from minerals in sulfuric acid solutions. Li et al. (2007) also investigated the

effect of leaching temperature ranging from 70 to 100 °C on milled samples.

47
Chapter 2: Literature Review

They found that the dissolution of iron and titanium were fast at the initial stages and

the extraction of iron increased with temperature while that of titanium decreased.

High temperature resulted in instability of the titanium in solution, therefore

enhancing the hydrolysis of titanium ions and lowering the apparent extent of

titanium extraction. The results obtained by Zhang and Nicol (2010) showed a

significant effect of temperature on the rate of ilmenite dissolution in the range of

85 °C to 100 °C. The rates of dissolution of titanium and iron were similar at all

temperatures, indicating the simultaneous dissolution of titanium and iron as

expressed by the overall reaction (Eq. 2.9). It was assumed that there was no solid

product formed during leaching at low concentrations of sulfuric acid.

iii. Mechanical activation

The influence of mechanical activation on subsequent leaching of finely

milled minerals has also attracted attention (Li et al., 2007; Sasikumar et al, 2004,

2007; Welham, 1996; Welham and Llewellyn, 1998; Wu et al., 2011). Mechanical

activation introduces phase transformation, structural defects, strain and changes in

the surface characteristics of minerals thus enhances the leaching rates. In the case of

an activated ilmenite sample at a solid to liquid ratio of 1:10 it was observed that the

hydrolysis of titanium ions during leaching occurred to a more significant extent at

higher temperatures (>70 °C), longer leaching times (> 45 min) and lower acid

concentration (Sasikumar et al., 2007). Welham and Llewellyn (1998) have reported

an eightfold increase in the dissolution rate of ilmenite in sulfuric acid, even at lower

temperature (80 oC), through mechanical activation. They have used a conventional

ball mill and carried out experiments for periods in excess of 100 hours.

48
Chapter 2: Literature Review

Studies by Sasikumar et al. (2004) on the leaching kinetics of mechanically

activated ilmenite with sulfuric acid revealed that the titanium as well as iron

dissolved differently and not according to the stoichiometry. Li et al. (2007) studied

the simultaneous milling and dissolution behavior and found that mechanical

activation of the ilmenite resulted in increased rates of leaching for both iron and

titanium. The enhancement in the leaching kinetics because of mechanical activation

can arise due to increase in surface area or structural ordering or both.

2.3.5. Kinetics of hydrochloric acid leaching

The overall leaching reaction in HCl is generally represented by the reaction

in Eq.2.11. A summary of findings reported by previous researchers is presented in

Table 2.7.

FeTiO3 + 4HCl → FeCl2 + TiOCl2 + 2H2O (2.11)

Tsuchida et al. (1982) studied the kinetics in highly concentrated HCl at the

temperature range of 30 °C- 80 °C (See Table 2.7). They have used the shrinking

sphere model on their data and reported that dissolution rates were well expressed by

the rate equation based on the rate determining step of the surface chemical reaction

(Eq. 2.7). The apparent activation energy was 81.2 kJ mol-1for titanium and 73.2 kJ

mol-1 for iron. van Dyk et al. (2002) also concluded that under conditions stated the

rate of the dissolution of ilmenite could be chemically controlled. However, the

dissolution rate above 60 °C was controlled by the diffusion of ions in the residual

solid layer formed according to Eq. 2.8a-2.8b (Tsuchida et al., 1982).

49
Chapter 2: Literature Review

Table 2.7
Kinetic parameters for ilmenite leaching in concentrated HCl solutions.

[HCl] Grain Temp kc x 106 De x 1010 Ea References


size
(mol dm-3) (oC) (m h-1) (m2 h-1) (kJ mol-1)
(m)
Ti Fe Ti Fe Ti Fe

30 1.23 1.72

40 3.83 4.29 81.2 73.2

50 9.2, 10.4

9.38a Tsuchida et
11 63-45
al., 1982
60 1.84 2.03

70 3.14 3.51
48.9 53.7
80 4.98,
6.09
6.03a

27-32b
Sarker et al.,
1.6-5.2 30-75 40-45c
2006

70,80,
Olanipekun,
7.2 53-74 67.1 62.4
1999
90

92.5 88.3 Jackson and


6 390-254 60-90 Wadsworth,
52.3d 56.1d 1976

58,67,
Orth and
3,4,5,6 -45+38 72,87 66.1 54.3
Liddell, 1986
95

2.3, 6.9,
Sasikumar et
50-100 48-68 45-64
al., 2007
9.2, 11.5

5.37,6.68, 65,75, Wang et


33-38 47.21
al.,2009
8.04, 9.44 85,95

Gireesh et
3.2 – 9.8 40-70 0.13
al.,2015
a
Values obtained using ilmenite ore of 106-63 m grain size.
b
Oxidized for 1h at 950 oC and reduced (in charcoal for 4 h at 1050 oC)
c
Reduced
d
Oxidized and reduced at 900°C

50
Chapter 2: Literature Review

Apparently, the rates after 2-3 hours at high temperature deviated from the

rate equation because of the precipitation of hydrous titanium(IV) oxide on the

surface of the ore by hydrolysis of the dissolved titanium(IV) ions.

Olanipekun (1999) found that two diffusion processes compete in

determining the reaction rate in the leaching of ilmenite in HCl: (i) diffusion of H+

and (ii) diffusion of Ti4+ and Fe2+ produced by the surface chemical reaction of the

unreacted ore. The diffusion rate of H+ was expected to be faster because of the

smaller size of H+ ion compared to Ti4+ and Fe2+.

Wang et al. (2009) proposed that during the early stage, when ilmenite comes

into contact with the hydrochloric acid solution, both iron and titanium dissolve. At

this point the rate is controlled by chemical reaction. After a certain period, the

titanium concentration in solution started to increase to a certain level and then it

precipitated in the pores of the leached solid or as fine particles in solution.

Furthermore, van Dyk et al. (2002) indicated that polymerization occurred when the

Ti(IV) concentration in solution exceeded the value of 10-3 M.

Further support for this view comes from the fact that the ions present in

aqueous solution are not only mononuclear species such as [Ti(OH)]3+, [TiO(OH)]+,

[Ti(OH)3]+ but also polynuclear species such as [(TiO)8(OH)12]4+ (Einaga, 1979).

The diffusion of the polymer chains away from the reaction interface becomes the

rate determining step. Thus, the formation of these polymeric species and the

transition from chemical controlled reaction to a diffusion controlled reaction is

dependent on the concentrations of titanium(IV) and chloride ions and on the solid to

liquid ratio.

51
Chapter 2: Literature Review

The speciation of titanium in chloride media shown in Fig. 2.11 is from van

Dyk et al. (2002). Based on this speciation diagram, van Dyk et al. (2002) suggested

that the rate of reaction may be significantly affected by the formation of solid

TiOCl2 in the pores of the particles. The rate determining step then changes to the

diffusion of the reacting species through the product layer. When a high solid to

liquid ratio is used, the diffusion of hydrogen ions to the reaction interface may be

rate determining at this stage.

Fig. 2.11. Ti(IV) chloride speciation diagram for 0.1 Ti(IV) activity at 298K (van
Dyk et al., 2002).

52
Chapter 2: Literature Review

2.3.6. Factors affecting hydrochloric acid leaching

(i) Particles Size

The rate of iron and titanium extraction increased with the decrease in

particle size (Olanipekun, 1999). The maximum iron dissolution from the particles

of size fraction 20-37 µm was about 67% after 300 minutes. During the same time,

58% of the titanium was dissolved (Table 2.8). Mahmoud et al. (2004) used ground

ilmenite ores of finer size of <75µm (P100) and produced synthetic rutile of <2.5µm

(P99) under optimum conditions. This product is finer than the required specification

for the chlorination process (>50 µm) (Mahmoud et al., 2004). However, few

researchers have reported the effect of particle size on the rate of leaching of ilmenite

in hydrochloric acid (Table 2.8). This may be due to the fact that there was no

observed effect of particle size (Jackson and Wadsworth, 1976; van Dyk, 2002).

Finer particles appear to offer faster leaching kinetics.

Table 2.8
Effect of ore particle size on ilmenite leaching.

Particle T Leaching [HCl] Solid Maximum References


size liquid extraction (%)
(m) (⁰C) Time (h) (M) ratio
(w/w) Fe Ti

20-37 70 6 7.2 0.3 67 58 Olanipekun,1999

37-53 60 (%) 52

53-74 54 46

53
Chapter 2: Literature Review

(ii) Acid concentration

The maximum extraction of iron and titanium from ilmenite obtained by

previous researchers at different acid concentrations is summarized in Table 2.9.

Mahmoud et al. (2004) found that an increase in acid concentration from 15% to

20% increased the iron dissolution from 72% to 99.9% under reducing conditions

discussed in more detail later. This was due to the fact that high acid concentration

prevented the hydrolysis reaction which assisted the dissolved titanium to remain in

solution. However, when the acid concentration was increased above 20%, the TiO2

recovery decreased. At 30% acid the TiO2 recovery was about 27.5% therefore

Mahmoud et al. (2004) chose 20% HCl as a suitable acid concentration. Li et al.

(2008) also suggested on optimum acid concentration of 20% because it is well

known, that the volatility of HCl increases with increasing acid concentration. While

the use of an excessively high acid concentration does not improve the leaching

efficiency, it would also impose a heavy burden upon the HCl regeneration system.

Lasheen (2005) studied hydrochloric acid leaching of Rosetta ilmenite which was

high in iron content. It was reported that a higher acid concentration increased the

iron removal from the lattice structure of the mineral. This was evident from the

tripled iron extraction from 17.5% to 55.4%, with the increase in acid concentration

from 8 to 12 M. Thus, more effective and rapid iron leaching from ilmenite mineral

grains was directly correlated with the molar concentration of acid. About 89% of

rutile was recovered from Rosetta ilmenite at 90 ºC, after 4 hours of agitation at a

solid to liquid ratio of 1:5 (w/v). El-Hazek et al. (2007) indicated that the dissolution

efficiency of TiO2 and total iron is strongly dependent on the acid concentration.

54
Chapter 2: Literature Review

In 12 M HCl about 92% of both titanium and iron were dissolved, but in

7 M HCl, iron dissolution was more than twice that of titanium. The low dissolution

of TiO2 in 7 M HCl is most probably due to hydrolysis and precipitation at this

relatively low acidity. Hydrolysis and precipitation of rutile is expected to be

inhibited at high acidity, as reported by Mahmoud et al. (2004).

Table 2.9
Effect of HCl concentration on ilmenite leaching.

[HCl] T Particle Leaching Acid to Maximum extraction (%) References


(⁰C) size Ore
(M) time ratio Fe Ti
(m)
(h) (w/w)

7.2 52.5 42.5 Olanipekun,


1999
8.4 70 53-74 6 1:300 70.0 60.0

9.6 82.5 75.8

1.7 47.0 Sarker et al.,


2006
3.4 75 1 1:100 48.5

5.2 49.5

3.4 8.8 85.1 Li et al.,


2008
5.2 3.2 91.4
100 4 1:5.5
6.9 1.5 93.3

8.6 1.6 93.1

5.2 17.3 70.1 Mahmoud et


al., 2004a,c
6.2 2.3 95.4

6.9 110 -75 6 1:7.23 0.5 97.5

8.6 0.1 75.8

10.3 0.1 72.4

6.9 10.0;56.0b 5.0; 23.0b Sasikumar et


80 4 1:0.01 al., 2007
9.2 18.0;70.0b 10.0;43.0b
a
Reductive leaching; iron was used as reducing agent
b
Ilmenite ore was mechanically activated
c
% recovery was calculated from residue sample analysis

55
Chapter 2: Literature Review

(iii) Temperature

Table 2.10 summarizes the beneficial effect of high temperatures on iron and

titanium leaching from ilmenite. Sarker et al. (2006) found that at all temperatures

investigated the rate of leaching was initially very fast but decreased after 20 minutes

of leaching. El-Hazek et al. (2007) investigated the effect of change in temperature

from room temperature to 100 °C under optimum leach conditions. The dissolution

of iron and titanium was only 25% and 38%, respectively, at room temperature, but

increased to 78% and 100%, respectively, at 50 °C. Although the titanium

dissolution was 98% at 80 °C, it decreased at 100 oC due to polymerization and

hydrolysis, while the iron dissolution was not affected (Table 2.10).

Therefore, El-Hazek et al. (2007) concluded that temperature plays an

important role in the hydrolysis of dissolved TiOCl2 and 80 °C was considered as the

optimum temperature for the dissolution of ilmenite. Other researchers (Olanipekun,

1999; van Dyk et al., 2002; Mahmoud et al., 2004; Lasheen, 2005; Lakshmanan et

al., 2008; Wang et al., 2009) also agreed that leaching efficiency was increased at

higher temperature but they have obtained different optimum temperatures in the

70 °C to 110 °C range depending upon other conditions.

56
Chapter 2: Literature Review

Table 2.10
Effect of temperature on ilmenite leaching.

T Particle [HCl] Leaching Ore to acid Extraction (%) References


size ratio (w/w)
(⁰C) (m) (M) Time
Fe Ti
(h)

30 45 Sarker et al., 2006

50 3.4 1 1:100 48

75 49

25 38 25 El Hazek et al.,
2007a,c
50 78 99
-75 4 2.5 1:7.23
80 98 100

100 100 30

65 40.0 Wang et al., 2009

75 64.0
33-38 9.44 4 1:50
85 83.0

95 99.0

85 49.2 19.6 Mahmoud et al.,


2004a
95 51.2 22.8
-75 7 6 1:7.23
105 52.7 27.1

110 99.4 97.5

70 50 46 Olanipekun, 1999

80 53-74 7.2 6 1:300 69 60

90 84 75

80 4.5 89.1 Li et al., 2008b,c

90 7 4 1:5.5 2.5 91.9

100 1.5 93.3

50 25.0 33.1 Sasikumar et al.,


2007
60 38.0 37.7

70 9.2 4 1:10 50.0 45.7

80 64.0 41.1

100 86.0 16.0


a
Reductive leaching; iron was used as reducing agent; b Ilmenite ore was mechanically activated
c
% recovery was calculated in residue samples

57
Chapter 2: Literature Review

(iv) Solid to liquid ratio

As shown in Table 2.11, the solid to liquid (S/L) ratio is another important

parameter that influences the maximum extractions of iron and titanium from

ilmenite. The experiments conducted by van Dyk et al., (2002) showed that a low

solid to liquid ratio of 1:4 resulted in only 5% of titanium dissolved at 105 °C.

However, at a solid to liquid ratio of 1:60, 45% titanium was dissolved at the same

temperature. According to van Dyk et al. (2002), the hydrolysis of TiOCl2 to

TiO2.H2O at higher Ti(IV) concentration in solution retarded leaching due to

precipitation of TiO2 in the pores of the leached solid or as fine particles in the leach

solution. This indicates that, at a relatively low initial acid/solid mole ratio and high

temperature, the polymerization reaction proceeds very rapidly in the existing

product layer.

El Hazek et al. (2007) used a lower solid to liquid ratio than the previous

two studies (Lasheen, 2005; Mahmoud et al., 2004) with the intention of dissolving

both iron and titanium and keeping the titanium dissolved in solution. Using 12 M

HCl at 80 °C, a solid to liquid ratio of 1:10 and 0.1 kg metallic iron / kg of ilmenite,

the titanium dissolution reached 90% after 1.5 hours. The increase in residence time

decreased the titanium concentration in solution due to hydrolysis. It was found that

only 42% of titanium was dissolved at a solid to liquid ratio of 1:7. Li et al. (2008)

also found that the quality of the synthetic rutile produced in the experiment

improved with decreasing ilmenite/acid ratio. A high ilmenite/acid ratio made the

solid liquid separation difficult due to extremely fine particle size of the precipitate.

58
Chapter 2: Literature Review

Table 2.11
Effect of solid liquid ratio (S/L) on ilmenite leaching.

Particle Maximum
Solid liquid T Leaching [HCl]
size extraction (%) References
ratio (w/w) (⁰C) Time (h) (M)
(m) Fe Ti

1:4 2.5 van Dyk et al.,


-106 105 8 6
1:60 50.0 2002c

1:6.1 97.3 86.6 Mahmoud et


al., 2004a,c
1:7.2 99.4 88.0

1:8.8 -75 110 6 6.9 99.5 89.8

1:10.8 99.6 90.1

1:11.5 99.7 90.4

1:5 75 10 El-Hazek et al.,


2007a,c
1:10 80 40
80 5 12
1:20 90 90

1:30 90 90

1:11 1.0 93.9 Li et al., 2008b

1:7.3 1.2 93.6


100 6.9
1:5.5 1.5 93.3

1:4.4 2.9 91.6


a
Reductive leaching with iron used as the reducing agent
b
Ilmenite ore was mechanically activated
c
% recovery was calculated from residue sample analysis

59
Chapter 2: Literature Review

2.3.7. Mixed brine leaching

Innovative chloride leaching followed by solvent extraction and precipitation

to produce high purity titanium dioxide has been studied by Lakshmanan et al.

(2008) and the flowsheet is shown in Fig. 2.12. Laboratory scale leaching

experiments were carried out using mixed hydrochloric acid and magnesium chloride

system as a lixiviant under atmospheric pressure at 70-73ºC for 4 hours. A high

titanium extraction (96.9%) was obtained for a solid loading of 8.8%, HCl

concentration 4 times the stoichiometric amount and MgCl2 concentration of 300

g/L. In another study published by Das et al. 2013, ilmenite leaching was also

conducted using HCl and CaCl2 and the extraction of titanium and iron was 96% and

97%, respectively.

Fig. 2.12. Flowsheet developed for production of TiO2 (Lakshmanan et al., 2008).

60
Chapter 2: Literature Review

A series of comparative experiments also were conducted, to show the

effect of hydrochloric acid in the absence of MgCl2. The result showed that low

titanium extraction of less than 20% was obtained even though the concentration of

hydrochloric acid was as high as 8 M and 3.83 times the stoichiometric amount for a

solid loading of 5% (w/w) (Lakshamanan et al., 2008). It seems that a mixed acid-

chloride leaching process has the potential to improve the production of titanium

dioxide in terms of recovery. It also allows recycling of hydrochloric acid,

magnesium chloride and magnesium oxide in the process. It was suggested that the

use of magnesium chloride allowed a reduction of the hydrochloric acid

concentration to no more than 20% (w/w). This also permitted use of azeotropic

distillation by pyrohydrolysis of recycled solutions without the addition of

substantial amounts of more concentrated hydrochloric acid, therefore avoiding

disposal of excess hydrochloric acid, which is an environmental problem. The use of

a lower concentration of hydrochloric acid is also expected to result in lower

extraction of impurities or gangue from the ore, and thus decrease the burden

downstream for the removal of impurities.

The flowsheet for the production of TiO2 described in Fig. 2.12 indicates

that the solvent extraction process consisted of two different stages of extraction in a

counter-current mode. Iron was removed in the first stage and titanium was

extracted in the second stage. The strip solution obtained from the second stage

contained the purified titanium dioxide and from this stream pigment grade rutile

(TiO2) was recovered. The barren solution was used for reagent recovery and the

regenerated leach liquor was recycled. It was noted by Lakshmanan et al. (2008)

that nano-sized TiO2 was produced by this process by controlling thermal

precipitation.

61
Chapter 2: Literature Review

2.3.8. Reductive leaching of ilmenite in HCl

(i) Advantages and flowsheet

The reductive leaching of ilmenite in hydrochloric acid by adding metallic

iron has been studied by several authors (Ibrahim et al., 2003; Mahmoud et al., 2004;

Lasheen, 2005; El-Hazek et al., 2007). According to Mahmoud et al. (2004), the iron

metal can be added as a pure powder or as scrap. This method has an advantage

compared to industrial processes in view of the avoidance of high temperature pre-

treatment or complicated pressure leaching. The acid consumption could be higher in

the proposed process, but recent advances in acid regeneration technology from the

spent leach liquor have made leaching processes more attractive and by-product iron

oxide may have commercial uses and add value to the process. Iron metal is

considered the most suitable reductant for this process since no foreign ions will be

introduced to the reaction medium and the ferrous chloride formed can be separated

from the solution and used for the regeneration of hydrochloric acid.

A promising route for synthetic rutile production can be outlined based on

the improved reactivity of ilmenite in hydrochloric acid after addition of iron metal

without pre-treatment. Fig. 2.13 shows a proposed flowsheet for the production of

synthetic rutile from ilmenite by the reductive leaching process in combination with

acid regeneration from spent leach liquor. It was designed to be a batch process

where leaching and hydrolysis take place simultaneously in one stage. The TiO2

concentrate can be separated by decantation with the help of a flocculating agent.

62
Chapter 2: Literature Review

Fig. 2.13. A proposed flowsheet of synthetic rutile production by reductive leaching


process (Mahmoud et al., 2004).

Synthetic rutile is produced by calcination of the TiO2 concentrate at about

900 ºC (Fig. 2.13). Hydrochloric acid can be regenerated from the spent liquor by

evaporation to separate FeCl2 from which the HCl can be produced by pyrolysis.

Ferric oxide is formed in a pellet which can be used as feed to steel making, cement

production or other uses depending on the location.

(ii) Comparison with non-reductive leaching

Mahmoud et al. (2004) have suggested that the different stages of the

proposed process should be investigated in detail. Ibrahim et al. (2003) and later

Mahmoud et al. (2004) conducted the most extensive testing of ilmenite leaching in

hydrochloric acid using metallic iron.

63
Chapter 2: Literature Review

They compared leaching under non-reducing and reducing conditions and found a

significant increase in the reaction rate when metallic iron was added.

Under non-reducing conditions, Ibrahim et al. (2003) and Mahmoud et al.

(2004) found that less than 50% of the iron and less than 10% of the titanium

dissolved after leaching for 5 to 6 hours in 20% HCl at S/L ratio of 1/7.3 at 110 ºC. It

was assumed that ilmenite was partially leached due to limited dissolution of

hematite in HCl. This made the leaching process difficult since the hematite was

finely disseminated in the ilmenite, as found by Mahmoud et al. (2004) by

microscopic investigation. Ibrahim et al. (2003) also found that at a high S/L ratio of

1:55, the extent of leaching continuously increased with time to values of 80% and

86% for TiO2 and total iron, respectively (Ibrahim et al., 2003). From these results,

Ibrahim et al. (2003) noted that significant ilmenite dissolution required an excessive

amount of HCl solution which is impractical in hydrometallurgical processes due to

economic reasons.

The hematite dissolution in HCl was found to improve under reducing

conditions (Lu and Muir, 1988). Mahmoud et al. (2004) added iron metal at a

Fe/ilmenite mass ratio of 0.075/1.000 or 1.000/1.000 into the leaching solution

resulting in more than 95% of iron dissolution in 4 hours with only 2% titanium

dissolution.

The effect of S/L ratio, acid concentration, temperature, iron powder addition

time, iron powder stoichiometry and retention time on the extent of reaction were

measured under reducing conditions. Baseline conditions of 20% acid, solid to liquid

ratio of 1:7.2, iron powder addition time after 90 min, iron powder stoichiometry of

1.5 and a leach time of 6 hours were used.

64
Chapter 2: Literature Review

(iii) Chemical reaction

Ibrahim et al. (2003) suggested that initially both ilmenite and hematite

dissolved in hydrochloric acid according to Eqs. 2.12-2.13.

Ilmenite reaction with HCl:

FeO.TiO2 + 4HCl → FeCl2 + TiOCl2 + 2H2O (2.12)

Hematite reaction with HCl:

Fe2O3 + 6HCl → 2FeCl3 + 3H2O (2.13)

Titanium species hydrolyse to form TiO2 precipitate:

TiOCl2 + H2O → TiO2 + 2HCl (2.14)

The titanium species then hydrolysed to produce a precipitate according to

Eq.2.14. A summary of the reaction equations for reductive leaching of ilmenite in

HCl with iron powder was provided by Mahmoud et al. (2004). The added iron

powder dissolved readily in HCl creating nascent hydrogen and part of the nascent

hydrogen forms hydrogen gas (Eqs. 2.15- 2.16). Then, the dissolved Fe(III) is

reduced by the nascent hydrogen and/or the Fe powder (Eqs. 2.17-2.18). Once all of

the dissolved Fe(III) has been reduced, any remaining nascent hydrogen and/or iron

powder reduces the dissolved titanium Ti(IV) species to the Ti(III) state (Eqs. 2.19-

2.20). As the Fe powder and nascent hydrogen will be rapidly removed from the

system, as soon as the iron powder is completely dissolved, the formed Ti(III) will

maintain the reducing conditions in the solution. That is, the formed Ti(III) will

reduce any further dissolved Fe(III) to the Fe(II) state (Eq. 2.21).

65
Chapter 2: Literature Review

This sequence is consistent with the Ti(III) contents in solution as shown in

Fig. 2.14 (Mahmoud et al., 2004). It shows the consumption of the Ti(III) from the

solution over time and indicates its continuous oxidation to the Ti(IV) state.

Hematite was dissolved in acid while being simultaneously reduced to Fe(II) by Eqs.

2.22-2.23.

Reaction of added iron powder with acid created nascent hydrogen:

Fe + 2HCl → FeCl2 + 2H● (2.15)

H● + H●→ H2(g) (2.16)

Dissolved Fe3+ reduced by nascent hydrogen or iron powder:

2FeCl3 + 2H● → 2FeCl2 + 2HCl (2.17)

2FeCl3 + Fe → 3FeCl2 (2.18)

Ti(IV) reduced by nascent hydrogen or iron powder:

2TiOCl2 + 2HCl + 2H● → 2TiCl3 + 2H2O (2.19)

2TiOCl2 + 4HCl + Fe → 2TiCl3 + FeCl2 + 2H2O (2.20)

Ti(III) reduced any further dissolved Fe(III):

TiCl3 + FeCl3 + H2O → TiOCl2 + FeCl2 + 2HCl (2.21)

The dissolved Fe(III) reduced to Fe(II) by reaction with Fe powder:

Fe2O3 + 6HCl → 2FeCl3 + 3H2O (2.22)

2FeCl3 + Fe → 3FeCl2 (2.23)

66
Chapter 2: Literature Review

(A) (B)

Fig. 2.14. Leaching of ilmenite ore with HCl only (A) and in presence of iron
powder.(B) [HCl] = 20%, ilmenite /acid ratio = 1/7.3, temperature = 110 oC, Fe
powder = 0.075 g/g ore, time of addition = 20 min (Mahmoud et al., 2004).

The reductive leaching of ferric oxide(s) increases the reactivity of the

ilmenite ore, due to the breaking up of the grain structure causing further diffusion of

the acid protons through the pores created in the ore particles. It was found that after

iron addition, the surface area of the leach residues increased due to partial

dissolution of hematite and ilmenite (Mahmoud et al., 2004). Continued rapid

leaching of iron beyond 60 minutes seems dependent on the hydrolysis reaction

where active HCl is released (Eq. 2.14). This was confirmed by the fact that the rapid

dissolution after 60 minutes coincided with the appearance of a white hydrated TiO2

precipitate in the reaction medium (Mahmoud et al., 2004).

67
Chapter 2: Literature Review

(iv) Effect of quantity and time of iron addition

Time of iron powder addition was varied from 0 (the start of the experiment)

to 90 minutes into the experiment conducted by Mahmoud et al. (2004). It was

found that adding the iron powder after 15 minutes of leaching improved the quality

of the rutile compared to addition at the start of the experiment.

The effect of quantity of iron powder on leaching efficiency was also

investigated by Mahmoud et al. (2004). The iron powder added to the leaching

system was according to the stoichiometric amount of iron powder from 0.0 to 1.5.

The stoichiometric amount of iron powder was corresponding to the content of Fe2O3

in the ilmenite ore that dissolved and reacted with Fe according to the reaction:

2FeCl3 + Fe → 3FeCl2 (Eq. 2.23). Without iron powder, the amount of rutile

produced was very low (26% TiO2). After 1.1 of stoichiometric amount of iron

powder was added, the rutile product quality increased to 97.9% TiO2.

From all the test work completed, Mahmoud et al. (2004) selected the

following conditions as optimum for the leaching of ilmenite: HCl concentration of

20%, temperature of 110ºC, iron powder stoichiometry of 1.1 added after 30 minutes

and retention time of 5 hours. These selected conditions agree well with reported

results by Ibrahim et al. (2003). It was also noted that the size of 99% of the rutile

particles produced was less than 2.5 µm. This is much finer than the size required by

the chlorination process (>50µm). Adding iron powder between 15 to 90 minutes

after the start had little effect on the quality of the rutile compared to addition at the

start of the experiment.

68
Chapter 2: Literature Review

It was reported that the increase in leaching time increased the rutile recovery

and iron removal efficiency (Mahmoud et al., 2004). Leaching experiments were

carried out using 20% HCl and 110 ºC for different periods (1.5 – 6.0 hours), with

iron powder added at a stoichiometry of 1.1 after 30 minutes. It was found that

during first 3 hours, the rutile recovery and iron removal reached 95.9% and 98.3%,

respectively. The rutile grade and recovery increased slowly as the leaching time was

increased from 3 to 6 hours. An optimum leaching time of 5 hours was

recommended corresponding to 97.7% and 99.4% rutile recovery and iron removal,

respectively.

In another study, Lasheen (2005) tested the effect of iron powder

stoichiometry on the rate of leaching of iron from Rosetta beach ilmenite

concentrates. The rate of leaching increased as iron powder dosage was increased

from 0 to 6%. The test work achieved 85% iron dissolution (0% aqueous titanium)

after leaching with 6% iron powder at 90 ºC using 12 M acid and a solid to liquid

ratio of 1:3 over a period of 8 hours.

(v) Effect of leaching time

El Hazek et al. (2007) had conducted an ilmenite dissolution study over a

shorter leaching period of 1 to 3 hours using 200 mesh size ore under optimum leach

conditions (12 M HCl, 80 ºC, S/L ratio 1/10 and 0.1 kg metallic iron/kg ilmenite).

The dissolution efficiency of titanium increased from 93% to 98% with the increase

in leach time from 1 to 2.5 hours. The extension of the leaching time from 2.5 to 5

hours, however, caused a slight reduction of the dissolution of titanium and total iron

due to hydrolysis.

69
Chapter 2: Literature Review

El-Hazek et al. (2007) measured the effect of iron powder dosage on the extent of

titanium dissolution. It was found that maximum dissolution was achieved when 0.1

kg metallic iron per kg of ilmenite was used. Increasing the amount of iron powder

to 0.8 kg/kg ilmenite decreased the dissolution efficiency. It was proposed that

increased acid consumption from the iron powder was the cause of this reduction.

2.4. Rotating disc study of ilmenite

2.4.1. Background

By studying dissolution from rotating disc, the reaction can be investigated

under closely controlled and well defined hydrodynamic conditions while

maintaining a relatively constant surface area. A few studies on different minerals

such as hematite, magnetite, nickel oxides, iron metal have been conducted using a

rotating disc method to study the kinetics of dissolution as shown in Fig. 2.15 (Lu

and Muir, 1988). Barton and McConnel (1979) also carried out experiments on the

dissolution of ilmenite in sulfuric acid by using the rotating disc method. The

advantage of using a rotating disc made of the reacting metal over the use of a

stationary surface in a stirred vessel is that, if the system is properly designed, the

rate at which reactants reach the surface can be accurately controlled and calculated

(Burkin, 2001). The rates of the different steps (adsorption, reaction, desorption) are

controlled by temperature, surface area and reactant concentration but are

independent of the solution hydrodynamics. Therefore, the results based on rotating

disc can be used to determine the chemical or diffusion controlled nature of the rate

controlling step.

70
Chapter 2: Literature Review

However, relatively few of the dissolution studies reported have employed

the rotating disc technique. Hence, ilmenite leaching in hydrochloric acid using

rotating synthetic ilmenite disc will be performed in this study. The details of the

rotating disc method and the approach will be discussed later in this thesis, but a

brief introduction follows.

Fig. 2.15. A comparison of the dissolution of the rotating discs of Fe3O4, Fe2O3 and
copper, nickel, and zinc ferrites in 1 M HCl (based on iron dissolved) at 25 oC,
400 rpm (Barton and McConnel, 1979).

2.4.2. Effect of disc rotation speed

The rotating disc system is a hydrodynamic system for which the equation

of convective diffusion can be solved. It consists of a disc, rotating about a central

axis perpendicular to its surface plane and of sufficient radius to make edge effects

insignificant (Barton and McConnel, 1979). Another requirement is that the fluid

flow adjacent to the disc is non-turbulent, which is the case if (i) the volume of

solution is large enough to avoid wall effects, and (ii) the disc is smooth and if the

71
Chapter 2: Literature Review

maximum rotational velocity is restricted so as not to exceed the critical value of the

Reynolds number (Barton and McConnel, 1979).

Under these conditions the Nernst diffusion layer thickness is given to a good

approximation by Eq. 2.24a (Burkin, 2001):

δ = 1.612 D1/3v1/6ω-1/2 (2.24a)

where v is the kinematic viscosity of the fluid, and ω is the angular velocity of the

disc. Rate of dissolution of metal per unit surface area is given by the Levich

equation Eq. 2.24b:

RM = 0.623 D2/3-1/6ω1/2 C = kC (2.24b)

where D is the diffusion constant, C is the concentration and k is the rate constant.

In the studies reported by Barton and McConnel (1979), it was found that

although straight lines were obtained in plots of dissolution rates against ω1/2 for

several systems, the lines did not pass through the origin as required by theory. This

was caused by non-uniformity of the disc due to the polycrystalline nature of the

samples. Similar results also were reported by Ward (1990) for the dissolution of a

pure iron disc and reduced ilmenite (See Fig. 2.16). As an additional check on the

nature of the dissolution process, the theoretical rate of dissolution for transport

control can be calculated using a known or estimated value of D, v, ω and c0 together

with the geometrical surface area of the disc. Such calculations are not precise, but if

the experimental dissolution rate is significantly lower than the theoretical rate it is

clear that the process is not limited by mass transport in the liquid. If the reaction is

under chemical control or limited by diffusion within the solid, then the rate will be

72
Chapter 2: Literature Review

independent of the hydrodynamic conditions, so that variation of the disc rotation

speed will not affect the observed rate.

(A)

(B)

Fig. 2.16. (A) Dissolution rate, dc/dt, at 25 oC of calcium ion from superphosphate as
a function of square root of the rotation velocity. (B) (▲) Reduced ilmenite disc;
25 oC, 0.1 M FeCl3, ( ● ) Pure iron disc; 25 oC, 0.02 M Fe2+/0.1 M HCl (Barton and
McConnel, 1979 (A); Ward, 1990 (B)).

73
Chapter 2: Literature Review

2.4.3. Effect of temperature

Evidence on the nature of dissolution reactions can be obtained also by

measuring the temperature dependence of the reaction rate. The apparent activation

energy of a reaction, E can be obtained by applying the Arrhenius equation given by

Eq. 2.25:

k = A exp (-E/RT) (2.25)

This equation describes the variation of the experimental value of the dissolution rate

constant k with absolute temperature T (where A is a pre-exponential constant and R

is the gas constant). For mass transport or diffusion control, the activation energies

normally ranging from 8 to 25 kJ mol-1, whereas for a chemically-controlled process

the value of activation energy is considerably higher than activation energy for mass

transport/diffusion. In some systems dissolution is controlled by diffusion of the

dissolved material through a porous insoluble layer which forms on the solid surface.

In such systems the rate becomes independent of stirring speed, in the case of a

suspended solid powder, or of rotation speed, in the case of a rotating disc, because it

is controlled by transport within the solid.

The value of the activation energy in such cases can assist in distinguishing

this situation from one of chemical control (Barton and McConnel, 1979). If stirring

is rapid and efficient (no transport control) the overall process will appear to be

chemically controlled (Burkin, 2001).

On the other hand, in stationary condition the overall reaction rate may well

be determined by transport control; and of course there will be an intermediate,

transition situation between the two regimes.

74
Chapter 2: Literature Review

In the case of a rotating disc, as the rotation speed is increased there may be

a transition from dependence on ω1/2 (liquid transport control) to independence of ω.

There are limitations to the rotational velocities of the disc, the upper limit being

determined by the critical value for transition to a turbulent flow, and the lower limit

being the requirement that contribution to mass transfer by natural convection is

negligible. Another complication is that many slow reactions require the use of large

surface areas in order to produce measurable dissolution in a reasonable time, and

here the rotating disc technique is of little use; it is necessary to grind the sample to a

powder to increase the surface area, at the expense of losing control of the

hydrodynamic conditions (Barton and McConnel, 1979).

2.4.4. Investigations relevant to this study

McConnel (1978) has outlined the details of disc preparation in his thesis.

The dissolution of ilmenite in sulfuric acid has been studied by McConnel (1978)

using particulate dissolution and the rotating disc method. Until now, there are no

reports on the dissolution of ilmenite in hydrochloric acid using a rotating disc. For

this research, the preparation of rotating disc will follow the method of preparation

reported by McConnel (1978). Barton and McConnel (1979) reported a few

experiments in which they studied the behaviour of solids which demonstrated

different types of rate control.

The dissolution of gypsum in water was identified as a transport controlled

process and the dissolution of anhydrous calcium sulfate indicated that the process

was intermediate between chemical and transport-controlled (Barton and McConnel,

1979).

75
Chapter 2: Literature Review

The dissolution of ilmenite in sulfuric acid was found to be a chemically

controlled process (McConnel, 1978). The effect of temperature (65, 75 and 85 °C)

was studied at different rotation speeds (12.4, 26.2, 36.7 and 71.2 rad s-1) in sulfuric

acid solutions of different concentrations (4.7 to 12.5 M). The dissolution rate was

reported to be independent of the rotation rate, and the calculated apparent activation

energy was 101±8 kJ mol-1. Both results proved that the rate of dissolution under

these conditions was chemically controlled (Barton and McConnel, 1978; Barton and

McConnel, 1979).

2.5. Electrochemical study of ilmenite

The study of the kinetics of electrochemical reactions involves

measurement of the rate or current as a function of the potential. A three electrode

system generally utilized for electrochemical measurements consists of a working

electrode, a reference electrode and a counter electrode. The platinum wire counter

electrode of 0.5 mm diameter is generally used to make an electrical connection to

the electrolyte so that a potential could be applied to the working electrode. The

reference electrode has a stable and well-known electrode potential which is used as

a reference which enables the working electrode potential to be established.

76
Chapter 2: Literature Review

Electrochemical studies of ilmenite can shed more light on the mechanism

of dissolution reactions involved. This understanding may assist the control and

optimization of the leaching process in terms of leaching rates of iron and titanium

and the purity of the titanium dioxide. The dissolution kinetics and electrochemistry

of oxide minerals has been the subject of detailed investigations for many

hydrometallurgical processes. There is not much information available in the

literature related to electrochemical studies of ilmenite, especially in acid media.

This is due to the difficulty of obtaining pure samples of ilmenite. However, a few

published studies related to ilmenite mineral are available. A number of studies have

been conducted on iron oxide minerals such as magnetite (Allen et al., 1979; Lu and

Muir, 1985; White at al., 1994), metal ferrite (Lu and Muir, 1988), reduced ilmenite

and natural ilmenite (McConnel, 1978; Ward, 1990; Zhang and Nicol, 2009).

Table 2.12 summarizes experimental conditions used in various published

electrochemical studies. McConnel (1978) carried out a preliminary study of the

electrochemical behavior of synthetic and natural ilmenite in various concentrations

of H2SO4 at temperatures of 20 °C and 40 °C using cyclic voltammetric and

potentiostatic techniques.

77
Chapter 2: Literature Review

Table 2.12
Study on oxide minerals by electrochemical method.

Electrolytes
Oxides Types of T Electrochemical
and References
source electrode (oC) techniques
concentrations
Compacted NaOH (1 M); Open circuit potential
Allen et al.,
Magnetite magnetite/ Percholric acid 22 (OCP), Cyclic
1979
graphite (pH 2-13) voltammetry
Iron(III)
oxide NaCl (3 M) Lu and Muir,
22 Cyclic voltammetry
Nickel HCl (1 M) 1988
Oxide
CuSO4
Carbon Acetonitrile
Lu and Muir,
Magnetite paste (4 M) 25 Cyclic voltammetry
1985
electrode HCl (1 M)
NaCl (3 M)
Magnetite OCP, White et al.,
NaCl (0.1 M)
Ilmenite Potentiodynamic 1994
Current potential,
Jayasekera et
Iron disc NH4Cl 30 Corrosion
al., 1995
measurement
Carbon
Reduced NH4Cl Marinovich et
paste 30 Cyclic voltammetry
ilmenite FeCl2 al., 1995
electrode
Carbon
Adriamanana
Ilmenite paste H2SO4 (1 M) 25 Cyclic voltammetry
et al., 1984
electrode
Carbon
Reduced NH4Cl Kumari et al.,
paste 30 Cyclic voltammetry
Ilmenite (1.5% at pH 4) 2002
electrode
OCP, Cyclic
Carbon
H2SO4 voltammetry and Zhang and
Ilmenite paste 60
(450 g/L) Coulometric Nicol, 2009
electrode
measurements

Adriamanana et al. (1984) studied the electrochemical behavior of several

natural ilmenite samples in 1 M sulfuric acid by cyclic voltammetry with a carbon

paste electrode. White et al. (1994) investigated the electrochemistry of magnetite

and ilmenite in anoxic solutions (pH 1-7) at temperatures of 2-65 °C. Recently,

Zhang and Nicol (2009) have investigated the dissolution of massive ilmenite and

powder ilmenite with carbon paste electrode in sulfuric acid at elevated

temperatures. They used different techniques to study the chemical reactions

involved in the leaching processes.

78
Chapter 2: Literature Review

2.5.1. Open circuit potential (OCP)

(i) Background

The electrochemical potential is one of the basic measurements in

electrochemistry. The equilibrium potential assumed by the metal in the absence of a

voltage applied to the test electrode is called the Open Circuit Potential (OCP)

(Holze, 2009).

The OCP is a parameter which indicates the thermodynamic tendency of the

material in the test electrode to electrochemical oxidation in a corrosive medium.

After a period of immersion it usually stabilizes around a stationary value. The goal

of OCP measurements is to measure the potential of the electrode without affecting,

in anyway, the electrochemical reactions taking place on the electrode surface. It is

necessary to make the potential measurements with respect to a stable reference

electrode so that any changes in the measured potential can be attributed to changes

at the electrode/solution interface (Holze, 2009).

The relationship between electrode potential, always considered to be at

equilibrium in this chapter, and the activities of the species involved in the electrode

reaction is given by the Nernst equation (Eq. 2.26) for the general reaction

ox + ne = red :

E= E’ + RT/nF ln (aox/ared) (2.26)

in which E (or Eh) is electrode potential; E’ is standard potential; n is number of

electrons; F is Faraday constant; R is universal gas constant; aox and ared represent

chemical activity of oxidants and reductant, respectively. The Nernst equation also

only applies when there is no net current flow through the electrode.

79
Chapter 2: Literature Review

The activity of ions at the electrode surface changes when there is current

flow, and there are additional terms for overpotential and resistive loss which

contribute to the measured potential. The electrochemical potential E (mV) at an

oxide surface in the presence of an electrolyte represents the three factors listed

below (Bard and Faulkner, 2001):

(1) solid state electrochemical reactions and electron transfer,

(2) interfacial electron transfer mechanisms dependant on the nature and

coordination of the surface ligands, and

(3) diffusion controlled transport through the double layer (interfacial) to the bulk

solution.

(ii) OCP of magnetite

White et al. (1994) measured the open circuit potential of magnetite and

ilmenite electrodes in 0.1 M NaCl at different pH values (3-7) to investigate the

reactions taking place at the surface of the electrodes. Apparently, the open circuit

potential (Eo) increased with time and decreased with solution pH, as shown in Fig.

2.17. The steady state potentials were reached 1 to 2 hours after immersion in the

solution. The potentials also increased with increasing dissolved oxygen (DO)

concentration in oxic solutions, indicating that they were influenced by an increase

in rates of surface oxidation and/or mixed aqueous reactions involving oxygen.

Steady state potentials values of magnetite and ilmenite, tabulated in Table 2.13

show slightly lower rest potential values for ilmenite than magnetite (White et al.,

1994).

80
Chapter 2: Literature Review

Fig. 2.17. Open circuit potential, Eo as functions of time for magnetite (ISH) and
ilmenite electrodes reacted in anoxic 0.1 M NaCl solutions at different pH (White et
al., 1994).

Table 2.13
Steady state potential values of magnetite and ilmenite.

pH ISH Magnetite, Eo (mV) Ilmenite, Eo (mV)


3 380 310
5 310 290
7 270 260
Obtained by White et al., (1994) in 0.1 M NaCl at different pH

(iii) OCP of massive ilmenite

Fig. 2.18 shows the open circuit potentials of the massive ilmenite and a

platinum electrode measured in 450 g L-1 sulfuric acid solution with or without Fe2+

and Fe3+ ions at 60 oC by Zhang and Nicol (2009). The potentials of the ilmenite and

platinum electrodes in equimolar solutions of Fe(III) and Fe(II) are comparable

(0.72 V in curves (d) and (e)).

81
Chapter 2: Literature Review

The Eh values representing the potential with respect to the standard

hydrogen electrode based on the HSC 7.1 data base (Roine, 2011) at 60 oC for the

redox couples Fe2O3/Fe2+ and Fe3+/Fe2+ in the absence of sulfate ions are 0.81 V and

0.74 V, respectively. Therefore, as noted by Zhang and Nicol (2009) the measured

value of 0.72 V in 450 g L-1 H2SO4 is close to the Fe3+/Fe2+ couple (0.74 V). The

initial increase in OCP in Fig. 2.18 is indicative of the increase in Fe(III) in solution,

due to the acid dissolution of Fe2O3 from ilmenite reaching an equilibrium in the

solution. This is further supported by the fact that the potential of the ilmenite

electrode in sulfuric acid solution alone is lower than that in the equimolar

Fe(III)/Fe(II) solution for the platinum electrode. Again, as interpreted by Zhang and

Nicol (2009) the increase in OCP in H2SO4 (curve (c)) showed a more rapid or

selective dissolution of Fe(III) than Fe(II) from the mineral.

The potential of the ilmenite electrode in acid solution containing only

Fe2+(curve (b)) is also higher than that of the platinum electrode in the same solution

(curve (a)), again suggesting a more rapid dissolution of Fe(III) from the mineral.

Thus, the potential of the ilmenite electrode is determined by that of the

Fe(III)/Fe(II) couple at the mineral surface (Zhang and Nicol, 2009).

82
Chapter 2: Literature Review

(e)

(d)
(c)

(b)

(a)

Fig. 2.18. Open circuit potentials of rotating ilmenite and platinum electrodes in
450 g L-1 H2SO4 solutions containing various concentration of iron at 60 °C. (a)
platinum [20 g L-1 Fe2+], (b) ilmenite [ 20 g L-1 Fe2+], (c) ilmenite [H2SO4 only], (d)
platinum [10 g L-1 Fe2+ and 10 g L-1 Fe3+], (e) ilmenite [10 g L-1 Fe2+ and 10 g L-1
Fe3+] (Zhang and Nicol, 2009).

2.5.2. Mixed potential theory

The mixed potential model was first introduced by Wagner and Traud (1938)

and extended by Nicol (1975, 1993) to explain the kinetics observed in many

hydrometallurgical processes involving solids which are electron conductors, such as

metals, metal sulfides and some oxides.

The mixed potential of an electrode is generally due to a reaction taking

place on the electrode. A particular case has to be distinguished where the mixed

potential measured on a mineral surface is due to a dissolution process (Lazaro-

Baez, 2001). Mixed potentials can be categorized into two types: mixed potential of

a solution and mixed potential of a mineral in solution.

83
Chapter 2: Literature Review

(i) Mixed potential in solution

The mixed potential of a solution arises when there are two or more redox

couples (Red1/Ox1 and Red2/Ox2) present in the solution and they are not in

equilibrium, i.e. the reversible potentials. The reversible potentials, as given by the

Nernst equation, are not equal (Eqs. 2.27-2.29).

ERRed1/Ox1 = E0Red1/Ox1 + (RT/n1F) ln ([Ox1]/[Red1]) (2.27)

ERRed2/Ox2 = E0Red2/Ox2 + (RT/n2F) ln ([Ox2]/[Red2]) (2.28)

ERRed1/Ox1 ≠ERRed2/Ox2 (2.29)

Thus, in this case the redox potential of the solution is not characterized by a unique

Eh value, but rather it is influenced by two potentials, i.e., those of the two redox

couples present (see Fig. 2.19). When an inert electrode such as platinum is

introduced into the solution, its electrons can exchange with two separate redox

couples and the measured potential will reach a value, lying between the equilibrium

potentials of the two couples at which the component anodic and cathodic processes

proceed at equal and opposite rates, as depicted in Fig.2.19.

84
Chapter 2: Literature Review

Fig. 2.19. Schematic representation of a mixed potential in solution (Rand and


Woods, 1984).

(ii) Mixed potential of a mineral in solution

The mixed potential of a mineral in contact with a solution, in this case the

electrode i.e. the mineral surface, forms an integral part of the components making

up the mixed potential. Hence, the mixed potential, as well as the current-potential

curves, in this case depend greatly upon the properties of the mineral surface. Figs.

2.20 and 2.21 show schematic diagrams that describe the mixed potential of oxide

minerals under reducing conditions.

In Fig. 2.21, the mixed potential Em represents the potential of the magnetite

electrode measured during reductive leaching of Fe3O4 to Fe2+according to the

reactions shown below:

Anodic half reaction:

Cu(I) → Cu(II) + e- (2.30)

85
Chapter 2: Literature Review

Cathodic half reaction:

Fe(III) + e-→ Fe(II) (2.31)

ia

Fe(II) = Fe(III) + e-
Current

Ee1 Em Ee2

Oxide + ne-= Reduced product

ic

Fig. 2.20. Mixed potential diagram.

Fig. 2.21. Schematic Evan diagram for the leaching of Fe3O4 in acidified Cu(II)-
Cu(I) solutions (Lu and Muir, 1985).

86
Chapter 2: Literature Review

2.5.3. Voltammetry studies

Voltammetry is the most widely used technique for acquiring qualitative

information about electrochemical reactions. It offers a rapid location of redox

potentials of the electroactive species (Bard and Faulkner, 2001). The technique

involves recording the current response of the system in which the applied potential

changes linearly with time (in case of linear sweep voltammetry). After reaching a

desired limit in current or potential, the scan is reversed (in the case of cyclic

voltammetry) and this can be repeated several times at various scan rates. A few

studies have been made to characterize the reactions taking place on ilmenite in

sulfuric acid solution using cyclic voltammetry techniques (Adriamanana et al.,

1984; McConnel, 1978; Zhang and Nicol, 2009).

(i) Cyclic voltammetry of ilmenite in sulfuric acid solution

The study of the behaviour of ilmenite in sulfuric acid solution has been

reported in literature using cyclic voltammetry techniques. Adriamana et al. (1984)

applied the voltammetric techniques to study the behaviour of ilmenite in sulfuric

acid using a carbon paste electrode and identified the peaks. Fig. 2.22 shows a

typical cyclic voltammogram for a carbon paste ilmenite electrode in H2SO4 (1 M)

solution. The following features were explained in terms of reactions taking place

during the scan from the rest potential in the positive direction and following

reversal back towards the negative direction.

87
Chapter 2: Literature Review

The curve obtained using 2 mg ilmenite at the scan rate of 10 mV/s showed

two anodic peaks A1 and A2 and two cathodic peaks C1 and C2. Peaks at A1 and

C1, as well as A2 and C2, are the reverse of each other, and assigned to the reactions

of titanium and iron, respectively.

Ti4+ + e- → Ti3+ (2.32)

Fe3+ + e- → Fe2+ (2.33)

During the forward sweep, oxidation of Fe(II) to Fe(III) occurs at peak A2.

Fe2+→ Fe3+ + e- (2.34)

Then, during reverse sweep, peak C1 was assigned to the reduction of Ti(IV) to
Ti(III).

TiO2+ + 2H++ e- → Ti3+ + H2O (2.35)

The areas measured under the peak of A2 (initial scan) and C1 are similar and are due

to partial dissolution of ilmenite according to reaction 2.36, which means that both

iron and titanium dissolve at the same rate:

FeTiO3 + 4H+ → Fe2+ + TiO2+ + 2H2O (2.36)

88
Chapter 2: Literature Review

Fig. 2.22. Voltammogram of Ilmenite (2 mg of Australian ilmenite ore) in H2SO4


(1M); scan rate = 10-3 V s-1; arrow indicates that the start of scan in the anodic
direction (Adriamanana et al., 1984).

Using similar techniques, Zhang and Nicol (2009) also presented a typical

linear sweep voltammogram in negative direction from the open circuit potential of

two different samples: (i) massive ilmenite electrode; (ii) ilmenite carbon paste

electode. In the voltammogram obtained for the massive ilmenite electrode shown in

Fig.2.23(A) (Zhang and Nicol, 2009) a small cathodic peak at 0.65 V was identified

by the authors due to the reduction of Fe(III) to Fe(II) either in solution (Eqs. 2.37)

or in the solid phase.

The reported measured potential of equimolar Fe(III)/Fe(II) sulfate couple in

1 M Na2SO4 at pH 0-0.5 at 25 oC on a platinum electrode is also 0.65 V (Senanayake

and Muir, 1988).

Fe3+ + e- → Fe2+ Eo = 0.69 V (1 M H2SO4 at 25⁰C) (2.37)

89
Chapter 2: Literature Review

The second cathodic peak at about 0.02 V was assumed to be associated with the

reduction of the mineral to Ti(III) in a reaction such as

2FeTiO3 + 6H+ + 2e-→ Ti2O3 + 3H2O + 2Fe2+ Eo = -0.10 V at 60 ⁰C (2.38)

There is an oxidation peak on the reverse sweep that could be the oxidation of Ti(III)

to Ti(IV). This peak is comparable with the results obtained by Adriamana et al.

(1984) as mentioned above. The second anodic peak is a oxidation peak for

Fe(II)/Fe(III) couple at 0.77 V (Zhang and Nicol, 2009).

(A) (B)

Fig. 2.23. Linear sweep voltammogram (5 mV/s) of massive ilmenite electrode (A)
carbon paste electrode (B) in 450 g/L H2SO4 solution at 60 oC. Negative sweep from
OCP, reversed at -0.4 V (Zhang and Nicol, 2009).

90
Chapter 2: Literature Review

(ii) Cyclic voltammetry of ilmenite in chloride solution

Cyclic voltammogram studies of ilmenite in chloride media, mainly in

hydrochloric acid, are not very common. There are few authors that discussed

ilmenite behaviour in chloride media as reported in Table 2.12 earlier. White et al.

(1994) studied the potentiodynamic polarization of ilmenite in 0.1 M NaCl solution

to investigate the reaction mechanisms of ilmenite under experimental conditions

applicable to natural weathering processes. Polarization of ilmenite electrodes to

positive and negative potentials relative to Eo produced increments in current

densities, as shown in Fig. 2.24.

Fig. 2.24. Potentiodynamic scans showing measured current plotted as function of


applied potential E for ilmenite electrode in anoxic pH 3 solutions. Vertical dashed
lines correspond to potentials defining the stability of water (White et al., 1994).

91
Chapter 2: Literature Review

The redox couples corresponding to the increase in anodic and cathodic currents are

reported as:

Anodic:

3FeTiO3 (ilmenite) → Fe2Ti3O9 (pseudorutile) + Fe2+ + 2e- (2.39)

And

Cathodic:

H2O + e-→ 1/2 H2 + OH- (2.40)

The authors stated that the anodic reaction (Eq. 2.39) passivated the

ilmenite surface, resulting in a relatively constant current density of 30 A cm-2 from

about 150 mV up to a potential around 1000 mV, at which O2 evolution takes place.

At low pH the application of negative potentials relative to Eo did not result in a

cathodic current peak for ilmenite. The absence of a reductive dissolution half

reaction for ilmenite under these conditions, compared to Fe3O4 which caused Eo to

shift to a much more negative potential in Fig. 2.21. reflects the cathodic reduction

of H2O to H2 (Eq. 2.40). Pseudorutile was found to be the weathering product of

ilmenite and this was confirmed by X-ray photoelectron spectroscopy (XPS) analysis

performed by the authors (White et al., 1994).

An anoxic weathering experiment conducted on ilmenite showed that Fe(II)

ions were released to the solution together with the formation of oxidized surface

products. For pure ilmenite, Fe(II) release does not involve a redox reaction (White

et al., 1994). However, ilmenite commonly forms a solid solution with hematite,

from which reductive dissolution can occur.

92
Chapter 2: Literature Review

Similar studies also have been conducted on reduced ilmenite in chloride

media, which also provided some information on the leaching behaviour of ilmenite.

Reduced ilmenite, produced in the process of reduction of weathered ilmenite using

coal at high temperature, is an intermediate product of the Becher process. It is used

to upgrade ilmenite to synthetic rutile (Becher et al., 1965). This reduced ilmenite

product contains a heterogeneous agglomeration of a poorly conducting rutile phase

and a highly conducting metallic iron phase. In the Becher process, the iron is

removed from reduced ilmenite by corrosion with oxygen in a 1% ammonium

chloride solution. Marinovich et al. (1995) and Kumari et al. (2002) conducted

electrochemical studies of reduced ilmenite using carbon paste electrodes in

ammonium chloride electrolytes. It was reported that the rutile phase catalyse the

anodic dissolution of iron.

2.5.4. Potentiostatic and coulometric study

McConnel (1978) conducted potentiostatic measurements on massive and

synthetic electrodes of ilmenite in sulfuric acid. Further studies were conducted by

Zhang and Nicol (2009) on massive and carbon paste electrodes of ilmenite. The

ilmenite was potentiostatically polarized at various potentials negative to the rest

potential for periods of up to several hours and the current was measured as a

function of time. On completion of the polarization, dissolved iron and titanium in

the resulting solutions were analyzed to calculate the average dissolution rate of iron

and titanium.

93
Chapter 2: Literature Review

Zhang and Nicol (2009) potentiostatically polarized a massive ilmenite

electrode at 0.1 V (SHE) for about 4 hours, and then examined the electrode surface

by optical microscopy. Fig. 2.25 shows the surface of massive ilmenite after it was

polarized at 0.1 V (SHE) in H2SO4 solution. The surface of the electrode indicated

distinctly non-uniform dissolution along the grain shown in Fig. 2.25 and it was

suggested that hematite was more reactive in this potential region, as consistent with

the thermodynamic data. The same observation was made with the massive ilmenite

electrode when it was kept at -1.2 V for an extended period in 2 M sulfuric acid at

40 oC (McConnel, 1978).

Fig. 2.25. Photomicrograph of the surface of a massive ilmenite electrode after


application of a potential of 0.1 V for 4 h in a 450 g/L H2SO4 solution at 60 oC.
Areas of A, B, C and D indicate corrosion of hematite lamellae (Zhang and Nicol,
2009).

94
Chapter 2: Literature Review

Application of a constant potential for a certain period of time to a

semiconducting ilmenite electrode in sulfuric acid solution had a marked influence

upon its dissolution rate (McConnel, 1978). There was no selective dissolution of

either iron or titanium from ilmenite at the applied potentials for the synthetic

ilmenite electrode (McConnel,1978). However, the massive ilmenite electrode

showed the initial presence of a more rapidly dissolving iron oxide phase compared

to titanium oxide. Similar results have been reported for natural ilmenite by Zhang

and Nicol (2009). This result can be misleading because of the influence of such

impurities associated with minerals (McConnel, 1978). Therefore, the use of

synthetic ilmenite free of such impurities in the electrochemical test work and

comparison of the results with those from massive ilmenite is of significant

importance in this thesis.

Fig. 2.26 shows the average dissolution rate as a function of potential for

massive and carbon paste ilmenite electrodes. The dissolution rates were higher at

lower applied potentials. The dissolution rate of iron was higher than that of titanium

for the first few hours. The calculated value of charge passed for a process involving

reduction with one electron per iron ion was found to be close to the rate of iron

dissolution, especially at 0.1 and 0.0 V (SHE) for massive ilmenite and at 0.3 V and

0.5 V for the carbon paste electrode. These observations are consistent with Eq. 2.41

or 2.42 (Zhang and Nicol, 2009):

{Fe2+ yFe3+ Ti4+(3+1.5y)O2-} + 3yH+ + ye-→ yFe2+ + FeTiO3 + 1.5yH2O (2.41)

Fe2O3 + 6H+ + 2e-→ 2Fe2+ + 3H2O Eo = 0.75 at 60 oC (2.42)

95
Chapter 2: Literature Review

These authors also suggest that the main reactions involved during reductive

dissolution of the ilmenite samples in the potential region from 0.5 V to 0.0 V are

given by Eqs. 2.41 and 2.42, together with the non-oxidative stoichiometric

dissolution of ilmenite given in Eq. 2.43. The value of ∆Go based on the HSC

database for Eq. 2.43 is -32.20 kJ mol-1 at 60 oC.

FeTiO3 + 4H+ → Fe2+ + TiO2+ + 2H2O ∆Go = -32.44 kJ mol-1at 60 oC (2.43)

(A) (B)

Fig. 2.26. Potentiostatic dissolution of massive (A) and carbon paste (B) ilmenite
electrode in 450 g/L H2SO4 at various potential at 60 oC; (▲) Ti, (◊) Fe, (■) Charge
(Zhang and Nicol, 2009).

Zhang and Nicol (2009) also studied the effect of temperature on the rates of

dissolution by applying a constant potential of 0.0 V to massive and carbon paste

ilmenite electrodes. As can be seen from Fig. 2.27 the rate of the dissolution of iron

and titanium increased with increasing temperature. The activation energy values for

the dissolution of massive and carbon paste ilmenite electrodes were reported to be

44 and 50 kJ mol-1, respectively.

96
Chapter 2: Literature Review

These reported activation energy values were lower than other published

values of 90 kJ mol-1 for free dissolution of ilmenite under similar conditions (Han et

al., 1987; McConnel, 1978; Zhang and Nicol, 2009). The same observation was

made by White et al. (1994) on Fe-Ti bearing minerals in natural weathering

environments, in which iron dissolved more quickly than titanium at 0.0 V (Zhang

and Nicol, 2009).

(A) (B)

Fig. 2.27. Effect of temperature on the rate of the dissolution of (A) massive ilmenite
and (B) carbon paste ilmenite at 0 V potential in 450 g/L H2SO4; (▲) Ti, (◊) Fe, (■)
Charge (Zhang and Nicol, 2009).

97
Chapter 2: Literature Review

2.6. Conclusion

The literature review has identified and compiled the established and

proposed ilmenite processing flowsheets for producing synthetic rutile and also

upgrading ilmenite ore to TiO2 pigment. Ilmenite is one of the most important ores

from which titanium can be recovered. Despite environmental problems and high

cost that producers must face, sulfate and chloride route still remain as the primary

route for producing titanium. Therefore, research on ilmenite upgrading using

hydrometallurgical route has increased significantly. The review also provides the

general reaction stoichiometry involving various reagents used in leaching system

such as sulfuric and hydrochloric acid. The advantages and disadvantages of HCl

and H2SO4 also have been discussed. Through this review it has been identified that

there is lack of understanding on reaction mechanism of ilmenite dissolution in HCl

solutions and the role of reducing agent. Although extensive electrochemical studies

relevant to sulfuric acid solutions have been conducted little attention has been paid

to HCl solutions which provides an opportunity for further studies

This study aims to understand the reaction mechanism of the dissolution

behaviours of iron and titanium in hydrochloric acid solutions which offer a more

aggressive but environmentally friendly route compared to sulfuric acid and produce

less waste or byproduct. It can easily be regenerated and recycled. This study

comprise of three different approaches to achieve the objectives in 8 chapters.

Chapters 1 and 2 demonstrated the theoretical knowledge to identify the gaps of

knowledge in this area through previous works by other researchers. Chapters 3 and

4 describe materials and methodologies for mineral characterization using standard

techniques such as XRD, SEM, EDX and optical microscopy, leaching and

collecting kinetic data using well defined synthetic surfaces and natural particles of

98
Chapter 2: Literature Review

ilmenite. Chapter 5 is dedicated to the study of dissolution of iron and titanium from

synthetic ilmenite rotating disc under reducing and non-reducing conditions to

investigate the reaction mechanism under reducing and non-reducing conditions

using synthetic ilmenite without any interference from other minerals. Chapter 6 is

focused on comparative electrochemical studies on massive, natural and synthetic

ilmenite, leading to important findings in reaction mechanisms, different behaviours

of iron and titanium dissolution with different leaching reagents and the effect of

weathering/alteration degree of natural ilmenite. The role of reducing agent and rate

control step of reducing reaction will be tested. Chapter 7 presents the leaching

reactivity of natural ilmenite samples with different degree of alteration and

weathering resulting from different origin and locations to examine and validate the

proposed kinetic models and reaction mechanism using different natural ilmenite

samples. Chapter 8 concludes and summarizes the findings from three chapters

above and suggests recommendations for future work.

99
Chapter 3 : Materials and Methods

3 CHAPTER 3

MATERIALS AND METHODS

3.1. MATERIALS

3.1.1. Concentrates

Three different ilmenite concentrates have been obtained from different

sources and labeled as Iluka (IL), Tiwest (TW) and North Capel (NC). These

samples were used for the electrochemical and leaching experiments.

The concentrates, as received, were in the form of heavy mineral sand. They

were mixed homogeneously, split and sampled to obtain a representative sample

prior to the analyses. About 100 g sample of each ilmenite concentrate was ground in

an iron-ring mill for 30 seconds. The crushed samples were then subjected for

screening and sieving to obtain four different size fractions (+45-53 µm, +53-63 µm,

+63-75 µm,+75-90 µm). All samples were stored at room temperature. Samples of

concentrates were analyzed chemically by the Ultra Trace laboratory in Perth,

Western Australia by the following method. The samples were dried and ground by

hand in a mortar and pestle. A known mass of the sample was fused with sodium

peroxide and the melt was subsequently dissolved in dilute hydrochloric acid for

analysis. Because of the high furnace temperature volatile elements were lost.

Elements analyzed were Ti, Fe, Mg, Mn, Al, Zn, Ca, Si, Nb, Cr, V, S, P and

moisture content.

100
Chapter 3 : Materials and Methods

3.1.2. Materials and reagents

All reagents used were of analytical grade. Millipore water was used

throughout the experiments. Table 3.1 gives a list of materials and reagents used in

the experiments described in this thesis.

Table 3.1
Lists of materials and reagents.

Reagents Formula Purity (%) Grade Manufacturer

Hydrochloric acid HCl 32 AR AJAX Fine Chem


Pty Ltd.

Sulfuric acid H2SO4 97-99 AR MERCK

Metallic iron Fe 99.0 Pro analysis MERCK

Tin powder Sn 99.0 AR MERCK

Tin(II) chloride SnCl2 >97 AR MERCK

Iron(III) chloride FeCl3 97.0 AR BDH


anhydrous

Ferrous chloride FeCl2.4H2O 99 AR Sigma-Aldrich


heptahydrate

Titanium(IV) Oxide TiO2 99.9 AR Sigma-Aldrich

Iron(III) Oxide Fe2O3 AR Sigma-Aldrich

Titanium(III) Oxide Ti2O3 99.9 AR Sigma-Aldrich

Platinum wire Pt 99.9 AR

Nitrogen gas N2 - High purity BOC


grade

Argon gas Ar - High purity BOC


grade

Heavy mineral oil - - AR AJAX Fine Chem


Pty Ltd.

Graphite powder C 98 High grade AJAX Fine Chem


Pty Ltd.

101
Chapter 3 : Materials and Methods

3.1.3. Synthetic ilmenite preparation

Synthetic ilmenite pellets for use in the rotating disc dissolution experiments

were prepared in the following manner. Fine powders of high purity ferric oxide,

titanium dioxide and metallic iron were mixed in the proportion required for the

reaction:

Fe2O3 + 3TiO2 + Fe = 3FeTiO3 (3.1)

The mixture was made up in a mortar and pestle and pressed into a large pellet using

low pressure. The pellets were then supported on a flat surface of Supercast Alumina

inside large diameter Vycor glass tubing in an electric furnace. The Vycor tubing

was fitted with removable gas inlet and outlet points at either end of the furnace. The

Vycor tube contained a baffle system, to ensure that the incoming gas rapidly

reached the working temperature of the furnace, and also to prevent excessive heat

loss by radiation along the tube. A thermocouple probe inside the Vycor tube,

adjacent to the pellets, was used to calibrate the furnace controller and at 1200ºC the

temperature could be maintained constant within ± 10ºC.

The pellets were heated initially by raising the temperature gradually to

850 ºC over two hours and maintaining that temperature for three hours. The

temperature was then slowly raised to 1200 ºC over 3 hours and held at 1200 ºC for

another 3 hours. After 3 hours of heating, the temperature was gradually reduced to

room temperature in 3 hours. Throughout all experiments, high purity dry argon was

passed through the Vycor tube at a rate of approximately 10-4 dm3 s -1.

102
Chapter 3 : Materials and Methods

After being ground in a ring grinder, 3.5 g portions of the black powder of

synthetic ilmenite were pressed into pellets in a 2.54 cm diameter die using a

hydraulic press. A pressure of 197 MPa was used, giving discs of approximately

2 mm thickness. These were supported on alumina and replaced in the furnace. The

temperature was raised to 1200 ºC over 4 hours and maintained for 12 hours under

argon, a long sintering time being required to ensure that the pellets were non-

porous. After sintering, the pellets had shrunk to 2.2 cm diameter and showed no loss

in weight. The pellets were then analyzed by X-ray diffraction (XRD) to confirm

only ilmenite (FeTiO3) was present. The chemical analysis of the ground pellets was

also conducted.

3.1.4. Rotating disc dissolution experiment

The synthetic ilmenite discs prepared in the manner described above were

used as a sample for the dissolution experiments in HCl solutions. The 2.2 cm discs

of synthetic ilmenite were cast into 2.7 cm diameter epoxy mounts. The disc holder

had an overall diameter of 2.7 cm at the bottom and a diameter of 1.2 cm at the top.

The disc holder was screwed into the shaft hole of rotating stand (labeled as C in Fig.

3.1) and the bottom of the holder was fitting inside the polyethylene lid of the

reaction vessel, to protect the exposed shaft from hydrochloric acid. Care was taken

to ensure that the holder was centered on the shaft. The illustrated diagram of

rotating disc setup is shown in Fig. 3.1.

The rotating disc of synthetic ilmenite was used to investigate the effect of

disc rotation speed, temperature, acid concentration and reducing agent upon the

dissolution rate of ilmenite in hydrochloric acid.

103
Chapter 3 : Materials and Methods

Fig. 3.1. An illustrated diagram of rotating disc dissolution experiment. A : rotating


motor; B : rotating stand; C : shaft; D : platinum electrode; E : reaction vessel; F :
synthetic ilmenite disc; G : tube; H : jack; I : temperature controller; J : water bath.

Most of the experiments were carried out at a temperature of 60.00±0.05oC

while temperatures of 70.0 oC and 80.0 oC were used for the determination of the

effect of temperature.

In all experiments, the reaction vessel was initially lowered and filled with

250 mL of hydrochloric acid, then raised and allowed to reach equilibrium with

nitrogen flowing. Before the start of each experiment, the disc was washed with de-

ionized water and dried under a stream of compressed air. The vessel was then

lowered, the disc fitted to the shaft and the electric motor turned on at the desired

rotation speed. The reaction was started by raising the vessel and clamping it,

starting the disc spinning before immersion to ensure that no gas bubbles were

trapped on the ilmenite surface. Aliquots (10.0 mL) of the reaction solution were

withdrawn through the sample port at known intervals and analyzed for total iron

and titanium.

104
Chapter 3 : Materials and Methods

After the completion of each experiment the disc holder was removed, washed and

dried, then examined for the disintegration of the pellets or leaks. All experiments

were duplicated and the mean values were reported as the results. All results of

rotating disc dissolution will be discussed in Chapter 5 and a summary of the results

is presented in Appendix A2.

3.2. Electrochemical measurements

3.2.1. Equipment, electrodes and electrochemical cell

The electrochemical studies were performed using a conventional three-

electrode system (reference, counter and test electrode). The complete

electrochemical set up is schematically shown in Fig. 3.2. A modified Metrohm glass

cell covered with a detachable lid was used as a reaction vessel. The working

electrode was a rotating disc fabricated at Murdoch University using pieces of the

massive ilmenite samples. The counter electrode was a 0.5 mm diameter platinum

wire.

For powdered ilmenite, a carbon paste electrode was used by placing carbon

paste on top of a specially built electrode (see Fig. 3.3). The carbon paste was made

by mixing ilmenite with graphite and spectroscopic grade paraffin oil according to

the ratio, ilmenite: graphite: paraffin = 12:3:3 by mass, until homogenous. The

carbon paste mixture then was filled into cylindrical cavity of 2.7 mm depth and

7 mm diameter. The surface was smoothed and the excess paste removed using

spatula.

105
Chapter 3 : Materials and Methods

The reference electrode was a saturated calomel electrode (SCE)/ saturated

KCl (SCE, E = +0.245 V versus SHE) which was joined to the main cell by a luggin

capillary passing through a screw fitting in the base of the cell. The tip of luggin

capillary was placed ~ 0.5 cm below the surface of the working electrode. All

potentials are reported with respect to the standard hydrogen electrode (SHE). The

disc electrode was rotated (RDE) using a direct current (dc) motor. The experimental

setup is shown in Fig. 3.2.

The electrodes were rotated at 200 rpm in all tests unless stated otherwise.

The experiments were carried out using a Princeton Applied Research Model 173

Potentiostat/Galvanostat and Model 376 Logarithmic Current Converter together

with an EG&G PARC Model 175 Universal Programmer. Analogue data from the

potentiostat were collected using a National Instrument data acquisition board

controlled by SoftCorr II software. The temperature of the electrolyte was controlled

by using a water thermoregulator (Rowe Scientific Pty Ltd).

106
Chapter 3 : Materials and Methods

Fig. 3.2. A schematic diagram of the electrochemical cell and experimental set up. (A) Metrohm glass cell; (B) water jacket (C) luggin capillary
(D) reference electrode ; (E) counter electrode; (F) working electrode; (G) water bath; (H) N2 cylinder; (I) rotater; (J) potentiostat; (K) computer;
(L) printer; (M) rotater controller; (N) stand ; (O) temperature controller; (P) platinum wire.

107
Chapter 3 : Materials and Methods

Fig. 3.3. A schematic diagram of special electrode used for carbon paste ilmenite
electrode.

3.2.2. Electrode materials and electrolytes

The hard rock ilmenite mineral was obtained from the Mineral Collection of

Murdoch University. The massive ilmenite pieces were cut into cylindrical rods and

bonded to a brass screw using conductive silver epoxy resin. The surface area of

exposed massive ilmenite electrode is 0.25 cm2. Following this, the electrode was

embedded in an epoxy resin (Araldite LC191) to insulate the body of the electrode,

leaving only a disc of known area exposed to the solution. Prior to all tests, the

exposed surface of working electrodes was wet polished using 240, 400, 600, 800

then 1200 silicon carbide papers and rinsed with deionised water.

The synthetic ilmenite was prepared according to the procedure explained in

Section 3.3 and used as a carbon paste electrode (Fig. 3.3).

108
Chapter 3 : Materials and Methods

The powder samples of ilmenite concentrate were obtained by Murdoch University

over the years from Iluka Resoucers Limited, Tiwest Resouces Limited and North

Capel Corporation (particle size is between 75-90 µm with specific gravity 4.3). The

exposed surface of the working electrode was circular and planar with geometrical

surface area of 1.52 cm2. The surface area of exposed ilmenite particles for the

carbon paste electrode was based on the procedure given by Zhang and Nicol (2009).

The experiments were conducted using 4 M solutions of hydrochloric acid with or

without addition of iron(II) and/or iron(III) ions and tin(II) ions, in the temperature

range of 50, 60, 70, 80°C. A 4 M hydrochloric acid solution was then used as the

base electrolyte for subsequent experiments which were conducted at 60 °C.

3.2.3. Batch leaching experiments

The batch leaching experiments were conducted on ground ilmenite

concentrates from three different sources as mentioned earlier. Batch leach tests were

conducted in a 500 mL cylindrical glass reactor, with the slurry agitated by a floating

magnetic stirrer. The stirrer was rotated at a speed that completely suspended the

ilmenite particles. The reactor was heated using a magnetic hot-plate stirrer

(IKAMAG) equipped with a temperature control (± 2°C) system. A condenser with

circulating water was employed to minimize evaporation of the solutions from the

reactor during the tests.

Hydrochloric acid solutions were then added and heated to the required

temperature. When the set temperature was reached, the samples of ilmenite were

transferred into the reactor.

109
Chapter 3 : Materials and Methods

The leaching experiments were carried out for 6 hours of leaching time or as

stated otherwise. Samples of 2 mL of the leaching solution were taken from the

reactor using syringe at various time intervals and filtered through Whatman

Celulose Nitrate membrane filters with pore size of 0.45 µm. The schematic diagram

of the leaching set up is shown in Fig. 3.4.

3.2.4. Solution potential measurements

The solution potential (Eh) during reaction was continuously monitored and

recorded with the LabView data acquisition system using platinum and ilmenite

electrodes. A platinum electrode combined with a Ag/AgCl/KCl reference electrode

(Methrom, Switzerland) was used for the determination of solution or slurry

potential. The potential response of the ilmenite electrode was also recorded

simultaneously with the same reference electrode. It should be noted that the

reported Eh is generally a mixed potential because there are many reactions taking

place at the electrode (see section 2.5.2 for mixed potential).

110
Chapter 3 : Materials and Methods

Fig. 3.4. A schematic diagram for dissolution experiment. (i) retort stand; (ii)
temperature regulator; (iii) condenser; (iv) reactor; (v) floating magnetic stirrer; (vi)
hotplate with temperature and magnetic controller; (vii) sample drawn vial.

3.3. Analytical measurement

3.3.1. Atomic absorption spectroscopy

Atomic absorption spectroscopy was used for the determination of the total

metal ion (Fe) concentration. The instrument employed was GBC 933 atomic

absorption spectrometer together with GBC Avanta AAS Version 2.0 software.

Calibration using standard solutions was done prior to the analysis. The leach liquor

samples were diluted from 10 up to 100 times in order to get the concentrations

within the linear calibration range.

111
Chapter 3 : Materials and Methods

3.3.2. UV/Visible Spectrophotometry

The titanium concentration in the samples was analysed using

spectrophotometry by the hydrogen peroxide method (Vogel, 1978) at a wavelength

of 410 nm. An Agilent 8435 UV-Visible spectrophotometer was used to scan from

900 to 350 nm at a rate of 600 nm/min. Standard solutions were prepared for the

calibration curve. Each leach liquor was diluted from 5 up to 50 times in order to get

the concentration within the linear calibration range. To measure the absorbance, the

leach liquors were transferred to a Unicam 1.0 cm Optical Path Length Quartz Glass

cell for immediate spectral analysis.

3.3.3. Inductive coupled plasma optical emission spectrometry (ICP-


OES)

The ICP-OES instrumental analysis technique was employed mainly for

determination of total dissolved iron, titanium and other metal ions in the solutions

used in this study. The instrument used was a Varian ICP model Liberty 200. The

uncertainty of the analysis was ± 5%.

3.3.4. Assay analysis

A representative sample of the feed from three different sources was sent to

Ultra Trace for analysis by X-ray fluorescence spectroscopy (XRF). The moisture

values shown are the chemically bound moisture determined gravimetrically

between 105 and 1000 °C.

112
Chapter 3 : Materials and Methods

As XRF gives only an elemental analysis, the components shown may not represent

the actual form of the metal present in the concentrate.

3.3.5. X-ray diffraction

This technique was employed to qualitatively analyse the composition and

structure of crystal or powder samples such as minerals and residues. The XRD

patterns were obtained using an EMMA Enhanced Multi Materials Analyzer (CuKα)

GBC X-ray diffractometer.

3.3.6. Particle size analysis

Particle size distribution is an important factor which could influence the

physical properties of a solid powder. Analysis of particle size is necessary and

sometimes liberation is a key step to achieve target size, which may influence

process reactions.

The particle size distribution of the concentrates was obtained on a Microtac

s3500 Particle Size Analyzer with Microtac FLEX Software. The particle size

distribution of ilmenite concentrates was expressed by span value. Span is defined as

(d90-d10)/d50, where d10, d50, d90 value represent the 10th, 50th and 90th of

cumulative volume passing, respectively. If a low span value was obtained (< 1), the

median value, d50 was used as mean diameter while if high span value was obtained,

it was described by the volume moment diameter (VMD).

113
Chapter 3 : Materials and Methods

3.3.7. Optical microscopy

The mineral surface was examined using a Nikon Optical Microscope (model

EPIPHOT 200).

3.3.8. Scanning electron microscopy

Scanning electron microscopy (SEM) investigations were obtained on a

Scanning Electron Microscope (Philips XL 20) with an Energy-Disperse X-ray

Analysis (EDX) system. The powdered samples were prepared by two methods:

1. The ilmenite samples and leach residue powder were mounted on

aluminium stubs using conductive two sided tape for the secondary electron

(SE) surface analysis.

2. The embedded samples were prepared by mounted resin method. The

samples were mounted on aluminium stubs using conductive tape and

analysed using the backscattered electron (BSE) technique.

3.3.9. Brunauer-Emmet-Teller (BET) surface area analysis

The BET analysis provides precise specific surface area evaluation of

materials by nitrogen multilayer adsorption measured as a function of relative

pressure using a fully automated analyzer. The technique encompasses external area

and pore area evaluations to determine the total specific surface area in m2/g yielding

important information in studying the effects of surface porosity and particle size in

many applications. All samples of ilmenite were sent for BET surface area analysis

at C.S.I.R.O. Australia to determine the relative surface area of different samples of

ilmenite.

114
Chapter 4: Characterization of synthetic and natural ilmenite

CHAPTER 4

4. CHARACTERIZATION OF SYNTHETIC AND


NATURAL ILMENITE

4.1. Introduction

As described in Chapter 3 (Section 3.1), three concentrates from natural

ilmenite samples collected from various locations were used in this study. Synthetic

ilmenite was synthesized in Murdoch University laboratories. These samples were

characterized in detail and described in the following sections.

4.2. Description of minerals

A representative sample of the feed from four different sources was sent to

Ultra Trace commercial laboratory for analysis by X-Ray fluorescence spectroscopy

(XRF). The results are shown in Table 4.1. The moisture content shown is the

chemically bound moisture determined gravimetrically between 105 oC and 1000 °C.

The collected ilmenite sample in this study has a TiO2 content ranging from 52.3%

to 62.9% (w/w) and a ferrous and ferric iron content which also covers a wide range,

from 10.3% to 50.0%.

115
Chapter 4: Characterization of synthetic and natural ilmenite

Table 4.1
Chemical composition of representative samples of ilmenite ores.
Theoritical Synthetic
Iluka Tiwest North Capel
Component ilmenitea ilmenite
% (w/w) % (w/w) % (w/w)
% (w/w) % (w/w)
TiO2 52.7 52.3 59.1 62.9 53.9
FeO 47.3 47.1 10.3 3.0 25.4
Fe2O3 34.6 31.6 45.0
Fe Total 36.8 36.7 24.2 22.1 31.5
Al2O3 1.21 0.98 0.5
MgO 0.86 0.2 0.16
MnO 0.57 1.01 1.5
CeO2 0.072 0.034 0.018
SiO2 1.12 0.71 0.59
SO3 0.05 0.02 0.01
Cr2O3 0.466 0.138 0.04
P2O5 0.152 0.065 0.023
V2O5 0.19 0.17 0.11
Nb2O5 0.17 0.169 0.126
CaO 0.06 0.01 0.01
Moisture 0.36 0.49 0.2
a
Egerton(2000)

4.2.1. Synthetic ilmenite

The synthetic ilmenite samples have been synthesized using the procedure

described in Chapter 3. Fig. 4.1 summarizes the XRD patterns of synthetic ilmenite

before and after sintering. The XRD pattern of powdered synthetic ilmenite before

sintering indicated a poorly crystalline nature as well as other phases including

hematite (Fe2O3) and rutile (TiO2). After the first stage of sintering, ilmenite peaks

appeared with a few minor peaks. After the second stage of sintering, the XRD

pattern indicated that only FeTiO3 was present, giving a pattern with sharp intense

peaks. The results of chemical analysis of the synthetic ilmenite shown in Table 4.1

indicate that the stoichiometry of synthetic ilmenite is close to that of theoretical

ilmenite.

116
Chapter 4: Characterization of synthetic and natural ilmenite

Fig. 4.1. XRD patterns for synthetic ilmenite(SI) (KαCo, λ = 1.788965).

Examination of the morphology of a synthetic ilmenite disc by SEM and

EDX showed that the surface of the disc was rough and EDX indicate only the

presence of Fe, Ti and O (Fig. 4.2(a)-(b)).

117
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.2. (a) SEM surface morphology of synthetic ilmenite disc and (b) EDX
analysis.

118
Chapter 4: Characterization of synthetic and natural ilmenite

4.2.2. Iluka concentrates

This sample was collected and concentrated from the black sand from a mine

site in Western Australia. The chemical analysis listed in Table 4.1 indicates that the

sample can be categorized as slightly weathered ilmenite containing 59.1% TiO2 and

approximately 34.6% of the iron in ferric form. The particle size distribution of Iluka

ilmenite concentrate illustrated in Fig. 4.3 showed the bimodal distribution. The

analysis of cumulative passing showed that 90% of the particles have a size below

249.3 µm, 50% below 213.8 µm and 10% below 118.8 µm. From these values, the

width of the particle size distribution expressed as span value is presented as 0.61.

This relatively low span value indicates a small size distribution. Therefore, the

value d50 was taken as mean diameter for this type of distribution. The XRD pattern

of Iluka concentrates indicated high intensity of ilmenite peaks (Fig. 4.4) which

overlap with hematite peaks. The presence of hematite, evident from XRD scans, is

consistent with the chemical analysis of Iluka ilmenite.

Fig. 4.3. Particle size distribution of Iluka ilmenite.

119
Chapter 4: Characterization of synthetic and natural ilmenite

Fig. 4.4. XRD patterns for Iluka ilmenite (KαCo, λ = 1.788965).

120
Chapter 4: Characterization of synthetic and natural ilmenite

The morphological study of the Iluka concentrate was conducted using SEM

and optical microscopic analysis. Fig. 4.5 shows the surface morphology of Iluka

concentrate from which the angular shape of the particles can be seen. The polished

section (Fig. 4.6(a)) shows some inclusion of non-titanium minerals (in dark colour)

and fractures. This was supported by EDX analysis which showed a small amount of

iron, vanadium and aluminium in Fig. 4.6(b). Fig. 4.7(a) and the EDX analysis (Fig.

4.6(b)) indicated the association of particles with other oxides such as chromium and

aluminium oxide, also evident from the chemical analysis (Table 4.1). This

examination showed that the Iluka concentrate has been altered or weathered by a

natural weathering process. Further examination of the morphology of this sample is

shown by the optical micrograph (Fig. 4.8). The particles showed various degrees of

alteration and weathering along the grain, fractures and a patchy pattern inside the

particles. This shows the complexity of the mineralogy and texture of the grains.

Fig. 4.5. Surface morphology of Iluka ilmenite concentrate.

121
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.6. (a) Surface morphology of polished section of Iluka ilmenite sample (b)
EDX analysis at low magnification.

122
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.7. (a) Surface morphology of polished section of Iluka ilmenite sample and (b)
EDX analysis on selected area, at higher magnification.

123
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.8. Optical micrograph showing (a) the weathering effect (b) altered particles
with patchy pattern on Iluka sample.

124
Chapter 4: Characterization of synthetic and natural ilmenite

4.2.3. North Capel concentrate

The Capel area of south-west Western Australia has extensive buried beach

sand deposits and this sample is a typical commercial concentrate from the region.

According to the chemical analysis, this sample contained 53.9% TiO2 and showed

some alteration, with 45.0% of the iron being in the ferric form. The sand is lighter

in colour than Iluka ilmenite, being greyish-black, but still retains the submetallic

lustre of ilmenite. The particle size distribution of North Capel ilmenite concentrate

is shown in Fig. 4.9. Based on cumulative passing size, 90% of the particles have a

size below 199.5 µm, 50% below 163.0 µm and 10% below 96.6 µm.

From these values, the width of the particle size distribution, expressed as

span value, is reported as 0.63 and d50 was taken as the mean diameter for this type

of distribution. This relatively low span value indicates a small size distribution.

From the surface morphology examination, the particles are more rounded but low in

sphericity (Fig. 4.10) compared to Iluka sample particles, which have an angular

shape (Fig. 4.5).

Examination of the XRD pattern of powdered North Capel ilmenite indicated

that no minerals other than ilmenite were present, as shown in Fig. 4.11. The

polished section of North Capel ilmenite in Fig. 4.12(a-b) and Fig. 4.13 (a-b) showed

the inclusion of other minerals by the patchy pattern with dark and grey appearance

under high magnification. Further EDX analysis identified the presence of other

elements such as vanadium, manganese and iron along with titanium. Further

investigation can be seen by the optical micrograph in Fig. 4.14 which also shows

the weathering effect on the mineral. Additionally, weathering effect on mineral was

also observed (Fig 4.14).

125
Chapter 4: Characterization of synthetic and natural ilmenite

Fig. 4.9. Particle size distribution of North Capel ilmenite.

Fig. 4.10. Surface morphology of North Capel ilmenite particles.

126
Chapter 4: Characterization of synthetic and natural ilmenite

Fig. 4.11. XRD patterns for North Capel ilmenite (KαCo, λ = 1.788965).

127
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.12. (a) Surface morphology of polished section of North Capel ilmenite
sample and (b) EDX analysis at low magnification.

128
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.13. (a) Surface morphology of polished section of North Capel ilmenite
sample and (b) EDX analysis at higher magnification.

129
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.14. Optical micrograph showing the weathering effect on North Capel sample
(a) altered particles (b) particles with patchy pattern.

130
Chapter 4: Characterization of synthetic and natural ilmenite

4.2.4. Tiwest concentrate

This sample is also from a commercial concentrate from North of Perth,

Western Australia. The concentrate is used in the Becher process to produce

synthetic rutile. According to the chemical analysis, this sample contains the highest

TiO2 content of about 62.9%, compared to Iluka and North Capel samples. About

31.6% of iron is in ferric form while only 3.0% of iron is in ferrous form. This

sample is very different from the other two samples discussed above. The particle

size distribution of the Tiwest sample is shown in Fig. 4.15. Based on cumulative

passing, 90% of particles have a size below 237.4 µm, 50% below 200.3 µm and

10% below 146.7 µm. Again, the value d50 was taken as mean diameter for this type

of distribution.

The XRD pattern of the Tiwest sample shown in Fig. 4.16 is complex,

containing intense peaks for ilmenite and many other weaker peaks. The

identification of other minerals present is difficult due to overlapping of peaks;

however, hematite is definitely present with possible traces of pseudorutile and

rutile. Tiwest ilmenite is best described as a highly altered ilmenite containing

ilmenite, hematite, pseudorutile and rutile, all of poor crystallinity.

131
Chapter 4: Characterization of synthetic and natural ilmenite

Fig. 4.15. Particle size distribution of Tiwest sample.

132
Chapter 4: Characterization of synthetic and natural ilmenite

Fig. 4.16. XRD patterns for Tiwest concentrate (KαCo, λ = 1.788965).

133
Chapter 4: Characterization of synthetic and natural ilmenite

From the surface morphology examination, the particles exhibit well-rounded

and homogeneous shapes and sizes (Fig. 4.17). The polished section of Tiwest

ilmenite in Fig. 4.18(a) and Fig. 4.19(a) shows the high sphericity of particles

compared to North Capel and Iluka sample. The EDX analysis indicated high

contents of titanium and iron with some traces of vanadium, as shown in Fig. 4.18(b)

and Fig. 4.19(b). Further investigation by optical micrograph is shown in Fig. 4.20,

which also shows the weathering effect on the mineral.

Fig. 4.17. Surface morphology of Tiwest ilmenite sample.

134
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.18. (a) Surface morphology of polished section of Tiwest ilmenite sample and
(b) EDX analysis at low magnification.

135
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.19. (a) Surface morphology of polished section of Tiwest ilmenite sample and
(b) EDX analysis at higher magnification.

136
Chapter 4: Characterization of synthetic and natural ilmenite

(a)

(b)

Fig. 4.20. Optical micrograph showing the weathering effect on Tiwest ilmenite
sample.

137
Chapter 4: Characterization of synthetic and natural ilmenite

4.3. Summary and conclusion

This chapter described the characterization of the three ilmenite samples that

have been collected from different locations in Western Australia and synthesized

ilmenite samples prepared in the laboratory. A summary of some of these results is

given in Table 4.2 which is useful in the discussion of chemical and physical

properties of the samples in conjunction with the results of the dissolution

experiments described later. Table 4.2 is only for general comparison/summary. The

ferric iron to total iron ratio is a measure of the degree of oxidation, while the

titanium to total iron mole ratio is a measure of the extent of the leaching of the iron

from the mineral during the weathering process, assuming that the titanium is not

leached. Both these ratios are calculated from the data given in Table 4.1. Also listed

in Table 4.2 are the identified phases present in each sample, which will be used in

further discussion in remaining chapters.

Table 4.2
Summary of certain properties of ilmenite samples.

Surface Mole ratio Mineral phase


Ilmenite Particles shape
areaa Fe3+/Fe Ti/Fe present

Theoretical - 0.00 1.00 Ilmenite


Synthetic - - 0.99 Ilmenite
Ilmenite,
Iluka Sub-angular 0.35 0.99 1.60
hematite
Ilmenite,
North Capel Sub-angular 0.47 1.00 1.20
hematite
Ilmenite,
hematite,
Tiwest Rounded 10.06 1.00 1.90
pseudorutile,
rutile
a
From B.E.T. surface area analysis (unit in m2/g)

138
Chapter 5: Rotating disc dissolution of SI in HCl

CHAPTER 5

5 DISSOLUTION OF SYNTHETIC ILMENITE


ROTATING DISC IN HYDROCHLORIC
ACID SOLUTIONS

5.1. Introduction

The dissolution of ilmenite in acid solution has been studied by many

researchers using various methods to improve the process, particularly in chloride

solution (Jackson and Wadsworth, 1976; Tsuchida et al., 1982; van Dyk et al., 2002;

Das et al., 2013). Studies using natural ilmenite demonstrated different reactivity

towards lixiviants (Sasikumar et al., 2007). Ideally, ilmenite has a composition of

52.7% TiO2 and 47.3% FeO. Many of the world’s heavy mineral beach sand

deposits contain minerals that are the result of the natural weathering and alteration

of ilmenite (Gambogi, 2009) and thus contain some iron in the form of iron(III)

(Egerton, 2000). The weathering process not only influences the titanium content of

the ore but it also appears to change its chemical behaviour. Ilmenite concentrates

that contain significant levels of impurities, either internal or external to the grains,

have a lower market value and so there is considerable incentive for producers to

carefully characterize the nature and mode of occurrence of the impurities.

139
Chapter 5: Rotating disc dissolution of SI in HCl

It has been suggested that the greater reactivity shown by some ilmenite ores

towards sulphuric acid may be a result of the mineral being relatively unaltered

(Barton and McConnel, 1978). The Murso process (Sinha, 1973) discussed in

Chapter 2 uses a 20% HCl solution to dissolve ilmenite under non-rigorous

conditions, producing a ‘synthetic ilmenite’ structure with enhanced reactivity.

Consequently, Barton and McConnel (1978) synthesized ilmenite and studied the

dissolution kinetics using a disc of constant surface area rotated in sulfuric acid. The

dissolution reaction was found to be chemically controlled with an apparent

activation energy of 90.0 kJ mol-1 in the temperature range from 65 to 85 oC.

Reductive leaching of ilmenite in HCl solutions has been studied by previous

researchers (Ibrahim et al., 2003; Mahmoud et al., 2004; Lasheen, 2005; El-Hazek et

al., 2007). The dissolution rate of ilmenite increases significantly in the presence of

reducing agents. The reaction of FeO.TiO2 with hydrochloric acid alone produces

FeCl2 and TiOCl2 (Eq. 5.1) but the consumption of acid is larger due to the reaction

with the oxidised iron (Fe2O3) present in the natural ore (Ibrahim et al., 2003).

FeTiO3 + 4HCl → Fe2+ + TiO2+ + 2H2O + 4Cl- (5.1)

Reducing agents such as iron powder can produce FeCl2 and TiCl3 but the

latter undergoes oxidation to TiOCl2 by Fe2O3 and re-precipitates as synthetic rutile,

TiO2. The enhanced kinetics of iron leaching and the production of synthetic rutile is

an advantage.

140
Chapter 5: Rotating disc dissolution of SI in HCl

Due to the fast reaction of iron powder with hydrochloric acid itself, the

delayed addition of iron powder (after the leaching of titanium) has improved the

synthetic rutile grade (Ibrahim et al., 2003). Therefore, the main objective of this

chapter is to understand the dissolution behaviour of synthetic ilmenite in HCl in the

absence of Fe2O3 using a rotating synthetic ilmenite disc and SnCl2 as the reducing

agent, as it does not interfere with the analysis of the iron concentration in solution.

A comparison under non-reducing and reducing conditions is also made.

5.2. Dissolution of synthetic ilmenite in hydrochloric acid using


rotating disc

The rotating disc (RD) technique was employed to study the dissolution of

synthetic ilmenite in 4 M HCl at 60 oC for 8 hours. The diameter of the rotating disc

was 2.2 cm as mentioned in section 3.1.4, chapter 3. The experiments for synthetic

ilmenite were designed to measure the effect of system variables including HCl

concentration, temperature, rotation speed and the presence of reducing agents on the

rate of ilmenite dissolution in order to determine the rate controlling reaction(s) and

mechanism(s). In these experiments, the disc-holder was removed after completion

at each experiment and the disc was re-polished before it was used for every new

experiment. All raw data are presented in Appendix A2.1.

141
Chapter 5: Rotating disc dissolution of SI in HCl

5.2.1. Effect of temperature

Fig. 5.1 and Fig. 5.2 show the iron and titanium dissolution over the

temperature range 50 oC to 80 oC from a synthetic ilmenite rotating disc in 7 M HCl

solution at 800 rpm. The extent of iron and titanium dissolution (mol m-2) increased

over time as well as with the increase in temperature. It has been reported that iron

and titanium dissolved in acid solution simultaneously at a stoichiometric ratio of 1:1

(Sinha, 1979; Jackson and Wadsworth, 1976; Imahashi and Takamatsu, 1976;

Tsuchida et al., 1982). However, in the present study the iron dissolved faster than

titanium as shown Fe-Ti correlation in Figure 5.3. For temperature at 50 oC the

slope of iron-titanium correlation is 2.4 indicating that iron dissolution is 2 times

faster than titanium. At higher temperatures (70 oC to 80 oC), the slope of Fe-Ti

correlation is less than 2 indicating that the difference in the extent of dissolution

between iron and titanium become less pronounce. This is due to simultaneous

dissolution and hydrolysis which occur at the same time. Titanium that dissolved in

solution tends to hydrolyse and precipitate as fine particles or form a product in the

pores of reacting interface resulting lower concentration of titanium dissolved as

discussed further in the next section.

142
Chapter 5: Rotating disc dissolution of SI in HCl

0.45
60 °C
0.40

0.35 70 °C

Iron dissolved (mol m-2) 0.30 80 °C

0.25 50 °C

0.20

0.15

0.10

0.05

0.00
0 2 4 6 8 10
Time (hours)

Fig. 5.1. Effect of temperature on iron dissolution from synthetic ilmenite disc in
7 M HCl solution at 800 rpm.

0.45

0.40 60 °C
70 °C
0.35
Titanium dissolved (mol m-2)

80 °C
0.30 50 °C

0.25

0.20

0.15

0.10

0.05

0.00
0 2 4 6 8 10
Time (hours)

Fig. 5.2. Effect of temperature on titanium dissolution from synthetic ilmenite disc in
7 M HCl solution at 800 rpm.

143
Chapter 5: Rotating disc dissolution of SI in HCl

0.45

0.4 50 °C y = 2.430x

0.35 60 °C y = 1.862x

Dissolved Fe (mol m-2) 0.3 70 °C y = 1.693x

0.25 80 °C y = 1.488x

0.2
Slope =
0.15
2.43
0.1

0.05

0
0 0.05 0.1 0.15 0.2 0.25 0.3
Dissolved Ti (mol m-2)

Fig. 5.3. Plot of iron dissolution as a function of titanium dissolution in 7 M HCl at


800 rpm at different temperatures.

The rates of dissolution of synthetic ilmenite in 7 M HCl at 50 oC, 60 oC,

70 oC and 80 oC were obtained from the slopes of Fig. 5.1 and Fig. 5.2 using the

polynomial method described in Appendix A2.2. Fig. 5.4 depicts the Arrhenius plot

of iron and titanium dissolution from synthetic ilmenite, which yields values for the

apparent activation energy, Ea, of 44.1 kJ mol-1 for iron and 75.6 kJ mol-1 for

titanium. The rate data and the activation energies determined from the slopes of the

Arrhenius plots are listed in Table 5.1. The different values of apparent activation

energy indicate that two different surface chemical processes control the rate of

dissolution of iron and titanium from ilmenite in hydrochloric acid solutions. Higher

apparent activation energies of ilmenite leaching in sulfuric acid and hydrochloric

acid have been reported previously and the values are listed in Table 5.2.

144
Chapter 5: Rotating disc dissolution of SI in HCl

McConnel (1978) reported values of 101 ± 8 kJ mol-1 for synthetic ilmenite and

87 ± 7 kJ mol-1 for natural ilmenite dissolution in solution of 7 M sulfuric acid over

the temperature range 65 – 85 oC using the rotating disc method. Jackson and

Wadsworth (1976) also reported the values of 92.5 kJ mol-1 and 88.3 kJ mol-1 for the

dissolution of natural ilmenite and hematite, respectively, in 6 M HCl using

particulate leaching experiments. None of the above studies report the apparent

activation energy for the dissolution of iron and titanium from synthetic ilmenite in

hydrochloric acid.

-8
-9
-10
Ln {rate (mol m-2 s-1)}

-11
-12
-13
-14 Fe y = 44.1x + 4.03
R² = 0.9
-15
Ti y = 75.6x + 14.5
-16
R² = 1.0
-17
-18
-0.4 -0.38 -0.36 -0.34 -0.32
-1000/RT (kJ/mol)
Fig. 5.4. Arrhenius plot for iron and titanium dissolution (T= 50, 60, 70, 80 oC,
[HCl] = 7 M, stirring speed = 800 rpm).

145
Chapter 5: Rotating disc dissolution of SI in HCl

Table 5.1
Rate data and activation energy.

Iron Titanium
o -1000/RT 6
T ( C) dn/dt 10 Ln dn/dt dn/dt 106 Ln dn/dt
(kJ mol-1)
(mol m-2 s-1) -2 -1
(mol m s )
-13.6
50 -0.372 4.8 -12.3 1.2

60 -0.361 5.2 -12.2 2.6 -12.9

70 -0.351 11.7 -11.4 7.8 -11.8

80 -0.341 17.4 -10.9 11.8 -11.3

Ea (kJ mol-1) 44.1 75.6

Table 5.2
Reported and current values of activation energy.

Temp. Ea ( kJ mol-1)
Ilmenite Acid con.
Acid range
type (M) Fe Ti Ilmenite
(oC)
Naturala H2SO4 7 65-85 87.0
Synthetica H2SO4 7 65-85 101.0
Naturalb HCl 6 92.5
Syntheticc HCl 7 60-80 44.1 75.6
a
McConnel (1979); b Jackson and Wadsworth (1976); c current study

5.2.2. Effect of disc rotation speed

The ilmenite leaching process includes the diffusion of reactants (H+, Cl-)

from the bulk solution to the ilmenite surfaces, adsorption at the surface, formation

of appropriate complexes of the iron and titanium with chloride, desorption of the

products such as complex species from the surface and their diffusion into the bulk

solution. If the process of ilmenite dissolution is limited by mass transfer of the

reactants, the reaction rate will be dependent on the disc rotational speed according

to the Levich equation (5.2):

RM = 0.623 D2/3-1/61/2 C = k C (5.2)

146
Chapter 5: Rotating disc dissolution of SI in HCl

where RM (mol cm-2 s-1) is the dissolution rate of iron or titanium, D (cm2 s-1) is the

diffusion coefficient for the reactant responsible for the rate limiting step,  (cm2 s-1)

is the kinematic viscosity of the solution,  (rad s-1) is the angular velocity of the

disc measured in revolutions per minute (rpm) and converted according to,  = rpm

x 2π/60 (s-1), and k can be considered as the apparent rate constant. If the surface

chemical reaction is the rate limiting step, the reaction rate should be independent of

the rotational speed.

Fig. 5.5 shows the effect of stirring speed on dissolution of iron as a function

of time. As noted in Chapter 3, the disc was polished before conducting each

experiment. The figure shows that the dissolution rate of iron from the synthetic

ilmenite disc is independent of the stirring rate. This indicates that the reaction rate

of iron is limited by the chemical reaction at the surface, rather than a mass transport

process. The titanium dissolution from synthetic ilmenite disc under the same

condition however is dependent on rotation speed from 400 rpm to 800 rpm (Fig.

5.6). The rate of titanium dissolution is less affected at rotation speeds above 800

rpm. Thus, the disc rotational speed was set to 800 rpm in subsequent experiments.

147
Chapter 5: Rotating disc dissolution of SI in HCl

0.12
400 rpm 600 rpm
0.10 800 rpm 1000 rpm

Fe dissolved (mol m-2)


0.08

0.06

0.04

0.02

0.00
0 10 20 30 40
Time (103 s)

Fig. 5.5. Extent of dissolution of synthetic ilmenite given by iron concentration as a


function of time at various stirring speeds in 7 M HCl at 60 oC.

0.12
400 rpm 600 rpm
0.10
800 rpm 1000 rpm
Ti dissolved (mol m-2)

0.08

0.06

0.04

0.02

0.00
0 10 20 30 40
Time (103 s)

Fig. 5.6. Extent of dissolution of synthetic ilmenite given by titanium concentration


as a function of time at various stirring speeds in 7 M HCl at 60 oC.

148
Chapter 5: Rotating disc dissolution of SI in HCl

A plot of the rate against the square root of the rotation speed in Fig. 5.7 is

based on the Levich equation. The linear relationship observed in Fig. 5.7 confirms

that the initial rate of titanium dissolution is essentially controlled by the diffusion of

reactant (H+). However, the line does not pass through the origin because of the non-

uniformity of the disc similar to Barton and McConnel (1979) finding as described

earlier in Fig. 2.16 (B) in Chapter 2. Nevertheless, the slope of the line can be used

to calculate a value for the diffusion coefficient for H+ based on the Levich equation

as described later.

Barton and McConnel (1978) indicated that the dissolution of synthetic and

natural ilmenite in the form of a rotating disc in sulfuric acid solutions was

independent of the rotation speed suggesting a chemically controlled surface

reaction. In the dissolution curve, the natural ilmenite disc showed some important

differences in its behaviour compared to synthetic ilmenite. The dissolved iron

concentration was slightly higher than that of titanium in the case of natural ilmenite

compared to synthetic ilmenite, where the dissolution of iron and titanium followed

the stoichiometric ratio of the ilmenite phase.

However, in the present study the results from the experiments conducted in

hydrochloric acid solutions show that the synthetic ilmenite disc resulted in a higher

concentration of dissolved iron compared to titanium, as shown in the plot of

dissolved iron as a function of dissolved titanium (Fig. 5.8). The slopes of the

different lines for different stirring speed decreased from ~ 2.0 at low rotation speeds

to 1.5 at higher rotation speed. Based on this behaviour of titanium, the titanium

dissolution appears to depend on diffusion of H+ through product layer where the

thickness of product layer is reduced at higher rotation speeds. This means that at

149
Chapter 5: Rotating disc dissolution of SI in HCl

higher rotation speeds more H+ ions reach at the reactive surface of ilmenite

facilitating titanium dissolution.

4.00E-06

3.50E-06

3.00E-06
Rate (mol m-2 s -1)

2.50E-06

2.00E-06
y = 0.0000003x + 0.0000006
1.50E-06 R² = 0.991

1.00E-06

5.00E-07

0.00E+00
0 5 10 15
1/2 ( s-1/2)

Fig. 5.7. Dissolution rate of titanium at 60 oC in 7 M HCl as a function of square root


of stirring rates.

0.14
400 rpm y = 1.86x 400 rpm
0.12 600 rpm y = 1.74x slope = 1.86
800 rpm y = 1.54x
Dissolved Fe (mol m-2)

0.1
1000 rpm y = 1.57x
0.08

0.06

0.04

0.02

0
0 0.02 0.04 0.06 0.08
Dissolved Ti (mol m-2)

Fig. 5.8. Plot of iron dissolution as a function of titanium dissolution in 7 M HCl at


60 oC at different stirring rates.

150
Chapter 5: Rotating disc dissolution of SI in HCl

5.2.3. Effect of hydrochloric acid concentration

(a) Dissolution curves

Fig. 5.9 and Fig. 5.10 illustrate some of the leaching curves obtained when

discs of synthetic ilmenite were leached at 60 oC and at 800 rpm in solutions of

different HCl concentrations. The dissolution of iron at 4 M HCl slowly increased as

leaching time proceeded for 8 hours as presented in Fig. 5.9. At higher acid

concentration (7 M and 9 M) iron was rapidly dissolved during the first hour of

leaching and continued until 8 hours. With the increase in HCl concentration from 4

M to 9 M a higher dissolution of iron was obtained. The same trend was also

observed for the titanium dissolution.

0.40
4 mol/L
0.35
7 mol/L
Iron dissolved x 10-5 (mol m-2)

0.30 9 mol/L
0.25

0.20

0.15

0.10

0.05

0.00
0 2 4 6 8 10
Time (hours)

Fig. 5.9. Effect of HCl concentration on iron dissolution at 60 oC and 800 rpm.

151
Chapter 5: Rotating disc dissolution of SI in HCl

0.40
4 mol/L
0.35
7 mol/L
Ti dissolved (mol m-2) 0.30 9 mol/L

0.25

0.20

0.15

0.10

0.05

0.00
0 2 4 6 8 10
Time (hours)

Fig. 5.10. Effect of HCl concentration on titanium dissolution at 60 oC and 800 rpm.

(b) Dissolution correlations

Fig. 5.11 shows the correlation of dissolved iron as a function of dissolved

titanium at different time intervals and acid concentrations.. At the low acid

concentration of 4 M HCl, the slope is 1.8 which means that iron dissolution was 1.8

times higher than that of titanium. At higher HCl concentration (7 M and 9 M) the

slope decreases to 1.2 reaching closer to the stoichiometric ratio of ilmenite phase of

1:1. Titanium(IV) in highly concentrated HCl forms complex species TiOCl 42- as

shown in the titanium speciation diagram (Fig. 5.12, based on the thermodynamic

data in Appendix A.1). As the leaching proceeds Ti(IV) will start to hydrolyse due to

the increase in concentration of dissolved Ti(IV) species. This will lead to the

production of white titanium oxides which precipitate as fine particles in the solution

or at mineral surface (Jackson and Wadsworth, 1976). In the case of a rapid leach

152
Chapter 5: Rotating disc dissolution of SI in HCl

reaction, titanium oxides can precipitate inside the pores of the particles. Therefore,

it is observed that iron leached faster than titanium at low values of temperature,

stirring rate and HCl concentration. The Eh-pH diagram in Fig. 5.13 also shows that

Ti(IV) dissolves in HCl as TiOCl42- only under oxidizing conditions and in a very

limited range of pH (up to ~ 1.5). Thus, the decrease in acid concentration inside

pores or at the surface can cause high pH leading to precipitation of Ti(IV) as

TiO2.H2O.

0.4

0.35 HCl = 4 M y = 1.9x


HCl = 7 M y = 1.3x
0.3
HCl = 9 M y = 1.2x
Dissolved Fe ( mol m-2)

0.25

0.2

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4

Dissolved Ti ( mol m-2)

Fig. 5.11. Plot of iron dissolution as a function of titanium dissolution at 60 oC , 800


rpm at different acid concentration.

153
Chapter 5: Rotating disc dissolution of SI in HCl

Fig. 5.12. Speciation diagram of Ti(IV) (total 0.001 mol/L) species in chloride media
based on HSC database version 7.1 (Roine, 2011).

Fig. 5.13. Eh-pH diagram of Ti-Cl-H2O at 25 oC (Vaughan and Alfantanzi, 2004).

154
Chapter 5: Rotating disc dissolution of SI in HCl

(c) Change of surface before and after dissolution

The changes of surface before and after dissolution can be observed on the

SEM images of synthetic ilmenite disc surface in 7 M at 80 oC. Figure 5.14 (A) and

(B) shows the surface of synthetic ilmenite disc before and after leaching. It can be

observed that the surface roughness of synthetic ilmenite has increased after

leaching, compared to the surface before leaching. The cracks on the surface of the

disc are due to the acid attack dissolving the iron matrix leaving behind the unreacted

surfaces. The EDX analysis in Figure 5.14 (B) and (D) was performed on the bulk

surface of synthetic ilmenite disc before and after leaching. The counts of titanium

are incredibly higher in (D) after leaching, than it was in (B), before leaching. It is

possible that iron was preferentially removed from the ilmenite surface, leaving

behind unreacted TiO2, which may dissolve in acid in secondary step. It is also

possible that a TiO2 layer was formed (precipitated from dissolved Ti(IV)) on the

surface of ilmenite which might dissolve slowly than ilmenite. Another possibility is

that both iron and titanium dissolved into solution simultaneously but at different

rates as controlled by the rate of chemical reactions. Again, the dissolved titanium in

the form of Ti(IV) may hydrolyse and precipitate as fine particles in solution or at

the surface of ilmenite (van-Dyk et al., 2002).

Thus, the SEM analysis on the same surface at higher magnification (500x),

showed that a precipitate was formed on the surface of the synthetic ilmenite disc

(Fig. 5.15) and EDX analysis on the white precipitate indicated the composition:

71.63% Cl, 13.31% O and 15.06% Ti (Table 5.3). Based on the atomic ratio

(Ti1.1O1.0Cl5.4) the white precipitate could be a hydrolysed product of the titanium

chloro-complex (see Fig. 5.13). Future studies need to be directed for proper

identification of the solution and solid species of Ti(IV).

155
Chapter 5: Rotating disc dissolution of SI in HCl

B
A

C D

Fig. 5.14 SEM images of synthetic ilmenite disc surface before dissolution (A) and
after dissolution in 7 M HCl at 80 oC (B) with EDX spectrum analysis before (C)
and after (D) dissolution.

Fig. 5.15. SEM images of synthetic ilmenite disc surface after dissolution in 7 M
HCl at 80 oC and EDX analysis of white precipitate (marked by white arrow).

156
Chapter 5: Rotating disc dissolution of SI in HCl

Table 5.3
EDX analysis of white precipitate in Fig. 5.15

Element % (w/w) % (atom)


Ti 20.77 15.06
O 6.13 13.31
Cl 73.10 71.63

(d) Kinetic studies

The curves for the iron and titanium dissolution as a function of time can be

used to determine initial rates as well as to quantitatively analyse rate data based on

the kinetic models. For instance, the initial rates (RM) can be used to determine the

reaction orders (n) by relating to the concentration of a reactant Y ([Y]) according to

Eqs. 5.3 and 5.4, where [M] is the concentration of dissolved metal ions in solutions,

X is the dissolved fraction of metal ions after time tand kap is the apparent rate

constant.

𝑑[𝑀] 𝑑𝑋
𝑅M = or = kap[Y]n (5.3)
𝑑𝑡 𝑑𝑡

log RM = n log[Y] + log kap (5.4)

Fig. 5.16 (a) and (b) show logarithmic plots of RFe and RTi as a function of

acid concentration [H+] based on the initial rate of synthetic ilmenite disc dissolution

from Fig. 5.9 and Fig. 5.10. Rates obtained from the slopes of linear regions in Fig.

5.9 and Fig. 5.10 are listed in Table 5.4. The lines of best fit in Fig. 5.16 (a) and (b)

correspond to slopes of 1.4 and 2.5, representing the reaction orders relative to [HCl]

for the dissolution of iron and titanium, respectively. The initial iron dissolution

from synthetic ilmenite disc shows approximately first order kinetics with respect to

H+ concentration, indicating the direct involvement of H+ in the surface reaction. The

initial dissolution of titanium in HCl solution shows a much stronger dependence on

157
Chapter 5: Rotating disc dissolution of SI in HCl

acid concentration compared to iron. McConnel (1978) reported a first order reaction

with respect to H+ concentration for initial dissolution of titanium from a synthetic

ilmenite disc in sulfuric acid solution. First order reaction with respect to H+ was

also observed for the dissolution of natural ilmenite in 3 M to 6 M hydrochloric acid

(Jackson and Wadsworth, 1976). Further analysis of reaction orders with respect to

H+ needs reliable information on ionic activities of H+ and Cl- at the temperatures in

question.

Table 5.4
Effect of HCl concentration on initial rates.

Rate (Iron) Rate (Titanium)


[H+] +
log [H ] dM/dt105 Log dM/dt dM/dt105 Log dM/dt
(mol m-3)
(mol m-2 s-1) (mol m-2 s-1)
4000 3.602 0.57 -5.24 0.27 -5.57
7000 3.845 1.32 -4.88 0.93 -5.03
9000 3.954 1.86 -4.73 2.19 -4.66

-4.7 -2
-2.5
-4.8
Log {RFe (mol m-2 s-1)}

Log {RTi (mol m-2 s-1)}

-3
-4.9 y = 1.4x - 10.25
R² = 0.99 -3.5
-5 -4 y = 2.5x - 14.66
R² = 0.99
-4.5
-5.1
-5
-5.2
-5.5
-5.3 -6
3.4 3.6 3.8 4 3.5 3.6 3.7 3.8 3.9 4
Log {[H+] (mol m-3)} Log {[H+] (mol m-3)}

Fig. 5.16. Logarithmic plot of initial rate of iron and titanium dissolution as a
function of [H+] at 60 oC and 800 rpm.

158
Chapter 5: Rotating disc dissolution of SI in HCl

5.2.4. Parabolic rate law

A leaching system which leaves a solid residue and of which the rate

determining step is diffusion of the reagent through a solid layer obeys the Parabolic

Rate Law, as described by Ray (1993) and Barton and McConnel (1979). If a flat

plate of area A and thickness d is considered, assuming that only one face of the

plate reacts, the area remains unchanged and only d diminishes. The reaction

produces a product of thickness y which increases gradually. If diffusion of reactant

through this product layer is rate controlling, then the rate of reaction is given by the

rate of diffusion of the reagent. Hence:

𝑑𝑊 𝐷𝐴 .𝐶
Rate = 𝑑𝑡
= 𝑦
(mol s-1) (5.5)

where W is the moles reacted after time t, A is the area (m2), D is the diffusion

coefficient (m2 s-1) and C (mol m-3) is the bulk reagent concentration (which can be

considered to remain constant, as it is generally maintained in excess).

Since, W = A y ρ / M (ρ is density of the product, M is the molar mass of the

product), dW/dt = (A ρ / M) dy/dt which leads to:

𝑑𝑦 𝑘′
= or y. dy = 𝑘′ dt (5.6)
𝑑𝑡 𝑦
DMC
Where k' = = constant;
ρ

Since y = 0 at t = 0, by integration

y2 = k t (5.7)
where k is a constant (= 2k’).

159
Chapter 5: Rotating disc dissolution of SI in HCl

Eq.5.7 is one of the general forms of the parabolic law which takes the form

of the general equation for parabola y2 = mx (Ray, 1993; Barton and McConnel,

1979). For a general stoichiometric reaction between iron in ilmenite and H+ in HCl

solution, Fe (in ilmenite) + b H+ = products, where the rate of dissolution of Fe and

consumption of H+ are related by equations RFe = (1/b)RH+, it can be shown that

(nFe/A)2 = (2 DH+ CH+ρ / M b).t, where nFe/A is the dissolved iron per unit area (mol

m-2), C is the concentration of H+(mol m-3), ρ is the density which may be taken

comparable to that of ilmenite, M is the molar mass of the product layer which may

be assumed to be the same as that of ilmenite.

Plots of the square of both iron and titanium dissolved (mol m-2) against time

are linear for the time periods from 180 to 420 minutes of dissolution, confirming the

applicability of the parabolic rate law (Fig. 5.17). This relationship reveals that the

rate controlling step is the diffusion of reagent through an insoluble product layer

that is formed on the surface of synthetic ilmenite. Thus, the diffusion coefficient can

be calculated from the parabolic law using slope(s) of linear relationships in Fig.

5.17. Based on the slopes of Fig. 5.17 for iron dissolution in 4 M HCl and b = 4 (see

Eq. 5.1), the calculated value of DH+ is 1.3 x 10-10 cm2 s-1. Likewise, the calculated

value of DH+ for titanium dissolution is 0.3 x 10-10 cm2 s-1. The order of these values

in the range of 3 x 10-11 to 13 x 10-11 cm2 s-1 are at the high end of the range of 10-11-

10-15 cm2 s-1 for previously reported values of DH+ relevant to the dissolution of

different minerals based on the diffusion through product layers (Georgiou and

Papangelakis, 1998; Senanayake and Das, 2004; Senanayake et al., 2015) indicating

the different porosities of such layers.

160
Chapter 5: Rotating disc dissolution of SI in HCl

0.120
HCl=4 mol/L y = 0.0000053x - 0.0271
0.100 R² = 0.99
HCl=7 mol/L

0.080 HCl=9 mol/L


y2 (Fe) (mol m-2)2

0.060
y = 0.0000017x - 0.0075
R² = 0.99
0.040

0.020 y = 0.0000004x - 0.0017


R² = 0.99

0.000
0 5000 10000 15000 20000 25000 30000
Time (s)

0.08
HCl=4 mol/L y = 0.0000033x - 0.0102
0.07
R² = 0.99
HCl=7 mol/L
0.06
HCl=9 mol/L
y2 (Ti) (mol m-2)2

0.05

0.04
y = 0.0000010x - 0.004
0.03 R² = 0.99

0.02

0.01 y = 0.0000001x - 0.0006


R² = 0.99
0.00
0 5000 10000 15000 20000 25000 30000
Time (s)

Fig. 5.17. Change in square of dissolved amount of Fe and Ti from synthetic ilmenite
disc at 60 oC and 800 rpm.

161
Chapter 5: Rotating disc dissolution of SI in HCl

5.3. Dissolution of ilmenite under reducing condition

At lower potentials in acidic solutions Ti(IV) can be reduced to Ti(III) which

precipitates as Ti(OH)3 at higher pH (Fig. 5.13). This implies that titanium leaching

in HCl can be affected by reducing conditions. In order to examine the effect of a

reducing agent on the rate of ilmenite dissolution, an experiment was carried out

with the addition of SnCl2 as a reducing agent. The effect of SnCl2 on the rate of iron

and titanium dissolution from synthetic ilmenite disc is presented in Fig. 5.18 and

Fig. 5.19.

Higher concentrations of SnCl2 accelerated the dissolution of iron and

titanium causing rates to increase by a factor of 4 to 10 compared to those in 7 M

HCl alone. The dissolution rates of iron and titanium also showed significant

increment with the increase of SnCl2 from 0.5 g/L to 2.0 g/L. Moreover, the iron and

titanium leaching in the presence of SnCl2 was found to be linear over 360 minutes,

compared to the parabolic behaviour observed in the absence of SnCl2 (Fig. 5.18-

5.19). Therefore, a logarithmic plot of RFe and RTi against log [Sn(II)] based on the

initial rates of synthetic ilmenite disc dissolution from Figs. 5.15 – 5.16 is shown in

Fig. 5.20. The dissolution rate of iron and titanium from ilmenite in HCl showed

reaction orders of 0.3 (Fe) and 0.4 (Ti). The mixed potential model for

electrochemical heterogeneous reactions suggests half order reaction kinetics (Nicol

and Lazaro, 2002). Further electrochemical studies are essential to rationalize the

low reaction order of 0.3 with respect to Sn(II) concentration observed in the present

study.

162
Chapter 5: Rotating disc dissolution of SI in HCl

1.40
Tin(II) chloride = 0.0 g/L
1.20
Tin(II) chloride = 0.5 g/L

Fe dissolved (mol m-2)


1.00 Tin(II) chloride = 1.0 g/L
Tin(II) chloride = 2.0 g/L
0.80

0.60

0.40

0.20

0.00
0 5 10 15 20 25
3
Time (10 s)

Fig. 5.18. Effect of tin(II) chloride on the rate of iron dissolution from synthetic
ilmenite disc in HCl (7 M) solution at 60 oC and 800 rpm.

0.90
Tin(II) chloride = 0.0 g/L
0.80
Tin(II) chloride = 0.5 g/L
0.70
Ti dissolved (mol m-2)

Tin(II) chloride = 1.0 g/L


0.60
Tin(II) chloride = 2.0 g/L
0.50
0.40
0.30
0.20
0.10
0.00
0 5 10 15 20 25
3
Time (10 s)

Fig. 5.19. Effect of tin(II) chloride on the rate of titanium dissolution from synthetic
ilmenite disc in HCl (7 M) solution at 60 oC and 800 rpm.

163
Chapter 5: Rotating disc dissolution of SI in HCl

-4

Iron y = 0.33x - 4.62


-4.2

Log {Rate (Ti) (mol m-2 s-1)}


R² = 0.99
Titanium

-4.4

-4.6
y = 0.42x - 4.89
R² = 0.98
-4.8

-5
0 0.5 1 1.5
Log {[Sn(II)] (mol m-3)}

Fig. 5.20. Logarithmic plot of initial rate of iron and titanium dissolution as a
function of Sn(II) concentration in 7 M HCl solution at 60 oC and 800 rpm.

5.4. Proposed reaction mechanism

The soluble species present in HCl solution could be identified by measuring

the Eh-pH values of the solution and comparing with the established Eh-pH

diagrams to shed more light on the reaction mechanism. The two electrodes

(platinum and reference (Ag/AgCl) electrode) were immersed in the solution and the

potential was recorded using LabViewTM data acquisition system. The values

obtained in this study as initial and final potentials are listed in Table 5.5. The

solution potential in 7 M HCl in the absence of SnCl2 (0.20-0.26 V) was in the region

of TiO2+ and Fe2+ (Fig. 5.21) and the potential value did not change significantly at

the completion of the experiment. This result is consistent with the dissolution of

ilmenite according to Eq.5.8.

164
Chapter 5: Rotating disc dissolution of SI in HCl

FeTiO3+ 4H+→ Fe2+ + TiO2+ + 2H2O (5.8)

Table 5.5
Potential data for synthetic ilmenite disc dissolution in 7 M hydrochloric acid at
60 oC.

SnCl2 Initial potential Final potential


(g/L) (V /SHE) (V/SHE)
0 0.20 0.26
0.5 0.23 -0.72
2.0 0.21 -0.35
4.0 -0.44 -0.84

Potentials of different redox couples at pH≤ 0: Fe3+/Fe2+ = 0.8 V (Fig. 5.21),


TiO2+/Ti3+ = 0.1-0.3 V (Fig. 5.21), TiOCl42-/Ti(OH)2+= 0-0.1 V (Fig. 5.13), SnCl62-
/SnCl42- = 0.15 (Fig. 5.22), H2/H+ = 0.

However, Fig. 5.12 and Fig. 5.13 show that the stable form of TiO2+ in

concentrated chloride solutions is TiOCl42-. Moreover the final potential of 0.26 V

(SHE) in Table 5.5, in the absence of SnCl2, is also in the region of stability of

TiOCl42- in Fig. 5.13. Mahmoud et al. (2004) reported that the addition of a reducing

agent during ilmenite leaching causes the reduction of Ti(IV) to Ti(III), which was

the main role of iron which accelerated the leaching of ilmenite, as described in

Chapter 2.

The measured Eh decreased in the presence of SnCl2 and the final potential

value in Table 5.5 (<-0.35 V) indicated that the dominant titanium species in solution

was Ti(III) (Fig. 5.21). This indicated that the reduction of TiO2 from ilmenite to

Ti(III) was caused by Sn(II). It can be assumed that if Sn(II) is directly involved in

the redox reaction, it will be oxidised to Sn(IV).

165
Chapter 5: Rotating disc dissolution of SI in HCl

Fig. 5.21. Eh-pH diagrams for Fe-Ti-H2O system at 25 oC, (excluding Ti2+ ions)
considering TiO2 (c,hydrated), Ti2O3 (c, hydrated) and Fe2O3 as the solid titanium
oxide and Fe(III) phases, and for 0.1 and 0.01 dissolved iron and titanium activities,
respectively, solid lines and dashed lines show iron and titanium species,
respectively (Kelsall and Robbins, 1990).

Although the Eh-pH diagram in Fig. 5.22 indicates the existence of the

chloro-species of Sn(II) and Sn(IV) the measured potential in Table 5.5 appears to be

lower than standard potential and warrants further studies on speciation. The

reductive role of Sn(II) could be the reduction of Fe(III) to Fe(II) and acting as a

reducing agent to convert TiO2 or the precipitated Ti(IV) to Ti(III).

166
Chapter 5: Rotating disc dissolution of SI in HCl

Fig. 5.22. Eh-pH diagram for the Sn/H2O-Cl system at 298 K, at a dissolved tin
activity = 10-3, aCl = 5, p (SnH4) = 1 (House and Kelsall, 1984).

In the absence of SnCl2, the dissolution follows the reactions in Eq. 5.9.

However, acid hydrolysis of the freshly formed TiO2 according to Eq. 5.10 can

produce blockage and retard the reaction, as evident from the applicability of the

parabolic rate law. In the presence of SnCl2, TiOCl2 is reduced to TiCl3 which

prevents surface blockage by hydrolysed Ti(IV) products and facilitates the reaction

according to Eq. 5.11. Further studies are essential to support these reaction

mechanisms.

In the absence of SnCl2

FeO.TiO2 + 2HCl → FeCl2 + TiO2 + H2O (5.9)

TiO2 + 2HCl → TiOCl2.H2O (hydrolysed Ti(IV) product) (5.10)

In the presence of SnCl2

2TiOCl2 + SnCl2 + 4HCl → 2TiCl3 + SnCl4 + 2H2O (5.11)

167
Chapter 5: Rotating disc dissolution of SI in HCl

The SEM image (Fig. 5.15) of the ilmenite following its dissolution in HCl

alone showed signs of titanium hydrolysis due to increase in Ti(IV) concentration

where the white precipitate was observed. The SEM image of the synthetic ilmenite

disc after leaching in the presence of Sn(II), shown in Fig. 5.23, indicated that the

surface of synthetic ilmenite was altered to a very porous surface resulting from

dissolution of Ti(IV) and Fe(II) from its surface. A few spot of white precipitate was

seen on the surface of synthetic ilmenite. Fig. 5.24 presented EDX analysis of bulk

surface of synthetic ilmenite disc after leaching with Sn(II) and specifically on white

materials marked by white arrow. Table 5.6 presents the elements that can be found

on EDX analysis of synthetic ilmenite surface after leaching with the reducing agent.

The EDX analysis of bulk surface indicated high titanium and iron but low in Cl, Sn

and O (Fig. 2.24.(A) and Table 5.6). The white precipitate spot on the surface of the

disc was identified to be consisting of from Cl, Fe and Sn (Fig. 2.24.(B) and Table

5.6). Therefore, in the presence of Sn(II) the white precipitate that represent

precipitation product of TiO2 was not formed due to the reduction of TiOCl2 to TiCl3

(Figs. 5.12).

168
Chapter 5: Rotating disc dissolution of SI in HCl

Fig. 5.23. SEM images of synthetic ilmenite disc surface after dissolution in 7 M
HCl in the presence of 2.0 g/L SnCl2 at 60 oC.

Fig. 5.24. (A)-(B) EDX spectrum and SEM analysis of bulk synthetic ilmenite disc
surface and (C)-(D) EDX spectrum and SEM analysis of white spot mark by red
arrow (Full scale of EDX spectrum are presented in Appendix A2.3).

169
Chapter 5: Rotating disc dissolution of SI in HCl

Table 5.6
EDX analysis on bulk surface of SEM (Fig. 5.24. (A)) and white spot mark by red
arrow (Fig. 5.24. (B)).
Bulk surface White spot
Elements
% (w/w) % (atom) % (w/w) % (atom)
O 44.30 71.57
Ti 26.26 14.17
Cl 3.03 2.21 32.03 51.20
Fe 25.73 11.91 30.43 30.88
Sn 0.67 0.15 37.54 17.92

The reaction mechanisms of synthetic ilmenite disc in hydrochloric acid can

be proposed with a few assumptions based on leaching, solution potential, kinetics

model and SEM/EDX analysis. The results presented in the previous section suggest

that initially, as predicted by the chemical reaction equation, when ilmenite comes

into contact with the hydrochloric acid leach solution, both iron and titanium

dissolve into solution. At this point, the rate is controlled by the rate of the chemical

reaction. The next step in the chain of events is determined by the titanium species

formed in solution. The Ti(IV) ions hydrolyse in hydrochloric acid solutions. The

Ti(IV) concentration increases as the leach reaction progresses. Once the Ti(IV)

concentration in solution exceeds the value of 10-3 M, polymerisation will occur.

The next critical event that influences the rate of the reaction is the eventual

formation of the solid species TiOCl2 or its hydrolyze product as the Ti(IV)

concentration in solution increases. The TiOCl2 can precipitate in the pores of the

leached particle or as fines in the leach solution. When the precipitation occurs on

the surface of ilmenite a product layer is formed. The rate-determining step in the

leach reaction now changes to diffusion of the reacting species through the product

layer. Any additives that are added to the leach solution can influence the titanium

species in solution and solid phase and consequently the rate-determining step. By

170
Chapter 5: Rotating disc dissolution of SI in HCl

stabilising soluble species of iron and titanium, the rate of dissolution can be

maintained. The final event that can influence the leach reaction is a significant

reduction in the hydrogen ion concentration. The decrease in acid concentration at

the surface and the resultant rise in pH can lead to the formation of a TiO2.H2O

precipitate in the existing product layer. It will also result in a change in the rate-

limiting step. The diffusion of hydrogen ions to the reaction interface will now

become the rate-limiting step.

171
Chapter 5: Rotating disc dissolution of SI in HCl

5.5. Summary and conclusions

Under non reducing conditions, the effect of rotation speed, temperature and

HCl concentration on the dissolution of iron and titanium from a rotating disc of

synthetic ilmenite was studied. It was found that the rate of iron dissolution was

independent of the stirring speed. However, the rate of dissolution of titanium was

found to be marginally dependent on stirring speed, indicating a transport controlled

reaction. The effect of temperature on iron and titanium dissolution indicated high

activation energies of 44.1 kJ mol-1 for Fe and 75.6 kJ mol-1 for Ti in the temperature

range of 60 oC to 80 oC. The acid concentrations also had a significant effect on

initial rates of synthetic ilmenite dissolution.

The reductive dissolution of a rotating disc of synthetic ilmenite in HCl with

SnCl2 showed an improvement compared with dissolution in HCl solution alone.

The dissolution rate of iron and titanium was accelerated by >10 times. The

dissolution of iron and titanium was also found to follow linear relationships with

respect to time. This indicated that in the absence of reducing conditions, the

dissolution behaviour of titanium was sensitive to the solution potential and the

hydrolysis and precipitation of titanium in solution, pores or on the surface which

could inhibit further dissolution. The formation of TiO2 layer on the surface of the

disc is a contributing factor to the limiting rate of diffusion of H+ through product

layer of Ti(IV). The addition of SnCl2 enhanced the dissolution by reducing Ti(IV) to

Ti(III) and thus avoiding the precipitation of surface/pore blocking Ti(IV)

oxides/hydroxide.

172
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

CHAPTER 6

6. ELECTROCHEMICAL STUDY OF ILMENITE

DISSOLUTION IN HYDROCHLORIC ACID

6.1. Introduction

McConnel (1978) conducted preliminary studies on the electrochemical

behavior of synthetic and natural ilmenite in solutions of various concentrations of

sulfuric acid at the two temperatures of 20 oC and 40 oC using cyclic voltammetric

and potentiostatic techniques. Later, Adriamanana et al. (1984) studied the

electrochemical behavior of several natural ilmenite samples in 1 M H2SO4 using

cyclic voltammetry of carbon paste electrode. They found that the oxidation state of

metal ions in FeTiO3 are Fe(II) and Ti(IV) and that small amounts of Fe(III) and

excess Ti(IV) can occur in natural ilmenite samples. The electrochemical behavior of

ilmenite in H2SO4 was studied by Zhang and Nicol (2009) using open circuit

potential (OCP), cyclic voltammetry (CV) and coulometry (CM). The dissolution of

ilmenite in acid takes place according to Eq. 6.1. The dissolution rate of ilmenite

significantly increases at potentials lower than the rest potential, where the

dissolution of ilmenite/hematite takes place according to reactions 6.2 -6.3 (Zhang

and Nicol, 2009).

173
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

At more negative potentials the dissolution rate of ilmenite increases due to

the reduction of Ti(IV) to Ti(III) (Eq. 6.3) and subsequent acid dissolution, as

described in Chapter 2.

FeTiO3 + 4H+ → Fe2+ + TiO2+ + 2H2O (6.1)

Fe2O3 + 6H+ + 2e- → 2Fe2+ + 3H2O (6.2)

2FeTiO3 + 6H+ + 2e- → 2Fe2+ + Ti2O3 + 3H2O (6.3)

Fe2+ → Fe3+ + e- (6.4)

The higher number of studies for ilmenite dissolution in hydrochloric acid solution

compared to those in sulfate systems reflects the importance of understanding the

leaching behavior of ilmenite in acidic-chloride systems. Although there have been

numerous recent leaching studies in this field, the electrochemical behavior of

ilmenite is not well established. This chapter describes an electrochemical study of

ilmenite dissolution in hydrochloric acid solution at concentrations and temperatures

relevant to leaching. The electrochemical measurements are based on spinning or

stationary electrodes of massive ilmenite, carbon paste-ilmenite and platinum in HCl

solutions at 50-80 oC. Some measurements were also made in H2SO4 solutions in

order to compare them with the literature data. All of the experimental techniques

have been described in Chapter 3.

174
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The open circuit potentials, cyclic voltammetry and coulometry measurements

were used to collect data which can shed more light on the acid dissolution and

reductive dissolution of ilmenite samples of different origin and composition. The

effect of scan rate and rotation speed on current density was used to determine the

diffusion coefficients of hydrogen ions as well as of reducing agents such as Fe(II)

and Sn(II).

6.2. Open circuit potential

6.2.1. Comparison of massive ilmenite and platinum electrodes

The potential attained by a massive ilmenite (MI) electrode in hydrochloric

acid solution (7 M) consists of a mixed potential because of the acid dissolution of

ilmenite/hematite producing both Fe(II) and Fe(III), as described in Eqs. 6.1-6.4.

Likewise, the potential of a platinum electrode, which measures the potential due to

the Fe(III)/Fe(II) redox couple (Eq. 6.5) becomes a mixed potential in the presence of

Sn(II), arising from the two redox couples Fe(III)/Fe(II) and Sn(IV)/Sn(II) in Eq. 6.6.

Moreover, the Fe(III) and Fe(II) ions added to the solution and/or produced by the

dissolution of ilmenite/hematite can also affect the concentration ratio of

Fe(II)/Fe(III) and hence the potential of the Pt electrode.

RT [Fe II ]
EFe(III)/Fe II = Eo – Ln (6.5)
F [Fe III ]

2Fe(III) + Sn(II) → 2Fe(II) + Sn(IV) (6.6)

175
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Therefore, open circuit potential measurements carried out by recording the

potential of both ilmenite and platinum electrodes in HCl solution can be used to

understand the role of reducing agents during the acid leaching of ilmenite in HCl

solution. The MI electrode used in the electrochemical studies was constructed from

a massive ilmenite sample obtained from a geological specimen collection in

Murdoch University. The analytical and mineralogical results reported by Zhang and

Nicol (2009) showed that the sample was ilmenite with interspersed titaniferous

hematite.

Figs. 6.1-6.7 show the potential of massive ilmenite and platinum electrodes

at 50 oC in 7 M hydrochloric acid at an electrode rotation speed of 200 rpm with

nitrogen gas sparging throughout the duration of experiment. A comparison of the

measured steady OCP values is made in Tables 6.1. A comparison of OCP values in

H2SO4 and in HCl (this work) is made in Fig.6.8.

700

650

600
Potential (mV)

550

500

450 MI Pt

400
0 10 20 30 40
Time (minutes)

Fig. 6.1. Open circuit potential of rotating massive ilmenite (MI) and platinum
electrodes in 7 M HCl solutions at 50 oC, rotated at 200 rpm.

176
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

650

600

550
Potential (mV)

500

450

400 MI Pt

350
0 10 20 30 40
Time (minutes)

Fig. 6.2. Open circuit potential of rotating massive ilmenite (MI) and platinum
electrodes in 7 M HCl in the presence of 10 g/L Fe(II) and 10 g/L Fe(III) at 50 oC,
rotated at 200 rpm.

700

600

500
Potential (mV)

400

300

200 MI Pt

100
0 10 20 30 40
Time (minutes)

Fig. 6.3. Open circuit potential of rotating massive ilmenite (MI) and platinum
electrodes in 7 M HCl solutions at 50 oC in the presence of 20 g/L Fe(II) rotated at
200 rpm.

177
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

900

800

700

600
Potential (mV)

500

400

300

200 MI Pt
100
0 10 20 30 40
Time (minutes)

Fig. 6.4. Open circuit potential of rotating massive ilmenite (MI) and platinum
electrodes in 7 M HCl solutions at 50 oC in the presence of 20 g/L Fe(III) rotated at
200 rpm.

600

500

400
Potential (mV)

300

200

100
MI Pt
0
0 10 20 30 40
Time (minutes)

Fig. 6.5. Open circuit potential of rotating massive ilmenite (MI) and platinum
electrodes in 7 M HCl solutions at 50 oC in the presence of 20 g/L Sn(II) solutions,
rotated at 200 rpm.

178
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.6. Open circuit potentials of rotating massive ilmenite (MI) electrode in 7 M
HCl solution containing various concentration of Fe(II) and Fe(III) ions at 50 °C.

Fig. 6.7. Open circuit potentials of platinum electrode in 7 M HCl solution


containing various concentration of Fe(II) and Fe(III) ions at 50 °C.

179
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Table 6.1
Summary of OCP of different electrodes under different conditions.

T Acid Metal ion (g/L) OCP (mV)


o RPM Acid Ref
( C) (g/L or M) Fe(III) Fe(II) Pt MI
60 200 H2SO4 50 g/L 0 0 900 690 (A)
60 200 H2SO4 50 g/L 20 0 690 920 (A)
60 200 H2SO4 50 g/L 0 20 600 640 (A)
60 200 H2SO4 50 g/L 10 10 720 720 (A)
50 200 HCl 7M 0 0 610 642 (B)
50 200 HCl 7M 20 0 790 755 (B)
50 200 HCl 7M 0 10 510 561 (B)
50 200 HCl 7M 0 20 580 415 (B)
50 200 HCl 7M 10 5 586 587 (B)
50 200 HCl 7M 10 10 610 610 (B)
50 200 HCl 7M 10 20 621 626 (B)
50 200 HCl 7M 5 10 585 592 (B)
50 200 HCl 7M 20 10 625 627 (B)
50 200 HCl 1M 0 0 - 660 (B)
50 200 HCl 4M 0 0 - 642 (B)
50 200 HCl 7M Sn(II) (20) 520 160 (B)

(A) Literature data from Zhang and Nicol (2009); (B) This work; Pt : Platinum electrode,
MI : Massive ilmenite electrode; Ref : References.

From Fig. 6.1, it is apparent that the open circuit potential of the MI electrode

(~ 642 mV) in HCl without other additives such as Fe(II), Fe(III) or Sn(II) is higher

than that of the Pt electrode (~ 610 mV). Further investigation was conducted by

measuring the mixed potential of MI and Pt electrodes in HCl solution in the

presence of Fe(II). The mixed potential of the Pt electrode in Fig. 6.3 is steady at 580

mV and lower than the value of 610 mV in Fig. 6.1 and 6.2. For MI electrode, the

mixed potential gradually decreased with time and reached a steady state value of

415 mV (Fig. 6.3).

180
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

This potential was lower than the potential of MI in an equimolar solution of

Fe(III)/and Fe(II) in Fig. 6.2. The lower mixed potential in the presence of Fe(II)

may correspond to an increase in the dissolution rate of ilmenite/hematite. The mixed

potential value of MI is also more negative than that of the Pt electrode in this

solution (Fig. 6.2) which could be indicating rapid dissolution of Fe(II) from the

mineral.

The open circuit potential of the MI and Pt electrodes was also measured in

HCl solution containing only Fe(III). As shown in Fig. 6.4, the potential of the Pt

electrode was higher than that in the other solutions mentioned previously. This

indicates the presence of more Fe(III) ions in solution near the surface of the Pt

electrode. However, the initial potential of the MI electrode in the same solution was

647 mV (Fig. 6.4), which was the same as the potential of MI in HCl alone (Fig. 6.1).

The potential gradually increased over time until it reached a steady state value of

755 mV. This change in potential indicated the dissolution process which occurred at

the mineral surface where Fe(III) and Fe(II) dissolved from the ilmenite/hematite.

The effect of the reducing agent SnCl2 in HCl solution on the OCP values is

shown in Fig. 6.5 and Table 6.1. The potential of platinum was slightly lower in the

presence of Sn(II) than in the HCl solution in the absence of Sn(II). The mixed

potential of the MI electrode also decreased due to the presence of Sn(II), compared

to that in HCl solution alone. The decrease in potential of MI in Fig. 6.3 and of the Pt

electrode in Fig. 6.5 indicates the reducing roles of Fe(II) and Sn(II), respectively.

The decrease in potentials was presumably caused by increased rates of surface

reduction due to the presence of these reagents as described in Chapter 5.

181
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Possible reactions of ilmenite in the absence or presence of Sn(II) in HCl

solutions have been discussed in Chapter 5. Such reactions will be considered again

later in this chapter and are given by Eqs. 6.15-6.20. The dissolution of hematite in

the absence or presence of Sn(II) is described by Eq. 6.18 and Eq. 6.21.

6.2.2. Comparison of OCP in HCl and H2SO4

The open circuit potential values obtained from the experiments with HCl

conducted in this study and data with H2SO4 published by Zhang and Nicol (2009)

are compared in Fig. 6.8(a) and (b). Zhang and Nicol (2009) reported that the

potential of ilmenite was likely determined by that of the Fe(III)/Fe(II) couple at the

mineral surface in sulfuric acid. The open circuit potential of the MI electrode in HCl

was lower compared to that in H2SO4 which suggested that HCl promoted higher

dissolution rates than H2SO4. As mentioned by White et al. (1994), although the

release of Fe(II) from pure ilmenite does not involve a redox reaction, natural

ilmenite commonly forms partial solid solutions with hematite, from which reductive

dissolution can occur.

Interestingly, in the presence of Sn(II), the OCP of the Pt electrode initially

decreases with time (Fig. 6.5), indicating that (i) it is affected by the changes in the

ratio of [Fe(II)]/[Fe(III)] which occur due to the reaction of ilmenite, (ii) it is also a

mixed potential arising from reaction in Eq. 6.6. The OCP values may also be

compared with the standard reduction potentials of various redox couples in the Fe-

Ti-oxide-water system reported by Zhang and Nicol (2009) listed in Table 6.2 and

the values based on the HSC 7.1 database (Roine, 2011) described later.

182
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.8. Measured OCP in (a) H2SO4 (4.6 M) (Zhang and Nicol, 2009) (b) HCl
(7 M).

Table 6.2
Standard reduction potentials.

Eo / V (60 oC)
Reaction
Zhang and Nicol (2009) HSC 7.1
3+ - 2+ o
Fe + e = Fe 0.69 (1 M H2SO4, 25 C) 0.81
+ - 2+
Fe2O3 + 6H + 2e = 2Fe + 3H2O 0.75 0.74
FeTiO3 + 6H+ + e- = Fe2+ + Ti3+ + 3H2O - 0.29
TiO2+ + 2H+ + e- = Ti3+ + 2H2O 0.046
2TiO2++ H2O + 2e- = Ti2O3+ 2H+ 0.24 -0.31

183
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The general comparison and trends of the steady potentials in Figs. 6.6-6.8 and Table

6.1 are listed below:

(i) In the presence of both Fe(II) and Fe(III) the OCP values of MI and

Pt electrodes are the same but higher in H2SO4 than in HCl (Table

6.1).

(ii) In H2SO4 the OCP of MI in the presence of additives follows the

descending order: Fe(III) > Fe(III)+Fe(II) > none > Fe(II). This trend

is generally similar to the descending order in HCl: Fe(III) > none >

Fe(III)+Fe(II) > Fe(II) > Sn(II).

(iii) The OCP of MI electrode is generally lower and more sensitive to

the additives in HCl compared to H2SO4 indicating higher reactivity.

A systematic and comparative discussion of these trends is presented in the next few

sections.

6.2.3. General trends and applicability of the Nernst equation

Previous OCP measurements in sulfuric acid solutions at 60oC conducted by

Zhang and Nicol (2009) showed that the highest OCP of MI and Pt electrodes in

4.6 M or 450 g/L H2SO4 (900 mV) was measured in the presence of Fe(III) (20 g/L,

Table 6.1). Zhang and Nicol (2009) also showed that the measured OCP of both MI

and Pt electrodes in the presence of Fe(III) and Fe(II) (10 g/L each) increased

initially and reached the steady value of ~720 mV after about 10 minutes. A similar

behavior was observed in the experiments conducted in this study, as shown in Fig.

6.2, but the steady state OCP reached after 5-10 minutes was lower, at ~610 mV.

184
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

This shows the relative effect of the anions (sulfate or chloride) on the dissolved

Fe(III) and Fe(II) species, due to the presence of chloride instead of sulfate resulting

in a lower OCP.

As previously shown in the case of the Cu(II)/Cu(I) redox system (Zhang et

al., 2008), the uncomplexed (Fe3+/Fe2+) redox system (Eq. 6.7) and the formation of

metal ion-ligand complexes of both Fe(III) and Fe(II) (Eq. 6.8) with a general ligand

L (SO42- or Cl-) in a redox equilibrium can be expressed by Eqs. 6.9-6.10.

Fe3+ + e- → Fe2+ (both ions uncomplexed) (6.7)

Fe(III)Ln + e- → Fe(II)Lm + (n-m)L (both complexed, charge omitted) (6.8)

RT [Fe2+ ]
EFe3+ 2+ = E° - Ln RT (6.9)
Fe F [Fe3+ ]

RT L n β Fe III L
E°Fe III Ln Fe II Lm = E°Fe3+ Fe2+ − Ln n
(6.10)
F L m βFe II L
m

According to the expressions for stability constants (β) of Fe(III)Ln and

Fe(II)Lmgiven by βFe(III)Ln = [Fe(III)Ln]/{[Fe3+][L]n} and βFe(II)Lm=

[Fe(II)Lm]/{[Fe2+][L]m}, Eq. 6.9 changes to Eq. 6.10 at standard conditions assuming

equimolar Fe(II) and Fe(III) in solution and unit activity coefficients. The

equilibrium constants (βn, βm) for the formation of complex species Fe(III)Ln and

Fe(II)Lm which involve dissolved species depend on a number of factors: (i) ionic

strength of the solution, (ii) temperature, (iii) charge of the complex ion species, and

(iv) activity coefficients of dissolved charged or uncharged species. This is evident

from some of the values reported in the literature listed in Table 6.3.

185
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The OCP values of the MI electrode in HCl solutions at 50 oC are lower than

those in H2SO4 at 60 oC. Assuming that OCP of MI and Pt measure the Fe(III)/Fe(II)

couple and that m = n, Eq. 6.10 simplifies to 6.11 and 6.12.

2.303 R T βFe III L


n
E°Fe III Ln Fe II Lm = E°Fe3+ Fe2+ - Log (6.11)
F βFe II L
m

EOCP( MI or Pt ) 

 o   Fe ( III ) Ln   [ Fe( II )] 

2.303RT 2.303RT (6.12)
 E Fe ( III ) / Fe ( II )  Log   Log  
   Fe ( II ) L   [ Fe( III )] 

nF  m 

nF

A plot of EOCP (MI or Pt) as a function of log {Fe(II)/Fe(III)} should give a straight line,

2.303RT
where slope =  , and (6.13)
nF


   Fe ( III ) Ln 
y-intercept =  E o Fe ( III ) / Fe ( II ) 
2.303RT
Log 

(6.14)
   Fe ( II ) L

nF  m 

Fig. 6.9 shows a plot of OCP for Pt and MI electrodes as a function of

Log{[Fe(II)]/[Fe(III)]} for the different mixtures of Fe(III) (10 g/L) and Fe(II) (5, 10

or 20 g/L) based on the data in Figs. 6.6-6.7. The predicted slope based on Eqs. 6.11-

6.12 is -2.303RT/nF= -64 mV (for n=1, R = 8.314 J mol-1 K-1, T = 423 K, F = 96485

C mol-1). This value closely agrees with the slopes in the range of -58 to -65 mV of

the linear relationships in Fig. 6.9. The y-intercepts in Fig. 6.9 are in the range 604-

606 mV, indicating that the OCP of MI and Pt electrodes are sensitive to the

concentrations of chloro-complexes of Fe(III) and Fe(II) in solutions according to

Eq. 6.12.

186
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Further analysis, which would require reliable information on the speciation in

concentrated chloride solutions, is beyond the scope of the present study.

Fig. 6.9. Effect of concentration ratio of Fe(II)/Fe(III) on OCP of MI and Pt


electrodes at 50oC (data from previous figures).

Table 6.3
Stability constants of relevant iron(III) and iron(II)complexes in aqueous media.

Species Ionic L = SO42- L = HSO4- L = Cl-


strength Log β1 Log β2 Log β1 Log β1 Log β2
Fe(III) 0 4.04 1.34 - 1.48 2.13
Fe(III) 1.2 2.33 4.23 0.78
Fe(III) 2.0 0.759 1.06
Fe(II) 0 2.3
Fe(II) 1 1.0
2 0.35 0.40

at 25 oC; Logβn or m represents the equilibrium constant for Fe(III or II) + (n or m)L =
FeLn or m (charge omitted) (Sillen and Martell, 1964).

187
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.2.4. Comparison with HSC data

Despite the fact that the measured OCP values represent mixed potentials, the

applicability of the Nernst equation to the OCP values (Fig. 6.9) allows the

comparison of stable OCP values listed in Table 6.1 with the standard electrode

potentials of various redox couples. Figure 6.10 shows the standard reduction

potentials of various Fe-Ti-H2O redox couples in the temperature range 0-100 oC

based on the HSC data base (Roine, 2011).

(a) Platinum electrode

As noted earlier and seen in Fig. 6.10 the measured OCP values of Pt

electrodes in H2SO4 at 60 oC (Zhang and Nicol, 2009) are generally higher than the

values in HCl at 50 oC. It is of interest to note that the OCP of the Pt electrode in

equimolar Fe(III) and Fe(II) solution in 4.6 M H2SO4 (q) is close to the Eo value of

the Fe2O3/Fe2+and FeCl2+/Fe2+ based on the HSC data base. The other OCP values in

4.6 M H2SO4 as well as in 4 M HCl which fall in between two lines may well be

indicating mixed potentials and warrant further studies for confirmation.

188
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.10. Open circuit potentials of platinum electrode in H2SO4 or HCl (Conditions:
4.6 M H2SO4 at 60 oC (Zhang and Nicol, 2009) or 4 M HCl at 50 oC (this work),
200 rpm under nitrogen (measured data points from Table 6.1). Note the lines are
reduction potentials from HSC 7.1(Roine, 2011) and the data points with symbols are
from OCP measurements.

189
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

(b) MI electrode

Fig. 6.11 shows a comparison of the OCP values of MI electrodes with the

calculated Eo values based on the HSC database. The OCP value of the MI electrode

obtained in H2SO4 at 60 oC was 720 mV denoted by (q) (Zhang and Nicol, 2009)

which may be compared with the values of Fe3+/Fe2+ (810 mV) and Fe2O3/Fe2+

(737 mV) plotted in Fig. 6.11.

It is apparent that the open circuit potential of the MI electrode (~ 645 mV) in

HCl without other additives such as Fe(II), Fe(III) or Sn(II) denoted by (r) is higher

than that of the Pt electrode (~ 610 mV, Table 6.1) and is close to the Fe2O3/Fe2+ and

Fe2O3/FeCl2+ couples in Fig. 6.11 at 50 oC based on the HSC 7.1 database.

Another important feature in Fig. 6.11 is that the OCP value of MI was lower

in the presence of Fe(II) or Sn(II) which supports the reductive role of these reagents.

The fact that the OCP of MI in the presence of Sn(II) denoted by (v) falls on the lines

representing Sn4+/Sn2+, Fe2O3/Fe3O4 or Fe3O4,TiO2 /FeTiO3 equilibrium boundaries

also warrants further studies. Dissolution of hematite from MI has occurred at the

electrode surface as shown by the close agreement of the measured value denoted by

(s) with line (b) and (c) in Fig. 6.11.

190
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.11. Open circuit potentials of massive ilmenite electrode in H2SO4 or HCl
(Conditions: 4.6 M H2SO4 at 60 oC (Zhang and Nicol, 2009) or 4 M HCl at 50 oC
(this work), 200 rpm under nitrogen. Note the lines are reduction potentials from
HSC 7.1 (Roine, 2011) and the data points with symbols are from OCP
measurements.

191
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.2.5. Effect of Fe(II), Fe(III) and Sn(II)

The change in open circuit potentials of MI and Pt electrode in 7 M HCl

solution in the presence of equimolar Fe(II) and Fe(III) at 50 oC in Fig. 6.2 shows

that the potential of the MI electrode rapidly increased from 500 mV to 605 mV after

about 45 seconds and remained at that potential throughout the experiment. The

potential of the platinum electrode gradually increased from an initial value of 450

mV to reach a steady value of 605 mV after about 7 minutes. The mixed potential of

the MI electrode obtained in this solution was lower than the mixed potential value in

HCl solution (640 mV) without additives (Fig. 6.1).

Fig. 6.9 highlighted the fact that the change in [Fe(II)]/[Fe(III)] ratio affected

the OCP of MI and Pt electrodes. In contrast, Fig. 6.12 compares the change in OCP

of MI and Pt electrodes due to the addition of Fe(II) or Sn(II) alone, in the absence of

added Fe(III). The OCP of MI (Fig.6.12a) is higher than that of Pt (Fig.6.12b) in the

absence of additives. The addition of Fe(II) increases the ratio of [Fe(II)]/[Fe(III)]

and thus decreases the OCP, as expected from Fig. 6.9. The effect of Sn(II) is much

larger than that of Fe(II) in Fig. 6.12b indicating the increase in [Fe(II)]/[Fe(III)]

ratio due to reaction in Eq. 6.6 i.e. 2Fe(III) + Sn(II) = 2Fe(II) + Sn(IV). The effect of

Fe(II) and Sn(II) on the OCP of MI is much larger (Fig. 6.12a) compared to that on

the OCP of Pt (Fig. 6.12b).

192
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

(a) MI (b) Pt

Fig. 6.12. Effect of Fe(II) and Sn(II) on open circuit potentials of (a) massive
ilmenite and (b) platinum electrodes in 7 M HCl in the absence of added Fe(III),
Fe(II) = 10 g/L, Sn(II) = 20 g/L (T = 50 oC, stirring rate = 200 rpm).

Based on the results of the study of the chemical dissolution of a synthetic

ilmenite disc described in Chapter 5, it was concluded that the addition of Sn(II)

facilitated the removal of the hydrolysed product layer of Ti(IV) by reducing it to a

soluble Ti(III) species. The addition of Fe(III) at a given concentration of Fe(II)

decreased the [Fe(II)]/[Fe(III)] ratio and increased the potential of both MI and Pt

electrodes.

Ilmenite reactions in the absence of SnCl2:

FeO·TiO2 + 2H+→ Fe2++ TiO2 + H2O (6.15)

TiO2 + 2HCl → TiOCl2·H2O (hydrolysed Ti(IV) product) (6.16)

FeO·TiO2 + 4H+→ Fe2+ + TiO2+ + 2H2O (6.17)

Fe2O3 + 6H+→ 2Fe3+ + 3H2O (6.18)

193
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Reactions in the presence of SnCl2:

2TiOCl2 + Sn2+ + 4H+→2Ti3++ Sn4+ +4Cl- + 2H2O (6.19)

2Fe3++ Sn2+→2Fe2++ Sn4+ (6.20)

Fe2O3 + Sn2+ + 6H+→ 2Fe2+ + Sn4+ + 3H2O (6.21)

Fig. 6.13 shows the effect of temperature on log K for various dissolution or

precipitation reactions based on the HSC 7.1 database. The acid dissolution of

ilmenite, either partial (curve a, in Fig. 6.13) or complete (curve b, in Fig. 6.13) is

thermodynamically favorable, with large K values in the range of 1012 to 103 over the

temperature range 0-100 oC. However, as evident from the results discussed in

Chapter 5, the TiO2 produced in Eq. 6.15 can easily react with HCl according to

reaction 6.16 and hydrolyse to insoluble products which block the surface retarding

further reaction. The equilibrium constants involving Ti3+ are not shown in Fig. 6.13

due to the unavailability of thermodynamic data for Ti3+ ions in the HSC 7.1

database. As shown by the large values of K (101.5 at 50 oC) in Fig. 6.13 (line d), the

Fe3+ ions produced in Eq. 6.18 can also be hydrolysed and precipitated as Fe2O3.

However, this is unlikely to occur in HCl, in which the chloride ions can also

stabilize Fe3+ by forming chloro-complexes such as FeCl2+, FeCl2+ and FeCl30

(Senanayake and Muir, 2003). The presence of Sn2+ can facilitate the reductive

leaching process.

194
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.13. Effect of temperature on equilibrium constants (Roine, 2011)

The acid dissolution of Fe2O3 according to reaction 6.18 (curve c in Fig. 6.13)

is unfavorable, as indicated by its low equilibrium constants (K = 10-1.5 at 50 oC).

Although the presence of strong HCl facilitates leaching, the rate is very low even in

strong acid solutions at 55 oC (Senanayake and Muir, 2003). Again, the reductive

leaching according to Eq. 6.21 can facilitate the leaching of Fe2O3 to produce Fe2+.

The formation of Fe3O4 is also a possibility in the presence of Sn2+ as predicted by its

equilibrium constant (K = 10-0.12 = 0.71) in Fig. 6.13 (curve e). As shown in Fig.

6.11, the measured OCP of the massive ilmenite electrode (Fig. 6.12) coincides with

the predicted value for the two redox couples Fe3O4/Fe2O3 and Sn4+/Sn2+ and

warrants further future studies.

195
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.3. Cyclic voltammetric study

Electrochemical methods have often been used to understand the reactions of

minerals involved during leaching processes. The electrochemical results are

applicable only if the experimental conditions closely simulate the real leaching

conditions. As mentioned previously in the literature review (Chapter 2), studies on

ilmenite have been conducted by several researchers using cyclic voltammetry, from

which the peak value of current and the corresponding potential gives useful

information about the reactions. The cyclic voltammetric behavior of ilmenite in

sulfuric acid solution and the corresponding reactions have been previously

determined (Zhang and Nicol, 2009; Adriamanana et al., 1984). None of the

electrochemical studies have focused on ilmenite leaching in HCl solutions.

Therefore, the aim of this study is to understand the behaviour of ilmenite in HCl

solution and compare it with that in H2SO4 solution.

The cyclic voltammetric experiments were also conducted in H2SO4 using

ilmenite (Iluka) to compare with the previous study by Zhang and Nicol (2009). The

ilmenite samples used in this experiment can be divided into two types which are

massive and powdered ilmenite. Ilmenite powder (synthetic and natural ilmenite)

was mixed together with carbon paste to make a carbon paste electrode. The method

used to prepare the electrodes was described in Chapter 3 (section 3.3.2). In all the

experiments, the potential sweep was initiated from the OCP to negative potentials

(-0.40 V) unless otherwise stated. The conditions used and the peak potentials from

the CVs reported in the next few sections are shown in Table 6.4.

196
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Table 6.4
Summary of peak potentials from cyclic voltammogram curves under different
conditions.

T [Acid] Metal Cathodic peak (V) Anodic Peak (V)


Acid E Ref
(oC) ( M) ion (s) C1 C2 A1 A2
0.20,
60 HCl 4 SI CP 0.61 0.70 0.10 (A)
0.0 (C3)
60 HCl 4 MI 0.5 0.10 0.80 - (A)

60 H2SO4 4 IL-CP 0.62 -0.009 0.71 0.19 (A)

60 HCl 4 IL-CP 0.59 - 0.62 - (A)

60 HCl 4 TW- CP 0.40 - 0.90 - (A)

60 HCl 4 NC- CP 0.23 - - - (A)


0.40
60 HCl 4 Sn(II) IL-CP 0.26 0.90 -0.14 (A)
(A3)
450
60 H2SO4 IL- CP 0.60 0.00 0.77 0.2 (B
g/L
450
60 H2SO4 MI 0.65 0.10 - -0.10 (B)
g/L

(A) This work ;


(B) Literature data from Zhang and Nicol (2009);
MI: Massive ilmenite electrode, SI: Synthetic ilmenite, CP: Carbon paste electrode; C: cathodic peak,
A: Anodic peak; Ref : References.; E : electrode

6.3.1. Cyclic voltammetry of synthetic ilmenite in HCl solution

Cyclic voltammetry studies of synthetic ilmenite shed light on the behaviour

of unaltered ilmenite in HCl solution without any interference from other associated

minerals such as hematite. Synthetic ilmenite used for preparing the rotating disc

employed for the dissolution study in Chapter 5 contained Fe(II)) and Ti(IV)

according to its formula FeTiO3.

197
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.14 shows the cyclic voltammogram (CV) of synthetic ilmenite carbon

paste electrode in 4 M HCl at 60 oC. A comparison of the peak potentials from the

CV measurements (Table 6.4) and the standard reduction potentials from HSC 7.1

data base is shown in Fig. 6.15.

Fig. 6.14. Cyclic voltammogram (CV) of synthetic ilmenite carbon paste (SI-CP)
electrodes recorded at 60 oC in 4 M HCl at a scan rate of 5 mV/s (0 rpm).

198
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.15. Peak potentials of cyclic voltammogram of synthetic ilmenite carbon paste
(SI-CP) electrode (Conditions: 4 M HCl, 60 oC, 0 rpm). Note the lines are reduction
potentials from HSC 7.1 (Roine, 2011) and the data points with symbols are from
CV peak potential measurements at 60 oC.

199
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The CV of synthetic ilmenite indicated that during the potential sweep from

the open circuit potential to negative directions, a first cathodic peak C1 appeared at

0.61 V (SHE) as shown in Fig. 6.14 and Table 6.5. This cathodic peak value when

compared with calculated standard reduction potential values based on the HSC 7.1

database (Fig. 6.15), is close to the Fe2O3/Fe2+ and Fe2O3/FeCl2+couples at 60 oC. It

indicates the reduction of Fe(III) at the synthetic ilmenite surface to Fe(II). This

could be described by the reactions mentioned in Table 6.2 which are also shown

below (Eqs. 6.21 and 6.22):

Fe3+ + e-→ Fe2+ Eo = 0.69 V (1 M H2SO4 at 25oC) (6.22)

Fe2O3 + 6H+ + 2e-→ 2Fe2+ + 3H2O Eo= 0.75 V (6.23)

As the scan proceeded towards more negative potentials, two small cathodic

peaks, C2 and C3, were observed at 0.2 V and 0.0 V, respectively. The potential

value of peak C2 is close to the standard value of the Fe2O3.TiO2/ FeTiO3 couple

plotted in Fig. 6.15. Peak C3 potential is close to that of FeTiO3/Ti2O3, Fe2+. The

cathodic peak (C3) is similar to that reported by Zhang and Nicol (2009) and it

corresponds to reduction of Ti(IV) to Ti(III) as represented by Eq. 6.23.

TiO2+ + 2H++e-→ Ti3+ + H2O Eo = 0.046 V at 60 oC (6.24)

In positive direction, peak A1 appears and based on the calculated value in Fig. 6.15,

peak A1 appears to be due to the oxidation of Ti(III) produced at the surface (Eq.

6.24) as identified by Zhang and Nicol (2009).

2TiO2++ H2O + 2e-→ Ti2O3+ 2H+ Eo = 0.24 V at 60 oC (6.25)

200
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

As the scan go through positive direction, a well defined peak (A2 at 0.7 V)

appears and based on calculated values in Fig. 6.15, this potential value is close to

that for the reaction assigned for peak C1 (Eq. 6.23, 6.24). This suggest that peak A2

corresponds to the oxidation of the product of the cathodic reduction of Fe(III) at

0.61 V (peak C1). Thus, the peaks in Fig. C2 and C3 indicate reductive leaching of

ilmenite and A1 and A2 indicates the oxidation.

Table 6.5
Cathodic peak potential and current density value for cyclic voltammogram of
stationary ilmenite under different conditions.

Peaks Current
Conditions
Electrode Peaks value density
(V) (mA cm-2)
C1 0.61 3.3
HCl (4 M), 60 oC, 0 rpm
SI- CP C2 0.20 2.1
C3 0.00 2.6
HCl (4 M), 60 oC, 0 rpm C1 0.50 6.1
MI
C2 0.14 9.8
HCl (4 M), 60 oC, 0 rpm
IL- CP C1 0.59 2.2
20 g/L SnCl2 + 4 M HCl 60 oC, 0
IL- CP rpm C1 0.30 50

According to the Table A1.1 in Appendix A1 the cathodic peaks can be assigned to:
C1: Fe3+/Fe2+; C2: TiO2+/Ti2O3; C3: TiO2+/Ti3+

201
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.3.2. Cyclic voltammetry of massive ilmenite electrode

During OCP measurements conducted on a massive ilmenite electrode as described

previously, it was found that the mixed potential of the electrode without Fe(II) or

Fe(III) additives was around ~640 mV. The mixed potential showed a decrease with

time in the presence of Fe(II) in the range from 400-600 mV. Cyclic voltammetry of

a massive ilmenite electrode was also conducted in 4 M HCl solution at 60 oC and at

a scan rate of 5 mV/s as described in this section. During the potential scan from the

OCP in the negative direction (Fig. 6.16), the first cathodic peak C1 appears at 0.5 V

(Table 6.5, Fig. 6.16). This potential value is lower than the FeCl2+/Fe2+ couple at

60 oC based on the HSC 7.1 database (Fig. 6.15). The OCP values obtained in the

presence of Fe(II) described previously are consistent with the dissolution of Fe2+

from ilmenite at this potential (Fig. 6.3).

Fig. 6.16. CV of a stationary massive ilmenite (MI) electrode recorded at 60 oC in


4 M HCl at a scan rate of 5 mV/s.

202
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

As the potential scan continued in the negative direction, a small peak (C2)

was observed at 0.1 V. This peak was attributed to the Fe2O3.TiO2/FeTiO3 couple

based on calculated reduction potentials from the HSC 7.1 database (Fig. 6.15). This

peak can also be attributed to Eq. 6.25, which describes the reduction of Ti(IV) from

the ilmenite surface to Ti(III).

TiO2+ + 2H++e-→ Ti3+ + H2O Eo = 0.046 V at 60 oC (Appendix A1) (6.26)

The potential of the broad peak at 0.8 V(A1) is close to the potential of the

Fe3+/Fe2+ couple, based on calculated values shown in Fig. 6.15. This indicated that

the oxidation process which occurs at this potential can be represented by Eq. 6.26:

Fe3+ + e-→ Fe2+ Eo = 0.81 Vat 60 oC (6.27)

6.3.3. Cyclic voltammetry of natural ilmenite powder in sulfuric acid

solution

Initially, cyclic voltammetric experiments of the carbon paste electrode of

ilmenite (Iluka sample) were performed using conditions similar to those used by

Zhang and Nicol (2009) for comparison (4 M H2SO4 solution at 60oC). The cyclic

voltammetry results from this study are compared with those reported by Zhang and

Nicol (2009) shown in Fig. 6.17(a) and 6.17(b). It can be seen that the comparison of

the CVs indicates the reproducibility.

203
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

During the potential scan in the cathodic direction the cathodic current

increased and peaked at 0.622 V (C1) as shown in Fig. 6.17(b). The peak potentials

are plotted in Fig. 6.18 to compare with the standard values calculated using the HSC

7.1 database. As evident from the comparison with the CV obtained by Zhang and

Nicol (2009), peak C1 corresponds to the reduction of Fe(III) to Fe(II) according to

Eq. 6.27. As the scan proceeded towards more negative potentials, current increased

in the cathodic direction and resulted in a defined cathodic peak at -0.009 (C2) which

corresponds to the reduction of ilmenite at the surface of the electrode (Eq. 6.28).

This peak potential is also shown in Fig. 6.18 which indicates that it is close to the

FeTiO3/Ti2O3, Fe2+ couple, confirming the reduction of ilmenite. On reversal of the

scan direction at -0.4 V, a small anodic peak appeared at 0.192 V (A1) which is

associated with the oxidation of reduced Ti(III) to Ti(IV) species as shown in Eq.

6.29 (Zhang and Nicol, 2009).

Further scanning towards positive potentials resulted in a second anodic peak

appearing at 0.708 V which represents to the oxidation corresponding to Fe3O4/Fe2O3

or FeTiO3/Ti2O3, Fe2+ couples, based on the calculated values in Fig. 6.18. The

obvious difference in the areas under anodic peak A2 and cathodic peak C1 indicate

that the reaction is not fully reversible.

Fe3+ + e-→ Fe2+ Eo = 0.69 V (6.28)

TiO2+ + 2H+ + e-→ Ti3+ + H2O Eo = 0.046 (6.29)

2TiO2+ + H2O + 2e-→ Ti2O3 + 2H+ Eo = 0.24 (6.30)

The measurement in 4.6 M H2SO4 solution confirmed that the behavior of

ilmenite was consistent with previous studies. Further studies were then conducted in

a 4 M HCl solution as described in the following sections.

204
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

(a)

(b)

Fig. 6.17. (a) CV of ilmenite carbon paste electrode in sulfuric acid (4M) at 60 ºC,
scan rate: 5 mV/s by Zhang and Nicol (2009), (b) CV of ilmenite (Iluka) carbon
paste electrode obtained in this study at 60 oC in 4 M H2SO4. Sweep initiated in the
negative direction from the open circuit potential and reversed at -0.400 V.

205
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.18. CV peak potentials of Iluka ilmenite/carbon paste electrodes in 4.0 or


4.6 M H2SO4 or 4 M HCl at 60 oC. Note the lines are reduction potentials from HSC
7.1 (Roine, 2011) and the data points with symbols are from CV peak potential
measurements.

206
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.3.4. Cyclic voltammetry of natural ilmenite in HCl solution

As discussed in section 6.3.3, two reactions which involve iron (peaks A2,

C1) and titanium (peaks A1, C2) have been identified during the CV scans of

ilmenite (Iluka) carbon paste electrodes in H2SO4 solution. However, it can be seen

from Fig. 6.19 that the CV curve of ilmenite (Iluka) in 4 M HCl showed only a single

reduction and oxidation peak during the scan. The cathodic peak occurred at 0.590 V

(C1) and the anodic peak at 0.628 V (A1). The reduction peak potential was similar

to that reported by Zhang and Nicol (2009) in sulfuric acid at 0.6 V (C1)

corresponding to the reduction of Fe(III) to Fe(II).

Fig. 6.19. CV of stationary ilmenite (Iluka) carbon paste electrode in (I) H2SO4 (4 M)
solution and (II) HCl (4 M) at 60 oC, scan rate = 5 mV/s, (III) carbon paste only,
sweep was initiated in the negative direction from the open circuit potential and
reversed at -0.400 V (see arrow).

207
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The differences in the two CVs of ilmenite in HCl and H2SO4 shed more light

on the leaching aspects of ilmenite. The dissolution of TiO2 in chloride and sulfate

media has been represented by the Eh-pH diagrams shown in Fig. 6.20, by previous

researchers. Fig. 6.20 (B) shows that TiO2 dissolves in H2SO4 at pH < 3.5 as

TiO(SO4)22- which is reduced to TiSO4+over a wide range of Eh from 0 V to -0.2 V

up to pH 2.75. Thus, the cathodic peak C2 in Fig. 6.19 appears to be representing the

reduction of acid dissolved TiO(SO4)22- to TiSO4+.

For chloride system as shown in Fig. 6.20 (A), TiO2 dissolves in HCl as

TiOCl4- in a very limited range of pH (0 to 1.5) and reduces to Ti(OH)2+ at the Eh

range 0.16 V to 0 V in this pH range. The cathodic peaks C3 of synthetic ilmenite

(carbon paste electrode) in Fig. 6.14 and massive ilmenite electrode in Fig. 6.16 are

in this range indicating reductive leaching.

At higher pH (>2.8), Ti(III) precipitates as Ti(OH)3 in both sulfate and

chloride systems (Fig. 6.20). Moreover the dissolved TiOCl42- or TiO(SO4)22- can

also precipitate as TiO2.H2O at higher pH, according to Fig. 6.20. Such precipitation

can also occur within pores, if the acidity becomes lower due to surface reactions

between surface oxides(s) and H+.

The absence of the cathodic peak for Ti(IV) to Ti(III) in ilmenite (Iluka)

carbon paste electrode in Fig. 6.19, compared to the massive ilmenite in Fig. 6.16

and synthetic ilmenite carbon paste electrode in Fig. 6.14 indicates the lower

exposure of TiO2 surface.

208
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.20. The Eh-pH diagram for (A) Ti-HCl-H2O (B) Ti-H2SO4-H2O system at
25 oC (Vaughan and Alfantazi, 2006).

The comparison of the CV peaks of massive ilmenite carbon paste electrode

with the calculated standard potential values in Fig. 6.18 shows that the peak

potential is close to the reduction potentials of the FeCl2+/Fe2+ or Fe2O3/Fe2+couples.

It could be assumed that this peak corresponds to the reduction of hematite mineral

which also has been identified as a component of the Iluka samples (Chapter 4). This

assumption could be supported with the statement by White et al. (1994) that

209
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

reductive dissolution could occur when ilmenite formed a solid solution with

hematite. Therefore, from these results, the reactions of natural ilmenite powder in

hydrochloric acid could only involve the reaction between the iron couple

[Fe(II)/Fe(III] either in solution or in the solid phase. It could also be due to the

reaction of hematite associated with ilmenite. In altered ilmenite, especially when

associated with hematite, Fe(III) has a higher tendency to be reduced compared to

Ti(IV) due to the higher potential value of the iron redox couple (0.77 V) compared

to the titanium redox couple (0.10 V). Table 6.5 presents the peak potential and

currents of the CV curve in 4 M HCl discussed above. The higher peak currents of

massive ilmenite electrode compared to carbon paste electrode is indicative of the

larger mineral (oxide) surface exposure of the former, compared to the latter.

Cyclic voltammetric experiments were also performed on other ilmenite

samples from Tiwest and North Capel to identify the voltammetric characteristics.

Fig. 6.21 depicts the CV of the Tiwest and North Capel ilmenite samples in 4 M HCl

at 60 oC. The CV of the Tiwest ilmenite samples shows very broad reduction peak at

around 0.4 V (C1) during the potential sweep from the rest potential to -0.4 V, and an

oxidation peak at 0.9 V (A1), during the reverse scan to 1.2 V.

The reduction peak corresponds to the Fe2O3.TiO2/FeTiO3 couple, as shown

in Fig. 6.21. The potential of the oxidation peak was close to that of the Fe3+/Fe2+

couple. For the North Capel sample, the potential of the reduction peak occurred at

0.23 V (C1) and no visible oxidation peak.

210
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.21. CV of ilmenite carbon paste electrode (TW & NC) in 4 M HCl solution at
60 oC, scan rate = 5 mV/s, sweep was initiated in negative direction from the open
circuit potential and reversed at – 0.400 V.

The reduction peak corresponds to the Fe2O3.TiO2/FeTiO3 couple, as shown

in Fig. 6.22. These differences in voltammetric characteristic are due to phase

alterations occurring in the ilmenite samples due to weathering. From the

characterization analysis described in Chapter 4, Tiwest concentrates have been

categorized as the highly altered because they contain other minerals associated with

ilmenite, such as hematite (Fe2O3), rutile (TiO2) and pseudorutile (FeTi3O9). The

existence of these minerals was confirmed by X-Ray Diffraction in Chapter 4.

211
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Based on the Ti/Fe content in each sample, the degree of alteration follows the

ascending order: North Capel < Iluka< Tiwest, with Tiwest being the most weathered

sample followed by Iluka and North Capel. For this reason, the Iluka sample,

categorized as moderately altered, displayed voltammetric behavior considered to be

standard in this chapter, unless stated otherwise.

Fig. 6.22. CV peak potentials of Tiwest/North Capel ilmenite/carbon paste electrodes


in 4 M HCl at 60 oC. Note the lines are reduction potentials from HSC 7.1 (Roine,
2011) and the data points with symbols are from CV peak potential measurements.

212
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.3.5. Effect of reducing agent

As discussed in Chapter 5, the presence of a reducing agent showed a

beneficial effect on the dissolution rates of a synthetic ilmenite disc when the

solution potential was changed from the rest potential to more negative potentials.

The OCP measurements show that the mixed potential of a massive ilmenite

electrode also decreased when Fe(II) and Sn(II) ions were added into solution,

supporting the reductive role of these reagents (Fig. 6.1-6.5).

Figs. 6.23 (A) and (B) show the CV of ilmenite (Iluka) in HCl in the absence

(Fig. 6.23A) or presence (Fig. 6.23B) of the reducing agent (SnCl2), respectively.

With the addition of Sn(II) ions, the cathodic peak C1 shifted to a more negative

potential and the peak current also increased (Fig. 6.23B). During the cathodic scan,

in the region from 0.0 V to -0.5 V, addition of Sn(II) resulted in a higher cathodic

current, about 20 times higher than in HCl alone. At more negative potentials (-0.5 V

to 1.2 V) a new anodic peak (A3) was observed at a potential of -0.14 V. This peak

was attributed to the oxidation of Sn(II) and identified on the CV of a carbon paste

electrode with no ilmenite, in 4 M HCl solution containing Sn(II), as shown in Fig.

6.23(C). It can be represented by half reaction 6.31 likely to correspond to processes

taking place at peaks A3 and C2:

Sn2+→ Sn4+ + 2e- (6.31)

Additional peaks at 0.4 V (A1) and 0.9 V (A2) were observed at more positive

potentials in the presence of Sn2+ in solution due to oxidation of the species reduced

at the cathodic peak at 0.2 V. This highlights the importance of considering the

limiting currents of Sn(II) oxidation.

213
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

(A)

(B)

(C)

Fig. 6.23. CV of (A) ilmenite (Iluka) in 4 M HCl alone at 60 oC; (B) in the presence
of SnCl2 (20 g/L) in 4 M HCl at 60 oC; (C) carbon paste electrode with no ilmenite,
under the same conditions as (B).

214
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.4. Limiting current of Sn(II) and Fe(II) oxidation

The experiments described in this section were conducted by measuring the

limiting current for the oxidation of Sn(II) to Sn(IV) or Fe(II) to Fe(III) as a function

of rotation speed () on the carbon paste electrode and applying the Levich equation:

iL = 0.62 F D2/3v-1/6½Cb (6.32)

IL = A*iL = 0.62 FA D2/3v-1/6½Cb (6.33)

Anodic currents were measured by using a carbon paste electrode in 1 M HCl

in the presence of tin(II) chloride or iron(II) chloride at concentrations varying from

0.25 g/L to 0.50 g/L at room temperature. For a Sn(II) or Fe(II) diffusion controlled

dissolution reaction, the diffusion coefficient of Sn(II) or Fe(II) is an important

parameter. The electrode potential was scanned from the OCP in the anodic direction

to 1.2 V. The carbon paste electrode with a surface area of 0.38 cm2 was used and its

rotation speed was varied from 200 to 1000 rpm.

Fig. 6.24 shows the anodic limiting current for Sn(II) oxidation at different

rotation speeds of the carbon paste electrode. According to the Levich equation the

plot of iL vs ω1/2 showed a linear relationship (Fig. 6.25). The slope of the linear

relationship was used to determine the diffusion coefficient,

Slope = 0.62 n F A D2/3 C v-1/6 (6.34)

where v is the kinematic viscosity for HCl at 25 oC is 0.9541 cSt, 1 cSt (centistokes)

= 1 x 10-6 m2 s-1(Nishikata et al. 1981).

215
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The calculated value of the diffusion coefficient DSn(II) is 1.7 x 10-6

cm2 s-1 (Appendix A3.1). It is slightly lower than a reported value for Sn(II) in

aqueous solution which is 6.5 x 10-6 cm2 s-1 (Low and Walsh, 2008) which may be

related to the higher viscosity in HCl solutions compared to water. For example, the

lowest value reported in non-aqueous ionic liquids is 2.3 x 10-7 cm2 s-1 due to high

viscosity (Yang et al., 2011). Higher values of Sn(II) diffusion coefficient are due to

the smaller complex ions formed with chloride or water molecules compared to

Sn(II) ions in ionic liquids, also considering that the viscosity of aqueous solutions is

lower than that of ionic liquids.

Fig. 6.24. Anodic limiting current curves measured at a rotating carbon paste disc
electrode in 1.0 M HCl solution in the presence of 0.5 g/L Sn(II) at 25 oC. The
potential was scanned from OCP to 1.2 V (scan rate: 5 mV/s).

216
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

7.0E-04

6.0E-04

Limiting current (A)


5.0E-04 y = 6.3E-05x
R² = 0.99
4.0E-04

3.0E-04

2.0E-04

1.0E-04

0.0E+00
0 5 10 15
ω1/2 (s -1/2 )

Fig. 6.25. Linear relationship of anodic limiting current for Sn(II) with square root of
rotation rate.

The anodic limiting currents of Fe(II) oxidation were also determined. Fig.

6.26 shows the corresponding anodic curves. Fig. 6.27 shows the linear plot of

current as a function of square root of the rotation speed. The calculated value of the

diffusion coefficient DFe(II) in HCl is 3.3 x 10-6 cm2 s-1. It is also lower than the

diffusion coefficient of Fe(II) in water at 25 oC (7.19 x 10-6 cm2 s-1) (Henry, 1994). In

both cases (Sn(II) and Fe(II)) the lower diffusivity compared to the values in water

can be related to the existence of chloro-complexes of Fe(II) and Sn(II) and lower

viscosity of concentrated HCl solutions. The calculated and reported diffusion

coefficient values are listed in Table 6.6.

217
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.26. Anodic limiting current curves measured at a rotating carbon paste disc
electrode in 1.0 mol/L HCl solution in the presence of 0.25 g/L Fe(II) at different
electrode rotation speeds. The potential was scanned from the OCP to 1.2 V; scan
rate = 20 mV/s.

6.0E-04

5.0E-04
y = 5E-05x
Limiting current (A)

4.0E-04 R² = 0.99

3.0E-04

2.0E-04

1.0E-04

0.0E+00
0 2 4 6 8 10 12
ω1/2 (s -1/2 )

Fig. 6.27. Linear relationship of anodic limiting current for Fe(II) oxidation with
square root of rotation rate ().

218
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Table 6.6
Diffusion coefficients of H+, Sn(II) and Fe(II).

Material / Medium Ion D (cm2 s-1) Reference


Bockris and Reddy,
Water H+ 9 x 10-5
1977
Water Fe2+ 7.2 x 10-6 Henry, 1994
Water Sn2+ 6.5 x 10-6 Low and Walsh, 2008

1-butyl-1-
methylpyrrolidinium
triflouro- Sn2+ 2.3 x 10-7 Yang et al., 2011
methanesulfonate
(BMPOTF) solution
This work
1 M HCl (aq) solution Sn2+ 1.7 x 10-6c
(Electrochemical)
This work
1 M HCl (aq) solution Fe2+ 3.3 x 10-6c
(Electrochemical)
Allen et al., 1979
Fe3O4 solid H+ 8 x 10-10a (Electrochemical-carbon
paste electrode)
Allen et al., 1979
NiO solid H+ 10 x 10-10 – 10-12
(Electrochemical)
Lu and Muir, 1985
Fe3O4b solid H+ 3.1 x 10-9
(Electrochemical)
a
Minimum possible value based on electrochemical studies with polycrystalline
magnetite.
b
at 0.1 M HCl and 25 °C.
c
Calculated from Levich equation

219
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.5. Coulometric measurement and dissolution study

To verify the extent of ilmenite dissolution in HCl solutions, coulometric

(CM) experiments were carried out by applying a constant potential to the stationary

electrode in HCl solution for several hours. The current passed through the cell was

recorded and the electric charge was calculated by integration of the current time

transient. The charge passed through the electrode was calculated by dividing the

integrated charge by the ilmenite surface area and the time. After the experiment, the

solution was analyzed for iron and titanium by ICP-EOS and the average dissolution

rate of iron and titanium was calculated in a similar fashion. All the result of solution

analysis is presented in Appendix A3.2. In these experiments, massive ilmenite and

ilmenite carbon paste electrodes were used to study the dissolution of iron and

titanium.

6.5.1. Massive ilmenite electrode

The massive ilmenite electrode was potentiostatically polarized at various

potential lower than the rest potential (~0.8 V) in 4 M HCl at 60 oC for a period up to

several hours. The resulting current was measured as a function of time. Current-time

transients obtained during these experiments are shown in Fig. 6.28. It is apparent

that the cathodic current density is relatively constant at the potential 0.5 V and 0.3

V. However, at lower potentials (-0.15, 0.0, 0.1 V), a high initial cathodic current

density is observed, which rapidly decreases reaching steady state after about 45

seconds (Fig. 6.29). This showed that the reductive dissolution of the sample

occurred at potentials lower than the rest potential, with the cathodic currents

indicating the reduction reactions.

220
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.28. Current-time transients during reduction of massive ilmenite electrode in


4 M HCl at 60 oC and different potentials (lower than rest potential) in 1 hour.

Fig. 6.29. Current-time transient of massive ilmenite electrode in 4 M HCl at 60 oC


and different potentials (lower than rest potentials) in 45 seconds.

221
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

The results based on the solution analysis for iron and titanium concentrations

are shown in Fig. 6.30 as plots of the average dissolution rate as a function of

potential. Also shown in Fig. 6.30 is the charge passed (in F m-2 s-1). The calculated

ratio of dissolution rates of iron and titanium are listed in Table 6.7. It is apparent

that under these conditions of acidity, temperature and time, selective dissolution of

iron occurs in this potential region with the rate of iron dissolution being 10 - 20

times faster than that of titanium (Table 6.7). The ratio of charge passed to iron

dissolution rate is close to one, showing that only iron dissolution took place in the

reactions at a potential of 0.5 V. However, the rate of dissolution of iron and titanium

increased significantly at potentials below 0.1 V, indicating that a reductive reaction

was involved.

The reductive leaching was demonstrated by the current-time transient

measurement, as described previously, in which the cathodic current increased at

more negative potentials. The results appear to show that ilmenite had a much lower

reactivity in hydrochloric acid solutions than hematite, so the dissolution behavior of

the mixed ilmenite-hematite ores is markedly dependent on the hematite and ilmenite

content. However, at potentials lower than 0.0 V, the dissolution of iron is higher

than the calculated charge, which may be due to a higher non reductive rate of (acid)

dissolution of hematite or ilmenite.

222
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

10
9 Fe

Rate ( 10-4 mol cm-2 s-1)


Ti
8
charge passed
7
6
5
4
3
2
1
0
-0.2 0 0.2 0.4 0.6
Potential (V)
Fig. 6.30. Potentiostatic dissolution of massive ilmenite electrode in 4 M HCl at
60 oC (Charge passed in 10-4 faraday cm-2 s-1).

Table 6.7
Effect of potential on the mean dissolution rates of Fe and Ti and electrical charge
for massive ilmenite electrode.

Potential Ratea Ratea Charge Fe/Ti Charge/Fe Charge/Ti Charge/(Fe+Ti)


(V) Fe Ti passedb
0.5 0.1 0.01 0.1 10.0 1.2 12.0 1.1
0.3 0.4 0.02 0.2 21.0 0.5 10.8 0.5
0.1 1.9 0.12 1.7 14.8 0.9 13.2 0.8
0.0 6.6 0.60 3.3 11.8 0.5 6.0 0.5
-0.15 8.4 0.70 5.4 12.0 0.6 7.7 0.6

In 4 M HCl at 60 oC
a. 10-4 mol m-2 s-1

b. 10-4 faraday m-2 s-1

223
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.5.2. Iluka ilmenite carbon paste electrode

Dissolution studies of Iluka ilmenite carbon paste electrodes have also been

carried out in 4 M HCl at 60 oC. Fig. 6.31 shows the average measured dissolution

rates of Iluka ilmenite over a period of 4 hours in 4 M HCl at 60 oC for potentials in

the range -0.15-0.3 V. Fig. 6.32 shows the current time transient over 60 seconds,

which indicated that the cathodic current response increased when lower potentials

were applied. It can be seen that the current is much higher at the initial stage and

after some time decreases to a lower value with negligible changes observed after

one minute. The increase in cathodic current indicated that the reductive dissolution

occurred at lower potential.

Fig. 6.31. Current-time transient during reduction of Iluka ilmenite carbon paste
electrode in 4 M HCl at 60 oC at different potentials (lower than rest potential) in 4
hours.

224
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

Fig. 6.32. Current-time transient of Iluka ilmenite carbon paste electrode in 4 M HCl
at 60 oC at different potentials (lower than rest potential) in 60 seconds.

The results of the solution analysis are shown in Fig. 6.33, as plots of the

average dissolution rate as a function of potential as well as the charge passed (in

faraday m-2 s-1). The calculated ratios of the dissolution rates of iron and titanium are

listed in Table 6.8. The rate of titanium dissolution from the carbon paste ilmenite

electrode was very low and consistent with the cyclic voltammetric results, in which

only the reductive/dissolution of iron was observed. The highest dissolution rate of

iron occurs at 0.0 V and the ratio of charge passed to the rate of iron dissolution

increased at lower applied potentials (Fig. 6.33). The higher rate of calculated charge

passed compared to the rate of iron dissolved could be due to the non-reductive

dissolution of hematite or ilmenite.

225
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

1.4
Ti
1.2

Rate (10-4  mol m-2 s-1)


charge passed
1.0 Fe

0.8

0.6

0.4

0.2

0.0
-0.2 -0.1 0 0.1 0.2 0.3 0.4
Potential (V)
Fig. 6.33. Potentiostatic dissolution of ilmenite (Iluka) carbon paste electrode in 4 M
HCl at 60 oC (Charge passed is in faraday m-2 s-1).

Table 6.8
Effect of potential on the mean dissolution rates of Fe and Ti from the Iluka carbon
paste electrode and charge passed.

Potential Rate Rate Charge Fe/Ti Charge Charge Charge


(V) Fea Tia passedb passed/Fe passed/Ti passed/
(Fe+Ti)
0.3 0.07 0.02 0.09 0.13 0.13 3.5 1.9
0.1 0.12 0.03 0.15 0.21 0.21 7.0 1.8
0.0 0.07 0.02 0.09 0.26 0.26 3.5 3.7
-0.15 1.28 0.11 1.39 0.32 0.32 11.6 0.3

In 4 M HCl at 60 oC
a. 10 -4 mol cm2 s-1
b. 10 -4 faraday cm-2 s-1

226
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.5.3. Comparison of rates of iron and titanium dissolution

The calculated rates of iron and titanium dissolution from ilmenite samples

based on coulometric measurements are listed in Table 6.9. The results from rotating

disc experiments described in Chapter 5 are also compared with results in this

chapter. The dissolution rates of massive, Iluka and synthetic ilmenite calculated

using coulometric and rotating disc measurements, agree reasonably well. The

dissolution rates of ilmenite in HCl solution also can be compared with dissolution

rates of ilmenite in H2SO4 solution (Zhang and Nicol, 2009). The reported rates of

dissolution of ilmenite in H2SO4 are lower than those measured in HCl in this study.

These results also agree with OCP measurements, which showed that the ilmenite

mixed potential was always lower in HCl than in H2SO4 which indicated that HCl

promoted higher dissolution rates.

Table 6.9
Dissolution rates of different ilmenite electrodes.

Ilmenite Conditions Methods Dissolving Rate


species (mol m2 s-1)
Fe(III) 0.1-8.4 x 10-4
Massive ilmenite 4 M HCl, 0.5 V to - CM
Ti(IV) 0.01-0.7 x10-4
0.15 V
Natural ilmenite Fe 0.07-1.28 x10-4
4 M HCl, CPE, 0.3 to CM
(Iluka) Ti 0.02-0.11 x10-4
-0.15 V
Synthetic Fe 0.06 - 0.19 x 10-4
4 M – 9 M HCl, 60 RD
ilmenitea o Ti 0.03- 0.22 x 10-4
C, 800 rpm
Synthetic Fe 0.38- 0.60 x 10-4
4 M HCl + SnCl2 RD
ilmenitea Ti 0.23- 0.41 x 10-4
(0.5-2 g/L) , 800 rpm
Natural ilmenite Fe 2.9 x 10-9
450 g/L H2SO4, CPE, CM
(Iluka)b Ti 1.9 x 10-9
0.0 V
a
Rotating disc of synthetic ilmenite (Chapter 5)
b
Zhang and Nicol (2009)
CM = Coulometry
RD = Rotating disc

227
Chapter 6:Electrochemical study of ilmenite dissolution in hydrochloric acid

6.6. Conclusion

The observed behaviour of natural ilmenite was different in HCl and H2SO4

solution as shown by open circuit potential and cyclic voltammetry analysis. In HCl

solutions, iron was selectively leached while in H2SO4 both iron and titanium were

dissolved from its lattice. However, it was found that both iron and titanium can be

dissolved from synthetic ilmenite in HCl solution. It was also found that HCl

promoted higher dissolution rates than H2SO4. The voltammetric characteristics of

ilmenite are dependent on the degree of alteration and weathering that have occurred

in natural ilmenite. Higher degrees of weathering or alteration resulted in lower

currents measured during the cyclic voltammetry scans. Hematite is always

associated with ilmenite and has a higher dissolution rate under cathodic conditions

compared to ilmenite. The scan under reducing conditions causes the following: i)

reduce Ti(IV) in the mineral phases to Ti(III), (ii) reduce Fe(III) to Fe(II) in iron-

bearing mineral phases, (iii) reduce Ti(IV) to Ti(III) at the electrode surface, (iv)

reduce Fe(III) to Fe(II) at electrode surface and (v) reduce Sn(IV) to Sn(II) at the

electrode surface if present. At the same time the acid reactions (Fe(II) and Ti(IV)

dissolution from ilmenite, possibly via a TiO2 intermediate) and Fe(III) dissolution

from iron minerals will continue. Therefore, this electrochemistry study has

identified the reactions of ilmenite and showed that the rate of reactions of both iron

and titanium can be enhanced at lower potential.

228
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

CHAPTER 7

7. REDUCTIVE AND NON-REDUCTIVE


LEACHING OF NATURAL ILMENITE IN
HYDROCHLORIC ACID SOLUTION

7.1. Introduction

In Chapter 5, it was shown that the dissolution of a flat surfaced synthetic

ilmenite disc in 4 - 9 M HCl solutions obeyed the parabolic rate law which indicated

that the reaction rate was limited by the diffusion of reactants/products through a

product layer formed on the surface. Further investigation showed that the leaching

rate was enhanced by adding a reducing agent (SnCl2) which inhibited the formation

of surface blocking solids. The significant effects of reducing agents in acid leaching

of ilmenite have been reported by a few researchers, as mentioned in Chapter 2

(Mahmoud et al., 2004; Lasheen, 2005 and El Hazek et al., 2007). These authors

used iron metal powder as a reducing agent and selectively leached iron from

ilmenite into solution. The reductive leaching of ferric oxide from the ilmenite

increased the reactivity of the ore, due to the breaking up of the grain structure which

allowed diffusion of the acid protons through pores created in the ore particles. It

was found that after the addition of iron powder, the surface area of the leach

residues increased due to partial dissolution of hematite and ilmenite (Mahmoud et

al., 2004).

229
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Moreover, the added iron powder dissolved readily in HCl creating nascent

hydrogen and part of the nascent hydrogen formed hydrogen gas (Mahmoud et al.,

2004). This was seen during the experiment where the gas bubbles were observed

after the reducing agent was added into the solution. These authors described the

significant role of proton diffusion through the pores, which determined the kinetics

of the reaction. All the reactions described by Mahmood et al. (2004) were reported

in Chapter 2.

Linear increase of the concentration of dissolved iron and titanium as a

function of time in the presence of SnCl2 indicated the importance of reductive

conditions. Under reducing conditions, the role of SnCl2 appeared to be the reduction

of Ti(IV) to Ti(III), thus facilitating the dissolution of ilmenite by avoiding the

surface blockage caused by the hydrolysed Ti(VI) species. The investigation was

extended in Chapter 6 where the electrochemistry of the reductive leaching of

ilmenite was studied.

This chapter describes the results from a study of the dissolution of natural

Australian ilmenite samples in hydrochloric acid which were conducted in batch

reactors to investigate the effect of ore particle sizes, acid concentration, temperature

and reducing agents. Three different ilmenite concentrates labelled as Iluka, Tiwest

and North Capel, which have different compositions and mineralogy as discussed

earlier in Chapter 4, were chosen in this study. The main differences between the

three ores are the content of FeO, Fe2O3 and TiO2, as presented in Fig. 7.1. These

differences had significant effects on the leaching efficiency of iron and titanium

from each ore which will be discussed and compared with previous results.

230
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

70

60 NC IL TW

50

Percentage (%)
40

30

20

10

0
FeO Fe2O3 TiO2

Fig. 7.1. Different composition of North Capel (NC), Tiwest (TW) and Iluka (IL)
ilmenite concentrates.

The kinetic models of particle dissolution will also be applied and compared

with results in Chapter 6. The standard initial conditions of the leaching experiments

are shown in Table 7.1.

Table 7.1
Parameters used in ilmenite leaching.

Pulp Size
Reducing
Variable studied HCl ( M ) T (oC) density fraction
agent (g/L)
(g/L) (m)

HCl concentration 4, 7, 9, 10, 80 4 +53-63 -


11

Temperature 7 80, 90, 100, 4 +53-63


110 -

+45-53
Particle size 7 80 4 +53-63 -
+63-75
+75-90
Sn,Fe
Reducing agent 7 80 4 +53-63 metal
powder
Concentration of
0.25, 0.5,
reducing agent 7 80 4 +53-63
1, 2, 4
(Sn metal powder)

231
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.2. Effect of particle size and mineralogy

7.2.1. Leaching efficiency

A study of the effect of particle size was conducted using four size fractions

of +45-53 µm, +53-63µm, +63-75 µm, +75-90 µm at 80 oC, solid to liquid ratio of

4 g/L in solutions of 7 M HCl. The extents of leaching of Fe and Ti from three

different ilmenite samples are shown in Figs. 7.2, 7.3 and 7.4. Particles of each size

fraction were analyzed for ferric, ferrous, total iron and titanium to evaluate any

changes in mineralogy. As shown in Table 7.2, the differences between the four size

fractions of a given ilmenite sample were not significant. Therefore, the

representative sample analyses were used in the leaching study.

Table 7.2
Chemical analysis of each ilmenite sample.

Species TiO2 FeO Fe2O3 Fe Total

Ilmenite type NC IL TW NC IL TW NC IL TW NC IL TW

+45-53µm 53.5 58.9 62.7 25.0 11.0 2.8 45.3 35.0 31.7 31.2 24.5 22.0

+53-63µm 53.4 59.1 62.7 25.2 11.1 2.9 45.2 35.0 31.5 31.3 24.5 22.2

+63-75 µm 53.4 59.0 63.0 25.3 11.2 2.8 45.2 35.4 31.5 31.4 24.8 22.3

+75-90µm 53.4 58.9 62.8 25.3 11.3 2.8 45.1 35.5 31.4 31.4 24.9 22.3

Representative 53.9 59.1 62.9 25.4 10.3 3.0 45.0 34.6 31.6 31.5 24.2 22.1

Footnotes : Iluka (IL), Tiwest (TW) and North Capel (NC)

232
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Fig. 7.2 shows the effect of particle size on the leaching of the North Capel

sample under condition stated above. For the first 60 minutes, it shows that iron was

leached at similar rates for particles of all size fractions. For the finest particle size

(+45-53 µm), the iron dissolution efficiency reached its maximum of 53% after 120

minutes. Under the same conditions, about 80% iron was dissolved from the North

Capel ilmenite sample of particle size +53-63 µm in 360 minutes. For Ti dissolution,

the efficiency increased with decreasing particle sizes (Fig. 7.2). The finest particle

size of +45-53 µm exhibited the highest Ti leaching efficiency of 96.2%.

In the case of Iluka and Tiwest samples, only small changes in leaching

efficiency were observed for particles of different sizes, as shown in Fig. 7.3. The Fe

leaching efficiency for the Iluka sample started to increase as the particle size

decreased, reaching a maximum dissolution efficiency of 32.8% after 240 minutes at

a particles of size range of +45-53 µm. The same trend occurred for Ti dissolution,

with the highest leaching efficiency of 30.0% being observed at a particle size of

+45-53 µm.

The Tiwest sample, however, displayed variable rates of Fe and Ti

dissolution from particles of different size ranges. The uncertainty of these results is

indicated by error bars shown in Fig. 7.4. Such behaviour is believed to be related to

the mineralogy of the samples which will be discussed later in this section.

Uncertainty in results of Fe and Ti dissolution studies in acid solutions at different

particle sizes have also been reported by previous researchers. McConnel (1978)

found that this behaviour was due to the degree of weathering or alteration of the

samples. In summary, the North Capel sample demonstrated the highest leaching

efficiency of Fe and Ti followed by the Iluka and Tiwest samples.

233
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100
90 90
80 80
Fe dissolved (%) 70

Ti dissolved (%)
70
60 60
50 50
40 40
30 size: +45-53 µm 30 size: +45-53 µm
20 size: +53-63 µm 20 size: +53-63 µm
size: +63-75 µm size: +63-75 µm
10 10
size: +75-90 µm size: +75-90 µm
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Time (min) Time (min)

Fig. 7.2. Effect of particle size on dissolution efficiency of Fe and Ti from North
Capel ilmenite sample ([HCl] = 7 M, pulp density = 4 g/L , T= 80 oC, t = 5 h).

100 100
size: +45-53 µm size: +45-53 µm
size: -63+53 µm size: +53-63 µm
80 80
size: +63-75 µm size: +63-75 µm
Ti dissolved (%)
Fe dissolved (%)

size: +75-90 µm
60 60 size: +75-90 µm

40 40

20 20

0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Time (min) Time (min)

Fig. 7.3. Effect of particle size on dissolution efficiency of (a) Fe and (b) Ti from
Iluka ilmenite sample ([HCl] = 7 M, pulp density = 4 g/L, T = 80 oC, t = 5 h).

234
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100
size: +45-53 µm
size:+45-53 µm
size: +53-63 µm size: +53-63 µm
80 80
Fe dissolved (%)
size: +63-75 µm size: +63-75 µm

Ti dissolved (%)
size: +75-90 µm size: +75-90 µm
60 60

40 40

20 20

0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Time (min) Time (min)

Fig. 7.4. Effect of particle size on dissolution efficiency of (a) Fe and (b) Ti from
Tiwest ilmenite sample ([HCl] = 7 M, pulp density = 4 g/L, T = 80 oC, t = 5 h).

7.2.2. Effect of alteration

In Chapter 4, Table 4.2 lists a summary of ilmenite properties obtained from

the characterization analysis. The degree of oxidation of the ferrous iron is expressed

by the mole ratio Fe3+/Fe while the degree of leaching of iron during weathering is

expressed by the mole ratio Ti/Fe. To express an overall degree of alteration of

ilmenite, a quantitative method has been proposed by McConnel (1978) described by

Eq. 7.1:

Fe 3+ 𝑇𝑖
M=( )x( ) (7.1)
𝐹𝑒 𝐹𝑒

where M is defined as the alteration factor. The value of M falls between the

theoretical limits of zero, for unweathered ilmenite (i.e. theoretical FeTiO3), to an

infinitely large number, for completely weathered material containing little or no

iron (e.g. leucoxene, TiO2).

235
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Therefore, the ilmenite samples used for the leaching showed different

degrees of alteration, as presented in Table 7.3. The degree of alteration followed an

ascending order of North Capel < Iluka < Tiwest.

Table 7.3
Alteration factor for ilmenite samples.

Fe3+/Fe Ti/Fe Alteration factor


Ilmenite
(a) (b) (M = a x b)
North Capel 1.00 1.2 1.20
Iluka 0.99 1.6 1.58
Tiwest 1.00 1.9 1.90

Based on the effect of particle size on the leaching results, the Iluka and

North Capel samples showed a general trend where the leaching rate of Fe and Ti

increased with decreasing particle size. This effect was less significant in the case of

the Tiwest samples. Pseudorutile (Fe2Ti3O9), which has a Ti/Fe mole ratio more than

1.5, is the major constituent of altered ilmenite, which has a lower reactivity than

that of ilmenite. This explains the low leaching efficiency observed for the Iluka and

Tiwest samples compared to the North Capel sample. The reactivity of rutile was

also reported to be even less than that of altered ilmenite (McConnel, 1978). The

gradual decrease in reactivity of ilmenite, altered ilmenite and rutile is related to the

stability of the Ti-O bond. It is general knowledge that altered ilmenite is more

difficult to dissolve than unaltered ilmenite. Physical changes accompanying the

alteration of ilmenite were described in Chapter 4. It was observed that the

weathering resulted in the formation of fine grained alteration products around the

outside of the particles and in some cases along fractures and cracks. This increases

the surface area greatly and hence the external surface area becomes unimportant.

Therefore, a significant effect of particle size would not be expected.

236
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.2.3. Fe-Ti correlation

Fig. 7.5 (a)-(c) shows the correlation between Fe and Ti extraction for the

three samples at different size fractions in 7 M HCl at 80 oC. A plot of percentage

dissolution of Ti against that of Fe, resulted in most of the data points for the North

Capel sample (Fig. 7.5 (a)) showing a linear correlation of slope 1. However, the

slope was higher than 1 in the case of fine particles (+53 - 45 µm) indicating that the

leaching efficiency of Ti was higher than that of Fe. The slopes of the linear

relationships are listed in Table 7.4.

A slope close to 1 indicated that the dissolution of Fe and Ti took place at the

same rate and can be represented by reaction 7.2.

FeO.TiO2 + 4HCl = FeCl2 + TiOCl2 + 2H2O (7.2)

The higher slope corresponding to the fine size fraction of +53-45 µm indicated

higher Ti leaching from smaller particles. This could be due to the higher surface

area of particles resulting in H+ having more access to unreacted ilmenite, which co-

exist with rutile and dissolve as TiCl4 (Eq. 7.3).

TiO2 + 4HCl → TiCl4 + 2H2O (7.3)

Table 7.4
Correlation data between Fe and Ti for ilmenite samples of different size fractions.

Size fraction (µm) and slope


Ilmenite Alteraction factor, M
+45-53 +53-63 +63-75 +75-90
North Capel 1.20 1.76 0.91 1.04 1.07
Iluka 1.58 0.88 0.93 0.86 0.70
Tiwest 1.9 0.81 0.99 0.74 0.69

237
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a)
100

80
y = 1.04x

Ti dissolved (%)
R² = 0.99
60

40
size: +45-53 µm
size: +53-63 µm
20
size: +63-75 µm
size: +75-90 µm
0
0 20 40 60 80 100
Fe dissolved (%)
(b)
35
size: +45-53 µm
30 size: +53-63 µm
size: +63-75 µm
25
Ti dissolved (%)

size: -+75-90 µm
y = 0.863x
20
R² = 0.97
15

10

0
0 10 20 30 40
Fe dissolved (%)
(c)
40
size: +45-53 µm
35
size: +53-63 µm
30 size: +63-75 µm
Ti dissolved (%)

25 size: +75-90 µm
20
15 y = 0.74x
R² = 0.88
10
5
0
0 10 20 30 40
Fe dissolved (%)

Fig. 7.5. Correlation between Fe and Ti extraction from ilmenite samples for
different size fractions (a) North Capel (b) Iluka (c) Tiwest ([HCl] = 7 M, pulp
density = 4 g/L, T = 80 oC, t = 5 h).

238
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

In the case of the Iluka and Tiwest samples, the linear correlation of slope ~1

was observed only for the smaller size fraction of +63-53 µm. The slope deviated

from unity for larger size fractions. This lower correlation between Fe and Ti (<1)

indicated an increase in the alteration factor (M). As shown in Table 7.4, North

Capel sample may be categorized as the least altered sample, followed by Iluka and

Tiwest. Therefore, the results obtained for the North Capel sample showed the co-

existence of Fe and Ti in the same oxide matrix, where H+ can penetrate to dissolve

them.

The most altered ilmenite sample (Tiwest) shows a more complex behaviour.

This could be because of the presence of altered mineral phases like hematite

(Fe2O3) and pseudorutile (Fe2Ti3O9). The reactions can be described by Eqs. 7.4 and

7.5:

Fe2O3 + 6HCl →2FeCl3 + 3H2O (7.4)

Fe2O3.3TiO2 + 18HCl → 2FeCl3 + 3TiCl4 + 9H2O (7.5)

7.3. Effect of acid concentration

7.3.1. Leaching efficiency

The effect of HCl concentration on the leaching efficiency of Fe and Tiwest

determined by varying the HCl concentration in the range of 4 M – 11 M using a

particle size fraction of +53-63 µm, 4 g/L pulp density at 80 oC for 5 hours. The

leaching curves are plotted in Fig. 7.6 - 7.8. The percentages of Fe and Ti dissolved

at different acid concentrations after 1, 2, 4 and 5 hours are listed in Table 7.5.

239
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

For the North Capel ilmenite sample, at low HCl concentration (4 M – 7 M),

the dissolution of Fe and Ti reached maximum values after 300 minutes of leaching

(Fig. 7.6). For acid concentrations higher than 7 M, the leaching efficiency reached a

maximum value after 120 minutes for Fe and after 60 minutes for Ti. Data given in

Table 7.5 show that the highest Fe leaching efficiency was achieved at 9 M HCl

(97%) after 120 minutes while 100% Ti dissolution was achieved at 11 M HCl after

60 minutes. The percentages of dissolved Fe for the NC sample given in Table 7.5

are generally lower compared to those of dissolved Ti at all HCl concentration. A

similar behaviour was reported by Das et al. (2013) and Laskhmanan et al. (2008) in

HCl- mixed-brine leaching of an ilmenite concentrate. They found that increasing the

salt concentration at a constant acid concentration resulted in a higher leaching

efficiency for Ti than for Fe.

100
100

80 80
Ti dissolved (%)
Fe dissolved (%)

60 60

40 40

HCl = 7 M HCl = 10 M HCl = 7 M HCl = 9 M


20 20
HCl = 11 M HCl = 9 M HCl = 10 M HCl = 11 M
HCl = 4 M
HCl = 4 M
0 0
0 100 200 300 0 100 200 300
Time (min) Time (min)

Fig. 7.6. Effect of acid concentration on Fe and Ti dissolution from North Capel
ilmenite sample. ([HCl] = 4 M -11 M, pulp density = 4 g/L, T = 80 oC, t = 5 h, size
fraction = +53-63 µm).

240
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.5
Effect of HCl concentration on efficiency of Fe and Ti leaching from natural
ilmenite.

Ilmenite dissolution
HCl
(M) 1h 2h 4h 5h
Fe Ti Fe Ti Fe Ti Fe Ti
North Capel(%)
4 35 29 45 46 63 65 67 79
7 40 43 55 60 67 74 79 83
9 86 95 97 98 97 98 97 98
10 90 98 97 98 97 98 97 98
11 94 100 97 100 97 100 97 100
Iluka (%)
4 13 7 20 12 29 21 31 21
7 21 13 36 23 51 33 55 37
9 27 22 45 40 61 51 62 62
10 61 48 75 53 86 61 85 59
11 66 48 82 53 88 61 88 59
Tiwest (%)
4 6 4 7 10 12 15 13 17
7 9 6 17 20 27 29 27 33
9 25 25 35 40 57 58 58 64
10 48 53 66 76 77 94 76 94
11 69 62 77 81 86 95 87 94
Other conditions: ([HCl] = 4 M -11 M, pulp density = 4 g/L, T = 80 oC, t = 5 h, size
fraction= +53-63 µm).

The effect of HCl concentration on the dissolution of the Iluka ilmenite

sample is presented in Fig. 7.7. It indicates that the dissolution of Fe increased with

time at HCl concentrations in the range of 7 M to 11 M. The maximum dissolution

of iron was 86.5% after 180 minutes of leaching in 11 M HCl. Similar results were

also obtained for the dissolution of titanium, which increased with increasing

concentration of HCl from 4 M to 11 M (Fig. 7.7). However, Ti leaching efficiency

was not affected when the HCl concentration was increased from 10 M to 11 M. The

maximum dissolution of titanium in 11 M HCl was about 57.2% after 180 minutes.

These maximum values are lower compared to those for the North Capel sample

under similar condition.

241
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100
HCl = 4 M HCl = 7 M
HCl : 4 M HCl : 7 M
HCl = 9 M HCl = 10 M
HCl = 11 M HCl: 9M HCl : 10M
80 80 HCl : 11 M

Ti dissolved (%)
Fe dissolved (%)

60 60

40 40

20 20

0 0
0 100 200 300 0 100 200 300
Time (min)
Time (min)
Fig. 7.7. Effect of acid concentration on Fe and Ti dissolution of Iluka ilmenite
sample. ([HCl] = 4 M -11 M, pulp density = 4 g/L, T = 80 oC, t = 5 h, size fraction =
+53-63 µm).

As shown in Fig. 7.8, the leaching efficiency of Fe from the Tiwest ilmenite

sample increased with increasing HCl concentration. In solutions of 4 - 11 M HCl

the fast initial rate of Fe dissolution leveled off after 60 minutes. The maximum

dissolution of 86% Fe was achieved at 11 M HCl after 240 minutes. The dissolution

of titanium also showed large increases with increasing HCl concentration. Initial

rapid dissolution was also observed during the first 60 minutes, leveling off after

reaching the maximum, which at 11 M was about 95% after 240 minutes.

242
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100 HCl = 4 M


HCl = 4 M HCl = 7 M
HCl = 9 M HCl = 10 M HCl = 7 M
HCl = 11 M HCl = 9 M
80 80 HCl = 10 M
HCl = 11 M
Fe dissolved (%)

Ti dissolved (%)
60 60

40 40

20 20

0 0
0 100 200 300 0 100 200 300
Time (min) Time (min)
Fig. 7.8. Effect of acid concentration on dissolution of Fe and Ti of Tiwest ilmenite
sample. ([HCl] = 4 M -11 M, pulp density = 4 g/L, T = 80 oC, t = 5 h, size fraction =
+53-63 µm).

Higher maximum values for Fe and Ti dissolution were achieved for the

Tiwest sample compared to the Iluka sample, but they were slightly lower than those

obtained for the North Capel sample. Based on the results obtained for all three

samples, the leaching efficiency of Fe and Ti increased with increasing HCl

concentration from 4 M to 11 M. The North Capel sample showed the highest

leaching efficiency for both Fe and Ti under these conditions, followed by the Tiwest

and Iluka samples. At higher HCl concentrations (10 M – 11 M), the North Capel

sample showed almost complete dissolution in less than 120 minutes.

243
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.3.2. Effect of alteration

The initial leaching of Fe from the North Capel sample was greater than that

of Ti but it did not remain so at longer times (Table 7.5). As the HCl concentration

increased, the percentage dissolution of Ti became higher than that of Fe. This

behaviour can be a result of the presence of hematite (Fe2O3) and rutile (TiO2) in the

sample which influence the initial and final dissolution behaviour. The alteration

factor in Table 7.3 showed that the North Capel sample was slightly altered ilmenite,

which contains rutile and fine grained hematite (in the altered region) as discussed in

Chapter 4. The results listed in Table 7.5 show that in 4 M HCl and 7 M HCl the iron

dissolution predominated over that of titanium for up to 60 minutes. This was

followed by a time at which some rates were similar. However, after 60 minutes Ti

dissolution became greater than that of Fe, until it reached complete leaching at

higher concentrations of HCl.

The Tiwest ilmenite sample, under similar condition, shows a similar

behaviour, where the percentage dissolution of Ti is greater than that of Fe (Table

7.5). The leaching reactions can be represented by Eqs. 7.2-7.4 described in the

previous section. However, the Iluka sample showed a lower Ti dissolution (%)

compared to the Tiwest sample, especially in 11 M HCl solution. A possible reaction

that may have caused a lower Ti dissolution in solutions of higher HCl concentration

is the precipitation of titanium, as represented in Eq. 7.6.

TiOCl2 + H2O → TiO2(s) + 2HCl (7.6)

244
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

As natural weathering proceeds, formation of rutile which is less reactive in

acid solutions, increases. In this case, Iluka sample rich in altered iron phase

(hematite) result preferential leaching of iron over titanium. When ilmenite ore

comes into contact with acid, this iron rich phase will dissolve rapidly and rutile if

present will slowly be dissolved followed by dissolution of the bulk of mineral. As

the leaching proceeded, Ti(IV) concentration increases and leads to the formation of

the solid species TiO2 or Ti(OH)4 (Eq. 7.6). At low pulp density and at high

temperature, the titanium polymerisation proceed very rapidly and this result

precipitation of Ti(IV) species in pores of particles. Polymerisation of titanium

would occur when the concentration of Ti(IV) reach > 10-3 M and H+ > 0.5 M (van-

Dyk et la., 2002). In the case of Iluka dissolution, titanium that dissolved in solution

is likely to reach the limits of Ti(IV) polymerisation resulting in lower Ti dissolution

compared to Tiwest and North Capel sample.

7.3.3. Fe-Ti correlation

The correlation between Fe and Ti dissolution from the North Capel, Iluka

and Tiwest samples at different acid concentrations is shown in Fig. 7.9 (a)-(c) and

the slopes of linear relationships and alteration factors are listed in Table 7.6. The

correlation data of all samples indicated that the slopes follow the order TW > NC

> IL. Based on Fig. 7.9(a) and Table 7.6, the North Capel sample showed a linear

correlation for Fe-Ti dissolution, at all HCl concentration, with slopes close to 1. The

dissolution of the Tiwest sample also showed linear correlations for Fe-Ti

dissolution which resulted in a slope of 1.22 at 4 M HCl which decreased to ~ 1 at

higher HCl concentrations (11 M).

245
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a)
100
4M 7M 9M
80 10 M 11 M

Ti dissolved (%)
60
y = 1.093x
40
R² = 0.99
20

0
0 20 40 60 80 100
Fe dissolved (%)
(b)
100
4M 7M 9M
80 10 M 11 M
Ti dissolved (%)

60

40
y = 0.832x
20
R² = 0.99

0
0 20 40 60 80 100
Fe dissolved (%)
(c)
100
4M 7M
80 9M 10 M
Ti dissolved (%)

11 M
60

y = 1.032x
40
R² = 0.968

20

0
0 20 40 60 80 100
Fe dissolved (%)

Fig. 7.9. Correlation between extraction of Fe and Ti from ilmenite samples at


different HCl concentrations (a) North Capel (b) Iluka (c) Tiwest. ([HCl] = 4 M -
11 M, pulp density = 4 g/L, T = 80 oC, t = 5 h, size fraction = +53-63 µm).

246
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.6
Slopes of linear correlation between Fe and Ti for ilmenite dissolution at different
HCl concentrations.

Alteration [HCl] (M) and slope (Ti/Fe)


Ilmenite
factor 4 7 9 10 11
North Capel 1.20 1.09 1.09 1.02 1.04 1.07
Iluka 1.58 0.67 0.67 0.83 0.73 0.69
Tiwest 1.9 1.22 1.16 1.06 1.17 1.03

The results obtained for the Iluka sample showed some deviation from

linearity at each HCl concentration, with the slopes being less than 1. Although the

Iluka and Tiwest samples have been classified as moderately altered ilmenite, the

leaching behaviour of these two samples differs significantly. It is possible that the

alteration products in these two samples consist of different phases. The main reason

for this view is that rutile and pseudorutile phases in Iluka samples are amorphous

compared to those in the Tiwest sample which were more crystalline, as seen from

the XRD scans described in Chapter 4. The XRD peaks of Tiwest ilmenite showed

the presence of a pseudorutile phase which was not observed in the Iluka samples.

The absence of pseudorutile peaks in XRD patterns of Iluka ilmenite is probably due

to an amorphous nature of pseudorutile. These amorphization products are likely to

promote the dissolution reactions (Sasikumar et al., 2004). Therefore, in the case of

the Iluka sample, Fe dissolution is favoured compared to Ti due to the presence of

pseudorutile, which contains Fe(III) and Ti(IV). The hydrolysis and precipitation of

the dissolved Ti caused a decrease in Ti leaching efficiency. This shows that

different reactions occurred because of changes in the solids composition and

solution concentration (van Dyk et al., 2002). However, further studies are essential

to confirm this dissolution behaviour.

247
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.4. Effect of temperature

7.4.1. Leaching efficiency

The three different ilmenite samples of size fraction +53-63 µm were leached

at 80 oC, 90 oC, 100 oC and 110 oC in solutions of 7 M HCl at 4 g/L pulp density,

over a period of 5 hours. The leaching efficiency of Fe and Ti at all temperatures is

presented in Figs. 7.10, 7.11 and 7.12 and Table 7.7. From Table 7.7, it can be seen

that as the leaching time progressed up to 1 hour, the extractions of Fe and Ti

increased with increasing temperatures from 80 oC to 110 oC. The highest leaching

efficiency was obtained for the North Capel sample, followed by the Iluka and

Tiwest samples. The North Capel sample showed the fastest leaching rate, with over

95% Fe and Ti being leached after 4 hours at 90 oC and after less than 2 hours at

100 oC and 110 oC.

It is apparent from Fig. 7.10 that temperature had a significant effect on the

rate of leaching of the North Capel sample. At 80 oC, the dissolved Fe increased over

time during the first 5 hours, reaching dissolution of 79%. At 90 oC, it reached 91%

after 180 minutes, levelling off thereafter. The same trend was observed for the

curves of 100 oC and 110 oC, where the maximum leaching (93% and 99%,

respectively) was reached after 120 minutes. As shown in Fig. 7.10, Ti was rapidly

dissolved during the first 60 minutes. At 80 oC, the dissolved titanium increased with

time and reached 79% after 5 hours. In contrast, at 90 oC, the Ti dissolution

increased rapidly and started to level off after 120 minutes, reaching 91%

dissolution. At the higher temperatures of 100 oC and 110 oC, 100% of the titanium

was dissolved into solution after 120 minutes.

248
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100

Fe dissolved (%) 80 80

Ti dissolved (%)
60 60

40 40
T = 80˚C T = 80˚C
T = 90˚C T = 90˚C
20 20
T = 100˚C T = 100˚C
T = 110˚C T = 110˚C
0 0
0 100 200 300 0 100 200 300
Time (min)
Time (min)
Fig. 7.10. Effect of temperature on dissolution of Fe and Ti for North Capel ilmenite
sample ([HCl] = 7 M, pulp density = 4 g/L, size fraction = +53-63 µm, leaching time
= 5 h).

Fig. 7.11 indicates that the leaching efficiency of Fe from the Iluka ilmenite

sample increased with the increasing temperature from 80 oC to 110 oC. A rapid

increase in the leaching efficiency of Fe was observed over the first 60 minutes of

leaching, slowing down after 5 hours. A maximum leaching of 96% was achieved at

110 oC after 5 hours. In the case of Ti dissolution, the leaching efficiency increased

with temperature and reached a maximum value of 51% at 100 oC. At 110 oC, the

titanium dissolution increased with time initially but decreased after 60 minutes of

leaching which resulted 40% dissolution.

249
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100
80 °C 90 °C

80 100 °C 110 °C
80

Ti dissolved (%)
Fe dissolved (%)

60 60

40 40

20 T= 80˚C 20
T= 90˚C
T= 100˚C
T= 110˚C
0 0
0 100 200 300 0 100 200 300
Time (min) Time (min)
Fig. 7.11. Effect of temperature on dissolution of Fe and Ti for Iluka ilmenite sample
([HCl] = 7 M, pulp density = 4 g/L, size fraction = +53-63 µm, leaching time = 5 h).

Result for the dissolution of Fe and Ti from Tiwest ilmenite presented in Fig.

7.12 show that the dissolution of Fe increased with increasing temperature from

80 oC to 110 oC. The highest dissolution of Fe, close to 100%, was obtained after

5 hours at 110 oC. However, the dissolution of Ti increased when the temperature

was increased to 100 oC and reached the highest dissolution of 74% after 5 hours

(Fig. 7.12). At 110 oC, the Ti dissolution only increased up to 40% after 60 minutes

and remained unchanged afterwards. This indicated that the precipitation of titanium

at higher temperature lowered the leaching efficiency.

250
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 100
T= 80˚C T = 80˚C
T= 90˚C T = 90˚C
80 T= 100˚C
80 T = 100˚C
T= 110˚C
Fe dissolved (%) T = 110˚C

Ti dissolved (%)
60 60

40 40

20 20

0 0
0 100 200 300 0 100 200 300
Time (min) Time (min)
Fig. 7.12. Effect of temperature on dissolution of Fe and Ti for Tiwest ilmenite
sample ([HCl] = 7 M, pulp density = 4 g/L, size fraction = +53-63 µm, leaching time
= 5 h).

7.4.2. Predicted equilibrium constants

Once again, differences were observed in the behavior of the tested samples

at various temperatures, as indicated by different leaching efficiencies of Fe and Ti.

Higher leaching efficiencies of Fe and Ti from all ilmenite samples were observed at

higher temperatures. Fig. 7.13 shows that the log K values (based on HSC 7.1

database) for various dissolution reactions generally decrease with increasing

temperature which can be used to predict the thermodynamic feasibility of the

reactions. For example, the values of log K for TiO2 dissolution to TiO2+ in acid

solutions increase at higher temperatures indicating that they are more favourable at

higher temperatures. Nevertheless, the low log K of < 0 indicates that this reaction to

TiO2 is not favoured.

251
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Fig. 7.13. Effect of temperature on equilibrium constants for oxide dissolution in


acid-chloride solutions (Roine, 2011).

Despite these predictions, the data given in Table 7.7 shows high Ti and Fe

leaching efficiency (%) indicating the involvement of chloro-complexes. Both ferric

and ferrous ions are quite soluble in HCl solutions, forming chloro-complexes, with

stronger complexing ability exhibited by the ferric ions. Formation constants for

Fe(III) complexes at 75 oC and 90 oC from data presented by Helgeson (1969) are

reported in Appendix A.1 (Table A.1.4). The formation of these complex species in

solutions of high acid and chloride concentration resulted in high leaching efficiency

of metal ions even in the absence of reducing agents.

252
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.7
Effect of temperature on efficiency of Fe and Ti leaching from natural ilmenite.
Time and leaching efficiency (%)
T (oC) 1h 2h 4h 5h
Fe Ti Fe Ti Fe Ti Fe Ti
North Capel (%)
80 35 43 55 60 67 74 67 83
90 64 69 81 88 94 90 95 91
100 85 95 93 98 96 99 97 100
110 92 99 99 99 100 100 100 100
Iluka (%)
80 13 7 20 12 29 21 31 21
90 30 19 42 26 63 36 64 38
100 54 37 68 46 79 49 83 51
110 74 42 84 43 91 41 96 40
Tiwest (%)
80 9 11 17 18 27 25 28 29
90 30 32 38 34 52 44 54 48
100 40 40 54 48 69 52 74 53
110 47 40 69 53 91 67 98 74
([HCl]: 7 M, 4 g/L pulp density, size fraction: +53-63 µm, 5 h)

7.4.3. Fe-Ti correlation

The relative leaching behaviour of Fe and Ti can be rationalised by

considering the correlation in Fig. 7.14. In the case of North Capel samples, Fe and

Ti dissolved completely. At the highest temperature (110 oC) both Fe and Ti

dissolved at stoichiometric ratio. This is evident from Fig. 7.14(a), which plots the Ti

extraction as a function of Fe extraction and shows a linear relationship with a slope

close to unity. In the case of the Iluka and Tiwest samples, as shown in Fig. 7.11 and

7.12, the Ti dissolution was lower even at 110 oC. This decrease was due to the

hydrolysis of the titanium chloro-complex resulting TiO2 precipitation throughout

the bulk of the solution and on the mineral surfaces.

253
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a)
100
80 °C 90 °C
80 100 °C 110 °C

Ti dissolved (%)
60

40 y = 1.013x
R² = 0.993
20

0
0 20 40 60 80 100
Fe dissolved (%)
(b)
100
80 °C 90 °C
80 100 °C 110 °C
Ti dissolved (%)

60

40
y = 0.652x
20
R² = 0.991

0
0 20 40 60 80 100
Fe dissolved (%)
(c)
100
80 °C 90 °C
80 100 °C 110 °C
Ti dissolved (%)

60

40
y = 0.986x
20 R² = 0.996

0
0 20 40 60 80 100
Fe dissolved (%)

Fig. 7.14. Correlation of Fe – Ti dissolution from ilmenite samples at different


temperatures (a) North Capel (b) Iluka (c) Tiwest.

254
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

This behavior was also observed by Verhulst at al. (2003) and Mahmoud et

al. (2004) at higher temperatures. The slopes of the linear relationships in Fig. 7.14

(a, b, c) are listed in Table 7.8. Unlike those corresponding to the North Capel

samples, which are close to 1 and relatively independent of temperature, the Tiwest

samples showed slopes close to ~1 at the temperature range 80 oC – 100 oC but a

lower slope of 0.61 at 110 oC. In this case, the Tiwest sample appears to have

resulted in the critical concentration of Ti for precipitation at 110 oC.

The results obtained for the Iluka sample showed a slope less than 1 and

followed the same trend as in different HCl concentrations as discussed previously.

Rapid dissolution of Fe from the more weathered sample, compared to Ti, indicated

that the iron dissolution took place from hematite, Fe2O3, which is known to dissolve

rapidly (Eq. 7.4) in HCl solutions of 7 M at temperatures close to 110 oC (Parida and

Das, 1996). Thus, based on the degree of alteration, North Capel samples having

lowest alteration factor resulted dissolution of Fe and Ti stoichiometrically. While

for Iluka sample having moderate alteration, however, Fe was preferentially leached

over Ti due to the weathering effect of rutile phase. Finally, the results for the Tiwest

sample indicated that Fe and Ti leached stoichiometrically at temperature from 80 oC

to 100 oC. However, at 110 oC Fe leached was higher than Ti due to hydrolysis and

precipitation of the latter.

Table 7.8
Slopes of linear correlation data between Fe and Ti dissolution from ilmenite
samples at different temperatures.
Slopes (Ti/Fe) at different temperatures
Ilmenite Alteration factor, M
80 oC 90 oC 100 oC 110 oC
North Capel 1.20 1.08 1.01 1.05 1.01
Iluka 1.58 0.67 0.59 0.65 0.49
Tiwest 1.9 0.99 0.89 0.99 0.61

255
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.5. Effect of reducing agent

7.5.1. Effect of iron metal powder

Results from the leaching study of ilmenite performed in the absence or

presence of reducing agents are given in Table 7.9 and Fig. 7.15. In these

experiments, metal powders of iron and tin were used as reducing agents (RA) for

the three ilmenite samples being studied. The comparison of the leaching efficiency

of Fe and Ti shows that there is a small improvement of the leaching of Fe and Ti.

For the North Capel sample, the effect of added iron metal powder on the leaching

efficiency was small, increasing only from 67.9% to 74.3% for Fe and from 65.3% to

73.3% for Ti.

Similar results were obtained for the Iluka sample, where the leaching

efficiency was increased from 28.9% to 41.7% for Fe and from 20.7% to 24.5% for

Ti. In the case of the Tiwest samples the opposite effect was observed, where Fe

leaching efficiency slightly decreased from 26.7% to 20.9% in the presence of iron

metal powder.

The dissolution of Ti from all ilmenite samples increased by only less than

10% in the presence of iron metal powder, compared to leaching without iron metal

powder. Generally, the trend in Fe and Ti extractions was similar to that obtained by

varying other parameters (HCl concentration and temperature). The leaching

efficiency followed the descending order of NC > IL > TW for Fe and NC > TW >

IL for Ti.

256
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.9
Effect of reducing agents on iron and titanium dissolution.

With RA
Ilmenite Without RA
RA = iron RA = tin
samples
Fe (%) Ti(%) Fe (%) Ti(%) Fe (%) Ti(%)
North Capel 67.9 65.3 74.3 73.3 91.2 100
Iluka 28.9 20.7 41.7 24.5 57.3 40.8
Tiwest 26.7 24.8 20.9 32.1 43.7 33.4
Experimental condition: [HCl] = 7 M, T = 80 oC, t = 240 minutes, [iron powder or tin
powder] = 1.0 g/L, pulp density = 4 g/L, RA = reducing agent, particle size = +53-63 µm,
time = 5 h.

(a) (b)

100 100 Ti (no RA)


Fe (no RA)
Fe dissolved after 4 h (%)

Ti (with Fe as RA)
Ti dissolved after 4 h (%)
80 Fe (with Fe as RA) 80
Ti (with Sn as RA)
Fe (with Sn as RA)
60 60

40 40

20 20

0 0
NC IL TW NC IL TW
Ilmenite sample Ilmenite sample

Fig. 7.15. Effect of reducing agent (RA) on dissolution of Fe and Ti for North Capel
(NC), Iluka (IL), Tiwest (TW) ilmenite samples ([HCl] = 7 M, T = 80 oC, pulp
density = 4 g/L, size fraction = +53-63 µm, time = 5 h).

7.5.2. Effect of tin metal powder

(i) Leaching efficiency

The presence of tin metal powder during ilmenite leaching resulted in higher

leaching efficiency in North Capel, Iluka and Tiwest samples, as shown in Table 7.9

and Fig. 7.15. Despite the higher molar mass of Sn (118.7 g/mol) compared to that of

Fe (55.8 g/mol) the effect of Sn was greater than that of Fe, at an added

concentration of 1.0 g/L.

257
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Further investigations were conducted to study the effect of dosage of tin

metal powder on the leaching of Iluka and Tiwest samples since these samples

showed lower Fe and Ti dissolution compared to the North Capel sample. The

dosage of Sn powder was varied from 0.25 g/L to 4.0 g/L in 5 hour leaching tests in

7 M HCl solutions at 80 oC. The leaching efficiency of Fe and Ti at different

concentrations of Sn metal powder is presented in Table 7.10. As the time of

leaching progressed from 1 hour to 5 hours, Fe leaching efficiency increased for both

the Iluka and Tiwest samples. It can also be seen that Fe leaching efficiency

increased with the increase in Sn addition. The Ti leaching efficiency also increased

for Iluka samples.

In the case of the Tiwest sample, the Ti leaching efficiency increased with the

increase in Sn from 0.48 g/L to 4.0 g/L during the first 4 hours. At higher Sn (2 –

4 g/L), the Ti leaching efficiency remained less affected after 4 hours of leaching.

The Iluka sample yielded the highest Fe and Ti leaching efficiency of 97%

and 90%, respectively at a Sn concentration of 4 g/L after 5 hours, while the Tiwest

sample yielded 100% Fe and 84% Ti. These results show the beneficial effect of Sn

as a reducing agent which enhanced the Fe and Ti dissolution from both Iluka and

Tiwest samples.

The effect of Sn addition is also shown in Figs. 7.16 and 7.17. Both samples

(Iluka and Tiwest) achieved high dissolutions of > 90% for iron and > 80% for

titanium in HCl solutions of lower concentration (7 M) and lower temperature

(80 oC) compared to previously published results (Zhang and Nicol, 2010).

258
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

The effect of Sn(II) as a reducing agent in the dissolution of synthetic

ilmenite is consistent with the rotating disc studies, where the leaching efficiency of

Fe and Ti was increased by a factor of about 4 to 10 times compared to leaching

without the reducing agent.

Table 7.10
Effect of Sn as a reducing agent on Fe and Ti leaching from natural ilmenite.
Sn (g/L) Fe and Ti leached (%) from Iluka sample
1h 2h 4h 5h
Fe Ti Fe Ti Fe Ti Fe Ti
0.00 13 7 20 12 29 21 31 22
0.25 22 20 32 29 41 41 46 43
0.48 25 18 41 27 58 41 61 43
1.00 31 30 50 47 77 81 80 80
2.00 36 27 61 52 82 84 86 88
4.00 45 18 67 54 91 78 97 90
Fe and Ti leached (%) from Tiwest sample
1h 2h 4h 5h
Fe Ti Fe Ti Fe Ti Fe Ti
0.00 9 14 17 21 27 29 28 36
0.48 22 14 17 23 45 34 47 36
1.00 32 35 58 50 79 66 79 70
2.00 43 35 69 58 93 79 100 79
4.00 49 42 75 83 96 83 100 84
([HCl] = 7 M, T = 80 oC, pulp density = 4.0 g/L, size fraction = +53-63 µm, time = 5
h).

259
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

100 Sn=0 g/L 100


Sn = 0 g/L
Sn= 0.25 g/L Sn = 0.25 g/L
Sn= 0.48 g/L
80 80 Sn= 0.48 g/L
Sn= 1.0 g/L
Sn = 1.0 g/L
Sn= 2.0 g/L

Ti dissolved (%)
Fe dissolved (%) Sn= 2.0 g/L
Sn = 4.0 g/L
60 Sn= 4.0 g/L
60

40 40

20 20

0 0
0 100 200 300 0 100 200 300
Time (min) Time (min)

Fig. 7.16. Effect of Sn metal on Fe and Ti leaching efficiency for Iluka ilmenite
sample. ([HCl] = 7 M, T = 80 oC, pulp density = 4.0 g/L, size fraction = +53-63 µm,
time = 5 h).

100 100
Sn = 0.0 g/L Sn=0.48 g/L
Sn = 0.48 g/L Sn = 1.0 g/L
Sn = 1.0 g/L Sn = 2.0 g/L
80 80 Sn = 4.0 g/L
Sn = 2.0 g/L
Fe dissolved (%)

Sn = 0 g/L
Ti dissolved (%)

Sn = 4.0 g/L
60 60

40 40

20 20

0 0

0 100 200 300 0 100 200 300


Time (min) Time (min)
Fig. 7.17. Effect of Sn metal on leaching efficiency of Fe and Ti for Tiwest sample
([HCl] = 7 M, T = 80 oC, pulp density = 4.0 g/L, size fraction = +53-63µm, time =
5 h).

260
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(ii) Fe-Ti correlation

Once again, the correlation of Fe and Ti was plotted at different Sn dosages

as shown in Fig. 7.18 (a)-(c) for all three samples. The slopes of these correlations

are presented in Table 7.11. The Fe and Ti correlation and slope data for the North

Capel sample are only available for 0.0 g/L Sn and 1.0 g/L Sn (Fig. 7.18(a), based on

the data in Appendix A4.1, Fig. A4.1). Both Fig. 7.18 and Table 7.11 show that the

slopes are much closer to unity in the experiments conducted in the presence of

added tin metal powder. However, in the case of the Tiwest sample, which had the

highest alteration factor of 1.9 (Table 7.11), the slope is closer to 1 in the absence of

added tin, but lower than 1 in the presence of added tin. The general trend in the

presence of 1 g/L Sn appears to be that, as the alteration factor increased in the order

NC < IL< TW, the slope of the linear correlation decreased in the order NC > IL >

TW. This suggests that the addition of 1 g/L Sn facilitates the dissolution of Fe over

Ti from highly altered ilmenite samples due to the presence of more oxidized iron.

This finding is consistent with previous studies where the addition of reducing

agents enhanced the dissolution of iron from natural hematite ore during leaching in

HCl solutions at elevated temperatures (Parida and Das, 1996).

Table 7.11
Slopes of linear correlation data between Fe and Ti dissolution from ilmenite
samples at different tin dosages.
Alteration Slopes (Ti/Fe) at different Sn dosages (g/L)
Ilmenite
factor, M 0.0 0.25 0.48 1.0 2.0 4.0
North Capela 1.2 0.97 1.1
Iluka 1.58 0.67 0.95 0.69 0.98 0.95 0.84
Tiwest 1.9 0.99 n/a 0.69 0.70 0.87 0.86
([HCl] = 7 M, pulp density = 4 g/L, size fraction = +53-63 µm, time = 5 h).
a
These result are from Appendix A4.1

261
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a)
100
Sn(II) = 0.0 g/L
80
Sn(II) = 1.0 g/L

Ti dissolved (%)
60
y = 1.1x
40
R² = 0.99
20

0
0 20 40 60 80 100
Fe dissolved (%)
(b)
100
Sn(II) = 0.0 g/L
80 Sn(II) = 0.25 g/L
Sn(II) = 0.48 g/L
Ti dissolved (%)

Sn(II) = 1.0 g/L


60 Sn(II) = 2.0 g/L
Sn(II) = 4.0 g/L y = 0.98x
40
R² = 0.99
20

0
0 20 40 60 80 100
Fe dissolved (%)
(c)
100
Sn(II) = 0.0 g/L
80 Sn(II) = 0.48 g/L
Ti dissolved (%)

Sn(II) = 1.0 g/L


60 Sn(II)= 2.0 g/L
y = 0.87x
Sn(II) = 4.0 g/L R² = 0.97
40

20

0
0 20 40 60 80 100
Fe dissolved (%)

Fig. 7.18. Correlation of Fe – Ti dissolution from ilmenite samples at different Sn


dosage (a) North Capel (b) Iluka (c) Tiwest.

262
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.5.3. Electrode potentials and leaching reactions

The effect of applied potential on the dissolution rate of synthetic ilmenite

has been investigated under non-reducing and reducing condition in previous

chapters. Therefore, in this leaching study of natural ilmenite, similar measurements

were performed using a platinum electrode to measure the potential of the solution

and a carbon paste electrode to measure the potential of ilmenite during the leaching

process. Both electrodes were immersed in the leach slurry in the reaction vessel

during leaching experiments. The potential of the two electrodes was recorded as a

function of time and the solution samples were also analyzed for dissolved iron and

titanium (Appendix A4, Table A4.1)

As shown in Fig. 7.19 for the leaching study of the Iluka ilmenite sample in

7 M HCl solution without a reducing agent, the initial potential was recorded at 0.1

V which gradually decreased to -0.3 V after 1 hour, then increased to reach a steady

state value at -0.2 V. The mixed potential of ilmenite recorded using a carbon paste

electrode during leaching is shown in Fig. 7.20. It shows that the measured potential

decreased slightly from 1.0 V to 0.8 V in the absence of Sn, but it was initially

lower, at 0.7 V, in the presence of Sn. The measured Pt and carbon paste electrode

potentials are listed in Table 7.12. The measured solution potential (-0.2 V) under

non reducing condition (in the absence of RA), indicated that the soluble species was

Ti3+ and Fe2+ as consistent with the Eh-pH diagram in Fig. 7.21.

263
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

0.2
0.1
without Sn with Sn
0
-0.1
-0.2
Ept (V)

-0.3
-0.4
-0.5
-0.6
-0.7
-0.8
0 1 2 3 4 5 6
Time (h)

Fig. 7.19. Measured potential of platinum electrode (Eh or Ept) during leaching of
Iluka ilmenite sample in 7 M HCl solution at 60 oC with and without the presence of
reducing agent (Sn).

1.1
1
0.9
0.8
Em (V)

0.7
0.6
0.5
0.4
0.3 without Sn with Sn

0.2
0 2 4 6
Time (h)

Fig. 7.20. Measured potential of carbon paste electrode (Em) during leaching of Iluka
ilmenite sample in 7 M HCl solution at 60 oC with and without the presence of
reducing agent (Sn).

264
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

The measured ilmenite potential in 7 M HCl is 0.8 V corresponded to

potential value obtained from electrochemical study for Iluka carbon paste electrode.

At the same conditions, the mixed potential of Iluka ilmenite is higher than the

solution potential (-0.2 V) because more Fe2+ enters into solution resulting in a

cathodic reaction on the mineral surface, hence depressing the mixed potential. The

solution potential measured in HCl solution in the presence of Sn was lower (-0.7 V)

initially but continued to increase until it reached a steady value at -0.5 V after 3

hours of leaching. As leaching progresses the acid dissolved Sn (as Sn(II)) becomes

oxidized and hence less reductive, hence depressing the mixed potential even further.

The Eh value is comparable with the values obtained during synthetic ilmenite

rotating disc dissolution in 7 M HCl in the presence of SnCl2 which indicated that

the dominant species formed in solution was Ti3+ according to the Eh-pH diagram in

Fig. 7.21. The measured ilmenite potential after stabilization was 0.9 V, which

indicated Fe3+ as the dominant species at the surface of the electrode, according to

Fig. 7.21.

265
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Fig. 7.21. Eh-pH diagram for Fe-Ti-H2O system at 25 oC, (excluding Ti2+ ions)
considering TiO2 (c, hydrated), Ti2O3 (c, hydrated) and Fe2O3 as the solid titanium
oxide and Fe(III) phases, and for 0.1 and 0.01 dissolved iron and titanium activities,
respectively (Kelsall and Robbins, 1990).

This result is consistent with that from synthetic ilmenite disc dissolution

under reducing conditions discussed in Chapter 5, which supported the reaction

mechanism proposed in Eqs. 7.7-7.11. Therefore, in the absence of a reducing agent,

the reaction of ilmenite generally follows Eqs. 7.7-7.8. Eq. 7.8 indicating that the

formation of a product layer will inhibit further dissolution of Fe and Ti from the

mineral. In the presence of tin, Sn will readily react with HCl to produce SnCl2 and

nascent hydrogen. This nascent hydrogen forms hydrogen gas and rapidly removed

from the reaction medium. The Sn(II) ions reduce TiOCl2 to TiCl3 and thus prevents

the surface blockage by hydrolysed Ti(IV) products and facilitates the reaction,

according to Eqs. 7.9 - 7.11.

266
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

In the absence of Sn:

FeO.TiO2 + 2HCl → FeCl2 + TiO2 + H2O (7.7)

TiO2 + 2HCl → TiOCl2.H2O (hydrolysed Ti(IV) product) (7.8)

In the presence of Sn:

Sn + 2HCl → SnCl2 + 2H (or H2) (7.9)

2FeCl3 + SnCl2 → 2FeCl2 + SnCl4 (7.10)

2TiOCl2 + SnCl2 + 4HCl → 2TiCl3 + SnCl4 + 2H2O (7.11)

Table 7.12
Potential data measured using platinum electrode and ilmenite electrodes under non-
reductive and reductive conditions.

Conditions Solution Ilmenite


[HCl] T potential potential References
Sample RA (Eh, V) (Em, V)
(M) (oC)
7 60 Iluka - -0.2 0.8 This studya
7 60 Iluka Sn -0.5 0.9 This studya
7 60 SI - 0.3 Chapter 5b
7 60 SI Sn(II) -0.7 Chapter 5b
4 60 ILCP - - 0.8 Chapter 6c
7 60 MI - 0.6 0.6 Chapter 6c
7 60 MI Sn(II) 0.5 0.2 Chapter 6c
RA = reducing agent; SI = synthetic ilmenite; MI = Massive ilmenite
a
Batch leaching of particles
b
Rotating disc dissolution of synthetic ilmenite
c
Open circuit potential of Massive ilmenite electrode

267
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

This study has shown the significant role of a reducing agent in improving

the recovery of Fe or Ti from ilmenite ores using a leaching process which reduces

the energy consumption as discussed in Chapter 2. The latest study published by

Zhang and Nicol (2010) showed that the addition of Ti(III) into the sulfate system

decreased the solution potential and thus increased the dissolution of both titanium

and iron. The authors also suggested that dissolution of ilmenite in the presence of

Ti(III) ions occurred as a result of reduction of iron and oxidation of Ti(III), based on

an electrochemical study (Zhang and Nicol, 2009). Mahmoud et al. (2004) reported

that the addition of a reducing agent during ilmenite leaching caused the reduction of

Ti(IV) to Ti(III), which was the main role of the dissolved iron, as Fe(II), which

accelerated the leaching of ilmenite, as described in Chapter 2. The Sn(II) produced

by the acid dissolution of Sn metal was directly involved in the redox reaction, in

which Sn(II) was oxidised to Sn(IV). The role of Sn(II) was to reduce Fe(III) to

Fe(II) and TiOCl2 to Ti(III). A study on the role of hydrogen produced by the acid

dissolution of Sn metal is beyond the scope of this study.

268
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.6. Residue analysis

The residues of selected ilmenite leaching experiments conducted in the

presence of a reducing agent have been analysed to provide more information on the

reaction products and mechanism. These experiments were conducted at a relatively

high pulp density of about 40.6 g/L (compared to 4 g/L) in 7 M HCl at 110 oC in the

presence of tin (2 g/L), in order to obtain sufficient masses of residues for analysis.

About 5.0 g of residue was collected, separated, washed, dried at 110 oC and

calcined at 900 oC to completely dehydrate the sample. This residue was subjected to

X-ray diffraction (XRD), chemical analysis, scanning electron microscopy (SEM)

and EDX analysis.

The XRD peaks of the Iluka sample before and after leaching are shown in

Fig. 7.22. The XRD pattern of the Iluka ilmenite sample measured before leaching

showed 9 high intensity peaks corresponding to the ilmenite (FeTiO3) phase.

However, there are a few low intensity peaks that correspond to the hematite, rutile

and pseudorutile phases.

269
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Fig. 7.22. XRD pattern of Iluka sample and Iluka leach residues after leaching in
7 M HCl, at 110 oC in the presence of tin (2 g/L), pulp density = 40.6 g/L. I =
ilmenite, R = rutile, P = pseudorutile, H = hematite.

270
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

The chemical analysis of the residue is shown in Table 7.13. The residue

contained about 85.6% TiO2 and only 7% total iron. Higher pulp density resulted in

the precipitation of titanium which remains in the solid residue. A similar behaviour

has been reported by Olanipekun (1999). About 10.5% of Fe2O3 (hematite) still

present in the residue is not dissolved during leaching. Tin was also associated with

the solid residue in very small amounts indicating that the added tin metal powder

was dissolved and some precipitated. Further evidence of titanium enrichment in the

solid residue after leaching at high pulp density in the presence of Sn metal powder

can be seen from the SEM images presented in Fig. 7.23 (a)-(d).

Table 7.13
Chemical analysis of feed and leach residues.

Ilmenite sample (Iluka) Residue


Component
% (w/w) % (w/w)
TiO2 59.1 85.6
FeO 10.3 0.10
Fe2O3 34.6 10.5
Fe Total 24.2 7.34
Al2O3 1.21 0.76
MgO 0.86 0.06
MnO 0.57 0.26
CeO2 0.072 0.002
SiO2 1.12 1.02
SO3 0.05 0.01
SnO2 nil 1.08
Cr2O3 0.466 0.14
P2O5 0.152 0.096
V2O5 0.19 0.06
Nb2O5 0.17 0.245
CaO 0.06 0.01
Moisture 0.36 0.05

271
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Fig. 7.23(a) shows the ilmenite particle prior to leaching. The electron

micrograph depicts the shape of Iluka ilmenite particles which are of almost

homogeneous sizes. After leaching for 2 hours, a few particles have become smaller

due to structural disruption during chemical attack by HCl, but some still retain their

shapes, as shown in Fig. 7.23(b). Fig. 7.23(c) shows the particles of ilmenite after

being leached for 5 hours. Surface irregularities suggested that iron dissolution took

place in the regions where the ilmenite or hematite was located. The image of higher

magnification (Fig. 7.23(d)) shows that some of the particles have been coated by a

white precipitate around the partially leached particles. This could be the titanium

hydroxide precipitated during leaching which inhibited further dissolution of

ilmenite.

(a) (b)
)

(c) (d)
)

Fig. 7.23. SEM images of particles obtained before and after leaching in 7 M HCl at
40.6 g/L pulp density, 2 g/L of tin metal powder and 110 oC (a) before leaching, (b)
after leaching (2 h), (c) after leaching (5 h), (d) after leaching (5 h) at higher
magnification.

272
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

The EDX analyses of the ilmenite feed and leach residue were carried out to

examine the compositional variation between them, as listed in Table 7.14. The %

(w/w) of Ti increased from 24.57% to 61.11% while that of Fe decreased from

29.85% to 7.18%. Beside Fe, the residue contained traces of Mg, Si, Mn, Cl and Al.

No traces of reducing agent (Sn) were detectable on the residues by SEM/EDX.

Table 7.14
EDX analysis of Iluka ilmenite particles before and after leaching.

Elements Composition (%)


Before leaching After leaching for 5 h
Fe 29.85 7.18
Ti 24.57 61.11
O 44.09 30.31
Mg 0.24 nil
Al 0.54 0.11
V 0.1 Nil
Cl nil 1.07
Sn nil nil

273
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.7. Reaction rates and orders

7.7.1. General equations

Curves representing the metal dissolution (%) as a function of time can be

used to determine initial rates as well as to quantitatively analyse rate data based on

suitable kinetic models. For example, initial rates can be used to determine reaction

orders, which in turn can be used to propose reaction mechanisms. In general, a

decrease in reaction rate of particles with time at given concentrations of reagents is

a result of:

(i) The decrease in surface area,

(ii) The build-up of an insoluble product layer on a particle surface which

affects the diffusion of reactants to the surface and products away from

the surface.

The rate data for HCl leaching of synthetic ilmenite under controlled

hydrodynamics (rotating discs) was determined using initial rates, the Levich

equation and the parabolic rate law, as discussed in Chapter 5. It was found that the

reaction rate determining step was governed by the diffusion of H+ through a product

layer on the mineral surface. The reaction scheme of ilmenite leaching in HCl

solutions is complex because of the high acid concentration used and the complex

mineralogy of ilmenite samples. The reactions between ilmenite and HCl change the

concentration of reagents with time and affect the kinetics of ilmenite dissolution.

Therefore, it is more appropriate to use the initial slope at t → 0 which gives the rate

of reaction corresponding to the initial concentration of reagents.

274
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Thus, this method can be used to determine the reaction orders (n) by

relating the initial rate (RM) to the concentration of a reactant, in this case H+,

according to Eqs. 7.12 - 7.16, where [M] is the concentration of dissolved metal in

solutions, X is the dissolved fraction of metal after time t and kap is the apparent rate

constant.

d[M] dX
RM = or = kap[H+ ]n (7.12)
dt dt

The rate RM in this case is a derivative at time zero (dx/dt) t=0. This can conveniently

be obtained by fitting a polynomial equation to the data of reacted fraction (X)

versus time (t) data, as shown in Fig. 7.24. Second or third order (degree)

polynomial regression using Excel was performed to obtain the parameters of the

best fit to the experimental data. The rate (dX/dt) was derived from the Eqs. 7.14 -

7.16 as shown below. Eq. 7.14 was obtained for the curve at 10 M HCl shown in Fig.

7.24:

X = -0.00018t2 + 0.02074t + 0.00969 R² = 0.99 (7.13)

Differentiation with respect to time, t gives:

dX/dt = 2( -0.00018)t + 0.0207 (7.14)

when t = 0 :

dX/dt = 0.0207 min-1 (7.15)

275
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

A plot of log initial rates (RM) at t = 0 against log [Y] when other variables remained

constant resulted in a straight line with a slope (n) which can be related to the

reaction order of reagent Y, as shown in the next section.

Log RM = n log[Y] + log kap (7.16)

0.7
y = -0.0002x2 + 0.0209x + 0.0225
R² = 0.99
0.6 HCl = 9 M
HCl = 10 M
Fraction dissolved (X), Fe

0.5 HCl = 11 M
HCl = 7 M
0.4 y = -0.00018x2 + 0.02074x + 0.00969
R² = 0.99853
0.3

0.2 y = -0.0001x2 + 0.0083x + 0.0173


R² = 0.97
0.1
y = -0.00003x2 + 0.00407x + 0.00560
R² = 0.99
0.0
0 10 20 30 40 50 60 70
Time (min)

Fig. 7.24. Fraction (X) of Fe dissolution with respect to time at different acid
concentration for polynomial fitting (sample = Iluka, [HCl] = 7 M – 11 M HCl, T =
80 oC, pulp density = 4 g/L, size fraction = +53-63 µm).

276
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.7.2. Reaction orders under non-reducing conditions

Fig. 7.25-7.27 show logarithmic plots of (dX/dt)Fe and (dX/dt)Ti as a function

of acid concentration [H+] based on initial rate data given in Table 7.15. The lines of

best fit in Fig. 7.25(a) and Fig. 7.25(b) have slopes of 4.2 and 6.5, which are the

reaction orders for the dissolution of iron and titanium, respectively. The reaction

orders for the Iluka, Tiwest and North Capel samples are shown in Table 7.16, along

with corresponding alteration factor calculated in Table 7.3. In Table 7.16, the

samples including synthetic ilmenite have been listed in the order of increasing

degree of alteration to allow the general trend to be seen more easily. Thus, from the

data given in Table 7.16, the reaction orders of ilmenite leaching in HCl solution

increased with the increase in degree of alteration. This means that extensively

weathered samples undergo more complex reactions compared to the synthetic

ilmenite. Similar reaction order (4th order) was also reported by Haverkamp et al.

(2016) by using the data from Olanipekun (1999) on reaction rate with respect to H+

concentration.

(a) (b)
-1.0 0.0
-1.2 -0.5
y = 4.2x - 5.9
Log dX/dt]Ti (min-1)
Log dX/dt]Fe (min-1)

-1.4 y = 6.5x - 8.5


R² = 0.94 -1.0
-1.6 R² = 0.95
-1.5
-1.8
-2.0
-2.0
-2.2 -2.5

-2.4 -3.0
-2.6 -3.5
0.7 0.8 0.9 1.0 1.1 0.8 0.9 1.0 1.1
Log [H+ ] (mol/L) Log [H+] (mol/L)
Fig. 7.25. Logarithmic plot of initial rates of Iluka sample (a) [dX/dt]Fe and (b)
[dX/dt]Ti as a function of acid concentration ([HCl] = 7, 9, 10, 11 M, pulp density =
4 g/L, T = 80 oC, time = 5 h, size fraction= +53-63 µm)

277
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a) (b)
0.0 0.0

-0.5
Log [dX/dt]Fe (min-1)

Log [dX/dt]Ti (min-1)


-0.5
y = 5.4x - 6.5
y = 3.7x - 4.9 -1.0 R² = 0.99
R² = 0.98
-1.0
-1.5

-1.5
-2.0

-2.0 -2.5
0.6 0.8 1 1.2 0.6 0.8 1 1.2
Log [H+ ] (mol/L) Log [H+ ] (mol/L)

Fig. 7.26. Logarithmic plot of initial rates of North Capel sample (a) (dX/dt)Fe and
(b) (dX/dt)Ti as a function of acid concentration ([HCl] = 7, 9, 10, 11 M, pulp
density = 4 g/L ,T = 80 oC, time = 5 h, size fraction = +53-63 µm).

(a) (b)
0.0 0.0

-0.5
Log [dX/dt]Fe (min-1)

-0.5
Log [dx/dt]Ti (min-1)

y = 3.7x - 4.9
R² = 0.98 -1.0 y = 4.4x - 6.3
R² = 0.99
-1.0 -1.5

-2.0
-1.5
-2.5
-2.0
-3.0
0.6 0.8 1 1.2
0.6 0.8 1 1.2
Log [H+] (mol/L) Log [H+] (mol/L)

Fig. 7.27. Logarithmic plot of initial rates of Tiwest sample (a) (dX/dt)Fe and (b)
(dX/dt)Ti as a function of acid concentration ([HCl] = 7, 9, 10, 11 M, pulp density =
4 g/L ,T = 80 oC, time = 5 h, size fraction= +53-63 µm).

278
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.15
Initial rate data (dX/dt) at various HCl concentrations.

North Capel Iluka Tiwest


HCl (M) dX/dt x 103(min-1) dX/dt x 103(min-1) dX/dtx 103(min-1)
Fe Ti Fe Ti Fe Ti
7 16.6 11.0 4.0 1.0 1.0 2.8
9 36.0 51.0 9.0 4.0 7.0 7.4
10 68.0 78.0 22.0 15.0 16.0 12.6
11 85.0 122.0 23.0 15.0 18.0 20.9
o
([HCl] = 7, 9, 10, 11 M, pulp density = 4 g/L, T= 80 C, t = 5 h, size fraction = +53-
63 µm)

Table 7.16
Reaction orders and degree of alteration.

Ilmenite sample Alteration factor, M nFe nTi


Synthetic ilmenite 1.00 1.4 2.5
North Capel 1.20 3.7 5.3
Iluka 1.58 4.2 6.5
Tiwest 1.90 6.7 4.4

7.7.3. Reaction orders under reducing conditions

The initial rates from Table 7.17 were obtained from Fig.7.16 and 7.17 in the

dX
previous section. A plot of logarithm of initial rate, as a function of the
dt t=0

logarithm of tin concentration, assuming that all the tin was dissolved rapidly and

other variables remained constant, resulted in a straight line with a slope n

representing the reaction order with respect to Sn(II), as shown in Fig. 7.28(a)-(b)

and Fig. 7.29(a)-(b).

279
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a) (b)
-1 -1

-1.5

log [dX/dt] (min-1)


-1.5
log [dX/dt] (min-1)

y = 0.34x - 1.54 y = 0.22x - 1.89


R² = 0.99 R² = 0.99
-2 -2

-2.5 -2.5

-3 -3
-3 -2.5 -2 -1.5 -1 -3 -2.5 -2 -1.5 -1
Log [Sn] (mol/L) Log [Sn] (mol/L)

Fig. 7.28. Initial rate of dissolution of Iluka ilmenite (a) Fe and (b) Ti as a function
of Sn concentration in HCl solutions ([Sn] = 0.5 g/L -4.0 g/L, [HCl] = 7 M, T =
80 oC, time = 5 h, size fraction = +53-63 µm).

(a) (b)
-1 -1

-1.5
-1.5
Log [dX/dt] (min-1)
Log [dX/dt] (min-1)

y = 0.45x - 1.33 y = 1.09x - 0.39


R² = 0.97 -2 R² = 0.97
-2
-2.5

-2.5
-3

-3 -3.5
-3 -2 -1 -3 -2.5 -2 -1.5 -1
Log [Sn] (mol/L)
Log [Sn] (mol/L)
Fig. 7.29. Initial rate of dissolution of Tiwest ilmenite (a) Fe and (b) Ti as a function
of Sn concentrations in HCl solutions ([Sn] = 0.5 g/L - 4.0 g/L, [HCl] = 7 M, pulp
density = 4 g/L,T = 80 oC, time = 5 h, size fraction= +53-63 µm).

280
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

All the initial reaction rates (dX/dt) and reaction orders (n) obtained under

reducing conditions for the leaching of natural and synthetic ilmenite rotating discs

under reducing conditions described in the previous chapter are listed in Table 7.17.

The reaction order for the Iluka sample with respect to tin (0.2-0.3) was

comparable with that for synthetic ilmenite (0.3-0.4), as listed in Table 7.17.

However, the Tiwest sample showed higher reaction orders of 0.5 and 1.0 for Fe and

Ti, respectively. This indicated that the Tiwest sample was more strongly affected by

the presence of Sn as a reducing agent compared to synthetic ilmenite and the Iluka

sample. This was also supported by higher reaction rates of Fe and Ti dissolution

from the Tiwest sample compared to those from the Iluka sample.

Table 7.17
Initial reaction rates (dX/dt) and reaction orders (n) with respect to Sn(II).

HCl (7 M) Synthetic ilmenitea


Iluka Tiwest
dX/dt x 105
SnCl2 dX/dt x 103(min-1) dX/dt x 103(min-1)
Sn (g/L) (mol m-2 s-1)
(g/L) Fe Ti Fe Ti Fe Ti
4.0 1.0 1.0 2.8
0.0
0.5 4.5 3.8 4.1 1.0
1.0 5.9 4.4 5.2 2.0
2.0 7.0 5.2 8.3 6.0
4.0 9.3 6.1 10.0 8.9
0.0 0.8 0.9
0.5 3.8 2.3
1.0 5.0 3.3
2.0 6.0 4.1
Reaction order, n 0.3 0.2 0.5 1.1 0.3 0.4
Initial rates (R) with respect to Fe and Ti dissolution based on graphical method (see
Fig. 7.24) at 80 oC under reducing conditions.
a
from Chapter 5

281
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

In general a half order reaction with respect to a reagent corresponds to an

electrochemical reaction (Nicol and Lazaro, 2002), which is also consistent with

other findings. In recent studies, kinetic models have been developed to rationalize

reaction orders in the range of 0.2-1.5 with respect to H+ ions for the acid dissolution

of various oxides/hydroxides/silicates assuming the interdependence of parallel

processes of the removal of material in the form of ions from the minerals surfaces

resulting in the dissolved cations and anions in solutions where in some cases

charged surfaces are involved due to the conditions which are far from equilibrium

and involve H+ as well as water molecules (Crundwell, 2016).

7.8. Activation energy

dX
Initial rates for different temperatures were calculated from the plots
dt t=0

in Figs. 7.10 – 7.12 and are given in Table 7.18. The activation energy can be

derived from the Arrhenius plot of Ln dX/dt as a function of -1000/RT as shown in

Fig. 7.30(a-f). The calculated values of activation energy are tabulated in Table 7.19

which also includes the values determined for synthetic ilmenite rotating disc and

other published data.

282
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a) (b)
0 0
Fe -1
-2 -2 Ti

Ln dX/dt
-3
Ln dX/dt

-4 -4
-5 y = 71.9x + 20.19
y = 90.08x + 25.85
-6 R² = 0.97
R² = 0.97 -6
-7
-8 -8
-0.35 -0.33 -0.31 -0.35 -0.33 -0.31
-1000/RT (kJ mol-1) -1000/RT (kJ mol-1)
(c) (d)
0 0

Ti y = 109.4x + 30.96
-2 Fe -2 R² = 0.99
y = 62.32x + 15.09
Ln dX/dt

Ln dXdt
R² = 0.97
-4 -4

-6 -6

-8 -8
-0.35 -0.33 -0.31 -0.35 -0.33 -0.31
-1000/RT (kJ mol-1) -1000/RT ] (kJ mol-1)
(e) (f)
0 0
y = 83.77x + 22.41
y = 39.31x + 8.114
Fe Ti R² = 0.96
-2 R² = 0.99 -2
Ln dX/dt
Ln dX/dt

-4 -4

-6 -6

-8 -8
-0.35 -0.33 -0.31 -0.35 -0.33 -0.31
-1000/RT (kJ mol-1) -1000/RT (kJ mol-1)

Fig. 7.30. Arrhenius plot for Fe and Ti dissolution on ilmenite samples (a) - (b)
North Capel , (c) - (d) Iluka, (e) - (f) Tiwest ([HCl] = 7 M, pulp density = 4 g/L, T =
80 oC, 90 oC, 100 oC, 110 oC, time = 5 h, size fraction = +53-63 µm).

283
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.18
Initial rate data at different temperatures.

North Capel Iluka Tiwest


o dX/dt x 103 (min-1) dX/dt x 103 (min-1) dX/dt x 103 (min-1)
T ( C)
Fe Ti Fe Ti Fe Ti

80 9.0 13.0 2.0 2.0 3.0 2.0

90 14.0 26.0 4.0 4.0 10.0 14.0

100 46.0 61.0 8.0 14.0 11.0 14.0

110 88.0 82.0 10.0 34.0 12.0 15.0

Table 7.19
Calculated activation energy values.

Activation energy
Conditions Material
(kJ/mol)
Reference
[acid] Temperature
Acid Ilmenite ore Fe Ti
(M) range (oC)
HCl 7 70-110 North Capel 71.9 90.1 Current study

HCl 7 70-100 Iluka 62.3 109.4 Current study

HCl 7 70-100 Tiwest 39.3 83.8 Current study


Synthetic
HCl 7 50-80 44.2 75.3 Current study
Ilmenite
Nigerian Olanipekun,
HCl 7.2 70-90 67.1 62.4
ilmenite 1999
Tsuchida et al.,
HCl 11 30-50 Ilmenite 73.2 81.2
1982
Jackson and
HCl 6 60-90 Allard lake 88.3 92.5 Wadsworth,
1976
Zhang and
H2SO4 4.6 85-100 Iluka 75.0
Nicol, 2010

284
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

The observed apparent activation energies are all in the same order of

magnitude and there is no obvious trend in the values with composition or degree of

alteration. Olanipekun (1999) obtained activation energy of 67.1 kJ mol-1 for

titanium and 62.4 kJ mol-1 for iron for the mixed chemical-diffusion controlled

reaction in hydrochloric acid leaching in the temperature range from 70 oC to 80 oC.

A similar high activation energy (75 kJ mol-1) was reported by Zhang and Nicol

(2010) for the dissolution of ilmenite in 450 g/L H2SO4 under mixed chemical-

diffusion controlled reaction conditions.

Fig. 7.31a, b plot the composition of the three oxides (TiO2, FeO and Fe2O3)

in the four ilmenite samples (SI, NC, IL and TW) as % (w/w) and mol/100 g as a

function of the alteration factor. Whilst the TiO2 content in all samples is larger than

FeO and Fe2O3, the FeO content is lower at a higher alteration factor. The lower

activation energy of iron leaching, compared to titanium leaching (Fig. 7.31c)

supports a mixed-diffusion controlled reaction for iron leaching. This warrants

further studies described in next section.

285
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

(a)

(b)

(c)

Fig. 7.31. Effect of alteration factor on (a) oxide composition (% w/w) in ilmenite
(b) mol/100g ilmenite and (c) activation energy for the leaching of four samples (SI,
NC, IL and TW) (Data from previous figures).

286
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.9. Kinetic models

7.9.1. Shrinking sphere and core models

The well-known kinetic models such as the shrinking sphere and the

shrinking core models for the dissolution of sized particles under the assumption that

the particles are homogeneous and spherical (Levenspiel, 1972) have been widely

used to rationalise the kinetics of mineral dissolution. In the shrinking sphere model,

the rate is governed by a chemical reaction at the surface of the reacting particles,

while in the shrinking core model the rate is governed by mass transport through a

thickening product layer on the surface. The models can be expressed as follows in

the form of integrated rate equations for the reaction with H+ ions (Levenspiel, 1972;

Senanayake and Das, 2004; Senanayake et al., 2015) for the reaction A(aq) + bB(s) =

products:

For the shrinking particle model:

1-(1-X)1/3 = b[H+]bulkkr-1ρ-1t = ksst (7.17)

For the shrinking core model:

1 - 3(1-X)2/3 + 2(1-X) =6 b [H+]bulk DH+ r-2ρ-1t = kplt (7.18a)

1 - 3(1-X)2/3 + 2(1-X) =6 b [H+]bulk DH+ r-2 {(1-ε) ρ}-1t = kplt (7.18b)

287
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

where X = the fraction dissolved at time t, kss, kpl = apparent rate constant, [H+] =

concentration of acid in bulk solution (mol cm-3), DH+= diffusion coefficient of H+

through the product layer (cm2 s-1), r = initial particle radius (cm), ρ= molar density

of the dissolving metal in the initial particle (mol cm-3), ε = particle porosity, b =

stoichiometric factor.

Examination of plots based on the equations and the two expressions on the

left hand side of Eqs. 7.17 and 7.18 as a function of time for the Iluka ilmenite

sample showed that only Eq. 7.18 (shrinking core model) gave a straight line (for the

time period of 0 to 300 min) with a higher correlation (R2≈1) compared to Eq. 7.17

(shrinking particles) as shown in Fig. 7.32, both for Fe and Ti dissolution. Similar

behaviour was observed for the Tiwest and North Capel samples. The slopes of the

linear relationships of these plots provide a value for kpl.

Fig. 7.32. Validity of shrinking core model for Fe (A) and Ti (B) for the leaching
results (Iluka ilmenite sample, in 7 M HCl at 80 oC with pulp density of 4 g/L).

288
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.9.2. Factors affecting apparent rate constants

Figs. A4.2 -A4.9 in Appendix A4.2 show the fitting of the shrinking core

model for North Capel, Iluka and Tiwest ilmenite leaching in HCl solution with

concentration in the range 7 M to 11 M in the absence or presence of a reducing

agent (tin metal powder) and at different temperatures from 80 oC to 110 oC. All the

rate constants (kpl values) obtained from the slopes are listed in Table 7.20. All the

kpl values obtained from the slopes are further examined by considering the relative

increase in kpl due to the increase in acid concentration, temperature or Sn addition.

Relative increase is the ratio of each kpl value divided by the corresponding baseline

kpl value at the lowest acid concentration (7 M HCl), lowest temperature (80 oC) or

lowest Sn addition (0 g/L). The relative values are also plotted in Fig. 7.33.

The general trend is that the relative kpl values increase with increasing HCl

concentration, temperature and addition of tin, but reached a plateau in some cases.

In general the effect of these variables on the relative kpl values of Fe and Ti leaching

is similar and follows the same trend (Fig. 7.33b1, b2, c1). The effect of the increase

in acid concentration on relative kpl of Tiwest much larger in the case of the IL and

TW samples compared to the NC sample (Fig. 7.33 a1, b1, c1). The effect of

temperature on relative kpl is much larger for Fe dissolution than Ti from the Tiwest

sample (Fig. 7.33 c2). The effect of tin addition is also larger in the case of Fe

compared to that of Ti for the Tiwest sample (Figs. 7.40 c3) but opposite was

observed for the Iluka sample (Fig. 7.40 b3).

289
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.20
Summary of effect of variables on apparent rate constants and diffusivity in HCl leaching of ilmenite samples.

T HCl kpl x103 DH+ x 1010


Variable Sn Relative kpla
Sample (oC) (M) (min-1) (cm2s-1)
(g/L)
Fe Ti Fe Ti Fe Ti
NC 80 7 0 1 1 1.0 1.0 0.2 0.2
Acid
concentration NC 80 9 0 8 10 8.0 10.0 1.7 1.7

NC 80 10 0 10 14 10.0 14.0 2.1 2.1

NC 80 11 0 11 24 11.0 24.0 2.1 3.3

IL 80 7 0 0.1 0.04 1.0 1.0 0.02 0.01

IL 80 9 0 0.7 0.2 7.0 5.0 0.08 0.4

IL 80 10 0 2.7 0.8 27.0 20.0 0.3 1.5

IL 80 11 0 3.3 1.9 33.0 47.5 0.4 3.4

TW 80 7 0 0.09 0.1 1.0 1.0 0.01 0.03

TW 80 9 0 0.48 0.7 5.3 7.0 0.05 0.1

TW 80 10 0 1.69 1.69 18.8 16.9 0.2 0.4

TW 80 11 0 4.17 4.17 46.3 41.7 0.4 0.8

NC 80 7 0 2 1 1.0 1.0 0.4 0.2


Temperature
NC 90 7 0 6 4 3.0 4.0 1.3 0.9

NC 100 7 0 8 12 4.0 12.0 1.7 2.6

NC 110 7 0 16 15 8.0 15.0 3.4 3.3

IL 80 7 0 0.1 0.04 1.0 1.0 0.02 0.01

IL 90 7 0 0.7 0.2 7.0 5.0 0.1 0.3

IL 100 7 0 2 0.8 20.0 20.0 0.3 0.1

IL 110 7 0 5.2 1.9 52.0 47.5 0.8 0.3

TW 80 7 0 0.1 0.1 1.0 1.0 0.01 0.03

TW 90 7 0 0.7 0.4 7.0 4.0 0.1 0.1

TW 100 7 0 2 0.9 20.0 9.0 0.3 0.2

TW 110 7 0 5.2 1.1 52.0 11.0 0.8 0.3

NC 80 7 0 1 1 1.0 1.0 0.2 0.2

NC 80 7 1.0 3 4 3.0 4.0 0.6 0.9


Sn addition IL 80 7 0 0.1 0.04 1.0 1.0 0.02 0.01

IL 80 7 0.5 0.6 0.3 6.0 7.5 0.09 0.05

IL 80 7 1 1.3 1.3 13.0 32.5 0.2 0.2

IL 80 7 2 1.6 1.6 16.0 40.0 0.3 0.3

IL 80 7 4 2.2 1.5 22.0 37.5 0.2 0.2

TW 80 7 0 0.09 0.1 1.0 1.0 0.01 0.03

TW 80 7 0.5 0.4 0.2 4.4 2.0 0.05 0.05

TW 80 7 1 1.4 0.9 15.6 9.0 0.2 0.3

TW 80 7 2 2.7 1.6 30.0 16.0 0.4 0.5

TW 80 7 4 3.1 1.9 34.4 19.0 0.5 0.6


a
Relative increase is the ratio of each kpl value divided by the corresponding baseline value at lowest acid concentration
(7 M HCl), lowest temperature (80 oC) or lowest Sn addition (0 g/L). DH+ values were calculated assuming ε = 0 and b = 4.

290
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

The general observation is that the relative effect of the variables is larger in

the case of the Tiwest and Iluka samples, compared to the less altered North Capel

sample. Moreover, in the case of the relatively less altered North Capel sample, the

effect of HCl concentration on relative kpl is larger for Ti compared to Fe (Fig. 7.33,

a1, a2). These relative effects can be related to the reaction of H+ and Cl- with the

surface oxides to dissolve them and the role of these ions on enhancing the porosity,

ε in Eq. 7.18b. This can be further examined by considering the diffusivity of H+, for

which the calculated values are listed in Table 7.20 and discussed later.

NC (a1) NC (a2) NC (a3)

IL (b1) IL (b2) IL (b3)

TW (c1) TW (c2) TW (c3)

Fig. 7.33. Effect of leaching parameters on relative kpl values (Data from Table 7.20)
(a) North Capel, (b) Iluka, (c) Tiwest.

291
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.9.3. Effect of particle size

The effect of particle size on the dissolution of Fe and Ti from the North

Capel, Iluka and Tiwest samples and the corresponding shrinking core rate constant

(kpl) were also investigated. The plot of kpl values as a function of reciprocal square

of the initial radius (1/r2) shows that kpl increases with the increase of 1/r2, as

expected from the Eq. 7.19 (all the kpl versus 1/r2 plots are shown in Appendix A4.3,

Figs. A4.10 -A4.12 and kpl data are listed in Table A4.2).

However, a large increase was not observed as for the kpl value of Fe

dissolution from North Capel and Ti dissolution from Iluka and Tiwest sample. In

this case it could be due to the particle being highly porous or weathered, in which

case the shrinking core rate constant is independent of particle size. This is the

reason for which in some cases the particle size did not affect the dissolution rate, as

stated by previous researchers (van Dyk et al., 2002; McConnel, 1978).

7.9.4. Diffusion coefficient of H+

The diffusion coefficient of H+ (DH+) was calculated using Eq. 7.19 and 7.20

with the substitution of the rate constant (kpl) obtained from the slopes of the

shrinking core plots (Appendix A4.2) listed in Table 7.20:

DH+ = (kpl ρ ro2)/(6 b[H+]) (7.19)

292
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Assuming b = 4, where r is the initial particle radius = 0.0029 cm; [H+] = 0.007 mol

cm-3 and ρ (molar density in mol cm-3) is related to weight percent of Fe or Ti in the

ilmenite sample and its density, ρm = 4.5 g cm-3 (Zhang and Nicol, 2009), the value

of ρ was calculated using Eq. 7.20:

ρ = ρm x % Fe or Ti in material / (100 x molar mass of Fe or Ti) (7.20)

For North Capel sample, Fe in sample = 31.5 %, Fe molecular weight = 56.0 g mol-1

Therefore, ρ = 4.5 g cm-3 x 31.5 / (100 x 56.0 g mol-1) = 0.0253 mol cm-3

The substitution of the parameter given above into Eq. 7.19 and 7.20 gives the

following DH+ values calculated from data obtained for the North Capel sample in

7 M HCl at 80 oC (Details calculation of DH+ values are given in Appendix A4.4) :

DH+(Fe-leaching) = 2.1 x 10-11 cm2 s-1

DH+(Ti-leaching) = 2.1 x 10-11 cm2 s-1

DH+ values calculated from data obtained for the North Capel, Iluka and Tiwest

samples of different acid concentration, temperature and in the presence of a

reducing agent are also listed in Table 7.20. As can be seen from Table 7.20, DH+

values of Fe and Ti from North Capel, Iluka and Tiwest samples (particle leaching)

in the absence and presence of Sn(II) are of the order 10-10 to 10-12 cm2s-1.

The DH+ values obtained in this study are compared with those from others

studies listed in Table 7.21. As described in Chapter 5, rate of dissolution of

synthetic ilmenite in HCl solutions obeyed the parabolic rate law, indicating that the

diffusion of reagent through the solid layer formed on the surface was the rate

controlling step. The constant of square concentration (4 M - 9 M HCl) with respect

293
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

to time was proportional to the diffusion coefficient (D) value, which was calculated

to be 1.3 x 10-10 cm2 s-1 for Fe and 0.3 x 10-10 cm2 s-1 for Ti. The DH+ values from

dissolution of rotating disc of synthetic ilmenite are comparable with DH+ values in

this chapter (ranging from 0.1 x 10-11 – 3.0 x 10-10 cm2 s-1 (Fe) and 0.1 x 10-11 –

3.0 x 10- 10 cm2 s-1(Ti)) at leaching conditions of [HCl] = 7 M – 11 M and T = 80 -

110 oC. In the presence of reducing agent (HCl] = 7 M; T = 80 oC; [Sn] = 1.0 – 4.0

g/L), the diffusion coefficient of H+ increased one-fold of magnitude from 10-11 to

10-10. For North Capel sample the DH+ values are also in the range of 10-11 - 10-10

cm2 s-1 at all parameters studied (Table 7.20) but higher than DH+ values for Iluka

and Tiwest sample. Higher values of DH+ in North Capel sample compared to other

samples is a result of the low degree of alteration. Thus, it can be stated that at more

favourable conditions of higher acid concentration, temperature and in the presence

of a reducing agent, the production of the residual layer that formed on the surface of

particles in the form of a white precipitate could be minimized hence accelerating the

dissolution of iron and titanium.

In addition, other DH+ values previously reported for the diffusion of H+

through solids such as Fe3O4 (Allen et al., 1979) are comparable with DH+ in this

study at high end of a range. The only reported values of DH+ for ilmenite (1 x 10-9

(Fe) and 2 x 10-9 (Ti)) are from the work conducted by Tsuchida et al. (1982). They

reported that the surface chemical reaction at temperatures above 60 oC became so

fast that the dissolution rates were limited by the diffusion in the residual layer.

Other reported values for solids such as for -MnO2 (Allen et al. 1979) and laterite

(Senanayake and Das, 2004) were one or two-fold higher than DH+ values obtained

for ilmenite in this study. Much higher values were reported for the diffusion of H+

294
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

in aqueous solution and for single crystal iron as listed in Table 7.21 (Bockris et al.,

1977).

Table 7.21
Dependence of proton diffusion coefficient on medium and materials.

Diffusion coefficient
Materials/Medium References
DH+ (cm2 s-1)
Aqueous solutiona 9 x 10-5 Bockris and Reddy, 1977
Single crystal ironb 8 x 10-5 Bockris and Reddy, 1977
Allen et al., 1979
c -10
Fe3O4 8 x 10 (Electrochemical-Carbon
paste electrode)
FeTiO3d 1 x 10-9 (Fe); 2 x 10-9 (Ti) Tsuchida et al., 1982
-MnO2 8 x 10-8 Allen et al., 1979

Product layer in laterite Senanayake and Das, 2004


0.5 x 10-9 to 4 x 10-9
leaching (acid leaching)
FeTiO3e 1.3 x 10-10 (Fe)
This study (Chapter 5)
(SI, parabolic rate law) 0.3 x 10-10 (Ti)
FeTiO3f
0.01 x10-10 – 3.0 x 10-10 (Fe)
(NC, IL, TW, based on This study (Table 7.20)
0.01 x10-10 – 3.0 x 10-10 (Ti)
particle leaching
FeTiO3g
(NC, IL, TW, in the 0.09 x 10-10 – 0.6 x 10-10 (Fe)
This study (Table 7.20)
presence of Sn, based 0.01 x 10-10 – 0.9 x 10-10 (Ti)
on particle leaching)
a
At infinite dilution and 25 °C.
b
Diffusion of electronated H+(atomic hydrogen) into the metal
c
Minimum possible value based on electrochemical studies with polycrystalline magnetite.
d
Ilmenite leaching with HCl = 11 M, pulp density = 25 g/L, grain size = 60-100 µm.
e
Synthetic ilmenite (SI) samples based on parabolic law equation (HCl= 4 M -7 M, T=
60 °C
f
Ilmenite leaching DH+ based on shrinking core model (pulp density = 4 g/L, size fraction =
+53-63 µm, ρ = 4.5 g/cm3, density reported by Zhang and Nicol 2009)
g
Ilmenite leaching DH+ based on shrinking core model in the presence of reducing agent (Sn
= 0.5 g/L – 4.0 g/L, pulp density = 4 g/L, size fraction = +53-63 µm, ρm = 4.5 g/cm3 (Zhang
and Nicol 2009)]

295
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

7.10. Summary and conclusion

The dissolution behavior of natural Australian ilmenite in HCl was found to

be related to the degree of alteration of the ores. The leaching efficiency of ilmenite

followed the descending trend North Capel > Iluka > Tiwest, Tiwest being the most

weathered sample, followed by Iluka and North Capel, in ascending order. The

results on overall dissolution of Fe and Ti are presented in Table 7.22. At high acid

concentration (11 M HCl) and high temperature (110 oC), almost 100% of iron and

titanium in the North Capel sample was dissolved into solution after 5 hours. In the

presence of a reducing agent, the dissolution of iron and titanium increased

significantly for all samples used in this study. In the case of the North Capel

sample, the maximum dissolution was achieved after 2 hours under reducing

conditions. The use of a reducing agent allows the use of lower acid concentration,

temperature and less time to obtain maximum extraction of Fe and Ti. The

activation energy for reaction of the different samples tested agrees well with the

reported values between 40 kJ mol-1 and 110 kJ mol-1. The reaction orders of

ilmenite dissolution with respect to [H+] showed higher values for samples with a

higher alteration factor. The reaction order of ilmenite dissolution with respect to

concentration of reducing agent (Sn) was around 0.5 and agrees well with results

obtained from the dissolution of synthetic ilmenite disc. The role of Sn appears to be

the reduction of Fe(III) oxides to Fe(II) and TiO2 to Ti(III). The values of diffusion

coefficient calculated on the basis of a shrinking core kinetic model are in the range

of 10-10 -10-11 cm2 s-1 and comparable with values from previously published work.

296
Chapter 7: Reductive and non-reductive leaching of natural ilmenite

Table 7.22
Maximum leaching efficiency of iron and titanium from ilmenite under different
experimental conditions.

Leaching time
Samples HCl (M) T(ºC) RA Fe (%) Ti (%)
(h)
NC 11 80 2 - 99.5 100
IL 11 80 5 - 86.5 59.9
TW 11 80 5 - 85.8 93.7
NC 7 110 2 - 100 100
IL 7 110 5 - 94.3 42.4
TW 7 110 5 - 96.9 61.1
NC 7 80 5 Fe 76.1 81.2
IL 7 80 5 Fe 44.1 25.8
TW 7 80 5 Fe 22.4 37.4
NC 7 80 3 Sn 92.2 100
IL 7 80 5 Sn 97.0 89.0
TW 7 80 5 Sn 100 83.7
Other parameters: particle size fractions = +53-63 µm; pulp density = 4 g/L in all
cases. RA = reducing agent or reductant;

For rapid and effective dissolution of ilmenite in chloride systems, it is

necessary to reduce TiO2+ to Ti(III) to inhibit the formation of a product layer and on

the surface of ilmenite and Fe(III) oxides formed by weathering to Fe(II) and also

enhance the porosity and hence the diffusivity of H+. Adding a reducing agent into

the system increased the rate of iron and titanium dissolution without using high

temperatures or HCl concentrations. Further work on chemical and electrochemical

kinetic modeling and product identification is essential for shedding more light on

the reaction mechanism.

297
Chapter 8: Summary, conclusions and future work

CHAPTER 8

8. SUMMARY, CONCLUSIONS AND FUTURE WORK

8.1. Summary and conclusions

This thesis studied the chemical and electrochemical dissolution of synthetic and

natural ilmenite carried out using flat surfaces in the form of rotating discs of carbon-

paste or massive ilmenite electrodes. Batch leaching studies of ilmenite particles of

three samples of different composition in HCl solution were also conducted to examine

the effect of acid concentration, temperature and particle size. In both cases the

experiments were conducted under both non-reducing and reducing conditions. The

feed materials and products were characterized using standard techniques such as XRD,

SEM, EDX and optical microscopy.

In the case of a rotating synthetic ilmenite disc, it was found that under non-

reducing conditions, the dissolution rate was controlled by the diffusion of reactants

through a product layer that formed on the surface of synthetic ilmenite. The initial rates

of dissolution of iron and titanium show reaction orders with respect to concentration of

HCl of 1.4 and 2.5, respectively, as a result of the involvement of both H+ and Cl- ions in

the complex surface reaction. The high apparent activation energy agrees well with the

literature data.

298
Chapter 8: Summary, conclusions and future work

Results from the rotating disc studies of synthetic ilmenite in HCl in the
o o
temperature range of 50 C to 80 C yielded apparent activation energies of

44.1 kJ mol-1 for iron and 75.6 kJ mol-1 for titanium, indicating that the iron dissolution

is mixed chemical-diffusion controlled. The linear relationship between the square of

dissolved iron and titanium concentrations and time confirmed the applicability of the

parabolic rate law. It also indicated that the diffusion of reagent(s) through the insoluble

product layer that formed on the surface of synthetic ilmenite was the rate controlling

step. A white precipitate formed on the surface of the disc appeared to be a hydrolysis

product of TiOCl2. The linear relationship between the square of dissolved iron and

titanium concentrations as a function of time with higher slopes in the presence of SnCl2

indicated the reductive role of Sn(II). The Sn(II) ions reduced TiOCl2 (insoluble

product) to TiCl3 (soluble) and thus prevented the surface blockage by hydrolysed

Ti(IV) products. In this case Sn(II) facilitated the reductive leaching of synthetic

ilmenite at moderate HCl concentration and temperatures over 50 oC, which enhanced

the rates by a factor of 20 compared to the rates in HCl alone.

The electrochemical studies were conducted on massive, natural and synthetic

ilmenite. Based on open circuit potentials and cyclic voltammetric analysis, the

behaviour of natural ilmenite in HCl was found to be different to that in H2SO4. In HCl

solutions, iron was selectively dissolved, while in H2SO4 both iron and titanium were

dissolved from the lattice. Moreover, HCl promoted dissolution with rates higher than

those in H2SO4.

299
Chapter 8: Summary, conclusions and future work

The voltammetric characteristics of ilmenite were affected by the degree of alteration

and weathering of the natural ilmenite. Hematite mineral is generally associated with

weathered ilmenite. The hematite phase in the mineral appeared to have a higher

dissolution rate under cathodic conditions than the ilmenite itself.

The dissolution rate of ilmenite in HCl at 60 oC was low at potentials above

0.1 V, where the free dissolution of ilmenite and the reductive dissolution of hematite

were the main reactions. In the presence of a reducing agent (Sn(II)), the dissolution of

ilmenite increased significantly due to reduction of titanium(IV) in the ilmenite solid

phase to titanium(III) in solution.

Leaching experiments of ilmenite concentrates from various locations in

Western Australia collected over a period, referred to as North Capel, Iluka and Tiwest

samples were conducted under different conditions to examine the effect of variables

such as particle size, HCl concentration, temperature and reducing agents. The three

ilmenite samples gave different results of leaching efficiency of iron and titanium. It

was found that the leaching efficiency followed the order North Capel > Iluka > Tiwest,

i.e. it was inversely proportional to the degree of alteration of ilmenite due to

weathering. A higher degree of alteration resulted in a lower leaching efficiency of iron

and titanium in HCl solutions. Under non-reducing conditions, the leaching efficiencies

of all ilmenite samples were directly proportional to acid concentration and temperature,

and inversely proportional to particle size.

300
Chapter 8: Summary, conclusions and future work

The North Capel ilmenite sample, having the least altered mineral, yielded

100% of Fe and Ti dissolution after 5 hours at high acid concentration (11 M HCl) and

high temperature (110 oC). The other samples, from Tiwest and Iluka, yielded a lower

leaching efficiency. The reducing environment created by the presence of metal powder

(M = Sn or Fe), which produces M(II), resulted in higher leaching efficiency of iron and

titanium.

The beneficial effect of Fe metal powder was lower than that of Sn, but in both

cases the leaching efficiencies followed the same order as before: North Capel > Iluka >

Tiwest. The leaching efficiency of iron and titanium increased by a factor of 4-10

compared to leaching under non-reducing conditions. This higher leaching efficiency

was achieved at lower acid concentrations and temperatures than in the absence of the

reducing agent. This indicated the beneficial effect of using reducing agents, which

could lower the acid and energy consumption.

The initials rates of ilmenite dissolution in hydrochloric acid showed reaction

orders higher than 1 with respect to H+ concentration, which increased with increasing

degree of alteration. The altered ilmenite samples have more complex dissolution

reactions compared to less altered ilmenite. Approximately half-order reaction rates

with respect to Sn(II) concentration were observed for synthetic/natural ilmenite,

indicating the electrochemical nature of the reaction. Ilmenite dissolution in

hydrochloric acid has activation energy of 40 – 70 kJ mol-1 for Fe and 85-110 kJ mol-1

for Ti in the temperature range from 80 oC to 110 oC. This suggests the mixed

chemical/diffusion controlled nature of the surface reaction.

301
Chapter 8: Summary, conclusions and future work

Ilmenite dissolution in hydrochloric acid solution conducted in batch leach tests

appeared to obey a shrinking core model in which the diffusion of reactant (H+) was the

rate determining step. The apparent rate constant, kpl, increased with the increase in HCl

concentration and temperature for all three samples. The kpl values also showed a

remarkable increase in the presence of Sn(II).

The values of diffusion coefficient, DH+, calculated using the shrinking core

model for the dissolution of iron and titanium from North Capel, Iluka and Tiwest

samples, were in the range of 10-11 to 10-9 cm2 s-1. These values are comparable with the

reported values for ilmenite and other solids, but further work is essential on

electrochemical, and chemical kinetic modelling and product identification.

Nevertheless, this work has shown that the different reactivities of ilmenite

samples of different origin from different locations is a result of their degree of

alteration due to natural weathering. It can generally be concluded that the dissolution of

ilmenite in HCl initially involves the removal of an altered surface layer rich in ferric

iron in the form of fine grained hematite followed by attack upon the bulk of the solid.

The slower rates observed were due to the involvement of product layers which can be

removed under reducing conditions to facilitate leaching.

302
Chapter 8: Summary, conclusions and future work

8.2. Recommendations for future work

A more extensive theoretical study to explain the reaction orders is essential,

which takes into account the theory of concentrated electrolytes. This should also

include a more extensive study of the relationship of actual solution/surface speciation,

Eh-pH relationships and reaction mechanism of the dissolution of ilmenite under

reducing conditions in order to improve the leaching rate of ilmenite by using different

reducing agents, based on further work on electrochemical and chemical kinetic models

and product identification. .

Extension of this work to higher solid-liquid ratio and reagent concentrations is

also essential to determine optimum condition for ilmenite dissolution.

303
References

REFERENCES

Adriamanana, A., Lamache, M., Bauer, D., 1984. Etude Electrochimique de Differentes
ilmenites. Electrochimica Acta 29 (8), 1052-1054.

Allen, P.D., Bignold, G.J., Hampson, N.A., 1979. The electrodissolution of magnetite:
Part 1. The electrochemistry of Fe3O4/C disc-potentiodynamic experiments.
Journal of Electroanalytical Chemistry 99, 299-309.

Bard, A.J., Faulkner, L.R., 2001. Electrochemical methods fundamentals and


applications. John Wiley & Sons, INC.

Barksdale, J., 1966. Titanium, its occurrence, chemistry and technology. Ronald Press
Company, New York.

Barth, T.F.W., Ponsjak, E., 1934. The crystal stucture of ilmenite. Zeitschrift fr
Kristallographie A88, 265-270.

Barton, A.F.M., McConnel, S.R., 1978. Rotating disc dissolution rates of ionic solids
Part 3. Natural and synthetic Ilmenite. 971-982.

Barton, A.F.M., McConnel, S.R., 1979. Dissolution Behaviour of Solids: The Rotating
Disc Method. Chemistry in Australia 46 (10), 427-433.

Baxter, J.L., 1977. Heavy Mineral Sand Deposits of Western Australia. Geological
Survey of Western Australia.

Becher, R.G., Canning, R.G., Goodheart, B.A., Uusna, S., 1965. A new process for
upgrading ilmenitic mineral sands, AusIMM. Australasian Institute of Mining
and Metallurgy, pp. 21-44.

Bedinger, G.M., 2015. Titanium mineral concentrate, Mineral Commodity Summaries,


U.S Geological survey, pp. 170-171.

Bedinger, G.M., 2016. Titanium mineral concentrate, Mineral Commodity Summaries,


U.S Geological survey, pp. 178-179.

304
References

Bockris, J.O' M., Reddy, A.K.N., 1977. Modern Electrochemistry. Plenum, New York,
pp. 1328-1330.

Burkin, A.R., 2001. Chemical hydrometallurgy: Theory and Principle. Imperial College
Press, London.

Chernet, T., 1999. Applied minerological studies on Ausralian sand ilmenite


concentrate with special reference to its behaviour in the sulphate process.
Minerals Engineering 12, 485-495.

Crundwell, F.K., 2016. The mechanism of dissolution minerals in acidic and alkaline
solution: Part IV a molecular viewpoint. Hydrometallurgy 161, 34-44.

Das, G.K., Pranolo, Y., Zhu, Z., Chen, C.Y., 2013. Leaching of ilmenite ores by acidic
chloride solutions. Hyrometallurgy 133, 94-99.

Deysel, K., 2007. Leucoxene study: a mineral liberation analysis (MLA) investigation,
The 6th International Heavy Minerals Conference 'Back to Basics'. The
Southern African Institute of Mining and Metallurgy, South Africa, pp. 167-
172.

Dimitrios, F., Guillaume, H., 2009. Iron removal and recovery in the titanium dioxide
feedstock and pigment industry, ProQuest Science Journal.
www.tms.org/jom.html, pp. 36.

Egerton, T.A., 2000. Titanium compounds, Inorganic, Kirk-Othmer Encyclopedia of


Chemical Technology. John Wiley & Son.

Einaga,H., 1979. Hydrolysis of titanium(IV) in aqueous (Na, H) Cl solutions. Journal of


Chemical Society. Dalton Trans, 1917-1919.

El-Hazek, N., Lasheen, T.A., El-Sheikh, R., Zaki, S.A., 2007. Hydrometallurgical
criteria for TiO2 leaching from Rosetta ilmenite by hydrochloric acid.
Hydrometallurgy 87, 45-50.

Gambogi, J., 2009. Titanium mineral concentrates. US Geological survey.

Georgiou, D., Papangelakis, V.G., 1998. Sulphuric acid pressure leaching of a limonitic
laterite: chemistry and kinetics. Hydrometallurgy 49, 23-46.

305
References

Gireesh, V.S., Vinod. V.P., Krishnan Nair, S., Georgee Ninan, 2015. Catalytic leaching
of ilmenite using hydrochloric acid: A kinetic approach. International Journal
of Mineral Processing 134, 36-40.

Grey, I.A., Reid, A.F., 1975. The structure of pseudorutile and its role in the natural
alteration of ilmenite. The American Minerologist 60, 898-906.

Han, K.N., Rubcumintara, T., Fuerstenau, M.C., 1987. Leaching behaviour of ilmenite
with sulfuric acid. Metallurgical Transaction B 18 (2), 325-330.

Haverkamp, R.G., Kruger, D., Rajashekar, R., 2016. The digestion of New Zealand
ilmenite by hydrochloric acid. Hydrometallurgy 163, 198-203.

Helgeson, H.C., 1969. Thermodynamics of hyrdrothermal systems at elevated


temperatures and pressures. American Journal of Science 267, 729-804.

Henry V. K., (1994). CRC handbook of thermophysical and thermochemical data


(Ed.).CRC press Inc. Boca Raton.

Holze, R., 2009. Experimental electrochemistry. Wiley-VCH.

House, C.I., Kelsall, G.H., 1984. Potentail -- pH diagram for Sn/H2O-Cl system.
Electrochimica Acta 29 (10), 1459-1464.

Hope, Matthew., 2016. Mineral sand price forecasts. Report for Equity research,
Diversified metal and mining. 28 July 2016.

Ibrahim I.A., Mahmoud M.H.H., Afifi A.A.I., El-Sayed B.A., 2003. Direct
hydrochloric acid leaching of an Egyptian ilmenite ore for production of
synthetic ilmenite, Fifth International Conference in Honor of Prof Ian Ritchie-
Volume 1: Leaching and Solution Purification. Edited by C.A. Young, A.M.
Alfantazi, C.G. Anderson, D.B. Dreisinger, B. Harris and A. James, TMS (The
Minerals, Metals and Materials society), pp. 555 - 564.

Imahashi, M., Takamatsu, N., 1976. The dissolution of titanium minerals in


hydrochloric and sulfuric acids. Bulletin of the chemical society of Japan 49
(6), 1549-1553.

Ishikawa, Y., 1958a. Electric properties of FeTiO3-Fe2O3. Journal of the Physical


Society of Japan 17, 37-42.

306
References

Ishikawa, Y., 1958b. An order-disorder transformation phenomenon in the FeTiO3-


Fe2O3 solid solution series. Journal of the Physical Society of Japan 13, 828-
837.

Ishikawa, Y., 1962. Magnetic properties of the ilmenite-hematite system at low


temperature. Journal of the Physical Society of Japan 17, 828-837.

Jablonski, M., Przepiera, A., 2001. Kinetic model for the reaction of ilmenite with
sulfuric acid. J. Therm. Anal. Calorim 65, 583-590.

Jackson, J.S., Wadsworth, M.E., 1976. Kinetic study of the dissolution of Allard Lake
ilmenite in hydrochloric acid. Light Metals, 481-540.

Jayasekera, S., Marinovich, Y., Avraamides, J.I., Bailey, S., 1995. Pressure Leaching of
reduced ilmenite:electrochemical aspects. Hydrometallurgy 39, 183-199.

Jha, A., Lahiri, Kumari, E.J., 2008. Beneficiation of titaniferous ores by selective
separation of iron oxide impurities and rare earth oxides for the production of
high grade synthetic rutile. Miner. Process. Extract. Metall. 117 (3), 157-165.

Kelsall, G.H., Robbins, D.J., 1990. Thermodynamics of Ti-H2O-F(-Fe) systems at


298 K. Journal of The Electrochemistry Society 283, 135-157.

Kumari, E.J., Berckman, S., Yegnaraman, V., Mohandas, P.N., 2002. An


electrochemical investigation of the rusting reaction of ilmenite using cyclic
voltammetry. Hydrometallurgy 65, 217-225.

Lakshamanan, V.I., Sridhar, R., Roy, R.R., Ramachandran, V., 2008. Recovery of
metals from ores by mixed chloride extraction, Proceeding of the sixth
International Symposium of Hydrometallurgy 2008, Colorado, USA, pp. 895-
902.

Lanyon, M.R., Lwin, T., Merritt, R.R., 1999. The dissolution of iron in hydrochloric
acid leach of an ilmenite concentrate. Hydrometallurgy 51 (3), 299-323.

Lasheen, T.A.I., 2005. Chemical benefication of Rosetta ilemenite by direct reduction


leaching. Hydrometallurgy 76, 123-129.

Lazaro-Baez, M.I., 2001. Electrochemistry of the leaching of chalcopyrite, Ph.D.


Thesis. Murdoch Universiti, Perth.
307
References

Levenspiel, O., 1972. Chemical reaction engineering. John Wiley & Son: New York,
USA, pp. 668.

Liang, B., Li, C., Zhang, C., Zhang, Y., 2005. Leaching kinetics of Panzhihua ilmenite
in sulfuric acid. Hydrometallurgy 76, 173-179.

Li, C., Ling, B., Lou, C., 2007. Dissolution of mechanically activated Panzhihua
ilmenite in dilute solution of sulfuric acid. Hydrometallurgy 89, 1-10.

Li, C., Liang, B., Wang, H., 2008. Preparation of synthetic rutile by hydrochloric acid
leaching of mechanically activated Panzhihua ilmenite. Hydrometallurgy 91,
121-129.

Lindsley, D.H., 1976. Oxide minerals: Petrologic and magnetic significance.


Minerology Society of America, Washington, DC.

Low, C.T.J, Walsh F.C., 2008. The Stability of an acidic tin methanesulfonate
electrolyte in the presence of hydroquinone antioxidant. Electrochimica Acta
53, 5280-5286.

Lu, Z.Y., Muir, D.M., 1985. A comparative study of the oxidative and reductive of
magnetite in acidified CuSO4-acetronitrile-H2O and CuCl2-NaCl-H2O leach
solutions. Journal of applied electrochemistry 16, 745-756.

Lu, Z.Y., Muir, D.M., 1988. Dissolution of metal ferrite and iron oxides by HCl under
oxidising and reducing conditions. Hydrometallurgy 21, 9-21.

Mahmoud, M.H.H., Afifi, A.A.I., Ibrahim, I.A., 2004. Reductive leaching of ilmenite
ore in hydrochloric acid for preparation of synthetic rutile. Hydrometallurgy
73, 99-109.

Marinovich, Y., Bailey, S., Avraamides, J., Jayasekera, S., 1995. An electrochemical
study of reduced ilmenite carbon paste electrodes. Journal of applied
electrochemistry 25, 823-832.

McConnel, S.R., 1978. Dissolution of ilmenite and related minerals in sulfuric acid,
Ph.D. Thesis. Murdoch University, Perth.

308
References

Nicol, M.J., 1993. The role of electrochemistry in hydrometallurgy, In hydrometallurgy:


Fundamentals, technology and innovation, Hiskey J. B. and Warren G. W. Eds.
Salt Lake City Utah, The Mineral and Materials Society of AIME, 43-62.

Nicol, M. J., 1975. Mechanism of aqueous reduction of chalcopyrite by copper, iron


and lead. Inst. Min. Metall. Trans., Sect C, 84, C206-C209.

Nicol, M.J., Lazaro, I., 2002. The Role of EH measurements in the interpretation of the
kinetics and mechanisms of the oxidation and leaching of sulfide minerals.
Hydrometallurgy 63, 15-22.

Nishikata, E., Ishil, T., Ohta, T., 1981. Viscosities of aqeous hydrochloric acid solutions
and densities and viscosities of hydroiodic acid solutions. Journal of Chemical
Engineering Data 26, 254-256.

Olanipekun, E., 1999. A kinetic study of the leaching of a Nigerian ilmenite ore by
hydrochloric acid. Hydrometallurgy 53, 1-10.

Ogasawara, T., Veloso de Araứjo, R.V., 2000. Hydrochloric acid leaching of a pre-
reduced Brazilian ilmenite concentrate in an autoclave. Hydromatllurgy 56,
203-216.

Orth, R.J., Liddell, K.C., 1986. Hydrometallurgical reactor design and kinetics. Kinetics
and solution in the leaching of ilmenite by hydrochloric acid- A partial
equilibrium model. The Metallurgical Society Inc, USA.

Parida, K.M., Das, N.N. 1996. Reductive dissolution of hematite in hydrochloric acid
medium by some inorganic and organic reductants: A comparative study.
Indian Journal of Engineering and Materials Sciences 3, 243-247.

Pownceby, M.I., Sparrow, G.J., Fisher-White, M.J., 2008. Minerogical characterisation


of Eucla Basin ilmenite concentrates-First results from a new global resource.
Minerals engineering 21, 587-597.

Premaratne, W. A. P. J., Rawson, N. A., 2003. The processing of beach sand from Sri
Lanka for the recovery of titanium using magnetic separation. Physical
Separation in Science and Engineering 12, 13-22.

309
References

Pugaev, D., Senanayake, G., Nicol, M., 2011. The mechanism of the passivation of
sulfide minerals in oxidative leaching. 6th Southern African Base Metals
Conference 2011. The Southern African Institute of Mining and Metallurgy,
South Africa, pp. 39-48.

Rand, D.A.J., Woods, R., 1984. Eh measurements in sulphide mineral slurries.


International Journal of Mineral Processing 13 (1), 29-42.

Ray, H.S., 1993. Kinetics of metallurgical reactions, International Science, pp.77-83.

Raymond, K.N., Wenk, H.R., 1971. Lunar ilmenite (refinement of the crystal structure).
Contributions to mineralogy and petrology 30, 135-140.

Reyneke, L., Wallmach, T., 2007. Characterization of FeTi-oxide species occuring in


the Ranobe heavy mineral deposit, Madagascar, The 6th International Heavy
Minerals conference 'Back to Basics'. The Southern African Institute of Mining
and Metallurgy, South Africa, pp. 151-158.

Rumble, D., 1976. Oxide minerals. Review in minerology. Mineralogical Society of


America, Washington, D.C.

Roine, A., 2011. HSC Chemistry Software (Version 7.1) Outotec Research Oy, Finland.
(www.outotec.com).

Sarker, M.K., Rashid, A.K.M.B., Kurny, A.S.W., 2006. Kinetics of leaching of oxidized
and reduced ilmenite in dilute hydrochloric acid solutions. International Journal
Mineral Processing 80, 223-228.

Sasikumar, C., Rao, D.S., Srikanth, S., Mukhopadhyay, N.K., Mehrotra, S.P., 2007.
Dissolution studies of mechanically activated Manavalakurichi ilmenite with
HCl and H2SO4. Hydrometallurgy 88, 154-169.

Sasikumar, C., Rao, D.S., Srikanth, S., Ravikumar, B., Mukhopadhyay, N.K., Mehrotra,
S.P., 2004. Effect of mechanical activation on the kinetics of sulfuric acid
leaching of beach sand ilmenite from Orissa, India. Hydrometallurgy 75, 189-
204.

310
References

Senanayake, G., Muir, D., 1988. Studies on liquid junction potentials in concentrated
chloride solutions and determination of ionic activities and hydration numbers
by the emf method. Electrochimica Acta 33, 3-9

Senanayake, G., Muir, D.M., 2003. Chloride processing of metal sulphide: Review of
fundamental and applications. Hydrometallurgy 2003: Proceedings of the 5th
International Symposium; Vancouver, BC; Canada, Volume 1 pp. 517-531.

Senanayake, G., Das, G.K., 2004. A comparative study of leaching kinetics of limonitic
laterite and synthetic iron oxides in sulfuric acid containing sulfur dioxide.
Hydrometallurgy 72, 59-72.

Senanayake, G., Das, G.K., de Lange, A., Li, J., Robinson, D. J., 2015. Reductive
atmospheric acid leaching of lateritic smectite/nontronite ores in
H2SO4/Cu(II)/SO2 solutions. Hydrometallurgy 152, 44-54.

Shirane, G., Pickart, S.J., Nathan, R., Ishikawa, Y., 1959. Neutron diffraction study of
antiferromagnetic FeTiO3 and its solid solutions with Fe2O3. Journal of Physics
and Chemistry of Solids 10, 34-43.

Sillen, L.G., Martell, A.E. 1964. Stability constant of metal complexes. Chemical
Society of Special Publication. Vol 17. Chemical Society, London.

Sinha, H.N., 1973. MURSO process for producing rutile subtitute. In: R.I. Jaffe and
H.M. Burte (Editors), Titanium science and technology. A publication of the
Metallurgical Society of AIME, Plenum Press, New York-London, Cambrige,
Massachusetts, pp. 233-244.

Stuart, A.D., Lawson, J.A., Ward, C.B., Peng, H., 2010. A sulphate process. WO
Patent 2010034083-A1.

Tsuchida, H., Narita, E., Takeuichi, H., Adachi, M., Okabe, T., 1982. Manufacture of
High Pure Titanium(IV) Oxide by the Chloride Process I. Kinetic study on
leaching of ilmenite ore in concentrated hydrochloric acid solution. The
chemical society of Japan, p. 55.

Valsquez-Yevenas, L., Nicol, M., Miki, H., 2010. The dissolution of chalcopyrite in
chloride solutions. Part 1. The effect of solution potential. Hydrometallurgy
103, 108-113.

311
References

van Dyk, J.P., Vegter, N.M., Pistorius, P.C., 2002. Kinetics of ilmenite dissolution in
hydrochloric acid. Hydrometallurgy 65, 31-36.

Vaughan, J., Alfantazi, A., 2004. The thermodynamics of titanium corrosion in acidic
systems. In: M.J. Collins and V.G. Papangelakis (Editors), Pressure
Hydrometallurgy 2004, 34th Annual Hydrometallurgy Meeting, Canada.

Vaughan, J., Alfantazi, A., 2006. Corrosion of Titanium and Its Alloy in Sulfuric Acid
in the presence of Chlorides. Journal of The Electrochemistry Society, 152:
B6-B12.

Verhulst, D., Sabacky, B., Spitler, T., Prochazka, J. 2003. New development in the
Altair hydrochloride TiO2 pigment process. Hydrometallurgy 2003: 5th
International Conference in Honour of Professor Ritchie, Vancouver, August
2003, vol. 1 (2003), pp. 565–575.

Vogel, A.I., 1978. A textbook of quantitative inorganic analysis. 4th Ed. Longman
group, London.

von Horn, D., 1993. Dry Mineral Separation Plant of Tiwest Joint Venture, Chandala,
WA. Australasian Mining and Metallurgy: the Sir Maurice Mawby Memorial 2
pp. 1294-1297.

Wagner, C., Traud, W., 1938. The interpretation of corrosion phenomena by


superimpostion of electrochemical partial reactions and the formation of
potentials of mixed electrodes. Z. Electrochem 44, 391-402.

Walpole, E.A., 1999. Synthetic rutile production from Murray Basin Ilmenite by the
Auspact ERMS process, Murray Basin Mineral Sands Conference. Australian
Institute of Geoscientist, Victoria, Australia.

Wang, Y., Yuan, Z., 2006. Reductive kinetics of reaction between a natural ilmenite and
carbon. International Journal of Mineral Processing 3 (81), 133-140.

Wang, Z., Xue, J., Wang, H., Jiang, X., 2009. Kinetic study of hydrochloric acid
leaching process of ilmenite for rutile synthesis, EPD Congress 2009 (The
Minerals, Metals & Materials Society), pp. 837-843.

312
References

Ward, C.B., 1990. The Production of Synthetic Rutile and By-Product Iron Oxide
Pigments from Ilmenite Processing, Ph.D. Thesis, Murdoch University, Perth.

Welham, N.J., 1996. A parametric study of the mechanically activated carbothermic


reduction of ilmenite. Minerals Engineering 9 (12), 1189-1200.

Welham, N.J., Llewellyn, D.J., 1998. Mechanical enhancement of the dissolution of


ilmenite. Minerals Engineering, 11 (9), 827-841.

White, A.F., Peterson, M.L., Hochella, M.F., 1994. Electrochemistry and dissolution
kinetics of magnetite and ilmenite. Geochimica et Cosmochimica Acta 58 (8),
1859-1875.

Wu, F., Li, X., Wang, Z., Wu, L., Guo, H., Xiong, X., Zhang, X., Wang, X., 2011.
Hydrogen peroxide leaching of hydrolyzed titania residue prepared from
mechanically activated Panzhihua ilmenite leached by hydrochloric acid.
International Journal of Mineral Processing 98 (1-2), 106-112.

Xu, S., Huang, Z.S., 1993. The kinetics of Panzhihua ilmenite with sulfuric acid. J. Min.
Metal. Eng. 13, 44-48.

Xue, T., Wang, L., Qi, T., Chu, J., Qu, J., Liu, C., 2009. Decomposition kinetics of
titanium slag in sodium hydroxide system. Hydrometallurgy 95, 22-27.

Yang, K.K., Mahmoudian, M.R., Ebadi,M., Koay, H.L., Basirun, W.J., 2011. Diffusion
coefficient of tin(II) methanesulafonate in ionic liquid and methane sulfonic
acid (MSA) solvent. Metallurgical and Material Transactions B, 42, 1274-
1279.

Zhang, X.M., Senanayake, G., Nicol, M.J., 2008. Beneficial effect of silver in
thiosulfate leaching of gold. Hydrometallurgy 2008: 6th International
Symposium; Phoenix, AZ; United States; 17 August 2008, pp. 801-810.

Zhang, S., Nicol, M.J., 2009. An electrochemical study of the reduction and dissolution
of ilmenite in sulfuric acid. Hydrometallurgy 97, 146-152.

Zhang, S., Nicol, M.J., 2010. Kinetics of the dissolution of ilmenite in sulfuric acid
solutions under reducing conditions. Hydrometallurgy 103 (1-4), 196-204.

313
References

Zhang, W., Zhu, Z., Chang C.Y., 2011a. A literature review of titanium metallurgical
processes. Hydrometallurgy 108 (3-4), 177-188.

Zhang, L., Hu, H., Liao, Z., Chen, Q., Tan, J., 2011b. Hydrochloric acid leaching
behavior of different treated Panxi ilmenite concentrations. Hydrometallurgy
107 (1-2), 40-47.

Resources from webpages:

http://www.geocities.jp/ohba_lab_ob_page/structure6.html
http://www.calctool.org/CALC/chem/electrochem/cv1

314
Appendix

APPENDIX

Appendix A1

Table A1.1
Thermodynamic data for possible reactions in the dissolution of ilmenite at 25 or
60 °C.

Reaction T (oC) E° No.


(V vs SHE)

Fe3+ + e- → Fe2+ 25 0.69 A1.1

2FeTiO3 + 6H+ + 2e- → Ti2O3 + 3H2O + 2Fe2+ 60 -0.10 A1.2

Fe2O3 + 6H+ + 2e- → 2Fe2+ + 3H2O 60 0.75 A1.3

FeTiO3 + 6H+ + e- → Fe2+ + Ti3+ + 3H2O 60 -0.29 A1.4

TiO2+ + 2H+ + e- → Ti3+ + H2O 60 0.046 A1.5

2TiO2+ + H2O + 2e- → Ti2O3 + 2H+ 25 0.24 A1.6


(Vaughan and Alfantazi, 2004, 2006; Zhang and Nicol, 2009)

315
Appendix

Table A1.2
Equilibrium constants.

Ionic
T
Reaction K strength No.
(oC)
(M)
Complexation reaction

Ti4+ + Fe2+ → Fe3+ + Ti3+ 3.19 x 10-10 3.0 25 A1.7

Fe2+ + Cl- → FeCl+ 2.29 x 100 2.0 20 A1.8

Fe2+ + 2Cl- → FeCl2 1.10 x 100 2.0 20 A1.9

Fe3+ + Cl- → FeCl2+ 6.31 x 100 3.0 25 A1.10

Fe3+ + 2Cl- → FeCl2+ 1.99 x 100 2.0 20 A1.11

Fe3+ + 3Cl- → FeCl30 8.71 x 10-1 2.0 20 A1.12

Ti4+ + H2O + Cl- → TiOCl+ + 2H+ 7.42 x 10-2 3.0 25 A1.13

Ti4+ + H2O + 2Cl- → TiOCl2 + 2H+ 8.32 x 10-2 3.0 25 A1.14

Ti3+ + Cl- → TiCl2+ 2.19 x 100 0.5 40 A1.15

Ti4+ + H2O + Cl- → TiOCl+ + 2H+ 7.42 x 10-2 3.0 25 A1.16

Hydrolysis
Fe2+ + H2O → FeOH+ + H+ 1.32 x 10-10 1.0 25 A1.17
Fe3+ + H2O → FeOH2+ + H+ 1.07 x 10-3 3.0 25 A1.18

Fe3+ + 2H2O → Fe(OH)2+ + 2H+ 5.53 x 10-7 3.0 25 A1.19

2Fe3+ + 2H2O → Fe2(OH)24+ + 2H+ 1.75 x 10-3 1.0 25 A1.20

Ti4+ + H2O → TiOH3+ + H+ 1.99 x 100 2.0 25 A1.21

Ti4+ + 2H2O → Ti(OH)22+ + 2H+ 2.40 x 101 2.0 25 A1.22

Ti4+ + H2O → TiO2+ + 2H+ 2.09 x 10-2 0.0 25 A1.23

Ti3+ + H2O → TiOH2+ + H+ 2.09 x 10-3 3.0 25 A1.24

2Ti3+ + 2H2O → Ti2(OH)24+ + 2H+ 2.77 x 10-4 3.0 25 A1.25


(Orth and Liddell, 1986)

316
Appendix

Table A1.3
Standard Gibbs free energy for titanium and iron species.

Species G° (kJ mol-1) S⁰ (kJ mol-1 K-1)


Ti 0 0.03029
Ti2+ 314.22
Ti3+ 349.78
TiO2+ 577.40
TiO2 467.20
489.2 0.03477
TiO (c)
495.0 0.050
Ti2O3 (c) 1432.2 0.07878
Ti2O3 1388
Ti(OH)3 1049.8
2314 0.1293
Ti3O5
2317.4 0.1294
Ti4O7 3213.266
TiO2 (c, hydrated) 821.3
TiO2.H2O (c) 1058.5
TiO2 (c) 888.4 0.05025
TiO2 (c, anatase) 884.5 0.04992
TiO2 (c, rutile) 889.5 0.04992
TiH (c) 3.59
4.92
TiH2 0.0297
80.3
1125.1
1168.7 0.1059
FeTiO3 (c) 1158.1 0.1059
1277.9 0.1059
1267.0
Fe2+ 91.5 -0.106
Fe3+ 16.7 -0.2803
Fe(OH)3 (ppt) 705.535 0.105
Fe2O3 743.608 0.0874
Fe3O4 1015.359 0.1464
Fe(OH)2 (ppt) 492.017 0.0879
H2O 237.178 0.06991
OH- 157.293 0.010175
(Kelsall and Robbins, 1990)

317
Appendix

Table A1.4
Thermodynamic data for possible reactions in the dissolution of ilmenite.


K
Reaction (V vs
at different temperatures
SHE)
25 oC 75 oC 90 oC 25 oC
Fe3+ + Cl- → FeCl2- 2.44 2.74
Fe3+ + 2Cl- → FeCl2+ 3.12 3.42
Fe3+ + 3Cl- → FeCl3 2.37 2,75
Fe3+ + 4Cl- → FeCl4- 0.84 1.32
Ti3+ + H2O → TiOH2+ + H+ 0.0028
Ti3+ + Cl- → TiCl2+ 0.2
TiOCl3- + Cl- → TiOCl42- 12
3.55 ±
TiO2+ + Cl- → TiOCl+
0.35
0.40 ±
TiOCl+ + Cl- → TiOCl2
0.06
10.6 ±
TiOCl2 + Cl- → TiOCl3-
0.20
TiO + 2H+ + 2e- → Ti + H2O -1.306
Ti3+ + 3e- → Ti -1.21
TiO2+ + 2H+ + 4e- → Ti + H2O -0.882
TiO2 + 4H+ + e- → Ti3+ + 2H2O -0.666
TiO2 + 4H+ + 2e- → Ti2+ + 2H2O -0.502
Ti3+ + e- → Ti2+ -0.368
TiO2+ + 2H+ + e- → Ti3+ + H2O +0.1
(Helgeson, 1969; Kelsall and Robbins, 1990; Vaughan and Alfantazi, 2004)

318
Appendix

Appendix A2
Appendix A2.1. Results of synthetic ilmenite rotating discs experiments :
Dissolution efficiency of Fe and Ti (Chapter 5)

Experimental conditions:

Disc surface area 0.00038 m2


[HCl] 7 mol/L
Temperature 60 oC
HCl volume 250 mL

Table A2.1
Effect of stirring rate
Other conditions:

319
Appendix

320
Appendix

Table A2.2
Effect of acid concentration
Other conditions:

321
Appendix

322
Appendix

Table A2.3
Effect of temperature.

Other conditions:

323
Appendix

324
Appendix

Table A2.4
Effect of reducing agent, SnCl2

Other conditions:

325
Appendix

326
Appendix

Appendix A2.2. Polynomial method


The rates of dissolution of synthetic ilmenite in 7 M HCl at 50 oC, 60 oC, 70 oC and

80 oC were obtained from the slopes of Figs. A2.1 and A2.2 using the polynomial

method. This can conveniently be obtained by fitting a polynomial equation to the

data of dissolved ions (Fe or Ti) versus time (t), as shown in Figs. A2.1-A2.2.

Second or third order (degree) polynomial regression using Excel was performed to

obtain the parameters of the best fit to the experimental data. The rate (dn/dt) was

derived from the Eq. A2.1 –A2.3 as shown below.

For T = 60 oC, from Fig. A2.1

n = -0.00007t2 + 0.00523t + 0.00499 R² = 0.99 [A2.1]

Differentiation with respect to time, t gives

dn/dt = 2( -0.00007)t + 0.00523 [A2.2]

when t = 0

dn/dt = 0.00523 s-1 [A2.3]

327
Appendix

Fig. A2.1. Effect of temperature on iron dissolution from synthetic ilmenite disc in
7 M HCl solution at 800 rpm.

Fig. A2.2. Effect of temperature on iron dissolution from synthetic ilmenite disc in
7 M HCl solution at 800 rpm.

328
Appendix

Appendix A2.3 EDX spectrum

Figure A2.3 EDX spectrum for bulk synthetic ilmenite disc for SEM image in
Figure 5.24 (B).

Figure A2.4 EDX spectrum on white spot mark for SEM image in Figure 5.24
(D).

329
Appendix

Appendix A3

Appendix A3.1. (Chapter 6)


Levich Equation

IL = A*iL = 0.62 FA D2/3 v-1/6 ½ C [A3.1]

The Levich equation plot iL as a function of ω1/2 and shows a linear relationship (Fig. 6.25

and Fig. 6.27). The slope of the linear relationship was used to determine the diffusion

coefficient.

Slope = 0.62 n F A D2/3 C v-1/6 [A3.2]

where v, kinematic viscosity for HCl at 25 oC is 0.9541 cSt = 9.54 x 10-7 m2 s-1, 1 cSt = 1

centiStokes = 1 x 10-6 m2 s-1(Nishikata et al. 1981); F, Faraday constant = 96485 A s/mol;

C[Sn], = 4.2 x 10-6 mol/cm3; C[Fe], = 4.5 x 10-6 mol/cm3; slope[Sn] = 0.00006; slope[Fe] =

0.00005; number of electrons, n for Sn(II) = 2 based on Eq. A3.3 and n for Fe(II) = 1 based

on Eq. A3.3 - A3.4:

Sn2+ → Sn4+ + 2e- [A3.3]

Fe2+ →Fe3+ + e- [A3.4]

The calculated value of the diffusion coefficient of Sn(II), DSn(II) = 1.7 x 10-6 cm2 s-1

The calculated value of the diffusion coefficient of Fe(II), D[Fe(II)] = 3.3 x 10-6 cm2 s-1

330
Appendix

Appendix A3.2.
(I) Inductive coupled plasma optical emission spectrometry (ICP-OES) data
for coulometric measurements with massive Ilmenite (Chapter 6)
Experimental conditions:

Electrode Massive ilmenite


Electrode surface area 2.5E-05 m2
[HCl] 4 mol/L
Temperature 60 oC
Volume 0.015 L
Time 3600 seconds

Table A3.1
Effect of potential on dissolution rate of massive ilmenite electrode.

Potential Time Fe Fe Fe Fe Rate


V s mg/L g/L mol/L mol mol m-2 s-1
-0.15 3600 280 0.28 0.00501 7.520E-05 8.36E-04
0 3600 220 0.22 0.00394 5.909E-05 6.57E-04
0.1 5400 93 0.093 0.00167 2.498E-05 1.85E-04
0.3 7200 27 0.027 0.00048 7.252E-06 4.03E-05
0.5 36000 24 0.024 0.00043 6.446E-06 7.16E-06

Potential Time Ti Ti Ti Ti Rate


V s mg/L g/L mol/L mol mol cm-2 s-1
-0.15 3600 20 0.02 0.000418 6.266E-06 6.96E-05
0 3600 16 0.016 0.000334 5.013E-06 5.57E-05
0.1 5400 5.4 0.0054 0.000113 1.692E-06 1.25E-05
0.3 7200 1.1 0.0011 0.000023 3.446E-07 1.91E-06
0.5 36000 1.7 0.0017 0.000036 5.326E-07 5.92E-07

331
Appendix

(II) Inductive coupled plasma optical emission spectrometry (ICP-OES) data


for coulometric measurement with carbon paste electrode of Iluka ilmenite
(Chapter 6)

Experimental conditions:

Electrode CPIL
Electrode surface area 0.14 cm2
[HCl] 4 mol/L
Temp 60 oC

Volume 0.015 L
Time 3600 s

Table A3.2
Effect of potential on dissolution rate of Iluka carbon paste electrode (CP IL).

Potential Time Fe Fe Fe Fe Rate


V s mg/L g/L mol/L mol mol m-2 s-1
-0.15 3600 24 0.0240 0.000430 6.45E-06 1.28E-04
0 3600 1.3 0.0013 0.000023 3.49E-07 6.93E-06
0.1 5400 3.4 0.0034 0.000061 9.13E-07 1.21E-05
0.3 7200 2.7 0.0027 0.000048 7.25E-07 7.19E-06

Potential Time Ti Ti Ti Ti Rate


V s mg/L g/L mol/L mol mol m-2s-1
-0.15 3600 1.7 0.00170 3.55E-05 5.326E-07 1.06E-05
0 3600 0.3 0.00030 6.27E-06 9.398E-08 1.86E-06
0.1 5400 0.81 0.00081 1.69E-05 2.538E-07 3.36E-06
0.3 7200 0.54 0.00054 1.13E-05 1.692E-07 1.68E-06

332
Appendix

Appendix A4

Appendix A4.1.

Fig. A4.1. Effect of the presence of tin powder on Fe and Ti leaching efficiency for
North Capel sample. ([HCl] = 7 M, T = 80 oC, [Sn] = 1.0 g/L, pulp density = 4.0 g/L,
size fraction = +53-63 µm, t = 5 h).

Table A4.1
Results of measured potentials during leaching.

Electrode potential (V vs SHE)


Time
Pt electrode Iluka ilmenite electrode
(hours)
Without Sn With Sn Without Sn With Sn
0 0.1286 -0.6917 0.9978 0.6853
1 -0.2864 -0.5403 0.9099 0.6853
2 -0.1985 -0.5012 0.8464 0.8366
3 -0.1692 -0.4719 0.8171 0.8513
4 -0.1546 -0.4671 0.7976 0.8611
5 -0.4622 0.8659

333
Appendix

Appendix A4.2. The fitting of shrinking core model for ilmenite leaching

(a)

(b)

Fig. A4.2. Fitting of shrinking core model for North Capel ilmenite in HCl solution
from 7 M – 11 M at 80 oC, (a) Fe and (b) Ti.

334
Appendix

(a)

(b)

Fig. A4.3. Fitting of shrinking core model for Iluka ilmenite in HCl solution from 7
M – 11 M at 80 oC, (a) Fe and (b) Ti.
.

335
Appendix

(a)

(b)

Fig. A4.4. Fitting of shrinking core model for Tiwest ilmenite in HCl solution from 7
M – 11 M at 80 oC, (a) Fe and (b) Ti.

336
Appendix

(a)

(b)

Fig. A4.5. Fitting of shrinking core model for Iluka ilmenite in the presence of tin
metal powder (Sn) in 7 M HCl at 80 oC, (a) Fe and (b) Ti.

337
Appendix

(a)

(b)

Fig. A4.6. Fitting of shrinking core model for Tiwest ilmenite in the presence of tin
metal powder (Sn) in 7 M HCl at 80 oC, (a) Fe and (b) Ti.

338
Appendix

(a)

(b)

Fig. A4.7. Fitting of shrinking core model for North Capel ilmenite in 7 M HCl
solution at temperature from 80 oC to 110 oC, (a) Fe and (b) Ti.

339
Appendix

(a)

(b)

Fig. A4.8. Fitting of shrinking core model for Iluka ilmenite in 7 M HCl solution at
temperature from 80 oC to 110 oC, (a) Fe and (b) Ti.

340
Appendix

(a)

(b)

Fig. A4.9. Fitting of shrinking core model for Tiwest ilmenite in 7 M HCl solution at
temperature from 80 oC to 110 oC, (a) Fe and (b) Ti.

341
Appendix

Appendix A4.3. Effect of particle size

Fig. A4.10. Variation of kpl as a function of initial particle radius, 1/r2 (North Capel,
size fraction of +45-53 µm, +53-63 µm, +63-75 µm, +75-90 µm, 80 oC, solid liquid
ratio = 4 g/L in 7 M HCl) Data from Fig. 7.2.

Fig. A4.11. Variation of kpl as a function of initial particle radius, 1/r2 (Iluka, size
fraction of +45-53 µm, +53-63 µm, +63-75 µm, +75-90 µm, 80 oC, solid liquid ratio
= 4 g/L in 7 M HCl) Data from Fig. 7.3.

342
Appendix

Fig. A4.12. Variation of kpl as a function of initial particle radius, 1/r2 (Tiwest, size
fraction of +45-53 µm, +53-63 µm, +63-75 µm, +75-90 µm, 80 oC, solid liquid ratio
= 4 g/L in 7 M HCl) Data from Fig. 7.4.

Table A4.2
Initial radius and kpl data of North Capel, Iluka and Tiwest for effect of particle sizes.

kpl (min-1)
Size Radius, North Capel Iluka Tiwest
Fraction
(µm) r (cm) r2 (cm2)

Fe Ti Fe Ti Fe Ti
+45-53 0.00245 6.0 x 10-6 0.00093 0.00421 0.000207 0.00018 0.000120 0.00015

+53-63 0.00290 8.4 x 10-6 0.00103 0.00144 0.000128 0.00017 0.000114 0.00010

+63-75 0.00345 1.2 x 10-5 0.00109 0.00134 0.000152 0.00016 0.000107 0.00014

+75-90 0.00412 1.7 x 10-5 0.00095 0.00075 0.00008 0.00004 0.000108 0.00004

343
Appendix

Appendix A4.4 Diffusion Coefficient of H+ calculation

The ilmenite dissolution reaction:


FeO.TiO2 + 4HCl = FeCl2 + TiOCl2 + 2H2O (A4.1)

ρ = ρm x % Fe or Ti in material / (100 x molar mass of Fe or Ti) (A4.2)

The diffusion coefficient of H+ (DH+) was calculated using Eq. A4.2 and A4.3

with the substitution of the rate constant (kpl) obtained from the slopes of the

shrinking core plots (Appendix A4.2) listed in Table 7.20 (Chapter 7):

DH+ = (kpl ρ ro2) / (6 b[H+]) (A4.3)

From Eq. A4.1, stoichiometry factor is b = 4, where r is the initial particle

radius = 0.0029 cm; [H+] = 0.007 mol cm-3 and ρ (molar density in mol cm-3) is

related to weight percent of Fe or Ti in the ilmenite sample and its density, ρm = 4.5 g

cm-3 (Zhang and Nicol, 2009), the value of ρ was calculated using Eq. A4.2:

Using data from North Capel sample, Fe in sample = 31.5 %, Fe molecular weight =

56.0 g mol-1

Therefore, ρ = 4.5 g cm-3 x 31.5 / (100 x 56.0 g mol-1) = 0.0253 mol cm-3.

The substitution of the parameter given above into Eq. A4.3 and calculated value of

ρ is 0.0253 mol cm-3, kpl = 1.67 x 10-5 s-1 gives the following DH+ values :

DH+ = 1.67 x 10-5 s-1 x 0.0253 mol cm-3 x (0.0029)2 / (6 x 4 x 0.009)

= 2.12 x 10-11 cm2 s-1

344

You might also like