+ All Categories
Home > Documents > REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

Date post: 11-Jan-2022
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
106
REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES AND THEORETICAL AND EXPERIMENTAL KINETIC STUDIES by Jaehun Cho A thesis submitted to the faculty of The University of Utah in partial fulfillment of the requirements for the degree of Master of Science Department of Metallurgical Engineering The University of Utah August 2016
Transcript
Page 1: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES

AND THEORETICAL AND EXPERIMENTAL

KINETIC STUDIES

by

Jaehun Cho

A thesis submitted to the faculty of

The University of Utah

in partial fulfillment of the requirements for the degree of

Master of Science

Department of Metallurgical Engineering

The University of Utah

August 2016

Page 2: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

Copyright © Jaehun Cho 2016

All Rights Reserved

Page 3: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

T h e U n i v e r s i t y o f U t a h G r a d u a t e S c h o o l

STATEMENT OF THESIS APPROVAL

The thesis of Jaehun Cho

has been approved by the following supervisory committee members:

Michael L. Free , Chair 05/06/2016

Date Approved

Zhigang Zak Fang , Member 05/06/2016

Date Approved

Amarchand Sathyapalan , Member 05/06/2016

Date Approved

and by Manoranjan Misra , Chair/Dean of

the Department/College/School of Metallurgical Engineering

and by David B. Kieda, Dean of The Graduate School.

Page 4: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

ABSTRACT

This thesis work mainly evaluates methods of obtaining pure titanium hydride

from different sources and methods and its theoretical and experimental kinetics using

model formulation.

The first route was to use Direct Reduction of Titanium Slag (DRTS) in which

metallic titanium can be obtained by dehydrogenation of titanium hydride after impurities

are removed. The leaching characteristics of iron removal from the reduced upgraded

titanium slag was studied with mild hydrochloric and boric acids under ambient pressure

and elevated pressure. Under the constraint that 1% (w/w) of titanium hydride loss is the

maximum amount tolerable, 0.1 M hydrochloric acid at 140 °C was found to be the most

effective condition for iron removal (87.63%). A factorial design of experiment for

equation modeling with three main factors (temperature, concentration of hydrochloric,

and boric acids) was performed and associated modeling results were in good agreement

with experimental data.

Additional study was carried out to justify the assumption, which utilized the

evaluation of the effects of three empirical size distributions, the Gate-Gaudin-Schuhmann

distribution, the Rosin-Rammler-Bennett distribution, and the Gamma distribution, on the

fluid-solid reaction kinetics. The expressions for overall conversion rate of entire particle

assemblages were derived mathematically, and calculated by a technical computing

language, “MATLAB”. According to the calculation, the assumption that particle size is

Page 5: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

iv

uniform can be valid in the determination of fluid-solid kinetics in the case where the

coefficient of variation (CV) is less than 0.5. Based on the theoretical kinetic study for the

effect of particle size distribution, it was assumed that the effect of particle size distribution

of reduced upgraded titanium slag (UGS) does not have to be considered in the calculation

of kinetics. Based on this calculation, a rate-controlling process can be found and it seems

to follow interfacial reaction controlled kinetics. The activation energy of the reaction was

determined to be 73.9 kJ/mole. Also, the other mechanism of the reaction-controlling

process (solid-state diffusion) is suggested.

An additional way of obtaining pure titanium involves the extraction of titanium

from ilmenite using tannic acid. The experimental and modeling results showed the

feasibility of the new process to obtain titanium.

Page 6: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

TABLE OF CONTENTS

ABSTRACT………………………………………………...........………………….…..iii

LIST OF TABLES………………………………………………………………………..vii

LIST OF FIGURES…………………………………………………………......……...viii

ACKNOWLEDGEMENTS……………………………………………………………….x

Chapters

1 INTRODUCTION………………..…………………………………………………...…1

1.1 Direct reduction of titanium slag……………………...………...…......…….1

1.2 Theoretical kinetic study………………………………………….……...…...4

1.3 Selective extraction of titanium from ilmenite using tannic acid………...……6

2 REVIEW OF THE LITERATURE…………………………………………………...….7

2.1 Processes for titanium production……………………………………...…...7

2.1.1 Kroll process…………………………………....…………...….….7

2.1.2 Alternative processes………………………………………...…….8

2.1.3 Extraction of titanium from ores………………………………......11

2.2 Effect of particle size distribution on kinetics…………………..……………12

2.3 Tannic acid……………………………………...……………...……………16

3 LEACHING CHARACTERISTICS OF IRON REMOVAL FROM REDUCED

UGS……………………………………………...………………………………….….18

3.1 Experimental methods………………………………………………….….18

3.1.1 Preparation of reduced UGS by reduction and acetic acid leaching.18

3.1.2 Experimental procedures........................................................……18

3.2 Experimental results and discussion……………………………….………19

3.2.1 Feed analysis……………………………………………………19

3.2.2 Experimental results…………………………………………….21

3.2.2.1 Effect of hydrochloric acid…………………….……….21

3.2.2.2 Effect of boric acid…………………………….……….25

3.2.3 Effect of temperature…………………..……….…………………30

Page 7: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

vi

3.2.4 Factorial design of experiments for equation modeling…………32

3.2.5 Removal of iron without acetic acid leaching……..…….......…….37

4 EFFECT OF PARTICLE SIZE DISTRIBUTION ON KINETICS…………………..…40

4.1 Model formulation…………..………………………………………………40

4.1.1 Shrinking unreacted core model………………………………...40

4.1.2 Mathematical formulation for size distribution…...………………44

4.2 Results and discussion……………………………….…………………...….50

4.2.1 Gate-Gaudin-Schuhmann distribution………………………….52

4.2.2 Rosin-Rammler-Bennett distribution…………………………...55

4.2.3 Gamma distribution………...…………………………………….57

4.2.4 Comparison of the effect of three size distributions on the kinetics.59

4.2.5 R2 values for the effect of particle size distributions………………61

4.2.6 Comments on particles with nonbasic shapes…………….…...…63

5 KINETIC STUDY OF REDUCED UGS WITH EFFECT OF PARTICLE SIZE

DISTRIBUTION………………………………………………………………….……64

5.1 Particle size distribution of reduced UGS…………...…………………….…64

5.2 Determination of rate-controlling process of the reduced UGS…...…………66

5.2.1 Interfacial reaction……………………………………………...66

5.2.2 Solid state diffusion……………………………………………….69

5.2.2.1 Mathematical formulation……………………….…….69

5.2.3 Effect of external mass transport………………………………….73

6 SELECTIVE RECOVERY OF TITANIUM FROM LEACHED ILMENITE USING

TANNIC ACID………………………..…………………………………………………78

6.1 Materials and methods…………………......………………………………...78

6.1.1 Preparation of titanium solution from ilmenite……………………78

6.1.2 Synthesis of titanium-tannic acid complex………………….…….78

6.1.3 Methodology of DFT simulation………………………………….79

6.2 Results and discussion……………………………………..………………...79

6.2.1 Experiment……………...………………………………………...79

6.2.2 Simulation…………………………...……………………………83

7 CONCLUSIONS……………………………………………………………………...87

REFERENCES…………………….……………………………………………………92

Page 8: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

LIST OF TABLES

Table Page

2.1 Mean, variance, and CV of three distributions……………………………………….14

3.1 Chemical composition of reduced UGS before and after acetic acid washing……….20

3.2 Fe removal % and Ti loss % at different leaching parameters with a solid to liquid ratio

of 1 to 100 and 4 h of leaching time……………...………………………....……......…....24

3.3 Leaching parameters………………………………………………………………...27

3.4 Chemical composition of reduced UGS leached by 0.1 M HCl at 140oC……….….33

3.5 Composition of reduced UGS leached by 0.4 M hydrochloric acid…………………...39

4.1 Mean, variance and CV of three distributions……………………………………....47

4.2 X(t) of three distributions for interface-reaction-controlled system…………………49

4.3 X(t) of three distributions for pore-diffusion-controlled system…………………….51

4.4 Maximum differences of conversion rate between uniform and nonuniform particles

with different distributions, particle shapes, CV, and rate controlling steps…………….54

4.5 R2 values of each case……………………………………….………………………62

5.1 Titanium loss with leaching time. Leaching parameters: .1 M HCl, 1000 rpm, 90oC, no

boric acid………………………………………….………………………....…………...74

Page 9: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

LIST OF FIGURES

Figure Page

1.1 Flow sheet of new route for production of metallic titanium …….………...……….3

1.2 The shrinking unreacted-core system …………………………………….…….….5

2.1 The scheme of TiO2 reduction in the FCC Cambridge process ……………...........10

2.2 The normal and log-normal distributions (CV=0.3)…………………………...15

2.3 The structure of (a) tannic acid and (b) dissociation of tannic acid in aqueous solution

as digallic acid……………………………………………...……………………...….….17

3.1 XRD pattern of the reduced UGS after acetic acid washing ……………….……...22

3.2 SEM image of reduced UGS after acetic acid washing …………………...……....23

3.3 Titanium loss versus HCl concentration …………………………………….……26

3.4 Comparison of iron removal and time for 0.05 M and 0.1 M HCl at 90oC……...….28

3.5 Comparison of iron removal versus time with/without boric acid at 90oC………...29

3.6 Fe removal and Ti loss versus temperature with 0.1 M HCl and with/without boric

acid………………………………………………...……………………………….......31

3.7 XRD pattern of the reduced UGS after 0.1 M HCl leaching at 140oC......................34

3.8 SEM image of reduced UGS after hydrochloric acid leaching……………………35

3.9 Residual plot for iron removal……………………………………………………36

3.10 Predicted and actual iron removal versus HCl concentration (a) Fe removal versus

HCl concentration at 90oC (b) Fe removal versus Temperature in 0.1 M HCl……………38

4.1 The three size distributions investigated in this work and the normal distribution

(CV=0.3)…………...…..........……………………….............….…………………...46

4.2 Fraction reacted versus normalized time for the GGS distribution (a) Interface reaction

Page 10: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

ix

control and (b) Pore-diffusion control………….………………………………………...53

4.3 Fraction reacted versus normalized time for the RRB distribution (a) Interface reaction

control and (b) Pore-diffusion control………...………………………………………….56

4.4 Fraction reacted versus normalized time for the Gamma distribution (a) Interface

reaction control and (b) Pore-diffusion control…………...………………………………58

4.5 Comparison for the three size distributions; Fp=3 and CV=0.3 and 1.5 (a) Interface

reaction control and (b) Pore-diffusion control………………………………………….60

5.1 Particle size of reduced UGS after acetic acid washing………………………….65

5.2 Rela t ionship be tween the values o f (1 - (1-X)) 1 / 3 and t ime a t t h ree

temperatures……………………………………..……………………………………….67

5.3 Arrhenius Plot………………………...………………………………........................68

5.4 Possible schematic diagram of the reduced UGS (Brown: TiH2, Black: FexOy, or Fe,

blue: other impurities, gray: hydrogen ion). In this scenario, the reaction is controlled by

solid state diffusion of iron……………………………...……………………………...70

5.5 Comparison of predicted values produced from the equation for solid state diffusion

and experimental value; 0.1 M HCl, 1,000 rpm, 90oC, no boric acid………….………….72

5.6 Comparison of predicted values produced from the equation for solid state diffusion

and experimental value until 4 h of leaching time…………………………..…………….75

5.7 Iron removal versus time at three rotation speed………………………………....…...76

6.1 (a) SEM image of titanium-tannic acid complex, (b) the composition of titanium-tannic

acid complex analyzed by EDS………………………………………………………...81

6.2 FT-IR spectra of (a) tannic acid (b) titanium-tannic acid complex……….…………...82

6.3 (a) The schematic diagram of titanium and tannic acid and (b) optimized structure by

DFT simulation at B3LYP with 6-31G: gray (carbon), red (oxygen), white and big atom

(titanium), white and small atom (hydrogen)………………………………...………...…84

6.4 The molecular orbitals’ energy levels of titanium-tannic acid complex at LUMO+1,

LUMO, HOMO, and HOMO-1: green (positive isosurface of molecular orbitals at the

energy level of each molecular orbital), red (negative isosurface of molecular orbitals at

the energy level of each molecular orbital)……….………………………………………85

Page 11: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

ACKNOWLEDGEMENTS

The research in this thesis was carried out with financial support of the Direct

Reduction of Titanium Slag (DRTS) by Advanced Research Projects Agency-Energy

(ARPA-E) of the US Department of Energy (DOE) (Grant # 55800707).

Firstly, I would like to thank my supervisor, Dr. Michael L. Free for his meticulous

guidance and support on the research. It was very fortunate to meet him and become one

of his students. What I learned from him will be one of the most lasting experiences and

lessons in my life. I would like to express sincere appreciation to Dr. H.Y. Sohn for giving

me an opportunity to work on one of the most fascinating fields, fluid-solid reaction

engineering, in my course work. Also, I would like to thank Dr. Z. Zak Fang and Dr.

Amarchand Sathyapalan for their academic guidance for two years and for helping me

complete this thesis.

Secondly, I would like to thank my lab mate, Syamantak Roy who works on the

same project as me. We have gone through several difficulties in the research and course

work together. Also, I would like to thank other lab mates, Dr. Prashant Sarswat, Yakun

Zhu, Gaosong Yi, Weizhi Zeng, Joshua Werner, Sayan Sarkar, and Erik Sundberg for their

encouragement, training, and priceless suggestions.

Page 12: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 1

INTRODUCTION

1.1 Direct reduction of titanium slag

Pure metallic titanium is of great industrial importance because of its use in various

applications such as aviation, aerospace, biomedical, marine, and nuclear waste storage due

to its corrosion resistance, high specific strength, light weight, relatively high melting point,

and chemical/heat stability.1-5 Moreover, it has great mechanical, and electronic properties,

as well as biocompatibility.6 Although it is predominantly found as rutile (TiO2) and

ilmenite (FeTiO3), which are abundant in the earth’s crust, the production cost is relatively

high compared to competing metals such as steel and aluminum.7 In general, metallic

titanium is produced from the Kroll process in which greenhouse gas is generated8 and it

requires an upgraded starting material, which increases the cost of production. Thus, it can

be said that an alternative route, which does not involve the environmental and cost

problems for production of metallic titanium, is desired. Recently, Fang et al. suggested

titanium production from titanium hydride (TiH2), which can be obtained by reduction of

upgraded titanium slag (UGS) or synthetic rutile under a controlled environment in the

presence of metallic magnesium and salts.9 After making titanium hydride, metallic

titanium is produced by dehydrogenation of titanium hydride at about 500oC9 with auxiliary

processes such as size, shape control, and deoxygenation.

In this new route, production cost and environmental problems can be significantly

Page 13: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

2

reduced and productivity can be improved compared to the Kroll process. The feed material,

upgraded titanium slag (UGS), is obtained by reduction of ilmenite ore in a blast furnace

followed by roasting and acid leaching. However, one of the concerns in this process is the

level of residual impurities such as iron oxide.

In this work, investigation of the leaching characteristics was carried out using

hydrochloric and boric acids associated with the removal of iron from UGS which has been

reduced using hydrogen gas, magnesium powder, and salts at 750oC for 5 h.9 Reduced UGS

is predominantly composed of titanium hydride and oxide impurities involving iron, silicon,

aluminum, and/or magnesium. According to the American Society for Testing and

Materials (ASTM), the maximum tolerable amount of impurities in titanium sponge for

general purpose is 0.08, 0.15, 0.05, and 0.04% by weight for magnesium, iron, aluminum,

and silica, respectively.10 Thus, one of the goals of this thesis research is iron removal from

reduced UGS to meet specifications. The flow chart shown in Figure 1.1 illustrates the

overall production of pure titanium from UGS. In this investigation, modest concentrations

of hydrochloric and boric acids were used to reduce product loss. Boric acid was used to

accelerate iron removal kinetics, because boric acid and borate tend to form complexes

with hydrous iron oxides as shown in Equation 1.1.11

𝐹𝑒3+ + 𝐻3𝐵𝑂3 + 𝐻2𝑂 = 𝐹𝑒𝐵(𝑂𝐻)42++ 𝐻+ (1.1)

The stability constant of iron borate complexes is 1.0 ± 0.2 × 10−2 25oC for an

ionic strength of 0.68.12

Page 14: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

3

Figure 1.1 Flow sheet of new route for production of metallic titanium

Page 15: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

4

1.2 Theoretical kinetic study

Fluid-solid reaction kinetics is important in chemical and metallurgical processes

such as the extraction of metals, the combustion of solid fuels, and coal gasification.

Various reaction models have been investigated for various fluid-solid reaction systems.

The shrinking unreacted-core model is one of the simplest models commonly used for an

initially nonporous solid producing a porous product layer.13-15 In this model the reaction

occurs at a sharp interface after reactant fluid diffuses through the porous product layer. As

the reaction proceeds, the size of the unreacted core diminishes. However, the overall size

of the solid remains constant. Figure 1.2 schematically shows the reaction of a nonporous

solid according to this model.

It is well known that the particle size of solid materials is rarely uniform. In most

cases, the particle size of solid materials follows certain distributions. Among, the most

commonly used forms of empirical size distributions in industry are the Gate-Gaudin-

Schuhmann (GGS) distribution, the Rosin-Rammler-Bennett (RRB) distribution, and the

Gamma distribution. All of the size distributions have two parameters that determine mean,

variance, and the coefficient of variation (CV), which is the ratio of the standard deviation

to the mean of the size distribution, which is also called relative standard deviation. The

best fit values of the parameters for each distribution can be determined directly from plots

or from computed values of the mean and variance.16-18

This study investigated the overall fluid-solid reaction kinetics when the particle

size distribution follows the above-mentioned three empirical size distributions with either

interface-reaction or pore-diffusion-controlled kinetics for the shrinking unreacted-core

model.

Page 16: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

5

Figure 1.2 The shrinking unreacted-core system

Page 17: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

6

The CV of the particle size was varied over a wide range from 0 to 1.5. Also, three

basic particle shapes were considered. The results of this study would be applied to the

leaching characteristics of reduced UGS.

1.3 Selective extraction of titanium from ilmenite using tannic acid

The merits of a metal production process have to be judged not only by the

metallurgical quality of the product obtained but also by its impact on environment and

health. It has been reported that abnormalities in pulmonary function and pleural disease

among people associated with titanium manufacturing can be correlated to exposure to

titanium tetrachloride, titanium dioxide, and duration of work in titanium production.19

In this regard, additional work was carried out to evaluate the ability of tannic acid

to selectively chelate with titanium ions from a multielement solution containing iron,

vanadium, silicon, aluminum, and other related ions, prepared by dissolving a complex ore

such as ilmenite. This route can decrease the significant number of processes for the

preparation of titanium pigment, which is a starting material for the route of direct reduction

of titanium slag (DRTS). That is, metallic titanium can be produced by dehydrogenation of

titanium hydride (TiH2), obtained by the reduction of titanium-tannic acid complex in the

presence of hydrogen. Using the special ability of tannic acid, titanium can be extracted

more environmentally and economically. Also, to acquire a deeper scientific insight into

the complexation and selectivity for titanium chelation, density functional theory (DFT)

simulations have been performed on the complex in addition to the experimental work.

Page 18: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 2

REVIEW OF THE LITERATURE

2.1 Processes for titanium production

2.1.1 Kroll Process

The current commercial method for production of metallic titanium is the Kroll

process in which metallic titanium is obtained by magnesiothermic reduction of titanium

tetrachloride (TiCl4).20 Titanium dioxide obtained by treating rutile and ilmenite is reduced

in the presence of cokes and chlorine gas in a fluidized bed reactor at 1,000oC. The resulting

titanium tetrachloride and impurities are separated using continuous fractional distillation.

Then, titanium tetrachloride is reduced to titanium sponge by magnesium at 850oC for

several days, depending on the size of the reactor, and liquidized magnesium chloride is

obtained as the main by-product. The separation of magnesium chloride from titanium

sponge can be performed by vacuum distillation.21 The chemical reactions can be expressed

as Equations 2.1 and 2.2.

𝑇𝑖𝑂2(𝑠) + 𝐶(𝑠) + 2𝐶𝑙2(𝑔) → 𝑇𝑖𝐶𝑙4(𝑔) + 𝐶𝑂2(𝑔) (2.1)

𝑇𝑖𝐶𝑙4(𝑔) + 2𝑀𝑔(𝑙) → 𝑇𝑖(𝑠) + 𝑀𝑔𝐶𝑙2(𝑙) (2.2)

Although the Kroll process allows industry to produce high-purity metallic

titanium with low oxygen levels, there are several challenges. Firstly, developing a

continuous reduction process is difficult due to the high vacuum, long-term processing, and

Page 19: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

8

the generation of titanium deposits on the inner wall of the reactor. Secondly, batch-type

processing for long periods of time results in high labor and power, and low productivity.21

Furthermore, environmental issues such as generation of greenhouse gas and corrosive

intermediate products like titanium tetrachloride are involved.6, 8 In summary, the Kroll

process has several major drawbacks such as extensive time and energy consumption, low

productivity, and environmental problems.

2.1.2 Alternative processes

To mitigate the challenges of the Kroll process, several new alternative techniques

with low cost and high purity, and productivity have been suggested.

The FFC Cambridge process is the one of the alternative processes developed to

replace the Kroll process, in which electrochemistry is utilized unlike other processes

which use pyrometallurgy.23 Titanium dioxide obtained from either sulfate or chloride

processes is used to form a sintered rectangular cathode using traditional ceramic

processing through which a conductive wire is inserted. The titanium dioxide cathode is

immersed in calcium chloride electrolyte with an anodic carbon electrode. As voltage is

applied, the oxygen in the titanium dioxide cathode is ionized in the molten calcium

chloride, transported to the carbon electrode and gasified as forms of oxygen, carbon

monoxide, and carbon dioxide.23 After that, pure metallic titanium remains in the cathode.

This process does not involve the complicated chemical reactions and corrosive titanium

tetrachloride to reduce titanium dioxide that are common in other pyrometallurgical

processes. However, the production cost of the sintered titanium dioxide cathode is fairly

high and the electrochemical processing time is long in order to achieve low levels of

oxygen in the titanium cathode. Moreover, this process includes greenhouse gases such as

Page 20: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

9

carbon monoxide and carbon dioxide. Figure 2.1. shows a schematic diagram of the FCC

Cambridge process.

Another alternative process to the Kroll processes is the Armstrong process. This

process adopted sodium as a reducing agent which is also utilized in the Hunter process.

However, the Armstrong process is a continuous process, rather than a batch process used

for the Hunter process.24 Titanium dioxide is reduced to titanium tetrachloride in the

presence of chlorine gas and cokes. The titanium tetrachloride is then reduced by liquid

sodium in a titanium reactor where continuous processing takes place. Sodium can be

removed using distillation and sodium chloride can be separated from titanium sponge by

washing. All sodium and sodium chloride are recycled in this process. The reaction that

takes place can be expressed as Equation 2.3.

𝑇𝑖𝐶𝑙4(𝑔) + 4𝑁𝑎(𝑙) → 𝑇𝑖(𝑠) + 4𝑁𝑎𝐶𝑙(𝑙) (2.3)

This process can become a viable commercial process if it can overcome

equipment durability, optimization of each process, and particle size and morphology

control.23 Titanium subchlorides such as titanium dichloride (TiCl2) and titanium

trichloride (TiCl3) have been investigated as potential replacements for titanium

tetrachloride (TiCl4) feed. Processing using subchlorides keeps the advantages of the Kroll

process such as an oxygen-free environment and easy impurity-control and may increase

the process speed22 However, concerns such as handling volatile titanium tetrachloride

require further technical and practical studies before industrial application.

Another alternative technique is direct reduction from titanium dioxide by

electrodeposition of titanium from ionic solutions, but this approach has several problems

like difficulties in handling the redox cycle of multivalent titanium ions and reactive

Page 21: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

10

Figure 2.1 The scheme of TiO2 reduction in the FCC Cambridge process

Page 22: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

11

dendritic products23. This process which directly deposits titanium on the cathode from the

titanium ion bearing solution is different from the FFC Cambridge process where oxygen

is removed from titanium dioxide cathode.

Although new technologies for production of metallic titanium without drawbacks

of Kroll process have been suggested and investigated extensively, none of these

technologies have been industrialized on a larger scale due to their intrinsic disadvantages

and limits.

2.1.3 Extraction of titanium from ores

In general, the extractions of titanium dioxide (TiO2) and titanium pigment are

performed using mainly two routes, sulfate and chloride processes. In the sulfate process,

low grade ilmenite (FeTiO3) can be used as a feed stock which is treated by concentrated

sulfuric acid (H2SO4). In the next step, impurities are precipitated by cooling down the

sulfuric acid solution and titanium dioxide can be obtained by performing hydrolysis on

titanyl sulfate (TiOSO4) at 95-110oC, and calcination at 1,000oC.25 On the other hand, the

chloride process requires a high grade feed stock such as rutile (TiO2). The oxygen in the

feed material rutile is removed by a reducing agent, usually carbon, and titanium

tetrachloride (TiCl4) is synthesized using chlorine gas. After that, pure titanium dioxide can

be obtained by oxidizing titanium tetrachloride at 1200-1700oC.26 However, there are

critical drawbacks in these processes such as environmental problems and high cost of

production. Alternate routes are still required to produce low cost, environmentally friendly

titanium from its naturally occurring ores such as ilmenite, by reducing prior treatments.

Alternatively, extraction of titanium using hydrochloric acid has been investigated

extensively due to its low cost, reduced environmental issues, and potential for recycling

Page 23: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

12

compared to competing lixiviants such as sulfuric acid. However, most of the work has

focused on extraction of titanium from ores, not reduced UGS, which is the starting

material for extraction of titanium in this thesis.27-29

Titanium can be extracted from ilmenite dissolution using hydrochloric acid. The

reaction can be described as Equation 2.4.

𝐹𝑒𝑇𝑖𝑂3 + 𝐻𝐶𝑙 → 𝐹𝑒2+ + 𝑇𝑖𝑂2+ + 2𝐶𝑙− + 2𝑂𝐻− (2.4)

According to El-Hazek et al., complete dissolution of titanium and iron could be

achieved using 12 M hydrochloric acid at a solid/liquid ratio of 1/20 at 80oC for 2.5 h.27

However, when the resistance of titanium hydride to acidic media is taken into account,

hydrochloric acid concentration should be much lower than 12 M in our work.

2.2 Effect of particle size distribution on kinetics

Generally, the most basic probability distributions are known to be normal and log-

normal distribution. The probability density function of the normal distribution can be

given as

ℎ(𝑅) =1

𝜎√2𝜋𝑒𝑥𝑝(−

(𝑅−𝜇)2

2𝜎2) (2.5)

where 𝜇 is the mean and 𝜎 is the standard deviation. Although the normal distribution is

important in statistics and social science, it is rarely encountered in a practical system of

engineering. The probability density function of log-normal distribution, the logarithm of

which is normally distributed can be expressed as Equation 2.6.

ℎ(𝑅) =1

𝜎√2𝜋𝑅𝑒𝑥 𝑝 (−

(𝑙𝑛𝑅−𝜇)2

2𝜎2) , 𝑅 > 0 (2.6)

If variable R is a particle size of a certain system that we are interested in, then the

Page 24: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

13

probability distribution can be understood as the particle size distribution. The adjustable

parameters for the normal and log-normal distribution can be used to determine the mean

and the standard deviation. To compare these distributions, the CV is used to normalize the

mean and variance of each distribution. Table 2.1 shows mean, variance, and mathematical

relationship among adjustable parameters and CV. Also, when CV is fixed at 0.3, two

probability distributions can be plotted as shown in Figure 2.2.

It is known that the effect of particle size distribution of solid materials treated in

mineral processing occupies an important position in determining the kinetics of the

reaction. Nonetheless, in most models, it is assumed that the particle size and shape of

reacting solid materials are uniform. Only few researchers have studied the effect of the

size distribution on the fluid-solid reaction kinetics. Hulburt, Katz, Randolph, and Larson

observed that a mathematical expression for a reaction taking place in a system cannot

represent the actual system which has a distribution of particle properties.30-32 They

suggested a concept of population balance where changes in the distribution of chosen

particle properties can be described. Mcllvried and Massoth narrowed down the focus of

the particle properties to the particle size, and particularly considered normal and log-

normal particle size distributions for nonporous spherical particles and either a chemical-

reaction-controlled or pore-diffusion-controlled process.33

The conclusion was that the overall reaction rate of a particle assemblage with a

normal or log-normal particle size distribution was not significantly different from those

with a uniform mass-average size as long as the CV of the distribution was less than 0.3.

The particle size distribution found in nature or formed by mineral processing rarely

follows a normal particle size distribution. In fact, it is convenient to use empirical

Page 25: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

14

Table 2.1 Mean, variance, and CV of three distributions

Name Normal distribution Log-normal distribution

Adjustable

parameters 𝜇, 𝜎 𝜇, 𝜎

Mean 𝜇 𝑒𝜇+𝜎2/2

Variance 𝜎 (𝑒𝜎2− 1)𝑒2𝜇+𝜎

2

CV 𝜎

𝜇 √𝑒𝜎

2− 1

Page 26: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

15

Figure 2.2 The normal and log-normal distributions (𝑪𝑽 = 𝟎. 𝟑)

Page 27: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

16

relationships which offer concise mathematical expressions of size distributions.34

2.3 Tannic acid

It has been reported that hydroxyaromatic compounds can selectively form a

complex with titanium compounds at pH 4-535-37. Tannic acid is such a compound and it

possesses carboxylic acid groups expected to show a strong affinity towards metal ions

having high oxidation states.35, 38 Also, tannic acid in water is hydrolyzed to digallic acid

and glucose.39 Figure 2.3 shows the molecular structure of tannic and digallic acids. By

adding tannic acid in the titanium bearing solution and adjusting the pH between 4 and 5

during precipitation, a titanium-tannic acid complex can be precipitated selectively from a

multielement solution, containing dissolved impurities with iron, magnesium, vanadium,

and manganese as ions. Moreover, tannic acid is commonly found in beverages, food, and

plant materials.40 Using such an ecofriendly material to separate titanium can minimize the

environmental and health concerns associated with current industrial titanium production

processing.

Page 28: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

17

Figure 2.3 The structure of (a) tannic acid and (b) dissociation of tannic acid in

aqueous solution as digallic acid

Page 29: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 3

LEACHING CHARACTERISTICS OF IRON REMOVAL

FROM REDUCED UGS

3.1 Experimental methods

3.1.1 Preparation of reduced UGS by reduction and acetic acid leaching

UGS is composed of about 96% titanium dioxide and 4% impurities. A mixture of

UGS (less than 40 μm ), magnesium powder, and salts, such as magnesium chloride

(MgCl2) and potassium chloride (KCl), with a ratio of 1:1:1 by weight were placed in a

furnace under hydrogen atmosphere and reduced for 5 h at 750oC. Excess magnesium and

salts are used to help increase kinetics.9 The product resulting from the reduction is washed

using 4.3 M acetic acid (CH3COOH) with a solid to liquid ratio of 1 g to 40 ml for 2 h at

70oC to remove magnesium compounds and salts from reduced UGS. After acetic acid

washing and filtration, the residue was washed with deionized water and 0.05 M

hydrochloric acid. The sample so obtained is used as the feed in this study.

3.1.2 Experimental procedures

A solution of 200 ml which contained 0.05 or 0.1 M hydrochloric acids with or

without 1 M boric acid was prepared in a flask made of semiclear polymethylpentene. A

chemical and temperature resistant plug with a hole that allows for a thermometer insertion

was applied on the top of the flask to minimize evaporation losses. The solution was mixed

Page 30: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

19

by a magnetic stirring bar at a rotational speed of 1,000 rpm and heated to the

predetermined temperature, which is lower than the boiling point of water (70, 80, and

90oC). Once the temperature reached the desired level, 2 g of reduced UGS obtained after

acetic acid washing was added into the flask. The leaching experiment was run for 4 h and

samples were collected every hour for the study of reaction kinetics. In the cases where the

experiments were conducted at higher temperature than the boiling point of water (140 and

190oC), a different experimental setup, which involved a pressure reactor, was used. A

solution of 200 ml which contained a low concentration of hydrochloric and boric acids

with 2 g of reduced UGS after acetic acid washing was prepared in a

polytetrafluoroethylene beaker and experiments were carried out in the pressure reactor.

The desired temperature was achieved after 1 h, and the experiments were carried out for

3 h at the desired temperature. The pressure reactor was slowly cooled for 2 h. After that,

the solution was filtered and residue was washed with deionized water and 0.05 M

hydrochloric acid. The solution samples were analyzed by inductively coupled plasma-

optical emission spectroscopy (ICP-OES). The solid samples were characterized using X-

ray diffraction (XRD) and scanning electron microscopy (SEM).

3.2 Experimental results and discussion

3.2.1 Feed analysis

Table 3.1 shows the composition of reduced UGS immediately after reduction.

Magnesium compounds and salts in Table 3.1 come from magnesium oxide, excess

magnesium powder, magnesium chloride, and potassium chloride. The composition of the

acetic acid washed reduced UGS is also listed in Table 3.1, showing the impurities which

predominantly consist of iron, aluminum, magnesium, and silicon oxides.

Page 31: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

20

Table 3.1 Chemical composition of reduced UGS before and after acetic acid washing

Component (%) Mg compounds

and salts Ti Fe Al Si

Before washing 79.79 19.52 0.22 0.07 0.41

After washing 0.36 96.20 1.07 0.35 2.02

Page 32: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

21

Figure 3.1 presents the mineralogical phases of the acetic acid washed; reduced

UGS analyzed by XRD shows that the crystalline phases of the products predominantly

consist of titanium hydride and ferric oxide.

Figure 3.2 shows the surface morphology of reduced UGS after acetic acid

washing as analyzed by SEM. This result can be used for comparison of surface

morphology with that of the product after leaching in hydrochloric and boric acids.

Although the sizes of individual particles seem very small, Figure 3.2 reveals larger

agglomerates.

3.2.2 Experimental results

Table 3.2 shows the effect of each parameter on iron removal and titanium loss

after 4 h of leaching. Again, the leaching temperature, hydrochloric acid concentration, and

boric acid concentration were chosen as main factors to verify the optimized leaching

parameters. Three values for leaching temperature, two values for hydrochloric acid, and

boric acid concentrations give 12 different experiments in total. By analyzing those

experimental results, the most important leaching parameter affecting iron removal can be

investigated.

3.2.2.1 Effect of hydrochloric acid

One of the difficulties in the leaching of reduced UGS is that titanium hydride is

easily dissolved in strong acid. In particular, titanium hydride tends to be converted to

titanium trichloride in high concentrations of hydrochloric acid. The dissolution of titanium

hydride can be expressed as41

2𝑇𝑖𝐻2 + 6𝐻+ + 6𝐶𝑙− = 2𝑇𝑖𝐶𝑙3 + 5𝐻2 (3.1)

Page 33: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

22

Figure 3.1 XRD pattern of the reduced UGS after acetic acid washing

Page 34: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

23

Figure 3.2 SEM image of reduced UGS after acetic acid washing

Page 35: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

24

Table 3.2 Fe removal % and Ti loss % at different leaching parameters with a solid to

liquid ratio of 1 to 100 and 4 h of leaching time

Temp (oC) HCl (M) Boric acid (M) Fe removal (%) Ti loss (%)

90 0.05 0 27.30 0.07

90 0.05 1 43.80 0.08

90 0.1 0 55.01 0.50

90 0.1 1 61.42 0.60

140 0.05 0 79.59 0.06

140 0.05 1 76.01 0.15

140 0.1 0 87.63 0.21

140 0.1 1 87.82 0.25

190 0.05 0 79.19 0.22

190 0.05 1 81.17 0.22

190 0.1 0 88.26 0.12

190 0.1 1 88.51 0.34

Page 36: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

25

Thus, before conducting the regular experiments with hydrochloric acid, the acceptable

quantity of titanium hydride loss during the leaching process needs to be determined. Our

research group determined arbitrarily that 1% (w/w) of titanium loss is the maximum

amount tolerable. Test results for titanium loss as a function of hydrochloric acid

concentration are presented in Figure 3.3 under the leaching parameters in Table 3.3.

Based on the results in Figure 3.3, the maximum concentration of hydrochloric

acid that meets the requirement of less than 1% titanium loss is approximately 0.1 M. Thus,

0.05 M and 0.1 M hydrochloric acids were chosen to study the influence of hydrochloric

acid concentration on the removal of iron as shown in Figure 3.4 under the leaching

parameters in Table 3.3. The leaching with 0.1 M hydrochloric acid shows 55.0% of iron

removal after 4 h along with 0.5% of titanium loss, whereas 0.05 M hydrochloric acid

removed only 27.3% of the iron. Much of the iron removal taking place during leaching

can be expressed generally as,

𝐹𝑒2𝑂3 + 6𝐻+ + 6𝐶𝑙− = 2𝐹𝑒𝐶𝑙3 + 3𝐻2𝑂 (3.2)

3.2.2.2 Effect of boric acid

Figure 3.5a shows the effect of boric acid in 0.05 M hydrochloric acid. The iron

removal efficiency significantly increased from 27.3% to 43.8% with the increase of boric

acid. However, in 0.1 M hydrochloric acid, the iron removal efficiency increased by only

6.4% when boric acid was added according to Figure 3.5b.

According to Peak et al., trigonal boric acid reacted with ferric iron would be

converted to tetrahedral surface complexes as a result of the Lewis acidity of the boron

metal center.11 Then hydrogen ions would be displaced from the tetrahedral boron, and

ferric borate complexes are obtained as shown in Equation 1.1.11, 12 However, this reaction

Page 37: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

26

Figure 3.3 Titanium loss versus HCl concentration

Page 38: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

27

Table 3.3 Leaching parameters

Temperature

(oC)

Time

(h)

Stirring speed

(rpm)

S/L

(g : ml)

Particle size

(𝛍𝐦)

90 4 1,000 1:100 0.05 – 0.2

Page 39: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

28

Figure 3.4 Comparison of iron removal and time for 0.05 M and 0.1 M HCl at 90oC

Page 40: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

29

Figure 3.5 Comparison of iron removal versus time with/without boric acid at 90oC

Page 41: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

30

is only favorable in weak acid, which might explain why boric acid does not affect iron

removal much at the higher concentration of hydrochloric acid (Figure 3.5).

3.2.3 Effect of temperature

The previous results show that the concentration of hydrochloric acid is restricted

to 0.1 M to prevent significant titanium loss, and boric acid does not help iron removal

much in the presence of 0.1 M hydrochloric acid in the solution. Thus, the effect of leaching

temperature on iron removal was studied in the broad temperature range of 70 – 190oC,

with 0.1 M hydrochloric acid and with/without 1 M boric acid. Figure 3.6 shows that the

iron removal increases with the increase of temperature. It is noted that once the

temperature reaches about 140oC, no additional iron was removed. Thus, it appears that the

limit of iron removal is near 90% if the titanium loss is kept below 1%. It is also noted that

iron removal without boric acid shows almost the same iron removal result as that with

boric acid as shown in Figure 3.6. This result suggests high temperature and high acidity

are not favorable for boric acid to form tetrahedral surface complexes with iron.

Therefore, the parameters that show the best results for iron removal based on the

series of tests presented are 140oC, 0.1 M hydrochloric acid and no boric acid. Weight

percentage of iron removal under those optimized parameters is 87.63%, which means that

only 0.13% of iron remains in the leached, reduced UGS. The percentage of titanium was

significantly changed from 96.2% to 99.45% after substantial impurity removal. According

to the ASTM, the maximum tolerable amount of iron impurities in titanium sponge for

general purpose is 0.15% by weight.10 Thus, the reduced UGS that has been purified under

the optimized parameters meets ASTM B299-13 iron specifications for general purposed

titanium sponge. The whole composition of the reduced UGS leached by 0.1 M

Page 42: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

31

Figure 3.6 Fe removal and Ti loss versus temperature with 0.1 M HCl and with/without

boric acid

Page 43: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

32

hydrochloric acid with/without boric acid at 140oC is tabulated in Table 3.4. The sample

was analyzed by ICP-OES.

Figure 3.7 shows the XRD pattern of the reduced UGS obtained after 0.1 M

hydrochloric acid leaching at 140oC. According to Figure 3.7, little titanium hydride was

oxidized to titanium dioxide or metallic titanium, and most of impurities were below the

detection limit for XRD.

Figure 3.8 shows the surface morphology of the reduced UGS after 0.1 M

hydrochloric acid leaching at 140oC. Comparing it to the SEM image of reduced UGS

before hydrochloric acid leaching (Figure 3.2), it is apparent that the surface morphology

is nearly the same. The particle shape remains granular, and agglomerates are formed.

3.2.4 Factorial design of experiments for equation modeling

A model that can predict iron removal with the three parameters was established.

Three levels of temperature (90, 140, 190oC) and two levels of concentration of

hydrochloric (0.05, 0.1 M), and boric acids (0, 1 M) were used as shown in Table 3.2. The

factorial design of experiments was analyzed by the statistical software “Minitab”. The

equation modeling for prediction of iron removal with three parameters is described in

Equation 3.3.

Iron removal (wt%) = -164.116 + 2.56685 T + 677.07 [HCl] + 18.0998 [H3BO3] -

0.00687205 T2 - 2.89343 T [HCl] - 0.103388 T [H3BO3] (3.3)

where T, [HCl], and [H3BO3] represent temperature in degrees Celsius, molarity of

hydrochloric and boric acids, respectively, and the R2(adj.) value of the equation is 0.9541.

Also, in Figure 3.9, the residual versus order plot shows randomness and unpredictability

where the sense of accuracy of the model is observed.

Page 44: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

33

Table 3.4 Chemical composition of reduced UGS leached by 0.1 M HCl at 140oC

Component (%) Mg Ti Fe Al Si

0.1 M HCl 0.05 99.45 0.13 0.12 0.25

0.1 M HCl + 1 M H3BO3 0.07 99.13 0.13 0.10 0.57

Page 45: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

34

Figure 3.7 XRD pattern of the reduced UGS after 0.1 M HCl leaching at 140oC

Page 46: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

35

Figure 3.8 SEM image of reduced UGS after hydrochloric acid leaching

Page 47: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

36

Figure 3.9 Residual plot for iron removal

Page 48: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

37

Figure 3.10a shows a comparison of iron removal at different concentrations of

hydrochloric acid for predicted and actual results at 90oC and Figure 3.10b presents a

comparison of iron removal as a function of temperature with 0.1 M hydrochloric acid.

Overall, Figures 3.10a and b show good agreement between predicted iron removal and

actual iron removal. Note that although some of the data in Figure 3.10a show high iron

removal, concentrations above 0.1 M hydrochloric acid result in significant titanium loss.

3.2.5 Removal of iron without acetic acid leaching

The starting material in this research was reduced UGS which was washed by

acetic acid to remove magnesium oxide and salts introduced from the reduction. Another

way of removing magnesium oxide and salts is to use hydrochloric acid instead of acetic

acid which allows us to remove iron and magnesium oxides together. This concept can help

reduce the number of leaching steps required for removal of impurities and results in

reduction of the cost for production of metallic titanium. Using this concept experimentally,

2 g of the powder taken out from the reduction furnace was prepared and leached in 0.4 M

hydrochloric acid at 70oC for 2 h to remove magnesium oxide and salts. After that, the

mixture of the leach liquor and the powder was put in the pressure reactor and leached at

140oC for 3 h to remove iron oxide. Table 3.5 shows the composition of reduced UGS

leached by 0.4 M hydrochloric acid at 70oC and 140oC for 2 h and 3 h, respectively. The

amount of iron and magnesium in the reduced UGS leached by 0.4 M hydrochloric acid

meets the ASTM B299-13 iron and magnesium specifications for general purposed

titanium sponge.10 Thus, the result indicates that this route has a strong potential to reduce

the production cost as well as produce the qualified titanium sponge which meets the

ASTM specifications.

Page 49: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

38

Figure 3.10 Predicted and actual iron removal versus HCl concentration (a) Fe removal

versus HCl concentration at 90oC (b) Fe removal versus Temperature in 0.1 M HCl

Page 50: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

39

Table 3.5 Composition of reduced UGS leached by 0.4 M hydrochloric acid

Composition Ti (%) Fe (%) Si (%) Al (%) Mg (%)

w/o acetic acid

leaching 99.25 0.06 0.49 0.20 <0.03

ASTM

Specific cation - 0.15 0.04 0.05 0.08

Page 51: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 4

EFFECT OF PARTICLE SIZE DISTRIBUTION ON KINETICS

4.1 Model formulation

4.1.1 Shrinking unreacted core model

Consider a reaction where the product solid forms a porous layer around an

unreacted nonporous core:

𝐴(𝑓) + 𝑏𝐵(𝑠) → 𝑐𝐶(𝑓) + 𝑑𝐷(𝑠) (4.1)

If the overall rate of the reaction is controlled by the interface chemical reaction,

the progress of the reaction is represented for the three basic geometries by13

𝑑𝜉

𝑑𝑡∗= −1 (4.2)

where

𝜉 ≡ (𝐴𝑝

𝐹𝑝𝑉𝑝) 𝑟𝑐 (4.3a)

𝑡∗ = (𝑏𝑘

𝛼𝐵𝜌𝐵) (

𝐴𝑝

𝐹𝑝𝑉𝑝) [𝐶𝐴𝑂

𝑛 −𝐶𝐶𝑂𝑚

𝐾] 𝑡 (4.3b)

where 𝐴𝑝 and 𝑉𝑝 are the original surface area and volume; 𝑟𝑐 the position of the

reaction interface in the distance coordinate perpendicular to the solid surface; and 𝐹𝑝 is

the shape factor, which has the value 1, 2, or 3 for infinite slabs, long cylinders, or spheres,

Page 52: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

41

respectively. 𝑘 is the heterogeneous rate constant, 𝛼𝐵 the fraction of pellet volume

occupied by solid reactant B, 𝜌𝑠 the molar density of the solid B, K the equilibrium

constant, 𝐶𝐴𝑂 and 𝐶𝐶𝑂 are the concentrations of A and C in the bulk fluid stream,

respectively. The fraction reacted of an individual particle, β, is given by

β = 1 − ξ𝐹𝑝 (4.4)

Integration of Equation 4.2 and substitution in Equation 4.4 gives

β(𝑡, 𝑅) =

{

1 − (1 −

𝑘𝐶𝑅𝑡)𝐹𝑝

at 𝑡 <𝑅

𝑘𝐶 (4.5a)

1 𝑎𝑡 𝑡 ≥𝑅

𝑘𝐶 (4.5b)

where

𝑘𝐶 = (𝑏𝑘

𝛼𝐵𝜌𝑠) [𝐶𝐴𝑂

𝑛 −𝐶𝐶𝑂𝑚

𝐾] (4.6)

𝑅 =𝐹𝑝𝑉𝑝

𝐴𝑝 (4.7)

𝑘𝐶 is a new constant incorporating system parameters, 𝑅 is the initial half thickness of

an infinite slab and initial radius of a long cylinder or a sphere, and β(𝑡, 𝑅) stands for the

fraction reacted of a particle with initial radius 𝑅 at a certain time 𝑡 . Equation 4.5b

represents the fact that once a particle is completely reacted, there is no longer any reaction

taking place in it.

If the overall rate of reaction is controlled by diffusion through the porous product

layer of a spherical particle, the mass balance for the reactant fluid species in the porous

product layer yields:

Page 53: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

42

𝑑

𝑑𝑟(𝑟𝐹𝑝−1𝑁𝐴) = 0 (4.8)

where

𝑁𝐴 = 𝑥𝐴(𝑁𝐴 + 𝑁𝐶) − 𝐷𝑒𝐶𝑇∇𝑥𝐴 (4.9)

𝑁𝐴 is the molar flux of species A, 𝑁𝐶 the molar flux of species C, 𝑥𝐴 mole fraction of

fluid A, 𝐷𝑒 effective diffusivity, and 𝐶𝑇 total molar concentration of fluid. The boundary

conditions are given by

𝐶𝐴 = 𝐶𝐴𝑂 𝑎𝑛𝑑 𝐶𝐶 = 𝐶𝐶𝑂 at 𝑟 = 𝑅 (4.10)

𝐶𝐴 = 𝐶𝐶/𝐾 at 𝑟 = 𝑟𝑐 (4.11)

The first boundary condition assumes rapid external mass transfer, which has been

shown to provide only a secondary effect to pore diffusion for the reaction of an individual

solid,12 and thus should not significantly affect the conclusions on the effect of particle size

distribution. For further mathematical simplicity, we will consider the case in which c in

Equation 4.1 is unity. Then, from the mole balance of species A and C, the following

equation is obtained:

𝐶𝐴 + 𝐶𝐶 = 𝐶𝐴𝑂 + 𝐶𝐶𝑂 (4.12)

Integrating Equation 4.8 with the boundary conditions and combining Equation

4.8 and Equation 4.12 gives42

𝑘𝐷

𝑅2𝑡 = 1 −

𝐹𝑝(1−𝛽)2/𝐹𝑝−2(1−𝛽)

𝐹𝑝−2 (4.13)

where

Page 54: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

43

𝑘𝐷 = (𝐾

1+𝐾)(𝐶𝐴𝑂 −

𝐶𝐶𝑂

𝐾)(2𝐹𝑝𝑏𝐷𝑒

𝜌𝑠) (4.14)

𝑘𝐷 is a new constant incorporating system parameters. The conversion versus time

relationship for spheres, long cylinders, and infinite slabs under pore-diffusion control can

be obtained in more familiar forms from Equation 4.13:

for 𝐹𝑝 = 3 (spheres),

𝑘𝐷

𝑅2𝑡 = 1 − 3(1 − β)

2

3 + 2(1 − β) (4.15)

for 𝐹𝑝 = 2 (long cylinders), by applying L’Hospital’s rule,

𝑘𝐷

𝑅2𝑡 = β + (1 − β)ln(1 − β) (4.16)

and for 𝐹𝑝 = 1 (slabs),

𝑘𝐷

𝑅2𝑡 = β2 (4.17)

It is necessary to convert Equations 4.15, 16, and 17 into β as an explicit function

of time and 𝑅 to allow integration with respect to particle size.

Equation 4.15 is thus converted to43

β(𝑡, 𝑅) =

{

1 − {sin [

1

3arcsin (1 −

2𝑘𝐷𝑅2

𝑡)] +1

2}3

𝑎𝑡 𝑡 <𝑅2

𝑘𝐷 (4.18a)

1 𝑎𝑡 𝑡 ≥𝑅2

𝑘𝐷 (4.18b)

The new forms of Equations 4.16 and 4.17 are

Page 55: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

44

β(𝑡, 𝑅) =

{

1 − exp[𝑊−1(

𝑘𝐷𝑡𝑅2

− 1

𝑒) + 1] 𝑎𝑡 𝑡 <

𝑅2

𝑘𝐷 (4.19a)

1 𝑎𝑡 𝑡 ≥𝑅2

𝑘𝐷 (4.19b)

and

β(𝑡, 𝑅) =

{

√𝑘𝐷𝑅2𝑡 𝑎𝑡 𝑡 <

𝑅2

𝑘𝐷 (4.20a)

1 𝑎𝑡 𝑡 ≥𝑅2

𝑘𝐷 (4.20b)

where 𝑊−1 is the 𝑊 ≤ −1 portion of the Lambert W function defined by44 𝑧 =

𝑊(𝑧)𝑒𝑊(𝑧).

Again, Equations 4.18b, 19b, and 20b represent the fact that once a particle is

completely reacted, there is no longer any reaction taking place in it.

4.1.2 Mathematical formulation for size distribution

As mentioned previously, the most commonly used empirical size distributions,

that is, the GGS distribution, the RRB distribution, and the Gamma distribution, were

chosen to investigate the effect of particle size distribution on the fluid-solid reaction. Let

ℎ(𝑅) represent the probability density function in terms of mass of radius of a long

cylinder or a sphere, and the half thickness of an infinite slab.

For the GGS distribution,45

ℎ(𝑅) = 𝑚𝑅𝑚−1

𝑅𝑚𝑎𝑥𝑚 (4.21)

for the RRB distribution,46

Page 56: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

45

ℎ(𝑅) =𝑚

𝑙𝑚𝑑𝑚−1𝑒𝑥𝑝[− (

𝑅

𝑙)𝑚

] (4.22)

and for the Gamma distribution,47

ℎ(𝑅) =𝑏𝑝𝑅𝑝−1𝑒−𝑏𝑅

𝛤(𝑝) (4.23)

where Γ(𝑝) is the gamma function of 𝑝 and {𝑚, 𝑅𝑚𝑎𝑥}, {𝑚, 𝑙}, 𝑎𝑛𝑑 {𝑏, 𝑝} represent

adjustable parameters for the three distributions, respectively. The mutual comparison

among the three size distributions can be achieved by adjusting the parameters that regulate

the mean (μ) and variance (𝜎2). For this study, the CV is used to standardize the effect of

the mean and variance of the different distributions. Figure 4.1 shows a plot of the three

size distributions and the normal distribution for comparison when the CV of each size

distributions is identical at 0.3 and Table 4.1 shows mean, variance, and CV expressed with

the adjustable parameters of the three distributions.

The mass fraction of particles within the range of radius 𝑅 and R+𝑑𝑅 is given

by ℎ(𝑅) ∙ 𝑑𝑅. As stated above, β(𝑡, 𝑅) stands for the fraction reacted of a particle with

initial radius R at a certain time t. Thus, the overall fraction reacted of the entire particle

assemblage at a certain time t, 𝑋(𝑡), is given by

𝑋(𝑡) = ∫ β(𝑡, 𝑅)∞

0ℎ(𝑅)𝑑𝑅 (4.24)

In order to determine the overall fraction reacted, β(𝑡, 𝑅) (depending on controlling

mechanism), ℎ(𝑅) (depending on the size distribution), and the particle shape factor must

be known. For interface-reaction control, the GGS distribution, and the particle shape

factor 𝐹𝑝, substitution of Equations 4.5 and 4.21 into Equation 4.24 gives the following

equation,

Page 57: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

46

Figure 4.1 The three size distributions investigated in this work and the

normal distribution (𝐶𝑉 = 0.3)

Page 58: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

47

Table 4.1 Mean, variance, and CV of three distributions

Name GGS RRB Gamma

Adjustable

parameters 𝑚,𝑅𝑚𝑎𝑥 𝑚, 𝑙 𝑏, 𝑝

Mean (𝛍) 𝑚

𝑚 + 1∙ 𝑅𝑚𝑎𝑥 𝑙 ∙ Γ(

𝑚 + 1

𝑚)

𝑝

𝑏

Variance

(𝝈𝟐) [

𝑚

(𝑚 + 2)−

𝑚2

(𝑚 + 1)2] ∙ 𝑅𝑚𝑎𝑥

2 𝑙2 ∙ [Γ (𝑚 + 2

𝑚) − Γ(

𝑚 + 1

𝑚)]

𝑝

𝑏2

CV (=𝝈

𝝁)

√𝑚

(𝑚 + 2)−

𝑚2

(𝑚 + 1)2

𝑚𝑚 + 1

√Γ(𝑚 + 2𝑚

) − Γ(𝑚 + 1𝑚

)

Γ(𝑚 + 1𝑚 )

1

√𝑝

Page 59: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

48

𝑋(𝑡) = ∫ β(𝑡, 𝑅)𝑅𝑚𝑎𝑥

0[𝑚 ∙

𝑅𝑚−1

𝑅𝑚𝑎𝑥𝑚 ] 𝑑𝑅 (4.25)

where β(𝑡, 𝑅) is given by Equations 4.5a and b. The final solution is given by

𝑋(𝜏) = 𝑚(𝑚

𝑚+1)𝑚 [∫ [1 − (1 −

𝜏𝐶

𝑎)𝐹𝑝] 𝑎𝑚−1

𝑚+1

𝑚𝑎𝛽=1

𝑑𝑎 + ∫ 𝑎𝑚−1𝑎𝛽=10

𝑑𝑎] (4.26)

where

𝑎 =𝑅

𝜇 (4.27)

𝜏𝐶 =𝑘𝐶𝑡

𝜇 (4.28)

𝑋(𝜏) is a function of three parameters, 𝑚,𝐹𝑝, and 𝜏. The adjustable parameter, 𝑚, can

be determined by fixing the CV, for mutual comparison with the effects of the other two

size distributions. For example, the value of 𝑚 is 2.48 or 1.24 when CV is 0.3 or 0.5,

respectively. Following a similar mathematical procedure, the final forms of 𝑋(𝜏) for the

RRB distribution and the Gamma distribution under the interface-reaction control were

derived and are tabulated in Table 4.2 together with Equation 4.26 for the GGS distribution.

Unlike the interface-reaction-controlled case, the pore-diffusion-controlled reaction has

different forms of 𝛽(𝑡, 𝑅) according to the particle shape factors, as shown in Equations

4.18, 4.19, and 4.20. For the case of pore-diffusion control, the GGS distribution, and

spherical particles, substitution of Equations 4.18 and 4.21 into Equation 4.24 and

expressing 𝑋(𝜏) with the normalized parameters in Equations 4.27 and 4.32 give the final

form of 𝑋(𝜏) as follows:

𝑋(𝜏) = 𝑚(𝑚

𝑚+1)𝑚 [∫ [1 − {𝑠𝑖𝑛 [

1

3arcsin(1 −

2𝜏𝐷

𝑎2)] +

1

2}3

] 𝑎𝑚−1𝑚+1

𝑚𝑎𝛽=1

𝑑𝑎 + ∫ 𝑎𝑚−1𝑎𝛽=10

𝑑𝑎] (4.31)

Page 60: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

49

Table 4.2 X(t) of three distributions for interface-reaction-controlled system

Distr. 𝑿(𝒕)

GGS 𝑚(𝑚

𝑚 + 1)𝑚 [∫ [1 − (1 −

𝜏𝐶𝑎)𝐹𝑝] 𝑎𝑚−1

𝑚+1𝑚

𝑎𝛽=1

𝑑𝑎 + ∫ 𝑎𝑚−1𝑎𝛽=1

0

𝑑𝑎]

RRB 𝑚{Γ(𝑚 + 1

𝑚)}𝑚

[ ∫ [1 − (1 −

𝜏𝐶𝑎)𝐹𝑝]

𝑎𝛽=1

𝑎𝑚−1𝑒−{𝑎Γ(

𝑚+1𝑚

)}𝑚

𝑑𝑎

+∫ 𝑎𝑚−1𝑒−{𝑎Γ(

𝑚+1𝑚

)}𝑚𝑎𝛽=1

0

𝑑𝑎]

Gamma 𝑝𝑝

Γ(𝑝)[∫ [1 − (1 −

𝜏𝐶

𝑎)𝐹𝑝]

𝑎𝛽=1

𝑎𝑝−1𝑒−𝑝𝑎𝑑𝑎 + ∫ 𝑎𝑝−1𝑒−𝑝𝑎𝑎𝛽=1

0

𝑑𝑎]

Page 61: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

50

where

𝜏𝐷 =𝑘𝐷𝑡

𝜇2 (4.32)

For the same rate-control and the size distribution but different particle shape factors, it can

be presented as

𝑋(𝜏) = 𝑚(𝑚

𝑚+1)𝑚 [∫ [1 − 𝑒

𝑤−1(𝜏𝐷 𝑎2⁄ −1

𝑒)+1] 𝑎𝑚−1

𝑚+1

𝑚𝑎𝛽=1

𝑑𝑎 + ∫ 𝑎𝑚−1𝑎𝛽=1

0𝑑𝑎] (4.33)

𝑋(𝜏) = 𝑚(𝑚

𝑚+1)𝑚 [∫ √

𝜏𝐷

𝑎2𝑎𝑚−1

𝑚+1

𝑚𝑎𝛽=1

𝑑𝑎 + ∫ 𝑎𝑚−1𝑎𝛽=1

0𝑑𝑎] (4.34)

where Equations 4.33 and 4.34 are for the particle shape of a long cylinder and an infinite

slab which takes the value of 2 and 1, respectively. 𝑋(𝜏) for the other two size

distributions with the three shape factors can be obtained in the same way and are tabulated

in Table 4.3.

4.2 Results and Discussion

The effects of size distributions and shape factors on the kinetics for different CV

values are visually illustrated in the following figures. The first five curves from the top

represent the fraction reacted versus normalized time for spherical particles (𝐹𝑝 = 3) with

CV values of 1, 0.75, 0.5, 0.3, and 0. The next five curves are for flat particles (𝐹𝑝 = 1). It

is noted that the behavior of cylindrical particles (not shown) is in between those of

spherical particles and slabs. It is also noted that the case of flat particles with a distribution

of thickness may be less often encountered in a practical problem than spherical particles.

On the other hand, modeling of spherical particles can be applied to real particles that have

similar sizes in the three dimensions even if they are not exactly spherical.

Page 62: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

51

Tab

le 4

.3 X

(t)

of

thre

e dis

trib

uti

ons

for

pore

-dif

fusi

on-c

ontr

oll

ed s

yst

em

X(t

)

𝑚(𝑚

𝑚+1)𝑚[∫

[1−{𝑠𝑖𝑛[1 3𝑎𝑟𝑐𝑠𝑖𝑛(1−2𝜏 𝐷 𝑎2)]+1 2}3

]𝑎𝑚−1

𝑚+1

𝑚

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑚−1

𝑎𝛽=1

0

𝑑𝑎]

𝑚(𝑚

𝑚+1)𝑚[∫

[1−𝑒𝑤−1(𝜏𝐷𝑎2

⁄−1

𝑒)+1]𝑎𝑚−1

𝑚+1

𝑚

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑚−1

𝑎𝛽=1

0

𝑑𝑎]

𝑚(𝑚

𝑚+1)𝑚[∫

√𝜏 𝐷 𝑎2𝑎𝑚−1

𝑚+1

𝑚

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑚−1

𝑎𝛽=1

0

𝑑𝑎]

𝑚{𝛤(𝑚

+1

𝑚)}𝑚

[∫[1−{𝑠𝑖𝑛[1 3𝑎𝑟𝑐𝑠𝑖𝑛(1−2𝜏 𝐷 𝑎2)]+1 2}3

]𝑎𝑚−1𝑒−{𝑎𝛤(𝑚

+1

𝑚)}𝑚

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑚−1𝑒−{𝑎𝛤(𝑚

+1

𝑚)}𝑚

𝑎𝛽=1

0

𝑑𝑎]

𝑚{𝛤(𝑚

+1

𝑚)}𝑚

[∫[1−𝑒𝑤−1(𝜏𝐷𝑎2

⁄−1

𝑒)+1]𝑎𝑚−1𝑒−{𝑎𝛤(𝑚

+1

𝑚)}𝑚

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑚−1𝑒−{𝑎𝛤(𝑚+1

𝑚)}𝑚

𝑎𝛽=1

0

𝑑𝑎]

𝑚{𝛤(𝑚

+1

𝑚)}𝑚

[∫√𝜏 𝐷 𝑎2𝑎𝑚−1𝑒−{𝑎𝛤(𝑚

+1

𝑚)}𝑚

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑚−1𝑒−{𝑎𝛤(𝑚

+1

𝑚)}𝑚

𝑎𝛽=1

0

𝑑𝑎]

𝑝𝑝

𝛤(𝑝)[∫

[1−{𝑠𝑖𝑛[1 3𝑎𝑟𝑐𝑠𝑖𝑛(1−2𝜏 𝐷 𝑎2)]+1 2}3

]𝑎𝑝−1𝑒−𝑝𝑎

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑝−1𝑒−𝑝𝑎

𝑎𝛽=1

0

𝑑𝑎]

𝑝𝑝

𝛤(𝑝)[∫

[1−𝑒𝑤−1(𝜏𝐷𝑎2

⁄−1

𝑒)+1]𝑎𝑝−1𝑒−𝑝𝑎

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑝−1𝑒−𝑝𝑎

𝑎𝛽=1

0

𝑑𝑎]

𝑝𝑝

𝛤(𝑝)[∫

√𝜏 𝐷 𝑎2𝑎𝑝−1𝑒−𝑝𝑎

𝑎𝛽=1

𝑑𝑎+∫

𝑎𝑝−1𝑒−𝑝𝑎

𝑎𝛽=1

0

𝑑𝑎]

Fp

3

2

1

3

2

1

3

2

1

Dis

tr.

GG

S

RR

B

Gam

ma

Page 63: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

52

4.2.1 Gate-Gaudin-Schuhmann distribution

The effect of the GGS distribution for interface-reaction-controlled and pore-

diffusion-controlled cases is shown in Figures 4.2. The fraction reacted for nonuniform

particles is higher than for uniform particles at the beginning of the reaction. After some

time, the fraction reacted for nonuniform particles is lower than for uniform particles. This

is due to the fact that reaction of small particles is completed rapidly at the beginning of

the reaction and large particles take longer to complete the reaction. The maximum

differences of conversion between uniform and nonuniform particles for different

distributions, particle shapes, CVs, and rate-controlling steps are tabulated in Table 4.4.

The difference of fractional conversion greater than 0.15 is considered as a significant

difference in this work. When the CV values are less than 0.5 or are greater than 0.75, the

differences of fractional conversion are much less than or are much greater than 0.15.

Therefore, only the differences at CV of 0.5 and 0.75 are shown in Table 4.4. For the

particle shape factor of 3, which is most likely to be encountered in practical systems, and

the overall rate control by interface reaction, the effect of particle size distribution is small

until CV increases to 0.5 where the maximum difference of conversion rates is 0.086 at a

normalized time of 0.1. Once CV increases to 0.75, the maximum difference is 0.171 which

is a significant difference. In case of cylindrical particles, the difference is somewhat higher

than those for spherical particles. The maximum difference is 0.091 at a CV value of 0.5.

When CV is 0.75, differences tend to exceed 0.15. Also, flat particles show the maximum

difference of 0.148 at the CV value of 0.5 and a normalized time of 1. Like spherical and

cylindrical cases, the difference is higher than 0.15 at the CV value of 0.75.

For pore-diffusion control, significant differences appear as CV increases to 0.75,

Page 64: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

53

Figure 4.2 Fraction reacted versus normalized time for the GGS distribution (a) Interface

reaction control and (b) Pore-diffusion control

Page 65: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

54

Table 4.4 Maximum differences of conversion rate between uniform and nonuniform

particles with different distributions, particle shapes, CV, and rate controlling steps

𝑭𝒑 CV

GGS RRB Gamma

Interface

reaction

Pore

diffusion

Interface

reaction

Pore

diffusion

Interface

reaction

Pore

diffusion

3 0.5 0.086 0.093 0.071 0.080 0.065 0.075

0.75 0.171 0.180 0.135 0.146 0.128 0.141

2 0.5 0.091 0.098 0.078 0.085 0.072 0.081

0.75 0.178 0.185 0.143 0.152 0.137 0.147

1 0.5 0.148 0.141 0.130 0.130 0.123 0.116

0.75 0.198 0.198 0.169 0.169 0.165 0.165

Page 66: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

55

which is the same trend as in the case of interface-reaction control. It is observed that the

difference of fractional conversion between uniform and nonuniform particles increases

with increasing CV value and decreasing shape factor. In brief, the differences caused by

the dispersion of particle size for the case of the GGS distribution are not significant until

CV increases to 0.5 irrespective of the type of rate-controlling step. When CV is higher

than 0.5, it is recommended to consider the effect of particle size distribution on the overall

kinetics.

4.2.2 Rosin-Rammler-Bennett distribution

Figure 4.3 represent the RRB distribution effects for interface-reaction-controlled

and pore-diffusion-controlled cases. The fraction reacted versus a normalized time for the

RRB distribution shows the same trend as in the case of the GGS distribution. However,

the differences of fractional conversion between uniform and nonuniform particles for the

RRB distribution is smaller than those for the GGS distribution with the same values of

CV and particle shape factor. For the case of interface-rate control in Figure 4.3a, the

maximum differences of 0.135, 0.143, and 0.169are observed for the CV value of 0.75

when the shape of particles are spheres, cylinders, and slabs, respectively. For pore-

diffusion control in Figure 4.3b, the maximum differences are 0.146, 0.152, and 0.169 for

the CV value of 0.75 for the three particle shapes. Thus, it can be said that the effect of the

RRB distribution on the kinetics for spherical and cylindrical particles under interface-

reaction control can be considered as insignificant until the CV is increased to 0.75. For

flat particles under interface-reaction control, the distribution effects are tolerable up to the

CV value of 0.5. In case of pore-diffusion control and spherical particles, the effect of

dispersion of particle size can be considered as tolerable up to the CV value of 0.75.

Page 67: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

56

(a) Figure 4.3 Fraction reacted versus normalized time for the RRB distribution (a)

Interface reaction control and (b) Pore-diffusion control

Page 68: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

57

However, for cylindrical and flat particles under pore-diffusion control, the effect of size

distribution is somewhat greater.

4.2.3 Gamma distribution

The effect of the Gamma distribution for interface-reaction-controlled and pore-

diffusion-controlled systems is illustrated in Figure 4.4. For interface-rate control in Figure

4.4a, 0.128, 0.137, and 0.165 are the maximum differences observed for the CV value of

0.75 when the shape of particles are spheres, cylinders, and slabs, respectively. For pore-

diffusion control and different particle shapes, the maximum differences are 0.141, 0.147,

and 0.165 at the CV value of 0.75. It shows very similar results to those of the RRB

distribution. The gamma distribution effects on the kinetics for spherical and cylindrical

particles under interface-reaction control are minor up to the CV value of 0.75. On the other

hand, in case of flat particles under interface-reaction control, the distribution effects are

acceptable only up to the CV value of 0.5. For spherical particles and cylindrical particles

under pore-diffusion control, the effect of particle size distribution is tolerable up to the CV

value of 0.75. However, for flat particles under pore-diffusion control, the effects are

insignificant only up to the CV value of 0.5. In summary for the Gamma distribution effects,

the effects are small until the CV increases to 0.75 for spherical and cylindrical particles

irrespective of the type of rate-controlling mechanism. When the CV value is higher than

0.75, it is recommended to consider the effect of particle size distribution on the overall

kinetics. For flat particles under both rate-controlling steps, the size variation effects on the

kinetics for a CV value higher than 0.5 should be considered significant.

Overall, the difference in conversion values between uniform and nonuniform

particles under pore-diffusion control is slightly higher than those under interface-reaction

Page 69: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

58

Figure 4.4 Fraction reacted versus normalized time for the Gamma distribution (a)

Interface reaction control and (b) Pore-diffusion control

Page 70: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

59

control. Thus, it can be surmised that when a system is rate-controlled simultaneously by

both chemical kinetics and pore diffusion along with the variation of particle size, the size

distribution effects in this system will be in between those for interface-reaction-controlled

case and pore-diffusion-controlled cases.

Another case frequently encountered in industry is that of porous solids. If the

overall reaction of a porous solid is slow and thus the fluid concentration is relatively

uniform throughout the porous solid, the fraction reacted will be the same regardless of the

pellet size. When pore diffusion controls the overall rate of a porous solid, the effects of

size will be the same as those for nonporous solids.12, 48-50

4.2.4 Comparison of the effect of three size distributions on the kinetics

The effects on reaction kinetics for the three types of size distributions are similar

as can be seen in Figures 4.2a, 4.3a, and 4.4a for interface-reaction control and Figures

4.2b, 4.3b, and 4.4b for pore-diffusion control although their probability density function

for size distribution presented in Figure 4.1 is quite different, especially for the GGS

distribution. If the CV value and the particle shape factor are the same, the mathematical

evaluation shows almost the same effect irrespective of the type of particle size distribution.

It is somewhat surprising especially for the GGS distribution, considering that its

probability density function is monotonous with particle size unlike for the other two size

distributions. Figure 4.5 describe the similarity of the reaction kinetics for the three size

distributions for spherical particles under interface-reaction control and pore-diffusion

control when the value of the CV is 0.3 and 1.5. For a CV value of 0.3 with both interface

reaction and pore-diffusion control, the difference among the three different curves is very

small. For a CV value of 1.5, the kinetics for the RRB and the Gamma distribution seems

Page 71: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

60

Figure 4.5 Comparison for the three size distributions; 𝐹𝑝 = 3 and 𝐶𝑉 = 0.3 and 1.5

(a) Interface reaction control and (b) Pore-diffusion control

Page 72: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

61

identical. The kinetics for the GGS distribution deviates rather slightly from the two other

curves. It is noted that the CV value of 1.5 is rarely encountered in practical systems. It has

thus been determined that the effect of particle size distribution on the reaction kinetics is

similar regardless of the type of distribution function as long as the CV values of the

distribution functions are the same. Thus, even when the effect of particle size distribution

is significant, it can be represented just by the CV value.

4.2.5 R2 values for the effect of particle size distributions

Another way to quantify the effects of particle size distribution on the kinetics is

to calculate R2 values which stand for the degree of the deviation of the overall conversion

rate from the case where the particle size is uniform. If R2 value is one, the conversion rate

of entire particle assemblages is the same as the case where the particle size is uniform.

When R2 value approaches to zero, the conversion rate shows the different trend from the

conversion rate of uniform particle size.

The R2 values in each case are tabulated in Table 4.5. As can be seen in the table,

R2 values are greater than 0.9 in every case where CV value is less than or equal to 0.5. If

we arbitrarily determine that R2 of 0.9 is the lowest value tolerable to be considered for the

particle size as uniform, the assumption that particle size is uniform can be valid in the

determination of fluid-solid kinetics in the case where CV is less than 0.5. On the other

hand, in the case where CV value is larger than 0.5, R2 values can be less than 0.9. Thus, it

can be said that the effects of particle size distribution should be considered in the fluid-

solid kinetics when CV is larger than 0.5.

Page 73: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

62

Tab

le 4

.5 R

2 v

alues

of

each

cas

e

CV

1.5

0.6

36

3

0.5

32

7

0.1

30

4

0.5

32

5

0.4

71

2

0.3

10

8

0.8

08

7

0.7

40

6

0.4

75

3

0.7

01

8

0.6

52

2

0.5

22

9

0.7

75

0

0.6

99

7

0.4

07

5

0.6

66

3

0.6

14

8

0.4

80

5

1

0.8

52

5

0.8

02

6

0.6

19

9

0.7

54

6

0.7

27

2

0.6

50

8

0.9

11

1

0.8

75

8

0.7

40

4

0.8

36

9

0.8

11

0

0.7

39

2

0.9

11

1

0.8

75

9

0.7

40

4

0.8

36

9

0.8

11

0

0.7

39

2

0.7

5

0.9

32

3

0.9

04

5

0.8

07

5

0.8

60

5

0.8

42

8

0.7

96

4

0.9

54

6

0.9

34

4

0.8

57

9

0.9

02

6

0.8

86

3

0.8

41

2

0.9

57

9

0.9

38

8

0.8

65

0

0.9

08

5

0.8

92

1

0.8

46

5

0.5

0.9

80

8

0.9

70

2

0.9

32

0

0.9

46

4

0.9

35

8

0.9

10

6

0.9

85

4

0.9

77

3

0.9

45

8

0.9

58

8

0.9

50

2

0.9

27

2

0.9

87

2

0.9

79

9

0.9

50

2

0.9

63

0

0.9

54

7

0.9

31

5

0.4

0.9

91

0

0.9

85

3

0.9

62

7

0.9

70

5

0.9

62

9

0.9

45

1

0.9

92

8

0.9

88

2

0.9

69

4

0.9

76

1

0.9

70

1

0.9

54

5

0.9

93

7

0.9

89

7

0.9

72

1

0.9

78

7

0.9

73

1

0.9

57

6

0.3

0.9

96

8

0.9

94

4

0.9

83

1

0.9

87

1

0.9

82

7

0.9

71

8

0.9

97

3

0.9

95

2

0.9

85

7

0.9

88

9

0.9

85

4

0.9

76

2

0.9

97

6

0.9

95

8

0.9

87

1

0.9

90

2

0.9

87

0

0.9

78

1

0.2

0.9

99

3

0.9

98

7

0.9

94

6

0.9

96

3

0.9

94

5

0.9

89

7

0.9

99

4

0.9

98

8

0.9

95

3

0.9

96

7

0.9

95

2

0.9

91

1

0.9

99

5

0.9

98

9

0.9

95

8

0.9

97

0

0.9

95

7

0.9

91

9

0.1

1.0

00

0

0.9

99

9

0.9

99

3

0.9

99

6

0.9

99

3

0.9

98

4

1.0

00

0

0.9

99

9

0.9

99

3

0.9

99

7

0.9

99

4

0.9

98

5

1.0

00

0

0.9

99

9

0.9

99

4

0.9

99

7

0.9

99

5

0.9

98

7

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

1.0

00

0

Fp

3

2

1

3

2

1

3

2

1

3

2

1

3

2

1

3

2

1

Rea

ctio

n

Ch

emic

al

Dif

fusi

on

Ch

emic

al

Dif

fusi

on

Ch

emic

al

Dif

fusi

on

Dis

tr.

GG

S

RR

B

Gam

ma

Page 74: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

63

4.2.6 Comments on particles with nonbasic shapes

Although many real particles have shapes that may be approximated by the three

basic shapes used in the above analysis, other particles have shapes that are intermediate

of the three basic shapes or more irregular. Examples are finite cylinders, parallelepipeds,

ellipsoids, completely irregular particles, or a mixture thereof. Assemblages made up of

particles with some of these shapes can be treated based on first principles, but even in such

cases the computation will be lengthy. Furthermore, with variations in particle shape, size

distribution, and reactivity even for the same solid as well as the errors typically present in

rate measurements, it is not worthwhile to spend such complicated computational efforts.

Here, the following suggestions for systems consisting of particles of nonbasic shapes can

be offered. Express the particle size in terms of 𝐴𝑝 and 𝑉𝑝 as defined by Equation 4.7.

This requires an estimation of the value of 𝐹𝑝, which depends on the ratios of the three

major dimensions of the particle and can be a noninteger between 1 and 2 or between 2 and

3. The exact appropriate value was found to be different by 0.1-0.3 depending on the rate-

controlling step.51 Further, differences of this extent in the value of 𝐹𝑝 have only modest

effects on the conversion versus time relationships regardless of the rate-controlling

mechanism. The use of such estimated 𝐹𝑝 values is made greatly more convenient by the

expression of conversion versus time relationships with 𝐹𝑝 as the parameter for different

particle shapes, especially for diffusion-controlled reactions. Thus, Equation 4.13 can be

used for any intermediate shapes, although this will involve evaluating the values of

β(𝑡, 𝑅) by an implicit method.

Page 75: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 5

KINETIC STUDY OF REDUCED UGS WITH EFFECT OF

PARTICLE SIZE DISTRIBUTION

5.1 Particle size distribution of reduced UGS

To study the effect of particle size of reduced UGS washed by acetic acid on the

reaction kinetics, the particle size distribution was analyzed as shown in Figure 5.1 by a

laser diffraction particle size analyzer. As can be seen in Figure 5.1, the particle size is not

uniform, but follows a particle size distribution. Again, the types of particle size distribution

do not affect the kinetics significantly according to Figure 4.5. Especially when CV value

is 0.3, the effect of types of particle size distribution is negligible. In this regard, the mean

and standard deviation of the distribution were found to be 0.096 and 0.032 μm ,

respectively, which gives the CV value (defined as standard deviation divided by mean) of

0.33. According to Table 4.5, R2 values for the CV value of 0.3 are larger than 0.97 in every

case. Especially, when we consider that the particle shape of reduced UGS washed by acetic

acid seems to be spherical based on the SEM image in Figure 3.2, the smallest R2 value for

shape factor of 3 is 0.9871. Thus, it will be assumed that the effect of particle size

distribution of reduced UGS washed by acetic acid is negligible in the calculation of

kinetics. In other words, it will be assumed that the particle size is uniform in the calculation

of the reaction kinetics of reduced UGS washed by acetic acid.

Page 76: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

65

Figure 5.1 Particle size of reduced UGS after acetic acid washing

Page 77: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

66

5.2 Determination of rate-controlling process of the reduced UGS

5.2.1 Interfacial reaction

The rate-controlling process and kinetic parameters were investigated on the basis

of the discussion in Chapter 5.1. If the overall rate of reaction is controlled by the interfacial

reaction, Equation 4.5 can be expressed for a normalized time parameter, t*, with respect

to conversion rate, X as follows13, 52 which is only valid under the assumption that the

particle size is uniform:

𝑡∗ = 1 − (1 − 𝑋)1/3 (5.1)

Experimental data obtained at three different temperatures, 70, 80, and 90oC were

applied to Equation 5.1 to evaluate interfacial reaction controlled kinetics as illustrated by

plotting 1 − (1 − 𝑋)1/3 versus time as shown in Figure 5.2. The value of 1 − (1 − 𝑋)1/3

is linearly related to leaching time, which suggests that the overall rate of reaction can be

controlled by the interfacial reaction in the leaching process.

The activation energy can be used to evaluate the rate-controlling step. The

Arrhenius equation, when rearranged for determining the activation energy is described in

Equation 5.2.52

𝑙𝑛𝑘 = 𝑙𝑛𝐴 −𝐸𝑎

𝑅𝑇 (5.2)

where k is the rate constant of a chemical reaction, A the preexponential factor, Ea the

activation energy, R the universal gas constant, and T the temperature in degrees Kelvin.

Figure 5.3 shows the relationship between lnk and 103/T in which the activation energy can

be determined from the slope of the line. The activation energy was determined to be 73.9

kJ/mole, suggesting interfacial reaction controlled kinetics.

Page 78: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

67

Figure 5.2 Relationship between the values of 𝟏 − (𝟏 − 𝐗)𝟏/𝟑 and time at three

temperatures

Page 79: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

68

Figure 5.3 Arrhenius Plot

Page 80: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

69

5.2.2. Solid state diffusion

It is worthwhile to note that the iron removal in the reduced UGS is not increased

although the leaching temperature was increased from 140oC to 190oC. If the overall

reaction was controlled by interface reaction, iron removal would be increased with

increase of temperature.

In this regard, the other possibility of rate-controlling process of the reduced UGS

is solid state diffusion of iron in the titanium hydride, which also has high value of the

activation energy. The average particle size of the reduced UGS is 0.1 μm, which is fairly

small for the reduced UGS to retain enough pore size. Thus, overall rate of reaction might

be controlled by the diffusion of iron in the titanium hydride. If leaching temperature is

high enough for iron to be stably dissolved in titanium hydride, increasing the leaching

temperature might oppress iron removal. In fact, it was observed that iron removal in 0.05

M hydrochloric acid rather decreased from 79.59% to 79.19% as temperature increased

from 140oC to 190oC.

Overall, several experimental evidences indicate that the rate-controlling process

of iron removal from the reduced UGS might be controlled by solid state diffusion. Figure

5.4 shows the probable micro-scale structure of the reduced UGS. This figure indicates that

the effectiveness of iron removal may be limited by locking up of iron as interstitial

substitutions.

5.2.2.1 Mathematical formulation

As can be seen in Figure 3.8, the overall particle shape is assumed to be sphere.

Solid state diffusion equation in a spherical particle according to Fick’s second law can be

given as,

Page 81: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

70

Figure 5.4 Possible schematic diagram of the reduced UGS (Brown: TiH2, Black: FexOy, or

Fe, blue: other impurities, gray: hydrogen ion). In this scenario, the reaction is controlled

by solid state diffusion of iron.

Page 82: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

71

𝜕𝐶

𝜕𝑡= 𝐷(

𝜕2𝐶

𝜕𝑟2+2

𝑟

𝜕𝐶

𝜕𝑟) (5.3)

where C is concentration and D is diffusivity. By setting u = Cr in Equation 5.3, it can be

rewritten as,

𝜕𝐶

𝜕𝑡= 𝐷

𝜕2𝑢

𝜕𝑟2 (5.4)

Under the condition that concentration changes with variation of time, the solution of

Equation 5.4 can be expressed as,53

𝐶−𝐶1

𝐶0−𝐶1= 1 +

2𝑎

𝜋𝑟∑

(−1)𝑛

𝑛𝑠𝑖𝑛

𝑛𝜋𝑟

𝑎exp(−

𝐷𝑛2𝜋2𝑡

𝑎2)∞

𝑛=1 (5.5)

where the boundary conditions and initial condition are

𝑎𝑡 𝑟 = 0, 𝑢 = 0, 𝑡 > 0 (5.6)

𝑎𝑡 𝑟 = 𝑎, 𝑢 = 𝑎𝐶𝑜 , 𝑡 > 0 (5.7)

𝑎𝑡 𝑡 = 0, 𝑢 = 𝑟𝐶1 (5.8)

where a is radius of a particle, Co is surface concentration, and C1 is initial uniform

concentration. When surface concentration of iron in the reduced UGS, Co, is 0 and the

initial uniform concentration if iron in the reduced UGS, C1, is 1, the total amount of

diffusing iron which leaves the titanium hydride can be given as,54

𝑴𝒕

𝑴∞= 𝟏 −

𝟔

𝝅∑

𝟏

𝒏𝟐𝐞𝐱𝐩(−

𝑫𝒏𝟐𝝅𝟐𝒕

𝒂𝟐)∞

𝒏=𝟏 (5.9)

The diffusion coefficient of iron in titanium hydride is not found in reality, so it is

difficult to judge whether iron removal from the reduced UGS follows solid state diffusion.

However, predicted and experimental iron removal curves were plotted in Figure 5.5 under

Page 83: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

72

Figure 5.5 Comparison of predicted values produced from the equation for solid state

diffusion and experimental value; 0.1 M HCl, 1,000 rpm, 90oC, no boric acid

Page 84: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

73

the assumption that the value of diffusion coefficient of iron in titanium hydride is 3E-16

cm2/s at 90oC, which gives the best fit of values using the Equation 5.9 to fit the

experimental data obtained under the condition in Table 3.3.

It should be noted that titanium loss was drastically increased at 8 h as shown in

Table 5.1. Decreasing the trend of titanium loss after 8 h can be due to the precipitation of

titanium ion as titanium dioxide. Dissolution of titanium hydride might allow hydrochloric

acid to contact with iron easily due to the disappearance of the titanium hydride shell, which

was supposed to prevent hydrochloric acid from reacting with iron. Therefore, predicted

and experimental iron removal curves until 4 h was plotted again in Figure 5.6 to mitigate

the effect of titanium dissolution by hydrochloric acid when the diffusion coefficient of

iron in titanium hydride is 1E-16 cm2/s, which gives the best fit. Figure 5.6 shows fairly

good agreement of experimental data with theoretical data. Thus, it can be said that solid

state diffusion of iron in titanium hydride can be the rate-controlling process which governs

the overall rate of reaction.

However, due to the fact that the additional reaction such as the dissolution of

titanium hydride in hydrochloric acid can affect the iron removal rate after 4 h of leaching

time, it is difficult to judge the rate-controlling process of iron removal from the reduced

UGS.

5.2.3 Effect of external mass transport

The influence of external mass transport on the reaction rate of leaching with 0.1

M hydrochloric acid at 90oC was investigated by varying the stirring speed in the range of

100 – 1,000 rpm. The data in Figure 5.7 show that the reaction rate did not change

significantly with increasing stirring speed, which implies that the overall rate of reaction

Page 85: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

74

Table 5.1 Titanium loss with leaching time. Leaching parameters: .1 M HCl, 1000 rpm,

90oC, no boric acid

Leaching time (hr) Titanium loss (%)

1 0.44

2 0.29

4 0.35

8 2.04

12 3.46

16 4.30

20 2.10

24 1.63

Page 86: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

75

Figure 5.6 Comparison of predicted values produced from the equation for solid state

diffusion and experimental value until 4 h of leaching time

Page 87: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

76

Figure 5.7 Iron removal versus time at three rotation speed

Page 88: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

77

is not controlled by mass transport through an outer boundary layer.

To sum up, the overall rate of reaction seems to follow interfacial reaction

controlled kinetics. The rate-controlling process of reaction was determined under the

assumption that the diameter of every single particle is identical. This assumption believed

to be valid due to the fact that the CV value of the particle size distribution is 0.33, which

gives R2 value of more than 0.97. When we consider that the reaction takes place based on

hydrometallugical processing, which is known as a slower process than pyrometallurgical

processing, the fact that the reaction is controlled by interfacial reaction seems to be

reasonable.

Page 89: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 6

SELECTIVE RECOVERY OF TITANIUM FROM LEACHED

ILMENITE USING TANNIC ACID

6.1 Materials and methods

6.1.1 Preparation of titanium solution from ilmenite

Sixteen molar sulfuric acid was chosen as a lixiviant to prepare a titanium

containing multielement solution from ilmenite. Ten grams of ilmenite was added into 50

ml of sulfuric acid in a 500 ml round-bottom boiling flask at the room temperature. Twenty

grams of ammonium sulfate ((NH4)2SO4) was also added to elevate the boiling point of the

sulfuric acid solution during leaching. A 200 ml Graham condenser which has an inner coil

for additional surface area was applied on the neck of the flask to maintain the volume. The

solution was refluxed and the leaching was performed for 8 h. A magnetic stirring bar at

1,000 rpm was used for homogenizing the solution. After the digestion process, the solution

is allowed to settle. A dark brownish multielement sulfate solution was obtained as

supernatant liquid was collected for chelation.

6.1.2 Synthesis of titanium-tannic acid complex

Fifty grams of tannic acid in 200 ml water was placed in a 500 ml three-neck

flask and constantly stirred by a magnetic stirring bar at a rotational speed of 500 rpm. A

200 ml separatory funnel fitted with a stopcock, containing titanium bearing solution, was

installed on one of necks of the flask. A pH probe was also inserted through another neck

Page 90: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

79

and the third neck was sealed. Ammonium hydroxide was added from the separatory funnel

so that the pH was maintained at 4.5 during the precipitation of the dark orange-colored

precipitate. The experiment was carried out at room temperature. The precipitate was

collected by vacuum filtration, washed with 2% hydrochloric acid solution, followed by

cold deionized water and isopropyl alcohol, dried at room temperature, and ground into

powder. The filtrate was investigated by ICP-OES. The solid samples were characterized

using SEM, energy-dispersive X-ray spectroscopy (EDS), and Fourier transform infrared

spectroscopy (FTIR).

6.1.3 Methodology of DFT simulation

DFT simulation using general ab initio quantum chemistry package, Gaussian 09,

was carried out to determine optimized geometry of titanium-tannic acid complex and the

stability based on highest occupied molecular orbital (HOMO) and lowest unoccupied

molecular orbital (LUMO). B3LYP55, 56 for the exchange-correlation energy functional

calculation was employed with a basis of 6-31G, which has been known as an effective

method for the simulation of organic molecules.57 Also, the optimization was conducted

without symmetry constraints.

6.2 Results and discussion

6.2.1 Experiment

The multielement solution, which was prepared by dissolving ilmenite in

concentrated sulfuric acid, was composed of 22.64% (w/w) of titanium, 71.28% (w/w) of

iron, and 6.08% (w/w) of other impurities determined by ICP-OES. After precipitating out

titanium-tannic acid complex with tannic acid using ammonium hydroxide to adjust the pH

Page 91: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

80

to 4.5, the yield of titanium was found to be 95.58% (w/w). Figure 6.1 shows the surface

morphology and the chemical composition of titanium-tannic acid. The particle shape looks

irregular and it forms an agglomerate. The sizes of individual particle are only few micro

meters. The powder has 98.6% (w/w) of titanium, 0.2% (w/w) of iron, and 1.2% (w/w) of

other impurities. When it is considered that iron is difficult to remove in titanium

production, Figure 6.1b shows an important capability of tannic acid to extract titanium

selectively from the solution containing impurities such as iron, magnesium, and silica.

Figure 6.2 shows FTIR spectra of (a) titanium and (b) titanium-tannic acid

complex. The IR spectra present peaks at a broad range of 3500~3200 cm-1 due to the O-

H stretching vibration in both tannic acid and titanium-tannic acid complex.58 However,

the positive shift and decreased intensity of the peak for O-H bond in the titanium tannic

acid complex were observed due primarily to coordination of phenolic –OH group with

titanium ion by means of deprotonation.59 This can suggest that tannic acid tends to form

complexes with titanium using hydroxyl groups attached to benzene rings. Several peaks

below 1800 cm-1 are due to tannic acid including benzene rings and carboxylic acids. C=O

stretching vibration at 1685 and 1706 cm-1 for tannic acid and titanium-tannic acid complex

was observed.55 The peaks 1500~1400 cm-1 account for C-C stretching vibration in

aromatic rings. Significant changes in the spectrum of titanium-tannic acid complex at a

range of 1700~1450 cm-1 may be due to the chelation process for the formation of metal

complex. The peaks due to C-O stretching vibration and O-H bending in benzene rings and

carboxylic acids appear at 1300~1000 cm-1 in both Figures 6.2a and b.55 The peaks below

900 cm-1 can be explained by C-H bonding in benzene rings. It is worthwhile to be noted

that peaks below 1452 cm-1 in tannic acid match with peaks in titanium-tannic acid complex,

Page 92: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

81

Figure 6.1 (a) SEM image of titanium-tannic acid complex, (b) the composition

of titanium-tannic acid complex analyzed by EDS

Page 93: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

82

Figure 6.2 FT-IR spectra of (a) tannic acid (b) titanium-tannic acid complex

Page 94: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

83

which may mean that there is no significant participation of carboxylic groups in

coordination with titanium ion.

6.2.2 Simulation

Based on the fact that titanium ion forms an ionic bonding with two –OH groups

in gallic acid,37 Figure 6.3 shows a probable schematic diagram of titanium-tannic acid

complex and optimized geometry determined by DFT simulation. The geometry was fully

stabilized.

A LUMO-HOMO gap is known as the important parameter that can describe the

degree of stability of the complexes.60 A molecule having a larger LUMO-HOMO gap is

more stable than a molecule having a smaller LUMO-HOMO gap. Also, according to a

concept of hard and soft Lewis acids and bases (HSAB), the global hardness of a molecule

is the representative stability index of its complex.61 A high value of the global harness

implies a highly stable complex. In the DFT simulation, the LUMO-HOMO gap was

calculated to be 0.08934 a.u (2.43 eV) as shown in Figure 6.4 and the global hardness (h)

was found to be 0.04467 a.u based on Equation 6.1.

h = (𝐸𝐿𝑈𝑀𝑂 − 𝐸𝐻𝑂𝑀𝑂)/2 (6.1)

where h is the global hardness, 𝐸𝐿𝑈𝑀𝑂 the energy of the lowest unoccupied molecular

orbitals, and 𝐸𝐻𝑂𝑀𝑂 the energy of the highest occupied molecular orbitals. The LUMO-

HOMO gap and the global hardness of titanium-tannic acid complex are large enough to

indicate the energetic feasibility and stability for the formation of titanium-tannic acid

complex.

Also, it is interesting to note that electron density mainly consists of carbon and

oxygen from a π-conjucated gallic acid unit located at the end of the molecule in both

Page 95: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

84

Figure 6.3 (a) The schematic diagram of titanium and tannic acid and (b) optimized

structure by DFT simulation at B3LYP with 6-31G: gray (carbon), red (oxygen), white and

big atom (titanium), white and small atom (hydrogen).

Page 96: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

85

Figure 6.4 The molecular orbitals’ energy levels of titanium-tannic acid complex at

LUMO+1, LUMO, HOMO, and HOMO-1: green (positive isosurface of molecular

orbitals at the energy level of each molecular orbital), red (negative isosurface of

molecular orbitals at the energy level of each molecular orbital).

Page 97: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

86

HOMO and HOMO-1 levels in which there is no contribution from the titanium atom.

On the other hand, in the LUMO level, molecular orbitals are uniformly distributed

around the titanium atom and the adjacent gallic acid group, which has empty π-orbitals

of benzene rings. It is also observed in the LUMO+1 level that the distribution of molecular

orbitals is somewhat similar to that in the LUMO level, but there is a more significant

localization in the titanium atom and less contribution from digallic acid group. This can

be explained by the relative energetic alignment between molecular orbitals of titanium

and gallic acid groups. Titanium atoms in LUMO are lower in energy than digallic acid

groups. Thus, titanium in titanium-tannic acid complex has a possibility of forming

titanium hydride in a hydrogen atmosphere as the titanium atom can easily accept excited

electrons.

Page 98: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

CHAPTER 7

CONCLUSIONS

In summary, this thesis work mainly evaluates methods of obtaining pure titanium

hydride from different sources and methods. The first route was to use DRTS in which

metallic titanium can be obtained by dehydrogenation of titanium hydride after impurities

are removed. The leaching characteristics of iron removal from the reduced upgraded

titanium slag was studied with mild hydrochloric and boric acids under ambient pressure

and elevated pressure. Under the constraint that 1% (w/w) of titanium hydride loss is the

maximum amount tolerable, 0.1 M turned out to be the most effective concentration of

hydrochloric acid for iron removal from the reduced UGS. Thus, 0.05 M and 0.1 M

hydrochloric acid were chosen as leaching solutions with or without 1 M boric acid. 90oC,

140oC, and 190oC were studied for optimization of leaching temperature.

Effect of boric acid on iron removal from reduced UGS was strongly dependent

on acidity and temperature of the leaching environment. Solutions of 1 M boric acid with

0.05 M hydrochloric acid at 90oC showed significant improvement of iron removal, but

with 0.1 M hydrochloric acid, a reduced effect was observed. At higher temperature (140

and 190oC), boric acid did not affect iron removal, regardless of the concentration of

hydrochloric acid.

The extent of iron removal increased significantly with the increase of temperature.

Once temperature reaches about 140oC, 87.63% of iron removal was achieved with 0.1 M

Page 99: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

88

hydrochloric acid. Additional iron removal was not observed at higher temperature (190oC).

To sum up, the best leaching conditions from those studied were found to be 0.1 M

hydrochloric acid, no boric acid, and 140oC.

A factorial design of experiment (FDE) for equation modeling was carried out to

study the influence of three factors (temperature and concentrations of hydrochloric, and

boric acids) on iron removal. The predicted iron removal obtained by the FDE model

showed good agreement with actual iron removal within the range of condition evaluated.

To verify the overall reaction kinetics, shrinking unreacted core model was

introduced with an assumption that every single particle has same diameter. Thus,

additional study was carried out to justify the assumption, which utilized the evaluation of

the effects of three empirical size distributions, the Gate-Gaudin-Schuhmann distribution,

the Rosin-Rammler-Bennett distribution, and the Gamma distribution, on the fluid-solid

reaction kinetics. The effects were analyzed for the two extreme rate-controlling steps,

either interface reaction or pore diffusion, for the shrinking unreacted-core model and three

basic particle shapes. The expressions for overall conversion rate of entire particle

assemblages were derived mathematically, and calculated by a technical computing

language, “MATLAB”.

According to the calculation, as the degree of spread of particle size distribution

increases, the reaction rates deviate progressively from the case of uniform particles. Also,

the effect of particle size distribution is somewhat greater for pore-diffusion-controlled

reactions than for interface-reaction-controlled cases, although the difference is rather

small. The case of mixed control is expected to be in between the two controlling steps.

In the case of spherical particles, the effect of the GGS distribution on the kinetic

Page 100: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

89

is small for both rate controlling steps until the CV increases to 0.5. For the RRB and the

Gamma distributions, the effect remains small up to a CV value of 0.75. The effects on

cylindrical and flat particles are similar, if slightly greater in that order. We can also

evaluate the effect of particle size distribution on the kinetics using R2 values. If we

arbitrarily determine that R2 of 0.9 is the lowest value tolerable to be considered for the

particle size as uniform, the assumption that particle size is uniform can be valid in the

determination of fluid-solid kinetics in the case where CV is less than 0.5. On the other

hand, in the case where CV value is larger than 0.5, the effects of particle size distribution

should be considered in the fluid-solid kinetics.

The effects of the types of size distribution on the overall rate is similar. When the

effect of particle size distribution is significant and must be taken into consideration, the

spread can be represented by the value of the CV regardless of the form of the distribution

function. Also, the effect of size distribution on the reaction rate of porous particles is

expected to be smaller than on nonporous particles, because there will be no effect of size

distribution if pore diffusion is fast unlike for nonporous solids, and the effect will be

identical to the case of nonporous particles if pore diffusion controls the overall reaction

rate of porous particles.

Based on the theoretical kinetic study for the effect of particle size distribution,

justification of the assumption suggested for the assessment of reaction kinetics of reduced

UGS can be evaluated. The coefficient of variation of reduced UGS defined as standard

deviation divided by mean was found to be 0.33. R2 values for the CV values of 0.3 are

larger than 0.98 in every spherical particle case. Therefore, it was assumed that the effect

of particle size distribution of reduced UGS does not have to be considered in the

Page 101: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

90

calculation of kinetics. Based on this calculation, a rate-controlling process can be found

and it seems to follow interfacial reaction controlled kinetics. The activation energy of the

reaction was determined to be 73.9 kJ/mole.

However, the reaction-controlling process may be controlled by solid state

diffusion of iron in titanium hydride based in part, on the fact that the iron removal in the

reduced UGS is not increased with increase of leaching temperature from 140oC to 190oC.

If leaching temperature is high enough for iron to be stably dissolved in titanium hydride,

increasing the leaching temperature might oppress iron removal. Therefore, the

effectiveness of iron removal may be limited by locking up of iron as interstitial

substitutions. However, the rate-controlling process of this mechanism is difficult to

analyze to completely due to the limited information about the diffusion coefficient of iron

in titanium hydride and involvement of other reactions such as dissolution of titanium

hydride.

An additional way of obtaining pure titanium involves the extraction of titanium

from ilmenite. The strong affinity of tannic acid towards metal ions having high oxidation

states was explored to recover titanium selectively from a multielement solution prepared

by dissolving ilmenite in sulfuric acid and ammonium sulfate. Titanium-tannic acid

complex was selectively precipitated from the multielement solution at pH 4.5. The

precipitate was separated by vacuum filtration and the filter cake was washed with 2%

hydrochloric acid solution, deionized water, and cold isopropyl alcohol after filtration. The

washed filter cake was dried at room temperature and ground into powder on drying. The

titanium-tannic acid complex so obtained was composed of 98.6% titanium, 0.2% iron, and

1.2% other impurities. The density functional theory simulation at B3LYP method with a

Page 102: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

91

basis of 6-31G was also conducted to identify optimized complex structure and its stability.

The result showed an energy gap of 2.43 eV between two frontier orbitals (highest occupied

molecular orbitals and lowest unoccupied molecular orbitals), which is large enough to

indicate energetic feasibility of the complex formation.

Page 103: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

REFERENCES

1. Zhou, Y.G.; Zeng, W.D.; Yu, H.Q. Mater. Sci. Eng. A 2005, 393, 204-212.

2. Boyer, R.R. Mater. Sci. Eng. A 1996, 213, 103-114.

3. Niinomi, M. Sci. Tech. Adv. Mater. 2003, 4, 445–454.

4. Gorynin, I.V. Mater. Sci. Eng. A 1996, 263, 112-116.

5. Hua, F.; Mon, K.; Pasupathi, P.; Gordon, G.; Shoesmith, D. Corrosion 2005, 61,

987-1003.

6. Amarchand, S.; Rama Mohan, T.R.; Ramakrishnan, P. Adv. Powder Technol. 2000,

11, 415–422.

7. Froes, F.H.; Eylon, D. Int. Mater. Rev. 1990, 35, 162–182.

8. Middlemas, S.; Fang, Z.Z.; Fan, P. Hydrometallurgy 2013, 131, 107–113.

9. Fang, Z.Z.; Middlemas, S.; Fan, P.; Guo, J. J. Am. Chem. Soc. 2013, 135, 18248–

18251.

10. ASTM B299-13; Standard Specification for Titanium Sponge In ASTM

International; West Conshohocken, Pennsylvania, 2013.

11. Peak, D.; Luther III, G.W.; Sparks, D. L. Geochim. Cosmochim. Acta 2013, 67,

2551–2560.

12. Elrod, J.A.; Kester, D.R. J. Solut. Chem. 1980, 9, 885–894.

13. Szekely, J.; Evans, J.W.; Sohn, H.Y.; Gas-Solid Reactions; Academic Press, New

York, 1976.

14. Sohn, H.Y.; Baek, H.D. Metall. Trans. B 1998, 20, 107-110.

15. Homma, S.; Ogata, S.; Koga, J.; Matsumoto, S. Chem. Eng. Sci. 2005, 60, 4971-

4980.

Page 104: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

93

16. Herbst, J.A.; Rate Processes In Multiparticle Metallurgical Systems In Rate

Processes of Extractive Metallurgy; Sohn, H.Y., Wadsworth, M.E., Eds.; Plenum,

New York, 1979; pp 53-112,

17. Ahmed, M.M.; Ahmed, S.S. J. Eng. Sci. 2008, 36, 147-166.

18. Macias-Garcia, A.; Cuerda-Correa, E.M.; Miaz-Diez, M.A. Mater. Cha. 2004, 52,

159-164.

19. Garabrant, D.H.; Fine, L.J.; Oliver, C.; Bernstein, L.; Peters, J.M. Scandi. J. Wo.

Envir. Heal. 1987, 13, 47-51.

20. Park, I.; Abiko, T.; Okabe, T.H. J. Phys. Chem. Solids 2005, 66, 410–413.

21. Kroll, W.J. Trans. Electrochem. Soc. 1940, 78, 35–47.

22. Takeda, O.; Okabe, T.H. Mater. Trans. 2006, 47, 1145–1154.

23. Chen, G.Z.; Fray, D.J.; Farthing, T.W. Nature Letter 2000, 407, 361-364.

24. Chen, W.; Yamamoto, Y.; Peter, W.H.; Gorti, S.B.; Sabau, A.S.; Clark, M.B.;

Nunn, S.D.; Kiggans, J.O.; Blue, C.A.; Williams, J.C.; Fuller, B.; Akhtar, K. Pow.

Tech. 2011, 214, 194-199.

25. Büchel, K.H.; Moretto, H.H.; Werner, D.; Woditsch, P.; Industrial Inorganic

Chemistry; Wiley-VCH, Mörlenbach, 2000; pp 548-558.

26. Froes, F.H.; Titanium: Physical Metallurgy, Processing, and Applications; ASM

International, Ohio, 2015.

27. El-Hazek, N.; Lasheen, T.A.; El-Sheikh, R.; Zaki, S.A. Hydrometallurgy 2007, 87,

45–50.

28. Ogasawara, T.; Veloso de Araujo, R.V. Hydrometallurgy 2000, 56, 203–216.

29. Van Dyk, J.P.; Vegter, N.M.; Pistorius, P.C. Hydrometallurgy 2000, 65, 31–36.

30. Hulburt, H.M.; Katz, S. Chem. Eng. Sci. 1964, 19, 555.

31. Randolph, A.D. Can. J. Chem. Eng. 1964, 42, 280.

32. Randolph, A.D.; Larson, M.; Theory of Particulate Processes; Academic Press,

New York, 1971; pp 50-76.

33. Mcllvried, H.G.; Massoth, F.E. Ind. Eng. Chem. Fundam. 1973, 12, 225-229.

34. Svarovsky, L.; Solid-Liquid Separation; Butterworths, London, 1977; pp 15-23.

Page 105: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

94

35. Iffat, A.T.; Maqsood, Z.T.; Fatima, N. J. Chem. Soc. Pak. 2005, 27, 174-178.

36. Surleva, A.; Atanasova, P.; Kolusheva, T.; Costadinnova, L. J. Chem. Tech. Met.

2014, 49, 594-600.

37. Araujo, P.Z.; Morando, P.J.; Blesa, M.A. Langmuir 2005, 21, 3470-3474.

38. Borgais, B.A.; Cooper, S.R.; Koh, Y.B.; Raymond, K.N. Inorg. Chem. 1984, 23,

1009-1016.

39. Hem, J.D.; Complexes of Ferrous Iron with Tannic Acid In Chemistry of Iron in

Natural Water; United States Government Printing Office, Washington, 1962; pp

75-94.

40. Finar, I.L.; Organic Chemistry; English Language Book Society & Longman

Group Limited, London, 1975.

41. Astrelin, I.M.; Prokofyeva, G.N.; Suprunchuk, V.I.; Knyazev, Y.V.; Panashenko,

V.M.; Morozov, I.A.; Morozova, R.A.; Interaction of TiHx (x < 2) with solutions

of some acids and alkalies In Hydrogen Materials Science and Chemistry of Metal

Hydrides; Veziroglu, T.N., Zaginaichenko, S.Y., Schur, D.V., Trefilov, V.I., Eds.;

Kluwer Academic Publishers, Netherlands, 2002; pp 133-140,

42. Sohn, H.Y.; Sohn, H.J. Ind. Eng. Chem. Process Des. Dev. 1980, 19, 237-242.

43. Perry, J.H.; Chemical Engineers’ Handbook; McGraw-Hill, New York, 1950; p

67.

44. Corless, R.M.; Gonnet, G.H.; Hare, D.E.G.; Jeffrey, D.J.; Knuth, D.E. Adv. Com.

Math. 1996, 5, 329–359.

45. Schuhmann, R. A.I.M.E. Tech. Publ. 1940, 1189, 1-11.

46. Rosin, P.; Rammler, E. J. Inst. Fuel 1933, 7, 29-36.

47. Belz, M.H.; Statistical Methods for the Process Industries; Wiley, New York, 1973;

p 98.

48. Sohn, H.Y.; Szekely, J. Chem. Eng. Sci. 1972, 27, 763-778.

49. Sohn, H.Y.; Chaubal, P.C. AIChE J. 1986, 32, 1574-1577.

50. Sohn, H.Y. Metall. Trans. B 1978, 9B, 90-96.

51. Sohn, H.Y.; Unpublished Work In Graduate course on Advanced Fluid-Solid

Reaction Engineering; University of Utah, Salt Lake City, Utah, 1981.

Page 106: REMOVAL OF IMPURITIES FROM TITANIUM-BEARING SOURCES …

95

52. Free, M.L.; Hydrometallurgy: Fundamentals and Applications; John Wiley &

Sons, lnc., New Jersey, 2013.

53. Porter, D.A.; Easterling, K.E.; Phase Transformations in Metals and Alloys; VNR

International, Berkshire, 1988.

54. Crank, J.; The Mathematics of Diffusion; Clarendon Press, Bristol, 1975.

55. Becke, A.D. J. Chem. Phys. 1993, 98, 5648-5652.

56. Lee, C.; Yang, W.; Parr, R.G. Phys. Rev. B 1988, 37, 785-789.

57. Tirado-Rives, J.; Jorgensen, W.L. J. Chem. Theo. Comp. 2008, 4, 297-306.

58. Larkin, P.J.; IR and Raman Spectroscopy: Principles and Spectral Interpretation;

Elsevier, Amsterdam, 2011.

59. Peres, R.S.; Cassel, E.; Azambuja, D.S. ISRN Corrosion 2012, 9-19.

60. Zhou, Z.; Parr, R.G. J. Am. Chem. Soc. 1990, 112, 5720–5724.

61. Pearson, R.G. J. Am. Chem. Soc. 1963, 85, 3533–3539.


Recommended