an electrochemical and sers study of the gold-thiosulfate

125
An Electrochemical and SERS Study of the Gold-Thiosulfate Interface in the Presence of Copper by Eric Nicol A Thesis Presented to The University of Guelph In partial fulfillment of requirements for the degree of Master of Science in Chemistry Guelph, Ontario, Canada © Eric Nicol, April 2013

Upload: others

Post on 25-May-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: An Electrochemical and SERS Study of the Gold-Thiosulfate

An Electrochemical and SERS Study of the Gold-Thiosulfate Interface in the Presence of Copper

by

Eric Nicol

A Thesis Presented to

The University of Guelph

In partial fulfillment of requirements for the degree of

Master of Science in

Chemistry

Guelph, Ontario, Canada © Eric Nicol, April 2013

Page 2: An Electrochemical and SERS Study of the Gold-Thiosulfate

ii

ABSTRACT

AN ELECTROCHEMICAL AND SERS STUDY OF THE GOLD-THIOSULFATE INTERFACE IN THE PRESENCE OF COPPER

Eric Nicol Advisor: University of Guelph, 2013 Prof. Jacek Lipkowski

Complementary electrochemical and spectroscopic techniques were used to characterize

the behavior and composition of the passive layer formed at the gold-thiosulfate interface in the

presence of copper. Raman studies of three different cationic (calcium, ammonium and sodium)

thiosulfate leaching solutions showed that the concentrations of sulfate, thiosulfate, trithionate

and tetrathionate remained constant.

Initial leaching current densities for the three systems were identical, however, significant

differences were noted in the open circuit potentials of these systems. Gold nanorod electrodes

were employed as substrates for Surface Enhanced Raman Spectroscopy (SERS) studies of the

gold-thiosulfate interface. The composition and behavior of the passive layer at the gold-

thiosulfate interface greatly differed from that of the bulk solutions. Higher order polythionate

species were not observed, and significant differences were noted in the behavior of species

common between the three thiosulfate leaching solutions. Passivation levels determined from

SERS indicate that in the presence of copper, the cation associated with thiosulfate may play a

key role in the extent of passivation on the gold surface.

Page 3: An Electrochemical and SERS Study of the Gold-Thiosulfate

iii

ACKNOWLEDGEMENTS

Many people contributed to the success of this project, and to the maintenance of my sanity

throughout. I would like to thank my supervisor Professor Jacek Lipkowski for being a fantastic

source of knowledge and guidance since I first emailed him asking for a job. His patience,

positive attitude and willingness to discuss challenges were instrumental in the success of this

work.

I would like to express my gratitude to the members of my advisory and examination

committees: Professors Peter Tremaine, Mark Baker, and Abdelaziz Houmam for their time

dedicated to this thesis, as well as Professor Paul Rowntree for chairing both my proposal and

defence.

I would like to recognize the financial support from Barrick Gold Corporation and in

particular Dr. Yeonuk Choi who was actively involved in the project.

I would also like to acknowledge the people who collaborated with the experimental

development and execution of this project: Dr. Janet Baron Gavidia without whom I would have

been lost and confused, and Scott Smith who has done an excellent job of filling Dr. Baron

Gavidia’s shoes. To past and present members of Dr. Lipkowski’s lab group at the University of

Guelph, I would like to express thanks for an unforgettable experience. Especially to Jay Leitch

and Ryan Seenath, who provided invaluable discussion and advice at coffee, and never let my

head swell too large. To my friends Michael Yacyshyn, Dave Sullivan and Christian Carello,

thanks for helping make my successes fantastic and my disasters not so bad.

Lastly, but not least, I would like to give my utmost gratitude to my amazing family and to

Sara Trayes who were cheerleaders when I wanted to quit, and without whom, achievement of

my Masters would not have been possible.

Page 4: An Electrochemical and SERS Study of the Gold-Thiosulfate

iv

TABLE OF CONTENTS

Acknowledgements ........................................................................................................................ iii  Table of Contents ........................................................................................................................... iv  List of Tables ................................................................................................................................. vi  List of Figures ................................................................................................................................ vi  List of Symbols and Abbreviations ................................................................................................ xi   Chapter 1: Introduction ................................................................................................................... 1  

1.1 Motivation and Impact: ......................................................................................................... 1  1.2 Objectives: ............................................................................................................................ 3  1.3 Scope: .................................................................................................................................... 4  

Chapter 2: Background and Literature Review .............................................................................. 6  2.1 Gold Ore and the Hydrometallurgical Process ..................................................................... 6  2.2 Leaching Mechanism and History ...................................................................................... 12  

2.2.1 Cyanidation .................................................................................................................. 13  2.2.2 Environmental Concerns .............................................................................................. 15  2.2.3 Thiosulfate Alternative ................................................................................................ 15  

2.3 Thiosulfate Leaching of Gold ............................................................................................. 16  2.3.1 Ammonia-Thiosulfate Leaching .................................................................................. 21  2.3.2 The Role of the Electrolyte on Ammonia-Thiosulfate Leaching ................................. 25  

Chapter 3: Experimental Techniques Theory ............................................................................... 28  3.1 Electrochemical Techniques ............................................................................................... 28  

3.1.1 Sweep Voltammetry Methods ...................................................................................... 28  3.1.2 Mixed Potential Theory ............................................................................................... 36  3.1.3 Hydrodynamic Methods ............................................................................................... 39  

3.2 Spectroscopic Techniques ................................................................................................... 42  3.2.1 Surface Enhanced Raman Spectroscopy ...................................................................... 42  3.2.2 SERS Substrates .......................................................................................................... 49  

Chapter 4: Methodology ............................................................................................................... 52  4.1 Reagents .............................................................................................................................. 52  4.2 Cleaning Methods ............................................................................................................... 52  4.3 Experimental Setup and Process ......................................................................................... 54  

4.3.1 Preparation of Gold Electrodes .................................................................................... 54  4.3.2 Electrochemical Experiments ...................................................................................... 54  

4.3.2.1 Leaching Current Measurements .......................................................................... 54  4.3.2.2 Open Circuit Potential Measurements .................................................................. 55  4.3.2.3 Solution pH Measurements ................................................................................... 56  

4.3.3 Preparation of Gold Nanorod Electrodes ..................................................................... 57  4.3.4 Raman Experiments ..................................................................................................... 58  

4.3.4.1 SERS ..................................................................................................................... 58  4.3.4.2 Solution Raman ..................................................................................................... 59  

Chapter 5: Results and Discussion ................................................................................................ 60  5.1 Preamble ............................................................................................................................. 60  5.2 Characterization of Bulk Solution ...................................................................................... 61  

5.2.1 pH Measurements ........................................................................................................ 62  

Page 5: An Electrochemical and SERS Study of the Gold-Thiosulfate

v

5.2.2 Solution Raman ............................................................................................................ 64  5.3 Characterization of the Gold-Thiosulfate Interface in the Presence of Copper .................. 71  

5.3.1 Initial Characterization ................................................................................................. 71  5.3.2 Leaching Current Measurements ................................................................................. 79  5.3.3 Open Circuit Potential Measurements ......................................................................... 81  5.3.4 Surface Enhanced Raman ............................................................................................ 83  

5.4 Summary ........................................................................................................................... 101  Chapter 6: Conclusions and Future Work ................................................................................... 106  

6.1 Conclusions ....................................................................................................................... 106  6.2 Future Work ...................................................................................................................... 107  

Page 6: An Electrochemical and SERS Study of the Gold-Thiosulfate

vi

LIST OF TABLES

Table 5.1: Characteristic Raman active vibrational modes of species expected in the Raman

spectra of the investigated leaching solutions.............................................................................65

Table 5.2: Final change in the normalized peak areas investigated in the calcium, ammonium

and thiosulfate systems.............................................................................................................102

LIST OF FIGURES

Figure 2.1: Flow chart for the processing of gold ore to the pure metal.....................................9

Figure 2.2. Schematic for the electrochemical model for leaching of gold...............................12

Figure 2.3: SERS spectrum of 0.1 M Na2S2O3 at open circuit potential...................................19

Figure 2.4: SERS spectrum of a gold nano-rod electrode submersed in a

0.1M Na2S2O3 + 1 x 10-4M NaOH at open circuit potential.......................................................20

Figure 2.5: Calculated voltammetric current for varying concentrations of ammonia in

thiosulfate leaching solution.......................................................................................................21

Figure 2.6: The effect of copper on gold dissolution in thiosulfate leaching............................23

Figure 2.7: Electrochemical model for the leaching of gold using ammoniacal thiosulfate leach

media...........................................................................................................................................25

Figure 3.1: a) Typical potential time curve for cyclic voltammetry b) Current response

curve............................................................................................................................................28

Figure 3.2: Schematic of the structure of the electrical double layer formed at the metal-

electrolyte interface.....................................................................................................................30

Figure 3.3: Tafel plot of a redox system....................................................................................33

Page 7: An Electrochemical and SERS Study of the Gold-Thiosulfate

vii

Figure 3.4: Diagram of the free energy curves of an oxidized and reduced species when a

potential is applied, shifting the system from equilibrium...........................................................35

Figure 3.5: Intersection of the locally linear potential energy surface of an oxidized and reduced

species..........................................................................................................................................35

Figure 3.6: Generic linear sweep voltammogram.......................................................................37

Figure 3.7: Tafel plot a gold electrode in contact with a 0.1 M Na2S2O3 + 0.01 M CuSO4 +

1.1µΜ Ca(OH)2 solution..............................................................................................................37

Figure 3.8: Plot of the current versus the overpotential in a ~40mV range around the mixed

potential for a 0.1 M Na2S2O3 + 0.01 M CuSO4 + 1.1µΜ Ca(OH)2 system................................39

Figure 3.9: a) Profile for a solution with laminar flow b) Profile for a solution with turbulent

flow...............................................................................................................................................40

Figure 3.10: Transitions seen in Raman spectroscopy................................................................45

Figure 3.11: Photon-driven charge-transfer process...................................................................47

Figure 4.1: Cyclic voltammogram of a clean, bare gold RDE in a solution of 0.1 M

NaF...............................................................................................................................................53

Figure 4.2: Schematic for the two cell configurations used.......................................................55

Figure 4.3: Calibration curve for the calculation of pH from the potential measured................56

Figure 5.1: Average pH over a period of 3 hours for calcium, ammonium and sodium thiosulfate

systems (0.1 M S2O3 + 0.01 M CuSO4 and adjusted to a pH of 8.0-8.5) and a blank solution of

pure MilliQ water.........................................................................................................................63

Figure 5.2: Average raw Raman spectra collected over a period of 3 hours for leaching solutions

of: a) 0.1 M CaS2O3 + 0.01 M CuSO4 b) 0.1 M (NH4)2S2O3 + 0.01 M CuSO4 c) 0.1 M Na2S2O3

+ 0.01 M CuSO4...........................................................................................................................66

Page 8: An Electrochemical and SERS Study of the Gold-Thiosulfate

viii

Figure 5.3: Deconvoluted spectra of an average of three 0.1 M CaS2O3 + 0.01 M CuSO4

solutions, adjusted to a pH of 8.0-8.5, 20 min after solution preparation...................................67

Figure 5.4: Deconvoluted spectra of an average of three 0.1 M CaS2O3 + 0.01 M CuSO4

solutions, adjusted to a pH of 8.0-8.5, 20 min after solution preparation, in the 900-1000 cm-1

region...........................................................................................................................................68

Figure 5.5: Normalized peak areas for band positions of 1037 cm-1 (tetrathionate = black), 425

cm-1 (trithionate = blue), 980 cm-1 (sulfate = green), 448 cm-1 (thiosulfate = red). a) 0.1 M

CaS2O3 + 0.01 M CuSO4, b) 0.1 M Na2S2O3 + 0.01 M CuSO4, c) 0.1 M (NH4)2S2O3 + 0.01 M

CuSO4...........................................................................................................................................69

Figure 5.6: Linear sweep voltammogram of a 0.1 M Na2S2O3 + 0.01 M CuSO4 solution, pH 8.0-

8.5.................................................................................................................................................71

Figure 5.7: Tafel plot of a 0.1 M Na2S2O3 + 0.01 M CuSO4 solution, pH 8.0-

8.5.................................................................................................................................................72

Figure 5.8: Linear regression of the cathodic reduction of Cu2+ for the determination of the

transfer coefficient........................................................................................................................73

Figure 5.9: Linear regression of the anodic oxidation of Au for the determination of the transfer

coefficient.....................................................................................................................................74

Figure 5.10: Linear regression of data acquired during a linear sweep voltammogram of a 0.1 M

Na2S2O3 + 0.01 M CuSO4 solution, pH 8.0-8.5 solution with a sweep rate of 1 mVs-1..............75

Figure 5.11: Sweep rate dependence of the leaching current measured in a 0.1 M Na2S2O3 +

0.01 M CuSO4 solution, pH 8.0-8.5, at sweep rates of 1, 2, 10, 20, 50 and 100 mVs-1..............77

Page 9: An Electrochemical and SERS Study of the Gold-Thiosulfate

ix

Figure 5.12: Rotation dependence of the leaching current of a gold electrode in contact with a

0.1 M CaS2O3 + 0.01 M CuSO4 solution, pH 8.0-8.5, at rotation rates of 300, 500, 700 and 1000

RPM..............................................................................................................................................78

Figure 5.13: Average calculated leaching current density for solutions of sodium, calcium and

ammonium thiosulfate solutions (0.1 M S2O3 + 0.01 M CuSO4 and were adjusted to a pH of 8.0-

8.5)................................................................................................................................................80

Figure 5.14: Average open circuit potential of a gold disk RDE as a function of immersion time

in solutions of calcium, sodium and ammonium thiosulfate (0.1 M S2O3 + 0.01 M CuSO4 and

adjusted to a pH of 8.0-8.5)...........................................................................................................82

Figure 5.15: Raw SERS spectra of gold nanorod electrodes exposed to leaching solutions of 0.1

M S2O3 + 0.01 M CuSO4, adjusted to pH 8.0-8.5. Solutions were a) CaTS b) ATS c)

STS................................................................................................................................................85

Figure 5.16: Fitted SERS spectrum of a gold nanorod electrode, after 5 minutes of exposure to a

0.1 M CaS2O3 + 0.01 M CuSO4 leaching solution, with an initial pH of 8.0-

8.5..................................................................................................................................................86

Figure 5.17: Fitted SERS spectrum of a gold nanorod electrode, after 15 minutes of exposure to

a 0.1 M (NH4)2S2O3 + 0.01 M CuSO4 leaching solution, with an initial pH of 8.0-

8.5..................................................................................................................................................88

Figure 5.18: Fitted SERS spectrum of a gold nanorod electrode, after 5 minutes of exposure to a

0.1 M Na2S2O3 + 0.01 M CuSO4 leaching solution, with an initial pH of 8.0-

8.5..................................................................................................................................................90

Figure 5.19: Normalized analytical peak areas for a gold nanorod electrode treated with 0.1 M

CaS2O3 + 0.01 M CuSO4 solution adjusted to pH 8.0-8.5. Peak areas tracked were for band

Page 10: An Electrochemical and SERS Study of the Gold-Thiosulfate

x

positions of a) 216 cm-1 (green circle), 255 cm-1 and 300 cm-1 (brown square), and b) 400 cm-1

(blue circle), 443 cm-1 (red square), and 610 cm-1 (pink diamond)............................................92

Figure 5.20: Analytical peak areas for a gold nanorod electrode treated with 0.1 M (NH4)2S2O3

+ 0.01 M CuSO4 solution adjusted to pH 8.0-8.5. Peak areas tracked were for band positions of:

a) 216 cm-1 (green circle), 255 cm-1 and 300 cm-1 (brown square), and b) 400 cm-1 (blue circle),

443 cm-1 (red square), and 610 cm-1 (pink diamond)..................................................................96

Figure 5.21: Analytical peak areas for a gold nanorod electrode treated with 0.1 M Na2S2O3 +

0.01 M CuSO4 solution adjusted to pH 8.0-8.5. Peak areas tracked were for band positions of: a)

216 cm-1 (green circle), 255 cm-1 and 300 cm-1 (brown square), and b) 400 cm-1 (blue circle), 443

cm-1 (red square), and 610 cm-1 (pink diamond)..........................................................................99

Figure 5.22: Normalized analytical peak area of a gold nanorod electrode treated with 0.1 M

Na2S2O3 solution adjusted to pH 10............................................................................................104

Page 11: An Electrochemical and SERS Study of the Gold-Thiosulfate

xi

LIST OF SYMBOLS AND ABBREVIATIONS

CHAPTER 1:

SERS – Surface enhanced Raman spectroscopy

CHAPTER 2:

REQCM – Rotating electrochemical quartz crystal microbalance

SCE – Saturated calomel electrode

ν  –  Stretching vibration

δ – Bending vibration

νsym  –  Symmetric stretching vibration

FSD – Fourier self-deconvolution

CHAPTER 3:

E – Applied potential

E1/E2/E3 – Potential limits

OHP – Outer Helmholtz plane

IHP – Inner Helmholtz plane

νf  – Rate of reduction reaction

νb – Rate of oxidation reaction

ic – Cathodic current

ia – Anodic current

CO – Surface concentration of oxidized species

CR – Surface concentration of reduced species

Page 12: An Electrochemical and SERS Study of the Gold-Thiosulfate

xii

n – number of electrons

A – Electrode area

F – Faraday’s constant

kf – Rate constant of reduction reaction

kb – Rate constant of oxidation reaction

α – Transfer coefficient

k0 – Standard rate constant

E0 – Standard potential

R – Gas constant

T – Temperature

i – Measured current

Eeq – Equilibrium potential

i0 – Exchange current

η – Overpotential

Rct – Charge transfer resistance

iM – Leaching current

EM – Mixed potential

αCu2+ – Transfer coefficient for the reduction of Cu2+

αAu – Transfer coefficient for the oxidation of Au

Re – Reynold’s number

υch – Characteristic velocity of a fluid

l – Characteristic length

ν – Kinematic viscosity

Page 13: An Electrochemical and SERS Study of the Gold-Thiosulfate

xiii

D – Diffusion coefficient

CO* - Bulk concentration of oxidized species

ω – angular velocity

ik – Current in the absence of mass transfer effects

il.c – Cathodic limiting current

EF – External electric field

E0 – Amplitude of external electric field wave

!ex – Frequency of exciting light

! – Induced dipole moment

!P – Polarizability of the molecule

!0 – Polarizability of the molecule at the equilibrium position

r – Internuclear distance

req – Internuclear distance at the equilibrium position

rm – Maximum internuclear distance

!v – Frequency of vibration of molecule

HOMO – Highest occupied molecular orbital

LUMO – Lowest occupied molecular orbital

Eout - Magnitude of the electric field produced outside a spherical nanoparticle of radius

a

(x, y, z) – Cartesian coordinates

(x, y, z) – Cartesian unit vectors

rd – Radial distribution of electric field

!M – Polarizability of the metal

Page 14: An Electrochemical and SERS Study of the Gold-Thiosulfate

xiv

εin – Dielectric constant of the nanoparticle

εout – Dielectric constant of the environment surrounding the nanoparticle

E(λ) – Extinction spectrum of a non-spherical nanoparticle

N – Finite number of polarizable elements within a nanoparticle

εi – Imaginary component of the dielectric constant of the nanoparticle

εr – Real component of the dielectric constant of the nanoparticle

λ – Wavelength

a – Radius of nanoparticle

! – Shape factor which accounts for geometries of the nanoparticle, other than spherical

LSPR – Localized surface plasmon resonance

CHAPTER 4:

RDE – Rotating disk electrode

CV – Cyclic voltammogram

RE – Reference electrode

CE – Counter electrode

E – Applied potential

i – Measured current

CHAPTER 5:

SERS – Surface enhanced Raman spectroscopy

FSD – Fourier self deconvolution

! – Rocking vibration

Page 15: An Electrochemical and SERS Study of the Gold-Thiosulfate

xv

νsym – Symmetric stretching vibration

νasym – asymmetric stretching vibration

δsym – Symmetric bending vibration

δasym – Asymmetric bending vibration

i – Measured current

V – Applied potential

!c – Cathodic transfer coefficient

!a  – Anodic transfer coefficient

F – Faraday’s constant

T – Temperature

R – Gas constant

i0 – Exchange current

n – Number of electrons

! - Overpotential

!Cu2+ – Transfer coefficient for the reduction of Cu2+

!Au – Transfer coefficient for the oxidation of Au

nCu2+ - Number of electrons transferred in the reduction of Cu2+

nAu – Number of electrons transferred in the oxidation of Au

iM – Leaching current at the mixed potential

SCE – Saturated calomel electrode

RDE – Rotating disk electrode

ik – Current in the absence of mass transfer effects

il.c – Cathodic limiting current

Page 16: An Electrochemical and SERS Study of the Gold-Thiosulfate

xvi

D – Diffusion coefficient

A – Area

! – Angular velocity

ν – Kinematic viscosity

CO* - Bulk concentration of oxidized species

CaTS – Calcium thiosulfate

ATS – Ammonium thiosulfate

STS – Sodium thiosulfate

OCP – Open circuit potential

CHAPTER 6:

SERS – Surface enhanced Raman spectroscopy

Page 17: An Electrochemical and SERS Study of the Gold-Thiosulfate

1

 

CHAPTER 1: INTRODUCTION

1.1 MOTIVATION AND IMPACT:

Metallic gold has application in a number of industries including the chemical catalysis

and microelectronics fabrication, as a result of exceptional properties such as its catalytic ability,

corrosion resistance, thermal conductivity, and ductility.1

Extraction from its ore is highly dependent on the state of gold within, and its response to

the use of leaching reagents.2 The leaching process involves the addition of a chemical agent to

oxidize the gold, and a complexing agent to bind the Au+ ion and bring it into solution. In most

instances, reduction of molecular oxygen is coupled with the oxidation of gold. The solution

containing the complexed gold ion is later submitted to a process such as electrowinning or

adsorption onto activated carbon to complete extraction of the gold.3

Cyanide is by far the most widely used lixiviant, or leaching reagent, in use today for

extraction of gold from its ores.3-6 However, due to recent concerns over its toxicity and

environmental impact, there has been an increased interest in alternative leaching reagents.3, 6

Stable thiosulfate-gold complexes have been known for over 100 years, but serious

investigations into its use in leaching has only begun in the last two decades. Thiosulfate is

attractive due to its remarkable ability to deal with refractory, preg-robbing and other ores that

may be difficult or non-responsive to cyanide.7 Due to its low toxicity, and comparatively high

initial rate of leaching, thiosulfate is perhaps the most promising alternative to cyanide

leaching.3, 4, 6 This is especially true in the presence of copper and ammonia in the leaching

solution. However, although initial rates of gold leaching from cyanide and thiosulfate solutions

Page 18: An Electrochemical and SERS Study of the Gold-Thiosulfate

2

are comparable, with extended time leaching, only 60-70% of gold is typically leached in the

presence of thiosulfate.4

The effectiveness of thiosulfate leaching in industry is limited by the formation of a

passive layer on the gold surface, composed of a combination of sulfide and other decomposition

products of the thiosulfate.8, 9 Recent research has shown that the passivating layer consisted of

sulfide and polythionate species10, including thiosulfate, tetrathionate and trithionate.11

The individual addition of both ammonia and copper into thiosulfate leaching solutions

has been shown to increase the rate of leaching and aid in the inhibition of passive layer

formation. Addition of ammonia in to thiosulfate leaching solution results in the formation of an

anionic aurocomplex leading to an increased rate of gold oxidation, providing stability over a

wide pH range, and inhibiting interference from species such as iron oxides, silica, silicates and

carbonates.4, 12 Ammoniacal solutions require the addition of copper due to the inability of O2 to

act as the oxidant.4 Copper, in the absence of ammonia, has also been shown to affect gold

dissolution rates. Zhang and Nicol found that the addition of copper in different forms resulted in

significant increases in the amount of gold dissolved.13

In aqueous thiosulfate solutions containing both copper and ammonia, an increased

leaching rate has been observed, in part due to the formation of an oxidizing copper tetraamine

complex that allows for rapid gold oxidation.6, 14, 15 Although the effects of copper and ammonia

addition on gold dissolution and oxidation are known, the mechanisms through which copper and

ammonia act individually, and in tandem, to affect passive layer formation remain unknown.

The composition of the electrolyte is also known to affect on the leaching rate of gold in

thiosulfate media. Chandra et al.16 observed a significant differences on the ease of oxidation of

gold between sodium, potassium and ammonium thiosulfate solutions. Potassium thiosulfate

Page 19: An Electrochemical and SERS Study of the Gold-Thiosulfate

3

showed considerable improvement over sodium thiosulfate; however, ammonium was the most

effective cation at increasing the leaching rate. Although explanations for the preference of

heavier alkali metals exist, no explanation for the effect on gold oxidation has yet been provided.

By using a combination of electrochemical techniques and Surface Enhanced Raman

Spectroscopy (SERS), the roles of both Cu2+ and the electrolyte cation in passive layer

composition during thiosulfate leaching of gold can be clarified. Application of the knowledge

gained in the study will ultimately aid in the understanding an improvement of the leaching

process.

Gaining an understanding of the leaching process in the presence of Cu2+ is invaluable

due to its direct application in industrial practices. Ideally, if the function of copper is

understood, optimization of the thiosulfate leaching system could be undertaken in an attempt to

obtain a viable competitor for cyanide leaching; one that is significantly more environmentally

responsible, shows improved gold recovery, and is economically sustainable.

1.2 OBJECTIVES:

Application of thiosulfate leaching of gold in industry requires a complete understanding

of the behavior and composition of the passive layer under applicable conditions. As such, the

role of copper in passive layer formation is vital. The specific objectives of this project are:

• To determine the species composing the passive layer at the gold-thiosulfate interface in

the presence of copper, and determine their behavior through time

• To investigate the effect of the cation in the electrolyte on the composition of the passive

layer at the interface

• To identify which species may be promoting leaching, and which promote passivation

through complementary electrochemical and spectroscopic techniques

Page 20: An Electrochemical and SERS Study of the Gold-Thiosulfate

4

• To identify dominating reactions at the gold-thiosulfate interface, and determine how

they influence the composition of the passive layer

1.3 SCOPE:

This work consists of 6 chapters. The first chapter is an introduction to the main body of

this thesis, including a brief background of the thiosulfate leaching system, and previous studies

on the effect of copper. The second chapter will include a detailed review of thiosulfate leaching

media, and previous studies for characterization of the passive layer, and its composition. Gold

dissolution rates in the presence and absence of copper will be addressed along with the effect of

ammonia, individually and in combination with copper.

Chapter 3 will summarize the theory behind the experimental techniques used in this

project; electrochemical theory for sweep voltammetry methods and mixed potential theory, as

well as hydrodynamic method theory will be discussed. Fundamentals of both Raman

spectroscopy and Surface Enhanced Raman Spectroscopy (SERS) are examined. Chapter 4

presents the experimental methodology used for all experiments, and the reagents and solutions

used for these investigations.

The fifth chapter presents the results acquired from electrochemical and spectroscopic

methods in calcium, sodium and ammonium thiosulfate leaching media in the presence of

copper. The results are analyzed and discussed in terms of studies of the bulk solution and the

passive layer at the interface, as a function of time. The sixth and final chapter presents the

conclusions of the study as well as suggestions for future work to further the understanding of

the industrial leaching process using thiosulfate media.

Page 21: An Electrochemical and SERS Study of the Gold-Thiosulfate

5

References

1. Sullivan, A. M.; Kohl, P. A. J. Electrochem. Soc. 1997, 144, 1686-1690. 2. Gupta, C.; Mukherjee, T. Boca Raton 1990, , 127-165. 3. Hilson, G.; Monhemius, A. J. J. Clean. Prod. 2006, 14, 1158-1167. 4. Abbruzzese, C.; Fornari, P.; Massidda, R.; Vegliņ, F.; Ubaldini, S. Hydrometallurgy

1995, 39, 265-276. 5. Zhang, S. C.; Nicol, M. J. J. Appl. Electrochem. 2003, 33, 767-775. 6. Jeffrey, M. I.; Linda, L.; Breuer, P. L.; Chu, C. K. Minerals Eng 2002, 15, 1173-1180. 7. Marsden, J.; House, I. In The chemistry of gold extraction; Society for Mining

Metallurgy: 2006; . 8. Chu, C. K.; Breuer, P. L.; Jeffrey, M. I. Minerals Eng 2003, 16, 265-271. 9. Pedraza, A. M.; Villegas, I.; Freund, P. L.; Chornik, B. J Electroanal Chem 1988, 250,

443-449. 10. Woods, R.; Hope, G. A.; Watling, K. M.; Jeffrey, M. I. J. Electrochem. Soc. 2006, 153,

D105-D113. 11. Baron Gavidia, J. Study of the Gold-Thiosulfate Interface Under Leaching Conditions,

University of Guelph, Guelph, ON, 2010. 12. Aylmore, M. G.; Muir, D. M. Minerals Eng 2001, 14, 135-174. 13. Zhang, S. C.; Nicol, M. J. J. Appl. Electrochem. 2005, 35, 339-345. 14. Aylmore, M. G.; Muir, D. M. Miner Metall Process 2001, 18, 221-227. 15. Senanayake, G. J. Colloid Interface Sci. 2005, 286, 253-257. 16. Chandra, I.; Jeffrey, M. I. Hydrometallurgy 2004, 73, 305-312.

Page 22: An Electrochemical and SERS Study of the Gold-Thiosulfate

6

CHAPTER 2: BACKGROUND AND LITERATURE REVIEW

2.1 GOLD ORE AND THE HYDROMETALLURGICAL PROCESS

The inert character of gold leads to very limited formation of naturally occurring

compounds within the Earth’s crust. The average gold content (0.005g/t) is far below that of

other metals such as copper and silver (50 g/t and 0.07 g/t respectively), and is mostly found in

residual hydrothermal fluids and metallic and sulfidic sub-phases1. Rocks with a high

concentration of clay, and low concentration of carbonates are considered some of the best

sources, partly due to ease of re-precipitation upon contact with a reducing environment (regions

with high carbonate, carbon or reducing sulfide).1

Native gold has been found to contain gold in concentrations up to 99.8%, however in

most cases gold content is in the range of 85 to 95%, with silver as the main impurity. If the

silver content is between 25 and 55%, the mineral is called electrum1, 2. The density of native

gold is less than that of pure gold (15 000 kg/m3 versus 19 300 kg/m3), and can consequently be

easily recovered with gravity concentration methods when combined with heavier minerals such

as in gangue minerals (i.e. quartz, silicates etc.)1. Gold tellurides are also a widespread form of

native gold, usually accompanied by free gold and sulfide minerals. Low concentrations of gold

are known to occur with bismuth and copper minerals, but such deposits are relatively rare.

Entrapment of ultrafine gold particles can occur within sulfide mineral grain structures,

occurring in concentrations as high as 15 000 g/t in arsenopyrite, for instance.

Gold ore can be classified into 3 main categories: placer deposits, free milling ores and

lastly, refractory ores3. Placers are deposits that have formed as a direct result of weathering and

Page 23: An Electrochemical and SERS Study of the Gold-Thiosulfate

7

hydraulic transport of gold particles away from the main source1, 3. They require a primary

source of gold (such as quartz-veins, auriferous sulfide deposits or former placers), a long period

of weathering (both chemical and physical) to liberate gold grains from the primary source,

concentration of gold particles by gravity, and a long period of stable bedrock and surface

conditions to allow for a significant concentration to conglomerate1. Leaching is rarely required

for processing of these ores due to the freedom and coarseness of the gold grains3.

Free-milling ores are defined as those which cyanidation (leaching with cyanide) can

extract approximately 95% of the gold, when 80% of the ore is ground to a size less than 75 µm1,

3. These ores contain gold that is finely distributed in a hard rock matrix (typically quartz) and

require thorough grinding in order to release the gold particles before they can be subjected to

subsequent extraction procedures3.

Refractory ores are unresponsive to the cyanide leaching method developed for

processing of free-milling ores. Generally, refractory ores can be grouped into three

subcategories: sulfide ores, carbonaceous ores and telluride ores. Sulfide ores are the largest

group of refractory ores. Gold can be contained within a wide range of host minerals, with pyrite

and arsenopyrite the most common3. Typically these ores contain gold as submicron sized

particles within sulfide grains, which are themselves usually finely distributed within a quartz or

other hard rock matrix1, 3, 4. Carbonaceous ores pose problems due to the preg-robbing effect

(adsorption of gold from solution as leaching is attempted) and formation of a chemical bond

between gold and the carbonaceous material. Gold tellurides, both in the presence and absence of

silver, dissolve very slightly, if at all in cyanide solutions. As the supply of both placer and free-

milling ores is depleted, greater focus is being placed on the extraction and processing of

refractory ores.

Page 24: An Electrochemical and SERS Study of the Gold-Thiosulfate

8

Ore preparation is the first process in gold extraction, and has a very large impact on the

overall recovery and entrapment of gold within the host mineral1, 3. Figure 2.1 displays a

flowchart description of the overall gold extraction and recovery process. Ore preparation can

involve size reduction, solid-liquid separation, and in some cases where floatation is to be used,

dewatering3. The optimum size is determined by the economics of the process, including reaction

kinetics, reagent consumption and grinding costs for a given processing method. Permeability

and separation efficiency can also play a large role in determining the particulate size1. For each

of the three main extraction methods (flotation, cyanide leaching and oxidative pretreatment

followed by cyanide leaching) particle size can play an important role in process time and

consumption.

In the process of flotation, ground ore, as a finely divided powder, is mixed with water to

form a slurry. Surfactant is added to bind the gold particles, rendering the surface hydrophobic.

Froth containing gold particles is created by introduction of the slurry into an aerated water bath.

The froth rises to the surface of the tank, where it is removed and further concentrated. Froth

flotation can be used for free gold and gold-bearing sulfide minerals. It allows for pre-

concentration, removal of sulfide (producing sulfide-free tailings), and removal of interfering

matrix components such as carbonates or other carbonaceous materials. Differential flotation (a

multi-stage process where specific minerals are targeted for flotation at each step) can also be

used for the separation of gold from host minerals such as pyrite, and arsenopyrite1.

Page 25: An Electrochemical and SERS Study of the Gold-Thiosulfate

9

Figure 2.1: Flow chart for the processing of gold ore to the pure metal. Adapted from reference5.

Oxidative pretreatment is required for ore that gives poor gold recovery through

conventional leaching methods1. The pretreatment can be performed through chemical oxidation,

dissociation or roasting, biooxidation or acid leaching (usually carried out at high pressure or

temperature in the presence of a strong oxidant)4. Mineralogy plays a large role in determining

Pit/Ore   Source

Ore  Preparation

Leaching

Solution Collection

Pregnant  Pond Barren  Pond

Solid/Liquid  Separation

Solution Application

Adsorption  on Charcoal

Solvent Extraction

Ion Exchange

Crystallization Ionic  Precipitation Gas Reduction

Electrochemical Reduction

Electrolytic Reduction

Pure  Metal  Compound

Metal  Compound Metal  or  its  oxide Impure  Metal Pure  Metal

Leach/Liquor

Page 26: An Electrochemical and SERS Study of the Gold-Thiosulfate

10

the degree of oxidation used; partial oxidation may suffice to passivate the surfaces of sulphide

minerals in refractory ores, or release gold that is associated with a specific mineral. Complete

oxidation is typically required if gold is finely dispersed within the ore in question, or locked

within sulfide minerals1.

Leaching is used in every hydrometallurgical extraction of gold to produce a gold bearing

solution by dissolving the constituents of the ore1, 5. It is a solid-liquid mass transfer process that

can be carried out at ambient conditions, or at elevated temperatures and pressures; the process

conditions are greatly dependent on the chemical reactions taking place, and their economics5.

Leaching reagents (also called lixiviants) must:

• Dissolve ore minerals rapidly enough for the process to be economical, and ideally be

chemically inert towards gangue minerals

• Be cheap and easily obtainable on the industrial scale

• Ideally, be regenerated in someway during the leaching reaction or process

The leaching process can be applied in four ways: agitated leaching, heap or dump leaching, vat

leaching, and intensive leaching1. Agitated leaching can be used for ore whose particle sizes

don’t allow for passage of the leach solution between the mineral and gold, particularly

submicron sized gold particles. These particles are added to a solvent within an agitated vessel,

where agitation is commonly incurred though gas injection or a rotating impeller. Agitation

ensures that the particles remain finely distributed in solution, increasing the percentage of gold

dissolved5. It is typically used for extraction of gold from unprocessed gangue ore, and may

require further processing, such as solid-liquid separation or the addition of carbon or ion-

exchange resins “in-pulp” for sequestering of the gold.

Page 27: An Electrochemical and SERS Study of the Gold-Thiosulfate

11

Heap (or dump) leaching is applied to ore which contains gold that be at least partially

liberated without grinding. As such, it is typically only usable with permeable ores 1. In this

process, the leaching solution is sprayed over the top of the ore pile, which rests on a pad with a

slight slope, and is allowed to drain through the pile, oxidizing and solubilizing gold into the

leach solution, which is collected at the bottom of the drainage channels5.

Vat leaching is similar to heap leaching, but the volume of the leaching media or solution

is greatly increased. The ore is contained within an impermeable vat or vessel, and is

successively treated with the leach solution. This can be achieved either through continuous flow,

or on a batch basis. Continuous flow vat leaching typically occurs with upward percolation,

where the leach solution exits through the top, whereas batch leaching generally uses downward

percolation, in which the leach solution exits at the bottom of the vat1.

Intensive leaching is a combination of vat and agitated leaching. The ore is contained

within a closed reactor system where mechanical agitation and high reagent concentrations

ensure high leaching kinetics. The reactors run continuously and have optimal volumes

dependent on the residence time. If necessary, they can also be run at elevated pressures and

temperatures1, 5.

Page 28: An Electrochemical and SERS Study of the Gold-Thiosulfate

12

2.2 LEACHING MECHANISM AND HISTORY

As stated previously, the leaching of gold involves a solid-to-liquid mass transfer process

across the gold-solution interface. The process is inherently electrochemical in nature,

composed of the anodic oxidation of gold to form a singly charged ion that reacts to form a gold

complex, and cathodic reduction of the oxidant (typically oxygen)1. A diagram of the dissolution

mechanism can be found in Figure 2.2.

Figure 2.2. Schematic for the electrochemical model for leaching of gold (adapted from reference6).

Once oxidized, gold(I) reacts with the lixiviant, or leaching reagent, to form a soluble

gold complex that undergoes mass transport to the bulk solution. Current industrial practices

make use of cyanide as the lixiviant for complexation. Chlorine-chloride leaching was applied

commercially in the 19th century, but use diminished greatly upon introduction of the cyanide

Page 29: An Electrochemical and SERS Study of the Gold-Thiosulfate

13

process is 18891. Other lixiviants such as ammonia, sulfide, thiocyanate, thiourea and thiosulfate

have been investigated, however, the complexity of these systems limits their application in

industry1.

2.2.1 Cyanidation

The process of cyanidation includes the dissolution of gold into an aerated cyanide

solution to form a dicyano-auro complex. The reaction proceeds via the Elsner equation 3, 7-9:

4Au + 8CN! + O2 + H2O ! 4Au(CN)2! + 4OH! (2.1)

The overall reaction provided by the Elsner equation is the sum of the electrochemical oxidation

of Au0 to Au+, and the reduction of oxygen. If both the anodic and cathodic half reactions are

considered, the system can most accurately be described by Equations 2.2 and 2.3, which

proceed in parallel:

2Au + 4CN! + O2 + 2H2O ! 2Au CN( )2! + H2O2 + 2OH! (2.2)

2Au + 4CN! + H2O2 ! 2Au(CN)2! + 2OH! (2.3)

Returning to Figure 2.2 with the cathodic and anodic reactions in mind, we can see that

the oxidant in the case of cyanidation would be oxygen, which is reduced to hydrogen peroxide

and the hydroxide ion, both of which will undergo mass transport away from the gold surface. In

terms of the anodic reaction, once gold is oxidized from Au0 to Au+, an aurocyanide complex is

formed, which also undergoes mass transport to the bulk solution from the gold-solution

interface1.

The gold dissolution rate in alkaline cyanide solutions is dependent on a wide number of

factors, including cyanide concentration, dissolved oxygen concentration, temperature, pH, and

surface area of exposed gold. Much research has gone into the optimization of the cyanidation

process over the last 100 years since its discovery and implementation. Based on calculated

Page 30: An Electrochemical and SERS Study of the Gold-Thiosulfate

14

diffusion coefficients for both cyanide and oxygen, as well as observed experimental and

practical values, the optimal molar ratio of CN-: O2 ranges from 4:1 to over 7:1. In practice,

ratios greater than 6:1 are used to ensure that the cyanide concentration is not the rate limiting

factor. 1

Temperature has a strong effect on the gold dissolution rate in aerated alkaline cyanide

leaching solutions. As the temperature is increased, due to increased diffusion rates of reacting

species, the gold dissolution rate also increases up to a maximum of ~85°C. Above this

temperature, oxygen solubility decreases and results in an overall decrease in the gold dissolution

rate1. The pH of the leaching solution is typically kept above 9.4 to avoid hydrolysis of cyanide

and excess consumption, however, in practice the pH conditions are dictated by other process

factors such as the ore composition, and solubility issues. 1

Exposed surface area is directly proportional to the gold dissolution rate. This is directly

related to particle size distribution and liberation characteristics of the ore. In general, with

decreasing particle size, the gold dissolution rate greatly increases, due to a larger amount of

exposed gold1.

Page 31: An Electrochemical and SERS Study of the Gold-Thiosulfate

15

2.2.2 Environmental Concerns

The presence of copper in ore poses a large problem during the cyanide leaching process,

not only because of the consumption of cyanide due to formation of copper cyanide complexes,

but also due to the increased toxicity of these complexes to birds, animals and fish 3, 9. Recent

concerns over environmental and human health issues involved with cyanide leaching as a result

of collapsed tailings dams in Guyana, and Romania has lead to increased interest in non-cyanide

lixiviants3, 7, 9, 10. One such alternative is thiosulfate, which is not only non-toxic, but also useful

for ores that consume excessive amounts of cyanide or absorb the resulting complex.9

Carbonaceous type ores are susceptible to thiosulfate leaching due to its ability to degrade sulfide

matrices and prevent preg-robbing10, 11. Jeffrey, Breuer and Choo12 compared the relative

leaching rates of chloride, cyanide and thiosulfate system for both pure gold and gold/silver

rotating electrochemical quartz-crystal microbalance (REQCM) electrodes. The authors found

that when comparing the initial rates of leaching, thiosulfate began with a much greater leaching

rate (3.8 × 10-5 mol m-2 s-1) than that of cyanide (0.1 × 10-5 mol m-2 s-1). Also, thiosulfate was

able to leach both pure gold and gold/silver alloy, whereas cyanide was only effective in leaching

of the gold/silver alloy.

2.2.3 Thiosulfate Alternative

Thiosulfate as a chemical has been widely used in both photography and in the

pharmaceutical industry7. It was first proposed for use in leaching of precious metals in the early

1900s in the Von Patera process, wherein gold and silver ores were subjected to a chloridising

roast, then leaching was performed using thiosulfate10. In the late 1970s a process for the

application of ammoniacal thiosulfate to gold ores containing copper and metal sulfides was

developed and patented for commercial use13.

Page 32: An Electrochemical and SERS Study of the Gold-Thiosulfate

16

2.3 THIOSULFATE LEACHING OF GOLD

Leaching of gold by thiosulfate proceeds through two coupled half reactions: oxidation

of gold (Equation 2.4) and reduction of oxygen (Equation 2.5) 9, 14, 15.

Au + 2S2O32! " Au(S2O3)2

3! + e! (2.4)

O2 + 2H2O + 4e! " 4OH! (2.5)

The overall reaction is reminiscent of the Elsner equation:

4Au + 8S2O32! + O2 + 2H2O " 4Au(S2O3)2

3! + 4OH! (2.6)

The only apparent change in the overall reaction is the lixiviant that acts to sequester the gold in

solution. The anodic reaction in Equation 2.4 has a standard reduction potential of 0.153 V,

while the reduction potential of the cathodic reaction (Equation 2.5) is 0.401 V14. The oxidant is

still O2, and the reaction is typically carried out in alkaline solution due to the acid catalyzed

decomposition of thiosulfate7, 9. Although the initial kinetics of the leaching reaction are rapid,

over extended periods of time the kinetics of this reaction are quite slow mainly due to oxidation

of thiosulfate by oxygen, leading to formation of species such as sulfite, and formation of a

passive layer from decomposition of the thiosulfate ion 6, 16.

The decomposition of thiosulfate under oxidizing conditions can proceed through a

variety of pathways 14, 15, 17:

2S2O32! " S4O6

2! + 2e! (2.7)

S2O32! + 6OH! " 2SO3

2! + 3H2O + 4e! (2.8)

S2O32! + 2OH! " SO4

2! + H2O + S2! (2.9)

Upon formation of these species, a number of additional reactions may occur in solution

including degradation and rearrangement of tetrathionate (S4O62! ) to form trithionateS3O6

2! ,

Page 33: An Electrochemical and SERS Study of the Gold-Thiosulfate

17

thiosulfate, pentathionate (S5O62! ), and sulfite (Equations 2.10-2.12).17 Subsequent degradation

of trithionate (Equation 2.13) leads to an even more complex and dynamic system. 18-20

S4O62! + S2O3

2! " S5O62! + SO3

2! (2.10)

S4O62! + SO3

2! " S3O62! + S2O3

2! (2.11)

2S4O62! " S3O6

2! + S5O62! (2.12)

S3O62! + H2O " S2O3

2! + SO42! + 2H+ (2.13)

Zhang et al. 17 studied the rearrangement and degradation reactions of both tetrathionate

and trithionate in near-neutral solutions. The presence of excess thiosulfate is known to catalyze

the rearrangement reactions of polythionates (Equation 2.15), which can be described by a

general disproportionation reaction 17:

2SxO62! " Sx!1O6

2! + Sx+1O62! (2.14)

2SxO62! + S2O3

2! H+

OH!" #""$ """ Sx+1O6

2! + SO32! (2.15)

It is important to note here, that unlike other polythionates, trithionate does not rearrange in

accordance with this disproportionation scheme, as dithionate cannot be formed by interaction of

polythionate species21. The direction of Equation 2.15 is highly dependent on the pH of the

solution. At pH > 7, the presence of the sulfite drives the reaction to the left. When the pH is less

than 7, disulfite drives the reaction to the right17. In alkaline solutions, higher order polythionates

show significantly decreased stability. The presence of polythionates higher than hexathionate

are not detected in solutions near neutral pH, or in slightly alkaline solutions17. Pentathionate,

and hexathionate themselves have been detected in solution, however, their degradation is much

more rapid than tetrathionate17.

2S5O62! + 6OH! " 5S2O3

2! + 3H2O (2.16)

Page 34: An Electrochemical and SERS Study of the Gold-Thiosulfate

18

S6O62! " S5O6

2! + S (2.17)

The presence of these fairly reactive species results in an extremely dynamic and complicated

system. Some, or all, of these species may be responsible for the passive layer on the electrode

surface, effectively inhibiting any further leaching of gold 22.

An investigation into the electro-oxidation of thiosulfate on gold by Pedraza et al. 23

showed that the passivating layer was composed of both elemental sulfur and some form of

oxidized sulfur species. Cyclic voltammetry in a solution of 0.01 M Na2S2O3 and 0.1 M Na2SO4

(the supporting electrolyte) displayed a prominent oxidation peak at 0.44 V versus a saturated

calomel electrode (SCE) that was later assigned to the formation of a sulfide film on the surface

of the gold electrode23. The adsorbed sulfide layer likely formed through the reactions described

by Equations 2.18-2.20: 23

S2O32! + 6OH! " 2SO3

2! + 3H2O + 4e! (2.18)

S2O32! + 2OH! " SO4

2! + H2O + S2! (2.19)

S2! " S + 2e! (2.20)

The second component of the film, oxidized sulfur species, arises not from electro-oxidation of

thiosulfate but from decomposition of thiosulfate at open circuit potential and subsequent

oxidation of these species23, as described by Equations 2.10-2.17.

The hypothesis of Pedraza and co-workers was verified by the experimental results of

Jeffrey et al. 24. Upon application of a triangular potential scan to a gold electrode in contact

with a solution of 0.1 M Na2S2O3, the authors noted an increase in the measured current,

beginning at approximately 0.4 V, with a peak at ~0.73 V. In parallel, a surface enhanced Raman

(SERS) spectrum was collected to track the expected νS-S and δS-S-S bands of elemental sulfur. As

the voltammetric current rose, evidence of peaks corresponding to the expected sulfur stretches

Page 35: An Electrochemical and SERS Study of the Gold-Thiosulfate

19

appeared in the collected spectrum, beginning at approximately 0.44 V and becoming more

intense as the potential was increased in the positive direction. The increasing intensity of these

bands was believed to correspond to the formation of some form of polymeric sulfur adsorbed on

the gold surface. 24

The presence of oxidation and decomposition products of thiosulfate on the gold surface

was also confirmed by Woods et al. 25 through the collection of a SERS spectrum of 0.1 M

Na2S2O3 at open circuit potential (Figure 2.3).

Figure 2.3: SERS spectrum of 0.1 M Na2S2O3 at open circuit potential. Modified from 25.

At open circuit potential, bands corresponding to the νS-S and νsym(S-O) vibrations of

thiosulfate are seen at 445 cm-1 and 999 cm-1. 25 The authors also noted the presence of bands at

378 cm-1 and 1033 cm-1, which are within 10 cm-1 and 7 cm-1 respectively of the νS-S and νsym(S-O)

vibrations of tetrathionate. Due to this difference in the band position, it was proposed that these

Page 36: An Electrochemical and SERS Study of the Gold-Thiosulfate

20

bands arose from a polythionate species adsorbed on the gold surface, with a longer chain length

than tetrathionate 25.

Recent work by J. Baron Gavidia proved that tetrathionate was present in the passive

layer forming on the gold surface 6. Time dependent SERS studies of a gold nanorod electrode in

a 0.1 M Na2S2O3 + 1.0 x 10-4 M NaOH solution were conducted at open circuit potential, the

results of which are shown in Figure 2.4. Using the Fourier self-deconvolution method (FSD),

bands at 258, 315, 378 and 1025 cm-1 were identified within the first five minutes of immersion.

These bands were assigned to tetrathionate in view of the fact that adsorption on to the gold

surface would shift the band positions from those typically seen in the solution spectra 6.

Figure 2.4: SERS spectrum of a gold nano-rod electrode submersed in 0.1M Na2S2O3 + 1x 10-4M NaOH

at open circuit potential (modified from 6).

After 630 minutes, the peaks corresponding to tetrathionate were no longer present, but

new bands at 265 and 420 cm-1 arose. These were attributed to the formation of trithionate on the

surface through decomposition of tetrathionate: 6, 19

2S4O62! + 3OH! " 5

2S2O3

2! + S3O62! + 3

2H2O (2.21)

Page 37: An Electrochemical and SERS Study of the Gold-Thiosulfate

21

2.3.1 Ammonia-Thiosulfate Leaching

The addition of ammonia into thiosulfate leaching solution results in the formation of an

anionic aurocomplex that is stable over a wide pH range. Moreover, ammonia helps to inhibit

interference from other species such as iron oxides, silica, silicates and carbonates, and increases

the rate of gold oxidation 16, 26. This was demonstrated in the work of Breuer and Jeffrey 27.

These authors studied the effect of the concentration of ammonia on gold dissolution rate by use

of a rotating electrochemical quartz crystal microbalance. The results distinctly showed that with

increasing ammonia concentration, the calculated current, and thus the dissolution rate, was

greatly increased (Figure 2.5) 27.

Figure 2.5: Calculated voltammetric current for varying concentrations of ammonia in thiosulfate

leaching solution. Adapted from 27.

The role of ammonia is still not fully understood, but it is believed that it prevents

passivation of the gold by forming a complex with Au+, and subsequently assisting in gold

dissolution13. However, it has been shown that in the absence of thiosulfate, no dissolution of

Page 38: An Electrochemical and SERS Study of the Gold-Thiosulfate

22

gold into the solution occurs in the presence of ammonia. Breuer and Jeffrey27 proposed that

rather than just ammonia and gold, it is a gold-amine-thiosulfate complex that affects passivation

of the gold surface during the leaching reaction.

Ammonia’s proposed catalytic role is defined by an adsorption/desorption/stabilization

process28. This mechanism corroborates the assumption of Breuer and Jeffrey that it is a gold-

amine-thiosulfate complex that is responsible for enhanced leaching27. The process begins with

the adsorption of ammonia and thiosulfate to the gold surface28:

a) Au(s ) + S2O3

2! + NH3 " Au(S2O3)(NH3)(ads )2!

b) Au(s ) + NH4S2O3! + NH3 " Au(S2O3NH4 )(NH3)(ads )

! (2.22)

Oxidation of the adsorbed species then occurs28:

a) Au(S2O3)(NH3)(ads )

2! " Au(S2O3)(NH3)(ads/aq)! + e!

b) Au(S2O3NH4 )(NH3)(ads )! " Au(S2O3NH4 )(NH3)(ads/aq)

0 + e! (2.23)

Followed by desorption of the now oxidized species28:

a) Au(S2O3)(NH3)(ads/aq)

! + S2O32! " Au(S2O3)2 (aq)

3! + NH3

b) Au(S2O3NH4 )(NH3)(ads/aq)0 + NH4S2O3

! " Au(S2O3)2 (aq)3! + NH3 + 2NH4

+ (2.24)

This mechanism was supported by experimental evidence showing that the rate of

oxidation of gold in ammoniacal thiosulfate solutions was highly dependent on the concentration

product ([M2S2O3][NH3])0.8, where M = Na, NH4+. Also, the concentration of the NH4S2O3

- ion

pair was equal to, or greater than, the concentration of free thiosulfate in solution. 28

Ammonia has also been shown to help in regeneration of thiosulfate from leaching

solutions through ammonolysis of degradation compounds such as trithionate:

a) S3O62! + NH3 ! S3O6 "NH3

2!

b) S3O6 "NH32! # S2O3

2! + SO3NH2! + H+

(2.25)

Page 39: An Electrochemical and SERS Study of the Gold-Thiosulfate

23

Addition of copper is required in ammonia solutions due to the inability of O2 to act as

the oxidant 16. However, copper is also known to affect gold dissolution without ammonia.

Zhang and Nicol 29 investigated the effect of several different forms, and concentrations of

copper on the gold dissolution reaction (Figure 2.6). The authors found that addition of any form

of copper increased the amount of gold dissolved, but the addition of 0.5 mM CuSO4 resulted in

the highest amount of gold recovered. Based on the work of Rabai and Epstein 30, Zhang and

Nicol 29 proposed that in the absence of ammonia, the oxidation of gold is accomplished through

a copper-thiosulfate-oxygen intermediate, [Cu(S2O3)3O2]5-.

Figure 2.6: The effect of copper on gold dissolution in thiosulfate leaching. Adapted from29.

The intermediate has been observed previously using UV-VIS, but not at the gold-

solution interface. Another suggested role of copper in aiding gold dissolution was the

scavenging of sulfide from the surface of gold, resulting in at least partial removal of the passive

layer. 29

Page 40: An Electrochemical and SERS Study of the Gold-Thiosulfate

24

One of the best-known methods to limit the formation of the passive layer, and ensure

relatively high gold oxidation rates, is the addition of both ammonia and copper into the solution,

where Cu2+ acts as the oxidant 13, 16, 26, 31. The presence of both of Cu2+ and NH3 is known to

increase the rate of gold dissolution31. Once in solution, Cu2+ and NH3 combine to

form Cu(NH3) 42+, which is believed to act as the oxidant 9, 13, 16, 26, 32. The overall reaction for

the dissolution of gold then becomes13, 15, 26, 33:

Au + 5S2O32! + Cu(NH3)4

2+ " Au(S2O3)23! + 4NH3 + Cu(S2O3)3

5! (2.26)

The reduction potential of the cupric tetra-ammine complex given by Watling 34 is +0.22V, and

can be described by19, 22, 33, 34:

Cu(NH3)42+ + 3S2O3

2! + e! " Cu(S2O3)35! + 4NH3 (2.27)

The reduction of Cu2+ in solution is coupled with the corresponding oxidation of gold, as seen in

Equation 2.4. Once generated, Cu+ can be oxidized back to Cu2+ through coupling to the

reduction of O2 13, 22, 33:

2Cu(S2O3)35! + 8NH3 + 1

2O2 + H2O " 2Cu(NH3)4

2+ + 2OH! + 6S2O32! (2.28)

The proposed mechanism for the dissolution of gold in thiosulfate media containing both

ammonia and copper involves adsorption and desorption of species involved in both the cathodic

and anodic reactions, with a rate-determining redox reaction26, 31. The electrochemical

mechanism is illustrated in Figure 2.7.

Addition of ammonia and copper into thiosulfate leaching solution is known to result in

an increased initial leaching rate, and increased gold recovery over the duration of leaching;

however, the mechanism and exact effect of each additive on the passive layer is still unknown.

Page 41: An Electrochemical and SERS Study of the Gold-Thiosulfate

25

Figure 2.7: Electrochemical model for the leaching of gold using ammoniacal thiosulfate leach media. Adapted from 26, 31.

2.3.2 The Role of the Electrolyte on Ammonia-Thiosulfate Leaching

The role of the metal cation of the thiosulfate salt was investigated by Chandra et al. 35,

using a polycrystalline solid gold electrode in combination with an REQCM. The authors

measured the gold oxidation polarization curves of the electrode in the presence of ammonium,

sodium and potassium thiosulfate. A significant increase in the current was noted as the metal

cation was changed from sodium to potassium. Two possible explanations were proposed:

i. Heavier cations (such as potassium) would be less hydrated compared to ions with

smaller atomic mass (such as sodium) and therefore would be more likely to be adsorbed

onto the gold surface. 35

Page 42: An Electrochemical and SERS Study of the Gold-Thiosulfate

26

ii. The dissociation constants for alkali metals with thiosulfate are known to decrease in the

order of Na > K > Rb > Cs, therefore, one would expect potassium ions to more readily

form an ion pair with either gold-thiosulfate, or free thiosulfate.35

However, no explanation yet exists for how either of these occurrences would influence the gold

oxidation reaction. 35

References

1. Marsden, J.; House, I. In The chemistry of gold extraction; Society for Mining Metallurgy: 2006; .

2. Kongolo, K.; Mwema, M. Hyperfine interactions 1998, 111, 281-289. 3. Gupta, C.; Mukherjee, T. In Hydrometallurgy in extraction processes; CRC:

1990; Vol. 2. 4. Feng, D.; van Deventer, J. S. J. Int. J. Miner. Process. 2010, 94, 28-34. 5. Gupta, C.; Mukherjee, T. Boca Raton 1990, , 127-165. 6. Baron Gavidia, J. Study of the Gold-Thiosulfate Interface Under Leaching

Conditions, University of Guelph, Guelph, ON, 2010. 7. Hilson, G.; Monhemius, A. J. J. Clean. Prod. 2006, 14, 1158-1167. 8. Jeffrey, M.; Breuer, P. Minerals Eng 2000, 13, 1097-1106. 9. Jeffrey, M. I.; Linda, L.; Breuer, P. L.; Chu, C. K. Minerals Eng 2002, 15, 1173-

1180. 10. Grosse, A. C.; Dicinoski, G. W.; Shaw, M. J.; Haddad, P. R. Hydrometallurgy

2003, 69, 1-21. 11. Feng, D.; van Deventer, J. S. J. Minerals Eng 2010, 23, 143-150. 12. Jeffrey, M.; Breuer, P.; Choo, W. L. Metallurgical and Materials Transactions B

2001, 32, 979-986. 13. Aylmore, M. G.; Muir, D. M. Minerals Eng 2001, 14, 135-174. 14. Zhang, S. C.; Nicol, M. J. J. Appl. Electrochem. 2003, 33, 767-775. 15. Wan, R. Y.; LeVier, K. M. Int. J. Miner. Process. 2003, 72, 311-322. 16. Abbruzzese, C.; Fornari, P.; Massidda, R.; Vegliņ, F.; Ubaldini, S.

Hydrometallurgy 1995, 39, 265-276. 17. Zhang, H.; Jeffrey, M. I. Inorg. Chem. 2010, . 18. Breuer, P. L.; Jeffrey, M. I. In A review of the chemistry, electrochemistry and

kinetics of the gold thiosulfate leaching process; Young, C., Alfantazi, A., Anderson, C., James, A., Dreisinger, D. and Harris, B., Eds.; 2003; , pp 154.

19. Byerley, J. J.; Fouda, S. A.; Rempel, G. L. J.Chem.Soc., Dalton Trans. 1973, , 889-893.

20. Naito, K.; Hayata, H.; Mochizuki, M. Journal of Inorganic and Nuclear Chemistry 1975, 37, 1453-1457.

Page 43: An Electrochemical and SERS Study of the Gold-Thiosulfate

27

21. Nickless, G. In Inorganic sulphur chemistry; Elsevier Publishing Company: 1968; .

22. Chu, C. K.; Breuer, P. L.; Jeffrey, M. I. Minerals Eng 2003, 16, 265-271. 23. Pedraza, A. M.; Villegas, I.; Freund, P. L.; Chornik, B. J Electroanal Chem 1988,

250, 443-449. 24. Jeffrey, M.; Watling, K.; Hope, G. A.; Woods, R. Minerals Eng 2008, 21, 443-

452. 25. Woods, R.; Hope, G. A.; Watling, K. M.; Jeffrey, M. I. J. Electrochem. Soc. 2006,

153, D105-D113. 26. Aylmore, M. G.; Muir, D. M. Miner Metall Process 2001, 18, 221-227. 27. Breuer, P. L.; Jeffrey, M. I. Hydrometallurgy 2002, 65, 145-157. 28. Senanayake, G. Hydrometallurgy 2005, 77, 287-293. 29. Zhang, S. C.; Nicol, M. J. J. Appl. Electrochem. 2005, 35, 339-345. 30. Rabai, G.; Epstein, I. R. Inorg. Chem. 1992, 31, 3239-3242. 31. Senanayake, G. J. Colloid Interface Sci. 2005, 286, 253-257. 32. Molleman, E.; Dreisinger, D. Hydrometallurgy 2002, 66, 1-21. 33. Arima, H.; Fujita, T.; Yen, W. Miner Metall Process 2003, 20, 81-92. 34. Watling, K. M. Spectroelectrochemical Studies of Surface Species in the

Gold/Thiosulfate System, Griffith University, Australia, 2007. 35. Chandra, I.; Jeffrey, M. I. Hydrometallurgy 2004, 73, 305-312.

Page 44: An Electrochemical and SERS Study of the Gold-Thiosulfate

28

CHAPTER 3: EXPERIMENTAL TECHNIQUES THEORY

3.1 ELECTROCHEMICAL TECHNIQUES

3.1.1 Sweep Voltammetry Methods

Sweep voltammetry methods can be invaluable for determination of the electrochemical

behavior of a system. A potential varying linearly with time is applied to the system, while the

output current is simultaneously recorded as a function of the applied potential. Figure 3.1a

shows a typical potential-time curve for cyclic voltammetry, and 3.1b the current response curve.

E1

E3

E2

t

E (-)

E0’ E!

A + e- A!-

A + e- A!-

a) b)

Figure 3.1: a) Typical potential time curve for cyclic voltammetry b) Current response curve. Adapted

from Bard and Faulkner1.

Sweep rate is defined as the slope of the potential applied as a function of time, and can

range from 1mV/s to 1000V/s. Starting from potential E1, a voltage ramp is applied to a higher

value, E21. At this point the potential is reversed and cycled back to either the same potential

(E1), or to a different potential, E3, where the sweep ends.

Page 45: An Electrochemical and SERS Study of the Gold-Thiosulfate

29

In an oxidation reaction or reduction reaction occurring at the surface of an electrode, the

electron transfer across the metal-solution interface produces a current that is defined as Faradaic

current. However, non-Faradaic processes such as charging of the capacitor at the interface may

also occur, resulting in current flow or electrode polarization without a net charge transfer across

the metal electrolyte-interface.

The metal-solution interface behaves as a capacitor. The solution side of the metal

electrolyte interface is composed of several layers, containing all charged species and dipoles.

The simplest model of this region is called the electrical double layer (Figure 3.2). It consists of

the inner and diffuse parts. The inner layer is a region between the metal surface and the outer

Helmholtz plane (OHP) that corresponds to a distance x2. This distance is the point of nearest

approach to the electrode surface, for solvated ions. The inner layer includes solvent molecules

and specifically adsorbed ions. The locations of the specifically adsorbed anions define the

position of the inner Helmholtz plane (IHP), at a distance x1. Beyond the OHP, and extending

into the bulk, is the diffuse layer. Here, solvated ions interact with the charged metal surface only

through long-range electrostatic interactions. Ions in the diffuse layer are treated as point

charges, meaning that their properties are essentially independent of the chemical properties of

the ion.

Current generated through electron transfer of redox reactions follows Faraday’s law:

the amount of a chemical consumed or produced in the electrode reaction is proportional to the

charge that has passed through the interface2. If we consider the reaction

O+ ne! " R (3.1)

Page 46: An Electrochemical and SERS Study of the Gold-Thiosulfate

30

Figure 3.2: Schematic of the structure of the electrical double layer formed at the metal-electrolyte

interface. Adapted from Bard and Faulkner1.

Where O is the oxidized species and R is the reduced species, the rates of the forward

reaction (reduction) and the reverse reaction (oxidation) are given by:1

! f = k fCO = icnFA

(3.2)

!b = kbCR =ianFA

(3.3)

Where νf is the rate of reduction; νb the rate of oxidation; CO is the surface concentration of the

oxidized species; CR is the surface concentration of the reduced species; n is the number of

electrons transferred in the reaction; A is the area of the electrode and F is Faraday’s constant (96

IHP OHP

Metal

+

+

+

+

-

-

Solvent molecule

x1 x2

Page 47: An Electrochemical and SERS Study of the Gold-Thiosulfate

31

485.3399 C mol-1). The rate constants of the forward (kf) and reverse reaction (kb) are potential

dependent, as given by Equations 3.4 and 3.5: 1

k f = k0 exp !"nF

RTE ! E0( )#

$%&'( (3.4)

kb = k0 exp

1!"( )nFRT

E ! E0( )#$%

&'(

(3.5)

where α, k0, and E0 are the transfer coefficient, standard rate constant and standard potential,

respectively. The standard rate constant is a measure of the forward or reverse reaction rates

when the system is at equilibrium at standard conditions. Its value indicates the kinetics of a

redox couple; a system with a large standard rate constant will rapidly achieve equilibrium,

whereas a system with a small k0 would be sluggish.

Measured current is the sum of the cathodic and anodic currents:

i = ic ! ia = nFA kfCO (0,t)! kbCR(0,t)"# $% (3.6)

Therefore, by combining the expressions for the rate constants and the potential dependence of

the two reactions, an equation for the current density at any potential can be derived1.

iA= nFk0CO exp

!"nFRT

E ! E0( )#$%

&'(! nFk0CR exp

1!"( )nFRT

E ! E0( )#$%

&'(

(3.7)

If we consider the interface at equilibrium with the solution, in which the bulk concentration of

oxidized species is equal to the surface concentration, and the surface concentration of the

reduced species is equal to the bulk, it follows that E = Eeq , k fCO = kbCR , and k f = kb = k0 . By

definition, the net current at this point is zero, and the electrode adopts a potential based purely

on the bulk concentrations of the oxidized and reduced species. However, this does not mean that

charge transfer across the interface has stopped. The magnitude of the anodic and cathodic

currents at equilibrium are equivalent, such that:

Page 48: An Electrochemical and SERS Study of the Gold-Thiosulfate

32

nFAk0CO exp!"nF( )RT

(Eeq ! E0 )#

$%&'(= nFAk0CR exp

1!"( )nFRT

(Eeq ! E0 )#

$%&'(

(3.8)

and are defined by the exchange current:

i0 = FAk0CO exp

!"nFRT

Eeq ! E0( )#

$%&'(

(3.9)

Dividing both sides of Equation 3.7 by the exchange current (Equation 3.9), and rearranging for

i, we achieve:

i = nFAk0CO exp!"nFRT

Eeq ! E0( )#

$%&'(exp !"nF

RTEeq ! E( )#

$%&'(! exp

1!"( )nFRT

Eeq ! E( )#$%

&'(

)*+

,-.

(3.10)

The first term in Equation 3.10 is the exchange current as seen in Equation 3.9, and using the

definition of overpotential (! = Eeq " E ), Equation 3.10 can be simplified to:

i = i0 exp!"nFRT

#$%&

'() ! exp

1!"( )nFRT

#$%&

'()

*

+,

-

./ (3.11)

Equation 3.11 is known as the Butler-Volmer equation, and is a description of the current when

the system is out of equilibrium, in the absence of mass transfer effects; that is, when i is < 10%

of the limiting currents. When the overpotential is small, the exponential terms can be estimated

as ex = 1+ x and e! x = 1! x . Using this approximation, Equation 3.11 can then be written as

follows1:

i = !i0nFRT

" (3.12)

The net current is linearly proportional to the overpotential in a narrow range near the

equilibrium potential. The negative reciprocal of the slope in this region has units of resistance,

and is often termed the charge transfer resistance:

Page 49: An Electrochemical and SERS Study of the Gold-Thiosulfate

33

! "#"i

= Rct =RTi0nF

(3.13)

A plot of log(i) vs. η yields a Tafel plot, a common data form used for analysis of the kinetic

aspects of a redox system. Figure 3.3 displays an example of a Tafel plot for a typical redox

system.

Figure 3.3: Tafel plot of a redox system, modified from Bard and Faulkner.1

When the overpotential is greater than ~60 mV on either the cathodic or anodic side, one

of these contributions becomes negligible. For example, at large negative overpotentials, the

cathodic contribution dominates, i.e. exp !"nFRT

#$%

&'(>> exp (1!" )nF

RT#$%

&'(

, and the anodic term in

Equation 3.11 can be ignored, meaning the Butler-Volmer equation can now be represented as:

i = i0 exp !"nFRT

#$%

&'(

or log i = log i0 !"nFRT

(3.14)

The exchange current can also be determined graphically, as shown in Figure 3.3, by

interpolation of the linear regions at large overpotential, to zero overpotential. The intersection of

the two linear regions gives the log of the exchange current.

-50 -100 -150 -200 50 100 150 200

slope = !"nF2.3RT

slope =1!"( )nF2.3RT

log i0

log |i|

-3.5

-4.5

-5.5

-6.5

!, mV

Page 50: An Electrochemical and SERS Study of the Gold-Thiosulfate

34

The values of the transfer coefficients can be calculated using the slope of the linear

regions of the Tafel plot. As in Figure 3.3, the slopes of the linear segments at large overpotential

are given by 1!"( )nF2.3RT

and !"nF2.3RT

for the anodic and cathodic branches respectively. From

these expressions, one can calculate the values of the transfer coefficients.

The transfer coefficient is a measure of the symmetry of the intersection of the free

energy curves versus the reaction coordinate for the products and reactants. The reaction

coordinate is variable, for example it could be the torsional angle around a bond, but in

electrochemistry, it is typically considered as the distance of the oxidized species from the

interface. Consider again a one-electron, one-step process in which an oxidized species, O, reacts

with a single electron to form the reduced species, R. The free energy curves of O + e and R are

represented in Figure 3.4. Using the equilibrium potential, Eeq, as a reference point, an applied

potential will shift the O + e energy curve either up or down according toF E ! Eeq( ) . Figure 3.4

shows a positive energy change. Consequently, the activation energy for the oxidation of the

reduced species, R, will also shift by a fraction of the total energy change, 1!!( )F E ! Eeq( ) .

The shift in the activation energy is therefore dependent on the value of the transfer coefficient,

which can range in value from zero to unity depending on the symmetry of the potential energy

curve intersection.

Page 51: An Electrochemical and SERS Study of the Gold-Thiosulfate

35

Figure 3.4: Diagram of the free energy curves of an oxidized and reduced species when a potential is

applied, shifting the system from equilibrium. On the right is a zoom in of the intersection of the two

potential energy surfaces. Adapted from Bard and Faulkner. 1

If the area of intersection is considered locally linear (Figure 3.5), a definition of the

transfer coefficient can be generated in terms of the geometry at that point.

Figure 3.5: Intersection of the locally linear potential energy surface of an oxidized and reduced species.

Adapted from Bard and Faulkner.1

!

! !

!FE

1!"( )FE

R

O + eE = E

E = 0

Length = x

Reaction Coordinate

Page 52: An Electrochemical and SERS Study of the Gold-Thiosulfate

36

Using the definition of the angles ϕ and θ:

tan! = "FEx

(3.15)

tan! =1"#( )FE

x (3.16)

The transfer coefficient can be defined as:

! = tan"tan# + tan"

(3.17)

A symmetric intersection of the two potential energy surfaces leads to the angles ϕ and θ being

equivalent, hence α is equal to ½. If the intersection is asymmetric, the value of α will either lie

between 0 and ½ or ½ and 1.

3.1.2 Mixed Potential Theory

The measured current is the sum of both the anodic (oxidation) and cathodic (reduction)

reactions. In the case of gold leaching, this current is the sum of the anodic current generated

from the oxidation of gold, and the cathodic current is generated by the reduction of oxygen or

Cu(II). Figure 3.6 is a representation of typical linear sweep voltammogram of a Au leaching

system. The point at which the curve crosses the x-axis corresponds to a value of zero for the

current. Here, the current contribution from gold oxidation is equivalent in magnitude, but

opposite in sign to the contribution of the reduction of oxygen. The potential at this junction is

defined as the mixed potential. The magnitude of either the cathodic or anodic current at the

mixed potential is known as the leaching current, iM. Using a Tafel-like plot, that is, the

logarithm of the current versus the overpotential, the mixed potential can be easily identified

through extrapolation of the respective currents and their intercept. This is displayed in Figure

3.7.

Page 53: An Electrochemical and SERS Study of the Gold-Thiosulfate

37

Figure 3.6: Generic linear sweep voltammogram. Dashed lines represent the individual contribution of

the cathodic and anodic currents, while the solid line represents the sum of these two contributions.

Figure 3.7: Tafel plot a gold electrode in contact with a 0.1 M Na2S2O3 + 0.01 M CuSO4 + 1.1µΜ

Ca(OH)2 solution.

Electrode potential

Cur

rent

den

sity

(EM, iM)

-0.26 -0.24 -0.22 -0.20 -0.18 -0.16 -0.14 -0.12 -0.10-2.5

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

Log

(i)

Potential / V vs. SCE

Mixed Potential

Page 54: An Electrochemical and SERS Study of the Gold-Thiosulfate

38

A modified version of the Butler-Volmer equation devised by J. Y. Baron3 can be used to

determine the leaching current. In the case of gold leaching, the exchange current (i0) in the

above equation is equivalent to the leaching current (iM). Because thiosulfate leaching of gold

involves different electron transfer reactions in the cathodic and the anodic processes, the charge

transfer coefficients are different: for copper and gold they are αCu2+ and αAu, respectively. Using

these transfer coefficients, the modified Butler-Volmer equation is:

i = iM exp!"

Cu2+nCu2+F

RT#

$%&

'()! exp

1!"Au( )nAuFRT

#$%&

'()

*

+,

-

./ (3.18)

Because the reaction involves oxidation of gold, and reduction of Cu(II) to Cu(I), the number of

electrons transferred, i.e. n, is equal to 1. When overpotentials are small, the equation can be

simplified to:

i = iM !"Cu2+

! 1!"Au( )( ) #FRT (3.19)

In some instances, the transfer coefficients can be approximated to 0.5, thus allowing a

simplification of the equation. 1 However, in the instance of gold leaching in the presence of Cu,

the transfer coefficients must be calculated as described previously, before the equation can be

further simplified.

By plotting the current versus overpotential in a small potential range (±20mV) on both

sides of the mixed potential, the leaching current can be obtained. A linear fit on the resulting

plot (as show in Figure 3.8) gives the slope, equivalent to the ratio of the current to potential.

Page 55: An Electrochemical and SERS Study of the Gold-Thiosulfate

39

Figure 3.8: Plot of the current versus the overpotential in a ~40mV range around the mixed potential for

a 0.1 M Na2S2O3 + 0.01 M CuSO4 + 1.1µΜ Ca(OH)2 system.

Combining equations 3.13 and 3.19, and using the calculated slope, the leaching current can be

found:

Rct = ! "#"i

= 1slope

= RTiMF !$

Cu2+ ! 1!$Au( )( )% iM =

slope RT( )F !$

Cu2+ ! 1!$Au( )( ) (3.20)

This method of calculating the leaching current, devised by J. Y. Baron3, will be used to interpret

data obtained from linear and cyclic sweep voltammetry experiments.

3.1.3 Hydrodynamic Methods

Hydrodynamic methods are those involving forced convection of the system, where

either the electrode itself is in motion (such as with a rotating disc electrode), or the solution is

-0.18 -0.17 -0.16 -0.15 -0.14

-1.5

-1.0

-0.5

0.0

0.5

1.0

j / µ

A c

m-2

Potential / V vs. SCE

Δi/Δη

Page 56: An Electrochemical and SERS Study of the Gold-Thiosulfate

40

forced past a stationary electrode. Techniques using hydrodynamic methods are useful because

the system reaches a steady state fairly rapidly, and effects from double-layer charging are

eliminated, while the contribution from mass transport on the electron transfer reaction is

minimized.

Two types of fluid-flow are typically considered. If the solution can be characterized by

smooth and steady flow in planar layers, the flow is said to be laminar (Figure 3.9a). Flow at the

edges will have the lowest velocity due to friction, while solution at the centre will have the

greatest velocity. Hence a velocity profile for a laminar solution will be parabolic in shape.

Figure 3.9: a) Profile for a solution with laminar flow b) Profile for a solution with turbulent flow.

Adapted from Bard and Faulkner.4

Turbulent flow is characterized by chaotic currents without well-defined layers, only an average

net flow in a particular direction.

The type of flow can be determined through calculation of a factor known as the

Reynold’s number. The Reynold’s number is a dimensionless measure of the ratio of inertial

forces of a fluid, to the viscous forces of the fluid, and is defined by Equation 3.21:

Re = !chlv

(3.21)

Laminar Flow Turbulent Flow

a b

Page 57: An Electrochemical and SERS Study of the Gold-Thiosulfate

41

Where υch is the characteristic velocity of the fluid, l the characteristic length, and ν the

kinematic viscosity. Although this number is dimensionless, the value is proportional to fluid

velocity, thus a high Reynold’s number implies either high flow rate or electrode rotation rates.

Determination of either laminar, or turbulent flow is based on a critical number, Recr. Below this

value, flow is considered laminar, while above flow is considered turbulent4. Typical values for

Recr lie around 105-106.5

Rotating disk electrodes are the most commonly used electrodes for hydrodynamic

methods. Typically they consist of a round metal disk inserted into a sheath of Teflon, or inert

material. The disk is connected to a motor, which allows rotation at the desired rate.

Three types of reactions are typically encountered during electrochemical hydrodynamic

experiments: those which are mass transport controlled, those which are under both kinetic and

mass transport control, and those which fall only under kinetic control. Reactions under only

kinetic control were discussed previously in terms of the Butler-Volmer model, and thus the

following discussion will only concern systems under mass transport control, or mixed control.

Limiting current for a mass transport controlled reaction is given by the Levich equation:

il ,c = 0.62nFADO2/3! 1/2" #1/6CO

$ (3.22)

Where n is the number of electrons involved in the redox reaction, D is the diffusion coefficient

of the electroactive species, ν is the kinematic viscosity, CO*

is the bulk concentration of the

oxidized species, ω is the angular velocity in s-1, F is the Faraday constant and A is the area of

the electrode. The Levich equation also dictates that the limiting current for the reaction is

proportional to the square root of the angular velocity. Thus a plot of i vs ω1/2 should provide a

straight line with an intercept of zero, if the reaction is indeed under full mass transport control.

Page 58: An Electrochemical and SERS Study of the Gold-Thiosulfate

42

Deviations from linearity in the above plot are indicative of the system moving from

under pure mass transport control, to that of mixed control. A system whose reaction rate is

limited by both mass transport and kinetic limitations can be described by the Koutecký-Levich

equation:

1i= 1iK

+ 1il ,c

= 1iK

+ 10.62nFADO

2/3! 1/2" #1/6CO$ (3.23)

Where iK is the current in the absence of mass transport effects, i.e. the current that flows only

under a kinetic limitation4.

3.2 SPECTROSCOPIC TECHNIQUES

3.2.1 Surface Enhanced Raman Spectroscopy

Raman spectra arise from the irradiation of a sample with photons in the visible or UV

regions of the electromagnetic spectrum. Scattered radiation may be collected parallel to, or at a

90° angle relative to the incident radiation6, 7. Interaction of the source light with the

polarizability of the molecule is responsible for the generation of an induced dipole moment.

This induced moment results in an excitation that is un-quantized, meaning that the molecule can

be in any one of an infinite number of virtual states between the ground electronic, and first

excited electronic state6.

If the collision between the photon and molecule is elastic, the escaping radiation is of the

same energy as the incident radiation and is defined as Rayleigh scattering. Elastic collisions

occur with the highest probability, thus the Rayleigh band is the central and most intense band in

the Raman spectrum. However, not all collisions are elastic, and thus radiation can be emitted at

either a higher or lower energy than the Rayleigh band6, 7. This is called Raman scattering. The

Page 59: An Electrochemical and SERS Study of the Gold-Thiosulfate

43

energy lost or gained in the radiation corresponds to the energy difference between the ground

and first excited vibrational energy of the molecule6.

Raman scattering can be explained by consideration of the behavior of a molecule in the

presence of an external electric field (in this instance, produced by the incident radiation):

EF = E0 cos 2!"ext( ) (3.24)

Where νex is the frequency of the exciting light, and E0 is the amplitude of the wave6, 7.

Interaction of the molecule’s electron cloud with the electric field of the incident radiation results

in the formation of an induced dipole moment. The magnitude of the dipole moment is

proportional to the molecule’s polarizability (αP):

µ =!PEF =!PE0 cos 2!"ext( ) (3.25)

In order to be classified as Raman active, the polarizability must vary as a function of the

internuclear distance (r):

!P =!0 + r ! req( ) "!P

"r#$%

&'( (3.26)

Where α0 is the polarizability at the equilibrium position req. The internuclear distance r ! req( ) varies with the frequency of vibration of the molecule, νv:

r ! req( ) = rm cos 2"#vt( ) (3.27)

In the above equation, rm is the maximum internuclear separation relative to the equilibrium

position. Substituting Equation 3.27 into Equation 3.26 we obtain:

!P =!0 +!!P

!r"#$

%&' rm cos 2!"vt( ) (3.28)

The induced dipole moment can then be defined by:

Page 60: An Electrochemical and SERS Study of the Gold-Thiosulfate

44

µ =!0E0 cos 2!"ext( )+ E0rm !!P

!r"#$

%&' cos 2!"vt( )cos 2!"ext( ) (3.29)

If we apply the trigonometric identity: cos xcos y = cos x + y( ) + cos x ! y( )"# $% 2 , the previous

equation for the induced dipole moment simplifies to:

µ =!0E0 cos 2!"ext( )+ E0rm2

!!P

!r"#$

%&' cos 2! "ex +!v[ ]t( )+ E0rm2

!!P

!r"#$

%&' cos 2! "ex (!v[ ]t( ) (3.30)

The three terms in the above equation represent the two types of scattering observed; the first

term represents oscillation of the dipole at the same frequency as that of the incident radiation,

resulting in Rayleigh scattering.This means that the emitted radiation also matches the frequency

of the incident. Rayleigh scattering accounts for the largest percentage of scattering that occurs.

The second and third terms in Equation 3.30 describe a modulation of the incident frequency by

the vibrational frequency of the bond, resulting in Raman scattering 6. The term containing

!ex "!v( ) represents what is known as Stokes lines, and that containing the term !ex +!v( ) , anti-

Stokes lines. Due to absorption of energy by the molecule, Stokes lines are found at lower

wavenumber than the Rayleigh line, and anti-Stokes lines, due to emission of energy by the

molecule, are found at higher wavenumber than the Rayleigh line 6. Figure 3.10 displays each of

the types of emission seen in Raman spectroscopy, and their origin.

Page 61: An Electrochemical and SERS Study of the Gold-Thiosulfate

45

Figure 3.10: Transitions seen in Raman spectroscopy. Adapted from6

Although useful to obtain fingerprint identification of molecules, a major limitation of Raman

spectroscopy is its low signal intensity. At most, the intensity of the scattered radiation is 0.001%

of the incident6.

Fleischmann et al. first discovered the benefits of surface enhanced Raman spectroscopy

in 1974 when a large increase in the signal strength of pyridine adsorbed on a roughened silver

electrode was observed 8, 9. At the time, this was attributed to the increase in surface area;

however, Albrecht and Creighton10, and independently Van Duyne and Jeanmaire11, realized that

the increased surface area was too small to cause enhancement over a factor of ~106 for the

predicted scattering cross section for pyridine12, thereby accrediting the extra increase to a

surface enhancement.9

Rayleigh Scattering Raman Scattering

Stokes Shift anti-Stokes Shift

E = h! E = h! + "E

!E

!S = !ex "!v!R = !ex !aS = !ex +!v

0 1 2 3

Page 62: An Electrochemical and SERS Study of the Gold-Thiosulfate

46

The greatest surface enhancement occurs on roughened metal surfaces of silver, gold,

copper and other coinage metals12; however, enhancement does still occur on metals such as

platinum, aluminum, rhodium and indium9, 12. It is believed that surface enhancement is a

combination of two contributions: an electromagnetic enhancement, and a chemical

enhancement13.

Chemical enhancement is believed to arise from the chemisorption of the adsorbate onto

the metal surface. When a molecule adsorbs onto a metal surface, it undergoes an electronic

rearrangement that allows enhancements for different vibrational modes, even in the ground

electronic state12. Upon chemisorption, it is proposed that a charge-transfer complex is formed

between the metal atoms on a rough surface and the molecule. Because of complex formation,

the energies of the HOMO and LUMO of the adsorbed molecule are close in energy to the Fermi

level of the metal. If the energy difference between the frontier orbitals and the Fermi level is

close to the same frequency of the exciting radiation, resonant enhancement may occur14. A

photon-driven charge transfer can also occur to increase enhancement, as shown in Figure 3.11.

Here, an adsorbed molecule is irradiated with light leading to the formation of an electron hole

pair. Electron tunneling occurs from the metal to the molecule, where vibrational information is

imparted to the electron, if the vibrational frequency matches the residence time within the

molecule. The electron then recombines with the hole, and a Raman photon is emitted with a

frequency modulated by the energy of the vibrational levels of the molecule14.

Page 63: An Electrochemical and SERS Study of the Gold-Thiosulfate

47

Figure 3.11: Photon-driven charge-transfer process. Adapted from14.

There is some debate in literature as to the importance of the chemical contribution to the

enhancement observed in SERS15, as electromagnetic theory can describe the major features,

including the exceptional intensity for alkali and coinage metals, a dependence on the nano-

structure of the system, the dependence on the aggregation of nanoparticles in the system, and

the polarization dependence16. The electromagnetic contribution to the enhancement arises from

a highly localized concentration of surface plasmons in metal surfaces with nano-sized surface

features17, 18. Plasmons are defined as the collective oscillations of the band of conduction

electrons against the background of the ionic metal cores16. When surface features of the metal

are smaller than the wavelength of the exciting light, dipolar and higher multipolar plasmons can

be excited, or undergo transitions. Many of the higher order multipolar plasmons are non-

radiative, and depending on the size of the feature, or particle, all plasmons other than the dipolar

Electron Hole

CT h!

After t = !

h! " h!v

Page 64: An Electrochemical and SERS Study of the Gold-Thiosulfate

48

plasmon, can be ignored. If the plasmon is resonant with the incident radiation, these plasmons

are sustained by the local electronic structure of the metal13, 16, 18. The more “free” the conduction

band electrons are in the metal, the sharper and more intense the excitation of dipolar plasmon’s

resonance will be. The excited plasmons may become confined within the nano-sized features,

creating a highly localized electromagnetic field9, 13, 16, 19, that can then interact with the incident

radiation.

The enhancement observed for a metal nanoparticle is highly dependent on a number of

factors, including the dielectric constants of both the environment of the particle and the particle

itself, as well as its aspect ratio9, 13, 17. Equation 3.31 describes the electric field produced outside

a spherical nanoparticle of radius a, when irradiated with z-polarized light of wavelength λ13:

Eout x, y, z( ) = E0 z !!ME0zrd3 !3zrd5 xx + yy + zz( )"

#$

%

&' (3.31)

Where x!, y!, z!( ) are the Cartesian unit vectors, rd the radial distribution, and αM the metal

polarizability given by!M = ga3 , where g is defined as13:

g = ! in " !out! in + 2!out( ) (3.32)

with εin, εout the dielectric constant of the nanoparticle, and the surrounding environment,

respectively. It can be seen from Equation 3.31 that the electric field enhancement decays with

rd-3, meaning there is a finite sensing volume around the nanoparticle, and that surface

enhancement is a near-field effect9, 13, 16, 17, 19.

Using Mie theory, an expression for the extinction spectrum of a non-spherical particle

can be derived:

Page 65: An Electrochemical and SERS Study of the Gold-Thiosulfate

49

E !( ) = 24"Na3#out

3/2

! ln 10( )# i !( )

# r !( ) + $#out( )2 + # i !( )2%

&''

(

)**

(3.33)

Where N is the finite number of polarizable elements (dipoles) within the nanoparticle, εi and εr

are the imaginary and real components of the dielectric constant for the nanoparticle, and χ is a

shape factor which accounts for geometries other than a spherical particle13. This shape factor is

also highly dependent on the dielectric of the external environment (εout), and so also acts as a

sensitivity factor of the localized surface plasmon resonance (LSPR) extinction spectrum to the

dielectric environment. For a sphere, χ has a value of 2, however, for other aspect ratios, it can

range as high as 20. 13, 20

3.2.2 SERS Substrates

SERS active substrates can be fabricated a number of ways including electrochemically

roughened electrodes and template nanostructures.

Electrochemically roughened electrodes are frequently used throughout the literature as

SERS active substrates. They can be generated through one or more oxidative reductive cycles (a

technique similar to cyclic voltammetry). During the oxidative cycle, the metal combines with a

salt (usually a halide) at the metal surface. When the reductive half cycle is begun the salt

dissociates and the native metal re-deposits on the surface in a non-uniform fashion. Using this

technique, metal clusters are formed with an apparent maximum size of ~200Å9. Although these

surfaces are fairly easily prepared, a common problem is the non-homogeneity of the sample,

subsequently leading to low reproducibility and stability5.

Templated nanostructures, in contrast with electrochemically-roughened surfaces,

provide a reasonably homogenous, and therefore reproducible surface. To achieve such a

reproducible surface, electrochemical reduction of various metals inside nanopores is perhaps

Page 66: An Electrochemical and SERS Study of the Gold-Thiosulfate

50

one of the best options due to its ability to confine and restrict the geometry of nanostructure to

that of the nanopores. To achieve such a result, anodic aluminum oxide films are commonly

used21. The use of gold and silver nanorods has been found to provide an enhancement up to 109

in a variety of systems19, 22. Pore dimensions, and thus dimensions of the nanorods, can be

controlled by the applied voltage, pH and composition of the electrolyte5.

The presence of a second nanoparticle within 1nm of other particles allows for even

greater enhancement, in the range of 1011. This effect can be attributed to the formation of an

effectively huge capacitive field from the viewpoint of a molecule located at the interstice

between two nanoparticles. As the particles are brought together, the dipoles in each are oriented

in opposite directions. This interaction with the incident light results in a greatly enhanced

Raman signal. 16

References

1. Bard, A. J.; Faulkner, L. R. In Kinetics of Electrode Reactions; Harris, D., Swain, E. and

Aiello, E., Eds.; Electrochemical Methods: Fundamentals and Applications; John Wiley & Sons, Inc.: New York, NY, USA, 2001; Vol. 2nd Ed., pp 87-132.

2. Skoog, D. A.; Holler, F. J.; Crouch, S. R. In Electroanalytical Chemistry; Kiselica, S., Short, M. A., Eds.; Principles of Instrumental Analysis; Thomson Brooks/Cole: Belmont, CA, USA, 2007, 1998; Vol. 6th Ed., pp 627-653.

3. Baron, J.; Szymanski, G.; Lipkowski, J. J Electroanal Chem 2011, 662, 57-63. 4. Bard, A. J.; Faulkner, L. R. In Methods Involving Forced Convection - Hydrodynamic

Methods; Swain, E., Harris, D. and Aiello, E., Eds.; Electrochemical Methods: Fundamentals and Applications; John Wiley & Sons, Inc.: New York, NY, USA, 2001; Vol. 2nd Ed., pp 331-364.

5. Baron Gavidia, J. Study of the Gold-Thiosulfate Interface Under Leaching Conditions, University of Guelph, Guelph, ON, 2010.

6. Skoog, D. A.; Holler, F. J.; Crouch, S. R. In Raman Spectroscopy; Kiselica, S., Short, M. A., Eds.; Principles of Instrumental Analysis; Thomson Brooks/Cole: Belmont CA, USA, 2007,; Vol. 6th Ed., pp 481-495.

7. Banwell, C. N. In Fundamentals of Molecular Spectroscopy; McGraw Hill Book Company (UK) Limited: United Kingdom, Vol. 3rd Ed., pp 338.

8. Fleischmann, M.; Hendra, P.; McQuillan, A. Chemical Physics Letters 1974, 26, 163-166.

9. Moskovits, M. Reviews of Modern Physics 1985, 57, 783-826.

Page 67: An Electrochemical and SERS Study of the Gold-Thiosulfate

51

10. Albrecht, M. G.; Creighton, J. A. J. Am. Chem. Soc. 1977, 99, 5215-5217. 11. Jeanmaire, D. L.; Van Duyne, R. P. Journal of Electroanalytical Chemistry and

Interfacial Electrochemistry 1977, 84, 1-20. 12. Tian, Z. Q.; Ren, B.; Wu, D. Y. J Phys Chem B 2002, 106, 9463-9483. 13. Stiles, P. L.; Dieringer, J. A.; Shah, N. C.; Van Duyne, R. R. Annual Review of Analytical

Chemistry 2008, 1, 601-626. 14. Brolo, A. G.; Irish, D. E.; Smith, B. D. J. Mol. Struct. 1997, 405, 29-44. 15. Mock, J.; Norton, S.; Chen, S. Y.; Lazarides, A.; Smith, D. Plasmonics 2011, 6, 113-124. 16. Moskovits, M. J. Raman Spectrosc. 2005, 36, 485-496. 17. Liao, Q.; Mu, C.; Xu, D. S.; Ai, X. C.; Yao, J. N.; Zhang, J. P. Langmuir 2009, 25, 4708-

4714. 18. Xu, H.; Käll, M. ChemPhysChem 2003, 4, 1001-1005. 19. Bok, H. M.; Shuford, K. L.; Kim, S.; Kim, S. K.; Park, S. Langmuir 2009, 25, 5266-

5270. 20. Joseph, V.; Matschulat, A.; Polte, J.; Rolf, S.; Emmerling, F.; Kneipp, J. J. Raman

Spectrosc. 2011, . 21. Matefi-Tempfli, S.; Matefi-Tempfli, M.; Vlad, A.; Antohe, V.; Piraux, L. Journal of

Materials Science-Materials in Electronics 2009, 20, 249-254. 22. Orendorff, C. J.; Gearheart, L.; Jana, N. R.; Murphy, C. J. Physical Chemistry Chemical

Physics 2006, 8, 165-170.

Page 68: An Electrochemical and SERS Study of the Gold-Thiosulfate

52

CHAPTER 4: METHODOLOGY

4.1 REAGENTS

All solutions were prepared using Milli-Q water with a resistivity of 18.2 MΩ cm with a

thiosulfate concentration of 100mM. Calcium thiosulfate solutions were prepared from calcium

thiosulfate (20-30% by weight) aqueous Captor® solutions from Tessenderlo Kerley Ltd. The

pH of the calcium and sodium thiosulfate solutions was adjusted to 8-8.5 using calcium

hydroxide (95%) from Sigma-Aldrich. Ammonium thiosulfate solutions were prepared from

ammonium thiosulfate (99%) from Alfa Aesar, and were pH adjusted using FisherBrand® ACS-

Pur ammonium hydroxide (28-30% NH3 w/w). Sodium thiosulfate pentahydrate (99.5%) salt

was acquired from Acros Organics. Copper was added to all leaching solutions as copper sulfate

pentahydrate (98+%) from Sigma-Aldrich to a concentration of 10 mM.

4.2 CLEANING METHODS

All glassware pieces were cleaned by soaking in a hot acid bath with concentrated H2SO4

and HNO3 in a ratio of 3:1 v/v for 60 min. After soaking in the acid bath the pieces were

thoroughly rinsed with Milli-Q water.

The gold disk rotating disk electrode was polished using 1 µm LECO® Premium

Diamond Suspension, and rinsed by rotation in Milli-Q water and methanol. Prior to use in

leaching current experiments the RDE was subjected to reductive desorption in 0.3 M NaOH

(semi-conductor grade 99.99% trace metals basis from Sigma-Aldrich). The solid polycrystalline

gold electrode was cleaned by flame annealing, followed by rinsing with Milli-Q water and

drying over a flame. Prior to each experiment, the counter electrode was flame annealed, rinsed

Page 69: An Electrochemical and SERS Study of the Gold-Thiosulfate

53

with Milli-Q water and dried over a flame.

Cleanliness of both the RDE and solid polycrystalline electrodes was checked by

cyclic voltammetry in a solution of 0.1 M NaF (≥99% from Sigma-Aldrich). The system was

considered clean when a cyclic voltammogram (CV) characteristic of a bare polycrystalline gold

electrode was achieved. Figure 4.1 displays a representative CV for either of the two gold

electrodes used. If the shape of the voltammogram matched those of literature1, 2, the electrodes

were considered clean.

Figure 4.1: Cyclic voltammogram of a clean, bare gold RDE in a solution of 0.1 M NaF. The inset graph

displays the double layer region of the gold electrode.

-1.0 -0.5 0.0 0.5 1.0 1.5

-20

-15

-10

-5

0

5

10

15

20

25

-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

i / µΑ

cm

-2

Potential / V vs. SCE

Page 70: An Electrochemical and SERS Study of the Gold-Thiosulfate

54

4.3 EXPERIMENTAL SETUP AND PROCESS

4.3.1 Preparation of Gold Electrodes

The polycrystalline gold electrode used for electrochemical experiments was made in-house

using the procedure outlined by Richer3.

4.3.2 Electrochemical Experiments

Electrochemical experiments were carried out using a HEKA (PG 590, Lambrecht/Pfalz,

Germany) potentiostat/galvanostat connected to an acquisition board from National Instruments

(PCI 6052E). Custom written software written by Professor Dan Bizzotto from the University of

British Columbia, and Professor Ian Burgess from the University of Saskatchewan, was used for

data acquisition.

4.3.2.1 Leaching Current Measurements

A typical three-electrode glass cell was used for the electrochemistry experiments. The

reference electrode (RE) was a saturated calomel electrode (SCE) placed in a separate cell in

order to avoid cross contamination between the investigated solution and the RE. The two cells

were connected with a salt bridge. Leaching current measurements using the RDE were achieved

with a gold coil as the counter electrode (CE), placed in a separate compartment in contact with

the solution through a glass frit. In the case of measurement using the solid polycrystalline

electrode (hanging meniscus configuration), the gold coil CE was submerged into the leaching

solution. Both the solid polycrystalline electrode and the RDE setups were left open to the

atmosphere, without purging of the solution. Figure 4.2 displays schematics of the two cells used

for the RDE and solid gold polycrystalline electrode.

Page 71: An Electrochemical and SERS Study of the Gold-Thiosulfate

55

Figure 4.2: Schematic for the two cell configurations used.

Linear sweep voltammetry was employed in order to probe the system and calculate the

current as described by J.Y. Baron4.

4.3.2.2 Open Circuit Potential Measurements

Measurement of the open circuit potentials for all three solutions (calcium, sodium and

ammonium thiosulfate with Cu(II)) were performed using the same three electrode setup as in

the leaching current measurements with an RDE (Figure 4.1). The rotating disk electrode was

cleaned according to the procedure described previously. Once cleanliness of the system was

ensured, the RDE was submerged into the leaching solution and the open circuit potential was

recorded for 3 hours using a custom potential-current (E-i) monitor program.

Salt bridge to

SCE

Au working

electrode Au

counter electrode

Au RDE (connected to motor)

Au counter

electrode Salt bridge to

SCE

Page 72: An Electrochemical and SERS Study of the Gold-Thiosulfate

56

4.3.2.3 Solution pH Measurements

The pH of the leaching solutions was monitored for 3 hours in order to understand possible

reaction pathways at the interface during the leaching reaction. Measurement of the solution

potential was accomplished using an Orion Expandable ionAnalyzer EA 920 pH meter with an

Accumet pH electrode. The electrode was immersed into a 150 mL beaker of the leaching

solution of interest, for 3 hours. Using the E-i monitor program, the potential output from the pH

electrode was recorded. The pH of the solution was calculated through the measurement of the

potential and pH of a set of standard buffer solutions, plotting each on a graph of potential vs. pH

(Figure 4.3). A linear regression was performed, generating a line of best fit with an R2 value of

0.99585. Using the equation of the line, pH = E !0.01836( ) + 6.90614 , the pH was calculated

and plotted as a function of time.

Figure 4.3: Calibration curve for the calculation of pH from the potential measured.

-200 -150 -100 -50 0 50 100 150 2003

4

5

6

7

8

9

10

11

pH

Potential / mV

Equation y = a + b*x

Weight No Weighting

Residual Sum of Squares

0.07405

Pearson's r -0.99862

Adj. R-Square 0.99585Value Standard Error

pHIntercept 6.90614 0.10022Slope -0.01836 6.83401E-4

Page 73: An Electrochemical and SERS Study of the Gold-Thiosulfate

57

4.3.3 Preparation of Gold Nanorod Electrodes

Gold nanorod electrodes used for SERS leaching experiments were prepared using 13mm

diameter anodized aluminum oxide filters with a 0.1 µm pore size, and a thickness of 60 µm as

templates for the electro-deposition of gold. Nanorods were deposited using two different

template treatments.

In the first method of growth, vapor deposition was employed to deposit a gold film, 70

nm thick, onto the back of the templates for electrical contact to a gold slide. The gold coated

filter was then pressed to the gold slide using a Kel-F® cell, where the uncoated side of the

template faced the open portion of the cell. This assembly was then attached to a potentiostat via

a gold wire and placed in a conventional 3-electrode glass cell, using gold foil as a CE, and a

SCE as the reference.

Uncoated templates were used in the second growth method in order to attach the

nanorods directly attached to the slide. A similar cell design as that for coated filters was used,

with the primary difference being a smaller diameter aperture in the top of the cell, and a greater

distribution of increased pressure on the template and gold slide for minimization of the space

where roughened gold would deposit.

Electrodeposition was carried out by submerging the template in TECHNIC gold solution

(TG-25 RTU), which was de-aerated by passing argon through the solution for 60 minutes. For

growth using coated filters a constant potential of -0.900 V was applied for 3-8 hours. Uncoated

filters required application of a lower potential (-0.800 V) for a significantly longer period of

time (15-20 hours). Once deposition was complete, the electrode was rinsed with Milli-Q water,

and the aluminum template dissolved by submerging the electrode in a 3 M NaOH (99.99% from

Sigma Aldrich) solution, for 3 hours.

Page 74: An Electrochemical and SERS Study of the Gold-Thiosulfate

58

Nanorod electrodes were cleaned using two different methods, depending on the growth

assembly used. Electrodes grown using coated templates were cleaned by gentle rinsing with

chloroform and a 30 min incubation in the UV/O3 chamber. Uncoated templates produced

electrodes that were mechanically attached to the underlying gold slide, greatly improving the

durability of the gold nanorods. Electrodes grown in such a fashion were cleaned by submerging

the whole cell into piranha solution for 15-30 minutes, followed by thorough rinsing with Milli-

Q water. The nanorod electrodes were then left to soak in Milli-Q water for 60 minutes before

use in experiments

4.3.4 Raman Experiments

A Renishaw Raman Imaging Microscope was used for Raman experiments. A NIR diode

laser with a wavelength of 785 nm and an output power of 300 mW was used for excitation. The

Raman instrument was equipped with a CCD array detector. The spectrometer was calibrated

using the Raman active vibration of silicon at 520 cm-1. A 63x immersion objective from Leica

was used for solution Raman and SERS experiments.

4.3.4.1 SERS

SERS spectra for all systems were collected at 10% power to avoid laser-induced

decomposition of species on the gold surface. For the calcium thiosulfate system the exposure

time was 5 s with 10 accumulations. For time dependent experiments, the static mode was used

to minimize the time required for data acquisition. In order to cover the whole spectral range of

interest, spectra centered at 400 cm-1 and 900 cm-1 were taken for each experiment and then

combined into one spectrum.

Spectra corresponding to the ammonium and sodium thiosulfate systems were collected

using the extended mode with a range of 100-2000 cm-1 and an exposure time of 15 s and 10

Page 75: An Electrochemical and SERS Study of the Gold-Thiosulfate

59

accumulations. Spectra were collected over a period of 3 hours after exposure of the nanorod

electrode to the leaching solution of interest.

4.3.4.2 Solution Raman

Solution spectra were collected using 100% power in extended mode, with an exposure

time of 15 s and 10 accumulations. Spectra were recorded for 3 hours after the time of solution

preparation.

References

1. Clavilier, J.; Van Huong, C. N. Journal of Electroanalytical Chemistry and Interfacial

Electrochemistry 1977, 80, 101-114. 2. Stolberg, L.; Richer, J.; Lipkowski, J.; Irish, D. Journal of electroanalytical chemistry

and interfacial electrochemistry 1986, 207, 213-234. 3. Richer, J. Measurement of Physical Adsorption of Neutral Organic Species at Solid

Electrodes, University of Guelph, Guelph, ON, Canada, 1985. 4. Baron, J.; Szymanski, G.; Lipkowski, J. J Electroanal Chem 2011, 662, 1, 57-63.

Page 76: An Electrochemical and SERS Study of the Gold-Thiosulfate

60

CHAPTER 5: RESULTS AND DISCUSSION

5.1 PREAMBLE

Gold leaching in industrial settings is carried out at the open circuit potential of the

system. At this potential, passivation of the gold surface can occur as a result of a number of

complicated reactive pathways that lead to the formation of oxidation and decomposition

products of thiosulfate, such as polythionates and elemental sulfur1-5. Addition of copper and

ammonia to the thiosulfate leaching system has been shown to have significant effects on the

gold leaching rate1-3, 6, 7. However, there is limited literature concerning experimental studies of

the thiosulfate leaching system in the presence of these additives individually. The work

described in this chapter was designed to provide a base understanding of the individual, and

cooperative effects of copper and ammonia on the passive layer.

Characterization of the systems of interest (calcium, sodium and ammonium thiosulfate)

was carried out using both electrochemical and spectroscopic analysis. Using the method devised

by Baron et al.8 the measured leaching currents of a non-passivated surface can be measured with

minimal perturbation to the system. This is an important characteristic of this methodology, as

application of potentials more positive than the mixed potential of the system can lead to electro-

oxidation of thiosulfate5.

Raman spectroscopy is particularly useful in the study of oxygenated inorganic sulfur

species due to their extensive characterization, and distinct intense vibrational modes that allow

for fingerprint identification of a multitude of species in a given sample9-11. This thesis employed

both solution Raman and Surface Enhanced Raman Spectroscopy (SERS) to study the bulk

Page 77: An Electrochemical and SERS Study of the Gold-Thiosulfate

61

solution and gold-thiosulfate interface in the presence of copper and ammonia to identify species

present at the gold surface.

Band shifting relative to the normal Raman spectrum (aqueous or solid) is expected upon

adsorption of a molecule to a SERS active surface12. The SERS band intensity arises from an

enhancement that is controlled by a number of factors, including: i) electric field localization at

the SERS active surface and ii) the orientation of the polarizability tensor as a result of the

geometry of adsorption. As a general rule, surface vibrations that affect the change in the

polarizability in a direction perpendicular to the surface will undergo the greatest enhancement.

Adsorbates with π electrons are also susceptible to shifts in the vibrational frequency as a result

of a change in the potential of the surface. Thus, the SERS spectra are extremely sensitive to the

open circuit potential of the system12.

Due to the complex composition of the passive layer, a SERS active substrate with very

high enhancement is required to identify various oxygen containing inorganic sulfur species

within the passive layer. The destructive nature of the leaching reaction makes a substrate with

long-term stability a desirable and necessary feature for long term studies. Gold nanorod arrays

have been shown to provide significant enhancement, and long-term stability when used as a

SERS active substrate to study the gold-thiosulfate interface under leaching conditions13, 14.

5.2 CHARACTERIZATION OF BULK SOLUTION

To fully interpret and understand the results acquired in characterizing the passive layer

formed at the gold-thiosulfate interface under leaching conditions, a thorough understanding of

the composition and behavior of the bulk solution must be achieved. Hence, both the behavior

and composition of the solution must be characterized.

Page 78: An Electrochemical and SERS Study of the Gold-Thiosulfate

62

5.2.1 pH Measurements

All leaching experiments were performed in solutions in equilibrium with ambient air,

and as such are sensitive to the effect of dissolved CO2. The pH of the leaching solution can have

a significant effect on decomposition pathways, and species present. Therefore, it was important

to determine the pH of all three systems over a period of 3 hours by recording the output from a

pH meter. Standard buffer solutions were used to calibrate the readout from the pH meter. The

results after calibration are plotted in Figure 5.1.

It was expected that the pH of the ammonium thiosulfate system would undergo the most

significant change due to possible evaporation of ammonia from the solution. However, the

buffering ability of NH4+ /NH3 can explain the fairly constant pH level in this solution. More

surprising are the behaviors of the calcium and sodium thiosulfate systems. Each of these

solutions displays an initial increase of approximately 0.4 pH units over a period of

approximately 30 min, followed by a drop in pH. After 3 hours the pH of the sodium thiosulfate

solution returns to its initial value (~ pH 8.2), a drop of approximately 0.6 pH units from the

maximum pH (~8.8). A drop of this magnitude is similar to that of the blank system (pure water),

where dissolved carbon dioxide is responsible for the change in pH. The behaviour of the

calcium thiosulfate system is far more drastic. The pH decreases by almost two pH units from its

initial value. A decrease of that magnitude could not solely be caused by dissolved carbon

dioxide.

Page 79: An Electrochemical and SERS Study of the Gold-Thiosulfate

63

Figure 5.1: Average pH over a period of 3 hours for calcium, ammonium and sodium thiosulfate systems

(0.1 M S2O3 + 0.01 M CuSO4 and adjusted to a pH of 8.0-8.5) and a blank solution of pure MilliQ water.

However, formation of a precipitate was observed in the calcium thiosulfate solution near

the end of the experiment. This could be explained by the precipitation of either CaCO3 or

CaSO4, leading to the observed drop in pH in the calcium thiosulfate solution. Both of these

species may be precipitating from solution; however, based on the solubility product constants of

these two salts (4.96 x 10-9 and 7.10 x 10-5 for calcium carbonate and calcium sulfate,

respectively), and evidence from industrial settings, it is more likely that calcium carbonate is

precipitating.15

0 25 50 75 100 125 150 175 2005.0

5.5

6.0

6.5

7.0

7.5

8.0

8.5

9.0

A

vera

ge p

H o

f Lea

chin

g So

lutio

n

Time / minutes

STS

ATS

CaTSMilliQ Water Blank

Page 80: An Electrochemical and SERS Study of the Gold-Thiosulfate

64

Significant changes in the pH of the leaching solutions were observed, depending on the

cation of the thiosulfate salt. These changes can greatly affect the composition, and subsequent

efficacy of the leaching process.

5.2.2 Solution Raman

Raman spectra of leaching solutions were collected for a period of 3 hours to correlate

any change in bulk solution species to those observed at the interface in SERS spectra. Many of

the decomposition products of these leaching solutions, such as tetrathionate, trithionate and

sulfite, have well characterized aqueous solution Raman spectra reported in the literature.

Characteristic vibrations for thiosulfate, tetrathionate, trithionate, sulfate and sulfite can be found

in Table 5.1. These vibrations can be used for fingerprint identification in the more complex

spectra of the leaching solutions. Figure 5.2 (a-c) displays the average spectra of 3 separate

Raman experiments for each of the calcium, ammonium and sodium thiosulfate systems,

respectively.

From these spectra, it is evident that there are two main regions of interest, 300 cm-1 to

700 cm-1, and 900 cm-1 to 1200 cm-1. Bands located at ~ 190 and 1640 cm-1 are attributed to

vibrational modes of water, and will not be further discussed in the scope of this chapter. Using

the Fourier Self-Deconvolution (FSD) method described by J. Baron13, 16, each of the regions

were analyzed by deconvoluting the broad bands into individual peaks that could be assigned to

a specific vibration. Fitting of peaks was performed using a mixed Gaussian and Lorentzian band

shape. It was found that the calcium, ammonium and sodium thiosulfate solutions displayed the

same bands, and upon peak fitting, the same component peaks. The deconvoluted spectrum of

the calcium thiosulfate system in the 400 cm-1 region, recorded 20 min after solution preparation,

is presented in Figure 5.3. The upper connected points correspond to the raw experimental

Page 81: An Electrochemical and SERS Study of the Gold-Thiosulfate

65

envelope, the solid line overlapping the raw data points is the simulated envelope, and the lowest

solid coloured lines represent the deconvoluted peaks.

Table 5.1: Characteristic Raman active vibrational modes of species expected in the Raman spectra of the

investigated leaching solutions10, 17-19.

Species Assignment Raman Shift / cm-1

Thiosulfate (S2O32-)

ρr (SSO) 334

νsym

(SS) 443

δasym

(SO3) 533

δsym

(SO3) 663

νsym

(SO3) 995

νasym

(SO3) 1122

Tetrathionate (S4O62-)

δ (SSinternal

) 260 δ (SS

terminal) 310

δ (SSterminal

) 390 δ

asym (SO

3) 532

δ sym

(SO3) 651

νsym

(SO3) 1040

Trithionate (S3O62-)

δ (SS) 264

δ (SS) 425

δsym

(SO3) 675

νsym

(SO3) 1055

Sulfate (SO42-)

δsym

(OSO) 448

δasym

(OSO) 620

νsym

(SO3) 982

νasym

(SO) 1110

Sulfite (SO32-)

δasym

(OSO) 470

δsym

(OSO) 620

νasym

(SO) 933 ν

sym (SO) 967

Page 82: An Electrochemical and SERS Study of the Gold-Thiosulfate

66

Figure 5.2: Average raw Raman spectra collected over a period of 3 hours for leaching solutions of: a)

0.1 M CaS2O3 + 0.01 M CuSO4 b) 0.1 M (NH4)2S2O3 + 0.01 M CuSO4 c) 0.1 M Na2S2O3 + 0.01 M

CuSO4. All solutions were adjusted to pH 8.0-8.5.

500 1000 1500 2000

Raman Shift / cm-1

10 min20 min33 min46 min57 min69 min

80 min110 min140 min170 min200 min

500 1000 1500 2000

Raman Shift / cm-1

10 min30 min

42 min

52 min

62 min72 min

92 min112 min

142 min

180 min

500 1000 1500 2000

Raman Shift / cm-1

5 min17 min

28 min39 min53 min

62 min

72 min102 min

142 min

182 min

a) b)

c)

Page 83: An Electrochemical and SERS Study of the Gold-Thiosulfate

67

Figure 5.3: Deconvoluted spectra of an average of three 0.1 M CaS2O3 + 0.01 M CuSO4 solutions,

adjusted to a pH of 8.0-8.5, 20 min after solution preparation.

Five peaks were identified, at positions of 387, 425, 446, 520, and 666 cm-1. The first

band at 387 cm-1 can be attributed to the presence of tetrathionate in solution10, 13, 17. Strong

bands at 425 and 446 cm-1 correspond to the stretches of trithionate and thiosulfate, respectively.

The assignment of bands at 520 and 666 cm-1 is uncertain, because they could correspond to

multiple oxygenated inorganic sulfur species. The band at 520 cm-1 is very near an asymmetric

stretch of thiosulfate in aqueous solution, 533 cm-1, and an asymmetric stretch of tetrathionate

(532 cm-1).10

300 400 500 600 700 800δ sy

m(O

SO) S

2O2- 3

ν sym

(SS)

S2O

2- 3

δ(SS

) S3O

2- 6

δ asym

(SO

) S2O

2- 3 / δ as

ym(O

SO) S

4O2- 6

Raman Shift / cm-1

δ(SS

term

inal) S

4O2- 6

Page 84: An Electrochemical and SERS Study of the Gold-Thiosulfate

68

Higher wavenumber sections of the spectra can also contain bands useful for

identification, and thus, cannot be ignored. Figure 5.4 shows the deconvolution of the 900-

1200 cm-1 region in the spectrum recorded 20 minutes after solution preparation.

Figure 5.4: Deconvoluted spectra of an average of three 0.1 M CaS2O3 + 0.01 M CuSO4 solutions,

adjusted to a pH of 8.0-8.5, 20 min after solution preparation, in the 900-1000 cm-1 region.

Four distinct peaks positioned at 980, 997, 1037, and 1130 cm-1 are easily identifiable in

this region of the spectrum. The symmetric stretch of aqueous sulfate matches the band observed

at 980 cm-1 in literature 18. Thiosulfate has a strong peak at 995 cm-1, and a weaker broad peak

near 1122 cm-1. Although there is a slight shift from these literature values, the peaks at 997 and

1130 cm-1 in Figure 5.4 can be assigned to thiosulfate13. The final peak at 1037 cm-1 can be

assigned to a symmetric stretch of tetrathionate10.

900 1000 1100 1200

ν sym(S

O) S

O2- 4

ν asym(S

O) S

2O2- 3

ν sym(S

O) S

2O2- 3

ν sym(S

O) S

4O2- 6

Raman Shift / cm-1

Page 85: An Electrochemical and SERS Study of the Gold-Thiosulfate

69

FSD was performed on each of the spectra acquired and the deconvoluted bands’

positions were matched with characteristic vibrations of species present in solution. Analytical

areas for each of the species of interest were tracked for the duration of the experiment in each of

the calcium, ammonium and sodium thiosulfate systems (Figure 5.5). Peak areas were

normalized with reference to the water band seen at 1640 cm-1 in each of the spectra in Figure

5.2.

Figure 5.5: Normalized peak areas for band positions of 1037 cm-1 (tetrathionate = black), 425 cm-1

(trithionate = blue), 980 cm-1 (sulfate = green), 448 cm-1 (thiosulfate = red). a) 0.1 M CaS2O3 + 0.01 M

CuSO4, b) 0.1 M Na2S2O3 + 0.01 M CuSO4, c) 0.1 M (NH4)2S2O3 + 0.01 M CuSO4. Solutions were

adjusted to pH 8.0-8.5.

0 50 100 150 200

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

νsym(SO) S4O2-6

νsym(SO) SO2-4

Nor

mal

ized

Ana

lytic

al P

eak

Are

a

Time / minutes

νsym(SS) S2O2-3

δ(SS) S3O2-6

0 50 100 150 2000.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

νsym(SO) SO2-4

νsym(SO) S4O2-6

Nor

mal

ized

Ana

lytic

al P

eak

Are

a

Time / minutes

νsym(SS) S2O2-3

δ(SS) S3O2-6

0 50 100 150 2000.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

νsym(SO) S4O2-6

νsym(SO) SO2-4

Nor

mal

ized

Ana

lytic

al P

eak

Are

a

Time / minutes

νsym(SS) S2O2-3

δ(SS) S3O2-6

a) b)

c)

Page 86: An Electrochemical and SERS Study of the Gold-Thiosulfate

70

It is evident from the data above that the composition of all three bulk solutions (CaTS,

ATS, and STS) remains fairly constant, with most species maintaining a steady state. Some

variation is observed in the peak area of trithionate in the ammonium thiosulfate system, leading

to a very slight decrease over the final 80 minutes of the experiment. However, the solutions’

compositions maintain a steady state concentration of thiosulfate, tetrathionate, and sulfate.

An important feature of these spectra is the appearance of trithionate early in the

experiment. In previous studies, trithonate was not seen until much later in the experiment13, 20.

Appearance of trithionate within the first 20 minutes after solution preparation is indicative of

catalysis of the decomposition or disproportionation of tetrathionate, likely by Cu2+ in the

solution:

2S4O62! " S3O6

2! + S5O62! (5.1)

4S4O62! + 3OH! " 5S2O3

2! + 2S3O62! + 3H2O (5.2)

In the step-wise reaction sequence for the decomposition of tetrathionate, thiosulfate is

both a product and a catalyst for the reaction. This is one possible reason why the thiosulfate

solutions containing copper maintained a constant thiosulfate concentration. Thiosulfate

solutions without copper can suffer a reduction in thiosulfate concentration over extended

periods of time, depending on the pH of the solution.

Page 87: An Electrochemical and SERS Study of the Gold-Thiosulfate

71

5.3 CHARACTERIZATION OF THE GOLD-THIOSULFATE INTERFACE IN THE

PRESENCE OF COPPER

5.3.1 Initial Characterization

Using linear sweep voltammetry, kinetic characterization was performed using a gold

electrode in contact with a solution of 0.1 M Na2S2O3 + 0.01 M CuSO4, adjusted to pH 8.0-8.5

using Ca(OH)2. Prior to leaching current measurement, it was necessary to determine the values

for the charge transfer coefficients. A linear sweep voltammogram was collected using a sweep

rate of 1 mVs-1 in a potential range of -0.32 V to 0.2 V, as shown in Figure 5.6.

Figure 5.6: Linear sweep voltammogram of a 0.1 M Na2S2O3 + 0.01 M CuSO4 solution, pH 8.0-8.5.

According to mixed potential theory, the point at which the i-V curve crosses zero

corresponds to the mixed potential (or open circuit potential) of the system. That is, the

magnitude of the current contributed by the cathodic reduction of Cu2+, and that from the

-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2

-3

-2

-1

0

1

2

3

4

5

Cur

rent

/ µA

Potential / V vs. SCE

Page 88: An Electrochemical and SERS Study of the Gold-Thiosulfate

72

oxidation of Au, are equal but opposite in sign. Thus, the recorded net current is zero. By

plotting the logarithm of the measured current as a function of applied potential, a Tafel-like plot

can be obtained, as shown in Figure 5.7. The values of the transfer coefficients can be

determined by performing a linear regression analysis of the linear sections of the curve, at large

overpotentials (Figures 5.8 and 5.9).

Figure 5.7: Tafel plot of a 0.1 M Na2S2O3 + 0.01 M CuSO4 solution, pH 8.0-8.5, used for calculation of

the transfer coefficients of the system.

The slopes of the cathodic and anodic linear regions are given by Equations 5.3 and 5.4,

respectively:

slope = !" cF2.3RT

(5.3)

slope =1!" a( )F2.3RT

(5.4)

-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2

-3.0

-2.5

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0AnodicLinear

Log(i)

Potential / V vs. SCE

CathodicLinear

Page 89: An Electrochemical and SERS Study of the Gold-Thiosulfate

73

Using the slope calculated through the linear regressions in Figures 5.3 and 5.4, the transfer

coefficients for the reduction of Cu2+ and oxidation of Au were found to be 0.35 ± 0.02 and 0.78

± 0.05, respectively.

Figure 5.8: Linear regression of the cathodic reduction of Cu2+ for the determination of the transfer

coefficient. A slope of -6.10 was calculated.

Quantification of the transfer coefficients is necessary for the calculation of leaching

currents in the systems of interest. The Butler-Volmer equation discussed in Chapter 3 can

provide an accurate description of the current-overpotential curve for a simple electron transfer

reaction:

i = i0 exp!" cnFRT

#$%&

'() ! exp

1!" a( )nFRT

#$%&

'()

*

+,

-

./ (5.5)

-0.32 -0.30 -0.28 -0.26 -0.24 -0.22 -0.20-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

0.4

0.5

Log

(i)

Potential / V vs. SCE

Page 90: An Electrochemical and SERS Study of the Gold-Thiosulfate

74

A modified version is required here since the values of the transfer coefficients are not the same

for the cathodic and anodic reactions:

i = iM exp!"

Cu2+nCu2+F

RT#

$%&

'()! exp

1!"Au( )nAuFRT

#$%&

'()

*

+,

-

./ (5.6)

Figure 5.9: Linear regression of the anodic oxidation of Au for the determination of the transfer

coefficient. A slope of 4.75 was calculated.

The term i0 in Equation 5.5 is equivalent to the leaching current, iM, in the case of the

gold leaching reaction, as in Equation 5.6. The terms nCu2+ and nAu in Equation 5.6 both refer to

one electron transfer redox reactions, the rengeneration of Cu2+ and oxidation of gold,

thus nCu2+ = nAu =1 . Because the potential region of interest is small, Equation 5.6 can be

simplified to:

-0.08 -0.06 -0.04 -0.02 0.00

-0.2

0.0

Log

(i)

Potential / V vs. SCE

Page 91: An Electrochemical and SERS Study of the Gold-Thiosulfate

75

i = iM !"Cu2+

! 1!" Au( )( ) nFRT # (5.7)

A plot of the current density response in a ~ 40 mV region around the mixed potential

should have linear characteristics (Figure 5.10). Application of a linear regression to the data

provided a slope with which to calculate the value of the leaching current. For the data shown in

Figure 5.10, a leaching current density of 2.14 µA cm-2 was obtained for a temperature of 293 K,

assuming a one-electron transfer reaction.

Figure 5.10: Linear regression of data acquired during a linear sweep voltammogram of a 0.1 M Na2S2O3

+ 0.01 M CuSO4 solution, pH 8.0-8.5 solution with a sweep rate of 1 mVs-1.

This method of calculating leaching current densities was used throughout this thesis for

analysis of all systems of interest. The current efficiency of thiosulfate leaching media has been

shown to be close to 100% up to 0.08 V vs SCE20, thus the calculation of the current density in a

-0.16 -0.14 -0.12

-1

0

1

j / µ

Αcm

-2

Potential / V vs. SCE

Page 92: An Electrochemical and SERS Study of the Gold-Thiosulfate

76

small potential range around the mixed potential can be directly related to the gold leaching

current density.

To gain an understanding of the kinetic behavior of the systems of interest, it is necessary

to observe under which sweep rates the system is under mass transport or kinetic control. The

thickness of the diffuse layer is inversely proportional to the square root of the sweep rate. At

low sweep rates, the diffuse layer extends further from the electrode, greatly decreasing the flux

to the electrode surface, and thus producing a relatively low current. Higher sweep rates reduce

the thickness of the diffuse layer, leading to increased current. Current produced by a system

under mass transport control (such as at low sweep rates) are proportional to the square root of

the sweep rate. Deviations from such a relationship indicate that the system is moving under

mixed control, that is, contributions to the current arise from both diffusion to the electrode

surface and kinetic limitations of the electron transfer reaction. To determine if the systems of

interest were susceptible to kinetic control, the leaching currents at 6 sweep rates (1, 2, 10, 20, 50

and 100 mVs-1) were measured and plotted as a function of the square root of the sweep rate

(Figure 5.11).

The linear behavior at low sweep rates in Figure 5.11 indicates that the system is under

mass transport control under these conditions. As the sweep rate increases to 20 mVs-1, deviation

from the line is observed. Hence, at this sweep rate, and higher, the system moves to mixed

control. Since this probing of the system did not involve forced convection of the solution, it

should be possible through use of an RDE to bring the system under full kinetic control.

Measurement of the leaching current while under full kinetic control is required in order to have

an accurate understanding of the gold leaching reaction rate, without the contribution of diffusion

effects.

Page 93: An Electrochemical and SERS Study of the Gold-Thiosulfate

77

Figure 5.11: Sweep rate dependence of the leaching current measured in a 0.1 M Na2S2O3 + 0.01 M

CuSO4 solution, pH 8.0-8.5, at sweep rates of 1, 2, 10, 20, 50 and 100 mVs-1.

Therefore, leaching current densities were also measured as a function of the angular

velocity of a gold disk RDE, using rotation rates of 300, 500, 700 and 1000 RPM. Due to rapid

passivation of the gold surface as a result of the forced convection in the solution, the potential

range of the linear sweep was decreased to a 120 mV region (-220 mV to -100 mV). The

dependence of the leaching current density was plotted as a function of the square root of the

angular velocity (calculated from the rotation rate), as shown in Figure 5.12.

Within experimental error, the data displayed below are independent of the angular

velocity (and thus rotation rate). According to the Koutecky-Levich equation discussed in

Chapter 3, at each sweep rate a maximum, or limiting current, exists.

0 1 2 3 4 5 6 7 8 9 10 110

2

4

6

8

10

j m / µA

cm-2

ν1/2

/ mV1/2 s-1/2

Page 94: An Electrochemical and SERS Study of the Gold-Thiosulfate

78

Figure 5.12: Rotation dependence of the leaching current of a gold electrode in contact with a 0.1 M

CaS2O3 + 0.01 M CuSO4 solution, pH 8.0-8.5, at rotation rates of 300, 500, 700 and 1000 RPM. Data was

recorded using a sweep rate of 5 mV s-1.

This limiting current is given by the term iK in Equation 5.8, and describes the current in

the absence of mass transfer effects:

1i= 1iK

+ 1il ,c

= 1iK

+ 10.62nFADO

2/3! 1/2" #1/6CO* (5.8)

When iK is small, such that:

1iK

>> 10.62nFADO

2/3! 1/2" #1/6CO* (5.9)

6 8 100.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

6.0

j M / µA

cm-2

ω1/2

Page 95: An Electrochemical and SERS Study of the Gold-Thiosulfate

79

and mass transfer is efficient enough to keep the surface concentration of species equal to that of

the bulk, a plot of i vs. ω1/2 will show that i is independent of the rotation rate, as seen in Figure

5.12 with respect to the leaching current densities. Thus, it can be concluded that with rotation

rates as low as 300 RPM, kinetic control of the system can be achieved. Since measurement of

the leaching current in the solutions of interest needs to be carried out in a regime where no mass

transfer effects are present, a rotation rate of 300 RPM and sweep rate of 5 mVs-1 were chosen

for characterization of the systems of interest.

5.3.2 Leaching Current Measurements

Leaching current densities of all three systems of interest (calcium, sodium and

ammonium thiosulfate) were calculated using a set solution of 0.1 M S2O32- + 0.01 M CuSO4, pH

adjusted to 8.0-8.5. Measurement of the current response of a gold disk RDE was carried out

using the 3-electrode cell described previously. The rotation rate was set at 300 RPM, and the

potential was scanned between -220 mV and -100 mV using a sweep rate of 5 mVs-1. The

average calculated leaching current density for calcium thiosulfate (CaTS), sodium thiosulfate

(STS) and ammonium thiosulfate (ATS) are shown in Figure 5.13.

The data show that there is no variation, within experimental error, in the leaching current

density measured in all 3 systems. These results show a different effect of the cation on the

leaching reaction to those presented by Chandra et al.21, who noted an increase in the gold

oxidation polarization curves upon changing the alkali metal cation from sodium to potassium.

Based on these results, it was expected that upon comparison of the leaching current in the

sodium and calcium thiosulfate systems, the leaching current would be greater in the calcium

system.

Page 96: An Electrochemical and SERS Study of the Gold-Thiosulfate

80

Figure 5.13: Average calculated leaching current density for solutions of sodium, calcium and

ammonium thiosulfate solutions (0.1 M S2O3 + 0.01 M CuSO4 and were adjusted to a pH of 8.0-8.5).

The disagreement with literature can be attributed to the introduction of copper into the

systems studied in this work. In both the calcium thiosulfate and sodium thiosulfate systems,

there is a lack of ammonia in the leaching solutions. Ammonia is normally required to stabilize

the Cu2+ ion as the copper tetraamine complex, which is the oxidant in the leaching system1-3, 22,

23. In an ammonia free solution, Zhang and Nicol6 proposed that Cu2+ may form a reactive

intermediate with thiosulfate and oxygen that can react with free thiosulfate, producing the

Cu(S2O3)35! complex and tetrathionate through Equation 5.10:

S2O3( )3 Cu !O2"# $%5&

+ 4S2O32& + 2H2O' 2S4O6

2& + Cu(S2O3)35& + 4OH& (5.10)

STS CaTS ATS0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

j M / µA

cm-2

Thiosulfate Salt

Page 97: An Electrochemical and SERS Study of the Gold-Thiosulfate

81

The formation of complexes of copper and thiosulfate is well documented in literature. The

Cu(S2O3)35! or Cu(S2O3)2

3! complexes are the most thermodynamically stable copper complexes

in the concentration, pH and potential ranges employed in this work2, 3, 24-26. The presence of this

copper-thiosulfate redox couple significantly changes the properties of the solutions and explains

the differences between the work of Chandra et al.21, and the results presented in this chapter.

5.3.3 Open Circuit Potential Measurements

The open circuit potential (OCP) of a stationary gold disk RDE submerged in the

solutions of interest was tracked for an immersion period of 3 hours. The resulting curves as a

function of time can be seen in Figure 5.9. At t = 0, the OCP of the system is 28.6 mV, 18.9 mV,

and -3.6 mV for STS, CaTS and ATS, respectively. The high positive values for these systems

may be the result of adsorption of thiosulfate to the clean electrode surface. Each system then

decays rapidly to more negative potentials, which may indicate that the gold leaching reaction is

favored during this time. In the calcium and ammonium thiosulfate leaching solutions, the

potential curve reaches a negative limit, and then rises in an asymptotic fashion. The OCP of the

sodium thiosulfate system, while it behaves similarly to the calcium and ammonium thiosulfate

systems in the first 15 minutes of immersion, differs in the rate at which the potential rises

toward positive potentials. Rather than rising asymptotically after reaching a minimum, the

potential curve rises in a quasi-linear fashion.

The relative magnitudes for the negative limits for the OCPs match the results of the

oxidation polarization curve results of Chandra et al.21. That is, a progressively more negative

limit is reached as the cation is changed from sodium, to calcium, to ammonium.

Page 98: An Electrochemical and SERS Study of the Gold-Thiosulfate

82

Figure 5.14: Average open circuit potential of a gold disk RDE as a function of immersion time in

solutions of calcium, sodium and ammonium thiosulfate (0.1 M S2O3 + 0.01 M CuSO4 and adjusted to a

pH of 8.0-8.5).

In order to replicate the conditions of the SERS leaching tests, measurement of the OCP

for all systems was performed while holding the disk electrode stationary. Under these

conditions, diffusion from the bulk solution to the electrode surface may directly affect the

potentials measured. The differences between the limiting values of the OCP may be a result of

ion pairing with thiosulfate in solution, as proposed by Chandra et al.21, or by the formation of

adsorbed mixed complexes at the gold-thiosulfate interface. Senanayake27 has shown that the

oxidation of gold is highly dependent on the concentration product [M2S2O3][NH3] (where M =

Na or NH4+), and that the concentration of the NH4S2O3

- ion pair was greater than or equal to the

concentration of free thiosulfate ions.

0 25 50 75 100 125 150 175 200

-0.125

-0.100

-0.075

-0.050

-0.025

0.000

0.025

0.050

Ave

rage

Ope

n C

ircui

t Pot

entia

l / V

vs.

SCE

Time / minutes

ATS

CaTSSTS

Page 99: An Electrochemical and SERS Study of the Gold-Thiosulfate

83

A greater percentage of available copper ions would be present as an oxidizing cupric

tetraamine complex in the ammonium thiosulfate leaching system. This should increase the rate

of gold electrode oxidation. In the sodium and calcium thiosulfate leaching systems, the stability

and subsequent effect of these ion pairs could vary greatly depending on the identity of the

cation, as no ammonia is present to stabilize the copper ions. The stability of the ion pairing

between the alkali metal cation and thiosulfate increases as one moves down the rows of the

periodic table21. Therefore, the heavier mass and increased charge of calcium could result in a

much more stable ion pair with thiosulfate than sodium. These differences in ion pairing may

lead to vastly different solvation and diffusion rates in each system; however, further work is

required to confirm these assumptions.

Typically, more positive open circuit potentials are indicative of passivation of the gold

surface. Thus, Figure 5.14 shows that in the first 15 minutes of exposure, the leaching process is

favored over passivation. At low exposure times, both the calcium and ammonium thiosulfate

solutions leach at a greater rate than the sodium thiosulfate leaching system. However, at

exposure times greater than 15 minutes, passivation occurs in these solutions. Differences in the

rate of passivation in the three systems could indicate that the identity of the cation may

influence the level and rate of passivation of the gold surface.

5.3.4 Surface Enhanced Raman

To characterize the effect of copper on the composition and properties of the gold-

thiosulfate interface, gold nanorod electrodes were employed as SERS active surfaces. Leaching

solution was introduced into a spectro-electrochemical cell with an assembled nanorod electrode.

Spectra were collected in situ during a 3 hour gold leaching experiment.

Page 100: An Electrochemical and SERS Study of the Gold-Thiosulfate

84

Figure 5.15 shows spectra collected for each of the calcium, ammonium and sodium

thiosulfate systems, as a function of exposure time. Upon comparison, it is evident that the SERS

spectra of species at the gold-thiosulfate interface are different from the Raman spectra of the

bulk solution (Figures 5.2-5.4). Unlike the bulk solution, the passive layer at the interface

undergoes significant changes in composition from the initial spectra.

Fourier Self-Deconvolution was used to identify peak positions, and peak fitting was

performed using a mixed Gaussian and Lorentzian band shape, as with the bulk solutions.

Fitted spectra of a gold nanorod electrode, 5 minutes after exposure to the calcium

thiosulfate leaching solution, can be seen in Figure 5.16.

Six bands are visible at positions of approximately 255, 400, 520, 650, 950 and 1010 cm-

1. Upon peak fitting, 10 individual peaks were identified. The broad band at 255 cm-1 was

separated into three peaks at positions of 216, 255, and 285 cm-1. The peak at 216 cm-1 was

assigned to the δ(S-S-S) of S8, while the peaks at 255 and 285 cm-1 could be attributed to either

ν(Cu-S) or ν(Au-S) 28. Although the solution spectra of both trithionate and tetrathionate display

vibrations near 260 cm-1, assignment of the 255 cm-1 peak to these species was avoided due to

the absence of vibrations definitively corresponding to tetrathionate (390, 1040, and 1233 cm-1),

and the lack of correlation in the intensity changes of the trithionate peaks (refer to Table 5.1).

However, the peak at 285 cm-1 is almost a direct match for NaHS on gold as shown by Jeffrey et

al.28

Page 101: An Electrochemical and SERS Study of the Gold-Thiosulfate

85

Figure 5.15: Raw SERS spectra of gold nanorod electrodes exposed to leaching solutions of 0.1 M S2O3

+ 0.01 M CuSO4, adjusted to pH 8.0-8.5. Solutions were a) CaTS b) ATS c) STS.

a)

b)

c)

Page 102: An Electrochemical and SERS Study of the Gold-Thiosulfate

86

Identification of the 255 and 285 cm-1 peaks as either ν(Cu-S) or ν(Au-S) is not possible

without further analysis of the samples with a complementary technique, due to the overlap

between the vibrational modes.

Figure 5.16: Fitted SERS spectrum of a gold nanorod electrode, after 5 minutes of exposure to a 0.1 M

CaS2O3 + 0.01 M CuSO4 leaching solution, with an initial pH of 8.0-8.5.

A peak centered at 405 cm-1 was assigned to the δ(S-S) mode of trithionate, although the

peak center was shifted 15 cm-1 lower than the corresponding vibration in aqueous solution. Such

a shift could be the result of adsorption to the gold surface13. A weak shoulder at 443 cm-1

directly correlated to the symmetric S-S vibration of the thiosulfate ion (νsym(S-S)). The weak band

at 530 cm-1 could either be assigned to the δasym(S-O) of the thiosulfate ion or the δasym(O-S-O) of the

tetrathionate ion. The asymmetric band near 650 cm-1 was composed of two strong peaks

centered near 620 and 660 cm-1. The first peak at 620 cm-1 was assigned to a δasym(O-S-O) vibration

of either the sulfate or sulfite ion18, 19. The peak at 660 cm-1 was assigned to the δsym(S-O) of the

thiosulfate ion or the δsym(O-S-O) of the trithionate ion.

100 200 300 400 500 600 700

δ(SS

) S3O

2- 6

ν sym(S

S) S

2O2- 3

δ asym

(SO

) S2O

2- 3 / δ as

ym(O

SO) S

4O2- 6

δ asym

(OSO

) SO

2- 4 / ν sy

m S

O2- 3

δ sym(S

O) S

2O2- 3

/ δ sy

m(O

SO) S

3O2- 6

Raman Shift / cm-1

δ(S-

S-S)

S8

ν(A

u-S)

/ ν(

Cu-S

)

700 800 900 1000 1100

ν sym S

O2- 3

ν sym(S

O) S

2O2- 3

Raman Shift / cm-1

Page 103: An Electrochemical and SERS Study of the Gold-Thiosulfate

87

The final region of interest, from 900 to 1100 cm-1, displayed two bands at 950 and 1010

cm-1. In conjunction with the peak at 620 cm-1, its most likely that the peak at 950 cm-1

corresponds to the νsym(S-O) of sulfite at the interface rather than sulfate19. As was seen in the bulk

solution, the most intense peak for sulfate should be found near 980 cm-1, but is distinctly lacking

in this spectrum18. Thus, it can be concluded that sulfate is not a species responsible for

passivation of the interface, at least at low exposure times. Based on literature, the strong peak at

1010 cm-1 can be assigned to the gold-thiosulfate complex, Au(S2O3)23! , with the downward shift

in peak position from the aqueous spectrum likely resulting from adsorption to the gold

surface29.

Throughout the duration of the experiment, peaks shifted to both higher and lower

wavenumbers from their initial positions. A shift to higher wavenumber could be the result of a

change in the ionic strength of the solution, or through a change in the dielectric constant of the

SERS active surface as a result of the leaching process30. For example, at longer exposure times,

the 285 cm-1 peak shifted to higher wavenumber centered closer 305 cm-1. However, it is well

documented that this shift occurs as a result of an increase in coverage of the gold surface20, a

process that significantly changes the local dielectric environment of the nanorod substrate31.

Peak shifting was also accompanied by the appearance and disappearance of bands. As

exposure time increased, a strong peak at 469 cm-1 developed, along with a weaker band near

140 cm-1. Both of these peaks may be assigned as vibrations of elemental sulfur (S8); specifically

as ν(S-S) and δ (S-S-S) modes, respectively. 28

Page 104: An Electrochemical and SERS Study of the Gold-Thiosulfate

88

Deconvolution and peak fitting was also carried out on the ammonium thiosulfate system;

the fitted spectrum of a gold nanorod electrode after 15 minutes of exposure can be seen in

Figure 5.17.

Figure 5.17: Fitted SERS spectrum of a gold nanorod electrode, after 15 minutes of exposure to a 0.1 M

(NH4)2S2O3 + 0.01 M CuSO4 leaching solution, with an initial pH of 8.0-8.5.

Below 670 cm-1, peak positions were similar to the initial spectrum of the calcium

thiosulfate system. Peaks were identified with positions of 216, 255, 300, 400, 443, 530, 610 and

640 cm-1. Assignment of these bands was the same as in the calcium thiosulfate system,

indicating that the passive layer after 15 minutes contained polymeric sulfur, a variety of either

Au-S or Cu-S bonds, trithionate, thiosulfate and either sulfate or sulfite. An additional band at

690 cm-1 was noted, which was tentatively assigned as a vibration of dithionate. Although

literature states that dithionate may not be formed from the interaction of polythionates32, it can

be generated via oxidation of hydrogen sulfite (bisulfite) or sulfite by Cu(II) 3, 32. Since the

leaching solution itself constitutes an oxidizing environment, and the presence of sulfite is

possible, as indicated by the spectra, it is possible that dithionate was generated in this system.

100 200 300 400 500 600 700

δ sym(S

O) S

2O2- 3

/ δ sy

m(O

SO) S

3O2- 6

δ asym

(OSO

) SO

2- 4 / ν sy

m S

O2- 3

δ asym

(SO

) S2O

2- 3 / δ as

ym(O

SO) S

4O2- 6

ν sym(S

S) S

2O2- 3

δ(SS

) S3O

2- 6

ν(A

u-S)

/ ν(

Cu-

S)

δ(SS

S) S

8

Raman Shift / cm-1

900 1000 1100 1200 1300 1400

ν sym(C

O2)N

H2C

OO

-

ν asym

(CO

) CO

2- 3 /

δas

ym(N

H2)

NH

+ 4

ν asym

(SO

) S3O

2- 6 /

HSO

- 3

ν sym(S

O) A

u-S 2O

2- 3

ν SO

3NH

- 2

ν sym(S

O) S

2O2- 3

Raman Shift / cm-1

ν SO

2 / S

O3N

H- 2

Page 105: An Electrochemical and SERS Study of the Gold-Thiosulfate

89

Significant differences were noted in the SERS spectra of the ammonium and calcium

thiosulfate systems, above 1000 cm-1. An asymmetric band centered near 1000 cm-1 was

composed of two peaks positioned at 998 and 1011 cm-1 that were assigned to the νsym(S-O) of

thiosulfate ion, and a vibration of the gold-thiosulfate complex, respectively. Five additional

peaks were observed between 1200 and 1500 cm-1 at positions of 1237, 1267, 1351, 1404 and

1443 cm-1. The first peak at 1237 cm-1 was assigned as either a νasym(S-O) of trithionate or bisulfite.

Since both of these species have corresponding vibrations in the rest of the spectrum, further

identification is not possible. Peaks located at 1267 and 1351 cm-1 matched literature values for

vibrations of the sulfamate ion, produced through reaction of trithionate with ammonia33:

S3O62! + 2NH3 ! S2O3

2! + SO3NH2! + NH4

+ (5.11)

Such a reaction pathway indicates that decomposition of trithionate at the interface was

occurring, while regenerating thiosulfate. The remaining peaks at 1404 and 1443 cm-1 were

assigned as symmetric vibrations of the carbamate, and carbonate ions, respectively.

The final system analyzed via deconvolution and peak fitting was the sodium thiosulfate

system with copper. The fitted spectrum of a gold nanorod electrode after five minutes of

exposure to the leaching solution can be found in Figure 5.18.

In the first five minutes of exposure, 7 peaks were identified at positions of 256, 300,

391, 443, 525, 613 and 1004 cm-1. As in the previous two systems, these bands were assigned to

ν(Cu-S) or ν(Au-S), δ(S-S) mode of trithionate, νsym(S-S) of thiosulfate, δasym(S-O) of the thiosulfate ion or

the δasym(O-S-O) of the tetrathionate ion, δasym(O-S-O) vibration of either the sulfate or sulfite ion and

νsym(S-O) of thiosulfate ion, respectively. An interesting note here is the absence of elemental

sulfur, which was observed in the initial spectra of both the calcium and ammonium thiosulfate

systems.

Page 106: An Electrochemical and SERS Study of the Gold-Thiosulfate

90

Figure 5.18: Fitted SERS spectrum of a gold nanorod electrode, after 5 minutes of exposure to a 0.1 M

Na2S2O3 + 0.01 M CuSO4 leaching solution, with an initial pH of 8.0-8.5.

In each of the three thiosulfate systems investigated, six peaks were common amongst the

Raman spectra; vibrational modes corresponding to δ(S-S-S) of S8 (~ 216 cm-1), ν(Cu-S) or ν(Au-S)

vibrations (at ~ 255 and 300 cm-1), δ(S-S) mode of trithionate (~ 400 cm-1), νsym(S-S) of thiosulfate

and the δasym(O-S-O) vibration of either the sulfate or sulfite ion were identified in the three systems

of interest, and their peak areas tracked as a function of exposure time of the gold nanorod

electrodes to the respective leaching solutions. Changes in the peak areas are directly related to

changes in the composition of the passive layer.

Analysis of the peaks’ area as a function of time was performed by normalizing the peak

area at time ‘t’, with respect to the initial value of the area. Thus, plotted values can be viewed as

a percent change. This normalization allows for a semi-quantitative comparison of the rate of

change in area between the 6 peaks of interest, and a comparison of changes in band intensity

100 200 300 400 500 600 700

Raman Shift / cm-1

ν(A

u-S)

/ ν(

Cu-

S)

δ(SS

) S3O

2- 6

ν sym(S

S) S

2O2- 3

δ asym

(SO

) S2O

2- 3 / δ as

ym(O

SO) S

4O2- 6

δ asym

(OSO

) SO

2- 4 / ν sy

m S

O2- 3

900 1000 1100 1200 1300 1400 1500

Raman Shift / cm-1

ν sym(S

O) S

2O2- 3

Page 107: An Electrochemical and SERS Study of the Gold-Thiosulfate

91

between the 3 electrolytes. Data for the calcium, ammonium, and sodium thiosulfate systems are

presented in Figures 5.19 to 5.21.

Two panels were used to present the peak analysis data; panel (a) groups data that display

large changes in concentration in the first 100 minutes of exposure, and panel (b) contains data

that show small changes in concentration, either in the first 100 minutes of exposure, or over the

duration of the experiment. Panel (b) in Figures 5.19-5.21 also display the open circuit potential,

discussed and shown previously in Figure 5.14, for each of the 3 leaching systems.

Page 108: An Electrochemical and SERS Study of the Gold-Thiosulfate

92

Figure 5.19: Normalized analytical peak areas for a gold nanorod electrode treated with 0.1 M CaS2O3 +

0.01 M CuSO4 solution adjusted to pH 8.0-8.5. Peak areas tracked were for band positions of a) 216 cm-1

(green circle), 255 cm-1 and 300 cm-1 (brown square), and b) 400 cm-1 (blue circle), 443 cm-1 (red square),

and 610 cm-1 (pink diamond).

0 50 100 150 200 250 3000

4

8

12

16

20

24

28

32

Nor

mal

ized

Pea

k A

rea

Time / minutes

ν(Au-S/Cu-S) [255 cm-1 + 305 cm-1]

δ(SSS) S8

0 50 100 150 200 250 3000

4

8

12

16

20

Nor

mal

ized

Pea

k A

rea

Time / minutes

δ(SS) S3O2-6

νsym(SS) S2O2-3

νsym SO2-3

-0.12

-0.10

-0.08

-0.06

-0.04

-0.02

0.00

Ave

rage

Ope

n C

ircui

t Pot

entia

l / V

a)

b)

Page 109: An Electrochemical and SERS Study of the Gold-Thiosulfate

93

Over the first hour of exposure in the calcium thiosulfate system, the amount of elemental

sulfur at the interface significantly increased to a maximum, approximately 16 times the initial

concentration. From 60 minutes to 120 minutes, almost 100% of the sulfur at the interface was

removed. Deposition was noted through the remainder of the experiment, resulting in a final

concentration approximately 2.5 times the initial concentration.

Absolute identification of the peaks at 255 cm-1 and 305 cm-1 as either Au-S or Cu-S

interactions was not possible based solely upon the SERS spectra acquired. However, each of

these bands contains information as to the level of passivation at the interface. Thus, the sum of

the areas of both of these peaks should represent the total number of Au-S or Cu-S interactions

making up the passive layer. The change in the normalized area of this sum provides an

indication of the level of passivation.

The change in the normalized area of the Au-S/Cu-S interactions paralleled that of

elemental sulfur at the interface, especially in the first hour of exposure. A maximum value of

approximately 24 times the initial concentration was reached after 90 minutes of exposure. From

approximately 120-300 minutes the normalized area decayed significantly, reaching a final value

approximately 8 times the initial. The large spike at 90 minutes of exposure was most likely the

result of a sudden increase in the surface enhancement from the gold nanorod substrate. The

leaching process is a destructive process, and the sudden increase in enhancement may arise

from a temporary ‘lightning rod’ effect resulting from the collapse of nanorods, bringing them

within 1nm of each other. 34, 35

A significant decrease in the amount of trithionate at the interface occurred in the first

hour of exposure, with almost 100% of the initial value removed either through decomposition or

desorption of trithionate from the gold surface. A spike in the amount of trithionate at the

Page 110: An Electrochemical and SERS Study of the Gold-Thiosulfate

94

interface was observed after 90 minutes. By 120 minutes, the concentration of trithionate had

returned to a level on par with the initial surface concentration; the final concentration at the

interface was approximately the same level as the initial.

Thiosulfate at the interface experienced an increase in the surface concentration (to

approximately 4 times its initial value) during the first hour of exposure. This may be a result of

the decomposition of trithionate36, 37:

2S3O62! + 6OH! ! S2O3

2! + 4SO32! + 3H2O (5.12)

At the same time, decomposition of thiosulfate to form elemental sulfur was occurring,

providing a competitive pathway between production and decomposition of thiosulfate at the

interface2:

2S2O32! + H2O ! 2SO4

2! + 4S + OH! (5.13)

3S2O32! + 6OH! ! 4SO3

2! + 2S2! + 3H2O (5.14)

S2O32! ! SO3

2! + S0 (5.15)

The thiosulfate concentration at the interface over the remaining time of the experiment

was observed to increase greatly, with a final concentration approximately 20 times the initial.

The behavior of sulfite in the first 90 minutes of exposure was similar to that of

thiosulfate; an increase to approximately 4 times the initial concentration was observed. This

production is further support of Equation 5.12. Sulfite at the interface experienced a significant

decay from 90 minutes to approximately 260 minutes of exposure; such an effect could be

produced by oxidation of sulfite by Cu(II) to form sulfate23:

2Cu2+ + SO32! + OH! ! 2Cu2+ + SO4

2! + H2O (5.16)

However, sulfate was only observed in solution and not at the interface. Therefore, if

sulfate is being formed as a by-product of sulfite oxidation, or other decomposition reactions, it

Page 111: An Electrochemical and SERS Study of the Gold-Thiosulfate

95

must be rapidly transported to the bulk upon formation. The final concentration of sulfite was

approximately the same as the initial value.

Figure 5.19b also displays the open circuit potential of the calcium thiosulfate system as a

function of exposure time. While the time scale of the SERS leaching experiment and open

circuit potential do not directly overlap, some correlation can be drawn between the species

present and the rate of passivation. The OCP potential curve displays a minimum, followed by an

increase between 15 and 60 minutes of exposure. During this time, an increase in elemental

sulfur was observed, along with a decrease in the concentration of trithionate; removal and

deposition of these species may be responsible for the curvature in this region of the potential

plot. Although almost all trithionate was removed in the first hour, the concentration of elemental

sulfur increases to approximately 16 times its initial concentration. Hence, the open circuit

potential experienced a steep rise as the surface became extensively passivated. Between 90 and

120 minutes, a plateau is observed in the potential curve. This corresponds to the achievement of

a steady state concentration of elemental sulfur, and trithionate. Such a result could indicate that

the rate of passivation depends on a combination of changes in the surface concentration of both

trithionate and elemental sulfur in this system.

Peak area analysis of the interfacial species for the ammonium thiosulfate system can be

found in Figure 5.20 (a) and (b). The overall changes in band intensity were much smaller in the

ammonium thiosulfate system than the corresponding changes in the calcium thiosulfate system.

The data are presented in two panels that plot the same normalized band intensities as panels (a)

and (b) in Figure 5.19.

Page 112: An Electrochemical and SERS Study of the Gold-Thiosulfate

96

Figure 5.20: Analytical peak areas for a gold nanorod electrode treated with 0.1 M (NH4)2S2O3 + 0.01 M

CuSO4 solution adjusted to pH 8.0-8.5. Peak areas tracked were for band positions of: a) 216 cm-1 (green

circle), 255 cm-1 and 300 cm-1 (brown square), and b) 400 cm-1 (blue circle), 443 cm-1 (red square), and

610 cm-1 (pink diamond).

0 20 40 60 80 100 120 140 160 180 2000

4

8

Nor

mal

ized

Pea

k A

rea

Time / minutes

ν(Au-S/Cu-S) [255 cm-1+ 305 cm-1]

δ(SSS) S8

0 20 40 60 80 100 120 140 160 180 2000

1

2

3

4

Nor

mal

ized

Pea

k A

rea

Time / minutes

δ(SS) S3O2-6

νsym(SS) S2O2-3

δasym(OSO) SO2-4 / νsym SO2-

3

-0.12

-0.10

-0.08

-0.06

-0.04

-0.02

0.00

Ave

rage

Ope

n C

ircui

t Pot

entia

l / V

a)

b)

Page 113: An Electrochemical and SERS Study of the Gold-Thiosulfate

97

More scatter was observed in the ATS data (Figure 5.20), however, the scale is expanded

which may contribute to the random appearance of the data. This scatter may be due to

fluctuations in the surface enhancement as a result of physical damage to the nanorods caused by

dissolution during the experiment.

The concentration of elemental sulfur at the gold-thiosulfate interface in the ammonium

thiosulfate system increases significantly over the course of the experiment, achieving a

maximum value approximately 9 times the initial concentration. Removal of sulfur from the

interface was observed over some periods of the experiment. This resulted in a final

concentration approximately 4 times the initial.

The normalized area for Au-S/Cu-S interactions at the surface increased over the duration

of the experiment, reaching a maximum value of 3 times the initial area. The final normalized

area at the interface is approximately 2 times the initial value.

Allowing for scatter, the data in Figure 5.20 (b) show a general increase in all peak

intensities over the course of the experiment, with the exception of the δ(S-S) mode of trithionate.

Over the first hour of the experiment the concentration of trithionate at the surface decayed. A

large spike was observed after 90 minutes, as in the calcium thiosulfate system. Again, this spike

may arise from a sudden increase in enhancement arising from a temporary lightning rod effect34,

35. Decay of trithionate at the interface continued from 90 minutes onward, resulting in a final

concentration one-tenth the initial. The final values for thiosulfate and sulfite at the interface

were 3 times, and on par with the initial values, respectively.

The open circuit potential of the ammonium thiosulfate system (Figure 5.20 b) displays a

much slower rate of passivation at longer exposure times than the calcium thiosulfate system.

This may be the result of constant removal of trithionate from the interface, and alternating

Page 114: An Electrochemical and SERS Study of the Gold-Thiosulfate

98

periods of deposition and removal of elemental sulfur at the interface; a slower, and smaller,

change in the amount of all adsorbed species may also contribute. Because neither trithionate nor

sulfur was completely removed, or achieved a steady state, passivation of the surface continued;

hence, the OCP of the system moved consistently towards more positive potentials.

Decay of trithionate at the interface may be the result of interaction with ammonia,

resulting in the production of sulfamate (Equation 5.11), which was observed at the interface at

early exposure times. This reaction also produces thiosulfate, which may contribute to the

periodic behavior observed for this species.

The composition of the passive layer formed on the gold surface in the sodium thiosulfate

system displayed much different behavior over the duration of the experiment than in the

previous two systems. Normalized peak area analysis for the sodium thiosulfate system can be

found in Figure 5.21 (a) and (b). Data in both panels displayed a sharp increase at 80 minutes of

exposure time. Assuming that this increase is due to a temporary lightning rod effect, as in the

calcium and ammonium thiosulfate systems, the data in Figure 5.21 (b) show a continuous quasi-

linear increase of all band intensities with time. The number of Au-S/Cu-S interactions at the

interface reached a final value approximately 18 times the initial. Elemental sulfur at the

interface also experienced a large increase, reaching a final value approximately 9 times the

initial. Final values for thiosulfate, trithionate were approximately two times the initial

concentration, and sulfite experienced a 3-fold increase over the duration of the experiment.

Page 115: An Electrochemical and SERS Study of the Gold-Thiosulfate

99

Figure 5.21: Analytical peak areas for a gold nanorod electrode treated with 0.1 M Na2S2O3 + 0.01 M

CuSO4 solution adjusted to pH 8.0-8.5. Peak areas tracked were for band positions of: a) 216 cm-1 (green

circle), 255 cm-1 and 300 cm-1 (brown square), and b) 400 cm-1 (blue circle), 443 cm-1 (red square), and

610 cm-1 (pink diamond).

0 20 40 60 80 100 120 140 160 180 2000

4

8

12

16

20

Nor

mal

ized

Pea

k A

rea

Time / minutes

ν(Au-S/Cu-S) [255 cm-1 + 305 cm-1]

δ(SSS) S8

0 20 40 60 80 100 120 140 160 180 2000

4

8

12

Nor

mal

ized

Pea

k A

rea

Time / minutes

νsym(SS) S2O2-3

δ(SS) S3O2-6

δasym(OSO) SO2-4 / νsym SO2-

3

-0.12

-0.08

-0.04

0.00

0.04

Ave

rage

Ope

n C

ircui

t Pot

entia

l / V

a)

b)

Page 116: An Electrochemical and SERS Study of the Gold-Thiosulfate

100

The sodium thiosulfate system showed the highest level of passivation at early exposure

times, as shown by the negative limit in the open circuit potential of the system in Figure 5.14

(the OCP is also displayed in Figure 5.21 (b)). However, it also displayed the slowest rate of

change over the duration of the leaching experiment; a linear increase in the OCP was observed

from 25-180 minutes of exposure. A gradual linear increase was also observed in the behavior of

trithionate, sulfite, thiosulfate, and up to 160 minutes, sulfur. Because no single species displays

behavior solely related to the trend in the open circuit potential, it can be concluded that in the

sodium thiosulfate system the passive layer is complex, and no single species can be assigned

responsibility for passivation of the gold surface.

It has been proposed that Cu(II) may aid in gold dissolution through the removal of sulfur

contaminants from the surface by scavenging sulfide, or through interaction with sulfur chains on

the surface6. If, in the work presented in this chapter, the peak observed at ~305 cm-1 in all 3

systems corresponds to a Au-S vibration28, then the peak at ~255 cm-1 could correspond to Cu-S

vibrations. Increases observed in the area of the Cu-S peak may have been a result of interaction

of Cu(II) with sulfur on the gold surface, and the observed decrease in the calcium and

ammonium thiosulfate systems could correspond to subsequent diffusion from the interface to

the bulk solution. This is possible, as leaching solutions left for several hours after preparation

reproducibly precipitated black copper sulfides.

Peaks corresponding to characteristic vibrations of tetrathionate were not observed in the

SERS spectra collected, thus, it can be concluded that in the presence of Cu(II) in these systems,

tetrathionate is not present at the interface, or if it is formed, undergoes rapid decomposition.

Page 117: An Electrochemical and SERS Study of the Gold-Thiosulfate

101

5.4 SUMMARY

The composition and behavior of species in the bulk leaching solution, and at the gold

thiosulfate interface in the presence of copper, were characterized through complementary

electrochemical and spectroscopic techniques. The bulk solutions for the calcium, ammonium

and sodium thiosulfate systems all maintained steady state concentrations of thiosulfate,

tetrathionate, sulfate, and trithionate. These systems were unaffected by the changes in pH

caused by carbon dioxide absorption and precipitation of calcium carbonate (or calcium sulfate).

Solubility values for each of these salts indicate that it is most likely calcium carbonate that is

precipitating from solution.

The gold-thiosulfate interface showed large differences between the three systems.

Tetrathionate and sulfate were found to be absent in each system at the interface; however,

elemental sulfur, and sulfite were present. The negative limits of the open circuit potentials

indicated that the level of passivation was most significant in the sodium thiosulfate system,

followed by calcium and ammonium thiosulfate, respectively. This result was also reflected in

the SERS results. Table 5.2 displays the final change in the normalized peak area of the six

bands analyzed in this thesis (to δ(S-S-S) of S8 [~ 216 cm-1], ν(Cu-S) or ν(Au-S) vibrations [at ~ 255

and 300 cm-1], δ(S-S) mode of trithionate [~ 400 cm-1], νsym(S-S) of thiosulfate and the δasym(O-S-O)

vibration of either the sulfate or sulfite ion), for the calcium, ammonium, and sodium thiosulfate

systems.

Page 118: An Electrochemical and SERS Study of the Gold-Thiosulfate

102

Table 5.2: Final change in the normalized peak areas investigated in the calcium, ammonium and

thiosulfate systems.

Overall Change in the Normalized Peak Area

Peak Calcium Thiosulfate Ammonium Thiosulfate Sodium Thiosulfate

δ(S-S-S) of S8 [~ 216 cm-1] 2.6 4.6 9.4

ν(Cu-S) / ν(Au-S) [~ 255 and 300 cm-1] 8.0 2.1 17.8

δ(S-S) of S3O62-

[~ 400 cm-1] 1.2 0.1 2.4

νsym(S-S) of S2O32-

[~ 445 cm-1] 19.8 3.2 2.4

δasym(O-S-O) of SO32-/SO4

2- [~ 610 cm-1] 1.4 0.8 3.1

The overall change in the normalized areas are representative of change in the passive

layer at the gold surface over the duration of the experiment. In terms of the relative increase

from the initial concentration at the surface, the sodium thiosulfate system experienced the

greatest overall increase in the amount of Au-S or Cu-S interactions at the interface, followed by

the calcium, and ammonium thiosulfate systems, respectively. A higher level of passivation (or

larger number of Au-S/Cu-S interactions at the surface) should result in a lower leaching rate.

The results in Table 5.2 match what is expected based on the results from literature21. The

calcium thiosulfate and sodium thiosulfate systems show significant differences in the behavior

of thiosulfate at the interface. An increase to 2.4 times the initial concentration was noted in the

sodium thiosulfate system; however, a massive increase in calcium thiosulfate system was

observed, with the final concentration almost 20 times the initial. Such a large concentration of

thiosulfate at the interface should lead to promotion of leaching.

Page 119: An Electrochemical and SERS Study of the Gold-Thiosulfate

103

Large amounts of sulfur were present at the interface at early exposure times in both the

calcium and ammonium systems; however, each of these two systems displayed a significant

ability for removal of sulfur from the interface (final values for the concentration at the interface

for the calcium and ammonium systems are 2.6 and 4.6 times the initial concentrations,

respectively). Comparably, trithionate was observed to decompose in both of these systems as

well. Almost 100% removal occurred over the course of the experiment in the ammonium

thiosulfate system (the final surface concentration was one-tenth the initial), while the calcium

thiosulfate system displayed decomposition only at early exposure times (final surface

concentration was 1.2 times the initial). The combined presence of both elemental sulfur and

trithionate at the interface in calcium thiosulfate may explain the higher level of passivation

when compared with the ammonium thiosulfate system, where elemental sulfur appears to be the

main component of the passive layer. The increased concentration of Cu(II) as a result of

stabilization by NH3 in the ammonium thiosulfate solution should lead to increased leaching, and

minimize degradation of thiosulfate.

Previous studies13, 20 of the gold-thiosulfate interface in the absence of copper displayed

evidence of thiosulfate and tetrathionate in the passive layer at early exposure times. Extended

exposure showed that tetrathionate at the interface degraded, forming trithionate and thiosulfate;

such behavior reflects known solution reactions. Normalized peak area analysis of the work by

Baron13 can be found in Figure 5.22.

Page 120: An Electrochemical and SERS Study of the Gold-Thiosulfate

104

Figure 5.22: Normalized analytical peak area of a gold nanorod electrode treated with 0.1 M Na2S2O3

solution adjusted to pH 10. Modified from13.

In the presence of copper, an enhanced degradation rate of higher polythionate species

was noted. Also, elemental sulfur was identified at early exposure times at the interface in some

systems. Thus, copper appears to be aiding in removal of polythionates from the interface, but

may also be contributing to enhanced degradation of thiosulfate to form elemental sulfur, a

species which seems to contribute to passivation. Introduction of copper is both beneficial and

detrimental to the leaching system, as leaching appears to be promoted, as well as passivation.

References

1. Abbruzzese, C.; Fornari, P.; Massidda, R.; Vegliņ, F.; Ubaldini, S. Hydrometallurgy

1995, 39, 265-276. 2. Aylmore, M. G.; Muir, D. M. Miner Metall Process 2001, 18, 221-227. 3. Aylmore, M. G.; Muir, D. M. Minerals Eng 2001, 14, 135-174. 4. Byerley, J. J.; Fouda, S. A.; Rempel, G. L. J.Chem.Soc., Dalton Trans. 1973, , 889-893. 5. Pedraza, A. M.; Villegas, I.; Freund, P. L.; Chornik, B. J Electroanal Chem 1988, 250,

443-449. 6. Zhang, S. C.; Nicol, M. J. J. Appl. Electrochem. 2005, 35, 339-345.

Page 121: An Electrochemical and SERS Study of the Gold-Thiosulfate

105

7. Breuer, P. L.; Jeffrey, M. I. Hydrometallurgy 2002, 65, 145-157. 8. Baron, J.; Szymanski, G.; Lipkowski, J. J Electroanal Chem 2011, 662, 57-63. 9. Meyer, B.; Ospina, M. Phosphorus Sulfur and Silicon and the Related Elements 1982, 14,

23-36. 10. Sato, S.; Higuchi, S.; Tanaka, S. Appl. Spectrosc. 1985, 39, 822-827. 11. Haigh, J.; Hendra, P.; Rowlands, A.; Degen, I.; Newman, G. Spectrochim. Acta, Pt. A:

Mol. Spectrosc. 1993, 49, 723-725. 12. Fanigliulo, A.; Bozzini, B. Trans. Inst. Met. Finish. 2002, 80, 132-136. 13. Baron Gavidia, J. Study of the Gold-Thiosulfate Interface Under Leaching Conditions,

University of Guelph, Guelph, ON, 2010. 14. Antohe, V. A.; Radu, A.; Matefi-Tempfli, M.; Attout, A.; Yunus, S.; Bertrand, P.; Dutu,

C. A.; Vlad, A.; Melinte, S.; Matefi-Tempfli, S.; Piraux, L. Appl. Phys. Lett. 2009, 94, 073118.

15. Chang, J. C. In Section 8: Analytical Chemistry; Lide, D. R., Ed.; CRC Handbook of Chemistry and Physics; CRC Press: Boca Raton, 1990; pp 8-39.

16. Surewicz, W. K.; Mantsch, H. H. Biochim. Biophys. Acta 1988, 952. 17. Meyer, B.; Ospina, M. Abstracts of Papers of the American Chemical Society 1982, 183,

56-INOR. 18. Daly, F. P.; Brown, C. W.; Kester, D. R. J. Phys. Chem. 1972, 76, 3664-3668. 19. Evans, J. C.; Bernstein, H. J. Canadian Journal of Chemistry-Revue Canadienne De

Chimie 1955, 33, 1270-1272. 20. Jeffrey, M.; Watling, K.; Hope, G. A.; Woods, R. Minerals Eng 2008, 21, 443-452. 21. Chandra, I.; Jeffrey, M. I. Hydrometallurgy 2004, 73, 305-312. 22. Jeffrey, M. I.; Linda, L.; Breuer, P. L.; Chu, C. K. Minerals Eng 2002, 15, 1173-1180. 23. Molleman, E.; Dreisinger, D. Hydrometallurgy 2002, 66, 1-21. 24. Arima, H.; Fujita, T.; Yen, W. Miner Metall Process 2003, 20, 81-92. 25. Wan, R. Y.; LeVier, K. M. Int. J. Miner. Process. 2003, 72, 311-322. 26. Senanayake, G. Hydrometallurgy 2004, 75, 55-75. 27. Senanayake, G. Hydrometallurgy 2005, 77, 287-293. 28. Jeffrey, M. I.; Watling, K.; Hope, G. A.; Woods, R. Minerals Eng 2008, 21, 443-452. 29. Watling, K.; Hope, G. A.; Jeffrey, M. I.; Woods, R. ECS Transactions 2006, 2, 121-132. 30. Moskovits, M. J. Raman Spectrosc. 2005, 36, 485-496. 31. Jain, P. K.; Huang, X.; El-Sayed, I. H.; El-Sayed, M. A. Plasmonics 2007, 2, 107-118. 32. Nickless, G. In Inorganic sulphur chemistry; Elsevier Publishing Company: 1968; . 33. Naito, K.; Hayata, H.; Mochizuki, M. Journal of Inorganic and Nuclear Chemistry 1975,

37, 1453-1457. 34. Tian, Z. Q.; Ren, B.; Wu, D. Y. J Phys Chem B 2002, 106, 9463-9483. 35. Stiles, P. L.; Dieringer, J. A.; Shah, N. C.; Van Duyne, R. R. Annual Review of Analytical

Chemistry 2008, 1, 601-626. 36. Zhang, H.; Jeffrey, M. I. Inorg. Chem. 2010, 49, 10273-10282. 37. Varga, D.; Horváth, A. K. Inorg. Chem. 2007, 46, 7654-7661.

Page 122: An Electrochemical and SERS Study of the Gold-Thiosulfate

106

CHAPTER 6: CONCLUSIONS AND FUTURE WORK

6.1 CONCLUSIONS

Overall this thesis has described the composition and behavior of the bulk solutions, and

the passive layer that forms on the gold surface, in three different thiosulfate-leach solutions

containing copper. In an attempt to identify passivating species at the interface, the open circuit

potentials and pH of all three systems have been characterized along with the kinetics of the

electron transfer reaction. This work was supported by NSERC and Barrick Gold Inc.

Beginning with the characterization of the bulk solutions of interest (calcium, ammonium

and sodium thiosulfate with Cu(II)), some conclusions can be drawn. Firstly, the identity of the

cation in thiosulfate salt has no effect on the composition of the bulk solution; all three solutions

displayed identical compositions. In each solution thiosulfate, trithionate, tetrathionate and

sulfate were identified. Secondly, the time dependent concentration of these species is stable, and

independent of pH. Each of the three solutions displayed steady state concentrations of

thiosulfate, tetrathionate, trithionate, and sulfate over a period of 3 hours. This is likely an effect

of Cu(II) in solution, as it is known to catalyze decomposition of both thiosulfate and

polythionates1, leading to regeneration of both.

At the gold-thiosulfate interface, in the presence of copper, little correlation to the

characteristics of the bulk solution was observed. Species identified at the interface included

elemental sulfur, trithionate, thiosulfate, and sulfite. Polythionates higher than trithionate were

not observed. The absence of higher polythionates is likely an effect of the local pH of the

interface. The cation of the thiosulfate salt was found to have no effect on the initial leaching rate

Page 123: An Electrochemical and SERS Study of the Gold-Thiosulfate

107

of gold; however, significant differences were noted in the open circuit potential curves of a gold

disc electrode exposed to the calcium, ammonium and sodium thiosulfate solutions. This trend in

the open circuit potentials indicates that there are differing levels and rates of passivation

between the three systems. The time dependent composition of the passive layer was drastically

different between calcium, ammonium and sodium thiosulfate systems. Both the calcium and

ammonium systems displayed the ability to remove trithionate from the interface, as well as

elemental sulfur. This was not seen in the sodium system, which displayed a slower rate of

passivation, once the negative limit was reached. However, Cu(II) has been proposed to

scavenge sulfur contaminants from the gold surface1; this interaction maybe responsible for the

removal elemental sulfur observed. Thus, more work is needed to confirm whether the behavior

of elemental sulfur at the interface is an effect of the cation, or copper.

One common feature between the systems was the behavior of trithionate at the interface.

In all 3 systems, a large spike in the concentration of trithionate is observed between 75-90

minutes exposure of the gold electrode to leaching solution. Rapid removal or decomposition of

trithionate was then observed, returning the concentration at the interface to near pre-spike

values. This large spike is likely the result of a temporary lightning rod effect created as a result

of dissolution of the gold nanorod electrode.

6.2 FUTURE WORK

Potential roles of both copper and the cation of the thiosulfate salt have been proposed in

the scope of this thesis. However, further comparative studies are required before definitive

conclusions can be drawn. In order to determine if the cation of the thiosulfate salt is having a

significant effect on the time dependent composition of the passive layer, electrochemical

measurements of the leaching current as a function of time should be performed. If the cation

Page 124: An Electrochemical and SERS Study of the Gold-Thiosulfate

108

effect arises from the extent or rate of passivation, a significant difference in leaching currents

should be noted for a non-passivated and sequentially passivated electrode surface.

Although open circuit potentials for the calcium, ammonium and sodium thiosulfate

systems in the presence of copper have been recorded, a polished gold disk electrode was

employed. This “smooth” gold electrode had a significantly different geometry than the gold

nanorod electrode employed in the SERS studies. Differing geometries could lead to different

rates of mass transport, and subsequently different rates of adsorption. Thus, measurement of the

open circuit potential of the gold nanorod electrode directly would allow for correlation of SERS

results and the open circuit potentials of the systems. It would allow for tracking peak shifts

within the SERS spectrum, giving insight into the adsorption and desorption characteristics of

each species. As well, SERS studies for all three salts must be performed in the absence of

copper in the leaching solution. These studies have been executed on the sodium thiosulfate

system, however, characterization of the ammonium and calcium thiosulfate systems must be

completed. This will clarify the interaction of copper with adsorbed elemental sulfur. If Cu(II) is

unable to scavenge sulfur or sulfide from the surface, a constant buildup would be expected in

the copper free experiments, in all solutions.

Addition of ammonia to each of the leaching solutions, in the absence of copper would

allow for an understanding of the individual effect of this additive. SERS results may indicate

whether ammonia/ammonium is adsorbing at the interface. Once complete, SERS studies of the

leaching solutions with both copper and ammonia can be performed. In this way, the individual

and tandem role of both additives can be identified.

Extending both SERS and solution Raman studies beyond the exposure limit within this

work would greatly enhance the industrial applicability of the results, as industrial leaching

Page 125: An Electrochemical and SERS Study of the Gold-Thiosulfate

109

processes are extremely long-scale (eg. up to 90 days). Identifying the end products at the

interface, and bulk solution would allow for greater understanding and optimization of the leach

system.

References

1. Zhang, S. C.; Nicol, M. J. J. Appl. Electrochem. 2005, 35, 339-345.