Sunteți pe pagina 1din 321

The Effect of Bentonite on External

Corrosion of Well Casings

A thesis submitted to The University of Manchester for the degree of


Doctor of Philosophy
in the Faculty of Engineering and Physical Sciences

2012

Mohammed M. Orayith

School of Materials
Corrosion and Protection Centre
List of contents

Abstract .......6
Declaration ..7
Copyright .....8
Dedication.....9
Acknowledgments .... 10
1 Introduction ...........................................................................................................12
2 Fundamentals of Corrosion..................................................................................15
2.1 Definition of Corrosion ...................................................................................15
2.2 Forms of Corrosion .........................................................................................16
2.2.1 Uniform Corrosion ..................................................................................16
2.2.2 Crevice Corrosion ...................................................................................17
2.2.3 Pitting Corrosion .....................................................................................18
2.2.4 Galvanic Corrosion .................................................................................19
2.3 Electrochemical Nature of Corrosion..............................................................19
2.4 Thermodynamics and Kinetics of Corrosion ..................................................21
2.4.1 Background .............................................................................................21
2.4.2 Gibbs Free Energy and Electrode Potential ............................................22
2.4.3 Application of Thermodynamics to Corrosion .......................................24
2.4.4 Kinetics of Corrosion ..............................................................................25
2.4.4.1 Faradays Law.....................................................................................25
2.4.4.2 Activation Polarization........................................................................26
2.4.4.3 Concentration Polarization..................................................................28
2.4.4.4 Resistance Polarization .......................................................................28
2.5 Corrosion Products and Passivity....................................................................29
2.5.1 Types of Corrosion Products on the Metal Surfaces...............................31
2.5.2 Calcareous Deposit..................................................................................32
Chapter 3 .................................................................................................................37
Literature Review: Part Two...................................................................................37
Well Casing.............................................................................................................37
3 Well Casing ............................................................................................................38
3.1 Corrosion Phenomena and Well Casing .........................................................39
3.2 Causes of Corrosion in Well Casings..............................................................40
3.3 Cathodic Protection of Well Casing...............................................................41
3.3.1 Main Problems in Application of CP on Well Casing ............................43
3.3.2 Criteria for Cathodic Protection and Current Requirements...................44
3.3.3 Casing Potential Profile ..........................................................................45
3.4 Corrosion by Soil ............................................................................................46
3.4.1 Factors Affecting Corrosion by Soil .......................................................48
3.4.1.1 Aeration...............................................................................................48
3.4.1.2 Water Content .....................................................................................49
3.4.1.3 Soil Acidity .........................................................................................50
3.4.1.4 Salt Content.........................................................................................50
3.4.2 Soil Chemical Analysis ...........................................................................51
3.4.3 Soil Electrical Resistivity........................................................................52
3.4.3.1 Soil Resistivity Measurement .............................................................53
3.4.3.1.1 Wenner Four- Pin Method ............................................................53
3.4.3.1.2 Soil Box Method ...........................................................................54
3.4.3.2 Temperature Effect..............................................................................56
3.5 Corrosion Prevention ......................................................................................56
3.5.1 Cathodic Protection.................................................................................56
3.5.1.1 How Cathodic Protection Works ........................................................57
3.5.1.2 Cathodic Protection Systems...............................................................58
3.5.1.2.1 Sacrificial Anode System..............................................................58
3.5.1.2.2 Impressed Current Cathodic Protection System ...........................60
3.5.1.2.3 Choice of Sacrificial or Impressed Current System......................62
3.5.1.2.4 Advantages and Uses of Cathodic Protection ...............................63
3.5.1.2.5 Cathodic Protection Monitoring....................................................65
4 Bentonite ................................................................................................................69
4.1 Background .....................................................................................................69
4.2 Uses of Bentonite ............................................................................................70
4.3 Classification and Chemical Composition of Bentonite .................................71
4.3.1 Water Content in Bentonite.....................................................................73
5 Experimental Methods and Instrumentation .....................................................78
5.1 Introduction .....................................................................................................78
5.2 Na-Bentonite ...................................................................................................80
5.2.1 Bentonite Resistivity Measurements.......................................................81
5.3 Preparation of Specimens................................................................................81
5.3.1 Surface Preparation of the Specimens.....................................................84
5.3.2 Reference Electrodes...............................................................................84
5.3.3 Auxiliary Electrode .................................................................................87
5.3.4 pH Measurements....................................................................................87
5.4 Application of Cathodic Protection (CP), Using Power Supply.....................88
5.5 Preparation of Solutions and Experimental Set-Up ........................................91
5.6 Weight Loss Test.............................................................................................95
5.7 Potential -Time Experiments...........................................................................99
5.8 Linear Polarization Resistance (LPR)...........................................................100
5.8.1 Experimental Cell and LPR Data Collection ........................................104
5.9 X-Ray Diffraction (XRD) .............................................................................105
5.10 Scanning Electron Microscopy (SEM) and Energy Dispersive X-ray Analysis
spectrometry, EDAX.................................................................................................110
5.11 The Impact of Zinc chloride..........................................................................111
6 Results and Discussion........................................................................................114
6.1 Introduction ...................................................................................................114
6.2 Bare Mild Steel Buried Vertically in Bentonite Mixed with 45% (w/w) 0.5 M
NaCl Solution............................................................................................................115
6.2.1 Open Circuit Potential (OCP) ...............................................................116
6.2.2 Impressed Current Cathodic Protection Polarization Potential.............119
6.3 Bare Mild Steel Exposed to 0.5 M Sodium Chloride Solution Containing
(Free of bentonite).....................................................................................................135
6.3.1 Introduction ...........................................................................................135
6.3.2 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution
Containing No Bentonite ......................................................................................138
6.3.2.1 Optical Microscopy of Specimens Immersed in 0.5M NaCl Solution
141
6.3.3 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution with No
Bentonite under CP at -0.8 V and -1.15V .............................................................150
6.3.4 SEM/EDAX and XRD Analyses of Specimens Immersed in 0.5 M NaCl
Free of Bentonite...................................................................................................156

3
6.4 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution Mixed With
1.0% (w/w) Na-Bentonite with no Protection Polarization (OCP)...........................163
6.4.1 Optical Microscopy of the Investigated Specimens..............................166
6.5 Bare mild Steel after immersed in 0.5 M sodium chloride solution, adding
1.0% (w/w) Na- bentonite under CP, at 1.15V Vs SCE ...........................................172
6.5.1 XRD and SEM Analyses for Specimens Immersed in 0.5 M NaCl
Solution Adding to 1.0% (w/w) Bentonite............................................................176
6.6 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution, Adding 10.0%
(w/w) Bentonite at Open Circuit Potential (OCP) ....................................................190
6.6.1 Optical Microscopy of the Investigated Specimens..............................193
6.7 Bare mild steel immersed in 0.5 M sodium chloride solution, adding 10.0%
(w/w) bentonite under cathodic polarisation protection............................................200
6.7.1 XRD Analyses of Specimens Immersed in 0.5 M NaCl Mixed with
10.0% (w/w) Bentonite .........................................................................................205
6.7.2 SEM/EDAX Analyses of Specimens Immersed in 0.5 M NaCl Mixed
with 10.0% (w/w) Bentonite .................................................................................214
6.8 Bare mild steel specimens after burying inside the bentonite layer below the
solution level of 0.5 M sodium chloride solution adding 10.0% (w/w) bentonite with
no cathodic protection (OCP) ...................................................................................219
6.8.1 Optical Microscopy for specimens buried inside the bentonite layer ...221
6.9 Bare mild steel buried inside the bentonite layer of 0.5 M sodium chloride
solution containing 10.0% (w/w) bentonite under CP ..............................................227
6.9.1 SEM and XRD Analyses for specimens buried inside the bentonite layer
230
6.9.1.1 XRD Analyses for Specimens buried inside 0.5 M NaCl solution added to
10.0% (w/w) Bentonite .........................................................................................230
6.9.1.2 SEM/EDAX analysis for specimens buried inside 0.5 M NaCl solution
with 10.0% (w/w) Bentonite .............................................................................233
6.10 Bare Mild Steel Immersed in 0.5 M NaCl solution adding 10.0% (w/w)
Bentonite and 500 ppm and 1000 ppm Zinc chloride (ZnCl2)..................................236
6.11 Mild steel specimens after immersing in 0.5 M NaCl solution, containing 10%
(w/w) bentonite and ZnCl2 at 500ppm and 1000ppm respectively at -01.15V CP for
14 days. .....................................................................................................................241
6.11.1 XRD of Steel Specimens Immersed in 0.5 M NaCl Contains 10.0% (w/w)
Bentonite and 500ppm and 1000 ppm ZnCl2 under Cathodic Protection at -1.15V,
after 14 days ..........................................................................................................244
6.11.2 SEM /EDX analyses for specimens immersed in 0.5 M NaCl adding
10.0% (w/w) bentonite and Zinc Chloride............................................................248
7 General Discussion ..............................................................................................254
7.1 Introduction ...................................................................................................254
7.2 Specimens tested inside a layer of bentonite, adding 45% (w/w) NaCl solution
at 0.5 M under constant current density from 0 (OCP) to 200mA.m-2 .....................257
7.3 Specimens Exposed to 0.5 M Sodium Chloride Solution, Containing Different
Concentration of Bentonite. ......................................................................................259
8 Conclusions ..........................................................................................................269
9 Future Work ........................................................................................................271
10 References ............................................................................................................272
11 Appendices ...........................................................................................................280
11.1 Appendix I: Bare mild steel buried in bentonite containing 45% (w/w) 0.5 M
NaCl Solution............................................................................................................280

4
11.2 Appendix II: Bare mild steel immersed in 0.5 M NaCl Solution free of
bentonite....................................................................................................................293
11.3 Appendix III: Bare mild steel immersed in 0.5 M NaCl Solution containing
1% (w/w) bentonite. ..................................................................................................303
11.4 Appendix IV: mild steel immersed in 0.5 M NaCl Solution containing 10.0%
(w/w) bentonite. ........................................................................................................309
11.5 Appendix V: mild steel buried inside the bentonite layer at the bottom of the
beaker below the solution level. The corrosive electrolyte was 0.5 M NaCl containing
10.0% (w/w) Bentonite. ............................................................................................311
11.6 Appendix VI: Mild Steel Immersed in 0.5 M Sodium Chloride Solution
adding 10.0% (w/w) Bentonite and ZnCl2. ...............................................................317

5
ABSTRACT

The overall goal of this research is concerned with understanding the effects of
bentonite on the external corrosion of bare mild steel well casing. Na-bentonite is
mainly used in enormous amounts in drilling processes, so it used as the main
electrochemical environment surrounding the casing at different condition.

The major part of the current study was divided into 3 stages; the first stage is constant
current cathodic protection (CP) with a range of 0.0 (Open Circuit Potential) to
200mA.m-2 was applied respectively to protect bare mild steel buried vertically inside
the bentonite layer contains 45% (w/w) 0.5 M NaCl Solution. This study was attempted
to investigate the polarisation potential distribution over depth. The second stage is
polarisation potential with a range of OCP to -1.15V/SCE was applied to protect mild
steel exposed to 0.5M sodium chloride solution containing different concentration of
bentonite, namely 0.0%, 1.0 and 10.0% (w/w). The third stage was concerning with the
ZnCl2 added at 500ppm and 1000ppm to the bentonite as cathodic inhibitor to
investigate its effects on the corrosion process of mild steel. CP at 0.0mV (OCP) and -
1.15 V against saturated calomel electrode (SCE) were applied for this experiment.

Weight loss, visual observation, Open Circuit Potential (OCP), cathodic protection and
linear polarisation resistance (LPR) techniques were employed for this study. Optical
microscope, Scanning Electron Microscopy (SEM) and Energy Dispersive X-Ray
analysis (EDX), and X-Ray diffraction (XRD) analysis are illustrated to study, examine
and analyse the films that formed on the metal surface.

It is demonstrated that the corrosion rate produced from LPR measurement data was
fairly lower than that obtained by experimental weight loss data. Low corrosion rate
was recorded for the specimens immersed in 0.5NaCl solution, containing bentonite
compared to that were obtained for solution free of bentonite.

Magnesium hydroxide and calcium carbonate were the main chemical compounds that detected
on the metal surfaces when cathodic protection (CP) was applied at -1.15V vs. SCE. It was
confirmed that, the addition of ZnCl2 at two different concentration, 500ppm and 1000ppm has
reduced the current density applied to a considerable value. This was ascribed to the formation
of compacted and uniform film on the protected surface. The examination of the specimens
using EDX and XRD has shown the formation of zinc containing compounds. Small amount of
NaCl in the form of halite for both concentrations was also detected.

6
DECLARATION

No portion of the work referred to in this thesis has been submitted in support of an
application for another degree or qualification of this or any other university or other
institution of learning.

Signature

7
COPYRIGHTS

The author of this thesis (including any appendices and/or schedules to this thesis) owns
any copyright in it (the Copyright) and he has given The University of Manchester the
right to use such Copyright for any administrative, promotional, educational and/or
teaching purposes.

Copies of this thesis, either in full or in extracts, may be made only in accordance with
the regulations of the John Rylands University Library of Manchester. Details of these
regulations may be obtained from the Librarian. This page must form part of any such
copies made.

The ownership of any patents, designs, trade marks and any and all other intellectual
property rights except for the Copyright (the Intellectual Property Rights) and any
reproductions of copyright works, for example graphs and tables (Reproductions),
which may be described in this thesis, may not be owned by the author and may be
owned by third parties. Such Intellectual Property Rights and Reproductions cannot and
must not be made available for use without the prior written permission of the owner(s)
of the relevant Intellectual Property Rights and/or Reproductions.

Further information on the conditions under which disclosure, publication and


exploitation of this thesis, the Copyright and any Intellectual Property Rights and/or
Reproductions described in it may take place is available from the Head of School of
Materials (or the Vice-President).

Copyright in text of this thesis rests with the author. Copies (by any process) either in
full, or of extracts, may be made only in accordance with instructions given by the
author and lodged in the John Raylands University Library of Manchester. Details may
be obtained from the Librarian. This page must form part of any such copies made.

8
To My

Beloved family

9
ACKNOWLEDGMENTS

I start with the name of Allah who is the most gracious and merciful. I am very
thankful to Allah for his blessing and guidance to give me strength to achieve and finish
this milestone in my life.

I would like also to express my heartfelt thanks to my supervisor, Prof. J. D.


Scantlebury, for all the guidance, encouragement, endless help and support towards the
completion of this work. I would like also to thank Prof. Bob Cottis and Dr. Elena
Koroleva for their guidance and fruitful suggestion during this study. At the same time,
I am deeply indebted to my wife and children for their support and patience. I am also
very thankful to my brothers, and sisters for their emotional and financial support.

The author also would like to express his special acknowledgement to the sponsor; the
Libyan Ministry of Education for giving the opportunity to follow the PhD degree in
one of the supreme universities in U.K, The University of Manchester.

I would like also to express my gratitude and special thanks to the following persons in
the School of Material Sandra, Judith, S. Blach for their help in different ways. In
addition, also a big thank to Abdulhamied Twier for his support and help in different
ways, I am also grateful to all the students in the department especially to my lab mates
in D6 for their help. Many thanks to my great friends; Mukhtar Shaglouf, Elsadig Eltai,
Abdullah Al-rfaie and Ali Adham for their support and help throughout various stage of
my studies. Ultimately, I will always be thankful to my friends Mohammed Benosman,
Elmahdi Eljadi and Mukhtar abusaa for their advices and support.

10
Chapter 1

Introduction
1 Introduction

Well casing corrosion is a considerable problem for many industries includes water, oil
and gas, and others. As far as this project is concerned, well casing suffers from severe
and undesirable corrosion. Unfortunately, the complexity of the problem, uncertainties
and lack of adequate data on the real conditions in oil wells makes it difficult to find
either a practical or a theoretical solution using traditional analytical or numerical
approaches. Even today, numerous simplification and assumptions are necessary to
compensate for the lack of necessary data. Therefore, the corrosion of well casing
remains a challenge and a major cause of concern for oil companies consequently more
investigation is needed.

As mild steel is one of candidate material used for well casings. Massive studies have
been conducted for mild steel corrosion behaviour in aqueous electrolyte under aerobic
and anaerobic conditions. Although, only limited publication has been conducted for
mild steel in a semi permeable materials, such as bentonite clays. However, the
corrosion process for steel in bentonite clay is expected to be quite complicated than
that in aqueous solutions, due to different factors, including the redox potential,
chemistry of clay and different species present in the buffer media.

In this research, the external corrosion of mild steel exposed to semi- permeable
materials has been investigated, because such semi-permeable materials have been used
for drilling wells process. In addition to few studies and the shortage of data about the
effect of low permeable material (bentonite clay) on the external well casing, bentonite
was chosen to be the main corrosive electrolytes surrounding the casing at different
conditions. This is to gain further insight into the possible use of bentonite and to
provide a better understanding about the corrosion behaviour of well casing at different
conditions. Furthermore, the different corrosion products on the metal surface due to
corrosion process of mild steel exposed to bentonite and/ or the other films that built up
on the metal surface or generated during the application of different cathodic protection
levels has been investigated.

Different techniques have been applied to characterise those films which formed and
precipitated on the metal surfaces. In this study, considerable amount of experiments
have been conducted under different conditions using bare steel and numerous data was
generated.

12
Wyoming Bentonite was added to 0.5 M NaCl solution at different concentrations
(1.0% (w/w), 10.0% (w/w), However, it should be noted that the bentonite is a
commercially material available at the time of this research and does not necessarily to
represent the bentonite used in the field.

It was also recommended to study the electrochemical behaviour for mild steel under
natural corrosion process and/or under protection, when zinc chloride ZnCl2 was added
to the bentonite. ZnCl2 was used as an inhibitor in the presence of bentonite to study its
effect on the deposited films that may be precipitated.

Most potential criteria in the literature that have been proposed for adequate protection
were developed by both experiments and field experiments, in order to reduce corrosion
reactions to an acceptable rate. In this study, two cathodic protection potentials were
applied; -800mV and -1150mV (where all potential have been measured with respect to
Saturated Calomel electrode (SCE), as applying potential in this range is generally
accepted by many standards and codes of practice that at or below this potential the
corrosion of steel is reduced approximately to zero [1].

The Open Circuit Potential (OCP) was plotted as a function of time for each test and the
protected current over time was also monitored and measured at each constant
polarization potential applied to all the samples tested under all different conditions.

A significant part of the study was monitoring corrosion. Specific measurements


including free corrosion potential, polarization current, weight loss and polarisation
resistance (Rp). Data was reported for 7 to 14 days for each experiment before samples
were used for any further investigation. The data and associated interpretations are
presented and discussed.

Further investigation was continued to study, examine and analysed the films that were
formed on the metal surface. Optical microscope, SEM/EDAX, and XRD techniques
were used for such analysis. Different compounds were detected; including magnesium
hydroxide Mg(OH)2 and calcium carbonate (CaCO3). Different products has also been
detected when zinc chloride was added to the bentonite including: Chinese white zinc
Zn O, Pentazinc hexahydroxide carbonate Zn5(OH)6(CO3)2 and zinc Zn. Mg and Ca
elements were only detected in limited amount when EDX technique was imployed.

13
Chapter 2

Literature Review: Part one


Corrosion
2 Fundamentals of Corrosion

2.1 Definition of Corrosion

The general definition of corrosion is the degradation of materials as a result of reaction


with the surrounding environment [2, 3]. This definition includes all materials: metals
and non-metals, such as concrete, plastic and ceramic as shown in Figure 2.1 (a), (b).
The interactions of a metal or alloy with its surrounding environment will generally
degrade properties of the materials that need to be preserved. Corrosion can be
separated into a direct combination of both chemical and electrochemical reactions. In
the case of metallic materials, the damage occurs mostly through electrochemical
processes [4]. Thermodynamically, most engineering metals are unstable and tend to a
state of lower energy; this process is called corrosion, and the environments which
cause corrosion to the metal include air, water and/or moist soil [4]. Different
characteristics of corrosion that can be observed include both mass loss and mass gain.
Sometimes no obvious change in the metal will be seen but still its properties might be
changed physically or mechanically; for example loss of ductility without any loss of
weight [5].

Figure 2.1: Corrosion of materials, (a) metal, (b) brick [5].


2.2 Forms of Corrosion

The observed forms of corrosion can be differentiated by the environment that attacks
the metal: non-aqueous liquid, gases and aqueous liquid [4]. The different forms of
corrosion can be classified according to the appearance of the corrosion products and
visual observation (including the optical microscope and SEM) can be applied to
identify the different forms of corrosion. The forms of corrosion are well established
and described in the literature and in this thesis uniform corrosion, crevice corrosion,
pitting corrosion and galvanic corrosion will be discussed in some detail.

2.2.1 Uniform Corrosion

As illustrated in Figure 2.2, uniform corrosion is a regular and uniform removal of


metal surface, and is considered the most common form of corrosion. In general, this
type of corrosion is due to chemical or electrochemical reaction and, for any uniform
attack to take place, the corrosive environment must have an equivalent contact at all
parts of the metal to avoid any local cathodic or anodic centres on the metal surface [6].
Uniform corrosion is also known as general corrosion, from a technical point of view
this corrosion develops at a regular and slow rate, until the metal becomes so thin it
finally fails and causes severe damage, unless suitable control measures are taken [4].

Figure 2.2: Schematic representation of uniform corrosion.

16
Uniform attack is due to the energy state of the metal. Energy is added to the raw
material during metal manufacturing but the metal has a natural tendency to return to its
normal state when is exposed to a homogeneous electrolyte. For this type of corrosion
to occur, the metal must be homogeneous in both composition and metallurgical
properties. The most common example of uniform corrosion is atmospheric corrosion
and corrosion of steel exposed to an acidic solution [7].

2.2.2 Crevice Corrosion

Crevice corrosion is caused due to a concentration cell, when an oxidizer, such as


oxygen is depleted inside the crevice and at the same time there is an oxygen rich
environment around it.

Figure 2.3: Schematic representation of crevice corrosion [8].

Crevices can be due to many reasons, such as when two metals are loosely joined
together, as in the threads of bolts and nuts, pipes connected by an incompletely
threaded joint, an inadequately fitting gasket, under rivets that are working loose, and
between the bundle of tubes and sheets in heat exchangers and boilers. All these allow
moisture or small volumes of solution to penetrate and become stagnant, while other
area of the metal are in a more corrosive environment in the presence of liquid flow [3].

17
3Figure 2.3 shows the progress of corrosion in the crevice, where the metal become
more active at the crevice (anode) causing a large driving forces between anode and
cathode. But due to the nature of the crevice, all oxygen will have reacted and finally
depleted and very little oxygen can enter the crevice [3], while the anode reaction
continues producing more positive metal ions which attract anions into the crevice.

Migrations of aggressive anions, particularly chloride ions into the crevice will form
metal chloride which is considered a soluble chloride hydrolyser.

2.2.3 Pitting Corrosion

Pitting corrosion can be any localised attack that leads to the formation of pits or holes
in the metal surface. The pits can be due to a number of different conditions, such as
crevice corrosion, corrosion by water droplets and galvanic corrosion of a reactive
phase in an alloy and failure of a noble metal coating. Pitting is considered as one of the
most destructive types of corrosion because it causes more unpredictable corrosion than
any other forms of corrosion attack. From a technical point of view this type of
corrosion is not favoured as it is not predictable and failure often occurs suddenly due to
deep, shallow or undercut pits. Figure 2.4 shows an example of pitting corrosion.

Figure 2.4: Schematic representation of pitting corrosion [6].

18
2.2.4 Galvanic Corrosion

Galvanic corrosion takes place, when two different metals are connected together and
immersed in a conductive solution. A potential difference is usually developed which
produces more positive metal ions due to the corrosion of the metal with less corrosion
resistance. However, electrons being supplied from the previous anodic reactions will
be reduced at the more resistant metal and decrease its attack, the more resistance metal
is considered as the cathode in the electrochemical cell [6, 9].

Figure 2.5: Schematic representation of galvanic corrosion.

An example of this type of corrosion is brass in contact with copper hot-water pipes.
Figure 2.5 shows how a galvanic couple forms when brass becomes anodic and suffers
the loss of its zinc atoms. Brass in contact with galvanized steel is protected, while the
zinc coating on the steel is first dissolved, leaving the steel open to be attacked for the
same reason. An obvious area of concern is the use of one type of metal as bolts, screws,
and welds to fuse together pieces of another metal. The combination to be desired is the
large anode-small cathode combination. Bolts, screws, and so on should be made of the
metal less likely to be oxidized so that the bolt or weld is cathodically protected [10].

2.3 Electrochemical Nature of Corrosion

Any corrosion scientist must have the ability to understand the electrochemical nature
of corrosion. Almost, all metallic corrosion processes involve transportation of
electronic charge in an aqueous environment, thus all aqueous corrosion reactions are
believed to be electrochemical [11, 12]. Carbon steel is the metal typically used to carry

19
out this type of study, the electrochemical nature of corrosion of metals can be illustrate
by considering an example of steel immersed in NaCl solution. The following
instantaneous reactions on the metal surface can to occur; the anodic and cathodic
reactions which can be summarised as follows:

Anodic reaction: this chemical oxidation reaction in an electrochemical cell occurs at


the anode, which also known as a metal dissolution reaction where the atoms of iron are
transformed into a metal ion which usually dissolves in the electrolyte. In this reaction,
the valence of the iron atom increases from zero to +2 as it gives up two electrons [6,
13]. A typical oxidation reaction for iron is illustrated by the following equation:

Fe Fe2+ + 2e.. (2.1)

Where, Fe is iron atom, Fe2+ is iron ion and e is electron

The cathodic reaction: this chemical reduction reaction, takes place at the cathodic side
in the electrochemical cell: an atom or molecule gains and consumes the electrons
liberated from the iron. According to the nature of the environment in the
electrochemical cell, there is possibility for two reactions to take place at the cathodic
side and the reduced atoms or molecules may be deposited on the cathode or may be
released as gas according to the following reactions:

O2 + 2H2 O + 4e 4OH- . (2.2)


Or
2H2O + 2e H2 + 2OH ..... (2.3)

If a metal is corroding in an acid solution the cathodic reaction will be as follows:

2H+ + 2e H2 .. (2.4)

It is generally accepted that oxidation and reduction reactions are associated by an


increase and decrease in valence respectively, and one of the basic principles in
corrosion science is that both reactions must occur simultaneously and at the same rate

20
[4]. Based on Equations 2.2, 2.3, and 2.4, it is obvious that the corrosion of mild steel is
associated with the movements of ions from anodic sites to cathodic sites. Consequently,
the presence of anodic and cathodic sites is vital for corrosion reactions to occur. It is
also important to have an electrical connection between both sites in order for the
corrosion current to pass, and the electrolyte in contact with the metal side as a medium
for ion migration [3, 14]. Figure 2.6 illustrates the steps of the corrosion process for
metal under aeration cell conditions. Aggressive and active anions such as chlorides
have a very strong chance to break down any passive films, and the corrosion will occur
spontaneously at localised points which may cause pitting or localised attack of the
metal. More details of the corrosion process are described in Section 2.5.

Figure 2.6: Schematic diagram showing the steps of the chemical and electrochemical
corrosion processes for aeration cell [15].

2.4 Thermodynamics and Kinetics of Corrosion

2.4.1 Background

There are different types of corrosion mechanisms which depend on many factors
including the local environment that is the electrolyte in where the metal is immersed or
buried. Corrosion mechanisms also depend on the type of metal and the composition of
the metal and the state of the metal surface which play vital roles in determining the rate
and the type of corrosion which may take place.

21
It is well known that the corrosion processes are of an electrochemical nature, therefore
any change of electrochemical potential or the availability of electrons on the metal
surface may lead to a change in the rate of corrosion, which can be explained
thermodynamically since the equilibrium state of any reaction can be predicted using
thermodynamics. However, the thermodynamic explanation is limited since, for
example, it does not give any information about how fast or slow the reaction will be.
To do that the kinetics of corrosion need to be employed.

It is clear that the change in the electrochemical potential between anodic and cathodic
sites causes the ions to pass between anodic and cathodic sites. By knowing the change
in corrosion potential significant information about the possibility of corrosion reaction
can be obtained [4]. This information can be used to stop corrosion from taking place.

Sections 2.4.2 and 2.4.3 below discuss thermodynamics and kinetics of corrosion in
more detail.

2.4.2 Gibbs Free Energy and Electrode Potential

In metallic corrosion, the driving force is known as the Gibbs energy change (G)
which is defined as the change of free energy and must be negative when the corrosion
reaction process takes place spontaneously. In other words, the reactant reactions should
have a higher energy than the products. In real G calculation, a variety of intermediate
reaction stages are ignored and only the final and initial states of the reaction are
included. The Gibbs energy change can be expressed as:

G = -nEF .. (2.5)

Where: F is Faradays constant (96490 Coulombs/mole),

n is the number of moles of electrons released per mole of metal included in the reaction,
(mol e-/ mol of metal corroded) and

E is the driving force in volts (V), which V = joules/ coulombs.

For simplicity a spontaneous reaction for corrosion of steel in sea water can be
considered as:

22
Fe + 2H2O + O2 2Fe2+ + 4OH- .. 2.6

Equation 2.6 can be divided into two half cell reactions namely cathodic and anodic
reactions as follows:

Fe Fe2+ + 2e- ..... 2.7

O2 + 2H2O + 4e- 4OH- . 2.8

The standard electromotive force series for the elements is always listed according to
their standard reduction potentials, (E). The potentials are known as the half cell
electrode potential at certain conditions. E = ea +ec where ea and ec are the anodic and
cathodic potentials respectively. The sum of these potentials is equal to E in Equation
2.5.

It is common to use the Nernst equation to correct the standard potential for reactions at
the real corrosion process conditions by using equation 2.7:

RT [reduced ]
= Ln (2.7)
nF [oxidized ]

Where: E is the half cell potential,

E is the standard half cell potential,

R is the gas constant (8.314 J.K-1.mol-1),

T is the absolute temperature, K,

[reduced] and [oxidized] are the concentrations or the pressures of the products on the
reduced side of the reaction and the reactants on the oxidized side of the equation,
respectively.

23
The following Nernst equation can also be used to correct the standard potential for any
reaction at the real corrosion process conditions by using equation 2.7 in the following
form:

0.059 [reduced ]
E = E - log (2.8)
n [oxidized ]

In general, the Nernst equation is used for correcting the standard reduction potential,
but not either the oxidation potential or the total cell potential.

2.4.3 Application of Thermodynamics to Corrosion

In the previous section it was shown that there is a relationship between the
electrochemical potential and the free energy change (G). It was also shown that the
sign of the free energy can be used to determine in which direction the reaction will
proceed.

Pourbaix diagrams can be used to predict the equilibrium state of any reaction. Such
diagram is considered as a map, showing the different possible phases that are stable in
an aqueous solution system. It can be deduced that the employment of a potential pH
diagram is useful to corrosion engineering and valuable information can be obtained and
used to reduce or stop corrosion. Thus a basic requirement for any corrosion engineer is
to understand and apply Pourbaix diagrams to help prevent corrosion from occurring.

It is clear from Figure 2.7 that each stage is separated by lines derived from the Nernest
equation. Figure 2.7 represents a Pourbaix diagram for Fe-H2O system which shows the
different stages for the corrosion of iron as follows:

1) Corrosion: in this region extensive corrosion takes place under conditions equivalent
to the domain labelled Fe2+ and the metal will be dissolved as soluble ions products
according to Equation 2.1.

2) Passivity: this stage is characterised by the formation of an insoluble layer on the


metal surface. The primary corrosion products formed by anodic dissolution are ferrous
ions. Further corrosion products of ferrous ions due to oxidation reaction and
precipitation are transformed as stable ferric oxide hydroxide products, FeOOH.

24
Thermodynamically, FeOOH products are considered as stable corrosion products. It
shields the metal from the corrosive medium and stops corrosion from taking place [16].

3) Immunity: this stage is characterised by the absence of corrosion, as the metal is


considered thermodynamically stable.

Figure 2.7: Pourbaix diagram for iron water (Fe-H2O) system at 25oC showing zones of
immunity, passivity, and corrosion for species at activities of 10-6 M (g-equiv/L) [17].

2.4.4 Kinetics of Corrosion

Because of the importance of corrosion kinetics in determining the rate of corrosion, it


is necessary to understand its basic concepts as discussed below.

2.4.4.1 Faradays Law

The transfer of electrons to, or from, a reacting interface can yield important
information about the rate of any chemical or electrochemical reaction. The transfer of
such electrons can be measured as current (I) in Amperes. The current can be
determined from the Faraday law [18]:

Ita
m= (2.16)
nF

25
Where: m is the mass reacted,

I is the current,

t is the time,

a is the atomic weight in gram (g) per mol of corroded metal),

F is the Faraday constant (96500 coulombs (C)/mole e-), and n is the number of
electrons transferred during the reaction (mol e-/ mol of metal corroded).

To explain the concept Faraday laws consider the anodic reaction for iron in Equation
2.1 as an example, here n is 2 mol of electrons per mol of Fe.

If we divided Equation 2.16 by time (t) and the surface area (A), the corrosion rate r in
g.m-2s-1 can be estimated:

m ia
r= = (2.17)
tA nF

I
Where: i is current density which is equal to (Amperes/m2).
A

Dividing the corrosion rate by the density of the metal gives the depth of penetration in
a given time. In fact the corrosion rate, r, expressed in millimetres per year (mm/y) is
common in corrosion engineering. However, different units can also be used to express
corrosion rate, including mils per year (mpy) or milligrams per square centimetre
(mg.cm-2.d-1).

As can be seen from Equation 2.17, the mass loss per unit area per unit time is
proportional to current density [4].

2.4.4.2 Activation Polarization

One of the conditions of having a reaction under activation polarisation is that, the flux
of electrons must be controlled by certain steps in the half cell reaction [3, 18]. The

26
concept of activation polarisation and the steps involved can be explained by taking the
corrosion of steel in an acid environment as an example. For iron, the hypothetical
succession is showing by the following oxidation reaction steps:

Fe Fe+ Fe2+ Fe3+ (2.18)

It is clear that the hydrogen evolution reaction is predominant reaction in the acid
solution. It is suggested to be controlled by the steps shown in Equations 2.19, 2.20 and
2.21 [3]. The final step is sufficient molecules of hydrogen needed to combine and form
hydrogen bubbles on the metal surface. In Equation 2.19, a reaction between a hydrogen
ion cation that has been adsorbed or attached on the metal surface and one electron must
occur, resulting in the formation of adsorbed hydrogen.

H + + e H ads (atomic hydrogen) (2.19)

In the second step (Equation 2.19), H2 gas will be formed as a result of two adsorbed
hydrogen atoms combined to each other.

H ads + H ads H2 (hydrogen molecules) (2.20)

In the third step (Equation 2.20), instance formation of H2 gas bubbles on the metal
surface due to combination of hydrogen molecules.

H2 + H2 H2 (Gas bubbles) (2.21)

The charge transfer for the activation polarization is controlled by one step of each
reaction; the oxidation reaction of the steel (equation 18) and the reduction reaction of
hydrogen ion (equation 2.19).

27
2.4.4.3 Concentration Polarization

The controlling factor in this type of polarisation is the movement or diffusion of the
reactants or products to and from the metal surface via the solution. To illustrate this
type of polarisation, consider the corrosion of steel in sea water where the controlling
factor is the concentration of dissolved oxygen and its diffusion rate to and from the
metal surface via the solution. The resulting corrosion current of mild steel in sea water
can be illustrated as a limiting current density as in Equation 2.21 [17, 19, 20].

icorr = ilim = (D n F CB)/x (2.22)

Where: icorr. is the corrosion current density,

ilim. is the limiting current density,


n is the number of electrons involved,
D is the diffusion coefficient,
F is Faraday constant,
CB is the bulk concentration of species, and x is the diffusion layer thickness.

2.4.4.4 Resistance Polarization

This type of polarisation depends on a factor associated with the solution resistance on
the metal surface. The formation of any layer on both anodic and cathodic sites, or in
one of them of electrodes in the corrosion cell can act as a barrier between the metal
surface and the solution and help suppressing corrosion. This type of layer can be
formed on cathodically protected mild steel in sea water due to the formation of calcium
carbonate and/ or magnesium hydroxide and hence the passage of corrosive species
towards the metal surface will be restricted, resulting in lower corrosion rate [20, 21].

It can be concluded that the process of corrosion is very complex and as it is not fully
understood there is the possibility of disastrous consequences. Full understanding of the
corrosion process and the means to prevent it can only be achieved through deep
knowledge of a wide range of scientific areas including thermodynamics, kinetics,
chemistry, electric and mechanical engineering, and materials science. For example,
conducting research in the area of materials science can help corrosion scientists and
engineers develop new materials and improve the characteristics of existing ones to help
prevent corrosion and its consequences.

28
2.5 Corrosion Products and Passivity

Different types of corrosion products may be formed due to the electrochemical


corrosion reactions in its aqueous electrolyte. The corrosion products might be soluble
species, insoluble deposits or gaseous [3, 17].

Both, dissolved products and gases, particularly in de-aerated electrolyte systems which
usually involve the generation of H2 gas, will not remain on the metal surface and will
not interfere with the corrosion reaction. For example, if a piece of carbon steel is
immersed in an acidic solution, it will corrode rapidly to form a hydrous oxide film on
the surface. This film is not a protective film because of the ease with which it dissolves.
So the metal surface is soon again exposed to the electrolyte and the corrosion continues
[3]. Thus the conductivity of the corrosion medium might increase due to the soluble
corrosion products and corrosion rate may stay constant as shown in Figure 2.8 (1),
though there is the possibility that it might increase.

Figure 2.8: Hypothetical change in deterioration rate of metals with time [17] (1) If the
corrosion product is soluble and/ or a gas in the electrolyte, then the corrosion rate of
the metal is almost constant. (2) If insoluble corrosion products are gradually deposited
on the surface of metals, the corrosion rate declines with time. (3) Rate drops swiftly if
a passive film forms on the metal surface.

Insoluble products may also be found on the metal surface. These come into existence
following the precipitation of hydrous ferrous oxide (FeO.nH2O) or ferrous hydroxide

29
form (Fe(OH)2) on the cathode, anode or even in between as a corrosion product
according to the reaction:

2Fe + 2H2O + O2 2Fe(OH)2 .. (2.22)

In general, Fe(OH)2 itself is a relatively highly soluble material and cannot form a
protective film on the corroded metal surfaces, however insoluble solid corrosion
products can form on the metal surface due to further oxidation reactions, leading to the
formation of a protective film of magnetite; Magnetite is generally found in the middle
of corrosion products [14].

3 Fe(OH)2 Fe3O4 + 2 H2O+ H2 ..... (2.23)

Further reactions of hydrous ferrous oxide or ferrous hydroxide on the metal surface
with oxygen may also expected to occur to convert to hydrous ferric oxide or ferric
hydroxide (Fe(OH3) [22] as given in the following equation:

4Fe(OH)2 + 2H2O + O2 4Fe(OH)3 ... (2.24)

The corrosion products depend on the mobility of the ions that may be produced at
either the cathode or anode due to electrochemical reactions; insoluble precipitate has a
tendency to deposit at the anode, provided negative ions such as OH- produced due to
cathodic reaction in the electrochemical corrosion cell reach their destination at the
anode, before the positive ions that were produced at the anode.

Any corrosion deposits on the metal structure or coating faults, even if it is inadequate
or permeable scale will interfere with the oxygen mass transfer to the cathode side and,
at the same time, slow the diffusion of metal ions from the anode side and, hence, will
slow down corrosion. Consequently, the corrosion rate is gradually decreased by the
growth of corrosion deposits on the metal surface, see Figure 2.8 (2).

30
A number of reactive metals exposed to corrosive environments do not appear to
corrode due to a thin and completely invisible film that forms as a consequence of the
reaction of these metals with the environment. Corrosion films are considered as a good
barrier that slows the corrosion rate. This phenomenon is shown in Figure 2.8(3), and is
known as passivation. The passive film is a few atoms thick and mainly composed of
the oxide or hydroxide of the metal in its electrolyte. For any passive metal such as in
stainless steel that may form a passive film; any defects, scratches or break downs of the
protective film, it will repair itself rapidly. But aggressive and active anions such as
chlorides have a very strong chance of breaking down such films, and corrosion occurs
spontaneously at localised points that may cause pitting or localised attack for metal [4].

2.5.1 Types of Corrosion Products on the Metal Surfaces

Carbon steel is one of the most widely and commonly used industrial materials due to
its low cost and ease of fabrication, machining and welding [23]. It is the most
commonly used metal for oil well casings. However, this material is not stable and
rapidly corrodes in an aqueous environment when different types of corrosion products
may be formed due to the electrochemical corrosion reactions. The influence of the
environments on the corrosion behaviour under different conditions is the subject of
considerable research because the composition of the corrosion products depends on the
specific conditions in which they are formed, and thus will differ in different types of
electrolyte. In neutral to basic electrolytes an Fe(OH)2 film may form on the surface.
But if the electrolyte is weakly acidic due to presence of any impurities such SO2, the
film does not precipitate.

Ferrous ions (Fe2+) may form due to anodic dissolution in any corrosion process and are
considered as the main product. Further concentrations of Fe2+ ions due to precipitation
and oxidation reactions can result in insoluble ferric hydroxide (FeOOH) film on the
metal surface [16]. Formation of green rust is considered as a transitional corrosion
product for the oxidation of Fe(OH)2 to FeOOH [24]. Four different crystalline forms of
the corrosion products of FeOOH can occur; -FeOOH (lepidocrocite), -FeOOH
(goethite), -FeOOH (akaganeite), and (feroxyhyte) -FeOOH [16].

31
-FeOOH (lepidocrocite) is considered as the first crystalline corrosion product that
may compose. With the existence of certain conditions such as a weak acidic
environment due to the presence of sulphates, -FeOOH products are converted to -
FeOOH. The solubility of -FeOOH is about 105 times higher than that of -FeOOH
[16]. The -FeOOH product can also be found in environments that contain chlorides,
as in marine environments; in one marine site this corrosion product was found to
include up to 5% by weight of chloride ions [25].

Magnetite, Fe3O4, may form by the oxidation of Fe(OH)2. Fe3O4 is generally found in
the middle of corrosion products by reduction of FeOOH, under certain conditions such
as a limited oxygen supply:

8FeOOH + Fe 3Fe3O4 + 4H2O . (2-25)

In this work, the corrosion behaviour for the external casing in bentonite as an
electrolyte surrounding the simulated well casing was studied using different
electrochemical techniques. The different corrosion products, particularly the
composition of the precipitated films, due to the corrosion processes were investigated
with a view to show how they behave under different operating conditions.

2.5.2 Calcareous Deposit

Calcareous deposits usually consist of a white/grey scale mixture of calcium carbonate,


CaCO3 and magnesium hydroxide, Mg(HO)2. In sea water, the scale can be found on
the steel surfaces naturally [26]. The calcareous film can also be produced by applying
cathodic protection [27, 28]. This film can act as a physical barrier and has the ability to
reduce the oxygen, reaching the metal surface. According to the NACE standard
PR0176-2003 [1], the final current density required for long term cathodic protection
systems used in offshore petroleum production can be reduced, if a calcareous film is
precipitated on the metal surface. The calcareous film can be generated rapidly by
applying a relatively high initial current to the cathodic protection system. More rapid
polarization for the protected structures, using potentials in the range of 950 mV to
1050 mV with respect to SCE, have been applied in both field and laboratory

32
experiments to achieve a faster and more protective calcareous deposit. If the applied
potential is less than -950 mV, the protective calcareous deposit achieved is more
slowly [1]. However, various protection potential ranges from 800 mV to -1200
mV/SCE was also applied to protect the steel structures in sea water [28]. However,
evolution of H2 bubbles from the metal surface can be observed, when high cathodic
polarisation potential was applied, causing cracks in the deposits layer and delay the
aragonite layer due to it is small size [27]. Calcareous deposits have been well
researched and published, for equipments and structures mounted or built in the sea [28-
30]. whereas calcareous deposits formed on landbased pipelines surfaces have not yet
subjected to such detailed research and study [22]. For this reason, landbased deposits
on well casings is one of the main areas included in this research, to gain a better
understanding of the principles of this phenomenon.

The formation of calcareous deposit on the metal surfaces on the sea- based equipment
has been studied by different researchers and the following general concepts established
[28-30]: If a metallic structure immersed in seawater is protected from corrosion by
applying cathodic protection and if the availability of oxygen to the metal surface is
high, the initial and most important cathodic reaction is oxygen reduction under
activation control, as shown by Equations 2.27 and 2.28:

O2 + 2H2 O + 2e H2O2 + 2OH . (2.27)

H2O2+ 2e 2OH (2.28)

When the net current is limited by mass transfer to the surface and the supply of oxygen
into the protected structure is limited, particularly at the interface; in this case only
hydrogen evolution becomes the dominant reaction as shown in Equation 2.29:

2H2 O + 2e H2 +2OH ... (2.29)

Significant production of hydroxyl ions (OH) due to reactions (2.27) - (2.29), causes an
increase in local pH value on the protected metal surface. Sea water, brackish and

33
connate water contain different ions in different concentrations including; bicarbonate,
calcium and magnesium ions. The bicarbonate and carbonate ions in that water are in
equilibrium. The direction of the reaction depends upon the pH as:

HCO3 H++CO32 (2.30)

Carbon dioxide (CO2) gas can be found naturally in the atmosphere in several hundreds
ppm. A weak carbonic acid will form when it is dissolved in the water. If the
concentration of carbon dioxide (CO2) in the sea water is low, up to 15 ppm, the
reaction of CO2 with the water will be slow and any subsequent reactions that generate
direct precipitation of calcium carbonate (CaCO3) onto the metal surface as a protective
layer against corrosion will be correspondingly slow [26]. This precipitation is formed
by direct reaction between calcium and carbonate ions due to the local increase in the
pH value at the metal surface according to the following mechanisms [31]:

CO2 + H2 O H2CO3 (2.31)

H2CO3 H+ + HCO3 . (2.32)

HCO3 H++CO32.. (2.33)

CaSO4+CO32CaCO3+SO42. (2.34)

As mentioned above, the production of hydroxyl ions (OH) according to Equations


2.27, 2.28 or 2.29, result in an increase in the local pH near the metal surface/
electrolyte interface of the protected metal, causes the reaction of Equations 2.30, 2.32
and 2.33 to take place in the direction of producing more carbonate ions. These ions will
react with the calcium ions that are soluble in the electrolytes and form calcium
carbonate which precipitates onto the protected metal surface as an insoluble layer as

34
shown in Equation 2.35 or by direct formation of calcium carbonate at the cathodic sites
according to the Equation 2.34:

CO32 +Ca2+ Ca CO3 . (2.35)

A local alkaline surface due to further concentration of hydroxyl ions near to the surface
solution, may also cause a precipitation of magnesium hydroxide Mg(OH)2 onto the
protected surfaces as shown by Equation 2.36:

Mg2+ + OH Mg (OH)2 (2.36)

It is believed that, cathodic protection can be used to protect immersed metallic


structures from corrosion. This technique promotes precipitation of calcareous deposits
on the metallic surface, creating a physical barrier against oxygen diffusion towards
exposed metal surface [27, 32]. The result is a decrease in the final protected current
density required or extension of the life of the sacrificial anode to maintain sufficient
protective cathodic potentials for the structure [32, 33].

Usually, the main compounds of the calcareous deposit are calcium carbonate, calcium
hydroxide and magnesium hydroxide, but the precipitation of those compounds is a
function of many different factors. For example, the solubility product for calcium
carbonate CaCO3 (a crystalline appearance of calcite or aragonite) in seawater is lower
than that of Mg(OH)2, and it is considered to be in a supersaturated condition near the
surface in seawater [34, 35]. However, magnesium hydroxide Mg(OH)2 is recognized to
be under-saturated and tends to precipitate at quite high pH values, 9.5 [31]. So, the
precipitation of protective compounds will depend on the interfacial pH. The initial bulk
pH of seawater is about 8.2 and quite high current density is required to shift the initial
interfacial pH. For example, a calculated local pH for completely protected steel surface
in sea water was shifted from the original value to a pH of 10.9 [31].

Barchiche, and Deslouis [28] carried out experiments on carbon steel in artificial
seawater and applied cathodic potentials in the range from -900mV to -1400mV/SCE,
and showed that the calcareous deposits contained only calcium carbonate in the form
of aragonite when the cathodic potential was between -900mV and -1100mV/SCE. Both

35
compositions, magnesium hydroxide as brucite and calcium carbonate as aragonite,
were found when applying potential at -1200mV against SCE, and only brucite was
found when potential applied was -1300mVSCE. It is also reported that, Imperfect
layer contains CaCO3 will be delayed due to presence of sulphate (SO4-2) ions in the
environment [27].

Calcium carbonate solubility was inversely proportional to pH and/or temperature [26].


Approximate calculations based on the pH of saturation of CaCO3 (pHs) can be used to
indicate scale formation by using the following relationships [26]:

Lagelier saturation index LSI = pH - pHs . (2.37)

Ryznar stability index (RSI) =2 pHs pH (2.38)

Where: pH of CaCO3 saturation (pHs) = (9.3+A+B) (C+D)

Where A, B, C and D values are constants, which can be found in literature, which
illustrated by C. P. Dillon, Table19, page 185 [26].

According to the above relationships and the pH value calculation; the zero value for
LSI indicates balanced condition for scale, positive value indicates scale formation and
a negative value indicates scale dissolution. However, when RSI is > 7.5, it indicates
scale dissolution rather than formation.

36
Chapter 3

Literature Review: Part Two

Well Casing
3 Well Casing

The well casing equipments that is buried vertically or horizontally might corrode from
both sides; internally and externally and so it is often difficult to know the exact reasons
for corrosion and why leakages occur. However, these equipments must be protected
against corrosion from both sides to minimise any unexpected and sudden failures.
Experience demonstrates still cathodic protection for internal surface is not practicable,
an inhibitor can be used for controlling the internal corrosion [36]. Nevertheless,
external corrosion for well casing, cathodic protection can be applied to minimise
corrosion.

Surveys of the cost of corrosion carried out in a number of industrial countries have
found that substantial amounts of money are spent on corrosion control, and that the
different forms of material corrosion lead to increased direct and indirect costs. Direct
costs includes changing or replacing corroded parts, re-tubing, repainting, labour and
the use of chemical inhibitors or any other form corrosion protection such as cathodic
protection (CP). The indirect costs include plant shutdowns due to failure of essential
equipment, reduced efficiency and increased maintenance costs. These sums are very
difficult to predict and determine, but indirect costs are generally considered as great as
direct costs.

Studies by different national governments have estimated the direct costs of corrosion
of well casings. In Canada the failure between May 1988 and February 1989 (about 9 to
11 years after completion) of four casings at a depth exceeding 4600 meters cost more
than 6 M Canadian dollars just for casing replacement, excluding any equipment
replacement [37]. In the USA the direct cost of corrosion has been estimated at about
4.9% of Gross National Product [17].

A well casing is a large tubular structure placed in a borehole to maintain and support
the main well opening and can be for natural gas, oil or water. Figure 3.1 illustrates a
typical well construction with its important components according to W. Von.
Baeckmann and Schwenk [15]. Most casings comprise lengths of carbon steel pipe with
matching male and female threads at either end. It is widely used due to its low cost,
easy to fabricate and good mechanical (machine and weld) and corrosion resistance
properties [23]. However, other material can be used, including stainless steel and PVC.
Depending on the situation, including depth of well, pressure in the well and other
operating parameters, several pipes can be fixed inside each other in the region near the
well head; starting at the top with a large diameter and moving ahead through smaller
diameters with increasing distance downward. Typically the narrowest pipe will have
ID 4.5 inches (114mm) and OD 5 inches (127mm).

Figure 3.1: Profile for a typical well construction showing important components [38].

3.1 Corrosion Phenomena and Well Casing

The bare steel will corrode from the moment of it is buried in moist soil is a common
knowledge. Well casings suffer severe corrosion in the field and general causes of
corrosion are spontaneous electrochemical potentials generated by variations of the
strata through which the casing passes; poor or partially cemented casings, differences
in concentration between connate water and drilling fluid that may distributed through
different parts of the casing. The result is corrosion cells and hence, which a driving
force between anodes and cathodes at different locations on the external surface of the
casing with corrosion taking place at the anodic sites [39].

39
3.2 Causes of Corrosion in Well Casings

Different forms of corrosion may occur to exposed metal surfaces buried in a moist soil
or any other environment. As mentioned early, most well casings are of bare carbon
steel which suffers from corrosion that may occur due to microscopic compositional
changes of the casing material, concentration cells caused by non-homogeneous soil,
differential aeration and/or microbiological attack. Galvanic corrosion can also occur
due to electrochemical cells that may be created by dissimilar metals, and stray current
due to exterior electrical sources[40].

In the field, most casings extend from several hundred to a few thousand meters into the
earth; passing through different regions of geological formations, which are usually
layers of sedimentary rock, but may also be metamorphic rocks and volcanic flows [15,
41]. Each of these particular strata may have their own different water and moisture
content. The concentration of different mineral materials in the earth layers can vary
significantly with depth for an individual single well, with large changes in composition
and physical properties due to different temperatures and pressures. These variations of
strata, differences in concentration between the connate water in the different layers,
and the circulation of drilling fluid that may be distributed through different parts of the
casing, can create corrosion cells and, hence, spontaneous electrochemical potentials
may be generated and cause pitting corrosion at different locations of the external
surface of the casing [15].

In new wells, the annular space between the external casing and the surrounding
environment is filled with cement through the whole depth up to the surface, the reason
is to protect the outer pipe from the pressure of the surrounding rocks and to avoid
direct contact of the casing with the different environments including sea water, fresh
water and the different geological soil zones. In reality the cement covering the casing
will not be homogeneous and might be either of poor quality or there might be only
partial filling, both of which can cause corrosion cells.

In contrast to the new wells, only the upper sections of old well casings are cemented
and the other, lower, parts are not. Those lower parts of the casings can pass through
different aggressive materials such as salts in high density water and deposits due to the

40
drilling circulation. The casing also passes through different geological zones. Both
these scenarios can result in a large number of corrosion cells [21].

In general, corrosion of casings that are partially cemented is a matter of serious


concern. Different conditions such as parts of the casing being contained in cement and
other parts being in different soils, brine, fluids containing dissolved aggressive gases
and micro organisms etc. is a recipe for the promotion of corrosion and will lead to cell
formation on the external wall of the pipe resulting in corrosion and leakage. When
leaks in casing occur, aggressive water (for example) will come into contact with the
inner, main production pipe and will continuously attack the casing from the outside.
Thus pitting and/or other types of corrosion phenomenon can cause cracks or holes to
the main pipe.

Corrosion fatigue and sulfide stress corrosion cracking (SSCC) may also occur due to
formation of hydrogen sulfide by bacteria and the high concentration of sulphate and
dissolved salts in soil.

Bentonite is one of the main recipes of drilling mud which may cause severe corrosion
for external well casing in different location. Therefore, in this experimental study, part
of our investigation aims to focus on measuring corrosivity of mild steel caused by
bentonite at static condition, which was achieved by weight loss methodology.

3.3 Cathodic Protection of Well Casing

Well casing that cross terrain with different resistivity is usually protected from external
corrosion, applying CP to control anodic current flow from the metal surface to the
surrounding environment. The CP technique can be used in two different ways;
sacrificial anodes and/or the impressed current system [42]. The sacrificial anode
system is not recommended for protecting well casings due to the limiting voltage
difference, in the region of 1V [11] . However, the impressed current system is
applicable and can be widely used due to its large range of voltage settings, between 0
to 100V, which allows the current required for protection to be adjusted, and can be
effective in environments of high resistivity [17].

41
The most current criteria in the literature that have been proposed for adequate
protection was developed by both experiments and field experiments, in order to
mitigate or reduce corrosion reactions to an acceptable rate, consequently, eliminated
all anodic areas in the casing [43].

The cornerstone of this method is the decision of what is the necessary CP current
required to eliminate all the anodic areas. The goal is to mitigate corrosion by
maintaining the external buried metal surface of a structure at constant potential; by
lowering its potential from the natural corrosion potential to a value where anodic
corrosion is unable to take place. An impressed current generated by using an external
power source is often used to protect the outside surface of the well casing from
corrosion. The negative terminal of a direct current power source is connected to the
corroding casing at the wellhead, while the positive terminal is connected to an auxiliary
buried anode. Current flows out from the anode, through the soil, and is picked up on
the well casing (Figure 3.2). The distribution of current on the well casing determines
the degree of corrosion protection.

The distribution of electrical potential is a function of soil properties, including oxygen,


pH, ion concentration, and moisture content. The protection potential is adjusted to be at
a sufficiently negative level in the range between 850 mV to -1300 mV instant OFF
potential when the IR component of the measured potential is assumed to be zero as
explained in Section 3.5 .1.2.4.1. The term IR drop comes from the formula E=IR and is
due to the current from the CP anode to the metal pipeline passing through the
resistance of the soil. When the system is switched on, the measured potential including
the IR drop is called the ON potential. However, in the case where the potential is more
positive than -850mV it is not yet definitively determined that CP is useful, with
different opinions being expressed [44]. In this study, cathodic protection of simulated
well casing has been studied.

In this work, impressed current cathodic protection (ICCP) is the particular technique
has been used to study the level of protection for the external well casings. Current
distribution and polarization potential at different locations for the simulated casing of
mild steel was determined. A 0.5 M sodium chloride solution with and without
bentonite at different content was used as a corrosive electrolyte. It should be mentioned
that, bentonite is the main fluid used in the well drilling; which known as a drilling

42
mud. Therefore, in this study, metal was subjected to a series of experiment to
investigate the effect of drilling mud (bentonite fluid) on the casing and to study the
corrosion behavior of the casing under different conditions (further discussion about
bentonite will be presented in Chapter 4).

In order to realize high accurate and real simulation, the casing was tested, using both
homogeneous and non-homogeneous media, by using Na-Bentonite of different
characteristics, to take into account the variability of ground resistivity with depth.

3.3.1 Main Problems in Application of CP on Well Casing

A major difficulty for the CP designer is the well casing itself, since it is not a simple
long vertical pipe, but a series of pipes assembled concentrically from the top to the
bottom of the borehole, as illustrated in Figure 3.1. The first problem is to decide
whether all the casing parts are electrically conducting or not because it is impossible
for those parts not linked to be cathodically protected.

An additional problem is the inner sections; near the top of the well, see Figure 3.1, the
inner and outer casings form two separate cylinders with a layer of cement between
them, although, the all pipes must be connected together to ensure conducting current
passes through all metal when CP applied. If the annular space between the casings is
not adequately cemented the inner pipe will not be completely protected from the
surrounding environment, and there will be a tendency for corrosion to accelerate.

Figure 3.2: Schematic diagram for current distribution from different groundbed [45]

43
To control or adjust the current distribution over the entire depth of the casing below
ground is difficult. Generally, the groundbed (the position of the anode in the ground)
should be at a sufficient remote distance from the well to ensure an optimal and equal
distribution of current to the all casing, see Figure 3.2, [46]. The depth at which the
anode is placed is also important and will depend on the relative resistivity of different
strata.

A further problem for well casing is the electrical potential distribution over the casing
because, in general, the potential at the bottom is less negative than at the top [42]. It is
due to the potential drop along the length of the casing which is taken to be equal to IR
where I is the mean current through the casing and R the electrical resistance of the well
casing. Because the minimum protection potential must be achieved at the bottom of the
borehole there must be an over potential elsewhere, but this must be controlled so that
damage due to hydrogen evolution caused by strongly negative potential can be
avoided. This is particularly the case where casing material is high strength steel to
avoid stress corrosion cracking. This might be achieved by determining an appropriate
location for the groundbed. Figure 3.2c shows the current distribution for an anode
placed below a high resistance stratum to allow more polarization current to be forced
into the deeper parts of the casing. In contrast, increasing the level of polarization
current for an anode located close to the surface above a high resistivity stratum can
cause a very negative potential at the top of the well casing (Figure 3.2(d)) [45, 46].

The cost of CP is directly proportional to the adequate current that maintains the
protected structure free of corrosion. However, the average amount of current supplied
to the CP system must be limited to avoid over protection. That current is a function of
many aspects including time of polarisation and the thickness, structure, and the nature
of any film (e.g. calcareous deposits, see Section 2.5.2) that may form on the metal
surface, when cathodic protection has been applied at adequate potential[47].

3.3.2 Criteria for Cathodic Protection and Current Requirements

Unfortunately in reality, a great problem is the direct measurement of the most critical
parameters; i.e. potential along the casing, as it is almost impossible to measure. Results
have been obtained in certain limited circumstances. Different methods have been

44
developed for predicting the potential at the bottom of the well casing based on the local
and remote potentials measured at the surface.

However, no economical method has been established to determine the current required
to afford full protection of well casings. The criterion given in NACE RP0186-2001
[43] for the application of CP for the external surfaces of steel well casings is intended
as a guide for minimum requirements for corrosion control. Corrosion initiation and
propagation on the external surfaces of steel well casings cannot be predicted from
theoretical estimates so the degree of corrosion risks from formation of cells and the
effectiveness of CP cannot be determined in the field solely by pipe/soil potential
measurement along the casing [48]. Additionally, the total actual protection current
needed cannot be predicted with any confidence from simple rules of thumb. However,
there are number of tools that can be used for determining and detecting the protection
current needed along the well casing including the Casing Potential Profile (CPP) tool,
Corrosion Protection Evaluation Tool (CPET), E-Log I method, calculation and
mathematical modeling methods. The tools are very expensive due to either direct cost
or loss of production because production must be halted while the inside surface of the
pipe is cleaned and the measurements take place [49]. None of these techniques can
provide direct information on the casing surface polarisation of very deep wells. In
addition the CP technique has significant limitations; it assumes a uniform casing
resistance and perfect contact resistance. Hence, as a significant variation in the actual
resistances can extensively affect the calculated axial current flow [50], hence, it is hard
to estimate both the degree of metal losses and the corrosion rate accurately. The E-logI
technique is the most common method used for a long time to determine the actual
current required for CP. It is described in detail in NACE standard Pro186-2001,
Appendix B- E-Log-I Page 22. Hence, the Casing Potential Profile method is the only
technique will be explained in brief in this thesis. The others can be found in literature.

3.3.3 Casing Potential Profile

The CPP tool is a device used to measure the voltage drop between two sets of electric
probes in good electric contact with the internal well casing, Figure 3.3(a), and so
evaluate the efficiency of CP. The probes are about 8m apart and electrically insulated
from each other [15]. Knowing the resistivity of the well casing, the magnitude and

45
direction of the current flow can be calculated from measured voltage drop. If the
current flows from bottom to top, the casing is probably protected. But if, at any
location, the current is found to flow in the opposite direction, it means that current is
leaving the casing, and the casing is corroding [37, 46]; see Figure 3.3(b) for cathodic
potential profile log.

Figure 3.3: Casing potential profile tool and potential profile plot [43].

3.4 Corrosion by Soil

Corrosion of metals in soils is aqueous in nature and the mechanism is electrochemical


[12]. Soil cannot be considered as a homogenous electrolyte due to its different material
content. The heterogeneity that exists in the soil depends on such conditions as the
compactness of the soil, water content and dissolved oxygen. Most metallic materials
(including steel) placed in moist soil, will corrode, with the corrosion controlled by
diffusion of the dissolved oxygen in that soil solution [48, 51].

In stagnant electrolytes, the diffusion of metal ions away from the metal surface and
oxygen to the metal surface is slow and the diffusion of oxygen that controls the
reaction is much slower in soil because the gas has to diffuse through the soil as a layer
of water on the metal surface. Thus, the corrosion rate in soil is considerably lower than
corrosion in aqueous electrolytes.

46
In fact, most soils consist of aggregates of particles within a medium of organic and
inorganic matter rather than separate individual particles. This aggregation gives the soil
its characteristics. Soils with a high proportion of sand have a very limited storage
capacity for water and will be less corrosive than clay. Clays are less permeable so that
they tend to be poorly aerated but are excellent at retaining water, therefore a soil with a
high percentage of clay, e.g. bentonite, will be a more aggressive environment than
other soils containing less clay [52, 53].

Soil contains different compositions of organic matter, moisture, gases and mineral
particles. The relative size range of the particles comprising that soil does not determine
the whole nature of the soil structure [54]. The variation in soil composition can result
in different environments acting on different parts of the same metal surface, and this
can give arise to differing electrical potentials at the metal-soil interface [52]. This will
result in the establishment of anodic or cathodic areas and the consequent passage of
current through the metal and through the soil.

Redox potential is also known as a reduction potential for any aqueous solution which is
defined as a measure for aqueous environment to gain or lose electrons. The redox
potential was measured in the field by using platinum vs. normal hydrogen electrode
(NHS at pH7), (either gold or graphite material can also be used as an inert electrode)
[55]. So, the potential of the inert electrode is not a measure for oxygen concentration in
the environment rather it is an evidence of the oxidizing or reducing capacity of the soil.
For example, under aerobic conditions the oxygen content will be higher than that in
anaerobic soil, the oxidation reduction potential (ORP) will be more positive in an
anaerobic environment [48]. Therefore, the high reduction potential value for any
electrolyte is as an indication for aqueous environment to have a tendency to gain
electrons rather than to lose. Table 3.1 below gives an indication of corrosivity in water-
saturated biological-active soil.

47
Table 3.1: Classification of water-saturated biological-active soil corrosivity based on
oxidation- reduction potential [48].

Redox Potential, mV Corrosivity


<100 Severe
100-200 Moderate
200-400 Slight
>400 Not aggressive

3.4.1 Factors Affecting Corrosion by Soil

Soil alone does not determine the corrosion process; there are additional factors that
contribute to the corrosion process such as dissolved oxygen, water content, acidity, and
dissolved salts [48].

3.4.1.1 Aeration

The aeration of soil depends on different physical characteristics, including particle size
(Table 3.2), particles distribution, and apparent specific gravity. Soils of fine texture due
to high clay content as in montmorillonite contains more very closely packed particles
and have less pore capacity for gaseous diffusion than an open type soil such as sand.

Table 3.2: Classification of soil particles according to size [56].

Class Diameter (mm)


Gravel >2
Fine Gravel 1-2
Sand 0.05 to 1
Silt 0.002 to 0.05
Clay < 0.002

Diffusion of gases through a soil is enhanced by a number of climatic factors such as


temperature changes from day to night and variations in barometric pressure. Factors

48
that causing reduce diffusion rate like rainfall which can lead to anaerobic conditions
within the soil [20]. The general conclusion that can be made: metals in soils with low
oxygen content are usually anodic when compared with those in soil with high oxygen
content.

3.4.1.2 Water Content

The variation in water content of soil usually depends on many conditions including
rainfall, snow, flooding and the climate; in completely dry soil there is no corrosion of
metals. Hence, water is an essential factor for corrosion to take place in soil and the
corrosivity of soil is related to its moisture content. There is insignificant corrosion of
mild steel when water content is 50% and substantial corrosivity is observed only
above that level. Some authors have observed that above 20% moisture content soil is
aggressive towards carbon steel, and general corrosion can be expected when water
content is greater than that value [22].

Hence, water is an essential factor for corrosion to take place in soil and the corrosivity
of soil is related to its moisture content. There is significant corrosion rate for mild steel
when the water content is about 10% and the corrosion was classified by localised
corrosion and general corrosion is only observed when water content in soil is above
that level [57]. It was also reported that, when water content in soil is above 20%, the
corrosion rate was dropped considerably and general corrosion can be observed. The
corrosion products on the metal surface are mostly found to be lepidocrocite (-FeOOH),
when water content is greater than that value, but -FeOOH, Fe2O3 and magnetite
(Fe3O4) is the corrosion products will be expected on the metal surface when water
content is about 10% [57].

There is an inverse relationship between water volume and oxygen concentration in soil;
when rainfall ceases, soil dries and becomes more aerobic, especially in colloidal and
porous soils. The oxygen diffusion rate increases due to the larger pores in the soil, and
it gradually fills with oxygen instead of water. Because the cathodic activity of oxygen
is most vigorous when the water content in soil is greater than 50% this wet-dry cycle
(anaerobic-aerobic) leads to an increase in the corrosion rate [58].

Brackish water has less salinity than sea water which may result when fresh water meets
with seawater or connate water, as in the coastal regions of Northern Libya [Prof. Ali
Elbasir; private communication], the salinity is in the range of 0.5 and 30 grams of salt

49
per litre [59], while the sea water salinity [JC Rowlands in Shreir Vol. 1; 3rd edition;
Shreir Jarman and Burstein] [60] is in the range of 20 to 40 grams of salt per litre. So,
brackish water salinity can vary considerably over space and time, but their salinity is
less than that for sea water.

3.4.1.3 Soil Acidity

Soil acidity can develop due to organic and inorganic acids due to wastes from
industries and carbonic acid formed by carbon dioxide [61]. Any small variation in pH
represents a great change in concentration of hydrogen ions [36]. Acidic soils with pH
3.5 to 4.5 are usually strongly corrosive even towards steel. An acidic conditions tends
to have a depolarising effect, by removing hydrogen in the polarization films as it is
formed by the cathodic protection process [62].

3.4.1.4 Salt Content

Large number of chemicals can be found in soils and most of them are combined in
complex compounds. Water is considered as the main solvent whereas those materials
create soil solutions. In temperate climates and moderate rainfall areas, the soil solution
is relatively mild; regions of extensive rainfall show lower concentrations of soluble
salts due to leaching. Conversely, soils in dry regions are usually quite high in salts as
the salts are carried to the surface layers of the soil by water movement due to surface
evaporation. The nature and amount of soluble salts, together with the moisture or water
content in the soil will largely determine the ability of the soil to conduct an electric
current.

Generally, the most common cations in the soil solutions are sodium, potassium,
calcium, and magnesium. Alkali soils are high in sodium and potassium, while
calcareous soils contain predominantly calcium and magnesium. Such materials tend to
form insoluble films such as calcium carbonate and magnesium hydroxide in non-acidic
conditions which is considered as protective layers on the metal surfaces and reduce the
corrosion rate.

In this project, a moist soil of Na-bentonite is used as one of the main electrolytes and
further investigation was done to validate its usage and determine its effect when it was
the main soil solution surrounding the well casings under different conditions;

50
particularly when cathodic protection is applied to protect the external surface of the
well casing from corrosion. Moreover, the influence of different cations that may found
on the metal surface that was exposed to different corrosive environment is investigated
for more understanding their effect on current applied and potential polarisation. Results
will be presented in details in chapter 6.

3.4.2 Soil Chemical Analysis

Soil samples can be measured in location but if collected for later measurements, they
should be kept in air-tight containers to retain the moisture content. However, dry soils
may be moistened with distilled water to obtain their resistivity under wet conditions.

A wide variety of soluble salts can typically be found in moist soils; two soils may have
the same resistivity, but may have significantly different corrosion characteristics,
depending on the specific ions present. The major constituents that accelerate corrosion
are chlorides, sulphates, and the acidity of the soil.

Table 3.3 shows the effect of chloride (Cl), sulphate (SO4), and pH on corrosion of
buried steel structures [2]. Acidity as indicated by the pH value of soil/ ground water is
another aggressive factor. At low pH levels corrosion of buried metallic structures will
occur, but as the pH value increases above the neutral value, conditions become
gradually more alkaline and less corrosive. Strongly alkaline environments are not
aggressive toward steel and do not cause any serious damage.

As mentioned in Section 3.4.1.4, calcium and magnesium tend to form insoluble films
in the form of CaCO3 and Mg(OH)2 precipitate, leading to protect the metal surface,
reducing the corrosion rate. Bicarbonates are not typically detrimental to buried metals,
although high concentration of bicarbonates found in soil and ground water tend to
lower resistivity but there is no increase in corrosion activity. In contrast to the alkaline
earth elements, the chloride ions tend to breakdown any passive or protective films that
may created and cause pitting corrosion for buried metallic structures.

During corrosion surveys, it is important to test areas where there is any possibility of
unusual chemical conditions: knowing the pH value could have considerable

51
consequences for the location of CP systems - especially in acid soil conditions which
need relatively high current densities to maintain the CP [48].

Table 3.3: Corrosivity of Cl, SO4 and pH on buried steel pipelines [2].

Concentration (ppm) Degree of corrosivity


Chloride (Cl)
> 5,000 Severe
1,500-5,000 Considerable
500- 1,500 corrosive
< 500 Threshold
Sulphate (SO4)
1,500-10,000 Severe
< 10,000 Considerable
150-1500 positive
0- 150 Negligible
pH
< 5.5 Severe
5.5- 6.5 Considerable
6.5-7.5 positive
> 7.5 Negligible

3.4.3 Soil Electrical Resistivity

Soil electrical resistivity will give a combined measurement of both the moisture
content and the dissolved materials in the wet soil. Soil resistivity is easy to measure, so
it is frequently used and it may give important indications of soil corrosivity. Soil with
resistivity smaller than 2500 .cm is classified as a dangerous environment [11]. Table
3.4 shows soil corrosivity, based on the single parameter; soil resistivity, However, to
estimate the corrosion that may be expected from a certain soil, other characteristics
must be included and considered, such as total acidity, conductivity, water- air
permeability, and drainage- texture.

52
Table 3.4: Soil Corrosivity depends on their resistivity measurement [48].

soil resistivity(.cm) corrosivity


>20,000 Essentially non-corrosive
10,000- 20.000 Mildly corrosive
5,000- 10,000 Moderately corrosive
3,000- 5,000 Corrosive
1,000- 3000 Highly corrosive
<1,000 Extremely corrosive

3.4.3.1 Soil Resistivity Measurement

As soil resistivity is an important factor affecting the corrosion rate, the corrosion
engineer should know how to measure it. Many relatively simple resistivity techniques
are available, including the Wenner four-pin technique and the soil box method [63].

3.4.3.1.1 Wenner Four- Pin Method

Here four pins are inserted into the soil at equal distances from each other as shown in
Figure 3.4. An alternating voltage (I) is then applied to the outer pins. The potential
drop across the inner pins is measured. Using Ohms law (V= IR), the soil resistance
can be calculated. The resistivity () in .cm can then be determined by using the
following empirical equation [64]:

Resistivity () = 191.5 R.L ... (3.1)

Where: is the resistivity in (.cm),

L is the pins spacing in (cm), and

R is the resistance in Ohm ()

53
Figure 3.4: Wenner 4-pin apparatus with a schematic diagram for measuring soil
resistivity onsite.

3.4.3.1.2 Soil Box Method

The resistivity tests using the soil box technique operates on the same principle as the
Wenner four-pin test. A direct current (DC) is passed through the soil in the box, see
Figure 3.5, using metal end plates to give a more uniform current distribution. The
voltage produced across a pair of pins separated by a distance L is measured and Ohms
law used to calculate the resistance. Here the resistivity will be calculated as:

RWxD
Resistivity ( ) = (3.2)
L

Where: is the resistivity in Ohm.cm (.cm),

W and D are the metal plate dimensions in centimeters (cm),

R is resistance in Ohm (), and

L is the plate spacing in centimeter (cm).

54
Figure 3.5: Schematic diagram for soil box method, using combination of meters and
D.C power supply to measure resistivity in laboratory[65].

Figure 3.6 shows a soil box method, using an earth resistivity measuring set,
manufactured by the Metro HM RS, 174-511 Model. It is a direct measurement method
for resistivity which operates by applying the same principle as that mentioned above.
This technique was used in this research for measuring the resistivity of bentonite at
different water content.

Figure 3.6: Soil box method using an earth resistivity measuring set, manufactured by
the Metro HM RS, 174-511 Model.

55
3.4.3.2 Temperature Effect

Soil resistivity is greatly affected by its temperature. Seasonal fluctuations in


temperature and moisture content can substantially change the resistivity. These affects
are particularly pronounced in a high resistivity environment and may affect the CP of
metal materials [66]. It is recommended to measure the resistivity over the temperature
range likely to be found at the site, to measure the resistivity at other temperatures,
would be to introduce unnecessary errors [64].

3.5 Corrosion Prevention

The most effective time to prevent corrosion is during the design stage of the well by
determining the factors affecting corrosion and giving due consideration to the material
to be used. However, there are techniques available to the industry to mitigate corrosion
in addition to material selection; the most common are use of inhibitors, coatings and
CP, but other methods are available [40, 52].

The purpose of applying coating to any structure is to reduce the area of metal that is
exposed to corrosive media. Coatings can be applied to the metal surfaces to act as a
barrier between dissimilar metals surfaces to protect them from direct chemical
corrosion or direct contact with each others to prevent galvanic corrosion. From an
economic and technical point of view, it is difficult to protect a bare pipe line using CP
alone because the magnitude of the current required for the protection is high, creating a
high voltage drop in the soil, therefore external coatings and cathodic protection are
used together to minimise corrosion [11].

3.5.1 Cathodic Protection

As discussed in section 3.3, CP is a technique used to reduce the corrosion rate of the
metal surface of a structure by making it the cathode of an electrochemical cell with the
aim of lowering its potential from the natural corrosion potential to a value where the
anodic corrosion process is unable to take place; it is one of the most effective means
used to control corrosion of external surfaces of buried or submerged structures, and can

56
be applied either by the use of sacrificial anode or impressed current CP system [67],
see Figure 3.7 (a) and (b).

Figure 3.7: Schematic diagram for cathodic protection (CP) of a pipeline by (a)
sacrificial anode, and (b) Impressed current method [68].

3.5.1.1 How Cathodic Protection Works

It was pointed out in Section 2.3 that corrosion is an electrochemical process and
involves the passage of electrical current. When a metal (M) corrodes in a dilute aerated
electrolyte solution, Figure 3.8 a, two reactions take place, anodic and cathodic:

2M 2M+2 + 4e - ............................................................................................. (3.3)

O2 +2H2O +4e- 4OH- ...................................................................................... (3.4)

As illustrated in Figure 3.8 (b), the anodic reaction (3.3) produces free electrons which
pass within the metal to the cathode in the same metal surface and are consumed by the
cathodic reaction (3.4). Both reactions occur simultaneously with the result that metal
corrodes by dissolution of surface metal atoms. The principle of CP is to externally
connect an anode to the metal to be protected and supply electrons to the metal so that
all areas of the metal surface becomes cathodic and do not corrode, the rate of cathodic
reaction is accelerated and the rate of anodic reaction is decreased.

If iron for example is cathodically polarised using the above system, the rate of the half
cell reaction will be reduced, and with the supply of more electrons, the equilibrium will

57
be driven from right to left, the excess of electrons also increases the rate of oxygen
reduction and OH- production in the reaction (3.4) [62].

3.5.1.2 Cathodic Protection Systems

As mentioned early in section 3.3, there are two general systems of cathodic protection;
sacrificial anodes and impressed current. Each of those methods will be explained in
brief in the following sections.

3.5.1.2.1 Sacrificial Anode System

As soon as two different metals are electrically connected through an electrolyte, a


current will flow between them due to the dissimilarity of their electrochemical
potentials and polarization of the system will begin immediately. The more noble metal
becomes the cathode and usually corrodes more slowly than previously, while the more
base metals becomes the anode and corrodes more rapidly. The current flowing between
the two electrodes will accelerate the corrosion of the anode which is sacrificed to
protect the cathode [40]. This method of protection is known as the galvanic cathodic
protection system and is based on galvanic cell formation. Figure 3.9 illustrates the
electrical arrangement for a sacrificial anode to protect a metal pipe in soil. It is
connected to the pipe to be protected by an insulated wire. A galvanic cell is set up and
electrons flow from the anode to the cathode which is the more noble metal [69]. Steel
is protected from corrosion using magnesium, zinc or aluminum as sacrificial anodes.

Figure 3.8: Cathodic protection of pipeline buried in soil using a buried sacrificial anode
in soil [70].

58
Theoretically, any metal that is more base than the metal to be protected can be used as
a sacrificial anode but there are other considerations and requirements as to which
material should be chosen as the anode to be used in the system. For example,
magnesium has a negative potential which makes it suitable for environments of quite
high resistivity, but it is expensive in relation to capacity. Both aluminum and zinc have
more positive potentials than magnesium so their use is restricted to low resistivity
environments. Generally aluminum is more economical than zinc but has the tendency
to become passive, which may be overcome by alloying with mercury or indium.
Another difference is that aluminum does not work well when immersed in mud or sand
and zinc is more desirable under these conditions.

In general, anodes are designed in different sizes and shapes depending on the current
output required, expected life and the environment in which they are to be used. They
are connected directly to the structure to be protected, either singly, or, more commonly,
in groups or clusters which are known as the anode beds. As the anodes are consumed at
a rate proportional to the current produced, they must be monitored and inspected
regularly to determine when to replace them.

In the sacrificial anode system, the driving potential available to produce the protective
current is limited to the open circuit potential between the anode and the structure to be
protected; therefore the use of the galvanic anodes is limited to applications where
current requirements are relatively low and environments of high conductivity. Typical
applications of this system are for protecting relatively small areas of steel structures
buried in soil of low resistivity, or protecting stationary marine structures. It is also used
to protect the ships, underground pipelines, floating ducks, and most industrial
equipment in contact with corrosive environments.

The advantages of the sacrificial anode CP system is that it is easy to install,


independent of any external power source so it can be installed in remote areas where
there is no electrical power, is most suited to localised protection and less likely to
cause interference on neighbouring structures, with minimum maintenance required.

59
Anode Backfill

Backfill is used to reduce the anode resistance to lower than that of the surrounding soil.
Using special chemical mixtures, the backfill provides a uniform environment
surrounding the anode, absorbing the soil moisture and preventing local corrosion of the
anode. Hence, obtaining maximum performance from the sacrificial anode [53]

Typically a chemical backfill for use with a sacrificial anode could be a mixture of 75%
calcium sulphate (CaSO4), 20% granular bentonite and 5% sodium sulphate (Na2SO4).
This mixture is very useful in high resistivity soils due to its low resistivity; 50.cm.
However, the percentage of bentonite is recommended to be increased to 75% if the
moisture content in the soil is very low. This ensures good contact between anode and
soil as the bentonite absorbs the water and expands [62]. The anode resistance to earth
can be determined using the Dwight equation (3.5) [71];

4L
R= ( Ln( ) 1) . (3.5)
2 L r

Where: R is the anode resistance to earth (),

is the resistivity of the anode backfill (.cm),

L is the length of the anode in cm, and

r is the radius of the anode in cm.

3.5.1.2.2 Impressed Current Cathodic Protection System

Where high electrolyte resistivity and/or high output current are required, an impressed
current using an external DC source is more practical. In this type of system, the DC
source is typically a transformer-rectifier operating from an AC supply. Transformer
rectifiers normally operate in the DC range 0.0 to 100V and current outputs of 100A are
common.

The structure to be protected is coupled to the negative terminal of a DC power source,


while the positive terminal is coupled to an auxiliary anode, in order to lower the
structures potential to the protective potential [72]. Figure 3.10 shows a schematic

60
diagram of an external power source providing the external cathodic polarization for the
protected structure.

An impressed current system requires; inert anodes connected together and surrounded
by a non carbonaceous backfill (usually called a groundbed), DC power source and
properly isolated connections; one between the power source and the groundbed, and
the other between the power source and the structure itself [72].

Anodes used in ICCP systems can be either consumable or permanent anodes. An


example of expendable material is scrap iron buried in the soil. However, noble
electrodes including platinum coated titanium, silicon-iron, and lead alloys anodes are
considered as permanent anodes.

The main advantage of the impressed current CP system is that fewer anodes are
needed. Also it can provide a high driving voltage and current to meet the requirement
of the structure to be protected and is usually sufficiently flexible to cope with changes
in requirements. The main disadvantage of the system is the need for the external
electrical power supply. However, mobile current sources such as diesel generators can
be used to produce an AC current which then operates a rectifier to produce DC current
where other AC supplies are not available, and galvanic anode systems are not suitable
[73]. Other power sources include heavy duty batteries and windmill generators but
these are less common due to very low power output.

Figure 3.9: Schematic diagram of Impressed Current Cathodic Protection (ICCP)


system for pipeline using graphite anodes [70].

61
Anode Groundbed

Carbonaceous backfill [53, 73] can be used to increase the effective size of the
impressed current anodes giving a lower resistance earth and enabling more current
discharge from the anode. The backfill has to be compacted around the anode to ensure
good electrical contact. The resistivity of the carbonaceous backfills is about 50cm.
Metallurgical coke breeze is often used for shallow groundbed anodes; depths of less
than 15m. However, calcined petroleum coke backfills are produced specially for deep
groundbed anodes. Table 3.5 shows the properties of the backfill materials.

Table 3.5: Carbonaceous backfill analysis: Type LORESCO SW/SWK/DW.1/SC.3,


Manufactured by Cathodic protection Engineering Equipment Co. Inc (USA).

Typical physical analysis Calcined petroleum coke Metallurgical coke


Resistivity (.cm) 15 50
Well packed, no load
-2
Specific gravity(g.cm ) 2.1 2.1
-3
Bulk density (Kg.m ) 1185 737
Porosity at 150psi (%) 55 55
Typical chemical analysis
Ash (%) 0.1 10.0
Moisture (%) 0.0 3.0
Volatile (%) 0.0 0.4
Sulphur (%) 0.0 0.7
Nitrogen (%) 0.0 0.5
Carbon (%) 99.7 85.0

3.5.1.2.3 Choice of Sacrificial or Impressed Current System

The decision of whether to use a sacrificial anode or impressed current system is based
on two major factors, feasibility and cost. Cost analysis should include structure life
time and operating, maintenance and replacement costs as appropriate. In general, the
sacrificial anode is considered a simpler system compared to impressed current. A small
CP current against the sacrificial system, with low or no maintenance required and the
reduced level of interference associated with low current and small anode to structure
distance. It is considered a high active system in the conductive environment [70].

62
However, the impressed current system is generally used where high protection currents
are required, or where there is high local electrolyte resistivity such as hard rock with a
resistivity more than 100,000 .cm [17]. However, it can cause stray current corrosion
for other structures positioned close to it.

3.5.1.2.4 Advantages and Uses of Cathodic Protection

One of the advantages of cathodic protection over the other methods of corrosion
control is, it can be applied simply by the application of DC current and its effectiveness
may be monitored continuously. CP is commonly applied to coated structures to provide
protection to areas of coating defects. It can also be applied to existing structures to
extend their service life. Cathodic protection minimizes the need for corrosion
allowance and thereby cost. CP can only be used to protect metallic structures buried in
moisture soil or immersed in water, but it cannot be used to avoid atmospheric corrosion
of metallic structures. However, it can protect atmospherically exposed, or buried,
reinforced concrete structures because the concrete contains sufficient moisture to act as
an electrolyte [49, 72].

CP is used to protect the external surfaces of pipelines, ships hulls, offshore platforms,
storage tank bases, and harbour structures. Also it is used to protect the internal surfaces
of ships tanks.

IR Drop

When cathodic protection (CP) is applied to a structure, current flows into the system
through the electrolyte. The potential needed to protect the structure from corrosion
when it is in the soil, includes the natural potential (OCP), polarisation potential due to
applying the cathodic protection, and the ground IR drop, Figure 3.11 (a). The sum of
these three is termed the ON potential [74]. However, the actual potential we have to
consider is the OFF potential which is the ON potential less the IR drop, Figure 3.11
(b). The IR voltage drop is considered as an error in the readings of the CP current
applied to the protected system. The IR drop error is difficult to predict and changes
with changing conditions such as the current applied, resistivity and the reference
electrode to structure location [40]. The ON potential (EON in Figure 3.11(a) includes
the IR soil drop (ON potential = true potential + IR drop) and will change as the
distance of the anode from the protected structure increases. However, when the anode

63
is sufficiently far from the protected structure, the potential is no longer changes with
distance; that means no more change in IR drop. This location is known as remote earth
and the ON potential no longer changes, and the true potential of the structure does not
vary with the location of the reference electrode. However, the ON potential depends on
the location of the reference electrode.

Figure 3.10: Schematic diagram for well casing potential measurements for (a) system
under CP ON, including OCP, polarization potential and IR drop and (b) CP OFF is
free of IR drop [74].

So, as it is schematically shown in Figure 3.10 (b), the system is instantaneously


interrupted and free of ohmic drop (IR) or instant OFF potential is commonly used for
measuring the protection potential of metallic structures [75]. When the current for the
protection system is instantaneously turned off, the protected structure to soil potential
is equal to the sum of the corrosion potential Ecorr. (OCP) and the polarization potential,
but the IR drop error is assumed to be removed, see Figure 3.11.

64
Figure 3.11: Schematic diagram for potential difference between CP protected pipe and
soil as function of time followed by instantaneous off current for protected system [2].

Data plotted in Figure 3.12 represents an actual cathodically protected pipe line. The
data is illustrates the potential difference between pipe and soil as a function of distance;
both when the protected current for the structure is ON and OFF.

Figure 3.12: Potential difference between pipe and soil for pipe protected by CP, as a
function of distance [2].

3.5.1.2.5 Cathodic Protection Monitoring

CP systems are like any other electrical or mechanical system. They require periodic
inspections, maintenance and adjustment to maintain the structure potential at the

65
required protection levels and avoid both under and overprotection conditions [69]. CP
systems can be monitored by measuring the potential difference between structure and
electrolyte using a high input impedance voltmeter and a half-cell reference electrode
such as a saturated copper sulphate reference electrode (CSE). For steel exposed to a
corrosive environment, a shift of corrosion potential (Ecorr.) in the range from -850mV to
-1150mV with respect to Cu/CuSO4 (CSE), instant OFF potential (free of ohmic drop
(IR)) can be used [75]. However, according to the NACE standard PR0169[1] for steel
and cast iron piping, cathodic polarization to the level of -850 mV with respect to CSE
as protection potential (CP) is considered as one of the most common and effective
criteria used for reaching a successful control of external surfaces corrosion for
submerged and buried pipes structures [72, 73].

Cathodic protection has been used since 1820s, when Sir Humphry Davy was in control
of the investigation of corrosion of British naval ships [70]. However, since then, the CP
researches was carried out and developed and can be considered as fully published and
understood.

The average current density required for fully cathodic protection of external well
casing structures is changeable, and according to the experience gained in the field by
the people who apply CP show the current density will be in the range of 1 to 20mA/m2
[36].. However, NACE standard PR0186-2001, related to the criteria applied for
external surface of steel well casing protection, recommended a current density in the
range 10 to 200mA.m-2 [43], consequently external corrosion can be eliminated and
monitored at different levels of polarization. Even though, no guarantees was provided
to assure the desired results in some situations, such as elevated temperature and a very
low pH environment [37, 76], the various ranges of current density mentioned earlier
for protecting well casing, may be not sufficient to protect certain structures. So, the
environmental conditions and temperature are an important factors that may affect the
cathodic polarization level for structures under CP [77].

Moreover, and as it was pointed out in Section 3.3, the well casing penetrates through
many different layers of differing electrical resistivity, and due to this non-
homogeneous environment the casing exhibits both galvanic potential and polarization
characteristics. If a casing is under CP; the casing current distribution and casing
polarization voltage in the field can be calculated using electrical models developed for

66
the purpose but, nevertheless, problems arise in well casing modeling due to the non-
homogeneous environment and linear representation of the casing current. Thus, it
should be noted that calculations using electrical models are not accurate in all respects
and have not been completely validated due to limited field testing on real wells.

Exact site conditions as in the field may be difficult to simulate in the laboratory.
However, Part of this study is, experiments on a vertical steel structures (well casing)
buried in different soil layers have been carried out. The average current density in
milliampere per square meter (mA/m2) was used to calculate the current required to
mitigate external corrosion. Different current as that was recommended by NACE
standard PR0186-2001 was applied to protect a simulated external surface of well
casings surrounded by a bentonite environment at different conditions for determining
the minimum and maximum polarization protection potential and different levels of
cathodic polarization was achieved to protect the well casings at that conditions. The
test procedure was started by measuring the natural potential of the casing before any
current is applied to the system. Then, an impressed current is applied to the well casing
through the groundbed; the polarisation potential of the casing will shift to more
negative potential. That potential will be interrupted by interrupting the current flow to
the system. Following the interruption, the potential will depolarize within a fraction of
second. The potential prior to the time of depolarization is recorded which known as the
instant OFF potential which is free of IR drop.

67
Chapter 4

Literature Review: Part Three


Bentonite
4 Bentonite

4.1 Background

Bentonite was so named in 1898 by Wilbur C. Knight, the American geologists who
first mined Wyoming bentonite clay [78]. While the Wyoming deposits comprise about
70% of the worlds known supplies, bentonite has been found in almost all countries.
The clay was found to be the main component of volcanic ash and contains 65 to 90 %
by weight of montmorillonite [79]. Bentonite clay is characterised by its fine size, high
hydration and negative electric charge sheets. It is high swelling, up to sixteen times of
its original volume, with a low permeability [80-82]. Wyoming bentonite clay is called
sodium bentonite when sodium is the dominant cations in the clay, but is known as
calcium bentonite or non-swelling bentonite if the dominant cations in the clay is
calcium or magnesium [78].

Sodium bentonite grade is often used in enormous amounts for drilling wells, which is
considered as one of the main compounds that added to the water base drilling mud
recipe for viscosity and filtration control [83] and at the end of drilling, large amounts
of those materials penetrate through the main borehole and settle at different location, of
external casing. Such materials have been used as a barrier or sealing materials due to
its low permeability [84, 85], where oxygen faces some difficulty to penetrate through it
which may causes corrosion cells for casing at different areas. Consequences, localised
corrosion might be expected [86].

However, corrosion process for steel in bentonite clay is expected to be quite


complicated than that in aqueous solutions, due to different factors, including the redox
potential, chemistry of the clay and different ions present in the clay.

It has been argued that uniform corrosion is likely to be the dominant form of corrosion
for mild steel in bentonite; so, it is easy to assess the general corrosion rate [87].
whereas, other argue mentioned that, localised corrosion is the dominant, if mild steel in
bentonite becomes passive [86]. Therefore, the influence of corrosion products layers on
the corrosion rate for mild steel are still under debate, due to the lack information about
its corrosion process [88].
Different experiments for carbon steel exposed to water-saturated compacted bentonite,
under aerobic conditions were carried out. Hunda et. al. in 1997 [89], pointed out that
no more corrosion process takes place, after the corrosion product layers created on the
metal surface. On the other hand, X. Xiaa, et. al. reported that the corrosion products
layers that built up on the metal surface due to corrosion process in bentonite under
aerobic condition, can not restrict carbon steel from corrosion to continue, under
anaerobic condition. An 0.1 m/year has been estimated from iron profiles, as a value of
corrosion rate, which is considered very small value for corrosion of carbon steel in
bentonite, compared to the conservative corrosion rate values (20m/year) that used in
most repository design in Japan [88].

because limited publication has been conducted for mild steel in semi permeable
materials, so in this study, moist bentonite was chosen as one of the main environment
surrounding the casing at different condition to gain further insight into the possible use
of bentonite and to provide better understanding about the corrosion behaviour of well
casing at different conditions, particularly when cathodic protection (CP) is applied to
protect the external surface of well casings from corrosion. Commercially available
Wyoming-bentonite (MX-80), supplied by the UK Company CTM, was used in this
experimental work under various controlled conditions. However, it should be noted
that the bentonite is a commercially material available at the time of this research and
does not necessarily to represent the bentonite used in the field. A full analysis of the
bentonite used was performed on two different random samples using x-ray diffraction.
See appendix I.

4.2 Uses of Bentonite

Bentonite of various types has wide commercial use. In particular, it is commonly used
as one of main compounds added to water base drilling mud used during drilling
process in the oil industry [82, 90]. The bentonite expands greatly and increases the
viscosity of the suspension solution. Table 4.1 shows the typical chemical composition
of bentonite used for drilling mud.

However, corrosion process for steel in bentonite clay is expected to be quite


complicated than that in aqueous solutions, due to different factors, including the redox

70
potential, chemistry of the clay and different anions present in the clay. It is known that,
bentonite contains different chemical composition in different ranges, including SiO2,
Al2O3, Fe2O3, K2O, Na2O, MgO and CaO. When bentonite is added to the 0.5 M NaCl
solution at any concentration, different anions will be partially dissolved and release
ions. Therefore, different cations and anions will be ionised in the solution. In some
cases, a protective oxide film may be formed.

Table 4.1: Chemical composition of bentonite used as drilling mud .


Component SiO2 Al2O3 Fe2O3 K2O Na2O CaO MgO
Percent % 63-65 13-15 2-4 1.5 2.5-3.5 1-1.5 1-2

Only sodium bentonite (Na-bentonite) grade or equivalent is used in the preparation of


bentonite slurry [91] which is introduced during drilling excavation to stabilize the
sidewalls by formation of filter cakes around the surface of the well to reduce the
likelihood of cave-in and to lower the permeability coefficient of the sidewall soil [91].

Bentonite slurry is also commonly used as a barrier in landfill to retain contaminated


waste material due to its superior water adsorption and very low permeability [91]. It is
frequently used in America in landfill construction by combined it with other materials
such as geosynthetics [92]. Countries, including Sweden and Finland, have used
compacted bentonite as a barrier material in radioactive nuclear waste disposal [93, 94].

4.3 Classification and Chemical Composition of Bentonite

The basic elements of mineral clays are two-dimensional sheets of Si-O tetrahedron
units which link and self-arrange as tetrahedral sheets. The different forms of the clays
are classified according to the number of layers in each sheet and the crystal structure of
the layers. The tetrahedral unit is composed of four atoms of oxygen positioned at
identical distance from each other to form a tetrahedron with a silicon atom at its centre.
A diagram sketch of tetrahedron (a) a single silica tetrahedron and (b) isometric view of
silica sheet is shown in Figure 4.1.

71
Also present are aluminium or magnesium oxygen hydroxyl (Al-, Mg-O-OH)
octahedron units linked in octahedral sheets. The octahedron units in the octahedral
sheet are composed of six equidistant hydroxyl groups or oxygen atoms arranged in the
form of an octahedron, with an aluminium, magnesium or other atom (such as iron) at
the centre, see Figure 4.1(c) and (d).

Figure 4.1: Sketch showing (a) a Single Clay Tetrahedral unit, (b) the tetrahedral sheet
structure, (c) a single aluminium or magnesium octahedral unit and (d) aluminium or
magnesium octahedral sheet structure [81].

The general structural formula for most montmorillonite mineral clays are; hydrated
sodium (Na), calcium (Ca), aluminium (Al), magnesium (Mg), silicate (Si4O10),
hydroxide (OH), ((Ca, Na)0.33(Al1.67, Mg0.33)Si4O10(OH)2nH2O) [80, 95]. Different
materials can also be found in mineral clays including iron and potassium, but the exact
concentration of these cations depends on the source of the clay.

Commercial Wyoming sodium bentonite (Na-montmorillonite) was used in this study


its main and key component is a phyllosilicate called montmorillonite with the general
structural formula Al2O3.5SiO2.6H2O [96]. Bentonite is a 3-sheet clay that contain two
tetrahedral sheets and one octahedral sheet in each layer, and so is known as a 2:1
mineral clay [79]. The crystal structure of this clay is one octahedral aluminium or
magnesium oxide sheet with a sheet of tetrahedral silicon on either side. Each sheet is
less than 1 nanometre thick. The three sheets form one layer of bentonite mineral clay,
see Figure 4.2.

72
Figure 4.2: Schematic diagram of repeated sheet for tetrahedral: octahedral: tetrahedral
(2:1) layer [79].

The arrangement of the internal aluminum sheet and the external silicon oxide sheets
has a net neutral electric charge, but the mineral clay sheets have a negative charge due
to the presence of magnesium (Mg2+) ions which frequently substitute for aluminum
(Al3+) ions. These negative charges are neutralized by drawing the cations to the
interlayer surfaces through the bentonite mineral sheet structures [78]. These sheets are
connected to each other by counter-ions, usually sodium (Na+), calcium (Ca2+) and /or
magnesium (Mg2+) cations as showing in figure 4.3.

Figure 4.3: Schematic diagram to illustrate montmorillonite [96, 97].

4.3.1 Water Content in Bentonite

The water content is varied in different montmorillonite clays. Bentonite is


characterized by its extreme swelling property when in contact with water, and the
degree of swelling increases with the initial dry density [98, 99]. The balance of
attractive and repulsive electrostatic forces between the phyllosilicate sheets controls the

73
free expansion [98, 100]. For several applications, the amount of water that can be taken
up by the bentonite is important and the water content of the bentonite is used to
characterise the clay, so it is important that the measurements are carried out in defined
conditions [101]. For example, Na-montmorillonite will swell much more than Ca-
montmorillonite because hydration of the interlayer sodium causes dramatic swelling, of
up to more than 20 times the original volume [78].

Two out of three of the central spaces in the octahedrons are filled by Al atoms. The
third central space is vacant. Due to isomorphous substitution, the trivalent Al3+ cations
in the octahedral sheet can be replaced by divalent cations, such as Mg2+, Fe2+, or Zn2+.
So, the clay will have different chemical compositions and properties [97]. If the Al3+ is
replaced by Mg2+ then it is called brucite. In fact different basic crystal structures will
occur depending on the different combinations of cations and sheets. The most
commonly occurring form is found to be montmorillonite, where one Mg2+ cation
occurs in every six Al3+ in the octahedral sheets, see Figure 4.6, but no substitution
occurs in the tetrahedral sheets. Obviously, these results in a net loss of positive charge
of about 0.66 per unit cell, which will exert an attractive force on positively charged
cations, normally alkali and alkaline earth cations in adjacent clay units acting to bring
them to the outer surfaces of the tetrahedral sheets [97]. See the data in the columns on
the left and right hand sides of Figure 4.4.

Figure 4.4: Montmorillonite charge distribution [81].

74
Ions in clay that can be easily replaced by ions in an electrolyte solution passing
through it are generally termed readily exchangeable. Naturally, there is a hierarchy
whereby cations hydrated when mixed with water can be displaced into solution by
other cations of higher replaceability. The degree of replaceability will be a function of
the relative proportions (concentrations) of the different types of ions in the solution, for
example the high concentration of monovalent ions such as Na+ can displace the
divalent ions of calcium (Ca2+). Moreover, the divalent ions can also be removed from
bentonite chemically by using soda ash or sodium hydroxide. The following is a typical
chemical reaction to replace calcium ions (Ca2+) with sodium by using soda ash
(Sodium carbonate): Ca2+ +2 Na+ + CO32- 2 Na+ + CaCO3

It is reported that the physical size of the hydrated ions is important with smaller
hydrated cations replacing larger hydrated cations of the same valence that are present
in exchangeable sites, and higher valence cations replace lower valence cations see the
replaceability series listed in Table 4.2 [102].

Table 4.2: Replace ability or lyotropic series [102].

Li+ < Na+ < K+ < Rb+ < Cs+ < Mg2+ < Ca2+ <Sr2+ < Ba2+ < Cu2+ < Al3+ < Fe3+

If there is a net negative charge on the clay particles then if water comes in the contact
with them, adsorption of cations on the surface of the clay particles and even in the
interlayer can take place.

Replaceability will be a dynamic balance between the different factors; however, there
is no universal relationship for applying the lyotropic series due to the surface- specific
effects that always present in the clay [102], therefore, a very high concentration of
cations of lower replacing power present in a solution will be more effective than a very
low concentration of cations of relatively higher replacing power and the small hydrated
ionic radii (e.g. hydrated K+ radii is smaller than Na+) is the ideal cation to fit into the
depression of tetrahedral (silicon oxide) layer [81]. Table 4.3 shows the dry and
hydrated radii of common cations from different groups. Ion exchange is similar to any

75
other inorganic chemical reaction and occurs due to broken bonds and long range
electrostatic forces of relatively low energy.

Table 4.3: Dry and hydrated ionic radii of some cations [102].
Ions Ionic radii () Hydrated ionic radii ()
Li+ 0.68 10.03
+
Na 0.98 7.90
+
K 1.33 5.32
2+
Mg 0.89 10.8
2+
Ca 1.17 9.6
2+
Sr 1.34 9.6
2+
Ba 1.49 8.8
3+
Al 0.47 0.61 0.47 0.61
3+
Fe 0.57 0.63 0.57 0.63

76
Chapter 5

Experimental Methods and Instrumentation

77
5 Experimental Methods and Instrumentation

5.1 Introduction

This chapter describes and details the materials used in this investigation and the
experimental methods carried out to simulate and investigate corrosion behaviour of an
external mild steel well casing. A solution of 0.5 M NaCl mixed with different amount
of bentonite was used as a corrosive electrolyte. This specific concentration was used as
the main aqueous corrosive environment because the salinity of brackish water is less
than sea water. It should be appreciated that the salinity of this water can vary
considerably over space and time. It is widely accepted that the average salinity of
brackish water is in the range of 0.5 and 3.0 grams of salt per litre [59].

As was mentioned earlier in chapter 1, bentonite was chosen as it is widely used in the
drilling of oil wells, but its behaviour is not well documented and a detailed
understanding of its behaviour in deep strata is needed. Different concentration of
bentonite ranged from 0.0 to 10 percent was used to show the effect of bentonite content
in drilling mud on corrosion behaviour of well casing.

The experimental work conducted in this research included a series of electrochemical


studies using the following techniques; weight loss measurements, open circuit potential
(OCP), potential-time measurements, impressed current cathodic protection (ICCP),
cathodic potentiostatic polarisation and linear polarisation technique (LPR), to study the
corrosion behaviour of mild steel well casing exposed to 0.5 M NaCl solution mixed
with different amount of bentonite under different conditions.

Optical microscopy, X-Ray Diffraction (XRD), scanning electron microscopy (SEM)


and energy dispersive X-Ray (SEM/EDAX) were also applied to examine the nature of
corrosion and different compositions on the mild steel surface. Surface morphology
analysis for specimens tested under different conditions was also studied. Cathodically
generated films that may have formed on the metal surface were studied to better
understand the behaviour of bentonite and its effect on the external casing of wells.

Two different electrochemical CP methods were used to protect the simulated casing.
The first series experiments were conducted using an ICCP technique to protect mild
steel buried in bentonite at 45% (w/w) water/moisture content. The water content
expressed as a fraction = [(mass of moist bentonite (g) mass of dry bentonite (g))/
mass of dry bentonite].

Mild steel pipes were polarised at constant current density, to study effects of cathodic
polarisation potential, Five constant current densities were each applied for 7 days;
10.5, 50, 100, 150 and 200mA.m-2. A power supply manufactured by Harvey Turner
Ltd was used.

Thereafter, a sequence of experiments was conducted, using a potentiostat, to polarise


cathodically the mild steel specimens at constant potential for electrochemical studies.
Two protection potentials were applied for 14 days; at 800mV and 1150mV with
reference to the saturated calomel electrode (SCE). A potential in this range is generally
accepted by many standards and codes of practice [1].

The electrolyte used was similar to the bentonite used for drilling wells in oil field.
Water content and resistivity of bentonite were measured and determined. The results
will be presented in Appendix I.

In this investigation we expected to improve the quality of cathodically generated layers


that may form on the specimens surface by adding ZnCl2 to the protected system.
Consequently the current applied to the system may be reduced. For more
understanding about its effect, the appearance, surface morphology and composition of
any deposit that can be formed on the metal surface were also investigated. A
combination of SEM/EDX and XRD techniques were employed and the results will be
presented in chapter 6.

0.5 M NaCl solution with and without bentonite was prepared to act as a corrosive
electrolyte. ZnCl2 at two different concentration, 500 ppm (0.05%), and 1000 ppm
(0.1%) was also recommended to use it as a cathodic inhibitor. it is added to the
bentonite as a cathodic inhibitor. Adding this inhibitor to the bentonite was to study the
effect of zinc chloride (ZnCl2) and how it behaves on corrosion of mild steel by
plugging the defective sites on the metal surface particularly when CP was applied. It is
expected to improve the quality of a cathodically generated layer where may form on
the metal surface, and consequently reduces the current applied to the system.

79
The following sections will explain the steps carried out to prepare the electrolyte
environments, surface of the working samples before the tests, the several steps that
were conducted, and the experimental techniques applied during the present study.

5.2 Na-Bentonite

In this study commercially available Wyoming Na-bentonite grade was used as the
main electrochemical environment. According to the data sheet provided by supplier,
Na-bentonite is a fine white/grey powder in colour. The moisture content is in the range
10.0 to14.0% and bulk density of 0.90-1.00kg/m3 [79]. The bentonite used in this study
was analysed chemically, using energy dispersive spectroscopy (EDS) and XRD, see
appendix I.

Na-bentonite grade is a particular clay that is often used in enormous amounts in


drilling processes [83]. It is considered as one of the main compounds where added to
the drilling mud recipe. The most essential functions for drilling mud is to remove the
cutting from the hole, controlling high pressure zones, forming a filter cake on borehole
walls and preventing caving of wells. At the end of drilling process, large amounts of
those materials penetrate to the main borehole and settle at different location of external
casing as a mud. Initially, the mud is an important material to partially support the
casing [103]. Nevertheless it is characterized as semi-permeable membrane material.
Oxygen faces quite difficulty to penetrate through the clay and carbon steel may
becomes passive in the presence of such material [86, 87]. Furthermore, corrosion cells
for casing at different areas can also be expected, consequently localised corrosion can
be predicted [86].

The range of potentials studied in this experimental work varied between natural
potential (OCP), -0.8V and -1150mV against saturated calomel electrode (SCE). The
reason for applying potential in this range is because it is generally accepted by many
standards and codes of practice and polarisation potential more negative than -0.8V,
corrosion is reduced approximately to zero [1].

80
5.2.1 Bentonite Resistivity Measurements

A soil box, designed and built in our laboratory was used to measure the resistivity of
bentonite at room temperature, for different known water contents. A known weight of
deionised water was added to a known weight of dry bentonite. The water content
varied from 10% to 80%, see Table 1. The resistivity measurements for bentonite,
followed the ASTM standard reference test method G 57- 95a [65]. The soil box
technique was used as described in Section 3.4.3.1.2, page 53 [104]. The resistivity in
.cm was calculated by using the following equation:

R W D
=
L

Where:

is the resistivity in Ohm.cm (.cm),

W and D are the metal plate dimensions in centimeters (cm),

R is resistance in Ohm (), and

L is the plate spacing in centimeter (cm).

The Values for resistivity of bentonite at different water contents is given in Appendix I,
table 1.1 and figure 1.1 and 1.2.

5.3 Preparation of Specimens


The material used in this study was commercial mild steel, since the mostly common
material used for oil well casings. Two main different geometries were used; cylindrical
shape, supplied by Merseyside Metal Services Ltd. and the flat shape, supplied by Q
Panel. Table 5.1 details the chemical composition of mild steel specimens as supplied
by the companies. In comparison to the flat mild steel composition the pipe had higher
quantities of most elements.

Table 5.1: Chemical composition of mild steel

Pipe shape composition (w%) supplied by Merselyside metal surface


Component C S P Mn Fe
Percent % 0.2 max 0.05 0.05 0.9 max Remainder
The mild steel flat geometry composition (w%) supplied by Q Panel Company

81
Component C S P Mn Fe
Percent % 0.08-0.13 0.05 0.04 0.3-0.6 Remainder

As has been mentioned, two main different geometries is designed and prepared as
specimen samples used in the experiments; flat and cylindrical. The first specimen used
was a steel pipe with 51mm (2 inch) outside diameter (OD) and overall length 345mm
divided into three sections each of length 115mm. the total external surface area for
each section of specimen that exposed to the electrolyte (As) in sq. meters = d L =
(3.1416 51 115 106 = 0.018425m2). Hence, the total surface area for the three
section pipes exposed to the bentonite (Ast = 0.018425 3 = 0.055275m2). A 50 cm
length of 0.4mm diameter insulated wire was spot welded externally to one end of each
section. The connection point between the wire and each section insulated from the
surrounding medium using a sealing lacquer which was allowed to dry overnight and
was tested to show there was no electrical contact with the electrolyte. The other end of
insulated wire for each sample was connected by a single jack plug as shown in Figure
5.1. The three pipe sections were used for the open circuit potential tests.

Figure 5.1: Samples of mild steel pipes for open circuit potential experiments.

CP was applied for another set of pipes shown in Figure 5.2, each of these was also
51mm OD and 115mm long. One end of a length of insulated wire was welded
externally to the end of tube, the other end of wire was soldered to a 1kohm resistance,
see Figure 5.6. After that, the three sections of pipe were electrically connected together
using 500mm single insulated wire, soldered to the other end side of each resistance.
Then, the single plug at the end of the insulated wire was inserted into a power socket.
By measuring the voltage drop (V) across the 1kohm resistance the current output

82
through the protected system could be found using Ohms law (V= IR). The circuit can
be seen in Figure 5.11(b).

Figure 5.2: Mild steel pipes for cathodic protection experiments.

Figure 5.3: 1kohm resistance shunt connected in series to the circuit.

Rubber stoppers, see Figure 5.4, were used to seal the ends of the pipes from both side,
to ensure the total external exposed surface area to the electrolyte for each section of
pipe is 0.018425m2.

Figure 5.4: Rubber stoppers of top side diameter 52 mm.

83
Subsequently, a rectangular flat shape was used in the remainder of the experiments, the
mild steel initially measuring 150mm by 100mm by 1.2mm thick; this was cut into
smaller rectangular shapes (coupons) with each test specimen 35mm long, 30mm wide
and 1.2mm thick. Each side of the mild steel coupon presented an area of 10.5cm2. The
coupon had a 500mm length of single insulated wire spot welded to the middle of one
of the longer side. Samples of the coupons used in this study are shown in Figure 5.5.
The back of the each coupon was insulated, using insulating lacquer 45.

Figure 5.5: Mild steel coupons used in experiment.

5.3.1 Surface Preparation of the Specimens

The working electrode specimens that were used to carry out the experiments were
polished according to the ASTM standard, G1-03 [105], first using sand paper up to
4000 grit, then using 2m diamond paste, then washing in DI water and finally rinsing
with ethanol. Samples were then quickly dried by warm air drier and stored in a
desiccator containing silica gel until wanted for use.

5.3.2 Reference Electrodes

The electrodes chosen for measuring the potential difference between structure and the
environment are widely used as reference electrodes in the field:

1. Saturated calomel reference Electrode (SCE) was used to measure the structure
to solution potential.

2. Saturated copper-copper sulphate (Cu/CuSO4) reference electrode (CSE) was


chosen for measuring the structure to soil potential.

84
Cu/CuSO4 reference electrodes are commonly used as reference electrodes in CP of
underground pipelines and storage tanks [106, 107]. A schematic representation of a
Cu/CuSO4 reference electrode is shown in Figure 5.6. It was built in the laboratory
using a plastic tube filled with saturated copper sulphate solution, sealed at the top by a
cork with a copper wire through it, and at the bottom a porous plug made of soft pine
wood.

In order to ensure the all reference electrodes provided the same potential, a saturated
copper sulphate solution was prepared in the laboratory with excess crystals to ensure a
stable electrolyte pH. The electrodes were then filled with the saturated solution. The
CSE electrodes have the operating potential, +314mV (Cu/CuSO4) with respect to a
standard hydrogen reference electrode (SHE) at 20C [107, 108].

Figure 5.6: Schematic diagram of Cu/CuSO4 reference electrode, CSE [70].

To ensure that the Cu/CuSO4 reference electrodes were within specification the
electrodes were calibrated by comparing their potential to a known reference electrode.
As shown in Figure 5.7, two new SCE were used for calibration and were initially
checked against each other. The potential difference between the two SCE was 0.0mV.

85
Figure 5.7: Calomel reference electrode comparison.

The SCE has an operating potential of 242mV with respect to the SHE electrode at
room temperature. I.e. the SCE electrode is 242mV more positive than the SHE
electrode. One of SCE electrodes that used above for calibration was then used as a
reference to measure the potential of five CSE electrodes, see Figure 5.8. The difference
in potential between CSE and SCE electrodes should have an absolute value of 72mV
[107, 108]. The measuring potential for five CSE versus SCE reference electrodes was
68.5, 68.3, 67.5, 65.6, 67.9 mV respectively. Figure 5.8 shows a photo of a comparison
being made between CSE and SCE electrodes.

Figure 5.8: Calibration of CuSO4/Cu reference electrode vs. SCE at room temperature.

86
5.3.3 Auxiliary Electrode

As illustrated in Figure 5.9, two different geometries of auxiliary electrode were used;
(a) rectangular undistorted mesh titanium anode coated with iridium/ tantalum / titanium
oxide (IrO2-Ta2O5) supplied by BAC Anti-corrosion. Its dimension is about
500mm 200mm. The auxiliary anodes provide large contact area and ideal for most CP
applications with high current density and there inert properties to resist corrosion in a
highly corrosive environment[109]. Thereafter, a cylindrical auxiliary electrode has
same material was used for a sequence of experiments that was conducted, using a
potentiostat instead of power supply. The dimension of cylindrical auxiliary was
100mm length and 10mm diameter welded to a 300mm*1mm wire of same material.

(a) An undistorted mesh auxiliary (b) A cylindrical auxiliary


Figure 5.9: A auxiliary electrode titanium anode coated with iridium/ tantalum/ titanium
oxide (IrO2-Ta2O5) supplied BAC Anti-corrosion; (a) a rectangular undistorted mesh
and (b) cylindrical auxiliary anode.

5.3.4 pH Measurements

A pH is a measure of solution acidity, which indicates the hydrogen ion (H+)


concentration in the solution in moles per liter and it is defined as a negative logarithmic
of H+ concentration (pH = -log[H+]). Because H+ associated with water molecules to
form hydronium ions (H3O), for that reason, pH is often expressed in terms of
concentration of H3O+ [36].

87
The pH was measured for electrochemical cells using a standard portable calibrated
HANNA pH meter. Figure 5.10 shows calibration of pH meter before making any
measurements, using pH buffers of 7 and 9. The initial pH of the solution for all
experiments conducted in this project was adjusted to a value around 8, by using 0.1 M
NaOH solution and then monitored over the period of the experiments. The initial pH of
the solution was around 8. This pH value was used for all experiments in this study,
because drilling mud pH that normally used in drilling oil field has a pH in the range 8
to 10 [83, 110]. The pH for weight loss and free corrosion potential tests was monitored
over the period of the experiment.

Figure 5.10: pH meter Calibration by using buffer solutions 7 and 9.

5.4 Application of Cathodic Protection (CP), Using Power Supply

In the early stage of this research, a series of preliminary experiments were conducted.
Cathodically polarising the specimens at constant current densities was carried out,
using a Harvey Turner power supply. Five different current densities in the ranges of
10.5, 50, 100, 150, and 200mA.m-2 were applied, for 7 days for each current applied.
The tests used a power supply for several reasons; firstly, in the field, a real CP system
is normally applied at constant current density [44], secondly, with structures under CP
investigation a current drain test is used at different constant current densities to provide

88
an estimate of the optimal current density needed to protect the casing for different
environmental conditions at suitable polarisation potentials, and finally, to measure the
polarisation potential, as function of depth for mild steel pipes buried in sodium
bentonite mixed with different water content.

After applying various currents (current drain test), a certain film formed on the external
surfaces of the pipes. That film caused the main current needed to protect the casing to
decrease. Therefore, the potentiostat was used in the rest of this research when studying
the behaviour and composition of that film, by fixing a certain potential and the decline
or decrease of current readings due to the film composing was recorded and studied for
a known period of time. Different potential values were applied to the samples as will
be explained later.

A Perspex box 380mm high, 250mm long, 200mm wide and 1mm thick was assembled.
Dry bentonite was moistened, with by 45% by weight of 0.5 M NaCl solution. Sample
was used to fill the Perspex box. Before commencing the experiment, the aqueous
corrosive environment resistivity was measured. All experiments were conducted at
room temperature, under stagnant and aerated conditions.

The surfaces of the pipes were polished using sand paper up to 4000 grit silicon carbide.
Two sets of three replicate samples of mild steel pipe were buried vertically one above
the other in the moistened bentonite in one Perspex box, see Figure 5.11. The total
depth of the three pipes in the bentonite was about 345mm. However, when the pipes
are buried in the box and covered by bentonite, there is no direct metallic contact
between them due to the rubber plugs in each end of each pipe and edge sealing, but
they can be electrically connected to each other by the electric wire spot welded to each
pipe, to work as a single electrode.

The experimental set up for (a) the system under OCP, and (b) for the system under
ICCP are shown in Figure 5.13. A three electrode cell was used in this experiment, the
three electrodes being the steel pipes as working electrodes, a home-made saturated
copper sulphate electrode (CSE) as reference electrode, and a mixed metal oxide as
auxiliary electrode. The location of auxiliary electrode was 120mm away from the mild
steel pipes in a vertical position parallel to the section pipes. The duration of the
experiments was 7 days. As stated above the tests were conducted in 45 % (w/w), 0.5 M
NaCl solution.

89
The open circuit potential (OCP) against depth was measured once a day for each of the
three buried pipes. The OCP was measured against depth using the CSE reference
electrode and the eighteen 2mm diameter holes drilled into the box. The 18 holes are
drilled into the box, (from 0mm depth at the top to 340mm depth at the bottom of the
box, see Figure 5.11. All 18 holes will be closed, using adhesive blu tack throughout
the experiment. Each hole was opened only for the time of measurement, that is for a
few seconds when the blu tack was taken from the hole and the potential measured,
after that it was re-closed. So, six readings were taken from each section pipe (depths of
0-100mm, 120-220mm and 240 to 340mm).

For the system under CP, five cathodic current densities were applied; current densities
in the ranges of 10.5, 50, 100, 150, and 200mA.m-2 (Current density (CD) in mA.m-2
was calculated by simply dividing the measured current that was recorded regularly in
mA due to applying constant polarisation potential by the total surface area in cm2 of
specimen in contact with corrosive electrolyte. Results for these measurements will be
shown in details later in chapter 6. The experiment began by applying the 10.5mA.m-2
current density to one set of three pipes in one box for 7 days. After day 7, the current
was increased to 50mA.m-2 for 7 days more, and so on until all five current densities
had been applied. A grey film was formed on the external surface for specimens under
the application of cathodic protection. This film was observed after the specimens were
taken out from the electrochemical cell. This film was composed due to application of
cathodic protection.

Figure 5.11: Schematic diagram to illustrate the experimental setup used in the early
stage of this research (a) system under OCP (b) system under CP.

90
5.5 Preparation of Solutions and Experimental Set-Up

The second stage was a sequence of experiments was conducted using a potentiostat.
The mild steel specimens were cathodically polarised at constant potential instead of
applying current for electrochemical studies. The experiment was performed in the
region of polarisation potentials between open circuit potential (OCP), 800mV(SCE) and
1150mV(SCE). Measurement of current was taken for each cell at 1, 3, 5, 7, and 14 days
at regular interval time.

For experiments that were conducted by applying polarisation potential, the auxiliary
electrode was located in a separate cell compartment from the working electrode
compartment; to ensure that the auxiliary was mounted at the same homogeneous
solution (0.5 M NaCl solution) for all experiments were carried out, to limit the
influence of any change in the electrolytes being used to test the working electrode and
to limit the effect on any cathodic generated film that may compose on the metal surface
as acidic medium generated due to chlorine gas evolved at the anode of electrochemical
cell.

The solution was prepared by using the following procedure: 60g sodium chloride was
weighed out to make up two litres of 0.5M NaCl solution using deionised (DI) water.
The solution was mixed and stirred thoroughly until the salt has been dissolved. The 0.5
M NaCl solution served as an aqueous corrosive electrolyte. Before the experiments
were carried out the initial pH was adjusted to 8. The experiments were conducted
under stagnant, aerated conditions and at room temperature which was varied in the
range of 18 to 21.5 C.

A set of three replicate specimens of mild steel were completely immersed in the NaCl
solution, one sample in each of three different beakers (see Figure 5.13 a). We carried
out the experiment at the same Immersion conditions were used for both weight loss
and electrochemical measurements. First of all, free corrosion potential (OCP) was
measured manually. Then potentials were measured regularly: after one hour (referred
as 0 day) after starting the test, and then measured daily during 14 days.

In this experiment, the potential was measured with respect to calomel reference
electrode (SCE). A TENMA 72-7730 and Fluke73 voltmeter were used to measure
potential. The voltmeters has an impedance of more than 106 to avoid drawing any

91
current from the system during the measurements [107, 111]. A standard cell,
manufactured by H. Tinsley and Co Ltd, Weston type 1268 was used to check the
accuracy of voltmeters. The standard cell has the following known value, 1.01859 V, at

20 C. The cell is shown in Figure 5.12.

Figure 5.12: Photograph showing a standard cell, manufactured by H. Tinsley and Co


Ltd, Weston type 1268.

We carried out the experiments by using three specimens for each test, and there were
five sets, fifteen mild steel specimens in all. The area of each specimen used was
10.5cm2 (3.5cm*3.0cm). Each specimen was immersed in 0.5 M NaCl solution for 24
hours and then one set of three samples was taken for further investigation (Optical
microscopy, SEM, XRD) in which the specimens surfaces were visually monitored to
see the progress of reaction and rust production. At the end of three days another set of
three samples were removed for further investigation. Similarly at the end of 5, 7, and
finally 14 days sets of three samples were tested and then removed for further
investigation. At the time of polarization potential applied to the system, current was
measured regularly; after 1h, then after a further 1, 3, 5, 7, and 14 hours. Results for
these measurements will be shown in details in the next chapter.

The first set of experiment was carried out with cathodic polarization potential of 0.8V
vs. SCE. The reason for applying potential in this range is because it is generally
accepted by many standards and codes in corrosion engineering science and they

92
assume that at -0.780V vs. SCE or below this potential the corrosion of steel is reduced
to approximately zero [1]. An exact similar set of previous experiment was performed at
cathodic polarization potential of 1.150V vs. SCE by using new set of specimens.

A set of parallel experiments was performed to investigate corrosion of mild steel in the
presence of bentonite powder and bentonite suspensions. Here, specimens of mild steel
were immersed in 3% NaCl solution containing different amount of bentonite, see Table
5.2. After adding appropriate amounts of bentonite to a given volume of 3% NaCl
solution, the solution was stirred thoroughly for 30 minutes using a glass rod. It was
then allowed to settle for a period of one hour before the coupons were introduced.
During this hour some bentonite settled out and formed a layer on the bottom of the
beaker, the thickness of the layer depended on the relative amount of bentonite added to
the NaCl solution, which saturated with bentonite at different concentration.

First 15 mild steel specimens were immersed in 1% (w/w) bentonite mixed in 3% NaCl
solution, see Figure 5.12b. All specimens were suspended in the solution and not buried
wholly or partially in the bentonite layer at the bottom of the beaker. Three samples
were removed for further investigation at the end of 1, 3, 5, 7 and 14 days. The series of
experiments were repeated for different concentrations of bentonite that was added to
0.5 M NaCl solution.

In a second set of parallel experiments new 15 specimens were buried inside the
bentonite layer at the bottom of the beaker below the solution level, see Figure 5.13c.
As above, these tests were carried out for 1.0 and 10 % (w/w) bentonite content that
mixed with 0.5M of NaCl solution.

93
Figure 5.13: Experimental setup for specimens (a) Immersed in 3.0% NaCl solution
without bentonite, (b) Immersed in the solution above the bentonite layer and (c) buried
inside the bentonite layer.

Using 800mV(SCE) and 1150mV(SCE) protection potential for mild steel in 0.5 M NaCl
solution, the OCP was measured daily for separate replicate samples in different beakers
for a period of up to 14 days and the specimens surfaces were visually monitored after
different immersion times to check the progress of corrosion. The OCP and LPR were
measured after 1 day, 3 days, 5 days, 7 days, and 14 days, the test samples were
removed from the experimental cells (solution, bentonite and both) washed as described
above and stored in a desiccator.

94
Table 5.2: Amounts of bentonite added to 0.5M NaCl (3%) solution.

Experiment No bentonite (g) Concentration (%w/w)


1 0 0%
2 20 1%
3 200 10%

5.6 Weight Loss Test

In general, corrosion is defined as deterioration of the metal due to its reaction with its
environment. In oil field, corrosion of well casing remains a challenge and a major
cause of concern for oil companies and needs more investigation. Water base drilling
mud considered as one of the main surrounding environment used in the drilling process,
which may cause an electrochemical cells (anode/cathode interaction) for well casings.
An appropriate corrosion allowance is important to be predicted to guarantee the
equipment life time [112]. An evaluation of corrosion allowance requires clarifying and
estimating the corrosion type and corrosion rate (CR) with time. In this study, weight
loss measurement was used to determine the corrosion rate (CR).

The immersion tests for mild steel specimens were conducted according to ASTM
standard practice (G31-72) [113], including specimen cut and preparation, test
conditions and methods of cleaning specimens. Specimens were cut from carbon steel
sheet to rectangular shapes. Each coupon was nominally 35mm x 30mm x 2mm. A
0.02mm diameter hole near to the top of the coupon was drilled, so specimens could be
suspended by string in the electrolyte solution. The specimens were abraded to 4000 grit
finish by using SiC paper, accurately measured, degreased in acetone, washed in DI
water, dried in dry air, and then stored in a desiccator.

To give a baseline for comparison the specimens prepared for weight loss test each
specimen was photographed, accurately weighed using a four digit balance (e.g. to
0.0001g) and its dimensions accurately measured using a micrometer, before being
exposed to the corrosion solution. The pH of the solution was measured using a
HANNA pH meter.

Weight loss methodology was carried out, in order to determine the corrosivity that may
caused by bentonite that normally used in a water base drilling mud. Two duplicate

95
coupons were used for each weight loss test. The tests were conducted in 2 litre glass
beakers with two coupons suspended in each beaker. The specimens were located in the
corrosive solution.

The experimental weight loss measurements is considered as time consuming process


and the variation of corrosion rate with time, might leading to cause inadequate results
[18]. Even though, the most common exposure time used for weight loss measurements
is in the range of 48 to 168h [113]. Following this practice, weight loss measurements
was carried out for 1, 3, 5, 7 and 14 days to evaluate the corrosion rate with time. Then,
Faradic corrosion rate in mm/y was also obtained by determining the polarisation
parameters, including polarisation resistance (RP) using linear polarisation resistance
technique (LPR). More details about LPR will be presented in section 5.8. Duplicate
coupons were exposed to the corrosive environment at room temperature
(approximately 22oC). At the end of the tests, an optical microscope, scanning electron
microscope (SEM) and X-ray diffraction (XRD) analysis were used to examine the
corrosion products on the test surfaces, and weight loss measurements used after the
corrosion products were removed from the surface. The corrosion rate for specimens
tested in different environments was determined from the experimental weight loss
measurement. The results were demonstrated in greater detail in chapter 6.

The ideal procedure must be followed to remove corrosion products without removing
any base metal. Corrosion products on the specimens were removed and cleaned
according to the ASTM standard (G1- 03), using Clarkes solution [105] which consists
of 1000ml of hydrochloric acid (HCl, Sp. Gr. 1.19), 20g of antimony trioxide (Sb2O3)
and 50g of stannous chloride (SnCl2) [114]. The procedure involves immersion of the
test coupon in the Clarkes solution which is designed to remove the reaction rust
products. The specimen in the solution was vigorously stirred for 4 minutes at 20 C0,
followed by light brushing, and finally thoroughly rinsed and immediately dried.

The mass loss was determined after cleaning by weighing the specimen after cleaning,
using the same four digit calibrated balance. To minimise the errors in mass loss
measurement, each tested specimen was weighed three times and the average reading
was determined. The cleaning procedure was repeated five times for each specimen.

Figure 5.14 shows a graph for hypothetical mass loss with number of identical cleaning
cycles. Two lines are obtained; AB and BC. The former corresponds to removal of

96
corrosion products, the latter to removal of base metal. The required corrosion mass loss
occurs at point B, the intercept of the two lines [105, 114].

Figure 5.14: Hypothetic mass losses for corroded sample due to repeated cleaning.

Accurate results for corrosion rate can be ensured by testing more than one specimen.
Calculating the weight loss (W) in g and then corrosion rate (CR) has been determined
by using the following equation [105, 113]:

KW
Corrosion rate = ` (5.1)
ADT

Where;

K = a constant (see ref. 64),

W = mass loss in grams,

A = surface area of coupon in cm2

D = density of coupon in g/cm3,

T = time of exposure in hours

Corrosion rates are commonly measured for uniform corrosion product (rust) over the
surface area, as a mass per unit area per unit time. However, if the corrosion attack is

97
pitting or localized, depth of penetration may give more meaning and a better
description. Numerous units can be used to express corrosion rate, by using the correct
values of K constant, including millimetre per year (mm/y), milli-inch per year (mpy),
and milligrams per square centimetre per day (mg/cm2.d). The units of mpy, can be
converted to mm/y or any other units, for example mm/y = 0.0254 mpy [115].

Massive studies have been conducted for carbon steel corrosion behaviour in aqueous
electrolyte under aerobic and anaerobic conditions. But, only limited publication have
been conducted for carbon steel corroding in a semi permeable materials, such as
bentonite clays. However, corrosion process for carbon steel in bentonite clay is
expected to be quite complicated than that in aqueous solutions, due to different factors,
including the redox potential, chemistry of the clay and different anions present in the
corrosive media.

It has been argued that the uniform corrosion is likely to be the dominant form of
corrosion for carbon steel in bentonite; hence, it is easy to assess the general corrosion
rate [87]. On the other hand, other argue that localised corrosion is the dominant, if
carbon steel in bentonite become passive [86]. Therefore, the influence of corrosion
products layers on the corrosion rate for carbon steel are still under debate, due to lack
of information about corrosion process [88]. In this experimental study, different
electrochemical techniques were employed including weight loss methodology to
determine the corrosivity of bentonite. The corrosion rate against immersion time and
the data and its associated interpretations will be presented and discussed in the
following chapter.

Different experiments for carbon steel exposed to water-saturated compacted bentonite,


under aerobic conditions were carried out. Hunda et. al. in 1997 [89], pointed out that
no more corrosion process takes place after the corrosion product layers created on the
metal surface, while X. Xiaa, et. al. reported that the corrosion products layers that built
up on the metal surface due to corrosion process in bentonite under aerobic condition
can not restrict carbon steel from the corrosion, under anaerobic condition and he was
estimated a value of corrosion rate from iron profiles, (0.1 m/year) [88]. That value is
considered very small value for corrosion of carbon steel in bentonite compared to the
conservative corrosion rate values (20m/year) that used in most repository design in
Japan [88].

98
5.7 Potential -Time Experiments

The free corrosion potential (Ecorr.) is also called rest potential or open circuit potential
(OCP). The study of the corrosion potential for the casing subjected to different
corrosive solutions of varying bentonite concentration was carried out as described in
Section 5.8. The experiments were performed in order to characterise and study the
behaviour of mild steel when exposed to the corrosive environment under different
conditions. It was recorded for samples exposed to the corrosion environment as a
function of time. In such measurements, two different reactions are taking place on the
metal surface, anodic and cathodic reactions, simultaneously. At equilibrium, the
reaction rates for the anodic and cathodic reactions are equivalent and the net current is
equal to zero.

The corrosion potential alone supplies limited information and informed conclusions
cannot be drawn from such measurements alone, other information is required.
Therefore, under certain conditions, using potentialtime studies for samples immersed
in bentonite of varying conditions with simultaneous visual observation of any surface
films produced on the test samples may give valuable information [116]. The change in
Ecorr with time may be helpful and useful information such as the changes in relative
anodic to cathodic surface area can be predicted[117]; an increase in anodic to cathodic
surface area is an indication of the shifting of corrosion potential in the active direction
and a sign of metal loss. But when the corrosion potential shifts into the noble direction
with time, it is considered as a sign for an increase in cathodic to anodic surface area
ratio. The corrosion behaviour of the steel will vary with changing corrosive
environment condition. So under oxidising condition, if steel was passivated, localized
corrosion in the form of pitting, crevice corrosion can be expected. But obvious
conclusions should not be expected for corrosion mechanisms [117].

The corrosion potential-time data was measured, using the Fluke 73 voltmeter [107,
111]. The data was recorded and documented manually; the corrosion potential values
against immersion time were plotted for each test, and the data and its associated
interpretations will be presented and discussed in the following chapters. The potential
values were measured for samples at different time; 0, 1, 3, 5, 7 and 14 days. The tests
under different conditions, such as change of bentonite concentration were carried out.
In this study, bentonite in the range of 0.0 to 10.0 % (w/w) was added to 0.5 M NaCl

99
solution. A similar to the experiments reported in previous sections was also conducted
with the only differences being that the bentonite concentration used was 10% and 500
(0.05%) and 1000 ppm (1.0%) of ZnCl2 were added to the solution containing bentonite.

For most specimens tested under natural potential, different corrosion products were
observed due to the corrosion reactions; the products may work as a protective barrier
and might partially and/or completely isolate the specimens surface from the
environment. Specimens were then stored in a desiccator at the end of each test, until
needed for further investigation.

5.8 Linear Polarization Resistance (LPR)

Linear polarisation resistance (LPR) is one of the most common electrochemical


techniques used for corrosion studies. It is a well developed and effective technique
used for determining the corrosion rate, and is based on the Stern-Geary equation [48].
The theoretical background for the technique has been well studied [115, 118, 119]. The
LPR technique measures the polarisation resistance (Rp) values for identical series of
samples exposed to different corrosive environments.

In general, the LPR technique determines the polarisation resistance RP by applying a


small potential scan with respect to the corrosion potential (E = E - Ecorr) [119] (Figure
5.15). A sufficient external potential close to the corrosion potential (Ecorr.), applied on
the metal surface being exposed to the electrolyte can be used and as a consequence, the
current can be recorded; several small ranges of over potential for metal electrode
exposed to corrosive environment were recommended to predict linear polarization
behaviour [19, 48]. For example, 5, 10, 20mV shift from the corrosion potential is
usually used. In general, at those external polarization potentials, two electrochemical
reactions take place at the same metal surface; metal anode solubility and oxygen
cathode recovery [118].

100
Figure 5.15: Hypothetical linear polarization resistance plot [115].

The polarization resistance, RP, for a sample exposed to a corrosive solution is defined
as the slope of potential change against current density, as in Equation 5.2 [118]:

RP = VI = B/icorr (5.2)

RP should be reported in .cm2 because I is current density (A/cm2).

For accurate measurements and steady state behaviour to obtain Tafel slopes values, a
slow scan of LPR must be run over the corroded sample used. For accurate
measurements for Tafel slope values using the LPR technique, the solution resistance
(RS) must be very small compared to the sample resistance (RP) so that the total
resistance (RT =RP+ RS) RP.

Evaluation of Stern-Geary constant (B) and polarisation resistance (Rp) leads to


calculate the corrosion current density, using equation 5.3 [19, 119]:

B
i corr . = .... (5.3)
RP

Where:

icorr. Is corrosion current density in A.cm-2,

101
B is the Stern- Geary constant and

RP is polarization resistance in .cm2

The Stern-Geary constant (B) must be calculated from known Tafel slopes under
activation cathodic and anodic reactions; at the linear regions close to the corrosion
potential of an E log I plot [4, 19, 118]. However, the Tafel slope constants ba and bc
required to be determined or estimated separately.

b b
= a c ...... (5.4)
2.3(b + b )
a c

Where: ba is the anodic Tafel reaction slope (V/ decade) and bc is cathodic Tafel
reaction slope in (V/decade). The ba and bc can be determined from potentiodynamic
polarization curves. But if the Tafel slopes regions are non- linear, a theoretical value
for constant B calculated from arbitrary ba and bc values can be predicted. A value of
0.026V/decade is frequently used [40, 120]. Moreover, if the electrochemical reaction is
diffusion controlled rather than kinetic controlled, the cathodic Tafel slope (bc) is
infinite and Equation 5.4 reduces to Equation 5.5 [118]. However, if the reaction
mechanism is known, Equation 5.6 can also be used to calculate the Tafel slope, [19]:

B = 2.3 b ...... (5.5)


a

K .R.T
b = .... (5.6)
n.F

Hence, Faradays law can be used to calculate the corrosion rate from icorr. (Equation
5.3), by using the following equation [18]:

icorr . Mw
Corrosion rate (cm / s ) = (5.7)
Fn

Where; K is a constant and usually has the value of 2.0 for simple electron reactions, R
is the ideal gas constant, T is the absolute temperature in Kelvin (K), n is the number of
electrons (e-) released by the metal or involved in the reaction in moles of electrons per
mole of metal corroded, F is Faradays constant (96490 C/mol.s), Mw is molecular

102
weight in g.mole-1 and is metal density in g.cm-3. As carbon steel is the main metal
used in this study which is basically iron (Fe). So when corrosion takes place for carbon
steel, the oxidation reaction is Fe Fe2+ + 2e- (n =2), Mw is 55.845 g. mole-1, =
7.86 g.cm-3 and F= 96,490 A . s / mol. Common units used for calculating corrosion rate
is mm/y and most designers and engineers prefer it instead of cm/s.

If Tafel slopes cannot be predicted, the Stern-Geary constants can be determined by


using the polarization resistance and measuring the mass loss. So, it is possible to
calculate the corrosion current density from the Stern-Geary constants and the
polarization resistance using Equation 5.3. Then the corrosion rate can be calculated
using the corrosion current in Equation 5.7.

In this study, the LPR technique was applied to determine the RP and then the values for
corrosion rate were calculated. It is obvious that most of the Tafel slopes regions are
non- linear, hence a theoretical value for constant B calculated from arbitrary ba and bc
values has been assumed. The value of 0.120 V/decade was used for the anodic Tafel
reaction slope ba and cathodic Tafel reaction slope bc in (V/ decade). (B = 0.026 V) to
calculate the corrosion current.

As mentioned above in section 5.6, the weight loss (W) was also determined
experimentally and then corrosion rate in mm/y has calculated by using equation 5.1.
After that Icorr. in A.cm-2 was calculated by using Faradays low, using equation 5.7.
Finally B coefficient was determined, using formula 5.3 and then compared with the
coefficient B that has been assumed for determining the corrosion rate from polarization
curve.

The corrosion rate in (mm/y) was calculated from the overall weight loss. To compare
the results of the weight loss (gravimetric) method with those obtained from the LPR
method the LPR results were integrated to weight loss (g) by applying the following
procedure:

The intervals t = n1, n2, n3, n4, n5, n6


where n1=0, n2 = 1, n3 =3, n4 = 5, n5 = 7 and n6 = 14 days
Starting the integration at t=1 (t=n2):

103
Then corrosion currents were obtained at each test point and averaged between every
two measuring points.

I average =[I (n2)-I (n1)]/2

I is the current obtained by LPR measurement


Then, The integrated volume using the average current between those two points:

V (cm3) = [Iaverage*M(g/mol)*(n2-n1)*24*60*60 (sec)] /[n*F(C/mol)*density of


steel(g/cm3)]

The above steps were repeated for each data point using the average current between
each data point and the previous point and using the time span between those two points.
The next step is to convert to the weight loss (multiply it by steel density). So, the
weight loss is cumulative. That means each value of integrated weight loss is added to
the previous weight loss, such that the last data point is the total weight loss over the
test period. This was done to allow comparison with the measured values of weight loss
which I believe is also cumulative. This procedure was applied to the the all LPR results.

5.8.1 Experimental Cell and LPR Data Collection

The main objective for carrying out the LPR experiments was to evaluate the increase in
polarisation resistant as a result of film formation on the metal surface for the
experiments conducted in the presence of bentonite. The aim of this experiment was to
prove that the improvement and the drop off in cathodic protection current applied are
due to the film formation on the metal surface rather than any other influences.

The LPR data for test samples were obtained using ACM (GILL AC) version 5,
consisting of a potentiostat and a frequency response analyser. The LPR spectra were
measured and analysed using a suitable fitting procedure as recommended by the
manufacture of the ACM (GILL AC).

104
The electrochemical cell set up is as shown in Figure 5.16. A three electrodes cell was
used in this experiment; the mild steel specimens as working electrode, saturated
calomel electrode as reference electrode, and a cylindrical titanium anode coated with
iridium/ tantalum/ titanium oxide (IrO2-Ta2O5) supplied BAC Anti-corrosion as
auxiliary anode.

The working electrode dimension used in LPR experiments is similar to the one used in
the electrochemical cathodic protection experiments; the exposed area was 10.5 cm2,
and they were immersed in solution at different bentonite concentration as shown in
Table 5.2. Throughout the immersion time, several LPR runs were made.

A potential ranges of -20mV to +20mV was applied to each sample. The LPR value was
measured for samples at different time; 1h, 1, 3, 5, 7 and 14 days as it will be shown in
the following chapter.

Figure 5.16: Experimental set-up of LPR measurements used in this study.

5.9 X-Ray Diffraction (XRD)

The basis on which modern X-Ray Diffraction machines work remains the
fundamental principles first reported by Laue [121] nearly a century ago. The
first practical demonstration of X-Ray diffraction was by Davey [122] in 1921.
Thus this is a topic well-grounded in the scientific literature.

105
When a beam of X-rays meets and interacts with a crystalline material, the
beam is scattered by the structure of the solid. Because this structure has a
definite and repeated pattern the scattering of the incident beam is not random
but is scattered preferentially in certain directions. Diffraction patterns appear
in certain directions when the scattered beams are in phase and add together to
reinforce each other. Thus constructive interference pattern will be produced by
the regularity of the structure of the material, and it has been found that the
diffraction pattern (intensities and angles) generated by any substance is
specific to and characteristic of that material. The corollary is that to produce
an X-ray diffraction pattern a solid must possess a regular structure which
means that X-ray diffraction will occur only with crystalline substances.

This section does not attempt a full discussion of X-ray diffraction theory but
provides a brief introduction to essential operating principles. For a full
account of the theory there are many excellent reviews available [123, 124]. In
particular there are several extensive reviews that fully discuss the specific
application of XRD to materials science, including identification of powders
and deposited layers [125, 126].

Crystalline materials may be considered as an infinite and regularly repeated


three dimensional array of an identical lattice which can comprised of atoms,
ions or molecules. The unit cell is the smallest unit of the crystalline structure
and determines the pattern/shape of the crystal. Theoretically there are only seven
ways in which atoms etc, can be packed together to form a space filling lattice so there
are seven basic shapes for the unit cell, and seven essential crystal systems:
cubic, hexagonal, monoclinic, orthorhombic, rhombohedral (trigonal), tetragonal
and triclinic. However, extra lattice points can occur such as when an atom is at the
centre of a face or in the centre of a unit cell. These form extra lattice points which
combine with the seven essential crystal systems and result in fourteen possible Bravais
lattices [127]. All crystal lattices are Bravais lattices, although within the constraints
imposed by symmetry they can have different aspect ratios and be of different sizes. In
total there are 230 so-called space groups consistent with an extended, regular and
repeated crystal lattice pattern.

Many unit cells pack together to form a crystalline array, so it is possible to construct

106
geometric planes passing through different layers of atoms, called lattice planes which
continue throughout the basic structure. X-ray diffraction is used to measure the spacing
(known as d-spacing) between these lattice planes, usually in Angstroms, 10-10 m. From
the d-spacings it is possible to determine the shape and size of the crystal. Because the
crystal structure is unique to a given material this process allows precise identification
of the material. This is usually done by comparing the diffraction pattern obtained for a
particular test substance with a data base of known patterns, see later in this section.

Successful XRD analysis requires a parallel, monochromatic beam of X-rays of known


frequency. To generate the required X-rays a beam of high-energy electrons is directed
at an anode in a sealed vacuum tube. The frequency of the X-rays produced will depend
on the anode material. Anodes for the X-ray diffraction of inorganic materials tend to be
copper (Cu), whereas a cobalt anode tends to be used where the X-rays are to
interrogate ferrous samples. Any anode will produce an X-ray spectrum which may be
considered as being in two parts; unwanted continuous radiation and the characteristic
lines which are monochromatic. The stronger the characteristic line the more useful it is
for X-ray diffraction. The strongest line is the K, and from a copper anode the K line
is 1.5418 . The unwanted lines and continuous radiation are removed using filters
or a monochromator.

X-Rays are normally characterised by their energy or frequency which are


related by the well-known Planck-Einstein equation:

E = h f ............................................................................................................ (5.8)

Where:

E = energy of the X-ray beam.


h = the Planck's constant (6.626069*10 -34 Js -1 ).
f = the frequency of the X-ray beam.

Here the X-rays are treated as a beam of photons, and each photon has its
particular energy which can be calculated according to Equation (5.8).

Figure 3.20 is a schematic diagram of the main components of a typical

107
diffractometer. The essential parts are:
(1) the X-ray tube which is the source of the X-ray beam.
(2) the primary optics between the X-ray tube and the sample, which is a tube
containing a series of slits that reduces axial divergence and controls the area of
the sample being irradiated by the beam.
(3) the holder containing the sample being irradiated.
(4) the secondary optics between the sample and the detector consists of a set
of receiving slits and a curved crystal monochromator that control resolution.
(5) the detector.

Figure 5.17: Schematic diagram of PANalytical diffractometer suitable for powders


[128].

The crystalline sample is irradiated by the X-ray beam, and the detector is
rotated through a specific angle, whilst continuously recording the intensity of
the X-ray diffraction pattern. The pattern will usually consist of background
noise on which are imposed clear peaks (known as Bragg peaks) of different
amplitudes. These peaks are classified according to their Miller indices which
are simply a code for the plane in the crystal that generated that particular peak.
A peak occurs when the conditions specified by Braggs Law are met. A peak
is observed at an angle in the diffraction pattern when (see Figure 5.18):

108
n= 2 d s i n .......................................................................................... (5.9)

Where = Bragg angle.


= is the incident wavelength.
n = is an integer determined by the order of the peak observed.
d = is the spacing between the planes in the crystal lattice.

Figure 5.18: Diagrammatic illustration of Braggs Law [129].

Consider Figure 5.18, using Bragg's law (Equation 5.9) and measuring the
Bragg angle (2 ) at which the peak occurs, then knowing the wavelength ( ), of
the X-rays, we can then calculate the value of d, which is the d -spacing (the
distance between different planes in the crystal). Since d -spacing is
characteristic of the material being investigated then the measured d -spacing
can be matched against a digital data bank of some 120,000 compounds stored
on the hard-drive of the XRD machine. The diffraction patterns are in the form
of a unique series of peaks so the exact composition of any compound can be
identified.

In this study XRD was employed for specimens exposed to the electrolytes at different
conditions at OCP and with CP. The main objective for XRD investigation was to
investigate the presence of chemical compounds as a result of film formation on the

109
metal surface for the experiments conducted in the presence of bentonite. The X-ray
diffraction (XRD) was used to identify the crystal structure of any unknown compound,
and by the knowledge of its crystal structure, a chemical composition can then be
inferred for compound as its principles are based on the theory of X-ray crystallography.
Moreover, a type of crystal structure can be distinguished, as in the calcium carbonate
(CaCO3) and magnesium hydroxide (Mg(OH)2) compounds which are the main
components in calcareous films. These compounds can be found in different forms but
the XRD technique can provide a precise identification of chemical and the type of
crystal structure present for each compound by its pattern.

All XRD analyses were obtained using a Philips analytical X-Ray. The machine used is
shown in Figure 5.19. As was mentioned early, the frequency of the X-rays produced
will depend on the anode material. In this study, Copper (Cu) anode in a sealed vacuum
tube was used for the X-ray diffraction. The frequency of the X-rays produced, K-Alpha
[K] with wavelength of 1.5406 angstroms power was used. The measurement were
taken from 1 to 80, at 2Th. using a 1/min continuous scan, an accelerating voltage
of 45kV and current of 40mA.

Figure 5.19: Philips XRD equipment used.

5.10 Scanning Electron Microscopy (SEM) and Energy Dispersive X-


ray Analysis spectrometry, EDAX
The SEM technique is used to examine and obtain images of corrosion products and
films might be formed on specimens surfaces. The secondary electron image was
emploed in this research to obtain images in high resolution with good depth for sample

110
surface topography. Generally, for SEM images, secondary electron (SE) system at 20
keV was applied. But the Energy Dispersive X-ray spectrometry (EDAX) in the
Scanning Electron Microscopy (SEM) was employed in this study to reveal the
chemistry of different compounds that might be formed on the specimens surfaces
when exposed to different corrosive electrolyte [109]. Then a quantitative analysis for
those elements, and their progress with time on the metal surface was also investigated,
by using EDAX technique. An XL30 Field Emission Gun Scanning Electron
Microscope (FEG SEM), manufactured by Philips was used (Figure 5.20). The SEM
and EDAX techniques are considered as one instrument.

At the end of immersion experiments, specimens with 35x30 mm2 were prepared for the
flat surface SEM examination and EDAX analysis, by mounting the flat surface on top
of a standard 25mm diameter aluminium SEM pin stubs by using a double side carbon
tabs. In order to reveal all elements that might be formed on specimens surfaces several
EDAX areas were analyzed. SEM and EDAX results for all experiments were presented
in the results sections in the following chapter.

Figure 5.20: Philips FEG SEM.

5.11 The Impact of Zinc chloride

To study the effect of zinc chloride (ZnCl2) and how it behaves on corrosion of mild
steel, particularly in the application of cathodic protection, it is added to the bentonite as
a cathodic inhibitor. It is expected to improve the quality of a cathodically generated
layer where may form on the metal surface, by plugging the defective sites on the metal

111
surface particularly when CP was applied. Hence, reduces the current applied to the
system.

500 (0.05%) and 1000 ppm (1.0%) of ZnCl2 were added to the solution containing
10%.bentonite, respectively. In this investigation, when ZnCl2 was added to bentonite,
only polarisation potential at OCP and at -0.115V against SCE were applied due to the
limited time.

For more understanding about its effect, the appearance, surface morphology and
composition of any deposit may form on the metal surface was also investigated. A
combination of SEM/EDX and XRD techniques were employed and the results will be
presented in chapter 6.

112
Chapter 6

Results and Discussion

113
6 Results and Discussion

6.1 Introduction
This research is an endeavour to study the external corrosion of well casing by exposing
mild steel specimens to 0.5 M NaCl adding bentonite at different conditions.

Corrosion behaviour of well casing in a porous media such as bentonite is not well
documented and only a few publications have been performed [88]. Consequently,
limited data can be found in the literature about the corrosion behaviour of well casing.
Therefore, it is decided to use bentonite as the main corrosive environment surrounding
the casing at different condition. Maximum and minimum corrosion that may cause by
bentonite to external corrosion of bare steel well casing was studied to gain further
insight about bentonite which considered as the main source for corrosion cells,
consequently, a source for initiation process for localised corrosion [86, 87]. We will
present the results in great details in this chapter.

The experimental setup for open circuit potential (OCP) and application of impressed
cathodic protection (ICCP) experiments has been illustrated early in the experimental,
chapter 5, section 5.4. Experimental procedure was started by measuring the natural
potential of the section pipes before any current was applied to the system. Then, an
impressed current was applied to the system.

As was mention earlier in chapter 5, The CP technique is the particular technique was
used in the first stage of this research to study the level of protection for the external
well casings [42] ; current density and polarization potential at different locations and
time were monitored; the polarisation potential of the pipes was shifted to more
negative potential. That potential was interrupted by interrupting the current flow to the
system. Following the interruption, the potential was depolarised within a fraction of
second. The potential prior to the time of depolarisation was recorded. That potential is
known as the instant OFF potential, which is free of IR drop [75]. More details about IR
drop, ON and OFF potential can be found in details in chapter 3. Different levels of
cathodic polarisation were achieved.
6.2 Bare Mild Steel Buried Vertically in Bentonite Mixed with 45%
(w/w) 0.5 M NaCl Solution
The first group of experiments, bare mild steel specimens were buried vertically in a
fish tank, containing bentonite mixed with 45 % (w/w) 0.5 M NaCl solution, as a
corrosive electrolyte. Greater details about dimensions and experimental setup were
illustrated in chapter 5. In brief, measurements including the weight loss, free corrosion
potential (open circuit potential (OCP)) coupled with impressed cathodic protection,
visual observation after exposure to tests and polarisation resistance (Rp) were recorded.
The progress of corrosion on the steel surface was visually observed and then, a
photograph was taken for the specimens at the end of experiment. Figure 6.1 shows the
photographs for specimens (a, b, c and d) in two different geometries that were used for
this experiment; (a) and (b) are section pipes and panel shapes respectively. Either are
bare mild steel specimens, as polished before contacting with bentonite. The first
geometry (a) in photographs, used for OCP tests and the panel shape in (b), used for
weight loss, LPR, SEM and XRD analyses. Photographs (c) and (d) are corrosion
specimens of mild steel buried in bentonite contain 45 % (w/w) 0.5 M NaCl solutions.
Those photographs were obtained after exposure to the bentonite for 7 days. The surface
of section pipes, (c) were almost covered with corrosion products. A similar corrosion
product was also observed for bare panels, (d). The corrosion deposit formed on the
metal surface, using the specimens in panel shape was analysed by SEM and XRD
technique. Results of SEM and XRD analysis will be presented later in this chapter.

115
(a) As polished for OCP test (b) As polished

(c) After OCP test for 7 days (d) After test for 7 days
Figure 6.1: Corrosion of mild steel specimens after buried inside bentonite contains
45% (w/w), 0.5 M NaCl solution for 7 day period (a) and (b) as polished for OCP test
and (c) and (d) after exposing.

6.2.1 Open Circuit Potential (OCP)

Figure 6.2 shows the natural corrosion potential variation with depth in cm, for 3
polished bare steel pipes buried vertically at different depths in bentonite contains 45 %
(w/w) 0.5 M NaCl solution. The OCP readings were taken after 1 day and after 7 days.
The OCP values illustrated in this figure were taken from Table 3 in Appendix I. The
OCP for the 3 specimens that were buried in bentonite after 1 day are nearly identical in
the range of ~-523mV to ~ -525mV with respect to saturated copper-copper sulphate

116
electrode (CSE); the average difference in potential between the 3 different samples was
close to 2mV. OCP after 7 days for the same 3 steel pipes was shifted to more negative
potential to settle at ~ -597mV at 0cm depth to ~ -595mV at depth 10cm for the first

pipe, ~ -599mv at depth 12cm to ~ -602mVat depth 22cm for the second pipe and ~-

581mV at depth 24cm and 34cm. The identical potential at different depths for the
buried pipes indicates that there is no large change in the potential status; they remained
almost steady at ~ -523mV to -525mV vs. CSE after 1 day and in the range of -581mV

at 34cm depth and -603mV in 18cm depth.

-500
P o te n tia l (m V / C S E )

-540

-580

-620

-660 Pipe1, 1d pipe2, 1d pipe3, 1d


Pipe1, 7d Pipe 2, 7d Pipe3, 7d
-700
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36
Depth (cm)

Figure 6.2 Graph showing plot illustrating Open Circuit potential (OCP) over time for
steel specimens after burying in bentonite contains 45 % (w/w) 0.5 M NaCl solution at
different depths for 1 and 7 days.

The potential over time measurements were recorded daily for the three different pipes
buried in bentonite at different location. As can be seen from the above figure, it is
evident that the potential was shifted to more negative with time. The OCP for the three
different pipes is approximately identical at all depths, which may be gave an indication
that the corrosion products on the metal surfaces were growing homogeneously with
time. This might be due to the limiting amount of oxygen that was trapped in bentonite
due to its initial stage of filling as bentonite [84, 85]. The O2 may be gradually
consumed by cathodic reduction reaction in the first stage of metal corrosion (O2 + 2H2
O + 4e 4OH). A spontaneous anodic reaction (oxidation reaction) for iron was

117
taken place (Fe Fe+2 + 2e). In bentonite, oxygen diffusion became very limited and
the reaction will slow down or cease.

The general conclusion that can be made is; a negligible change in corrosion potential
for steel specimens in bentonite was recorded. This behaviour can be attributed to the
limiting and identical amount of oxygen in bentonite which may be consumed by
oxygen reduction reaction.

As the oxygen consumed due to cathodic reaction and the reduced atoms or molecules
may be deposited on the cathode in the form of Fe(OH)2. In general, Fe(OH)2 it self is a
relatively highly soluble material and cannot form a protective film on the corroded
metal surfaces, however insoluble solid corrosion products can form on the metal
surface due to further oxidation reactions, leading to the formation of a protective film
[14]. Therefore, as a sequence of corrosion processes, oxygen was consumed at the
metal/bentonite interface due to oxygen reduction reaction. So, oxygen diffusion
became very limited and decreased (diffusion control), see Figure 7 in appendix I.

If we assumed there is no passive film built on the metal surface due to corrosion
process in presence of bentonite, as that generally found due to corrosion process for
carbon steel in neutral and natural water [130]. So the existence of water in the
corrosive environment may cause further corrosion process and the cathodic reaction is
most likely water reduction (2H2O + 2e H2 + 2OH). The result is corrosion products
layers were built up on the metal surfaces; uniform corrosion is likely to be the
dominant form of corrosion for mild steel in bentonite [87]. It is will known that
bentonite is used as a barrier or sealing materials due to its low permeability and oxygen
faces some difficulty to penetrate through it [84, 85]. So, magnetite (Fe3O4) was also
expected to form due to the reduction reaction of FeO(OH): 8FeOOH + Fe 3Fe3O4 +
4H2O. Even though, the Fe3O4 crystal was not detected when XRD analysis was
performed on the specimens surfaces that were exposed to the bentonite for 7 days.

It is reported that [86] corrosion process for mild steel in bentonite clay is more
complicated than that in aqueous solutions. On the other hand, localised corrosion might
be expected to be occurred for mild steel in bentonite if it is passivated [86]. The
passive film is a few atoms thick and mainly composed of oxide or hydroxide of the
metal in its electrolyte. When passive film forms; any defects, scratches or breaks down
will be repaired rapidly. The aggressive and active anions such as chlorides have a very

118
strong chance to breaking down such films. As a result, corrosion occurs spontaneously
at localised points where may cause pitting or localised attack of the metal [4].

More investigation might be required to reveal the nature and behaviour of those
corrosion products. Therefore, XRD analysis has been employed in this study. At the
end of the experiment, the specimens were taken out from the cells, photographed
(Figure 6.1 (c) and (d)) and for further investigation XRD analysis was employed to
reveal the crystal structure of any unknown compound, and by the knowledge of its
crystal structure a chemical composition can then be inferred for compound.

6.2.2 Impressed Current Cathodic Protection Polarization Potential

The setup for this experiment was described in Chapter 5, Section 5.4. The potential of
pipes buried inside damp bentonite containing 45% (w/w) 0.5 M NaCl solution was
studied, both when the impressed current cathodic protection (ICCP) was ON and
instant OFF. Different current densities of 10.5, 50, 100, 150, and 200mA.m-2 were
applied (in ascending order) to the system. The variation in the ON and OFF potentials
vs. copper sulphate reference electrode (CSE) was measured after 1 day and after 7 days.
The results were shown in Figures 6.3 to 6.7. The values for those results are shown in
Appendix I, Tables 4 to 8, respectively. For protection the potential must be greater than
-850mV and usually less than -1300mV, see Section 3.4.

When cathodic protection (CP) is applied to the system, current flows into the structure
through the surrounding damp bentonite. The resistivity of bentonite wetted with a 0.5
M NaCl solution is given in Appendix 1, Table 1 and Figure 1. For water content 45%
w/w the resistivity was measured as 440cm.

The ON potential is the sum of the natural corrosion potential (OCP) before applying
the CP, the polarisation potential due to applying cathodic protection and the ground IR
drop [74]. When the CP system is de-energised, the pipe-to-environment potential
undergoes an instantaneous positive shift as a result of elimination of the IR voltage
drop in the damp bentonite. As previously described in Section 3.5.1.4, the potential
measured immediately after the interruption of the current source is the instant OFF
potential. This instant OFF potential is the sum of the natural corrosion potential and
cathodic polarisation potential.

119
Figure 6.3 shows the polarisation potential (CP ON/OFF) for 3 sections of pipe at the
depths indicated in damp bentonite containing 45% (w/w) 0.5 M NaCl solution, after 1
day and 7 days when the applied current density was 10.5mA.m-2.

The potential was measured after 1 day at 1cm intervals (0.0cm to 34.0cm depth) for the
3 sections of pipe. The average value of the ON potential for the length of pipe closest
to the surface, from 0cm to 10cm deep, was -1005mV, and the least negative value was
about -888 mV for the middle section (12.0 to 24.0cm inclusive) while the deepest
section of pipe between 24.0cm and 34.0cm deep had an average potential of -901mV.
The range of measured values over each length of pipe was such that no significant
difference in potential existed over any of the individual lengths though, of course, the
differences in potential between the three lengths were significant.

After 7 days immersion, average polarisation potentials for all three pipe sections had
shifted to slightly more negative values -1028mV, -902mV and -940mV respectively.
The difference in average ON polarisation potentials between the top and bottom
lengths of pipe decreased from 104mV on day 1 to 88mV on day 7.

The averages instant OFF potentials for the three sections of pipe, on day 1, were found
to be similar: -888mV for the first length of pipe, -863mV for the second length pipe
and -871 mV for the bottom pipe. Again it was found that the range and standard
deviations of measured values over each individual length of pipe were such that no
significant differences in potential existed over any of the individual lengths. However,
there was a significant difference in mean potentials between the three lengths of pipe.

After seven days immersion the corresponding instant OFF average potentials for the
three sections of pipe, were: -930mV for the first pipe, -884mV for the second pipe and
-912 mV for the bottom pipe. The changes in instant OFF potentials over the seven days
between corresponding lengths of pipe were small but significant, -42mV for the first
pipe, -21mV for the second pipe and -40mV for the bottom pipe.

There was no general trend for the potentials to become more or less negative with
depth for this current density.

The values for ON potentials shown in Figure 6.3 are more negative than the instant
OFF potentials at end of both day 1 and day 7. However, it is seen that after seven days
the values obtained for the instant OFF potential were greater than the nominal accepted
protection criterion (-850 mV (CSE)).

120
-600
ON pipe 1, 1d OFF pip1, 1d ON pipe 2, 1d
OFF pipe2, 1d ON pipe 3, 1d OFF pipe 3, 1d
-700 ON pipe 1, 7d OFF pip1, 7d ON pipe 2, 7d
Potential (mV /C SE ) OFF pip2, 7d ON pipe 3, 7d OFF pip3, 7d
-800

-900

-1000

-1100

-1200
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36
Depth (cm)

Figure 6.3: ON and OFF polarisation potential Vs. CSE for bare steel pipes buried in
bentonite containing 45 % (w/w), 0.5 M NaCl solution at different depths under CP at
10.5mA.m-2 for 1 and 7 days respectively.

Figure 6.4 shows the results for a similar test to that described above but here the
applied current density was 50mA.m-2. Again the potential was measured after 1 day at
1cm intervals (0.0cm to 34.0cm depth) for the 3 sections pipe. The first observation is
that both sets of ON and OFF average potentials decrease with depth. For the ON
potential: the most negative average value was about -1138mV for the uppermost pipe,
with -981mV for the middle pipe and -946mV for the bottom length of pipe. The
corresponding average instant OFF potentials were -933mV, -926mV and -889mV.

After 7 days immersion, average polarisation potentials for all three pipe sections had
shifted to slightly more negative values; the average ON polarisation potential for the
uppermost pipe had shifted to a slightly more negative value -1319mV. It can be seen
that for the middle and bottom pipes the average potentials after the seven days of the
experiment were -1052mV and 1041mV respectively. The average OFF potentials were;
-1002mV for the top length of pipe, -980mV for the middle and -956mV for the bottom
length of pipe.

The changes in ON potentials over the duration of the experiment were top length
181mV, middle 71mV, and 95mV for the bottom length. The corresponding changes for
the OFF potentials were 69mV, 54mV and 67mV.

121
For all three lengths of pipe there was a significant change between ON and instant OFF
polarisations with the ON polarization more negative than the instant OFF potential.
For the top length of pipe the difference in average potentials on day 1 was -205mV,
and on day 7 was -317mV. For the middle length of pipe the corresponding differences
were -55mV and 72mV, and for the bottom length of pipe 57mV and 85mV.

A general trend was observed for the potentials to become less negative with depth.

-700
ON pipe 1, 1d OFF pip1, 1d ON pipe 2, 1d
-800 OFF pipe2, 1d ON pipe 3, 1d OFF pipe 3, 1d
ON pipe 1, 7d OFF pip1, 7d ON pipe 2, 7d
P o ten tia l (m V / C S E )

-900 OFF pip2, 7d ON pipe 3, 7d OFF pip3, 7d


-1000
-1100
-1200
-1300
-1400
-1500
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36
Depth (cm)

Figure 6.4: ON and OFF polarisation potential Vs. CSE for bare steel pipes buried in
bentonite containing 45 % (w/w), 0.5 M NaCl solution at different depths under CP at
50mA.m-2 for 1 and 7 days respectively.

Figure 6.5 shows OFF/ON polarisation potential with depth for bare steel pipes buried
in the same environment as described above but with current density increased to
100mA.m-2. A similar variation in polarisation potential in mV/CSE over depth was
observed as previously. The average ON potential on day 1 for the uppermost pipe was
-1217mV, for the middle length of pipe was -1091mV and for the bottom length of pipe
was -1056mV. The corresponding average instant OFF potentials were -1000mV, -
994mV and -975mV.

On day 7 the corresponding average ON potentials were -1408mV, -1111mV and -


1076mV, and the corresponding average OFF potentials were -1003mV, -1000mV and -
967mV.

122
-700
ON pipe 1, 1d OFF pip1, 1d ON pipe 2, 1d
-800 OFF pipe2, 1d ON pipe 3, 1d OFF pipe 3, 1d
ON pipe 1, 7d OFF pip1, 7d ON pipe 2, 7d
Potential (mV / C SE) -900 OFF pip2, 7d ON pipe 3, 7d OFF pip3, 7d

-1000

-1100

-1200

-1300

-1400

-1500
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36
Depth (cm)

Figure 6.5: ON and OFF potential for bare steel pipes buried in bentonite contains 45 %
(w/w), 0.5 M NaCl solution at different depths under CP at 100mA.m-2 for 1 and 7 days
respectively.

A general trend was observed for the instant OFF potentials to be less negative than the
ON potentials and for potential to become less negative with depth. But the pattern of a
significant increase in average instant OFF potential between day 1 and day 7 is NOT
maintained. There was no significant difference between the average potential for the
uppermost pipe on day 1 and the uppermost pipe on day 7, nor between the uppermost
pipe on day 1 and the middle pipe on day 7. (T Test used for comparison of means,
level of significance 5%.)

Figure 6.6 shows OFF/ON polarisation potential with depth for bare steel pipes buried
inside the same environment as described above with current density increased to
150mA.m-2. The general trend observed above for the instant OFF potentials to be less
negative than the ON was again seen here. It was observed that the ON potentials
became less negative with depth, but this pattern was seen only for the day 7 readings
for the instant OFF potentials. For day 1 the instant OFF potentials became more
negative with depth.

In this case, for day 1, the average ON potential was -1257mV for the uppermost pipe, -
1173mV for the middle pipe and -11430mV for the bottom pipe. The corresponding
average instant OFF potentials were -1006mV, -1019mV and -1023mV. It should be

123
noted that the difference in mean OFF potentials between the middle and bottom pipes
are not significant here.

After day 7, the average ON polarisation potential for all three lengths of pipe appeared
to shift to more negative values, -1373mV for the top length of pipe, -1177 mV for the
middle length and -1147mV for the bottom section. However, while the shift in
potential for the top length between day 1 and day 7 was significant, the changes for the
middle and bottom lengths were not significant.

On day 7 the instant OFF potentials for the three lengths of pipe were -1036mV at the
top of the pipe, -1025mV for the middle length of the pipe and -1016mV for the bottom
section of pipe.

The differences in average ON values between day 1 and day 7 for top, middle and
bottom lengths of pipe are respectively; 121mV, 4mV, 4mV (these latter two
differences are not significant), and for the OFF values; 30mV, 6mV and 7mV.

-700
ON pipe 1, 1d OFF pip1, 1d ON pipe 2, 1d
-800 OFF pipe2, 1d ON pipe 3, 1d OFF pipe 3, 1d
ON pipe 1, 7d OFF pip1, 7d ON pipe 2, 7d
P o ten tia l (m V / C S E )

-900 OFF pip2, 7d ON pipe 3, 7d OFF pip3, 7d


-1000
-1100
-1200
-1300
-1400
-1500
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36
Depth (cm)

Figure 6.6: ON and OFF potential for bare steel pipes buried in bentonite contains 45 %
(w/w), 0.5 M NaCl solution at different depths under CP at 150mA.m-2 for 1 and 7 days
respectively.

Figure 6.7 shows OFF/ON polarisation potential with depth for bare steel pipes buried
in the same environment as described above but with current density increased to
200mA.m-2. The general trend for the instant OFF potentials to be less negative than the

124
ON was again seen here. However, unlike previous observations the potentials became
more negative with depth for day 1, though the general trend had re-asserted itself by
day 7. It was also observed that all potentials became more negative by day 7.

As previously the CP ON potential is more negative than the instant OFF potential in all
cases, but it is obvious is that there is greater variation in the results than for the smaller
current densities. The differences between average ON and instant OFF potentials on
day 1 were 26mV, 25mV and 34mV for the top, middle and bottom sections of pipe,
respectively. On day 7 the corresponding values were 300mV, 230mV and 197mV.

-750 ON pipe 1, 1d OFF pip1, 1d ON pipe 2, 1d


OFF pipe2, 1d ON pipe 3, 1d OFF pipe 3, 1d
-850 ON pipe 1, 7d OFF pip1, 7d ON pipe 2, 7d
P o ten tia l (m V /C S E )

OFF pip2, 7d ON pipe 3, 7d OFF pip3, 7d


-950
-1050
-1150
-1250
-1350
-1450
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36
Depth (cm)
Figure 6.7: ON and OFF potential for bare steel pipes buried in bentonite contains 45
% (w/w), 0.5 M NaCl solution at different depths under CP at 200mA.m-2 for 1 and 7
days respectively.

It can be observed from the above results that there is a significant change in
polarisation potential values between CP ON and OFF, and the shift was in a less
negative direction when the cathodic protection potential was removed, see Figures 6.3
to 6.7.

Figure 6.3, Figure 6.4 and Figure 6.5 are interesting; the ON/OFF potentials seem to
indicate a band of lower resistance in the two pipe sections below the top one; at higher
current densities the bands of low resistance disappear. One is tempted to suggest that
electro-osmotic movement of water can be at least part of the explanation for this [131].
Higher current densities would cause greater water movement, but this is only a

125
suggestion and it would need further investigation to determine whether the difference
between the ON and OFF potential can be explained by solution resistance effects.
However, the difference between the ON and OFF potentials will include errors
associated with the ON potential which may be attributed to the variation of resistivity
of the electrolyte and/or to the fact that potential measurements were made with the
CSE reference that may not be in good contact with the damp bentonite directly over the
pipe structure and human error during manual measurements.

The ON potential depends on the location of the reference electrode [40] and it is
necessary to make appropriate arrangements to ensure uniformity of currents and
potentials. In the literature [2, 40, 75], the ON potential does not change when the anode
is sufficiently far from the protected structure and the true potential of the structure does
not vary with the location of the reference electrode as there is no significant change in
the IR voltage (Remote earth). However, in our case it was not possible for the anode to
be mounted sufficiently remotely for it to be considered a Remote earth, it was placed
relatively close to the protected structure (cathode) and an important consequence was
that the ON potential changed both with location and with current. If this hypothesis is
correct it would be expected that the difference between ON and instant OFF potentials
would be approximately proportional to the current density.

If we consider only the middle length of pipe the variations in potential for the
uppermost length of pipe make it unsuitable here then we obtain the graphs below.

The relationship assumed is that the IR drop is proportion to electrode separation:

IR drop, mV = Resistivity, .m*Current density, mA.m-2 * length, m

IR drop is directly proportional to the current density if the geometry and resistivity
remain constant. The former is true but there was some change in resistivity between
day 1 and day 7. The gradients of the two lines in the figures imply that the resistivity
increased between day 1 and day 7, which agrees with the observation that the bentonite
tended to dry out slightly over the time of the experiment.

126
The difference in potential can be said to be a straight line relationship within
experimental error which is a good indication that the difference (at least for the middle
section of pipe) is due to a resistivity effect.

(a) Middle pipe day 1

(b) Middle pipe day 7.

Figure 6.8: Difference in ON instant OFF potentials (in mV) as a function of current
density (in mA/m2); middle pipe section only for days 1 and 7.

127
The solution resistivity was calculated, using the following equation:
Rs (.m) = (IR drop (mV)* distance between the reference electrode and cathode
(0.02m)/(current density (mA/m2)*Area of cathode (0.01843m2)).
Rs (m) = 1.085 (m-1) * IR drop (mV) / current density (mA/m2)

3
day 1 day 7
2.5
R s (o h m .c m )

1.5

0.5

0
0 50 100 150 200 250
-2
Current density (mA.m )

Figure 6.9: The average variation in the solution resistivity with applied current density
for the middle pipe in damp bentonite.

The equation on p128 is: Rs (m) = 1.085 (m-1) * IR drop (mV) / current density
(mA/m2)
For Middle pipe, day 7, 200 mA condition Rs (m) = 1.085 x 230/200 = 1.25 m, and
for Middle pipe, day 1, 200 mA condition Rs (m) = 1.085 x 230/200 = 1.25 m
The values of Rs calculated using the equation on p128 are the same for each condition.
That is why the points overlap in Figure 6.9.
The results show that the solution resistivity decreases with the increase of the applied
current density on both day 1 and day 7. After 7 days of exposure, the solution
resistivity increased slightly as would be expected if the bentonite underwent any drying
out.

Since the variation in the solution resistivity with the applied current density follow
almost the same trend, this phenomenon can be related to the decrease in resistance of
the working electrode. One is tempted to suggest that electro-osmotic movement of
water can be the explanation as the applied current may cause more water migration
towards the metal surface. This explanation is in agreement with a conclusion reached

128
by Schwenk [131], who suggested that the application of cathodic current may cause the
resistance to decrease which can be understood by electro-osmosis which is responsible
for the enhanced water uptake . To have a proof of this more work is needed. Further
studies are also needed to establish whether the difference between the ON and OFF
potential can be explained by solution resistance effects. Other researchers such as
Sekioka, et, al. [132] have linked the applied current and the soil resistivity, they noted
that as the applied current is increased the resistance decrease which is in agreement
with the current research outcome. They [132] went further and explained this by the
ionisation of the soil around the metal as a result of the applied current.

Figure 6.10 shows the variation in polarisation potential in mV/CSE due to applying
different current density in mA.cm-2 for 3 identical pipes buried vertically at different
depths in bentonite contains 45 % (w/w), 0.5 M NaCl solution. The values illustrated in
Figure 6.10 were taken from Appendix I, Tables 4 - 8 which was summarised from
Table 6.1 below. Two instant OFF potential were presented for each pipe after 1 day of
applying CP for each current density applied to the system; at 2cm and 10 cm for the
first section, at 12cm and 22cm for the second pipe and at 24cm and 34cm for the third
one. The most negative potential was recorded at the top of the first pipe at 2cm and the
less negative one was observed approximately at the lowest part of pipe 3, at about 34
cm depth at the most current densities applied to the system. However, when the current
applied to the system was increased, the polarisation potential was gradually decreased
at the surface of the first pipe compared with the negative potential at 34cm depth for
pipe 3 that has more negative potential particularly when 200mA.m-2 was applied.

-800
-820 2cm 10 cm 12 cm
-840 22 cm 24 cm 34 cm
Potential (mV/C SE)

-860
-880
-900
-920
-940
-960
-980
-1000
-1020
-1040
-1060
-1080
0 50 100 150 200 250
-2
Corrent density (mA.m )
Figure 6.10: Current density dependency in mA.m-2 for polarisation potential in

129
mV/CSE (at instant OFF potential, IR= 0) for steel pipes buried at different depths in
bentonite contains 45 % (w/w), 0.5 M NaCl solution after 1 day.
Table 6.1: Current density dependency in mA.m-2 for polarisation potential in mV/CSE
(at instant OFF potential, IR= 0) for steel pipes buried at different depths in bentonite
contains 45 % (w/w), 0.5 M NaCl solution after 1 day.

Depth At 10.5mA.m-2 50mA.m-2 100mA.m-2 150mA.m-2 200mA.m-2


pipe 1
2 -886 -940 -1011 -1010 -1030
10 -888 -928 -993 -1010 -1035
pipe 2
12 -859 -920 -982 -1012 -1001
22 -867 -931 -1000 -1015 -1010
pipe3
24 -875 -903 -980 -1025 -1034
34 -863 -872 -966 -1014 -1055

Figure 6.11 describes a similar variation in polarisation potential in mV/CSE due to


applying different current density in mA.cm-2 as that explained in Figure 6.10 except
that the time of polarisation was after 7 days instead of 1 day. The values illustrated in
Figure 6.10 were summarised in Table 6.2. The instant OFF potential recorded and
plotted is at 2cm and 10 cm for the first section, at 12cm and 22cm for the second pipe
and at 24cm and 34cm for the third one at the five different current density applied to
the system. In this case, it is obvious that, the more negative potential observed after 7
days of ICCP polarisation was at the top of the first pipe for the most currents applied to
the system and the less negative one was approximately at the lowest part of pipe 3.

-870
-890 2 cm 10 cm 12 cm
-910 22 cm 24 cm 34 cm
Potential (mV/CSE)

-930
-950
-970
-990
-1010
-1030
-1050
-1070
-1090
-1110
-1130
-1150
0 50 100 150 200 250
-2
Current density (mA.m )
Figure 6.11: The polarisation potential in mV/CSE Vs. current density in mA.m-2 for
identical pipes buried vertically at different depths in bentonite contains 45 % (w/w),
0.5 M NaCl solution after 7 day.

130
Table 6.2: Current density dependency for polarisation potential in mV/CSE (at instant
OFF potential, IR= 0) for steel pipes buried at different depths in bentonite contains 45
% (w/w), 0.5 M NaCl solution after 7 day.

Depth (cm) OFF potential in mV at different current density in mA.m-2


10.5mA.m-2 50mA.m-2 100mA.m-2 150mA.m-2 200mA.m-2
pipe 1
2 -917 -1013 -1030 -1052 -1085
10 -940 -980 -980 -1015 -1047
pipe 2
12 -883 -972 -983 -1021 -1049
22 -887 -991 -1002 -1028 -1048
pipe3
24 -911 -973 -984 -1020 -1034
34 -909 -933 -965 -1014 -1055

Figure 6.12 summarised the polarisation potential in mV/CSE Vs. current density in
mA.cm-2 for identical pipes after buried vertically at 2 different depths (2cm and 34cm)
in bentonite contains 45 % (w/w), 0.5 M NaCl solution for 1 day and 7 days at the top
of pipe 1 and at the bottom of pipe 3. As it is illustrated in this figure, It is obvious that
the more negative polarisation potential was observed for both, pipe 1 and pipe 3, after
7 days for the most currents applied to the system compared to the polarisation potential
recorded after 1 day of applying the CP. Nevertheless, the less negative polarisation
potential was observed approximately at the lowest part of pipe 3 at 34 cm depth for
both time of polarisation compared to the polarisation potential at 2cm depth from the
top of the first pipe which was almost more negative for the all current applied to the
system. The difference in polarisation potential at 2 different depths (2cm and 34cm) for
the different current density applied was 23, 80, 65, 38 and 30mV respectively.

131
-810
-830 2 cm After 1 d 2 cm after 7 d
-850 34 cm after 1 d 34 cm after 7 d
-870
P otential (m V /C S E ) -890
-910
-930
-950
-970
-990
-1010
-1030
-1050
-1070
-1090
-1110
-1130
-1150
0 50 100 150 200 250
-2
Current density (mA.m )

Figure 6.12: Graph showing plot of polarisation potential in mV/CSE over current
density in mA.m-2 for identical pipes buried vertically at different depths in bentonite
contains 45 % (w/w), 0.5 M NaCl solution after 1 day and 7 days.

Various levels of current density in the range of 10.5 to 200mA.m-2 were applied to the
protected system at different depths. Table 6.1 and Table 6.2, illustrating the values of
polarisation potential in mV/CSE (at instant OFF potential) for steel pipes buried in
bentonite contains 45 % (w/w), 0.5 M NaCl solution at different depths after 1 day and
7days respectively.

The minimum polarisation potential obtained for the three pipes at different depths was
about -860 mV, after 1 day and 7 days of polarisation. This potential was recorded when
minimum constant current was applied to the system; 10.5mA.cm-2. Polarization
potential of -850 mV criteria relative to CSE is suggested to be an adequate protection
can be achieved [62]. It is accepted by different standards and codes in corrosion
engineering science and field experience [1]. Thus, no corrosion products were
observed on the specimens surface at the end of the test.

On the other hand, the greatest OFF potential obtained was -1055 and -1085 after 1day
and 7 days of polarisation respectively, when 200mA.cm-2 was applied to the system.
As can be seen from values illustrated in the tables mentioned above, it is evident that
the potential was shifted to more negative with time (Figure 6.12 and Figure 6.13). The
shift of potential to more negative with time may be given an indication that a certain

132
products may be formed on the metal surfaces were gradually growing with time. It is
possible for any deposit to form on the metal surface while it is exposed to the electrolyte
solution. However, in order to create an active and uniform protective film, sufficient
cathodic protection potential must be applied. As hydrogen evolution can be the main
source for any defect in the protective film due to over protection potential.

Figure 6.13 shows the photographs for mild steel specimens (a) and (b) in two different
geometries. Those shapes (a) and (b) are 3 bare section pipes and 4 panel shape
specimens, respectively. Those specimens were buried in bentonite mixed with 45 %
(w/w) 0.5 M NaCl and CP was applied for 7 days for each current was applied to the
system. A film was observed after the specimens were taken out from the
electrochemical cells. The photograph in (a) shows a thick grey film that was formed
and covered almost the all external surface area of 3 section pipes. A similar film was
also observed and composed on 4 bare panels, (b). The time was an essential factor for
the specification and quality of any film that may compose on the metal surface. The
deposit that was formed on the metal surface was analysed by SEM and XRD technique,
using specimens in panel shape. Further results of SEM and XRD analysis will be
presented later in this chapter.

(a) after CP (at 200 mA.m-2) (b) CP, ( at 200 mA.m-2)


Figure 6.13: Samples buried in bentonite mixed with 45 % (w/w) 0.5 M NaCl, under
cathodic protection for 7 days.

XRD Analyses

Figure 6.14 shows the XRD pattern along with its chemical analysis list for corrosion
deposit where formed on the specimens surface. They were buried in bentonite
contains 45 % (w/w), 0.5 M NaCl solution at free corrosion potential for 7 days. As can
be seen, different compounds were illustrated in the pattern list. The main crystalline
material detected in two reflections on the specimens surface is mostly metal substrate

133
in the form of ferrite. A reflection that might be related to Goethite in the form of (-
FeOOH) and lepidocrocite (-FeOOH) was also detected. A weak reflection that may
come from Beidellite 12 A Clay in the form Na0.3 Al2 (Si, Al )4 O10 (OH )2 2 H2 O was
also detected.

O ra yith_ S 1 2 b

6000

4000

2000

0
10 20 30 40 50 60 70 80

P o sitio n [2 The ta ] (C o b a lt (C o ))

Visi-ble Ref. Sco-re Compound Displacement Chemical


Ref. Code Score Compound Displacement Scale Factor Chemical
Name [2Th.] Formula
6-696 43 -Fe 0.000 1.016 Fe
010810463 55 Goethite, 0.000 0.106 FeO(OH)
000441415 20 Lepidocrocite 0.000 0.041 Fe+3 O(OH)
000430688 21 Beidellite12 0.000 0.011 Na0.3Al2

Figure 6.14: XRD pattern for steel coupon exposed to bentonite mixed with 45% (w/w)
0.5 NaCl solution, at OCP potential for 7 days.

Figure 6.15 shows SEM image for 2 specimens buried in bentonite contains 45 % (w/w),
0.5 M NaCl solution under cathodic protection, 200mA.cm-2. This layer was only
observed after the specimens were taken out from the electrochemical cells at the end of
the tests. The specification and morphology of this film for this stage was briefly
analysed, using SEM/EDX and XRD technique. This work was extended to the next
stage of our study.

134
Figure 6.15: Secondary electron, SE, micrographs showing layers that formed due to
cathodic protection application for 7 days for 2 specimens buried in bentonite contains
45% of 0.5 NaCl solution.

6.3 Bare Mild Steel Exposed to 0.5 M Sodium Chloride Solution


Containing (Free of bentonite)

6.3.1 Introduction

The next stage of this research was decided to study:

1- Corrosion behaviour for mild steel exposed to sodium chloride solution, free of
bentonite.

135
2- Corrosion behaviour for mild steel exposed to sodium chloride solution mixed
with different contents of bentonite.

3- Corrosion behaviour for mild steel due to adding ZnCl2 to NaCl solution in the
presence of bentonite.

It was recommended to study the electrochemical behaviour of mild steel under natural
corrosion process and under cathodic protection when the previous material was used as
corrosive electrolytes. The rational explanation for these series of experiments has
already been provided early in chapter 5.

The main objective for carrying out those experiments was to study the characterisation
and behaviour of those films that may compose on the metal surface, consequently, their
influence on the level of protection of external well casings. As was mentioned early, to
achieve this study different electrochemical techniques have been employed;
measurements including weight loss, open circuit potential (OCP) coupled with
cathodic polarisation protection potential and polarisation resistance (Rp) were applied.

Cathodic protection was applied on mild steel exposed to bentonite at different


condition, using a potentiostat instead of power supply, to study the variation of current
over time. A polarisation potential at natural corrosion potential, -0.8 V and -
1.15V/SCE were applied, to protect mild steel coupons exposed to 0.5 M NaCl solution
environment. At the same time current was measured regularly.

As was mentioned earlier in chapter 5, section 5.1, 0.5M sodium chloride (NaCl)
solution free of bentonite was initially used as a blank solution. It was chosen as the
main aqueous corrosive environment due to the maximum salinity of brackish water is
less salinity than sea water which may result from mixing of fresh water with seawater
and their salinity can vary considerably over space and time; the average salinity is in
the range of 0.5 and 30 grams of salt per litre. Different weights of bentonite mixed
with 0.5M NaCl solution were used. Table 5.2, chapter 5, shows the various amount of
bentonite that mixed with 0.5M NaCl solution, which was used in this research. ZnCl2
as a cathodic inhibitor in two different concentration, 500 ppm (0.05%) and 1000ppm
(0.1%) was also used. It was recommended to study the electrochemical behaviour for
mild steel under natural corrosion process and/or under protection when the previous
inhibitor mixed with bentonite.

136
As it is pointed out in chapter 5, section 5.6, an appropriate corrosion allowance is
important to be predicted to guarantee the equipment life time [112]. Therefore, weight
loss measurement is one of the most effective methods used to evaluate the corrosion
rate (CR) [113]. Hence, series of tests were carried out. Then, CR for specimens
subjected to different environment was extracted from the weight loss measurements. A
series of experiment were carried out to study the free corrosion potential (Ecorr.)
behaviour for the casing subjected to different corrosive environment as described in
chapter 5, section 5.7. Individual experiments were conducted by using separate
corroding specimens, were prepared for natural corrosion potential measurements. The
results will be presented as potential versus time plots and it will be described in this
chapter.

Linear polarization resistance (LPR) technique was also applied, using ACM (GILL AC)
version 5. The LPR data obtained was used to support the results that obtained from
corrosion potential-time experiments. As it is illustrated in chapter 5 section 5.8.1 and
throughout the immersion time, several LPR measurements were made. Some of the
obtained LPR measurement were analysed and it will be presented in this chapter.

For reproducibility and to support the results, and guarantee the measurements for all
experiments; weight loss, Open Circuit potential (OCP) and polarisation resistance
measurements were repeated in high accuracy.

To support the previous electrochemical results, surface morphology of the metal


surfaces for the coupons being subjected to different polarisation potential (Natural
corrosion potential, -0.8mV/SCE and 1.15mV/SCE) was also investigated using the
following techniques:

Optical microscopy

SEM and EDAX

XRD

XRD analysis and SEM observation were employed, in order to recognise and analyse
the elements that might build up on the metal surface exposed to various different
corrosive electrolytes. Consequently, discover and illustrate the various layers that may
compose on the metal surface at different polarisation potential; at free corrosion

137
potential, -0.8 V and -1.15 V. Hence, compare with data obtained due to applying
electrochemical techniques in this study.

The data generated due to employing the previous techniques will not be presented all
in this study. However, some of the plots, images, Energy dispersive X-Ray analyses
(EDAX) quantitative data, spectrum, and patterns will be presented in this chapter. The
rest of the results will be presented in appendixes.

6.3.2 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution


Containing No Bentonite

As mentioned earlier in experimental work chapter 5, bare mild steel specimens were
immersed in a beaker containing 0.5 M sodium chloride solution, as a blank corrosive
electrolyte. The area of the specimens was 10.5 cm2. Measurements including the
weight loss, free corrosion potential (open circuit potential (OCP)) coupled with visual
observation and polarisation resistance (Rp) were periodically recorded for the
experiment that illustrated early. Each sample for the most experiments was placed in an
individual beaker. The progress of corrosion on the specimens surface was visually
observed by looking at them and then, a photograph was taken at the different
immersion time, to notice the progress of corrosion. At the end of the experiment, the
samples were also used for SEM/ EDAX and XRD analyses. Figure 6.14 (b), (c), (d), (e)
and (f) showing photographs for mild steel specimens, after they were exposed to 0.5M
NaCl solutions for 1, 3, 5, 7 and 14 days respectively. Since, for weight loss
measurements the most common exposure time used is in the range of 48 to 168h [113].
Therefore the tests were carried out for periods of 1, 3, 5, 7 and 14 days. The
experiments were conducted under stagnant, aerated conditions and at room temperature
and initial pH = 8. It is obvious that, the progress of corrosion products on the metal
surface at the different immersion time can be observed due to continuously change of
the appearance of specimens. The mild steel specimens observed in the photograph
were used for weight loss experiments, two duplicate specimens were immersed in a
solution for each period.

138
Figure 6.16 (a) shows photograph for duplicate specimens after immersed in the
corrosive solution for 1 day. Using the same procedure, other specimens have been
immersed in corrosive electrolyte free of bentonite for different period of time that was
mentioned early. In general, most specimens were corroded and appeared as a powdery
uniform orange porous rust and dark brown film cover the most areas, particularly for
specimens after immersing for more than 5 days. As shown in photographs, black layers
were also appeared, particularly for specimens exposed to the solution for long period.

(a) samples in 0.5M NaCl (b) After 1 day (c) After 3 days

(d) After 5 days (e) After 7 days (f) After 14 days


Figure 6.16: Photographs showing corrosion of mild steel specimens after immersed in
3.0% NaCl solution containing no bentonite for different periods of time.

In this study, one of our aims is the corrosion rate of mild steel under the same
conditions mentioned early was studied. An appropriate corrosion allowance is
important to be predicted to guarantee the equipment life time [112]. The weight loss
technique mentioned in chapter 5, section 5.6 was followed to calculate the corrosion
rate. It is reported that [18, 113] corrosion rate (CR) is not always constant with
exposure time. As a result, the variation of corrosion rate with time, might leading to
cause inadequate results [112]. The effect of time on corrosion rate was also
investigated by using specimens exposed to different corrosive environment for 14 days.

139
Figure 6.17 showing plot of variation of corrosion rate (CR)) over time for specimens
after 14 days of immersion. The corrosion rate over time values for Figure 6.15, which
was extracted from the weight loss measurements, data where presented in Appendix II,
Table 9. It is obvious that the average corrosion rate was about 0.270 mm/y. this value
was derived from the weight loss after 1 day of immersion. After that the corrosion rate
was found gradually decreased from the initial value rather than to increase until day 7
which was found about 0.163 mm/y. The decrease in corrosion rate might be due to the
continuous accumulation of corrosion products, where covering partially or completely
the specimens surface with certain corrosion products. Any corrosion products on the
specimens surface or coating faults, even if it is inadequate or permeable scale will
interfere with the oxygen mass transfer to the cathode side and, as a result, slow the
diffusion of metal ions from the anode side and, hence, will slow down corrosion.
Consequently, the corrosion rate is gradually decreased by the growth of corrosion
deposits on the metal surface. However, the corrosion rate was increased after day 7,
which was found to be closed to the initial value (0.206 mm/y).

The instability in corrosion rate over time might be attributed to the instability of the
corrosion product on the metal surfaces due to the change in the characterisation of
corrosive electrolyte. The dropped off in the bulk solution pH over time to somewhat
from the initial pH value can lead to instable corrosion rate. The drop in pH might be
due to the carbonic acid formation, as atmospheric CO2 dissolved in the solution. It is
reported that [3], the corrosion products are more soluble and not remain on the metal
surface in the acidic solution. Hence, the metal surface is soon again exposed to the
electrolyte, more attack and corrosion continues. Thus the conductivity of the corrosion
medium might also increased due to the soluble corrosion products, so corrosion rate
may be increased by time [17].

For accurate measurements, the average corrosion rate obtained by weight loss
measurements, conducting for several times. The results were shown in appendix II,
Table 9. Then, optical microscopy was also employed to assess the morphology of the
substrate.

140
0.35

0.3
C o rro sio n ra te (m m /y ) 0.25

0.2
0.15

0.1

0.05 Sample 1 Sample 2 Avg. C.R


0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.17: Graph showing plot of corrosion rates (CR) (mm/y) over time determined
by weight loss measurements for bare mild steel specimens after immersed in 0.5M
NaCl solution, at OCP. The data extracted from weight loss experiments for 1, 3, 5, 7
and 14 days respectively.

6.3.2.1 Optical Microscopy of Specimens Immersed in 0.5M NaCl


Solution

Optical Microscopy is suggested to partially confirm the corrosion product type by the
colour changes during corrosion process. It was employed for more substrate
morphology assess and to examine the nature of corrosion products on the metal surface
before and after immersion for different exposure period. The magnification used was
20X. Figure 6.18 (a) shows a photograph for a sample for specimens surface before
immersing in the solution. It is used for weight loss experiments (silver grey colour), b,
d, f, h, and j for samples after immersed in the 0.5 M NaCl for a periods of 1, 3, 5, 7 and
14 days before the corrosion products were cleaned and c, e, g, i and k, after the
corrosion product was cleaned by Clarkes solution, applying ASTM standard (G1- 03)
[105]. The procedure was illustrated in chapter 5, section 5.6.

It is obvious that the progress of corrosion can be observed due to continuously change
of specimens surfaces that was exposed to the same solution for different periods. The
corrosion product on the surface for most samples before cleaning has almost the same
form and colour. It is reported in the literature that [112], one of the advantages of
carbon steel is localised corrosion is rarely expected to take place in neutral and natural

141
solutions. General corrosion due to oxygen reduction will take place and/or water
reduction in the absence of oxygen may also occur if corroded metal is not passivated.
However, the morphology for all specimens after cleaning was corroded and attacked at
almost exposed areas. Even though, pitting and chickenpox appearance can also be
observed in the most specimens subjected to the solution. More attack and pitting were
started to be increased by the time of immersion and can be exhibited, especially for
samples immersed in solution for more than 3 days.

100 m

(a) As polished

100 m 100 m

(b) 1day before cleaning (c) 1day after cleaning

142
100 m 100 m

(d) 3 days before cleaning (e) 3 days after cleaning

100 m
100 m

(f) 5 days before cleaning (g) 5 days after cleaning

100 m 100 m

(h) 7 days before cleaning (i) 7 days after cleaning

143
100 m 100 m

(j) 14 days before cleaning (k) 14 days after cleaning


Figure 6.18: Optical micrographs; magnification is 20x for mild steel specimens after
immersed in 0.5 M NaCl solution. (a) as polished, b, d, f, h and j for specimens before
cleaning from corrosion products, for 1, 3, 5, 7 and 14 days respectively and c, e, g, i
and k after cleaning for specimens for 1, 3, 5, 7 and 14 days respectively.

In this study, OCP varied with time was recorded. Then it is followed by LPR
measurements. The corrosion potential was measured with respect to saturated calomel
electrode (SCE). To avoid a drawing of any current from the system during the
measurements [107, 111], a high impedance, more than 106 voltmeter, Fluke73
Model was used. A set of 15 specimens, polished to a mirror finish were used for this
experiment; 3 replicate specimens for each test for 1, 3, 5, 7 and 14 days. The area of
each specimen was 10.5 cm2. Measurements were taken for the OCP and LPR at 1, 3, 5,
7, and 14 days at regular interval time.

Figure 6.19 shows the plot of natural corrosion potential variation over time for
polished bare mild steel specimens after immersed in 0.5 NaCl solutions for 1, 3, 5, 7
and 14 days. The values for corrosion potential are shown in appendix II, Table 10. In
this case, the open circuit potential was measured after 5 minutes, -0.350 V. It is started
to drift quickly into the negative direction as a result of spontaneous corrosion
(electrochemical) reaction process (oxidation and reduction reaction); the anodic
(oxidation) reaction is probably oxidation of mild steel to produce ferrous ions (Fe
Fe2+ + 2e-), whether the initial and most important cathodic reaction is oxygen reduction
(O2 + 2H2O + 4e- 4OH-). It has been reported that [14, 112] bare mild steel materials
is not stable and rapidly corrodes in an aqueous environment and Fe(OH)2 film may

144
form on the specimens surface exposed to neutral to basic electrolytes. This film is
relatively highly soluble material and cannot form a protective film on the corroded
metal surfaces; however insoluble solid corrosion products can form on the metal
surface due to further oxidation reactions, leading to form partially or totally a
protective film on the metal surface. The colour of the specimens was observed to
change from silver grey to green yellowish and corrosion deposit was appeared at the
top of steel sample.

After 1 day of immersion, the surface of steel specimens is mostly covered by bright
brown or orange deposit and surrounded by black colour products at the edges. The
potential was gradually shifted to more negative to settle at about -0.687 V and further
visible corrosion products were appeared on the mild steel surfaces and at interface
between the solution and metal surface. It is mostly dark brown in colour was observed
on the surface. After day 3 until day 7, soluble corrosion products were developed in the
bottom of the beaker and the water became yellowish and the corrosion potential falls to
more negative to settle at about -0.721V. Then the potential shifted toward positive
potential to reach a steady state value at about -0.693V after 14 days of immersion. The
difference in potential from day 7 to day 14 was about 28 mV.

The magnitude of this shift might be as a result of insoluble ferric hydroxide (FeOOH)
film formed on the specimens surface [16] or due to formation of a passive film.

So, the formation of such corrosion products on both anodic and cathodic sites or in
one of them, acting as a physical barrier against oxygen diffusion towards exposed
metal surface. Hence, anodic dissolution of metal can be reduced, resulting in lower
corrosion rate [20]. More investigation was carried out to reveal the nature and
behaviour of those corrosion products, using XRD and SEM/EDAX analysis. The
results will be presented in section 6.3.4.

The bare mild steel specimens were removed from solutions after the end of tests, dried,
photographed and stored in a desiccator for more investigation. The progress of
corrosion product over time for specimens exposed to the 0.5 M NaCl solution was
photographed and shown in Appendix II, Figure 8. For reproducibility, the OCP
experiment was conducted for several times. The values for previous Figure were
extracted from Appendix II, Table 10.

145
-300
-350 Sample 1 Sample 2 Sample 3 average
P o ten tia l V s. S C E (m V ) -400
-450
-500
-550
-600
-650
-700
-750
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.19: Corrosion potential for mild steel specimens immersed in aerated 0.5 M
NaCl solution at OCP.

Figure 6.20 shows the initial pH solution, the pH was about 8. The bulk solution pH was
found to be decreased from the initial values rather than to increase. The pH decrease is
most likely, due to the carbonic acid formation. As mentioned earlier in the
experimental part, section 5.3.4, the initial pH of the solution for all experiments
conducted in this project was adjusted to a value around 8. Bearing in mind that pH was
monitored for free corrosion potential and weight loss tests over the period of the
experiment.

Figure 6.20: The pH for solution at initial time of the experiment (8).

146
Corrosion rate in mm/y was also obtained, using RP values produced by linear
polarisation resistance measurement (LPR). The rationale for LPR procedure has
already been provided in section 5.8. The LPR measurement was run at regular intervals
at freely corrosion potential, repeated for several times (minimum 3 times), to guarantee
the measurements have been repeated in high accuracy. Figure 6.21 shows the plot of
Rp variation with time. The Rp values presented in this graph have been produced from
linear polarisation resistance measurement (LPR runs) over time; examples for Rp
curves were presented in Appendix II, Figure 9 to 14 and the data extracted from the
plots was also shown on Table 11. The initial Rp was recorded after 1h (day 0) of the
test. It was found to be slight increase from 2705 to 2742 Ohm.Cm2 over 1 day. It rises
by 273 Ohm.cm2 after 3 days to stabilise at 2978 Ohm. Cm2. Rp was continuously fall
down after day 3 to settle at range 2422 to 2175 Ohm.cm2, for 5 day to 14 days period.

4000
3500
3000
R p (O h m .c m2 )

2500
2000
1500
1000
500 Sample 1 Sample 2 Sample 3 average
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day

Figure 6.21: Graph showing plot of Rp values over time for mild steel specimens after
immersing in an aerated 0.5 M NaCl solution without bentonite for 14 days.

As was mentioned above, the results of the average values of Rp produced from linear
LPR measurement were used to calculate the corrosion current (icorr.) in A.cm-2 and
corrosion rate (CR) in mm.y-1 [18]. Evaluation of the proportionality constant B and Rp
values, leading to calculate corrosion current density by using equation 5.3, section 5.8
[19, 119]. Faradays law has been used to determine the corrosion rate from icorr., using
equation 5.7, page 103.

147
The corrosion rate in mm/y was also determined, using the experimental weight loss (W)
measurement by applying equation 5.1 in page 98. Then, icorr. in A.cm-2 was also
determined, using Faradays law (equation 5.7). The results were used to determine the
actual coefficient B at different period of immersion, using equation 5.3.

Since cathodic reaction is probably controlled by diffusion of oxygen reduction reaction,


a theoretical value for coefficient B calculated from arbitrary ba and bc values can be
predicted, a value of 0.026V/decade is frequently used [40, 120].

Comparison was made between the actual Stern-Geary constant (B) extracted from the
weight loss measurement with the assumed Stern-Geary constant, 0.026. The value was
used to calculate the corrosion rate from the results of the average values of Rp obtained
from LPR measurement. Comparison between the corrosion rates obtained from LPR
and weight loss measurement has been made.

Table 6.3 gives a summary for average corrosion current values (icorr.) in A/cm2,
corrosion rate values (CR) in mm/y estimated from LPR measurement and corrosion rate,
CR exp., corrosion current (icorr. exp.) and constant B obtained from weight loss
measurements. Interestingly, it is quite considerable different between the values for CR
exp., icorr. exp. and constant B obtained from the experimental weight loss measurements
compared to those were evaluated from LPR measurement, particularly for tests after 1
day and 3 days .

The trapezoidal rule was applied to determine the time-weighted average value. It was
found by determining the total area under the curve from day 0 to day 14 and then
dividing by the total time (14 days).

The assumed value for day 0 was determined by extending line joining the value for
day 3 to day 1 back to the CR , icorr, etc. axis. The intercept was taken as the value of the
relevant parameter for day 0.

It was noted that the contribution of the day 0 value to the time-weighted average was
not substantial because the base of the first trapezium was only 1 day. Thus taking CR
for example, a relatively large change in the day 0 value from 0.118 to 0.130 caused the
time-weighted average to change by less than 1%.

148
The values of corrosion rate and corrosion current were almost constant with immersing
time. Even though, the values for CR and icorr., obtained from the experimental weight
loss measurements are greater than those were obtained from LPR measurement. The
average value of constant B was also evaluated. It was ~2 times greater than that has

been assumed to calculate the corrosion rate values obtained from the linear slopes of Rp
curves (0.026V/decade). Clearly, lower corrosion rate was exhibited by using the LPR
measurement compared to that were obtained by experimental weight loss
measurements. Corrosion rate obtained from weight loss measurement is function of
exposing time change. Conversely, corrosion rate obtained from LPR measurement is
produced instantly in a small time. Corrosion rates as measured by weight loss are
approximately twice that obtained by LPR. This is hardly surprising since LPR is an
instantaneous measurement whereas weight loss is cumulative.

A continuous change in our corrosive electrolyte may be occurred due to the natural
dissolve of CO2 gas present in the atmosphere to produce carbonic acid [36].

So, it can conclude that, the formation of corrosion product on the metal surfaces was
expected to act as a barrier and will increase the polarisation resistance to somewhat.
However, the Rp values for the specimens were tested at OCP were in the range of 2705
to 2175 Ohm.cm2 which was considered almost same.

Table 6.3: Summary table of parameters determined for mild steel specimens after
immersing in aerated 0.5 M NaCl solution free of bentonite for 14 days at OCP.

Time C.R icorr CR exp. icorr. exp. Bexp.


-1 2 -1 2
day mm.y (A/cm ) mm.y (A/cm ) (V/decade)
Assumed value Day 0* 0.118 10.34E-06 0.280 23.5E-06 0.062
1 0.112 9.88E-06 0.270 2.3E-05 0.063
3 0.100 8.97E-06 0.251 2.2E-05 0.066
5 0.123 1.08E-05 0.175 1.5E-05 0.036
7 0.100 8.12E-06 0.163 1.4E-05 0.046
14 0.130 1.143E-05 0.208 1.8E-05 0.039
Average value using trapezoidal 0.113 9.72E-06 0.177 17.58E-06 0.048

As it was mentioned above, after the tests, specimens were removed from the
electrochemical cells and safely stored in a desiccator (RH) to avoid any contamination.
The specimens were taken for further investigations. XRD and SEM/EDAX quantitative

149
analysis were carried out. The results will be presented shortly, together for the
specimens after exposing to 0.5 NaCl solutions at natural corrosion potential and under
cathodic polarisation potential, at -0.8V and -1.15V/ SCE.

6.3.3 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution with
No Bentonite under CP at -0.8 V and -1.15V

The rational for the following series of experiments has already mentioned early in
section 5.5 to 5.11 in chapter 5. Briefly, cathodic protection was applied to protect bare
mild steel specimens exposed to aerated 0.5 M NaCl solution. The electrolyte was free
of bentonite and the initial pH value was adjusted to 8, using 0.1 M NaOH solution. The
resistivity was also measured, using soil box procedure; () = 23 .cm at room
temperature. In this study, NaCl solution was used as a blank corrosive electrolyte to
interact the results obtained from those tests with the other results that achieved during
the course of this study; by using different corrosive environment; 0.5 M NaCl solution
contains bentonite and zinc chloride at different concentration will be used.

Different criteria have been proposed in literature to present both the experiments and
field experience potential values [133]. A Polarization potential of -0.780 V criteria
relative to SCE with cathodic protection applied to the system is suggested to be an
adequate protection can be achieved. This value is accepted by different standards and
codes in corrosion engineering science and field experience including NACE Standard
[1, 134]. The second stage for this experiment is cathodic polarization potential at
0.8mV vs. SCE was firstly applied. Then, an exactly similar set of experiments was
performed with cathodic polarization potential of 1.150V against SCE. Applying quite
high initial potential significant reduction in current density can be achieved as a result
of growing up any film a protected metal surfaces [44].

Figure 6.22 shows variation in current density (CD) in mA.m-2 over time in days for
polished bare steel specimens immersed in 0.5 NaCl solutions. Current that measured
and recorded regularly due to applying -0.80 V polarisation potential was plotted in CD.
The values for current density are shown in appendix II, Table 13. The first current
measured after the sample was immersed will not be presented in the graph because it
was enormous value compared to the latter one. The first Current density that measured
after 1 h, expressed as day 0 in the figure. It was 247mA.m-2. A much faster decline in

150
current density was recorded during the first day of polarisation. The current was
dropped swiftly to 147mA.m-2. Then, the system shows a steady drop in current density
over the other periods of the experiment. A slight drop in CD was observed in the next 7
days with a fluctuation up to ~ 24mA.m-2, to stabilise at about 123.2mA.m-2. The

specimens exposed to the pure sodium chloride solution at -0.8V polarisation potential
were observed. The specimens surface was examined; the silver grey colour
appearance was changed to dark grey. There was no products on the specimens surface
can be revealed with the naked eye. For reproducibility the experiment was conducted
for several times. The current density over time values was extracted from Table 13 in
Appendix II.

275
250 Sample 1 Sample 2 Sample 3 Average
225
200
C D (m A .m- 2 )

175
150
125
100
75
50
25
0
0 1 2 3 4 5 6 7 8
Time (day)

Figure 6.22: Graph showing a plot of current density in mA.m-2 against time in days for
mild steel after immersed in aerated 0.5 M NaCl solution free of bentonite, at -
0.80V/SCE cathodic polarisation potential.

Figure 6.23 shows variation in current density (CD) in mA.m-2 over time in days for
polished bare steel specimens immersed in 0.5 NaCl solutions at -1.15V cathodic
polarisation potential. Values for current density are shown in Appendix II, Table 14.
The first current measured after the sample was immersed will not be presented in the
graph because it was enormous value compared to the latter one. The first Current
density that measured after 1 h, expressed as day 0 in the figure was 2584 mA.m-2.
The biggest change in current was taken place during the first day of polarisation. The
current dropped swiftly by more than 68.0% to settle at 1733mA.m-2. Then, a slight
drop in CD was observed in the next 7 days to stabilise at about 488mA.m-2. Finally, a

151
small shift was recorded at the end of day 14 of polarisation, and the final current
density recorded was quite high, 417mA.m-2. The specimens exposed to the pure
sodium chloride solution at -1.15V polarisation potential were observed and it was
exhibited that the shiny silver grey colour, unaffected, and there is no film on the
specimens surfaces under CP. The high current density recorded may be due the metal
surfaces were exposed directly to the solution and no barrier against oxygen diffusion.

As was expected, the SEM and XRD analysis, steel only was detected and no
reflections for any products were revealed for specimens have been under cathodic
protection, at -1.15V for 14 days. For reproducibility the experiment was conducted for
several times.

3000
2750 Sample 1 Sample 2 Sample 3 Average
2500
2250
C D (m A .m- 2 )

2000
1750
1500
1250
1000
750
500
250
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.23: Graph showing plot of current density in mA.m-2 over time in days for mild
steel specimens immersed in 0.5 M NaCl Solution under CP, at -1.15V/ SCE.

Figure 6.24 shows the pH variation with time. The initial bulk solution pH was about 8.
When the CP was applied, the pH was found to be increased somewhat from the initial
value. In 0.5 M sodium chloride solution, the pH increased by, at most 1.5, but at the
end it settled at 9.2. This increase in pH can, most likely due to the formation of OH-
ions, caused by the cathodic reaction at the working electrode causing pH to increase at
the interface of metal surface (cathode) and electrolyte. The pH values for 3 specimens
for 14 days were provided in Appendix II, Table 15.

152
9.8
9.6
9.4
9.2
9
pH

8.8
8.6
8.4
8.2
pH for sample 1 pH for sample 2
8 pH for sample 3 Average pH
7.8
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.24: Graph showing pH- Time plots for 3 specimens, after immersed in 0.5 M
NaCl solution for 14 days, under CP, at -1.15V/ SCE.

Figure 6.25 (a) and (b) and (c) show a photograph for bulk solution at the initial and at
the end of the experiment consequently.

(a) Initial pH (b) After CP (c) After CP


Figure 6.25: photograph shows pH for bulk of 0.5 M NaCl solution for samples under
CP, at -1.15V/ SCE,(a) Initial pH for solution = 8, (b and C) pH =9.3 after CP.

Linear polarization resistance (LPR) technique was also applied, using ACM (GILL AC)
version 5. The LPR data obtained was used to support the results that obtained from
corrosion potential-time experiments and by applying the cathodic protection technique
(The LPR measurement for specimens under cathodic protection was applied at OFF
potential at the end of each experiment), to emphasize that the drop off in cathodic
protection current applied to the system over time is due to the film formation that

153
formed on the metal surface rather than any other influences. As it is illustrated in
chapter 5 section 5.8.1 and throughout the immersion time, several LPR runs were made.
Some of the LPR measurements were analysed and it will be presented in this chapter.

The LPR experiment was also carried out for the specimens exposed to 0.5 M NaCl
solution and was under cathodic polarisation potential, at -1.15V/ SCE for 14 days (The
LPR measurement for specimens under cathodic protection was applied at OFF
potential at the end of each experiment). The CP system was switched OFF before any
LPR experiment was run to enable the system to reach stability. The following will
represent the most results obtained for specimens exposed to the corrosive electrolyte
that was free of bentonite.

Figure 6.26 shows a potential in mV versus current density in mA.cm-2 for mild steel
specimen immerse in aerated 0.5 M NaCl solution for 14 days at polarisation potential -
1.15V. The polarisation resistance (Rp) value presented in this graph was generated by
LPR run for the test at OFF CP. It was found to be about 2376.9 Ohm.Cm2.

The LPR was also run for two more samples were exposed to cathodic polarisation
potential at -1.15 V for 14 days and for more accuracy, the LPR was run and repeated
for several times for the same specimen as well. The polarisation resistance (Rp) plots
for those tests were presented in appendix II, Figure 15, 16 and 17 for specimen 1, 2 and
3 respectively. The polarisation resistance (Rp) values for those specimens were
illustrated in Table 6.4 below.

Table 6.4: Rp values in Ohm.cm2 for 3 mild steel specimens after immersing in aerated
0.5 M NaCl solution exposed to cathodic polarisation potential at -1.150 V/ SCE for 14
days. The test was run for 3 times for each sample at OFF CP.

Sample 1 (. Cm2) 2 (. Cm2) 3 (. Cm2) Av. (. Cm2)


Run 1 2376 2340 2901 2539.0
Run 2 2273 2211 2241 2241.6
Run 3 2120 2163 2057 2113.3

154
Figure 6.26: Potential in mV versus current in mA.cm-2 for mild steel immersed in
aerated 0.5 M NaCl solution under cathodic polarisation potential at -0.8V for 14 days.

As it is shown in Table 6.4, after 14 days of immersion for the 3 specimens were under
cathodic polarisation, a slight change in polarisation resistance values can be observed.
The Rp values for the same sample decreased after each run. However, the Rp for the all
specimens were immersed in NaCl solution free of bentonite and exposed to cathodic
polarisation for 14 days have almost same values; about 2273, 2211and 2241Ohm.cm2
for specimens 1, 2 and 3 respectively after the second run LPR. In general, the Rp values
for specimens were under cathodic polarisation potential and that were reported in
section 6.3 for specimens were exposed to the same corrosive solution at natural
corrosion potential (OCP), for the same period of time have approximately same
behaviour and the Rp was a little bit higher in the case of OCP tests.

So, it can conclude that, a shiny and clean surface for specimens was under cathodic
protection compared to those were under free corrosion potential. But in the case of
natural corrosion potential experiment, the formation of corrosion product on the metal
surfaces was expected to act as a barrier and will increase the polarisation resistance to
somewhat. However, the Rp values for the specimens were tested at OCP were in the
range of 2705 to 2175 Ohm.cm2 which was considered almost same as for the
specimens were under cathodic protection.

155
6.3.4 SEM/EDAX and XRD Analyses of Specimens Immersed in 0.5 M
NaCl Free of Bentonite

A brief introduction of SEM/EDAX technique was illustrated in section 5.9. The SEM
technique was employed for this experiment to reveal different cations that may be
accumulated on the metal surface, after exposing to the corrosive electrolyte with and
without cathodic polarisation potential. Then a quantitative analysis for those elements,
and their progress with time on the metal surface was also investigated, by using EDAX
technique.

Figure 6.27 shows two images in two different magnifications (500x and 1000x),
spectrum, and EDAX quantitative data for specimen that was exposed to 0.5 M NaCl
for 14 days. No cathodic protection was applied for this sample. As can be seen
different elements were listed in the spectrum for this sample in different percentages.
The highest percentage that was detected by EDAX is mostly iron, which was about
69.35%, oxygen at about 17.39%. Sodium and chloride elements at different percentage
were also detected; 7.86% and 5.39% respectively.

156
Figure 6.27: SEM and EDAX element analysis for steel metal exposed to 0.5 M NaCl at
OCP.

Figure 6.28 shows SEM/EDAX analysis for two images at 500x and 1000x
magnification and element analysis for specimen exposed to 0.5 NaCl solution, under
cathodic protection (-1.15 V) for 14 days. As can be seen, the only elements were
detected and showed in the spectrum for this sample was iron and carbon. The images
and EDAX analysis for the sample was immersed in 0.5 M NaCl, free of bentonite can
be used as an evidence that no films or any other products can be accumulated on the
metal surface when CP at (-1.15V) was applied. Consequently, the average current
density recorded from the results that was obtained due to the experiments that were
conducted in the laboratory, using 0.5 M NaCl solution free of bentonite was quite high;
123mA.m-2, and 417mA.m-2 for samples exposed to more than 7 days, when -0.8V and -
1.15V as a constant polarisation potential was applied to the system respectively. So, we
can conclude that, as it was expected, no scale formation was detected on the
specimens surfaces exposed to the solution free of bentonite due to applying cathodic
protection at any range. Hence, there is no indication for current density to decrease
over time.

157
Element Series C norm. [wt.-%]
Fe 26 K-series 91.1898
O 8 K-series 8.8101
Total: 100.00
Figure 6.28: SEM and EDAX element analysis for specimen exposed to 0.5 NaCl
solution without bentonite, under cathodic protection (-1.15 V) for 14 days.

The XRD technique was used, in order to recognise and illustrate the various
compounds that may form on the metal surface when they were exposed to 0.5 M NaCl
solution with and without bentonite for different period of time, at OCP, and under CP.
Furthermore, the results achieved by using XRD and SEM/EDAX techniques were also

158
used to support the results that were attained from the other techniques used in this
study.

Figure 6.29 shows XRD pattern with its chemical analysis list for mild steel species
exposed to 0.5 M NaCl solution at OCP. As can be seen, various products were listed in
the pattern list below. The crystalline material on the specimens surface is mainly halite
in the form of NaCl and metal substrate in the form of ferrite (Fe). The pattern was also
shown a reflection that may be related to goethite (-FeOOH), lepidocrocite (-FeOOH)
and Magnetite, Fe3O4.

The general conclusion can be made is, the spontaneous corrosion reaction process is
the anodic (oxidation) reaction which is probably oxidation of mild steel specimens to
produce ferrous ions (Fe Fe2+ + 2e-), whether the initial and predominant cathodic
reaction is oxygen reduction (O2 + 2H2O + 4e- 4OH-). It was mentioned early that,
bare mild steel is not stable and rapidly corrodes in an aqueous environment and Fe(OH)2
film may form on the specimens surface exposed to neutral to basic electrolytes [14].

The Fe(OH)2 film may form on the metal surface as a green rust, as a transitional
corrosion products according to the following reaction, Fe2+ + 2 OH- Fe(OH)2. The
Fe(OH)2 itself is relatively highly soluble material and cannot form a protective film on
the corroded metal surfaces. Such phenomenon can be easily interpreted to the
corrosion rate of the metal is almost constant [17]. See Figure 2.8(1) in chapter 2.
However, gradually, insoluble solid as a corrosion products were formed on the metal
surface due to further oxidation reactions [15]. The insoluble film that composed on the
metal surface may caused the corrosion rate to decline with time[31]. A slight drop in
corrosion rate for the first 7 days was illustrated in Figure 6.15.

6 Fe(OH)2 + O2 2 Fe3O4 + 6 H2O (2.23)

4 Fe(OH)2 + O2 4 FeOOH + 2 H2O . (2.24)

In general, FeOOH can have different crystalline forms of corrosion products [16];
lepidocrocite (-FeOOH) which was one of the compounds that was detected on the
metal surface. It is considered as one of the first crystalline corrosion products that may
form, but the high solubility of lepidocrocite (-FeOOH) products which is about 105
times higher than that of goethite -FeOOH [16]. So, the weak reflection for the pattern

159
of -FeOOH shown in Figure 6.29 may attributed to the high solubility of that crystals
which might part of that products converted to goethite (-FeOOH).

The dark green colour that can be revealed in the optical micrographs for corrosion of
steel specimens after immersed in 0.5 M NaCl solution shown in Figure 6.16 page 132
before and after cleaning is might be related to magnetite, Fe3O4 [88]. Fe3O4 was
detected on the specimens surface and proved by XRD analysis. As was mentioned
early, excluding the first crystalline corrosion products in the form of -FeOOH, the
forms Fe(OH)2 forms is believed to be insoluble and stable corrosion products in the
absence of O2 [88]. Even though, Fe3O4 might be formed according to the following
reactions [14]:

3 Fe(OH)2 Fe3O4 + 2 H2O+ H2 ........ (2.23)

8FeOOH + Fe 3Fe3O4 + 4H2O . (2-25)

Hence, it is obvious from the corrosion rate and polarisation resistance trends shown in
Figure 6.17 and 6.21 respectively were varied with time. After immersing the mild steel
in 0.5 M NaCl with no cathodic protection, there is no visible sign for any significant
protective film that may built in. This can be attributed to the instability and
continuously falling of corrosion products that from on the metal surface. The corrosion
products can be easily observed to settle at the bottom of electrochemical cells. Figure 8
(a) (f) in Appendix II, demonstrating continuous soluble corrosion products in the
bottom of the beakers which has an orange colour.

160
40000

10000

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C o p p e r (C u))

Ref. Scoe Compound Displacem-ent Scale Chemical


Code Name [2Th.] Factor Formula

000050628 66 Halite 0.000 0.747 NaCl


000060696 51 ferrite, 0.000 0.366 Fe
000290713 42 Goethite 0.000 0.023 -FeO(OH)
000190629 74 Magnetite 0.000 0.315 Fe3O4
000080098 56 Lepidocrocite 0.000 0.056 - FeO (OH)

Figure 6.29: XRD patterns for steel coupon immersed in 0.5 M NaCl at natural
corrosion potential.

Figure 6.30 and Figure 6.31 show X- Ray pattern with its chemical analysis list for mild
steel specimens exposed to 0.5 M NaCl solution for 14 days, with CP at -0.8V and -
1.15V, respectively. As can be seen from the patterns in both figures mentioned above,
a reflection that corresponds to the metal substrate in the form of ferrite (Fe) steel was
only identified for both; specimens were protected at -0.8V and -1.15V, respectively.
Though, The XRD and SEM/EDAX analysis, steel only was detected and no
unidentified reflections for specimens were been cathodically protected.

161
C o unts
O ra yith_ 3 _ Na c l
90000

40000

10000

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C o b a lt (C o ))

Visible Ref. Code Score Compound Displacement Scale Chemical


* 000060696 100 bainite, ferrite 0.000 1.001 Fe

Figure 6.30: XRD pattern for steel coupon immersed in 0.5 M NaCl at -0.8V/SCE
polarisation potential.

C o unts
O ra yith_ S 1 _ 3 p e r_ N a C l_ 1 1 5 0

40000

10000

0
10 20 30 40 50 60 70 80

P o sitio n [2 The ta ] (C o b a lt (C o ))

Visible Ref. Code Score Compound Displacm- Scale Chemical


Name Factor Formula
ent.2Th.
* 000060696 99 bainite, ferrite 0.000 0.996 Fe

Figure 6.31: XRD pattern for specimen exposed to 0.5 NaCl solution under cathodic
protection at -1.15 mV for 14 days.

162
6.4 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution
Mixed With 1.0% (w/w) Na-Bentonite with no Protection
Polarization (OCP)

The rational explanation for the following experiments has already been provided early
in chapter 5. In this experiment, bare mild steel specimens were immersed in a beaker
containing 0.5 M sodium chloride solution, adding 1.0% (w/w) Na-bentonite surved as a
corrosive electrolyte and no cathodic protection was applied to the system (OCP). For
this experiment, the same steps that were applied for the experiment that was illustrated
early in section 6.3.2 were followed.

Figure 6.32 (b), (c), (d), (e) and (f) showing photographs for mild steel specimens, after
immersing in 0.5M NaCl solutions , adding 1.0% (w/w) bentonite for 14 days. Since,
for weight loss measurements the most common exposure time used is in the range of
48 to 168h [113]. Therefore the tests were carried out for periods of 1, 3, 5, 7 and 14
days. For weight loss experiments, two duplicate specimens were immersed in solution
for each period that was mentioned early. They were conducted at stagnant, aerated
conditions and at room temperature and initial pH value was 8. In general, most
specimens were corroded in different levels. A powdery porous uniform corrosion
product was observed on the specimens surfaces. It was covered most specimens
surface, particularly those immersed for more than 5 days. It is obvious that the
corrosion products was more stable on the metal surfaces for specimens exposed to the
solution containing bentonite compared to the one free of bentonite (0.5M NaCl
solution). On the other hand, no corrosion products can be observed to settle at the
bottom of the cells. The solution in the cells for this experiment was more clearly
(Figure 20, Appendix III) compared to the solutions in the previous experiment, Figure
8, Appendix II.

Corrosion process for steel in bentonite clay is expected to be more complicated than
that in aqueous solutions, due to different factors, including the redox potential,
chemistry of the clay and different anions present in the clay. It is will known that, Na-
bentonite clay contains different chemical composition in different ranges, including
SiO2, Al2O3, Fe2O3, K2O, Na2O, MgO and CaO [130]. When bentonite is added to the

163
solution at any concentration, different chemicals species will be dissolved. So,
different cations and anions will be ionised in certain limit in the solution.

More stable, insoluble and protective oxide films were observed on the specimen
surface exposed to the 0.5 M NaCl solution added 1% bentonite. This film may slow
down the corrosion. Hence, corrosion rate might be gradually decreased. Weight loss
experiment will be conducted and weight loss measurement will be presented later in
this section.

(a) Sample in solution (b) After 1 day (c) After 3 days

(d) After 5 days (e) After 7 days (f) After 14 days


Figure 6.32: Corrosion of mild steel specimens after immersed in 3.0% NaCl solution
containing 1.0% bentonite under natural corrosion potential for 1, 3, 5, 7 and 14 days.

The bulk solution pH was slightly decreased from the initial values to settle at about 7.2
at the end of the experiment. As carbon dioxide (CO2) gas can be found naturally in the
atmosphere in several hundreds ppm. This gas can be dissolved naturally to a certain
limit in the aerated corrosive electrolyte to produce carbonic acid. So, pH value
decreasing is most likely due to the carbonic acid formation in the solution. Any small
variation in pH represents a great change in concentration of hydrogen ions [36].
Monitoring the pH was illustrated and can be found in Appendix III, Figure 21.

164
Figure 6.33 showing a plot of corrosion rate in mm/y over time in days for bare mild
steel specimens, after immersing in 0.5 M NaCl solution, containing 1.0% (w/w)
bentonite under natural corrosion potential for 1, 3, 5, 7 and 14 days. The values for
previous Figure were extracted from Appendix II, Table 16. They were derived from
weight loss measurements. The corrosion rate was about 0.115 mm/y after 1 day of
immersion. After that the corrosion rate was found slightly increased at the end of day 3.
As was mentioned earlier, corrosion rate is not constant always with exposure time
[113]. Therefore, the effect of time was investigated by using a massive number of
specimens, exposed for different periods of time; 1, 3, 5, 7 and 14 days in corrosive
electrolyte mention above. It is obvious; the trend of corrosion rate was almost constant
for the period after day 3 up to day 7 to settle at about 0.195 mm/y. However, the rate
was slightly decreased after day 7, which was found to be about 0.184 mm/y at the end
of day 14.

It is important to mention that, the cathodic reaction was mainly oxygen reduction
reaction (O2 + 2 H2O + 4 e- 4 OH). Other possible reduction reaction might be
occurred is reduction of water. Solution pH value is important factor in the corrosion
process [135]. In the first day, the pH value decreased by about 0.8. The drop in pH can
be due to the presence of CO2 in the atmosphere, so carbonic acid can form due to the
dissolved of CO2 gas in the solution [61]. In general, pH was monitored over the time of
the weight loss experiments. The bulk solution pH was decreased over time somewhat
from the initial value. Measurements for weight loss were repeated for several times for
reproducibility. Values are given in Appendix III, 16. Clearly, lower corrosion rate
values were recorded for specimens immersed in 0.5 M NaCl solution, containing 1.0%
bentonite, compared to those were tested in the solutions free of bentonite.

This decrease might be due to the corrosion products formed on the specimens surfaces,
which forms stable, insoluble and protective film. Even if products on the metal surface
were permeable films and partially cover the surface, but it will interfere with the
oxygen mass transfer to the cathode side and delay the diffusion of metal ions from the
anode side. As a result, corrosion rate will be slowed down.

165
0.25

Corrosion rate (mm/y)


0.2

0.15

0.1

0.05
Sample 1 Sample 2 Avg. C.R

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.33: Graph showing corrosion rate plot for bare mild steel specimens after
immersed in 0.5M NaCl solution mixed with 1.0% (w/w) bentonite with no protection
polarisation (OCP). The data extracted from weight loss experiments for 1, 3, 5, 7 and
14 days respectively.

6.4.1 Optical Microscopy of the Investigated Specimens

The optical microscopy was also employed for the specimens to access substrate
morphology. The magnification used was 20X. Figure 6.34 (a) shows a specimens
before immersing in the solution for the experiments carried out for weight loss (silver
grey colour), b, d, f, h, and j for specimens after immersed in 0.5 M NaCl mixed with
1.0% (w/w) bentonite for a periods of 1, 3, 5, 7 and 14 days before the corrosion
products were cleaned and c, e, g, i and k, after the corrosion product was cleaned by
Clarkes solution, applying ASTM standard (G1- 03) [105]. The procedure was
illustrated in chapter 5, section 5.6.

It is obvious that, the progress of corrosion can be observed due to the continuously
change of specimens surface that was exposed to the same solution for different
periods. The corrosion product on the surface for most specimens before cleaning has
almost the same form and colour. The green colour was dominant for specimens
immersed for 1 and 3 days, but dark brown colour was dominant for the other
specimens, immersed for 5, 7 and 14 days. However, the morphology for all specimens
after cleaning was corroded and attacked at most exposed areas. Even though, localised

166
corrosion is rarely expected to take place for carbon steel in neutral and natural water
[112]. But pitting and chickenpox appearance has been also observed in most coupons
subjected to the solution containing bentonite. The more attack and pitting was started
to be increased by the time of immersion and can be exhibited, particularly for specimen
immersed in solution for 7 and 14 days.

100 m

(a) One sample as polished

100 m 100 m

(b) After 1day before cleaning (c) After 1day after cleaning

167
100 m 100 m

(d) After 3 days before cleaning (e) After 3 days after cleaning

100 m 100 m

(f) 5 days before cleaning (g) 5 days after cleaning

100 m 100 m

(h) After 7 days before cleaning (i) After 7 days after cleaning

168
100 m 100 m

(j) After 14 days before cleaning (k) After 14 days after cleaning
Figure 6.34: Photographs showing optical micrographs for corrosion of steel specimens
after immersed in 0.5 M NaCl solution, adding 1.0% (w/w) bentonite at OCP. (a) one
example for sample as polished, b, d, f, h and j for samples before cleaning from
corrosion products and c, e, g, i and k after cleaning for samples after 1, 3, 5, 7 and 14
days respectively.

Figure 6.35 shows the corrosion potential variation over time, for polished bare steel
specimens after immersed in 0.5 NaCl solutions mixed with 1.0% (w/w) bentonite for 1,
3, 5, 7 and 14 days. The values for corrosion potential were shown in Appendix III,
Table 17. The OCP was measured after 5 minutes, -0.306V vs. SCE. The silver grey
colour for mild steel surface was changed and corrosion product was observed at the top
of steel specimens surface. After 1 day of immersion, the surface of the steel specimen
was partially covered by light and yellowish to orange film, surrounded by dark brown
colour at the edges. The potential was gradually shifted to more negative to settle at
about -0.737V. The same corrosion potential was recorded for the sample immersed for
3 days, but further porous corrosion products was observed on the metal surface with
slight change in colour for more darker than that for specimen in Figure 6.32 (b). The
cathodic reaction was mainly oxygen reduction. Another possible reduction reaction
that might occur is water reduction.

169
-250
-300
Sample 1 Sample 2 Sample 3 average

P o ten tia l V s. S C E (m V )
-350
-400
-450
-500
-550
-600
-650
-700
-750
-800
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.35: Natural corrosion potential for mild steel specimens after immersed in
aerated 0.5 M NaCl solution contains 1% bentonite for 14 days.

The progress of corrosion products were observed on the metal surface for sample
immersed for 5 days and the OCP shifted to more positive to settle at about -0.730V.
Quite insoluble dark brown corrosion products were developed on the metal surface.
Only small amount of corrosion products were observed at the bottom of the beakers for
all periods and the water was slightly changed in colour to become slightly yellowish
compared to the corroded samples that were immersed in 0.5 M NaCl solution free of
bentonite. After day 7, the corrosion potential slightly shifted to more negative again to
settle at -0.723V. Then finally, after day 14, Ecorr potential declined from 0.722V to
reach a steady state value at about -0.714V.

Figure 6.36 shows the polarisation resistance (Rp) variation with time. The experiment
was conducted for several times and the Rp values presented in this graph were
extracted from the Rp plots generated by LPR runs over time shown in appendix III,
table 18. The initial polarisation resistance curve was recorded after 1h (day 0) of the
test which was about 1801 Ohm.cm2. The Rp value was slightly fall down by 243
Ohm.cm2 at the end of day 1 to settle at 1558 Ohm.cm2. After that, Rp was gradually
increased after day 1, to settle at 2472 Ohm.cm2 after day 7. Finally, it was slightly
decreased to 2150 Ohm.cm2 at the end of the experiment. Examples for LPR
measurement runs after 1h, 1 day and 14 days period were presented in Appendix III,

170
Figure 22 24, respectively and the data extracted from the plots was also shown on
Table 18.

3000

2500
R p (O h m .c m2 )
2000

1500

1000

500 Sample 1 Sample 2 Sample 3 average

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.36: The Rp values for mild steel specimens immersed in aerated 0.5 M NaCl
solution, contains 1% bentonite at natural corrosion potential.

Table 6.5 gives a summary for (a) corrosion current values (icorr.) in A.cm-2, corrosion
rate values (CR) in mm/y produced from LPR measurement and corrosion rate, CR exp.,
corrosion current (icorr. exp.) and Stern-Geary constant (B) obtained from weight loss
measurements.

Tables 6.3 and 6.5 show that the corrosion rates obtained from LPR measurement was
consistently higher for the 0.5NaCl solution, containing 1% bentonite. The results from
the weight loss experiments are not so clear cut and it is observed that the corrosion
rates are generally less for the 1% bentonite solution, though the individual values at 5
and 7 days are greater. It is reported by many authors [18, 113] that the corrosion rate
varies over time and this was certainly observed in this experiment. For example, for the
1% bentonite solution the values of CRexp ranged from 0.115 mm/yr from day 1 to 0.197
mm/yr for day 5. The average value of constant B was also evaluated, and found to be
0.023V/decade, very similar to that used to calculate the corrosion rate using Rp curves,
(0.026V/decade).

171
The slight change in the corrosion rate and corrosion current for both measurements can
be attributed to dissimilar surface condition at the time of measurement. As was
mentioned early in the previous experiment, corrosion rate obtained from weight loss
measurement is a function of exposing time. Conversely, corrosion rate obtained from
LPR measurement is produced instantly in a short time period. Corrosion rates as
measured by weight loss are greater than that obtained by LPR. This is hardly surprising
since LPR is an instantaneous measurement whereas weight loss is cumulative.

Continuous change in the surface conditions as a result of continuous change in


corrosion process is may be due to variation in corrosion electrolyte. This change may
be occurred due to the natural dissolution of CO2 gas present in the atmosphere to
produce carbonic acid. As a result pH decreased from the initial values.

Table 6.5: Summary table of parameters determined for mild steel specimens after
immersing in aerated 0.5 M NaCl solution containing 1.0% bentonite over 14 days
period at OCP.

Time C.R icorr CR exp. icorr. exp. Bexp.


-1 2 -1 2
day mm.y (A/cm ) mm.y (A/cm ) (V/decade)
0* 0.203 17.5E-06 0.078 6.73E-06 0.010
1 0.190 16.4E-06 0.115 9.95E-06 0.015
3 0.164 14.2E-06 0.190 16.4E-06 0.025
5 0.122 10.6E-06 0.197 17.0E-06 0.025
7 0.141 12.2E-06 0.195 16.9E-06 0.039
14 0.156 13.5E-06 0.184 16.0E-06 0.024
Average value using trapezoidal 0.174 13.2E-06 0.179 15.51E-06 0.028

6.5 Bare mild Steel after immersed in 0.5 M sodium


chloride solution, adding 1.0% (w/w) Na- bentonite
under CP, at 1.15V Vs SCE
It is believed that, cathodic protection can be used to protect immersed metallic
structures from corrosion. This technique promotes precipitation of calcareous deposits
on the metallic surface, creating a physical barrier against oxygen diffusion towards
immersed metal surface to the corrosive solution [27, 32]. The result is a decrease in the

172
final current density required to protect the structure or extension of the life of the
sacrificial anode to maintain sufficient protective cathodic potentials for the structure
[32, 136].

Figure 6.37 shows variation in CD in mA.m-2 over time in days for polished bare steel
specimens immersed in 0.5 NaCl solutions containing 1.0% (w/w) Na-bentonite at -
1.15V cathodic polarisation potential. Current was measured and recorded regularly.
The values of current density are shown in Appendix III, Table 20. The first current
measured after the sample was immersed will not be presented in the graph, as it is
enormous value compared to the latter one. The current was measured after 1 h,
expressed as day 0 in the figure which was 2927mA.m-2. Then, current was monitored
and current density was calculated; it was dropped swiftly at about 50% to settle at
1494.7mA.m-2. The continuous growing insulating layer on the specimens surface over
time has worked as physical barrier against oxygen diffusion. The system shows a faster
decline in current density over 2 day period. This event is due to the continuously
growing of the film on the protected specimens surface. A steady fall for current was
observed over time of the test. More than 68% was the drop in current after 3 days
period to stabilise at about 637mA.m-2. The final measure was recorded after 14 days
period, 224.8mA.m-2.

According to NACE standard PR0176-2003 [1], the final current density required for
long term cathodic protection systems used in offshore petroleum production can be
reduced, if a calcareous film is precipitated on the metal surface. In this case, the current
was gradually decreased by time, but the final current was quite high.

It must be pointed out that, there is no indication for current to tend in the direction of
zero [33, 136]. This event is probably due to the porous nature of Mg(HO)2 film in the
form of brucite which was gradually cover the most specimens surfaces [27]. Further
more, if the amount of species (cations) present in the corrosive solution is limited in
the environment, the progress of the film on the metal surface will be imperfect. So, it is
difficult to repair any porous or defects that may be caused by evolution of H2 bubbles
from the cathode surface, when applying quit high cathodic potential [27].

173
3000
Sample 1 Sample 2 Sample 3 Average
2750
2500
2250
2000

CD (mA.m )
-2
1750
1500
1250
1000
750
500
250
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.37: Current density in mA.cm-2 for mild steel specimens immersed in aerated
0.5 M NaCl Solution mixed with 1.0% Na-bentonite for test 1, under CP at -1.15V/SCE.

At the end of each test, the samples were taken out from the solution dried and
photographed, then stored in a desiccator for SEM/EDAX and XRD analysis.

Figure 6.38 shows the progress of the film composed on the specimens with time due to
applying cathodic polarisation potential at -1.15V. The specimens exposed to the 0.5 M
NaCl solution mixed with 1.0% (w/w) bentonite for 14 days period were observed.
They were completely covered by Shiny silver grey film. The film was generated
rapidly by applying a relatively high initial current to the system under CP.

More rapid polarization for structure under CP, using potentials in the range of 800
mV to -1200 mV/SCE to protect steel structures in sea water [28]. However, evolution
of H2 bubbles from the metal surface was also observed, when high cathodic
polarisation potential was applied. H2 bubbles continuous remove part of the scale or
cause cracks for the layer formed on the metal surface and delay the aragonite layer due
to it is small size [27]. In our case, shiny fine white products were observed on the top
of bentonite layer of the electrochemical cell under CP. They fall down from the film
has been formed on the metal surface. As a result, unstable porous layer can be formed
and partially blocking the specimens surface.

As was mentioned earlier, the structure was under CP at -1.15V and the current was
gradually decreased by time due to continuous formation of different products on the
specimens substrate. Even though, quit high current density value was recorded after

174
14 days. This could be due to the porous film formed on the metal surface. Magnesium
hydroxide in the form of aragonite is characterized by its small size which may block
the porous layer but it was not detected by XRD in this case. More results about the
crystal compounds detected by XRD will be presented later in section 6.5.1.

(a) After 1 day (b) After 3 days (c) After 5 days

(d) After 7 days (e) After 14 days


Figure 6.38: Film composed on the mild steel specimens surface immersed in 0.5 M
NaCl solution containing 1.0% bentonite, the system was under CP at -1.15 V/SCE) for
different time.

The alkalinity at the cathodic regions in the electrochemical cells possibly will cause
precipitation of insoluble hydroxide species at the cathodic sides [137]. The process is
taken place due to a certain cations that may present in the electrochemical solution;
Magnesium hydroxide (Mg(OH)2) is one of the most common compounds that may
formed on the specimens under CP if Mg2+ was soluble in the solution (section 2.5.2 ).
For more investigation, the SEM/ EDAX and XRD analyses was applied.

Figure 6.39 shows the solution pH value for first day of the experiments and at the end
of the experiment. The initial bulk solution pH was about 8 which were close to the
initial bulk pH of seawater, 8.2. When the CP was applied at -1.15V, the pH was

175
gradually increased somewhat from the initial value. In 0.5 M NaCl solution mixed with
0.1 (w/w) bentonite , the pH increased by, at almost more than 2 to settle at 10.1 as
shown in figure 6.37(f). The increase in pH is attributed to the formation of OH- ions
due to the cathodic reactions at the working electrode. The cathodic reaction was
illustrated below in section 6.5.1.

(a) After 1 day (b) After 3 days (c) After 5 days

(d) After 7 days (e) After 14 days (f) pH after 14 days


Figure 6.39: Photographs for pH for mild steel specimens immersed in 3.0% NaCl
solution mixed with 1.0% bentonite under CP at -1.15 V/SCE for different time.

6.5.1 XRD and SEM Analyses for Specimens Immersed in 0.5 M NaCl
Solution Adding to 1.0% (w/w) Bentonite

In this experiment, XRD analysis was carried out for specimens exposed to 0.5 M NaCl
solution adding 1.0% (w/w) bentonite at free corrosion potential, and when exposed to
cathodic polarisation potential, in order to recognise and illustrate the various crystalline
compounds formed on the metal surface. The information obtained from the XRD was
used to be compared with data and results that was achieve form electrochemical
measurements in this study.

176
Figure 6.40 shows XRD pattern with its chemical analysis list for mild steel species
exposed to 0.5 M NaCl solution mixed with 1.0% bentonite at natural corrosion
potential after 14 days. It is possible for deposits to form on the metal surface while it is
exposed to the electrolyte solution. As can be seen, various corrosion products are listed
in the pattern list below the figure. The crystalline material layer on the metal surface is
mostly lepidocrocite with goethite; both are in the form of FeOOH. Halite in the form of
NaCl was also present. The layer was nominally 10 to 20 microns thick, as the substrate
in the form of ferrite (Fe), was not visible. The mechanism and electrochemical
reactions for all chemical products mentioned above were illustrated in detail in section
6.3.4, page 156. In this experiment, the layer formed on the specimens surface is
thicker than that formed when the specimens were exposed to the 0.5M NaCl solution
which contained no bentonite. However, the substrate in the form of ferrite is not
detected in the XRD patterns for the electrolyte that contains bentonite compared to the
solution that contains no bentonite. As a result, the corrosion rates are greater than those
recorded for the specimens immersed in the solution free of bentonite. If the corrosion
rate using LPR is considered more important then this result may be attributed to the
corrosion products layer though it was thinner was a better barrier to the diffusion of
oxygen to the metal surface. However, if the measured corrosion rate using weight loss
is the important result then the thicker layer gives better protection.

Ref. Code Score Compound Displacement Scale Chemical


000050628 53 Name
Halite [2Th.]
0.000 Factor
0.206 Formula
NaCl
000080098 57 Lepidocrocite 0.000 0.727 - FeO (OH)
000290713 38 Goethite 0.000 0.316 -FeO(OH)

Figure 6.40: XRD pattern for steel specimens immersed in 0.5 M NaCl solution
contains 1.0% (w/w) bentonite at natural corrosion potential.

177
Figure 6.41 to Figure 6.44 shows XRD pattern with its chemical analysis list for mild
steel species exposed to 0.5 M NaCl solution mixed with 1.0% (w/w) bentonite. The
specimens were under cathodic protection (-1.15V) for 3, 5, 7 and 14 days, respectively.
As can be seen, various products were listed in the pattern list below. Identical
reflections at different strength, at different degrees for all figures below can be
observed. A strong reflection related to Mg(OH)2 in the form of brucite was detected at
approximately 18.5 and 38 degrees 2theta, in the main graphics, although there are 6
more reflections at different strength at different degrees, represent the brucite. In sea
water, Mg(OH)2 is recognized to be under-saturated condition and tends to precipitate at
quite high pH values, 9.5 [31]. So, the precipitation of any protective layer on the
metal surface will depend on the local pH, at the metal/electrolyte interface of the
protected metal and the concentration of different cations in the solution containing
bentonite.

The brucite layer was gradually increased by the time of the test, see Figure 6.44. The
continuous formation of brucite can be ascribed to the local increase in pH at metal/
electrolyte interface caused by the production of hydroxyl ions (OH ), due to applying
CP in the presence of magnesium. Magnesium is a common impurity material can be
found in bentonite. The initial and most important cathodic reaction is oxygen reduction
under activation control, if the availability of oxygen to the metal surface is high [28-30]
(Equations 2.27, 2.28).

O2 + 2H2 O + 2e H2O2 + 2OH . (2.27)

H2O2+ 2e 2OH (2.28)

When the net current is limited by mass transfer to the surface and the O2 at the
interface is limited; in this case only hydrogen evolution becomes the dominant reaction
due to reduction of water as shown in Equation 2.29:

2H2 O + 2e H2 +2OH ... (2.29)

The result is a significant production of hydroxyl ions (OH), causes an increase in local
pH value. The bulk solution pH values were settled at about 10.1. So, A local alkalinity
at the specimens surface due to further concentration of hydroxyl ions near to the
surface solution, may be caused the precipitation of magnesium hydroxide Mg(OH)2 on
the specimens surfaces according to the following reaction:

178
Mg2+ + OH Mg (OH)2 .. (2.36)

A quite considerable reflection was also detected for calcium carbonate (CaCO3) in the
form of calcite at approximately 29.5 degree, 2theta. But after day three, brucite was
gradually decreased and only traces of it were detected in a very weak reflection for
samples tested for 5 and 7 days. However, no calcite has been detected on the sample
was tested for 14 days. This is may be due to the conversion of the original calcium
based bentonite to its soluble sodium form would not have proceeded to completion.

The electrochemical reactions that may be suggested to occur for CaCO3 to form were
illustrated in detail in section 2.5.2. In brief, high local pH at the metal/solution
interface due to the large amount of OH ions produced by the reactions illustrated
above in Equation 2.27, 2.28, 2.29, encouraging the reaction between calcium and
carbonate ions present in the solution, according to the following reaction [31]:

CO32 +Ca2+Ca CO3 (2.35)

Only small amount of calcium carbonate in the form of calcite was precipitated on the
specimens surface. The limited amount of calcite that was detected might be due to the
low concentration of carbon dioxide (CO2) in the solution, containing bentonite, up to
15 ppm. Hence, the reaction of CO2 with the water will be slow and any subsequent
reactions that generate direct precipitation of calcium carbonate (CaCO3) onto the metal
surface as a protective layer against corrosion will be correspondingly slow [26].

A considerable amount of halite in the form of NaCl was also revealed by XRD. The
presence of NaCl crystalline on the metal surfaces might be ascribed to the specimens
were not rinsed by deionised water at the end of the experiment. So, NaCl may be
precipitated on the metal surface while specimen dried by dry air.

The substrate in the form of ferrite (Fe) was detected by XRD, its reflection at 45, 65
and 83 degrees 2theta was gradually decreased by time of the test, particularly for the
system under CP for 5 days ( Figure 6.43, Figure 6.44). This continuous decrease of
substrate reflection can be suggested to the continuous growth of the film for the system
under CP [27, 28].

179
22500

10000

2500

0
10 20 30 40 50 60 70 80

P o s i ti o n [2 T h e ta ] (C o p p e r ( C u ))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
010822453 78 Brucite 0.000 0.519 Mg (OH)2
000050628 70 Halite 0.000 0.040 (H)2
Na Cl
000060696 55 Ferrite 0.000 0.399 Fe

Figure 6.41: XRD pattern for steel specimens immersed in 0.5 M NaCl solution
contains 1.0% (w/w) bentonite under cathodic protection at -1.15 mV for 3 days.

C o u n ts

6400

3600

1600

400

0
10 20 30 40 50 60 70 80

P o s i ti o n [2 T h e ta ] (C o p p e r (C u ) )

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula

180
010822453 74 Brucite 0.000 0.949 Mg(OH)2
000050628 68 Halite 0.000 0.673 NaCl
000060696 51 ferrite 0.000 0.886 Fe
040078659 84 Calcite 0.000 0.312 CaCO3

Figure 6.42: XRD pattern for steel specimens immersed in 0.5 M NaCl solution
contains 1.0% (w/w) bentonite under cathodic protection at -1.15 mV for 5 days.

O ra yith_10 2

64 00

36 00

16 00

4 00

0
10 20 30 40 50 60 70 80

P osition [2 Theta ] (C opper (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
010822453 73 Brucite 0.000 0.780 Mg(OH)2
000050628 74 Halite 0.000 0.511 Na Cl
000060696 52 ferrite 0.000 0.228 Fe
040078659 63 Calcite 0.000 0.085 CaCO3

Figure 6.43: XRD pattern for steel specimens immersed in 0.5 M NaCl solution
contains 1.0% (w/w) bentonite under cathodic protection at -1.15 mV for 7 days.

181
6400

3600

1600

400

0
10 20 30 40 50 60 70 80

P o s i ti o n [2 T h e ta ] (C o p p e r ( C u ))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
010822453 78 Brucite 0.000 0.888 Mg(OH)2
000050628 74 Halite 0.000 0.307 Na Cl
000060696 54 Ferrite 0.000 0.825 Fe
040078659 49 Calcite 0.000 0.105 CaCO3

Figure 6.44: XRD pattern for steel specimens immersed in 0.5 M NaCl solution
contains 1.0% (w/w) bentonite under cathodic protection at -1.15 mV for 14 days.

Figure 6.45 represents the reproducibility for reflections of compounds that were
accumulated on the metal surface under CP. As can be seen, identical reflections at
different strength, at different degrees in the main graphics for each compound can be
distinguished. The substrate in the form of ferrite was gradually decreased by time of
the test, particularly after day 5. This behaviour may be attributed to the continuous
formation of different deposit on the specimens surface, including Mg(OH)2 in the
form of brucite and calcium carbonate in the form of calcite.

In our case, calcium carbonate, CaCO3 in the form of aragonite was not detected by
XRD. The absence of aragonite is difficult to explain but one tempted to suggest is
possible that the magnesium compound are screened by the other deposit which is
mainly brucite and calcite.

We can conclude that, the XRD results indicate similar reactions have been occurred,
leading to create almost same composition on the metal surface. It is confirmed that, the
time was an essential factor for specification and quality of any film growing on the

182
metal surface. It is possible for any deposit to form on the metal surfaces while exposed
to the electrolyte solution. However, in order to create an active and uniform protective
film, cathodic protection at precise potential must be applied. More investigation must
be done by applying different polarisation potential.

Figure 6.45: XRD pattern for specimens after immersing in 0.5 M NaCl solution
contains 1.0% bentonite under cathodic protection at -1.15 mV for 3, 5, 7 and 14. days.

The following is the data generated due to employing the SEM/EDAX techniques. The
EDAX was carried out for specimens at free corrosion potential, and when exposed to
cathodic polarisation potential. The technique was employed to reveal the chemistry of
different compounds that might be formed on the specimens surfaces. Then a
quantitative analysis for those elements, and their progress with time on the metal
surface was also investigated, by using EDAX technique.

Figure 6.46 and Figure 6.47 show SEM images and EDX patterns with its chemical
analysis list for corrosion products formed on mild steel surface exposed to 0.5 M NaCl
containing 1.0% (w/w) bentonite at natural corrosion potential for 14 days. As can be
seen, various elements were illustrated in the pattern list. The SEM technique used does
not give the corrosion products as compounds; it gives the major element present as Fe
as shown in Figures 6.46 and 6.47. This was also detected by XRD, see Figure 6.40; the
product on the metal surface is mostly corrosion products of lepidocrocite with goethite;
both are in the form of FeOOH. The patterns were also shown a reflection that may be
related to O2+ and Na+. A weak reflection that may be related to Cl- was also detected.

183
Element Fe Cl Na O
Wt % 74.60 3.82 2.61 18.97
Figure 6.46: SEM and EDAX element analysis for steel metal exposed to 0.5 M NaCl
solution contains 1.0% (w/w) bentonite at natural corrosion potential. After 14 day.

184
Element Fe Na Cl O
Wt % 74.53 4.22 4.08 21.18
Figure 6.47: SEM and EDAX element analysis for specimens after exposed to 0.5 M
NaCl solution contains 1.0% (w/w) bentonite at natural corrosion potential for 14 days.

Figure 6.48 to Figure 6.52 show the SEM images and EDAX element analysis for
deposits on steel metal exposed to 0.5 M NaCl contains 1.0% (w/w) bentonite under
cathodic protection (-1.15V) for 1, 3, 5, 7 and 14 days day, respectively. The deposits
obtained on the metal surface at -1.15V/SCE are made of crystals assembled in different
morphology. EDX analysis indicates that the precipitate contains magnesium, calcium,
Fe substrate, Carbon, Chloride, sodium and oxygen in different concentrations. It was
also confirmed by XRD analysis; the products detected on the metal surfaces are
magnesium hydroxide in the form of brucite calcium carbonate in the form of calcite
and ferrite as metal substrate and sodium chloride in the form of halite. Obviously, the
main deposit for all figures was magnesium hydroxide in the form of brucite (see XRD
results). Traces of calcite and halite were detected by EDX. The substrate in the form of
ferrite (Fe) was gradually decreased by time of the test due to growing up of the brucite
film.

185
Element Mg Ca Fe Na Cl C O
Wt % 27.84 0.66 46.01 0.23 0.66 0.59 24.01
Figure 6.48: SEM and EDAX element analysis for specimens after exposed to 0.5 M
NaCl solution contains 1.0% (w/w) bentonite under CP at -1.15V for 1 day.

186
Element Mg Ca Fe Na Cl C O
Wt % 31.50 0.38 18.64 2.83 2.04 1.55 43.06
Figure 6.49: SEM and EDAX element analysis for specimens after exposed to 0.5 M
NaCl solution contains 1.0% (w/w) bentonite under CP at -1.15V for 3 day.

187
Element Mg Ca Fe C Cl Na O
Wt % 30.21 0.68 11.15 1.67 4.21 3.98 48.10
Figure 6.50: SEM and EDAX element analysis for specimens after exposed to 0.5 M
NaCl solution contains 1.0% (w/w) bentonite under CP at -1.15V for 7 days.

188
Element Mg Ca Fe C Cl Na O
Wt % 31.85 0.81 7.90 1.25 3.78 4.07 50.32
Figure 6.51: SEM and EDAX element analysis for specimens after exposed to 0.5 M
NaCl solution contains 1.0% (w/w) bentonite under CP at -1.15V for 7 days.

189
Element Mg Ca Fe Cl Na C O
Wt % 31.32 0.68 3.01 0.35 0.65 1.42 62.57
Figure 6.52: SEM and EDAX element analysis for specimens after exposed to 0.5 M
NaCl solution contains 1.0% (w/w) bentonite under CP at -1.15V for 14 days.

6.6 Bare Mild Steel Immersed in 0.5 M Sodium Chloride Solution,


Adding 10.0% (w/w) Bentonite at Open Circuit Potential (OCP)

Figure 6.53(b), (c), (d), (e) and (f) show photographs for mild steel specimens, after
they were immersed for 1, 3, 5, 7 and 14 days in 0.5M NaCl solutions mixed with
10.0% (w/w) bentonite, under stagnant, aerated conditions and at room temperature and
initial pH value was 8. For weight loss experiments, two duplicate specimens were

190
immersed in solution for each period mentioned early. In general, the most specimens
were corroded in different levels. After 24 hours of immersion, partially corrosion at the
edges of specimens was observed and there was no clear corrosion products in the
corrosive electrolyte can be observed. A stable orange film of corrosion products was
observed on the metal surface for specimens after they were tested for 3, 5 and 7 days of
immersion. Even though, corrosion products covered most of the specimens surface,
the corrosion solution in the cell was quite clear and there is no corrosion products can
be observed on the top of bentonite layer. This may be due to the solid and stable
insoluble corrosion products that formed on the metal surface. More dark brown stable
corrosion products were been exhibited on the specimens immersed for 14 days.

Even though, the bentonite content in the 0.5 M NaCl solution for this experiment is
about 10 times greater than the previous experiment was illustrated in section 6.4, but
the results were quite similar. The corrosion products may be continuous to develop to a
certain limit. The film may slow down the corrosion. Hence, corrosion rate might be
gradually decreased. Weight loss experiments were conducted and weight loss
measurement will be presented later in this section. As was mentioned early, Na-
bentonite clay contains different chemical compounds in different ranges, including
SiO2, Al2O3, Fe2O3, K2O, Na2O, MgO and CaO [130]. For this experiment, the Na-
bentonite was 10% in the solution compared to the bentonite added to the previous
experiment, 1%. So, the concentration of different chemicals species dissolved in the
solution will be about 10 times greater than that for the solution containing 1%
bentonite. So, more different cations and anions will be ionised in certain limit in the
solution. The result was more stable, insoluble and protective oxide films were observed
on the specimen surface exposed to the NaCl solution with 10% bentonite added. This
film may slow down the corrosion. Hence, corrosion rate might be gradually decreased.
Weight loss experiment will be conducted and weight loss measurement will be
presented later in this section.

191
Specimens in 0.5M NaCl (1day) (3 days)

(5 days) (7 days) (14 days)


Figure 6.53: graph showing Corrosion of mild steel specimens after immersed in
10.0% bentonite mixed with 3.0% NaCl solution for different time.

Figure 6.54 shows the plot of corrosion rate for bare mild steel specimens against time
immersed in 0.5 M NaCl solution contains 10.0% (w/w) bentonite. The values for this
graph were illustrated in appendix IV, Table 21. After 24 hours of immersion, the
corrosion rate was about 0.115mm/y for the samples exposed to the solution. This value
was derived from the weight loss after 1 day of immersion. After that the corrosion rate
was found slightly increased at the end of day 3, to settle at about 0.181mm/y. However,
it is obvious that the trend of corrosion rate was almost constant for 5, 7 and 14 days to
settle at about 0.187mm/y.

Similar trend was also observed for specimens immersed in 0.5 M NaCl solution
containing 1.0% (w/w) bentonite. It is may be due to the presence of bentonite in the
solution for both experiments (Figure 6.33).

In this case, the corrosion rate was almost stable after 3 days of immersion. A negligible
variation corrosion rate values was recorded for the specimens exposed to the solution
for 5, 7 and 14 days. This might be due to stable, insoluble and protective films that
were formed. Even if it is permeable or partially covers the surface, it will interfere with
the oxygen mass transfer to the cathode side and delay the diffusion of metal ions from

192
the anode side. As a result, corrosion rate will be slowed down. In bentonite, the main
cathodic reaction is oxygen reduction reaction and the other possible reduction reaction
that might be occurring is reduction of water. Solution pH value is important factor in
the corrosion process [135]. A drop in pH value over time for the corrosive electrolyte
is expected to occur due to presence of CO2 in the atmosphere, so carbonic acid can
form due to dissolved of CO2 gas in solutions [61] . Acidic solution might accelerate the
corrosion process.

0.25
C o r ro sio n ra te (m m /y )

0.2

0.15

0.1

0.05
Sample 1 Sample 2 Avg. C.R

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.54: Graph showing plot of corrosion rate over time for mild steel specimens.
The data obtained from weight loss measurement for specimens after immersing in 0.5
M NaCl solution contains 10.0% (w/w) bentonite for different time.

6.6.1 Optical Microscopy of the Investigated Specimens

The optical microscopy was also employed for the specimens for more substrate
morphology assessment, using the same magnification that was used for the previous
experiments (20X).

Figure 6.55 (a) shows a specimen as an example before immersing in the solution for
the experiments carried out for weight loss. The surface was silver grey in colour. An b,
d, f, h, and j for samples immersed in the 0.5 M NaCl mixed with 10.0% (w/w)
bentonite for periods of 1, 3, 5, 7 and 14 days before the corrosion products were
cleaned and c, e, g, i and k, after the corrosion product was cleaned by Clarkes solution.

193
The progress of corrosion can be observed due to continuously change of specimens
surfaces that was exposed to the identical solution mentioned above for different period.
The corrosion product on the most specimens surface before cleaning has almost the
same form and colour. The green rust was dominant for specimens immersed for 1 and
3 days, but dark green with brown rust was mostly covered the other specimens
immersed for 5, 7 and 14 days and the green rust still can be detected. The corrosion
products according to that colour suggested to be ferrous hydroxide (FeO(OH)2) and /or
mixture magnetite (Fe3O4). XRD and EDX analysis will be conducted to reveal the
different crystal structure formed on the specimens. The morphology for all specimens
after cleaning was corroded and attacked at most exposed areas. General corrosion can
be observed for most specimens immersed in the solution. Pitting appearance was also
detected in different levels in some areas. Nevertheless, more attack and pitting were
started to develop by time of immersion and can be exhibited, especially for samples
exposed to solution for more than 1 day. It is clear that, the attack for specimens
exposed to the solution containing 10% bentonite was not as much as that was observed
for the others where immersed in the solution containing 1% bentonite and the one was
free of bentonite, particularly after 5 days of immersion. This can be ascribed to the
continuous growth of corrosion products on the metal surfaces. This film may be more
thicker and less porous compared to the others where detected on the specimens
exposed to the 0.5 NaCl solution free of bentonite and the one contains 1% bentonite. In
some cases, the different cations and anions dissolved in the corrosive solution create a
protective oxide film.

(a) As polished

194
100 m 100 m

(b) 1day before cleaning (c) 1day after cleaning

100 m 100 m

(d) 3 day before cleaning (e) 3 day after cleaning

100 m 100 m

(f) 5 day before cleaning (g) 5 day after cleaning

195
100 m 100 m

(h) 7 day before cleaning (i) 7 day after cleaning

100 m 100 m

(j) 14 day before cleaning (k) 14 day after cleaning


Figure 6.55: Optical micrographs for corrosion of steel specimens immersed in 0.5 M
NaCl solution mixed with 10% bentonite. (a) sample as polished, b, d, f, h and j samples
before cleaning from corrosion products and c, e, g, i and k after cleaning for samples
after 1, 3, 5, 7 and 14 days respectively.

Figure 6.56 shows the corrosion potential variation over time, for polished bare steel
specimens immersed in 0.5 NaCl solutions mixed with 10.0% (w/w) bentonite for 1, 3,
5, 7 and 14 days. The values for corrosion potential are shown in Appendix IV, Table
22. The OCP was measured after 5 minutes, -0.314V vs. saturated calomel electrode
(SCE). After 24 hours, the potential was gradually shifted to more negative to settle at
about -0.745V and the silver grey colour for mild steel surface was changed and light
corrosion product was observed at the edges near to the top of steel samples. Almost,
same corrosion potential values were recorded with a slightly shift to more positive for
samples immersed for 3, 5, 7 and 14 days; -0.743V, -0.740V, -0.737V and -0.732V,
consequently. The development of insoluble corrosion products on the metal surface

196
was also observed. It is thicker and darker brown in colour for samples immersed for 7
and 14 days compared to other time periods.

-250
-300
Sample 1 Sample 2 Sample 3 average
-350
P otential V s. SC E mV

-400
-450
-500
-550
-600
-650
-700
-750
-800
-850
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.56: Graph showing corrosion potential at OCP for mild steel specimens after
immersing in aerated 0.5 M NaCl solution, containing10% bentonite for 14 days.

Figure 6.57 shows the plot of polarisation resistance (Rp) variation with time for mild
steel specimens after immersing in aerated 0.5 M NaCl solution at OCP. The
experiment was conducted for several times and the Rp values presented in this graph
were extracted from the Rp plots generated by LPR runs over time that shown in
Appendix IV, Table 23. The initial polarisation resistance curve was recorded after 1h
(day 0) of the test which was about 4362 Ohm.cm2. The polarisation resistance was
dramatically fall down by 242.7 Ohm.cm2 at the end of day 1 to settle at 3396 Ohm.cm2.
After that, the Rp was continuously turned down for samples immersed for 3, 5 and 7
days, to settle at 3207, 2657 and 2247 Ohm.cm2, respectively. Finally, the Rp for the
test for 14 days was increased to settle at 2322 Ohm.cm2.

197
5000
4500
4000
3500

R p (O h m .c m2 )
3000
2500
2000
1500
1000
500 Sample 1 Sample 2 Sample 3 average
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.57: Graph showing plots of Rp values for mild steel specimens after immersed
in aerated 0.5 M NaCl solution, adding 10% bentonite for 14 days at OCP.

A summary for results, comparing the corrosion rate that was calculated obtained from
LPR measurement (Table 24, Appendix 24) and the one were experimentally
determined by using weight loss measurement (Table 21, Appendix 24) . Comparison
between the actual Stern-Geary constants B extracted from the corrosion rate that was
determined experimentally with the one that has been assumed (0.026 V/decade) was
also reviewed.

Corrosion current (icorr.) in A.cm-2 and corrosion rate (CR) in mm.y-1 were calculated,
using the same procedure illustrated earlier for the previous experiments. In brief,
evaluation of constant (B), assuming 0.026 V/decade and polarisation resistance (Rp)
leads to calculate the corrosion current density by using equation 5.3, section 5.8 [18].
(Note: we assumed bc = , since cathodic reaction is probably controlled by diffusion of
oxygen reduction, a value of 0.026V/decade is commonly used [40, 120].

The corrosion rate in mm/y was also determined, using the experimental weight loss
measurement, applying equation 5.1 in page 98. Then, icorr. in A.cm-2 was also
determined, using Faradays law (equation 5.7). The results were used to determine the
actual constants B at different period of immersion, using equation 5.3.

Table 6.6 illustrate a summary for values of corrosion rate and corrosion current
evaluated, using LPR measurement and those determined experimentally, using weight

198
loss measurements. The values for CR and corrosion current for specimens were
immersed in the solution for 1 day period for both cases almost have the same values.
However, a negligible change in corrosion rate and current density were recorded for
the one where determined by LPR measurement. On the other hand, excluding
specimens immersed in the solution for 1 day period, identical values were obtained for
corrosion rate and corrosion current determined, using weight loss measurement. The
average value of constant B was also evaluated, and found to be 0.041V/decade, greater
than that used to calculate the corrosion rate using Rp curves, by a factor that ranged
between 1.31 and 1.92.

A negligible change in corrosion rate over time was observed for specimens exposed to
the solution containing bentonite. However, low corrosion rate was recorded for the
specimens immersed in the 0.5NaCl solution, containing bentonite compared to that
obtained for solution free of bentonite. The change in the corrosion rate and corrosion
current for both measurements can be attributed to dissimilar surface condition at the
time of measurement. Corrosion rates as measured by weight loss are greater than that
obtained by LPR. This is hardly surprising since LPR is an instantaneous measurement
whereas weight loss is cumulative.

Table 6.6: Summary of parameters determined for mild steel specimens, after
immersing in aerated 0.5 M NaCl solution containing 10.0% bentonite for 14 days at
OCP.

Time C.R icorr CR exp. icorr. exp. Bexp.


-1 2 -1 2
day mm.y (A/cm ) mm.y (A/cm ) (V/decade)
0* 0.091 7.85E-06 0.082 7.11E-06 0.011
1 0.093 8.02 E-06 0.115 9.94E-06 0.015
3 0.097 8.36E-06 0.181 15.6E-06 0.024
5 0.112 9.65E-06 0.191 16.5E-06 0.025
7 0.137 11.8E-06 0.190 16.4E-06 0.025
14 0.130 11.2E-06 0.187 16.2E-06 0.024
Average value using trapezoidal rule 0.120 10.3E-06 0.176 15.2E-06 0.023

199
6.7 Bare mild steel immersed in 0.5 M sodium chloride solution,
adding 10.0% (w/w) bentonite under cathodic polarisation
protection.

Figure 6.58 shows variation in current density (CD) in mA.m-2 over time in days for
polished bare steel specimens immersed in 0.5 NaCl solutions mixed with 10.0 % (w/w)
bentonite at -0.8V cathodic polarisation potential. Current was measured and recorded
regularly and the values for current density are shown in Appendix IV, Table 25. The
system show a steady fall in current over 7 day period after that the system almost
reached a steady state. The first current was measured after 1 h, expressed as day 0 in
the figure was 149mA.m-2. Then, current was monitored and current density was
calculated. After 1 day, the current was slightly dropped to settle at 135mA.m-2. Current
was gradually dropped to settle at 43.5mA.m2 after 7 days of polarisation. Almost same
current density at the end of the experiment was recorded, 42.8 mA.m-2. The system
was reached a steady state. This incident might be due to the convenient polarization
potential applied to the system; as no evolution for H2 bubbles were observed at the
metal surface, the result is no defect for the polarisation protective film that might be
created on the metal surface due to applying CP.

300
275
Sample 1 Sample 2 Sample 3 Average
250
225
C D (m A .m- 2 )

200
175
150
125
100
75
50
25
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.58: Graph showing current density plot in mA.cm-2 for specimens after
immersed in aerated 0.5 M NaCl solution with 10% bentonite added, under CP at -
0.8V/SCE.

200
Figure 6.59 shows variation in current density (CD) in mA.m-2 over time in days for
polished bare steel specimens immersed in 0.5 NaCl solutions mixed with 10.0 % (w/w)
bentonite, at -1.15V cathodic polarisation potential. Current was measured and recorded
regularly and values for current density are shown in appendix IV, Table 26. The first
current was measured after 1 h, expressed as day 0 in the figure was 2614.4mA.m-2.
Then, current was monitored and current density was calculated. After 1 day, the current
was dropped swiftly at about 50% to settle at 1381.6mA.m-2. More drops in CD were
observed after 3 days to stabilise at about 375.6mA.m-2. Current was gradually dropped
and final measure was recorded after 14 days of polarisation. The minimum current
density at the end of the experiment illustrated was about 118.6mA.m-2.

A rapid fall in current at the initial time might be due to the film that has been formed
and gradually covered the surface of the specimens to work as a physical barrier against
oxygen diffusion. The system shows a faster decline in current density after 2 day
period. This event is due to the continuously growing of the film on the specimens
surface. A steady fall for current was observed over time of the test. It must be pointed
out that, there is no indication for current to tend in the direction of zero [33, 136]. This
event is probably due to the pours nature of Mg(HO)2 film in the form of brucite which
gradually covered most of the specimens surfaces [27]. porous and/or defects may be
caused by the evolution of H2 bubbles from the metal surface, when applying quite high
cathodic potential [27]. At the end of each test, the samples were taken out from the
solution dried and photographed, then stored in a desiccator for SEM and XRD analysis.

201
3000
2750 Sample 1 Sample 2 Sample 3 Average
2500
2250

C D m A .m- 2
2000
1750
1500
1250
1000
750
500
250
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.59: Graph showing plots of current density values in mA.cm-2 for specimens
after immersed in 0.5 M NaCl solution with 10.0% bentonite added, under CP at -
1.15V/SCE.

Figure 6.60 shows pH variation with time for 0.5 M NaCl solution contains 10.0% (w/w)
bentonite at the bulk. The initial pH was about 8. When CP was applied at -1.15V the
pH was gradually increased to somewhat from the initial value. The pH was monitored
for 14 days. After 1 day of polarisation, the pH was 10.2. The greatest pH was about
10.6 which observed after 3 days of polarisation. This increase in pH is attributed to the
formation of OH- ions, due to the cathodic reaction at the working electrode causing an
increase at the interface of metal surface (cathode) and electrolyte. Then, trend was
slightly decreased rather than to increase, and almost has the same trend for the other
periods of the experiment. The lowest pH was recorded after 14 days; the pH was 9.7.
the values for pH was reported in table 27, Appendix IV. It is reported that, any small
variation in pH represents a great change in concentration of hydrogen ions [36]. An
acidic conditions tends to have a depolarising effect, by removing hydrogen films that
has been formed due to applying cathodic protection [62].

202
11
10.8
10.6
10.4
10.2
10
9.8
9.6
pH

9.4
9.2
9
8.8
8.6
8.4
8.2 PH s1 PH s2 PH s3 pHAverage
8
7.8
0 2 4 6 8 10 12 14 16
Time (day)

Figure 6.60: pH for mild steel specimens immersed in 3.0% NaCl solution mixed with
10.0% bentonite under CP (-1.15 V/SCE) for different period.

Figure 6.61 shows the development of films on the specimens surface that were under
cathodic protection at -1.15V for 14 days. In this experiment, the growth of those layers
has almost the same behaviour like that was revealed for the experiment conducted for
the specimens exposed to the solution contains 1.0 (w/w) bentonite (See figure 6.36).
But, a thicker film was observed on most specimens surfaces in this experiment than
those where illustrated in section 6.5.

So, any film detected for both experiment mentioned above expected to form by the
same process and mechanisms that was mentioned early in section 6.5. The alkalinity at
the cathodic regions in the electrochemical cells can cause different possible effects
including, the precipitation of insoluble hydroxide species, at the cathodic sites [137].
For more investigation, SEM/ EDAX and XRD analyses was applied.

203
(a) 1 day (b) 3 days (c) 5 days

(d) 7 days (e) 14 days


Figure 6.61: Film composed on the specimens surface, after immersed in 0.5 M NaCl
solution with 10.0% bentonite, applying CP at -1.15V/SCE for different periods.

As was mentioned before the LPR measurement for specimens under cathodic
protection was applied at OFF potential at the end of each experiment. The CP system
was switched OFF before any LPR experiment was run to enable the system to reach
stability.

Table 6.7 demonstrating the parameters determined for mild steel specimens, after
immersing in aerated 0.5 M NaCl containing 10.0% bentonite, applying CP at -1.150.
The average polarisation resistance (Rp), corrosion current (icorr.) in A.cm-2 and
corrosion rate in mm/y. In this case, cathodic protection at -1.15V/SCE was applied for
different periods; 1 h referred to day 0, 1 day, 5 days, 7 days and 14 days. Stern-Geary
coefficient (B) was assumed 0.026 V/decade and polarisation resistance (Rp) leads to
calculate the corrosion current density by using equation 5.3, section 5.8 [18].

Referring to the data given in this Table, the resistance was increased over time due to
applying CP. The increase in Rp values might be due to the continuous formation of
precipitates on the metal surface. This film may work as a physical barrier against
oxygen or any electrochemical active species to diffuse to the metal surface.
Consequently, corrosion current and corrosion rate decreases. Interestingly, the values

204
of corrosion rate and corrosion current values are fairly small, particularly for
specimens were immersed for more than 1 day. The corrosion rate was in the range of
0.01 to 0.003mm/y. These values can give a good indication that high quality protective
film has been created on the specimens surface due to applying convenient cathodic
protection.

Table 6.7: Summary of parameters determined for mild steel specimens, after
immersing in aerated 0.5 M NaCl containing 10.0% bentonite, applying CP at -1.150 V/
SCE (CP OFF).

Time (day) Average Rp .cm2 Average icorr (mA/cm2) Average CR (mm/y)


0 10730 0.003 0.03
1 11890 0.002 0.03
3 27002 0.001 0.01
5 55132 0.0008 0.01
7 71778 0.0004 0.01
14 87621 0.0003 0.003

6.7.1 XRD Analyses of Specimens Immersed in 0.5 M NaCl Mixed with


10.0% (w/w) Bentonite

In this experiment, the XRD analysis was carried out for specimens exposed to 0.5 M
NaCl solution mixed with 10.% (w/w) bentonite at free corrosion potential, and when
exposed to cathodic polarisation potential, in order to recognise and illustrate the
various compositions for the layers that have been formed on the metal surface. The
information obtained from the XRD was used to compare with data and results that
were achieved form electrochemical measurements in this study.

Figure 6.62: shows XRD pattern with its chemical analysis list for mild steel species
exposed to 0.5 M NaCl solution, adding 10.0% at natural corrosion potential after 14
days. As can be seen, various corrosion products were listed in the pattern list below. It
is confirmed that, the crystalline material layer on the metal surface is mostly
lepidocrocite with goethite; both are in the form of FeOOH. A small amount of halite in
the form of NaCl is also detected. The layer is quite thick as the substrate in the form of
ferrite (Fe) is not visible. Samples immersed in 0.5 M NaCl solution, containing 1% and

205
10% bentonite for the same period (14 days) are very similar. Both contain the same
corrosion products.

C o unts
O ra yith_ 9 0 _ Run2 ; S a m p le 9 0 S e c o nd R un

1600

900

400

100

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula

000080098 51 Lepidocrocite 0.000 0.298 - FeO (OH)


000290713 34 Goethite 0.000 0.255 -FeO(OH)
000050628 71 Halite 0.000 1.021 NaCl
000060696 00 ferrite 0.000 0.000 Fe
Figure 6.62: X-Ray pattern for specimens immersed in 0.5 M NaCl contains10.0%
(w/w) bentonite under cathodic protection at -0.8V at the end of the experiment (14d).

Figure 6.63 and figure 6.64 shows XRD pattern with its chemical analysis list for mild
steel species exposed to 0.5 M NaCl solution mixed with 10.0% bentonite. The
specimens were under cathodic protection, at -0.8V for 14 days. As can be seen, the
following crystalline material was detected on the metal surface; the substrate in the
form of ferrite (Fe) and CaCO3 in the form of calcite. In general, there is no obvious
layer can be exhibited on the specimens surface, but if there is any layer on those
specimens might be very thin, as the substrate pattern is very strong. The diffraction
pattern from the layer is very weak and difficult to identify and the calcite detection was
very limited.

206
C o un ts
O ra yi th_ 1 0 B 3 N a C l_ 8 0 0 _ S 2

10000

2500

0
10 20 30 40 50 60 70 80

P o s i ti o n [2 T h e ta ] (C o b a lt (C o ))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula

000050586 59 Calcite 0.000 0.213 Ca (CO3)


000060696 32 ferrite 0.000 0.412 Fe

Figure 6.63: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -0.8V for 14d).

C o un ts
O ra yi th_ B _ 8 0 0 C p _ S 2

10000

2500

0
10 20 30 40 50 60 70

P o s i ti o n [2 T h e ta ] (C o b a lt (C o ))

Ref. Code Score Compoud Displacement [2Th.] Factor Formula


000050586 17 Calcite 0.000 0.203 Ca(CO3)
000060696 No matching Ferrite 0.000 0.000 Fe

Figure 6.64: X-Ray pattern for steel specimens after immersed in 0.5M NaCl, adding
10.0% (w/w) bentonite under cathodic protection at -0.8 for 14 days.

207
Figure 6.65 - Figure 6.69 show XRD pattern with its chemical analysis list for mild steel
specimens exposed to 0.5 M NaCl solution mixed with 10.0% bentonite. The specimens
were under cathodic protection, at -1.15V for 1, 3, 5, 7 and 14 days, respectively. As
can be seen, various products were listed in the patterns list below. The following
crystalline material was detected on the metal surface; Mg(OH)2 in the form of brucite,
CaCO3 in the form of calcite, small amount of halite in the form of NaCl and the
substrate in the form of ferrite (Fe). Portlandite in the form of Ca(OH)2 was also
detected in some circumstances. Nevertheless, Mg(OH)2 was the main products has
been revealed on the surface of all specimens were under protection for different period.
The strongest reflection related to brucite was detected at approximately 18.5, 38, 51
and 58.5 degree 2theta, in the main graphic, even though there are more than 6 other
reflections at different strength at different degrees represents the brucite. The brucite
layer has gradually increased by the time of the test. Hence, the brucite reflection was
also found to be increased. The continuous formation of brucite can be ascribed to the
increase of the interfacial pH due to the production of hydroxyl ions (OH ) due to
applying cathodic protection.

Calcite was also revealed in different degrees and the strongest reflection was exhibited
at approximately 29 degree, 2theta, particularly for specimens were protected for more
than 1 day. The Ca(OH)2 was only detected in a very weak reflection at approximately
34 degree 2 theta for samples tested for 5 and 14 days (Figures 6.65 and 6.67).
Nevertheless, no portlandite has been detected on the specimens surfaces were tested for
other periods and brucite was detected instead of that reflection at 34 degree 2 theta as
shown in figure 6.67, 6.68, and 6.69.

The crystalline material layer on the metal surface is quite thick, even though, the
substrate in the form of ferrite (Fe) was detected by XRD. But its reflection at about 45,
65 and 83 degrees 2theta was gradually decreased by time of the test, particularly after 3
days period. The layer on the specimen exposed to CP for 7 days may be slightly thicker
as the pattern from the ferrite substrate is weaker compared to the other one that was
protected for 14 days, not what we would expect; the layer may have fallen off from
specimens. Different Ca/MgCO3 pattern chosen as different lattice parameters i.e. Ca:
Mg ratio is varying slightly from one sample to another and this may cause a change to

208
the position of the reflections very slightly. This decrease can be attributed to the
continuously formation of various crystalline materials on the surface of specimens
were under CP, consequently overall increase of thickness of the film. A considerable
amount of halite was also revealed by XRD. Their reflections were at different strength
and different degrees.

C o unts
O ra yith_ 7 9 ; S a m p le 7 9

10000

2500

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Displacement [2Th.] Factor Formula


000070239 77 Brucite 0.000 0.920 Mg(OH)2
000050586 96 Calcite 0.000 0.706 Ca(CO3)
000060696 49 ferrite 0.000 0.019 Fe
000050628 73 Halite 0.000 0.220 NaCl

Figure 6.65: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 1 day.

209
C o unts /s
O ra y i th_ 7 3 ; S a m p le 7 3

900

400

100

0
10 20 30 40 50 60 70 80

P o s iti o n [2 T he ta ] (C o p p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
000070239 87 Brucite 0.000 0.740 Mg(OH)2
000050586 96 Calcite 0.000 0.213 CaCO3
000060696 63 ferrite 0.000 0.412 Fe
000050628 83 Halite, 0.000 0.514 NaCl

Figure 6.66: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 3 days.

C o unts
O ra y i th_ 7 7 ; S a m p le 7 7

40000

10000

0
10 20 30 40 50 60 70 80

P o s iti o n [2 T he ta ] (C o p p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
010822453 74 Brucite 0.000 0.770 Mg(OH)2

210
040078659 83 Calcite 0.000 0.217 CaCO3
000040733 35 Portlandite 0.000 0.065 Ca(OH)2
000060696 49 ferrite 0.000 0.125 Fe
000050628 71 Halite, 0.000 0.218 NaCl

Figure 6.67: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 5 days.

C o unts
O ra yith_ 8 2 ; S a m p le 8 2

10000

2500

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
000070239 83 Brucite 0.000 0.879 Mg(OH)2
000050586 97 Calcite 0.000 0.659 CaCO3
000060696 58 ferrite 0.000 0.050 Fe
000050628 73 Halite, 0.000 0.167 NaCl

Figure 6.68: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 7 days.

211
C o unts
O ra yith_ 8 5 ; S a m p le 8 5

40000

10000

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


[2Th.] Factor Formula
010822453 74 Brucite 0.000 0.992 Mg(OH)2
010713699 87 Calcite 0.000 0.276 CaCO3
000040733 36 Portlandite 0.000 0.097 Ca(OH)2
000060696 47 ferrite 0.000 0.103 Fe
000050628 70 Halite 0.000 0.146 NaCl

Figure 6.69: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 14 days.

Figure 6.70 and Figure 6.71 represent the reproducibility for reflection patterns of
compounds that were accumulated on the metal surface were under CP, for different
period. The change of reflections at different strength, at different degrees in the main
graphics for each compound can be distinguished. For example, the increase in the
reflection related to Mg(OH)2 in the form of brucite can be exhibited, particularly for
specimens exposed to CP for 14 days compared to those were protected for 1 day
(figure 6.70). Furthermore, the specimens where under cathodic protection at -1.15 for 1,
3, 5 and 14 days is very similar to each other, except that the peak at 34 degrees 2theta
related to Ca(OH)2 that only detected for the specimens under CP after 5 days and 14
days (figure 6.71).

212
Figure 6.70: X-Ray pattern for specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 1 and 14 days.

Figure 6.71: X-Ray pattern for specimens after immersed in 0.5 M NaCl contains
10.0% (w/w) bentonite under cathodic protection at -1.15V for 7 and 14 days.

213
6.7.2 SEM/EDAX Analyses of Specimens Immersed in 0.5 M NaCl Mixed
with 10.0% (w/w) Bentonite

SEM/EDAX in the Scanning Electron Microscopy was carried out for specimens
exposed to 0.5 M NaCl solution adding 10.0% (w/w) bentonite at free corrosion
potential, and when exposed to cathodic polarisation potential. As was mentioned early
in this study, the technique was employed to reveal the chemistry of different
compounds that might be formed on the specimens surfaces. Then a quantitative
analysis for those elements, and their progress with time on the metal surface was also
investigated, by using EDAX technique.

Figure 6.72 shows SEM images and X- Ray pattern with its chemical analysis list for
corrosion products formed on mild steel surface exposed to 0.5 M NaCl contains 10.0%
(w/w) bentonite at natural corrosion potential for 14 day. As can be seen, the various
deposits obtained on the metal surface at natural corrosion potential are made of crystals
assembled in different morphology. EDX analysis indicates that the precipitate contains
various elements were illustrated in the pattern list below. The crystalline material on
the surface is mostly corrosion products in the form of ferrite. A strong reflection that
may be related to oxygen was also detected. The pattern was also shown a weak
reflection that may be related to chloride, sodium, carbon and oxygen. It was also
confirmed by XRD analysis; the products detected on the metal surfaces are mostly
lepidocrocite with goethite; both are in the form of FeOOH and quite fair reflection
referred to sodium chloride in the form of halite (See XRD results). It is similar
corrosion products to those detected for specimens were exposed to 0.5 M NaCl
solution containing 1.0% bentonite (See figure 6.46 and Figure 6.47).

214
Element Fe Na Cl O
wt.-% 70.53 4.22 4.08 21.18

Element Fe Cl Na C O
Wt % 62.93 3.39 1.86 0.94 30.88
At % 34.03 2.89 2.44 2.36 58.28
Figure 6.72: SEM and EDAX element analysis for specimen after exposed to 0.5 M
NaCl solution with 10.0% (w/w) bentonite at natural corrosion potential for 14 days.

Figure 6.71 to Figure 6.75 show the SEM images and EDAX element analysis for
deposits on steel metal exposed to 0.5 M NaCl contains 10.0% (w/w) bentonite under
cathodic protection (-1.15V) for 1, 3, 5, 7 and 14 days day, respectively. The deposits
obtained on the metal surface are made of crystals assembled in different morphology.

215
EDX analysis indicates that the products contain magnesium, calcium, Fe substrate,
Carbon, Chloride, sodium and oxygen in different concentrations. A weak reflection
shown for the most patterns related the substrate in the form of ferrite particularly after
3 days period. It was gradually decreased by time of the test due to growing up of the
brucite and calcite film.

It was also confirmed by XRD analysis; the products detected on the metal surfaces are
magnesium hydroxide in the form of brucite calcium carbonate in the form of calcite
and ferrite as metal substrate and sodium chloride in the form of halite. Obviously, the
main deposit for all figures was brucite and calcite.

Element Mg Fe C Cl Ca Na O
Wt % 30.25 20.08 3.15 1.10 0.83 0.64 43.95
At % 26.53 7.66 5.58 0.66 0.44 0.59 58.53
Figure 6.73: SEM and EDAX element analysis for specimen after exposed to 0.5 M
NaCl solution contains 10.0% (w/w) bentonite under CP at -1.15V for 1 day.

216
Element Mg Ca Fe Cl Na C O
Wt % 34.93 7.47 6.18 2.79 1.83 0.00 46.80
At % 29.83 3.87 2.30 1.63 1.65 0.00 60.72
Figure 6.74: SEM and EDAX element analysis for specimen after exposed to 0.5 M
NaCl solution contains 10.0% (w/w) bentonite under CP at -1.15V for 3 days.

Element Ca C Fe Cl Mg Na O
Wt % 61.73 4.96 2.51 1.72 0.74 0.35 27.99
At % 40.09 10.75 1.17 1.26 0.79 0.39 45.54
Figure 6.75: SEM and EDAX element analysis for specimen after exposed to 0.5 M
NaCl solution contains 10.0% (w/w) bentonite under CP at -1.15V for 5 days.

217
Element Ca Mg Fe C Cl Na O
Wt % 30.23 21.78 5.14 3.30 0.98 0.08 38.49
At % 16.93 20.12 2.07 6.16 0.62 0.08 540.1
Figure 6.76: SEM and EDAX element analysis for steel specimen after exposed to 0.5
M NaCl solution contains 10.0% (w/w) bentonite under CP at -1.15V for 7 days.

218
Element Mg Ca Fe Cl Na C O
Wt % 34.54 6.71 3.77 2.45 1.04 0.72 50.77
At % 28.40 3.35 1.35 1.38 0.90 1.20 63.42
Figure 6.77: SEM and EDAX element analysis for steel specimen after exposed to 0.5
M NaCl solution contains 10.0% (w/w) bentonite under CP at -1.15V for 14 days.

6.8 Bare mild steel specimens after burying inside the bentonite
layer below the solution level of 0.5 M sodium chloride solution
adding 10.0% (w/w) bentonite with no cathodic protection
(OCP)

For this experiment, we followed the same procedure has been conducted for the
previous experiments. The only different is the specimens for this experiment were
buried inside the bentonite layer below the solution level (See Figure 6.78 (a)).

Figure 6.78 shows photographs for mild steel specimens, after they were buried inside
bentonite layer for 1, 3, 5, 7 and 14 days respectively, in 0.5M NaCl solutions adding
10.0% (w/w) bentonite. The tests were under stagnant, aerated conditions and at room
temperature. The initial bulk solution pH was 8. Two duplicate specimens were buried
inside bentonite layer for each period that was mentioned early. In general, the
specimens surfaces were slightly corroded. Corrosion products can be seen and the
smooth, shiny grey colour has been changed to dark grey colour for all specimens as

219
shown in Figure 6.78. The corrosion rate was determined based on the weight loss
measurements.

(a) Specimens buried (b) 1d (c) 3d


inside bentonite layer

(d) 5d (e) 7d (f) 14d

Figure 6.78: Photographs of surfaces of mild steel specimens after exposed for
different periods in 10.0% bentonite mixed with 3.0% NaCl solution for different
time. The specimens were buried inside the bentonite layer at the bottom of the beaker
below the solution level.

Figure 6.79 shows the corrosion rate at OCP for specimens buried for the same periods
as mentioned before. The values for this graph were illustrated in Appendix VI, Table
28. The corrosion rate was about 0.03 mm/y at the end of day 1. This value was derived
from the weight loss after 1 day of the test. After that the corrosion rate decreased
rapidly, which was about 0.012 mm/y at the end of day 3. It is obvious that the trend of
corrosion rate decreased gradually and slowly; 0.009 for day 5, 0.008 for day 7 and
finally settled at about 0.006 mm/y at the end of the experiment which was after 14 days.

As we expected, the corrosion rate of carbon steel falls with time. The decrease in
corrosion rate can be attributed to the corrosion products on the interface which work as
a barrier against diffusion of active species. Furthermore, bentonite clay characterised

220
by its low permeability, so oxygen faces some difficulty to penetrate through the clay
[84, 85]. Different experiments undertaken to study the corrosion of carbon steel
exposed to water-saturated compacted bentonite, under aerobic conditions. Hunda with
his work-group in 1997 [89] have reported no corrosion process taken place after the
corrosion products form on the metal surface. For this experiment, the corrosion rate of
0.006mm/y after 14 days of exposure at OCP was recorded. This value was remarkably
lower than the conservative setting value used by most repository designers in Japan
[88]. The corrosion rate of carbon steel in compacted bentonite was also reported as
0.1m/year by X. Xiaa with his work-group at late (2005) [88]. The corrosion rate value
agreed unpredictably well with the corrosion rate, 0.006mm/y after 14 days of exposure
that was experimentally measured in our experiment.

0.04
0.035 Sample 1 Sample 2 Avg. C.R
C o rro sio n ra te (m m /y )

0.03
0.025
0.02
0.015
0.01
0.005
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.79: Graph showing plot of average corrosion rate/ time measurements over a
14 days tests period for mild steel specimens after burying inside the bentonite layer at
the bottom of the beaker below the solution level. The electrolyte was 0.5M NaCl
solution, with addition of 10.0% (w/w) bentonite.

6.8.1 Optical Microscopy for specimens buried inside the bentonite


layer

Optical Microscopy at magnification X20, and thermodynamic observation are


proposed for classifying the various corrosion products according to their colour
changes during the corrosion process and electrochemical reactions. These techniques
were used for substrate morphology assessment and to examine the nature of the

221
corrosion products on the metal surface before and after immersion for different
exposure periods.

Figure 6.80 (a) shows a cleaned specimen prepared for weight loss test carried out
before it was buried in bentonite. It is a shiny, silver grey in colour. The specimens b, c,
d and e for specimens buried inside the bentonite layer at the bottom of the beaker
below the solution level for periods of 1, 3, 5, 7 and 14 days consequently after they
were cleaned by Clarkes solution. The procedure was illustrated in chapter 5, section
5.6. The progress of corrosion can be observed due to continuously change of
specimens surface morphology. After cleaning, the silver grey specimens were changed
in colour to darker grey. General corrosion combined with different black spots was
detected for most specimens buried in bentonite side. Nevertheless, more attack and
black spots were started to develop by time of exposing. Pitting shape, particularly for
specimens buried in bentonite for 7 days and 14 days period were exhibited. It is well
known that, bentonite is characterised by its low permeability. Therefore it has been
used as a barrier or sealing material, so oxygen faces some difficulty to penetrate
through it which may causes corrosion cells at different areas [84, 85]. Consequences,
localised corrosion might be expected [86].

(a) Before test (b) 1d

222
(c) 3d (d) 5d

(e) 7d (f) 14d


Figure 6.80: Optical micrographs for specimens were buried inside the bentonite layer
at the bottom of the beaker below the solution level for different time. The electrolyte
was 10.0% (w/w) bentonite mixed with 3.0% NaCl solution.

Figure 6.81 shows the natural corrosion potential variation over time for polished bare
steel specimens buried inside the bentonite layer below the solution level of 0.5 NaCl
solutions containing 10.0% (w/w) bentonite for 1, 3, 5, 7 and 14 days. The values of
corrosion potential are shown in Appendix V, Table 29. The OCP was measured after 5
minutes, -0.312V vs. SCE. It was gradually shifted to more negative potential to settle
at about -0.797V after 1 day. Then slightly change was observed to be in the range of -
0.807 to -0.811V for specimens buried for 5 and 7 days, consequently. The trend of
corrosion potential-time was slowly shifted to more negative to settle at about -0.805V
after 14 days of exposing.

223
The first stage of corrosion for mild steel specimens were buried inside bentonite layer
could be controlled by the reduction of Oxygen to form free hydroxyl ions, due to the
main following electrochemical reaction:

O2 + 2H2 O + 4e 4OH- (2.2)

The amount of O2 that was trapped in the bentonite due to its initial stage of filling may
be gradually consumed due to the above reaction. It is well known that bentonite is
characterised by its low permeability and oxygen faces some difficulty to penetrate
through the clay. So the above reactions could be slowed or ceased due limiting amount
of oxygen [84, 85].

Further corrosion process due to presence of water in the bentonite layer is most likely
to occur due to the following reduction reaction to produce hydrogen gas and to form
free hydroxyl ions:

2H2O + 2e H2 + 2OH (2.3)

The continuous production of insoluble corrosion products in the form of ferrous


hydroxide Fe(OH)2 might be formed on the specimens surface according to the
following reaction:

2Fe + 2H2O + O2 2Fe(OH)2 . (2.22)

Further reactions of hydrous ferrous oxide or ferrous hydroxide on the metal surface
with oxygen may also expected to occur to produce hydrous ferric oxide or ferric
hydroxide (Fe(OH3) [22] as given in the following equation:

4Fe(OH)2 + 2H2O + O2 4Fe(OH)3 ... (2.24)

224
SEM/EDAX and XRD analysis was performed on the specimens were buried in the
bentonite side for 14 days period. The results will be presented in section 6.9.1.

Uniform corrosion is most likely to be the dominant form of corrosion for mild steel in
bentonite, as the OCP is probably stable [87]. On the other hand, pitting has been
exhibited at different location on the specimens surface. This can be due variation in
the solution chemistry at different location, causing corrosion cells at different areas.
localised corrosion is also suggested to be the dominant, if mild steel in bentonite
becomes passive [86].

-250
-300
-350 Test 1 Test 2 Test 3 average
P o ten tia l V s. S C E m V

-400
-450
-500
-550
-600
-650
-700
-750
-800
-850
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.81: Graph showing plot of corrosion potential for mild steel specimens buried
inside the bentonite layer, below the solution level of 0.5 M NaCl solution contains 10%
(w/w) bentonite for 14 days.

Figure 6.82 shows the polarisation resistance (Rp) variation with time. It was run at
regular intervals at freely corrosion potential for several times and the Rp values
presented in these trends were extracted from the Rp plots generated by LPR runs over
time that shown in Appendix V, Table 31. The initial polarisation resistance curve was
recorded after 1h (day 0) of the test which was about 5004.2 Ohm.cm2. The trend was
dramatically increased with time. The Rp was 23401.5ohm.cm2, at the end of day 1.
After that, the Rp for day 3 was gradually increased to settle at about 59364.1ohm.cm2.

225
Significant increase for Rp was recorded for specimens were buried inside the bentonite
layer for 5, 7 and 14 days and the Rp values were 81451.3, 90622.8 and
111842.4ohm.cm2 consequently. For the same specimen the Rp was run for several
times (minimum 3 times), to guarantee the plots of Rp have been repeated in high
accuracy; graphs showing plots of Rp were presented in Appendix V, Figure 25 to 27.
Polarisation resistance was run at regular intervals at the free corrosion potential.

120000

100000
R p (O h m .c m2 )

80000

60000

40000

20000
Sample 1 Sample 2 Sample 3 average
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.82: Graph showing the values of Rp/time measurements for mild steel
specimens buried inside the bentonite layer below the solution level of 0.5 M NaCl
contains 10% (w/w) bentonite for 14 days.

A summary of results, comparing the corrosion rate has been produced from LPR
measurements and the one was experimentally determined by using weight loss
measurement will be offered. The experimental Tafel constants B extracted from
corrosion rate was also compared with the one that has been assumed (0.026 V/decade).

Table 6.8 demonstrating the parameters determined for mild steel specimens, burying in
bentonite layer for 14 days at OCP. As was expected, the greatest corrosion rate and
corrosion current values for both cases were recorded for specimens buried inside the
bentonite layer for 1 day period. After that, a steady fall of corrosion rate and corrosion
current was observed for both cases. Negligible corrosion rate and corrosion current
were observed over time of exposing for both cases, but experimental corrosion rate
values were mostly greater than the corrosion rate values obtained from LPR; for
example, after1 day period the experimental corrosion rate was 3 times greater than that
was obtained from LPR and it was 1.5 times after 14 days period. Identical values for

226
constant B were obtained for most tests. The average value was ~ 0.079 V/decade,
which was slightly high compared to that has been assumed to calculate the corrosion
rate, 0.026V/decade.

The data related to Rp given in Figure 6.80 for specimens buried inside bentonite at
OCP was dramatically increased with time. Consequently, corrosion current and
corrosion rate decreased (See Table 6.8). The average value of Rp for specimens
exposed to the environment for 14 days was ~111842.cm2. So, corrosion rate was
~0.005mm/y compared. In this case, corrosion process was probably controlled by
oxygen diffusion [48, 51]. Obviously, low corrosion rate was due to the limiting amount
of oxygen, as it is facing difficulty to penetrate through the clay. So, trapped oxygen in
bentonite was gradually consumed. Hence, oxygen reduction reactions could be slowed
or ceased [84, 85]. Corrosion rates as measured by weight loss are approximately three
times that obtained by LPR. This is hardly surprising since LPR is an instantaneous
measurement whereas weight loss is cumulative.

Table 6.8: Summary of parameters determined for mild steel specimens buried inside
bentonite layer at the bottom of the beakers below the solution level of an aerated 0.5 M
NaCl Solution containing 10.0% bentonite for 14 days at OCP.
Time (a) Av CR (a) icorr (A/cm2) (b) CR exp. (b) icorr. .exp. (c) Bexp.
day mm/y (B=0.026V) mm/y (A/cm2) (V)
1 0.011 9.38E-07 0.030 2.97E-06 0.070
3 0.005 4.26E-07 0.012 1.89E-06 0.112
5 0.003 3.24E-07 0.009 8.91E-07 0.073
7 0.003 2.56E-07 0.008 7.92E-07 0.072
14 0.003 2.56E-07 0.006 5.94E-07 0.066

6.9 Bare mild steel buried inside the bentonite layer of 0.5 M sodium
chloride solution containing 10.0% (w/w) bentonite under CP

Figure 6.83 showing variation in CD in mA.m-2 at -0.80V vs. SCE for mild steel
specimens buried inside bentonite layer for 14 days. Current was measured and
recorded regularly. The values for current density are shown in Appendix V, Table 32.
The average current density measured after 1 h, expressed as day 0 was ~50.2 mA.m-2.

227
Then, it was dropped swiftly for more than 17 times than the first current to settle at
about 3mA.m-2, at the end of day 1. Steady fall was observed after day one to stabilise
at ~-2.0 mA.m-2 after 5 days of polarisation. Slight increase in current density was

observed for the other period of the test. The system was reached a steady state after day
7. The final current was measured after 14 days of polarisation, -1.7mA.m-2.

The pH of bulk solution above the bentonite layer was monitored for specimens under
CP at -0.8 V. The minimum pH was recorded after 14 days. The pH value was about 6.
The drop in pH might be due to anodic reaction that has been taken place on specimens
electrodes as polarisation potential applied to protect the system is more positive than
natural corrosion potential. The values for pH were illustrated in Appendix V, Table 33.

As we expected, system under CP at -0.8V, negative current was observed, -0.3mA.m-2


after 2 days. This was attributed to OCP been more negative than the applied potential.
So, in this case, corrosion might have being taken place as the protection system works
in a reverse. XRD analysis has shown no corrosion product crystals were formed on the
metal surfaces rather than the substrate in the form of ferrite (Fe) for specimens under
CP at -0.8 for 14 days. More XRD results will be offered in Section 9.1.

70
Sample 1 Sample 2 Sample 3 average
60
50
C D m A .m- 2

40
30
20
10
0
-10
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Time (day)

Figure 6.83: Graph showing plot of current density/ time for mild steel specimens
buried inside bentonite layer at the bottom of the beaker below the solution level of an
aerated 0.5 M NaCl Solution adding 10.0% bentonite, under CP at -0.80V vs. SCE.

228
Figure 6.84 showing CD in mA.m-2 for specimens buried for the same periods were
mentioned early in the previous Figure under CP at -1.15V instead of -0.8V vs. SCE.
Current was measured and recorded regularly and the values for current density are
shown in Appendix V, Table 34. The average current density measured after 1 h,
expressed as day 0 in the figure which was ~ 1068mA.m-2. Then, it was dropped swiftly

to settle at about 207.4mA.m-2 at the end of day 1. The system show a steady fall in
current after day 1 period but the system almost reached a steady state at the end of the
experiment. The current was dropped for more than 19 times than the initial current to
settle at about 60mA.m-2.

1250
Sample 1 Sample 2 Sample 3 average
1000
C D m A .m- 2

750

500

250

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.84: Current density in mA.m-2 for mild steel specimens buried inside the
bentonite layer at the bottom of the beaker below the solution level to an aerated 0.5 M
NaCl Solution mixed with 10.0% bentonite, under CP at -1.15V/SCE.

The pH of bulk solution above the bentonite was monitor and recorded; the values for
pH were illustrated in Appendix V, Table 35. The pH was slightly increased for the first
5 days and the greatest pH recorded was only 0.4 from the initial value, 8. Then it was
slightly dropped over time to settle at about 7.8 for most cells.

At the end of each test, the samples were taken out from the solution dried and then
stored in a desiccator for SEM/EDAX and XRD analysis. No obvious layers were

229
observed after the specimens were taken out from the cells at the end of the experiment.
Even though, the current was extremely dropping over time. This drop might be
ascribed to the following:

Initially, hydrated sodium bentonite may swell more than 20 times its original volume,
so bentonite layer increased [78]. As bentonite is characterised by its low permeability,
oxygen faces difficulty to penetrate through the thick barrier towards the specimens
surface. As a result, current will be decreased.

6.9.1 SEM and XRD Analyses for specimens buried inside the bentonite
layer

6.9.1.1 XRD Analyses for Specimens buried inside 0.5 M NaCl solution
added to 10.0% (w/w) Bentonite

Figure 6.85 shows XRD pattern with its chemical analysis list for mild steel species
buried inside bentonite layer below the solution level of 0.5 M NaCl solution, adding
10.0% bentonite at natural corrosion potential for 14 days. As can be seen, the layer on
this sample is very thin, as the substrate pattern is very strong. The diffraction pattern
from the layer is very weak and difficult to identify. Akaganeit and a very small amount
of NaCl may be present; however it is difficult to identify such a weak diffraction
pattern with any degree of certainty. The akaganeit -FeOOH product normally found in
environments that contain chlorides, as in marine environments [25].

230
C o unts
O ra yith_ 3 7 ; S a m p le 3 7

10000

2500

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
000060696 73 Ferrite 0.000 0.926 Fe
000341266 11 Akaganeit 0.000 0.004 - FeO(OH)
000050628 28 Halite 0.000 0.011 NaCl

Figure 6.85 X-Ray pattern for specimen buried in 0.5 M NaCl contains 10.0% (w/w)
bentonite with no protection polarisation (OCP) for14 days

Figure 6.86 and Figure 6.87 show XRD pattern with chemical analysis list for mild steel
specimens buried inside bentonite layer below the solution level of 0.5 M NaCl solution,
adding 10.0% bentonite. The specimens were under cathodic protection, at -0.8V and -
1.15 for 14 days respectively. As can be seen, there is no obvious layer can be exhibited
on the specimens surfaces for both cases, but if there is any layer it might be very thin,
as the substrate pattern is very strong. Even though, no obvious film was detected in
both cases but extreme drops in current densities were observed. This might be due to
the thick layer of bentonite insulating the specimens from oxygen and other aggressive
species. More details can be found in Section 6.9. The negative average current density
was recorded for the specimen was under CP at -0.8 V was attributed to the natural
corrosion potential was more negative than the potential applied to protect the system.
The current density for specimens under CP at -1.15V was also very small compared to
the previous results for all experiments conducted before. A negligible corrosion rate
was obtained from weight loss test after 14 day period, 0.006mm/y and the corrosion
current was 5E-07 (see table 6.7).

231
C o unts
O _ S 2 _ 1 0B B _ 8 5 0

40000

10000

0
10 20 30 40 50 60 70 80

P o sitio n [2 The ta ] (C o b a lt (C o ))

Visible Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
* 000060696 90 ferrite 0.000 0.992 Fe

Figure 6.86: X-Ray pattern for specimen buried in 0.5 M NaCl contains 10.0% (w/w)
bentonite under cathodic protection at -0.80 V for14 days.

C o unts
O ra yith_ 4 4 ; S a m p le 4 4
10000

2500

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
00-006- 75 ferrite 0.000 0.945 Fe
0696
Figure 6.87: X-Ray pattern for specimen after burying inside the bentonite layer of 0.5
M NaCl contains 10.0% (w/w) bentonite under cathodic protection at -1.15 V for 14
days.

232
6.9.1.2 SEM/EDAX analysis for specimens buried inside 0.5 M NaCl solution with
10.0% (w/w) Bentonite

Figure 6.88 shows two images in two different magnifications (500x and 2000x),
spectrum, and EDAX quantitative data for specimen after burying inside the bentonite
layer, below the solution level of 0.5 M NaCl solution contains 10.0% (w/w) bentonite
at OCP for 14 days. As can be seen the relative percentage of different elements
obtained using EDAX quantification are listed in below for this sample. The highest
percentage that was detected by EDAX is mostly iron, which was about 65%, oxygen at
about 17.7%. Sodium and chloride elements at different percentage were also detected;
11.4% and 6.6% respectively. Traces of Ca and Mg were also detected. Only iron
related to the substrate was detected using XRD analysis. This might be due to the very
then layer which could not be detected.

233
Element Fe Na Cl Ca Mg C O
Wt % 65.10 7.56 6.75 1.55 1.39 0.00 17.65
At % 40.43 11.40 6.60 1.34 1.98 0.00 38.25
Figure 6.88: SEM and EDAX element analysis for specimen after burying inside the
bentonite layer, below the solution level of 0.5 M NaCl solution contains 10.0% (w/w)
bentonite at natural corrosion potential (OCP) for 14 days.

Figure 6.89, Figure 6.90 show SEM and EDAX element analysis for mild steel
specimens buried inside bentonite layer below the solution level of 0.5 M NaCl solution,
adding 10.0% bentonite. The specimens were under cathodic protection, at -0.8V and -
1.15 for 14 days respectively. As can be seen, there is no obvious layer can be exhibited
on the specimens surfaces for both cases, but if there is any layer might be very thin, as
the substrate pattern is very strong. The relative percentage of different elements
obtained using EDAX quantification is listed below for these specimens; iron, oxygen
and carbon. The highest percentage detected by EDAX for both specimens were under
CP at -0.8 and -1.15V was for iron, which was related to the substrate. This was also
confirmed early, using XRD analysis.

234
Element Fe O C
Wt % 89.11 7.06 3.82
At % 67.74 18.74 13.52
Figure 6.89: SEM and EDAX element analysis for specimen buried inside the bentonite
layer at 0.5 M NaCl solution contains 10.0% (w/w) bentonite at -0.80 polarisation
potential for 14 days.

235
Element Fe Ca Mg O
Wt % 98.59 0.62 0.32 0.46
At % 96.84 0.85 0.73 1.58
Figure 6.90: SEM and EDAX element analysis for specimen after burying inside the
bentonite layer at 0.5 M NaCl solution contains 10.0% (w/w) bentonite at -1.15V
polarisation potential for 14 days.

6.10 Bare Mild Steel Immersed in 0.5 M NaCl solution adding 10.0%
(w/w) Bentonite and 500 ppm and 1000 ppm Zinc chloride
(ZnCl2)

ZnCl2 added to the bentonite as a cathodic inhibitor to investigate its effect and how it
behaves on corrosion of mild steel. For this set of experiments a similar investigation as
that was illustrated early in section 6.6 and 6.7 was followed. The procedure used for
this experiment has been already described in chapter 5. Briefly, 500ppm and 1000 ppm
of ZnCl2 was added to the 0.5 NaCl solutions containing 10% bentonite, respectively.
Polarisation potential at OCP and -0.115V against SCE were applied.

Figure 6.91 and 6.92 show the corrosion potential vs. time plots for specimens at OCP
obtained after 2 weeks of immersion in 500 ppm and 1000 ppm of ZnCl2 were added to
0.5 NaCl solution containing 10% bentonite. It is a very rapid change for corrosion
potential values toward the negative direction for the first hour of immersion. The
values for corrosion potential are shown in Appendix VI, Table 36 and Table 37. The

236
potential was measured periodically; after 24 hours, the average potential was gradually
shifted to more negative to settle at about -0.700V and -0.707V for specimens immersed
in the solution contains 500ppm and 1000ppm ZnCl2 respectively. Almost, same
corrosion potential values were recorded with a slight shift to more negative for samples
immersed in solution contains 500ppm ZnCl2 particularly for specimens immersed for 1,
3 and 5 days period. The development of insoluble corrosion products on the metal
surface was also observed. The specimens surfaces were covered by thick and drake
brown deposits. The deposits were considerably different from that were visually
appeared in the previous experiments for specimens exposed to different solutions
contain no ZnCl2.

-600

-620 Sample 1 Sample 2 Sample 3 average


P o ten tia l m V V s. S C E

-640

-660

-680

-700

-720

-740
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.91: Graph showing plots of potential- time curves at open circuit for mild steel
specimens immersed in aerated and stagnant 0.5 M NaCl solutions, contains 10%
bentonite and ZnCl2 at 500ppm.

237
-580

-600 Sample 1 Sample 2 Sample 3 average

P oten tial m V V s. S C E
-620

-640

-660

-680

-700

-720
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.92: Graph showing plots of potential- time curves at open circuit for mild steel
specimens immersed in aerated and stagnant 0.5 M NaCl solutions, contains 10%
bentonite and ZnCl2 at 1000ppm.

Figure 6.93 shows photographs for 2 sets of mild steel specimens tested at open circuit
obtained after 2 weeks of immersion in a solution contains (a) 500 ppm and (b) 1000
ppm of ZnCl2. The tests were carried out at stationary, aerated conditions and at room
temperature. The initial pH for the solution used was about 8. In general, the specimens
were corroded in different levels. A thick and stable brown corrosion product was
formed and covered the most specimens surfaces; particularly for the specimens were
immersed in the solution contains 1000ppm ZnCl2. It is obvious that the corrosion
products was more stable on the metal surfaces for all specimens corroded in 0.5M
NaCl solution contains bentonite and ZnCl2 compared to the specimens exposed to the
0.5 M NaCl solution free of bentonite.

238
(a) (b)
Figure 6.93: Photographs of surfaces of the 4 mild steel specimens after immersion in
0.5 M NaCl solution adding 10% bentonite and ZnCl2 at (a) 500ppm and (b) 1000ppm
respectively at open circuit potential for 14 days.

Figure 6.94 and Figure 6.95 show the average values of polarisation resistance (Rp)
variation with time. The experiment was conducted for several times and the Rp values
presented in these graphs were extracted from the Rp plots generated by LPR runs over
time that shown in Appendix VI, Table 38 and Table 39. The initial Rp values were
recorded after 1h (day 0) of the test which was about 2597 and 2509 Ohm.cm2
respectively for specimens after exposing to the solutions contains 500ppm and
1000ppm ZnCl2. After that, the Rp values were continuously increased for samples after
immersed for 1, 3, 5 and 7 days in the solution that contains 1000ppm but the Rp was
only increased for the first 3 days for specimens were immersed in the solution
containing 500ppm ZnCl2. Then, Rp values was gradually decreased for the other
periods of the tests to settle at 2560.7 Ohm.cm2 and 3867.3 Ohm.cm2 respectively after
14 days of immersion in the both solutions that contains 500ppm and 1000ppm ZnCl2.

239
6000

5000
R p (O h m .c m2 )
4000

3000

2000

1000
Sample 1 Sample 2 Sample 3 average
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)
Figure 6.94: Plots of Rp values for specimens after immersing in 0.5 M NaCl solution
adding 10% bentonite and ZnCl2 at 500ppm at OCP for 14 days.

7000
6000
5000
R p (O h m .c m2 )

4000
3000
2000
1000
Sample 1 Sample 2 Sample 3 average
0
0 2 4 6 8 10 12 14 16
Time (day)
Figure 6.95: Plots of Rp values for specimens after immersing in 0.5 M NaCl solution
adding 10% bentonite and ZnCl2 at 1000ppm at OCP for 14 days.

Table 6.9 illustrates a summary for values of corrosion rate and corrosion current has
been produced from LPR measurement for mild steel specimens after immersing in 0.5
M NaCl solution adding 10% bentonite and ZnCl2 at 500ppm and 1000ppm respectively
at OCP for 14 days. constant (B) was assumed 0.026 V/decade and polarisation
resistance (Rp) leads to calculate the corrosion current density by using equation 5.3,
section 5.8 [18]. Note: we assumed bc = , since cathodic reaction is probably
controlled by diffusion of oxygen reduction, a value of 0.026V/decade is commonly

240
used [40, 120]. The values for CR and corrosion current for specimens were immersed
in the solutions contains ZnCl2 at both concentrations almost have similar values for all
period. In general, negligible change and low corrosion rate was recorded over time.
The average corrosion rate was 0.100mm/y and 0.0713mm/y for specimens after
exposed to the 0.5NaCl solution, containing bentonite and ZnCl2 at 500ppm and
1000ppm respectively for 14 days. The slight change in corrosion rate and corrosion
current for specimens in both solutions was attributed to dissimilar surface condition at
the time of measurement. It is reported by many authors [18, 113] that the corrosion rate
is varied over time.

Table 6.9: Summary of parameters determined for mild steel specimens after immersed
in 0.5 M NaCl Solution containing 10.0% bentonite and ZnCl2 at 500ppm and 1000ppm
at OCP for 14 days.

Time icorr (A/cm2) Av CR mm/y icorr (A/cm2) Av CR mm/y


day 500ppm 500ppm 1000ppm 1000ppm
1 7.26E-06 0.0843 5.32E-06 0.0617
3 6.23E-06 0.0720 6.30E-06 0.0733
5 8.34E-06 0.0966 6.04E-06 0.0700
7 1.01E-05 0.1263 5.68E-06 0.0707
14 1.042E-05 0.1206 7.00E-06 0.0813

6.11 Mild steel specimens after immersing in 0.5 M NaCl solution,


containing 10% (w/w) bentonite and ZnCl2 at 500ppm and
1000ppm respectively at -01.15V CP for 14 days.

Figure 6.96 and Figure 6.97 showing plots of CD variation in mA.m-2 over time in days
for polished bare steel specimens after immersing in 0.5 M NaCl solution adding 10%
bentonite and ZnCl2 at (a) 500ppm and (b) 1000ppm respectively at -01.15V CP for 14
days. Current was measured and recorded regularly and the average current density was
shown in Appendix VI, Table 42 and Table 43 for specimens after immersing in the 0.5
M NaCl solution adding 10% bentonite and ZnCl2 at 500ppm and 1000ppm respectively.
The average current density measured after 1 h and it is expressed as day 0 in the
figures which was 549.2mA.m-2 and 753.8mA.m-2 respectively for specimens after
exposed to the solutions contains 500ppm and 1000ppm ZnCl2. Then, current was

241
continuously decreased and current density was calculated. The minimum CD was
recorded after 7 days of polarisation; 22.1mA.m-2 and 19.05mA.m-2 respectively for
specimens immersed in the solution contain 500ppm and 1000ppm ZnCl2. After 14 days
of polarisation CD was slightly increased to settle at about 27.3mA.m-2 for specimens in
exposed to both solutions. The system almost reached to a steady state. The system
shows a faster decline in current density after 2 day period. This event is due to the
continuously growing of the film on the specimens surface. A steady fall for current
was observed over time of the test. It must be pointed out that, there is no indication for
current to tend in the direction of zero [33, 136]. This event is probably due to the pours
nature of Zn(OH)2 film in which was gradually cover the most specimens surfaces.
porous and/or defects may be caused by the evolution of H2 bubbles from the metal
surface, when applying quit high cathodic potential [27].

700
Sample 1 Sample 2 average
600
500
C D m A .m- 2

400
300
200
100
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.96: Graph showing plots of values for current density in mA.cm-2 for mild steel
specimens after immersed in 0.5 M NaCl solution adding 10% bentonite and ZnCl2 at
500ppm at -01.15V CP for 14 days.

242
800
700 Sample 1 Sample 2 average

C D m A .m- 2 600
500
400
300
200
100
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Time (day)

Figure 6.97: Graph showing plots of values for current density (CD) in mA.cm-2 for
mild steel specimens after immersed in 0.5 M NaCl solution adding 10% bentonite and
ZnCl2 at 1000ppm at -01.15V CP for 14 days.

Figure 6.98 Showing photographs for specimens after immersed in 0.5 M NaCl solution
adding 10% bentonite and ZnCl2 at (a) 500ppm and (b) 1000ppm respectively at -
01.15V CP for 14 days. The growth of layer on the surfaces has almost same
performance. The only different was a thicker film was observed for the 2 specimens in
the Figure 6.100 (b). So, any film detected for both experiment mentioned above
expected to form by the same process and mechanisms that has been mentioned in
section 6.5. The Alkalinity at the cathodic regions in the electrochemical cells can cause
different possible effects including, the precipitation of insoluble hydroxide species, at
the cathodic sides. For more investigation, SEM/ EDAX and XRD analyses was applied.

243
(a) (b)
Figure 6.98: Figure 6.100: Photographs showing surfaces of 4 mild steel specimens
after immersion in 0.5 M NaCl solution adding 10% bentonite and ZnCl2 at (a)
500ppm and (b) 1000ppm respectively at -01.15V CP for 14 days.

6.11.1 XRD of Steel Specimens Immersed in 0.5 M NaCl Contains 10.0%


(w/w) Bentonite and 500ppm and 1000 ppm ZnCl2 under Cathodic
Protection at -1.15V, after 14 days

In this experiment, the XRD analysis was carried out for specimens immersed in the
solution contains 500 and 1000 ppm ZnCl2 respectively at free corrosion potential and
when cathodic polarisation potential applied at -1.15V, in order to recognise and
illustrate the various compositions for the layers that have been formed on the metal
surface. The information obtained from the XRD was used to compare with data and
results that were achieved form electrochemical measurements in this study.

Figure 6.99 and Figure 6.100 show XRD patterns together with the chemical analysis
list for mild steel species exposed to 0.5 M NaCl solution, adding 10.0% at natural
corrosion potential after 14 days. As can be seen, the only crystalline phases present for
the 500ppm concentration specimen are halite & goethite. However, various crystalline
phases for the 1000ppm concentration ZnCl2 specimen were listed in the patterns list
including lepidocrocite, goethite, the zinc oxide and halite. The substrate was also
detected for both concentrations. The layer is quite thick as the substrate in the form of
ferrite (Fe) in the case of 500 ppm not visible compared to the specimen in 1000 ppm
concentration.

244
C o unts

10000

2500

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Name Scale Chemical Formula


00-005-0628 77 Halite, syn 0.936 NaCl
01-081-0462 48 Goethite, syn 0.154 FeO(OH)
00-006-0696 33 ferrite, substrate 0.039 Fe
Figure 6.99: X-Ray pattern for steel specimens after immersion in 0.5 M NaCl
containing 10.0% (w/w) bentonite and ZnCl2 500 ppm at OCP for 14 days.

C o unts

1600

900

400

100

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C op p e r (C u))

Ref. Code Score Compound Name Scale Factor Chemical Formula


00-005-0628 66 Halite, syn 1.001 NaCl
00-006-0696 50 Ferrite, substrate 0.667 Fe
00-032-1100 17 -Na Fe2 O3 0.866 NaFe2O3
00-036-1451 36 Chinese white, zinc 0.561 ZnO
00-008-0098 28 Lepidocrocite 0.181 FeO(OH)
00-029-0713 27 Goethite 0.263 FeO(OH)
Figure 6.100: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 1000 ppm at OCP for 14 days.

245
Figure 6.101 and Figure 6.102 shows XRD pattern with its chemical analysis list for
mild steel species exposed to 0.5 M NaCl solution contains 10.0% (w/w) bentonite and
ZnCl2 1000 ppm for specimens were under cathodic protection at -1.15V for 14 days,
respectively. As can be seen, various products were listed in the pattern list below. The
layer is mostly Pentazinc hexahydroxide carbonate. The ZnO may also present but the
pattern is almost completely overlapped for both cases. A very complex XRD trace of
metallic Zn appears to be present. These peaks are particularly clear at about 71 degrees.
A small amount of NaCl as halite is also present in small amount. Even though, the
crystalline material layer on the metal surface is quite thick, the substrate in the form of
ferrite (Fe) was detected by XRD. Its reflection at 45, 65 and 83 degrees 2theta was
weak for both tests. This can be attributed to the overall increase of thickness of the film
due to applying cathodic protection.
It was confirmed that, the addition of ZnCl2 at two different concentration, 500ppm and
1000ppm have reduced the current density from 549 mA/m2 to 27 mA/m2 and from 753
mA/m2 to 27.3 mA/m2 between days 1 and 14 respectively. This was ascribed to the
continuous formation of uniform film on the metal surface. Identical formation of
different products was also illustrated in figure 6.103 for the solutions contain ZnCl2 at
the two different concentrations.

C o unts

3600

1600

400

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C o p p e r (C u))

Ref. Code Score Compound Factor Chemical Formula


00-0060696 54 Ferrite,substrate 0.205 Fe
01-070-2509 56 Halite, syn 0.279 NaCl
00-036-1451 51 Chinese white zinc 0.461 ZnO
00-004-0831 69 Zinc, syn 0.548 Zn
01-072-1100 46 Pentazinc hexahy-droxide carbonate 0.837 Zn5(OH)6(CO3)2
Figure 6.101: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 500 ppm under CP at 1.15V for 14 days.

246
C o unts

3600

1600

400

0
10 20 30 40 50 60 70 80

P o s itio n [2 The ta ] (C o p p e r (C u))

Ref. Score Compound Name Scale Chemical


Code Factor Formula
00-0060696 55 Ferrite, substrate 0.152 Fe
00-0050628 69 Halite, syn 0.358 Na Cl
00-0361451 56 Chinese white 0.450 Zn O
zinc
00-0040831 78 Zinc, syn 0.546 Zn
01-0721100 66 Pentazinc hexahy- 0.692 Zn5(OH)6(CO3)2
droxide carbonate

Figure 6.102: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 1000 ppm under CP at 1.15V for 14 days.

Figure 6.103: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 at 500 and 1000 ppm, respectively under
CP at 1.15V for 14 days.

247
6.11.2 SEM /EDX analyses for specimens immersed in 0.5 M NaCl adding
10.0% (w/w) bentonite and Zinc Chloride

As it was mentioned early in this study, the technique was employed to reveal the
chemistry of different compounds that might be formed on the specimens surfaces.
Then a quantitative analysis for those elements, and their progress with time on the
metal surface was also investigated, by using EDAX technique.

Figure 6.104 and 6.105 show SEM images and X- Ray pattern with its chemical
analysis list for corrosion products formed on mild steel surface exposed to 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 at 500 and 1000 ppm at natural corrosion
potential for 14 days. As can be seen, the various deposits obtained on the metal surface
at natural corrosion potential are made of crystals assembled in different morphology.
EDX analysis indicates that the same precipitate contains various elements were
illustrated in the pattern list below for both cases at 500 and 1000 ppm of ZnCl2
concentration respectively. The crystalline material on the both cases was identical in
the SEM/EDAX analysis. However, In the case of XRD analysis the products detected
on the metal surfaces are mostly goethite and no lepidocrocite were detected in the case
of 500 ppm ZnCl2 concentration.

248
Element AN [norm. wt.-%] [norm. at.-%] Error in %
Iron 26 59.23338 36.106106 1.609701
Zinc 30 13.16727 7.023039 0.638325
Chlorine 17 2.78484 2.739629 0.169697
Sodium 11 0.9113105 2.024586 0.043007
Oxygen 8 23.9032 52.10664 4.039511
Sum: 100 100

Figure 6.104: SEM/EDX pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 500 ppm at OCP for 14 days.

249
Element AN Net [norm. wt.-%] [norm. at.-%] Error in %
Iron 26 7481 50.16857 28.3781 1.315055
Zinc 30 1179 14.66886 7.086577 0.564701
Chlorine 17 2100 7.112442 6.337542 0.291507
Carbon 6 570 5.407618 14.22258 7.904269
Sodium 11 103 1.218357 1.674138 0.862051
Oxygen 8 1895 21.42415 42.30106 3.674302
Sum: 100 100

Figure 6.105: X-Ray pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 1000 ppm at OCP for 14 days.

Figure 6.106 to Figure 6.107 show the SEM images and EDAX element analysis for
deposits on steel metal exposed to 0.5 M NaCl contains 10.0% (w/w) bentonite and
ZnCl2, for both cases at 500 and 1000 ppm of ZnCl2 concentration under cathodic
protection (-1.15V) for 14 days respectively. The deposits obtained on the metal surface
are made of crystals assembled in different morphology. EDX analysis indicates that the
products contain calcium, zinc, magnesium, Fe substrate, chloride and oxygen in
different concentrations for both specimens at both concentration of ZnCl2. A weak
reflection shown for the most patterns related the substrate in the form of ferrite (Fe)
particularly.

XRD analysis not shown any crystals related to Mg and Ca products. Obviously, the
films generated on the metal surface for both concentration is compacted and uniform
and it works as an active barrier against diffusion of species into the metal surface.
Hence, it can be considered as the main cause for great drop in current density.

250
Element AN [norm. wt.-%] [norm. at.-%] Error in %
Calcium 20 25.39567 20.12078 0.904343
Chlorine 17 21.91993 19.63275 0.786146
Zinc 30 18.4512 8.959929 1.07156
Iron 26 8.027682 4.564377 0.606141
Magnesium 12 7.795581 10.1846 0.723646
Oxygen 8 18.40992 36.53756 15.37164
Sum: 100 100

Figure 6.106: SEM/EDX pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 500 ppm under CP at 1.15V for 14 days.

251
Element AN [norm. wt.-%] [norm. at.-%] Error in %
Zinc 30 55.67849 25.68673 1.551814
Iron 26 4.178011 2.256848 0.202021
Chlorine 17 2.767978 2.355299 0.152767
Sodium 11 1.344496 1.76424 0.93568
Oxygen 8 36.03103 67.93689 5.263708
Sum: 100 100

Figure 6.107: SEM/EDX pattern for steel specimens after immersed in 0.5 M NaCl
contains 10.0% (w/w) bentonite and ZnCl2 1000 ppm under CP at 1.15V for 14 days

252
Chapter 7

General Discussion

253
7 General Discussion

7.1 Introduction

The overall goal of this research was concerned with understanding the effects of
bentonite on corrosion of external surface of bare mild steel well casings. Corrosion of
casings remains a challenge and a major cause of concern for oil companies. Therefore,
more investigation needs to be carried out. Na-bentonite was used in this research which
mainly used in enormous amount in drilling processes, as it is one of the main recipes of
drilling mud. It may cause severe corrosion for external structures in different location
since starting the drilling process to erect the first pipe. This can be attributed to
spontaneous electrochemical potentials that may be generated due to the presence of
corrosion cells at different location through all depth of casing [39].

As mentioned earlier, corrosion process for mild steel in wet-bentonite environments is


classified to be quite complicated compared to that in solutions. Different experiments
for carbon steel exposed to water-saturated compacted bentonite, under aerobic
conditions were carried out and several mechanisms have been proposed [86, 87].

Casings suffer from corrosion at both sides during drilling process and after completing
drilling process. Severe corrosion might occur unless sufficient protection has been
applied to minimise any unexpected failures. Well casings are usually protected from
external corrosion by applying sufficient CP current. Impressed current cathodic
protection (ICCP) is the only recommended technique to protect casings [11, 42].

Different forms of corrosion might be expected to occur for structures buried or


immersed in water-saturated bentonite environment; including homogeneous and
localized corrosion at different locations [15, 21, 39]. Even though, Hunda with his
work-group in 1997 conclude that no more corrosion process in bentonite takes place
after corrosion product has built up on the metal surface [89]. In contrast, A. Pourbaix
and V. LHostis in his study in 2006 suggested that, uniform corrosion is likely to be the
dominant form of corrosion for mild steel in bentonite [87]. Whereas, others mentioned
that localised corrosion is the dominant if mild steel in bentonite becomes passive [86].
Even though, the real effects of corrosion products on the corrosion process for mild
steel in bentonite are still under debate due to lack information related to the corrosion
process of steel in bentonite [88]. Therefore, further work is required to clarify the

254
different opinions which have been suggested by different researches about the
corrosion process of steel in bentonite. In this research, we have studied the corrosion of
bare mild steel; simulating a well casing exposing to various water-saturated Na-
bentonite environment. Part of our investigation was obviously to investigate bare mild
steel at OCP and under full cathodic protection conditions.

For high accuracy, specimens that were tested in this study were prepared by applying
the same procedure recommended by ASTM standard, G1-03 to minimise the change in
surface finishing conditions [105]. Hence, errors results due to surface finishing can be
eliminated to a certain limit. More details about specimens preparation were illustrated
in Section 5.3 page 81. All the experiments were carried out under aerobic conditions
and ambient temperature. 0.5 M NaCl solution was the main solution added to the
bentonite at different conditions.

A full analysis of the bentonite used in this study was performed for three different
random samples, using SEM/EDX and XRD. Results are illustrated in Appendix I.
Number of chemical species in different ranges was presented in bentonite. These
chemical species were the main source for the deposits which have been formed on the
metal surfaces by applying CP at different levels of protection.

A total of three groups of experiments have been carried out. Details for these groups of
experiments have been already provided in Chapter 5. In brief, the major part of current
research was divided into 3 stages.

The first stage is different levels of constant current cathodic protection (CP) with a
range of 0.0 (Open Circuit Potential) to 200mA.m-2 was applied to protect the simulated
structure buried vertically inside the bentonite layer that contains 45% (w/w) NaCl
Solution at 3.0%. This study was attempted to study the polarisation potential
distribution over depth and time, the ON-OFF potential when CP was applied.

From the first current applied to the system (10.5mA.cm-2), full protection was obtained
for the structure buried in wet-bentonite environment.

In the first stage of this study, the specimens were completely buried inside a bentonite
layer. So, there is no possibility to observe the progress of corrosion process on the
metal surfaces over time unless they have been taken out from the electrochemical cell.
Even though, when CP was applied at different levels to the system at constant current,

255
continuous shift in our polarisation potential over time to negative direction was
observed, after 1 day and 7 days period.
Initially, this shift was expected and it was attributed to the thick layer of damp
bentonite that surrounding the specimens surface in the electrochemical cell. This layer
may act as a barrier against the diffusion of O2 or any other aggressive species. At the
end of the experiment, an interesting thick layer was generated on all specimens
surfaces that had been tested under cathodic protection. This layer was only observed
after the specimens were taken out from the electrochemical cells at the end of the tests.
The characteristics and morphology of this film was analysed using SEM/EDX and
XRD techniques. This work was extended to the next stage of our study.

In this experiment polarisation potential with a range of OCP to -1.15V/SCE was


applied to protect mild steel exposed to 0.5M sodium chloride solution containing
different concentration of bentonite, namely 0.0%, 1.0 and 10.0% (w/w). The progress
of corrosion on the specimens surface was visually observed by looking at them at
regular intervals, photographs were then taken at the end of each experiment, to monitor
the progress of corrosion, The study also involved the evaluation of current drop over
time due to applying constant potential and related that to the results obtained due to the
formation of deposit on the specimens surface. The experiments performed were also
including visual observation, weight loss measurement, LPR. SEM/EDX and XRD
techniques were employed to examine and analyse the films that were composed on the
metal surface. A comparison between the results obtained for this experiment will be
provided later in the general discussion.

The third stage was concerning with ZnCl2 added at 500ppm and 1000ppm to the
bentonite as a cathodic inhibitor to investigate its effects on the corrosion of mild steel.
Polarisation potential at 0 mV (OCP) and -1.15 V against saturated calomel electrode
(SCE) were applied for this experiment.

256
7.2 Specimens tested inside a layer of bentonite, adding 45% (w/w)
NaCl solution at 0.5 M under constant current density from 0
(OCP) to 200mA.m-2

From data given in Figure 6.2, the OCP for 3 pipes buried vertically inside bentonite
showed an identical potential profile over depth. A stable corrosion potential profile
shift to more negative with time was recorded. OCP after 7 days for the same 3 steel
pipes was shifted to more negative potential to settle at ~ -597mV at 0cm depth to ~ -
595mV at 10cm depth for the first pipe, ~ -599mv at depth 12cm to ~ -602mVat depth
22cm for the second pipe and ~-581mV at depth 24cm and 34cm. The values for OCP
over time are given also in Appendix I, table 3.

The general conclusion that can be made is that a negligible change in corrosion
potential values at all depth for steel specimens in bentonite was recorded between day
1 and 7. This behaviour may give an indication that the corrosion products on the metal
surfaces were growing homogeneously with time. This behaviour can be attributed to
the limited amount of oxygen and also to uniform distribution of oxygen in bentonite
which may be consumed by oxygen reduction reaction [4].

According to the visual observation, it is suggested that the corrosion products on the
metal surfaces were uniform due to a sequence of corrosion processes. Details of
corrosion process were explained early in Section 6.2.1. The result is more corrosion
products were built up on the metal surfaces; uniform corrosion is likely to be the
dominant form of corrosion for mild steel in bentonite. This might be due to the limited
amount of oxygen that was trapped in bentonite due to its initial stage of filling [84, 85].
The O2 may gradually consumed by cathodic reduction reaction in the first stage of
metal corrosion (O2 + 2H2 O + 4e 4OH). A spontaneous anodic reaction for iron
was taken place (Fe Fe2+ + 2e). In bentonite, oxygen diffusion became very limited
and the reaction slowed down or ceased. Further oxidation reactions due the existence
of water in the bentonite may caused further corrosion process [87]. The insoluble
products may lead to formation of a protective film [14]. The cathodic reaction is most
likely to be water reduction (2H2O + 2e H2 + 2OH) in the absence of oxygen [87].

A similar insoluble corrosion product was observed on bare specimens in a panel shape
as that on surface of section pipes. Most specimens were covered with insoluble

257
corrosion products in the form of Fe(OH)2. At this stage it is worth mentioning that no
analysis was conducted the corrosion product using SEM and XRD techniques. As
mentioned above in section 7.1. For the rest of experiment the SEM and XRD analysis
were presented in the second and third stage of our research. In brief, the main
crystalline corrosion products detected on the specimens surfaces tested at OCP is
mostly metal substrate in the form of ferrite (Fe). A reflection that might be related to
goethite (-FeOOH) and lepidocrocite (-FeOOH) was also detected.

Based on the results in section 6.2.2, it can be said that there is a significant change in
polarisation potential values between CP ON and OFF, it was noted that the shift was in
a less negative direction when the cathodic protection potential was removed, see
figures 6.3 to 6.7.

However, the difference between the ON and OFF potentials will include errors
associated with the ON potential which may be attributed to one of the following: the
variation of resistivity of the electrolyte and/or to the fact that potential measurements
were made with the CSE reference that may not be in good contact with the damp
bentonite directly over the pipe structure and human error during manual measurements.

The ON potential depends on location of the reference electrode [40] and here space
limitations meant that the ON potential changed both with location and with current in a
way that was proportional to the current density, see Figure 6.8. The results also appear
to confirm the observation that the bentonite tended to dry out slightly over the time of
the experiment. This approach also suggests that the solution resistance in ohm.m
should be proportional to (IR drop / current density) and Figure 6.9 shows that this is
indeed the general trend observed. Further work is needed to investigate whether the
difference between the ON and OFF potentials can be explained by solution resistance
effects.
The decrease in solution resistance with applied current density may be related to
decrease in resistance of the working electrode which could be due to the electro-
osmotic movement of water towards the metal surface, but as the processes relating to
mild steel in damp bentonite clay are much more complicated than that in aqueous
solutions, further work is needed to on this question [86].

258
7.3 Specimens Exposed to 0.5 M Sodium Chloride Solution,
Containing Different Concentration of Bentonite.

The values of CR in mm/y and icorr in A/cm2 in Table 7.1 were calculated, applying the
trapezoidal rule for determining the area under a curve, for 0.5M NaCl solutions free
of bentonite, with 1% and 10% bentonite at OCP (Table 6.3 shows data for solution free
of bentonite, Table 6.5 for solution containing 1% bentonite, and Table 6.6 for solution
containing 10% bentonite).

It is reported by many author [18, 113] that corrosion rates vary over time, and this is
confirmed for the specimens used in this series of tests, at least for the limited time
during which the specimens exposed to the different strength solutions at open circuit
potential, OCP.

Table 7.1 shows that the specimens in solutions containing bentonite showed an average
measured weight loss lower than those specimens in the bentonite-free solution.
However, no significant difference was observed as bentonite concentration increased
from 1% to 10%. The average corrosion rates were 0.214 mm/y, 0.176 mm/yr and 0.173
mm/yr for specimens exposed to 0.5 M NaCl solution containing no bentonite and 1%
and 10% bentonite respectively

The effectiveness of the bentonite in reducing corrosion rates in the weight loss
experiments is explained by the corrosion products layer that remain on the metal
surface of the specimens immersed in 0.5M NaCl solution containing bentonite being
less soluble and better barrier to the diffusion of oxygen to the metal surface than those
obtained in the same solution with no bentonite.

Consistent results were not obtained using the LPR technique, if anything the corrosion
rate increased in the presence of bentonite. Thus the evidence is not definitive on
whether a lower corrosion rate may be attributed to the addition of bentonite.
It was noted that the corrosion rate was higher for those specimens immersed in 0.5 M
NaCl solutions with no bentonite, compared to those where the solutions contained
adding bentonite.

259
Table 7.1 Average values for corrosion rates and corrosion currents for specimens
exposed to 0.5M NaCl solution with different concentrations of bentonite at open circuit
potential (OCP), using trapezoidal rule.

CR/CR
Concentration CR exp. Constant exp.
In 0.5 M NaCl CR icorr mm.y- icorr. exp. Bexp.
solution mm.y-1 (A/cm2) 1
(A/cm2) (V/decade)
0% Bentonite 0.113 9.72E-06 0.177 17.58E-06 0.048 0.64
15.51E-06 0.97
1% bentonite 0.174 13.2E-06 0.179 0.028
10% bentonite 0.120 10.3E-06 0.176 15.2E-06 0.023 0.68

Where icorr is the value of corrosion current and CR is the corrosion rate obtained from LPR
measurements. CR exp., icorr exp. and Bexp are the values of corrosion rate and corrosion current and
proportional constant B that were obtained from the experimental method of weight loss
measurements. Here the value of B used to calculate the corrosion rate obtained from the linear
slopes of Rp curves was 0.026V/decade.

Two methods were used to determine the corrosion rate. The first was the direct method
of weight loss measurement which gives more reliable values for corrosion rate, but has
the practical disadvantage that it takes a long time to produce results as weight loss is a
cumulative process [141]. The second method is the corrosion rate directly obtained
from Linear Polarisation Resistance (LPR) measurement which has the great advantage
that corrosion rates obtained are produced instantly.
The literature consistently reports differences between the LPR and weight loss
measurements. As long ago as 1960, Evans claimed that a factor 2 between the two
methods were common and ascribed these as being due to uncertainties in B values [12].
More recently Law and Carins [142], ], reported that LPR consistently differed from the
weight loss method for corrosion rates of steel in concrete by a factor that ranged
between 0.5 and 2.0, but their findings are vague putting the overestimation down to
unspecified environmental factors. These measurement results are relevant here
because bentonite is a semi-permeable material similar to concrete. As late as 2005
Aquatrac instruments in their online application notes stated [143]: If you have weight
loss coupons in the same stream as the LPR sensor, you should expect the LPR rate to
be within a factor of 2 of the weight loss rate. This correlation between monitoring
methods is applicable for all vendors of LPR.
It is suggested that the differences in LPR and weight loss measurements found in this
work are within the ranges commonly found in practice by other researchers.

260
The corrosion rates in (mm/y) calculated from the overall weight loss assuming an
average (and constant) corrosion rate over the test period. To compare the results of the
weight loss (gravimetric) method with those of the LPR method, the corrosion currents
were calculated at each test point and averaged between every two measuring points.
The average current in Amps was then related to weight loss using Faradaic conversion
and compared to directly measured weight loss. Figure 7.1 for the case of a bentonite-
free solution and Figure 7.2 for the case of a solution containing 10% bentonite, as
examples, show that the measured gravimetric weight loss values were always greater
than those calculated based on LPR measurements. The average values for the
integrated weight loss obtained from LPR and experimental weight loss illustrated in
Figure 7.1 and 7.2 were presented in Appendix VII, Table 44 and 45, respectively.

Figure 7.1: Weight loss: integrated from LPR measurements (g) and gravimetric values
directly measured, for specimens immersed in 0.5 M NaCl solution free of bentonite
exposed for 14 days

261
Figure 7.2 Weight loss: integrated from LPR measurements (g) and gravimetric values
directly measured, for specimens immersed in 0.5 M NaCl solution containing 10%
Bentonite exposed for 14 days.

The results for corrosion rate and corrosion current for specimens exposed to 0.5M
NaCl solution containing bentonite at different concentrations are summarized in Table
7.1. Clearly, a lower corrosion rate was obtained using the LPR measurement than
experimental weight loss measurements. The corrosion rates and corrosion currents as
measured by weight loss are approximately twice that obtained by LPR, particularly for
the 0.5 M NaCl solution free of bentonite.

It is worth noticing that the integrated weight loss values based on LPR tend to become
closer to the gravimetric values as the Bentonite concentration increases. This is likely
to be due to the fact that there may be only a few data points for averaging of current
values. This difference may be due to one or more reasons including; the change of
surface condition, the instability of corrosion products on the metal surfaces, the change
in the character of the corrosive electrolyte, and the drop in the pH of the bulk solution
due to the dissolution of CO2 in solution. Uncertainty in the value of constant B can
also cause a difference. A value of 26mV was used for LPR measurement.
Overestimation of Rp by electrochemical techniques such as LPR leads to an
underestimation of corrosion current and can create the so-called current confinement
effect which can lead to non-uniform distribution of excitation current and
overestimation of the value of Rp.

262
Corrosion potential was more negative in the case of specimens that exposed to the
solutions containing bentonite compared with the one without bentonite. Almost the
same corrosion potential values were recorded with a slight shift to more negative for
solution contains 10.0% bentonite. On the other hand, corrosion potential was more
negative for specimens inside bentonite compared to the others; it was at about -0.797V
after 1 day and -0.805V after 14 days. Similar corrosion potential trend was also
observed for the specimens that were immersed in the solution that contains bentonite.

Furthermore, the results achieved by using XRD and SEM/EDAX techniques were also
used to support the results that were obtained from the other techniques which used in
this study. Goethite (-FeOOH), lepidocrocite (-FeOOH) and magnetite, Fe3O4 were
detected for specimen exposed to 0.5 M NaCl solution without bentonite. On the other
hand no magnetite was detected on the other experiments that were conducted in the
presence of bentonite in the solution. It is believed that, cathodic protection can be used
to protect immersed metallic structures from corrosion. This technique promotes
precipitation of classic calcareous deposits on the metallic surface in seawater, creating
a physical barrier against oxygen diffusion towards immersed metal surface to the
corrosive solution [27, 32]. The result is a decrease in the final current density required
to protect the structure or extension of the life of the sacrificial anode to maintain
sufficient protective cathodic potentials for the structure [32].

In our study, two different polarisation potential was applied at -0.80 mV and -1.15V
against SCE. No corrosion products have been observed while specimens were under
the application of cathodic protection. Under condition of more negative potential, -
1.15V, the system showed a steady fall in current for most tests. Relatively high current
density was recorded for specimens under cathodic protection (-1.15 V) for specimens
in the solution free of bentonite, a current density of 417mA.cm-2 was recorded
compared with the others which contains bentonite. Current density of 225 mA.m-2 and
118.6 mA.m-2 were recorded when specimens were tested at 1% and 10% bentonite
respectively. It is worth mentioning that the values of 417, 225, and 118.6 were
obtained when the specimens were in the solution. When specimens were buried
completely in the bentonite a steady state was reached at the end of the experiment (14
days). The current was dropped for more than 19 times than the initial current and settle
at about 60mA.m-2.

263
In spite of the absence of magnesium and calcium in the chemical analysis of bentonite
solution it is clear from the X-ray analysis that they are present. This is hardly
surprising hence the conversion of the original calcium based bentonite to its soluble
sodium form would not have proceeded to completion. Magnesium is also a common
impurity in bentonite.

XRD analysis was performed in order to assess the relative quantities of compounds
that were generated on specimens surfaces after immersion in solutions containing 1%
and 10% bentonite. Figure 7.3 shows intensity vs. detector angle 2-Theta (2); the
relative quantities of the species obtained from X-ray results are shown in Tables 7.2
and 7.3. The X-ray diffraction quantitative analysis requires an accurate and exact
determination of the diffraction pattern for a sample, both in terms of peak positions
(Degree (2-theta)) and intensities (counts/sec.). The selection of where the peak starts
and ends for each compound will be presented according to the time of immersion.

80000
day 1 da3 day 5
70000 day 7 day 14
In te n sity c o u n t/se c .

60000
50000
40000
30000
20000
10000
0
17.5 18 18.5 19 19.5
Angle degree (2 Theta)

Figure 7.3: A plot of diffraction pattern for brucite reflected intensities versus the
detector angle at 18.5 degree (2), Cu radiation anode. The data was obtained from X-
ray results for specimens after immersion in 10% (w/w) bentonite solution at -1150 mV.

264
25000
day 1 da3
day 5 day 7
20000
In te n sity c o u n t/S e c . day 14

15000

10000

5000

0
28.5 29 29.5 30 30.5
Angle degree (2 Theta)

Figure 7.4: A plot of diffraction pattern for calcite reflected intensities versus the
detector angle at 29.43 degree (2), Cu radiation anode. The data was obtained from X-
ray results for specimens after immersion in 10% (w/w) bentonite solution at -1150 mV.

Table 7.2: Results summarizing the intensity (count/sec.) and ratios of brucite at 18.5
(2), calcite (1) at 29.43 (2), calcite (2) at 50.85 (2) and ferrite at 44.65 (2). The
data was obtained from X-ray data for specimens after immersion in solutions
containing 1% (w/w) bentonite solution at -1150 mV.
Time Intensity, Count/Sec. Ratio
(day) brucite calcite 1 brucite : calcite 1
3 day 27000 200 135
5 days 5500 1000 5.5
7 days 7000 800 8.75
14 days 25000 8200 3.049
brucite calcite 2 brucite:calcite 2
3 day 27000 3000 9
5 days 5500 1800 3.056
7 days 7000 1900 3.684
14 days 25000 3200 7.813
brucite ferrite brucite : ferrite
3 day 25000 11800 2.119
5 days 5500 6400 0.859
7 days 7000 1200 5.833
14 days 27000 - 8.471

265
The variation in intensity over time obtained for brucite confirms the continuous
formation of those compounds on the metal surface due to applying cathodic protection.
The intensity of brucite at 18.5 (2) was much higher than the most other species
intensity at different peak reflections. The relative quantity of calcite 1, 2 and ferrite
with respect to the brucite at 18.5 degree 2 theta (using the integral intensity of
diffraction peaks in the X-ray diffraction patterns) at different time was calculated and
their intensities were less than the brucite located at 18.5 (2).

Table 7.3: Results summarising the intensity (counts/sec.) and ratios of brucite at 18.5
(2) to calcite (1) at 29.43 (2), calcite (2) at 50.85 2 and ferrite at 44.65 (2)
respectively. The data was obtained from X-ray results for specimens after immersion in
10% (w/w) bentonite solution at -1150 mV.

Time Intensity (counts/sec.) Ratio


(day) brucite calcite 1 brucite : calcite 1
1 day 13000 12000 1.083
3 days 23000 13000 1.769
5 days 28000 14000 2
7 days 9600 7600 1.263
14 days 72000 21000 3.429
brucite calcite 2 brucite : calcite 2
1 day 13000 5000 2.6
3 days 23000 10000 2.3
5 days 28000 13000 2.154
7 days 9600 3600 2.667
14 days 72000 22000 3.273
brucite ferrite brucite : ferrite
1 day 13000 - 2.6
3 days 23000 8000 2.875
5 days 28000 8500 3.294
7 days 9600 750 12.8
14 days 72000 8500 8.471

Calcareous deposits in sea water have been well researched for equipments and
structures mounted in the sea [22, 29, 30]. In brief, the film consists of a white/grey
scale mixture of calcium carbonate (CaCO3) in the form of calcite and aragonite and
magnesium hydroxide (Mg(HO)2) in the form of brucite. The film can be created on the
metal surfaces by applying cathodic protection [27, 28]. In sea water, Mg(OH)2 is
recognized to be under-saturated condition and tends to precipitate at quite high pH

266
values, 9.5 [31]. So, the precipitation of any protective layer will depend on the local
pH, at the metal/electrolyte interface, of the protected metal and the concentration of
different cations in the solution containing bentonite. The film can be generated rapidly
by applying a relatively high initial current to the system under cathodic protection. The
specifications of that film depend on their chemical composition. It is known to work as
a physical barrier and have the ability to reduce the oxygen approaching the metal
surface [27, 32]. Hence, the final current density applied to protect the structure can be
reduced [1, 33]. The general concepts for calcareous film formation in sea water were
explained in greater detail in Chapter 2, Section 2.5.2.

In the current case, the precipitations observed on the specimens surface that under
cathodic protection after immersion in bentonite solutions can be compared with
calcareous films created in seawater. The results obtained from the XRD for the
specimens that under cathodic protection showed different precipitations deposited on
the metal surfaces. The XRD results indicate similar reactions that have been occurred
that led to create almost same compounds on the metal surface. The resulted compounds
include Mg(HO)2 in the form of brucite and CaCO3 in the form of calcite. The brucite
reflection was gradually increased with time. The continuous formation of brucite can
be ascribed to the local increase in pH at metal/ electrolyte interface caused by the
production of hydroxyl ions (OH ) due to applying CP. Only small amount of calcium
carbonate in the form of calcite was precipitated on the specimens surface. The limited
amount of calcite that was detected might be due to the low concentration of carbon
dioxide (CO2) in the solution. Hence, the reaction of CO2 with water will be slow and
any subsequent reactions that generate direct precipitation of calcium carbonate (CaCO3)
onto the metal surface, as a protective layer against corrosion, will be correspondingly
slow [26].
It is confirmed that the time was an essential factor for specification and quality of any
film growing on the metal surface. The system showed a steady fall in current for most
tests and the final applied current density was decreased. It can be concluded that a
classic calcareous film is formed (CaCO3, Mg(OH)2) [44].

According to Barchiche, and Deslouis [28], Calcium carbonate in aragonite form can be
created on a metal surface immersed in seawater when cathodic potential in the range
of -900mV to -1100mV/SCE is applied. In the current study, the aragonite was not
detected. The absence of aragonite is difficult to explain and needs further investigation.

267
It is clear from the results that the lack of difference in behaviour observed for 1% and
10% bentonite solutions can not be interpreted by the concentration difference, but by
the interfacial pH which is governed by oxygen concentration in solution which should
not be affected by the bentonite concentration.

The third stage was concerning with the ZnCl2 added at 500ppm and 1000ppm to the
bentonite to investigate its effects on the corrosion process of mild steel. For this
experiment, specimens were exposed to the electrolyte at natural corrosion potential
(OCP) and at -1.15 V polarisations potential against saturated calomel electrode (SCE).

Considerable current drop was observed when ZnCl2 was used; the possible
interpretation for this drop is the formation of different types of compounds containing
zinc, the formation of such compounds were confirmed by SEM and XRD techniques
(see the result section). Because of the limited time of this research the corrosion rate
was not measured, this part can be investigated as part of any future work that may be
conducted.

268
8 Conclusions
Full protection was obtained for the specimens buried in wet-bentonite
environment even when current density at 10.5mAm-2 was applied to the system.
This current was the minimum current density applied to the system.

Both ON and OFF potential values were determined. High values for ON
potential was always recorded compared to the instant OFF potential values.

The weight loss measurements were also utilized for days 1, 3, 5, 7 and 14, and
then corrosion rate was calculated. low corrosion rate was recorded for
specimens exposed to solution saturated with bentonite

The corrosion rate and corrosion current obtained from LPR measurement were
lower than the corrosion rate and corrosion current which obtained using weight
loss measurement in all experiment conducted in this research.

Different corrosion products were detected by EDX and the same results was
confirmed by XRD analysis.

Considerable drop in current density has been recorded due to the formation of
film on the metal.

It was confirmed that, the time was an essential factor for the specification and
quality of any film that may compose on the metal surface.

It is possible for any deposit to form on the metal surface while it is exposed to
the electrolyte solution, but in order to create an active and uniform protective
film, cathodic protection at precise potential must be applied.

Almost similar products were detected by SEM/EDX and confirmed by XRD


analysis for specimens tested under the application of cathodic protection at high
potential typically -1.15V.

No precipitates were observed on the specimens that immersed in the NaCl


solution with no bentonite.

When the bentonite was added to the solution the XRD results were almost the
same for different immersion time which indicates that the Mg(OH)2 in the form
of brucite is the main crystalline layer which was detected.

269
Calcium carbonate in the form of calcite was also detected in a limited amount.

Sodium chloride in the form of halite was also detected on the surface.

The results obtained by SEM/EDX showed different crystals on the specimens


surface which in agreement with the results that detected by XRD technique.

270
9 Future Work

It is suggested to repeat the same experiment conducted in this research using


different polarisation potential for long period of study focusing on the
following:
 The appearance and characterisation of deposit generated on the metal
surface.
 The current drop with time to confirm if it is reached a steady state with
time.
 The potential depolarisation with time when protected system at OFF
potential.
 The characterisation of the film after depolarisation.
When ZnCl2 is used it is suggested to apply potential in the range of -0.800mV
to -1000mV vs. SCE to avoid hydrogen evolution. Then, SEM/EDX and XRD
analysis might be applied to study the appearance, surface morphology and
crystal compounds for the new generated film due to presence of ZnCl2 in the
solution saturated with bentonite.

It is suggested to measure the corrosion rate using weight loss measurement


when ZnCl2 was added to the solutions.

Further investigation needs to be carried out under anaerobic conditions.

271
10 References
1. NACE Standard RP0176-2003, Item No. 21018, Standard Recomended Practice
"Corrosion Control of Steel Fixed Offshore Structures Associated with
Petroleum Production".
2. A. W. Peabody, Control of Pipeline Corrosion. 2nd ed., ed. R.L.Bianchetti.
NACE International: Houston, Texas,USA. 2001.
3. Mars G. Fontana and Norbert D. Greene, Corrosion Engineering , Second
Edition, seies in Material Science and Engineering Series, McGrow-Hill, USA.
1978.
4. D. A. Jones, Principles and Prevention of Corrosion, Second Edition,
Department of Chemical and Metallurgical Engineering, Prentice Hall, NY.
1996.
5. www.chemia.odlew.agh.edu.plcho_imir...lecture7_bez_anim.pps, AGH-
University os Science and Tech.
6. Mars G. Fontana, Corrosion Engineering. 3rd ed. : McGraw- Hill,1986
7. J.C. Scully, The Fundamentals of Corrosion. 3rd ed. : Oxford, UK. book, 1999.
8. Chemical and process Technology.
http://webwormcpt.blogspot.com/2009/01/crevice-corrosion-mechanism
prevention.html. 2009.
9. D.J. Mills and R.A. Akid, comparison between conventional macroscopic and
novel microscopic scanning electrochemical methods to evaluate galvanic
corrosion. Corrosion Science, 2001. 43(7): p. 1203-1216.
10. J. L. Luo, C.J.L., Q. Yang, and S.W. Guan, Progress in Organic Coating 31.
Science direct, 1997. 31: p. 289-293.
11. Ana Osella and Alicia Favetto, Effects of soil resistivity on currents induced on
pipelines. Journal of Applied Geophysics, 2000. 44(4): p. 303-312.
12. Evans Ulick Richardson, F. R. S., and F. I. M., The corrosion and oxidation of
metals: scientific principles and practical, London 1960.
13. U. R. Evans, Metallic Corrosion, Passivity and Protection, London. 1946.
14. The US department of defense, Cathodic protection military handbook, Mil-
HDBK-1004/10. 2003.
15. W.Von Baeckmann, W. Schwenk, and W. Prinz, Handbook of Cathodic
Corrosion protection, Gulf Publishing Company, Houston, TX. USA. 1997.
16. Florian Mansfeld, Chemical industries/28. Corrosion mechanisms Department of
materials science, University of Southern California Los Angeles, California.
1986.
17. Samuel A. Bradford, corrosion control, Second Edition, University of
Alberta.CASTI, Canada. 2001.
18. Nestor Perez, Department of mechanical engineering University of Puerto Rico,
Electrochemistry and corrosion science, Klower acadimic Puplishers, New york.
2004.
19. ASTM, G 102 89, Standard Practice for Calculation of Corrosion Rates and
Related Information from Electrochemical Measurements.
20. L.L.Shreir, R. A. Jarman, and G. T. Burstein, Corrosion, (Metal/Environment
Reactions). 3rd ed., Oxford,UK: Butterworth-Heinemann. 1994.
21. R. G. J. Edyvean , A.D.M. and N.J.S. C. J. Hutchinson , L. V. Evans,,
Interactions between cathodic protection and bacterial settlement on steel in

272
seawater. International Biodeterioration & Biodegradation, 1992. 29(3-4): p.
251-271.
22. Leeds, S.S., Influence of Surface Films on Cathodic Protection, A thesis
submitted to The University of Manchester for the degree of Doctor of
Philosophy (PhD) in the Faculty of Engineering and Physical Science,Corrosion
and Protection Centre School of. 2007.
23. Abdel Salam Hamdy, A.G.S.e. M.A. Shoeib, and Y. Barakat, The
Electrochemical Behavior of Mild Steel in Sulfide Polluted NaCl at Different
Velocities. Int. J. Electrochem. Sci., 2008. 3 p. 1142 - 1148.
24. Yahalom J., K.H.J.B., Robert, W. Cahn, Merton, C. Flemings, Bernard, Ilschner,
Edward, J. Kramer,Subhash, Mahajan and V. Patrick, Corrosion Protection
Methods, in Encyclopedia of Materials: Science and Technology. 2001, Elsevier:
Oxford. p. 1710-1713.
25. T. Misawa, T.K., W. Sutaka and S. Shimodaira, The mechanism of atmospheric
rusting and the effect of Cu and P on the rust formation of low alloy steels(28c).
Corrosion Science, 1971. 11: p. 35-48.
26. C. P. Dillon, Corrosion control in the chemical process industries, Second
edition , Material Technology institute of the Chemical Process, Inc. 1994.
27. Chems Barchiche, C.D., Otavio Gil, Philippe Refait, and B. Tribollet,
Characterisation of calcareous deposits by electrochemical methods: role of
sulphates, calcium concentration and temperature,Electrochimica
Acta,2004,49:p.2833-2839
28. Ch. Barchiche, C.D., D. Festy, O. Gil, Ph. Refait, S. Touzain, and B. Tribollet,
Characterization of calcareous deposits in artificial seawater byimpedance
techniques3*/Deposit of CaCO3 in the presence of Mg(II),Electrochmica
Acta,2003, 48:p.1645-1654
29. H. Cachet , T.E.M., D. Herbert-Guillou ,, D.F. b, S. Touzain , and B. Tribollet,
Characterization of deposits by direct observation and by electrochemical
methods on a conductive transparent electrode. Application to biofilm and scale
deposit under cathodic protection, Electrochimica Acta,46:2001, p.3851-3857
30. Claude Deslouis and B. Tribollet, Recent developments in the electro-
hydrodynamic (EHD) impedance technique,Electrochimical Chemistry, 2004
572:p.589-598
31. S.Elbeik, A. C. C. Tseung, and A. L. Mackay, The formation of calcareous
deposits during the corrosion of mild steel in sea water. Corrosion Science, 1986.
26 (9), P. 669-680
32. H. Mller, E.T. Boshoff, and H. Froneman, The corrosion behaviour of a low
carbon steel in natural and synthetic seawaters,The Journal Of South
Africa,2006, 106:p.585-592
33. Claud Deslouis, P.F., M.J. O. Gil, V. Maillot,V. Maillot, , and B. Tribollet,
Influence of clay on calcareous deposit in natural and artificial sea water,
Electrochimica Acta. 2006.51: p.3173-3180
34. H. Mller, The influence of Mg2+on the formation of calcareous deposits on a
freely corroding low carbon steel in seawater. Corrosion
Science.2007.49:p.1992-2001.

273
35. K. E. Mantel, W.H.H. and T.Y. Chen, Substrate, Surface Finish, and Flow Rate
Influences on Calcareous Deposit Structure,Paper Number 92060489 corrosion,
1992. 48.
36. Well casing cathodic protection, Prolonging metal
life,research.nigc.ir/files/articles/protection/TP-LR-017. 2003.
37. Jerry Bauman, Well casing cathodic protection utilizing puls current technology,
Paper No. 04049. Corrosion 2004.
38. Jafar Abdollahi, Analysing Complex Oil Well Problems through Case-Based
Reasoning, Doctoral Thesis for the degree of Philosophiae Doctor, Norwegian
University of Science and Technology Faculty of Engineering Science and
Technology Department of Petroleum Engineering and Applied Geophysics.
Journal of applied electrochemistry, 2007. 15.
39. Leendert de Witte, Analysis of the principal component of external casing
corrosion in deep wells. Journal of Applied electrochimistry 1984. 15: p. 325-
334.
40. UFC, Unified facilities criteria. Operation and maintenance: Cathodic protection
systems. UFC 3-570-06. JANUARY 31 2003.
41. Wikipedia on-lie reference on geological formations.
42. I.A. Metwally and A. Al-Badi, Factors affecting pulsed-cathodic protection
effectiveness for deep well casings Aiti-corrosion Methods and Materials, 1996.
56(4): p. 196.
43. NACE Standard RP0186-2001, Item No 21031, Standard Recommended
Practice "Application of Cathodic Protection for External Surfaces of Steel Well
casting".
44. Y. F. Yang, J. D. Scantlebury, and Elena Koroleva, Underprotection of mild
steel in seawater and the role of the calcareous film, NACE 2009, Atlanta. 2009.
45. Saibal Mitracorrpro Companies INC, Basics of corrosion control and well casing
protection presinted. 13011 Florence Ave Santa Fe Springs, CA 90670562-254-
6205smitra@corrpro.com.
46. H. James and H.B. Caproco International Ltd, Well casing cathodic protection,
UK(12).
47. Brenda J. Little and P.A. Wagner, The interrelation between marine biofouling
and cathodic protection. NACE 93, P. 525.
48. L. L. Shreir, Corrosion,Vol. 2, Corrosion Control, PhD, FRIC, FIM, FICorrT,
FIMF. Head of Department of Metallurgy and Materials, City of London
Polytechnic, printed in London by Newnes Butterworths group. 1976.
49. S. H. Lee, D. W. Townley, and K. O. Eshun, A boundary Element Model of
Cathodic Well Casing Protection Journal of Computational Physics, 1993. 107:
p. 338-347.
50. P. Eng. Bauman and CIMARONN Integrity Ltd., Well casing cathodic
protection, Utilizing Pulse Current Technology, Calgary, Alberta, Canada,
NACE 2004, Paper No. 04049.
51. Melvin Romanoff, Underground Corrosion, Chapter 3, Characteristics of Soil,
pp3, first edition,1957.
52. E. Escalante, Corrosion Tests and Standards ,Chapter 14, Soils, First Edition,
p.137, 1995.
53. J. Morgen, Cathodic protection of iron and steel, sacrificial anodes, [second
edition], p. 113. 1993.
54. V. Chaker, Corrosion Tests and Standards, Soils, First Edition 1995.

274
55. J. D. A. Miller and A. K. Tiller, Microbial Aspects of Metallurgy, , New York, .
American Elsevier, 1970: p. 61.
56. J.A. Von Fraunhofer, Concise Corrosion science, First Edition 1974.
57. X. H. Nie and co-reseachers, Raman spectroscopy. Technology and Business
Journal via NewsRx.com 17 March 2009: p. 176.
58. Costa Wranglen, Introduction to corrosion and protection of metals, Published
by Chapman and Hallin, chapter 10. 1985.
59. C. Den Hartog and R. Leiden, Brackish-water classification, its development
and problems Aquatic Ecology, 1974. 8(1-2): p. 15-28.
60. JC Rowlands in Shreir Vol 1; 3rd edition; Shreir Jarman and Burstein, Corrosion,
(Metal/Environment Reactions). 3rd ed., Oxford,UK: Butterworth-
Heinemann,1994.
61. B. J. Little, R. I. Ray, and R. K. Pope, The relation between corrosion and the
biological sulfur cycle, Paper No. 00394 Corrosion 2000.
62. A. W. Peabody, Control of pipeline corrosion, Survey method and techniques,
first edition 1967.
63. ASTM Standard, G 57 95a (Reapproved 2001), Standard Test Method for
Field Measurement of Soil Resistivity Using the Wenner Four-Electrode
Method1.
64. TM 5-811-7 Headquarters Department of the Army Washington, DC.22 April
1985, Manual Suersedes Section VII, VIII and IX.
65. ASTM standard, G 57 95a (Reapproved 2001) Standard Test Method for Field
Measurement of Soil Resistivity Using the Wenner Four-Electrode Method1.
66. Fraser King, G.V.B., N.T. Kurt Lawson, and R.R. P. Neckols, Cathodic
Protection of Pipelines in High Resistivity Soils and the Effect of Seasonal
Effect, NACE Corrosion, Paper No. 06163, p. 2. 2006.
67. C. Rousseau, F.B., L. Leleyter, and O. Gil, Cathodic protection by zinc
sacrificial anodes: Impact on marine sediment metallic contamination. Journal of
Hazardous Materials, 2009. 167: p. 953-958.
68. Serdar Gundogdu, Ozge Sahin, and E.M.I, Effects of Cathodic Protection on
Electromagnetic Flowmeters, http://www.mdpi.org/sensors. sensors, 2007.
7(1424-8220): p. 75-83.
69. B. A. Martin, Telluric effects on a buried pipeline NACE Corrosion, 1992. 49.
70. ASM Handbook Committee, Corrosion: Fundamentals, Testing, and Protection.
2003. 13A.
71. J. D. Scantlebury, Corrosion control engineering MSc. course- unit 6 lecture
notes, The University of Manchester. 2006.
72. R. L. Kean and K.G. Davis, Cathodic Protection, New version of DTI
publication first issued in (1981). 2004.
73. M. M. Ben Haji, Sacrificial Anode for Cathodic Protection In Sea Water, A
Thesis Presented to the Victoria University of Manchester, Degree of Doctor of
Philosophy, October. 1978.
74. John Dabkowski, Assessing The Cathodic Protection Levels Of Well Casings,
Contract PR-151-106, Prepared for the Pipeline Reseach Supervisory Committee,
Pipeline Research Committee of Pipeline Research Council International, Inc.
1983.
75. N. N. Glazov, I.I.R. S. M. U. , and I .I. A. A. N. Podobaev, Soil Corrosion of
Differential Aeration Cells and Conditions of Their Operation. Protection of
Metals. 2006. 42: p. 238243.

275
76. Schlumberger, CPET Corrosion Protection Evaluation Tool, Produced by
Marketing Communications, Houston. 2004.
77. NACE Standard RP0169-2002, Item No. 21001, Standard Recommended
Practice,Control of External Corrosion on Underground or Submerged Metallic
Piping Systems.
78. William J. Miles, Miles Industrial mineral research, Denver, CO. Bentonite
commodity markets from 1990 and future trends.
79. Michael H. Bradbury and B. Baeyens, Porewater chemistry in compacted re-
saturated MX-80 bentonite. Journal of Contaminant Hydrology, 2003. 61: p.
329 338.
80. M. Jansson, Diffusion of Radionuclides in Bentonite Clay- Laboratory and in
situ Studies Mats Jansson Doctoral Thesis, Department of ChemistryNuclear
ChemistryRoyal Institute of TechnologyStockholm, Sweden. 2002. p. 5.
81. Naim Muhammad, Hydraulic, Diffusion, and Retention Characteristics of
Inorganic Chemicals in Bentonite. A dissertation submitted in partial fulfillment
of the requirement for the degree of Doctor of Philosophy in Civil Engineering
Department of Civil and Environmental Engineering, College of Engineering
University of South Florida,2004.
82. A. Besq, C.M., A. Pantet, P. Monnet, and D. Righi, Physicochemical
characterisation and flow properties of some bentonite muds(16B). Applied Clay
Science, 2003. 23(5-6): p. 275-286.
83. R and M Energy Systems, Technical Bulletin -T0103: Drilling Mud pH and
Effect on Elastomers. WWW.rmenerg.com. 2003.
84. Birgersson Martin and K. Ola, Ion equilibrium between montmorillonite
interlayer space and an external solution--Consequences for diffusional transport.
Geochimica et Cosmochimica Acta, 2009. 73(7): p. 1908-1923.
85. M.V. Villar a, M. Snchez , and A. Gens, Behaviour of a bentonite barrier in the
laboratory: Experimental results up to 8 years and numerical simulation
Physics and Chemistry of the Earth, Parts A/B/C, 2008: p. s447-s485.
86. Gen Nakayama and Masatsune Akashi, The critical condition for the initiation
of localised corrosion of mild steel used for nuclear waste package, Materials
Research Society, 1991.
87. A. Pourbaix and V. LHostis, Passivation, localised corrosion and general
corrosion of steel in concrete and bentonite. Theory and experimentals,J. Phys.
IV France,136, p: 71-78, 2006.
88. X. Xiaa, K.I., T. Arima, Y. Inagaki, T. Ishidera, S. Kurosawa, K. Iijima, H.
Sato,, Corrosion of carbon steel in compacted bentonite and its effect on
neptunium diffusion under reducing condition, Applied Caly Science, 28, p: 89-
100, 2005.
89. Honda A., T.N., Ishikawa H., Fujiwara K., Experimental research on corrosion
behavior of carbon steel in anaerobic condition. PNC Tech. Rev. 104, P: 125
134. Japanese, 1997.
90. N. Yildiz, A.C. and Y. Sarikaya, The characterization of Na2CO3 Activated
Kutahya Bentonite. Turk J Chem, 1999 23: p. 309-317.
91. D. J. DAppolonia, Soil-Bentonite Slurry Trench Cutoffs.Geotechnical
Engineering.1980.106.p.GT4.

276
92. M. Garca-Gutierrez, T.M. and J.S. M. Mingarro, Z. Dai, J. Molinero, Solute
transport properties of compacted Ca-bentonite used in FEBEX project.Journal
of Contaminant Hydrology.2001.47.p.127-137.
93. F. M. I. Hunter, F.B., T. G. Heath, A. Hoch and L. O. Werme, Investigation of
iron transport into bentonite from anaerobically corroding steel: A geochemical
modeling study. Clays in natural and engineerd barriers for radioactive waste
confinement, 2007.
94. Anh Minh Tang, Y.J.C. and Trung Tinh Le, A study on the thermal conductivity
of compacted bentonites. Applied Clay Science 2008. 41: p. 181-189.
95. S. K. Kawatra and S.J. Ripke, Developing and understanding the bentonite fiber
bonding mechanism Minerals Engineering, 2001 14(6): p. 647-659.
96. Abdullah A. Al-shahrani, Layered Silicate Nanocomposites. A thesis submitted
to the university of Manchester for the degree of Doctor of Philosophy in the
Faculty of Science and Engineering.Supervisor: Professor David
Scantlebury.School of MaterialsCorrosion and Protection Centre,2008.
97. Suprakas Sinha Ray and Masami Okamoto, Polymer/layered silicate
nanocomposites: a review from preparation to processing, Progress in Polymer
Science, www.Sciencedirect.com. 2003. 28: p. 1539-1641.
98. Sandra Garc-Garci, S.W. Claude Degueldre, Sabrina Frick, and M.A. Glaus,
Determining pseudo-equilibrium of montmorillonite colloids in generation and
sedimentation experiments as a function of ionic strength, cationic form and
altitude. Journal of Colloid and Interface Science, 2009.
99. Villar M. Victoria and A. Lloret, Influence of dry density and water content on
the swelling of a compacted bentonite. Applied Clay Science 2008. 39: p. 38-49.
100. A. K. Ashmawy, N.S. D. El-Hajji, and N. Muhammad, Hydraulic performance
of untreated and polymer-treated bentonite in inorganic landfill leachates. Clays
and Clay Minerals, 2002. Volume 50(7): p. 546-552.
101. G. Montes-H, J. Duplay, L. Martinez, Y. Geraud, and B. Rousset-Tournier,
Influence of interlayer cations on the water sorption and swelling-shrinkage of
MX80 bentonite. Applied Clay Science, 2003. 23(5-6): p. 309-321.
102. L. M. McDonald, W.V.M., V. P. Evangelouy, and M. A. Chappell, CATION
EXCHANGE. Elsevier Ltd., 2005.
103. Omar Almisned, Assessing the corrosivity of drilling mud on well casings, King
Abdulaziz City for Science and Technology, Riyadh, Saudi Arabia,Emerald
Group Publishing Limited Anti-Corrosion methods and materials, 2008. 55: p.
278282.
104. ASTM standard, D2216-98, Standard Test Method for Field Measurement of
water content in Soil.
105. ASTM Standared, G1-03 Standard Practice for Preparing, Cleaning, and
Evaluating Corrosion Test Specimens.
106. F. M. Song, S. Brossia, and N.S. D. Dunn, A New Permanent Reference
Electrode for Field Use with Cathodic Protection in Protection of Underground
Pipelines and Storage Tanks, NACE, Paper No. 05040. Corrosion, 2005.
107. ASTM Standard, G 8 96 (Reapproved 2003), Standard Test Methods for
Cathodic Disbonding of Pipeline Coatings.
108. Neil G. Thompson, Joe H. Payer, and B.C. Syrett, DC Electrochemical Test
Methods by NACE International. 1988.
109. J. D. Scantlebury and L.K. Xu, A study on the deactivation of an IrO2-Ta2O5
coated titanium anode. Corrosion Science,. 2003. 45(12): p. 2729-2740.

277
110. John H. Berry, Drilling fluid properties and functions,
Hydrogeologist/Geologist, Foundation Division Manager, CETCO Drilling
Products.
111. ASTM standard,G 5 94 (Reapproved 1999) Standard Reference Test Method
for Making Potentiostatic and Potentiodynamic Anodic Polarization
Measurements.
112. N. Taniguchi, A.H. and H. Ishikawa, Experimental investigation of passivation
behavior and corrosion rate of carbon steel in compacted bentonite.Power
Reactor and Nuclear Fuel Development Corporation, Tokai-mura, Ibaraki 319-
11, JAPAN.
113. ASTM Standard, G 31 72, Practice for Laboratory Immersion Corrosion
Testing of Metals, (Reapproved 2004).
114. Gardner S. Haynes and R. Baboian, Laboratory corrosion tests and standards. p.
507.
115. NACE, NACE Corrosion Engineer's Reference Book, Third Edition, Library of
Congress,Editor: Robert Baboian, R. S. Trseder. 2002.
116. J. D. Scantlebury and F.H. Karman, Potential-time measurements on mild steel
in etch primer solutions. Corrosion Science, 1993. 35(5-8): p. 1305-1309.
117. Henry Leidheiser and JR, Electrical and electrochemical measurements as
predictors of corrosion at the metal.Organic coating interface,Progress in
Organic Coatings. 7: p 79-104, 1979.
118. G. P. Turmov, A.N.M., E. N. Minaev, and E.A. Petrov, Corrosion control
Investigation if cathodic protection is used, Paper No. 11. The NACE Annual
Conference and corrosion show, 1993.
119. ASTM standard, G 59 97(Reapproved 2003), Standard Test Method for
Conducting Potentiodynamic Polarization Resistance Measurements.
120. F. Mansfield, Electrochemical techniques for corrosion, R. Baboian, editor.
NACE, 1977: p. 18 - 26.
121. Wanklyn, J.N., Input data for modelling marine cathodic protection. Cathodic
protection: Theory and practice by V. Ashworth and C. J. L. Booker, London,
1986.
122. Davey, W.P., A new X-Ray Diffraction Apparatuse'' . J. Op. Soc. Am. s. pp.
479-492. 1921.
123. F. H. Chung and D.K. Smith, Industrial Applications of X-Ray Diffraction''
Marcel Dekker Pub. pp1006. 1999
124. B. E. Warren, X-Ray Diffraction'' (2nd Edition). Dover Publication.U.S.A.
pp381. 1990.
125. V. K. Pecharsky and P.Y. Zavalij, ''Fundamentals of powder Diffraction and
structural characterisation of Materials'' Springer Verlag. pp. 713. 2005.
126. C. Suryanarayana and M.G. Norton, ''X-Ray Diffraction: A Practical Approach''.
Plenham Press. New York. pp. 273. 1998.
127. J. P. Glusker and K.N. Trueblood, Crystal Structure Analysis: A Primer, Oxford
University Press. 1972.
128. www.fbi.gov/about-us/lab/forensic-science-
communications/fsc/april2007/research/2007_04_research03_figure01.htm.
129. http://www.britannica.com/EBchecked/topic/76973/Bragg-law.
130. TNRCC Regulatory Guidance Municipal Solid Waste Division, Soil-Bentonite-
Slurry Trench Cutoffs in Solid Waste Land Disposal Site Application. April
1997: p. RG-282.

278
131. W. Schwenk, Adhesion loss of organic coatings causes and consequences of
corrosion protection: in Corrosion Control by Organic Coatings, ed. H.
Leidheiser. 1981, NACE Publication: Houston, Texas, USA. 103-110.
132. Shozo Sekioka, M.I.L., Maria P. Philippakou, and John M. Prousalidis, Member
IEEE, , Current-dependent grounding resistance model based on energy balance
of soil ionization. EEE transanction on power delivery 2006. 21(1): p. 202-209.
133. Wankalyn, J.N., Input data for modilling marine cathodic protection. Cathodic
protection: Theory and practice by V. Ashworth and C. J. L. Bookker, London,
1986.
134. Beavers, J.A. and Kevin C. Garrity, Criteria for Cathodic Protection, Control of
Pipeline Corrosion by A. W. Peabody, ed. R.L.Bianchetti. NACE International:
Houston, Texas,USA. 2001.
135. Sudhakar P.V. Mahajanam, Application of hydrtalcites as corrosion -inhibiting
pigment in organic coatings, the Degree Doctor of Philosophy in the Graduate
School of The Ohio State University. 2005.
136. Claud Deslouis , P.F., O. Gil, , V.M. M. Jeannin , and B. Tribollet, Influence of
clay on calcareous deposit in natural and artificial sea water, Electrochimica
Acta. 2006.51: p.3173-3180.
137. J. D.Scantlebury, The dynamic nature of underfilm corrosion. Corrosion Science,
1993. 35(5-8): p. 1363-1366.
138. Klinghofer O., Riislund E., Frolund T., Elsener B., Schiegg Y., Bohni H.:
Assessment of Reinforcement Corrosion by Galvanostatic Pulse Technique,
Proceedings International conference on Repair of Concrete Structures, Norway,
1997, pp 391-400.
139. James B. Bushman and I.V.V.V.H.U.A.C. Dr. Bopinder S. Phull Bushman and
Associates, Corrosion Rate Measurement and Evaluation of Pitting Corrosion in
Water Distribution Piping Systems and Water Storage Tanks.
140. A. Pardo, S.F.J., 2 M. C.Merino,1 R. Arrabal,1 and E.Matykina3,
Electrochemical Estimation of the Corrosion Rate of Magnesium/Aluminium
Aloys. International Journal of Corrosion. 2010: p. 8.
141. Yunny Meas and Jorge Fujioka, The determination of the corrosion rate in
cathodically protected systems. Corrosion Science, 1990. 30(8-9): p. 929-940.
142. D.W. Law and J.J. Cairns, Non-Destructive Testing in Civil Engineering
International Symposium NDT-CE 2003.
143. http://www.aquatrac.com/literature/multiflex/AN_CorrosionRate. pdf.

279
11 Appendices
The Appendices associated to the previous work has been conducted throughout this
research consist of the following:

11.1 Appendix I: Bare mild steel buried in bentonite containing 45%


(w/w) 0.5 M NaCl Solution.

Table 1: Resistivity of bentonite with different water content.

Test number Water content (% Wt.) The average resistivity .cm


1 15 27294
2 20 4850
3 25 1450
4 30 1142
5 35 1044
6 40 766
7 45 440
8 50 247
9 60 185
10 80 172

28000
27294
26000

24000

22000

20000
Resistivity in .cm

18000

16000

14000

12000

10000

8000

6000
4850
4000
1450 1142 1044
2000 766 440 247 172
185
0
0% 10% 20% 30% 40% 50% 60% 70% 80% 90%
water content wt.%

Figure 1: Resistivity of bentonite at different water content in .cm.

280
5
4.5
4
log. resistivity in ohm .cm
3.5
3
2.5
2
1.5
1
0.5
0
0% 10% 20% 30% 40% 50% 60% 70% 80% 90%
water content wt.%

Figure 2: Resistivity of bentonite with water content using a logarithmic scale.

281
Figure 3: EDX analysis of bentonite used in this study (sample 1).

Ref. Code Score Compound Displacem- Scale Chemical Formula


Name ent [2Th.] Factor

282
000120204 75 Montmor- -0.160 0.509 Nax(Al, Mg)2Si4
illonite O10(O H)2z H2O
000471743 58 Calcite -0.160 0.076 CaCO3
010820511 62 Quartz, syn -0.160 0.202 SiO2
000060047 56 Gypsum -0.160 0.112 CaSO42H2O
010841302 65 Muscovite2M1 -0.160 0.204 KAl2.9Si3.1O10(OH2)
Figure 4: XRD pattern for bentonite used in this research (sample 1).

Figure 5: SEM pattern for bentonite used in this research (Sample 2).

283
Ref. Code Score Compound Displacem- Scale Chemical Formula
Name ent [2Th.] Factor
000120204 75 Montmor- -0.160 0.509 Nax(Al, Mg)2Si4
illonite O10(O H)2z H2O
000471743 58 Calcite -0.160 0.076 CaCO3
010820511 62 Quartz, syn -0.160 0.202 SiO2
000060047 56 Gypsum -0.160 0.112 CaSO42H2O
010841302 65 Muscovite2M1 -0.160 0.204 KAl2.9Si3.1O10(OH2)
Figure 6: X-Ray pattern for bentonite used in this research (sample 2).

Table 2: Data sheet for Wyoming Bentonite


Chemical Typical analysis
SiO2 63.50%
Al2O3 21.40%
Fe2O3 3.80%
MgO 2.00%
NaO 2.70%
Physical properties
Appearance Fine white gray
Moisture 10.0-14.0%
ulk density 0.9-1.00 Kgm-3

284
Table 3: Open Circuit Potential (OCP) in mV Vs. CSE for identical pipes buried at
different depths in bentonite contains 45% (w/w), 0.5 M NaCl solution; (a) after 1 day
and (b) after 7 days.
Depth (cm) (a) OCP after 1 day (b) OCP after 7 day
pipe1
0 -523 -597
2 -523 -599
4 -524 -601
6 -524 -598
8 -524 -596
10 -524 -595
pipe 2
12 -524 -599
14 -525 -600
16 -525 -603
18 -525 -603
20 -525 -602
22 -525 -602
pipe 3
24 -523 -581
26 -523 -584
28 -523 -583
30 -523 -582
32 -523 -581
34 -523 -581

O2 / OH-
E Fe/ Fe2+

Ecorr (1)
Ecorr (3)
Ecorr (2)

Time

Figure 7: Schematic diagram of potential vs. time.

285
Table 4: ON and OFF polarisation potential in mV Vs. CSE for 3 identical pipes buried
vertically at different depths in bentonite adding 45 % (w/w), 0.5 M NaCl solution.
Current density applied was 10.5mA.m-2, after day 1 and day 7, respectively. The ON
and OFF potential readings were taken three times for each depth (r1, r2 and r3) and the
average ON and OFF were calculated and Rs, is the calculated solution resistance, Rs.

ON potential, day 1 OFF potential, day 1

Pipe 1
St. St. Rs
Depth (cm) r1 r2 r3 av. r1 r2 r3 av.
dev. dev. .m
0
-1011 -1008 -1008 -1009 1.7 -887 -877 -875 -880 6.4 13.34
2
-1010 -1009 -1005 -1008 2.7 -890 -886 -883 -886 3.5 12.61
4
-1010 -1003 -993 -1002 8.5 -902 -891 -887 -893 7.8 11.27
6
-1008 -1008 -999 -1005 5.2 -900 -890 -881 -890 9.5 11.89
8
-1009 -1008 -1007 -1008 1 -901 -888 -888 -892 7.5 11.99
10
-1004 -1004 -995 -1001 5.2 -895 -887 -883 -888 6.1 11.68
pipe 2
12
-882 -882 -882 -882 0 -859 -864 -854 -859 5 2.38
14
-886 -885 -887 -886 1 -858 -860 -856 -858 2 2.9
16
-886 -883 -889 -886 3 -861 -860 -862 -861 1 2.59
18
-889 -888 -890 -889 1 -863 -865 -861 -863 2 2.69
20
-891 -890 -892 -891 1 -867 -870 -864 -867 3 2.48
22
-894 -893 -895 -894 1 -867 -873 -861 -867 6 2.79
pipe 3
24
-902 -903 -901 -902 1 -868 -861 -859 -863 4.7 4.03
26
-903 -902 -904 -903 1 -875 -884 -866 -875 9 2.9
28
-904 -903 -905 -904 1 -875 -885 -865 -875 10 3.0
30 -872 -879
-903 -902 -904 -903 1 -874 -875 3.6 2.89
32
-898 -898 -898 -898 0 -866 -875 -875 -872 5.2 2.69
34
-900 -894 -888 -894 6 -863 -872 -863 -866 5.2 2.9

ON potential, day 7 OFF potential, day 7


Pipe 1
St. St. Rs
Depth (cm) r1 r2 r3 av. r1 r2 r3 av.
dev. dev .m
0 .
-1024 -1023 -1022 -1023 1 -904 -895 -910 -903 7.6 12.41
2
-1018 -1016 -1015 -1016 1.5 -913 -916 -923 -917 5.1 10.24
4
-1025 -1024 -1024 -1024 0.6 -938 -930 -936 -935 4.2 9.2
6
-1030 -1030 -1029 -1030 0.6 -947 -941 -948 -945 3.8 8.79
8
-1037 -1036 -1037 -1037 0.6 -944 -930 -943 -939 7.8 10.13
10
-1039 -1039 -1037 -1038 1.2 -940 -933 -940 -938 4.0 10.34
pipe 2
12
-901 -900 -898 -900 1.5 -889 -878 -881 -883 5.7 1.76

286
14
-901 -900 -898 -900 1.5 -891 -876 -877 -881 8.4 1.96
16
-903 -903 -899 -902 2.3 -886 -877 -876 -880 5.5 2.27
18
-902 -900 -901 -901 1 -888 -883 -877 -883 5.5 1.86
20
-904 -900 -903 -902 2.1 -893 -889 -881 -888 6.1 1.45
22
-910 -907 -904 -907 3 -889 -887 -884 -887 2.5 2.07
pipe 3
24
-936 -935 -933 -935 1.5 -918 -909 -907 -911 5.9 2.48
26
-941 -939 -938 -939 1.5 -919 -911 -908 -913 5.7 2.69
28
-942 -942 -938 -941 2.3 -921 -912 -910 -914 5.9 2.79
30 -918 -914 -908 5.0
-944 -943 -939 -942 2.7 -913 3.0
32
-944 -942 -938 -941 3.1 -915 -912 -907 -911 4.0 3.10
34
-944 -945 -939 -943 3.2 -914 -908 -906 -909 4.2 3.51

The solution resistivity was calculated, using the following equation:

Rs (.m) = (IR x L)/( CD x A),

Where:

Rs is the resistivity of the solution between the working and reference electrodes.

IR drop is the difference between the ON and OFF potential, mV.

L is a distance between the Cathode and reference electrode, 0.02m.

A is the surface area of cathode, 0.018425m2.

CD is current density, mA.m-2

Table 5: ON and OFF polarisation potential in mV Vs. CSE for 3 identical pipes buried
vertically at different depths in bentonite adding 45 % (w/w), 0.5 M NaCl solution.
Current density applied was 50mA.m-2, after day 1 and day 7, respectively. The ON and
OFF potential readings were taken three times for each depth (r1, r2 and r3) and the
average ON and OFF were calculated and Rs, is the solution resistivity, Rs.

ON potential, day 1 OFF potential, day 1

Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev dev. .m
0 .
-1162 -1162 -1160 -1161 1.2 -936 -927 -919 -927 8.5 5.08
2
-1157 -1153 -1151 -1153 3.1 -947 -939 -934 -940 6.6 4.62
4
-1136 -1132 -1130 -1133 3.1 -945 -938 -932 -938 6.5 4.23
6
-1126 -1125 -1123 -1124 1.5 -938 -934 -928 -933 5.0 4.15
8
-1129 -1127 -1128 -1128 1 -939 -929 -935 -934 5.0 4.21

287
10
-1133 -1129 -1127 -1129 3.1 -932 -928 -925 -928 3.5 4.36
Pipe 2
12
-964 -961 -960 -962 2.1 -924 -921 -916 -920 4.0 0.91
14
-971 -968 -967 -969 2.1 -928 -918 -915 -920 6.8 1.06
16
-984 -980 -978 -980 3.1 -932 -922 -916 -923 8.1 1.24
18
-990 -986 -981 -986 4.5 -939 -931 -925 -932 7.0 1.17
20
-996 -992 -990 -992 3.1 -938 -928 -926 -931 6.4 1.32
22
-1004 -997 -995 -999 4.7 -934 -935 -925 -931 5.5 1.48
pipe 3
24
-956 -951 -945 -950 5.5 -912 -936 -956 -903 22.0 1.02
26
-955 -951 -948 -951 3.5 -910 -907 -897 -905 6.8 1.0
28
-948 -944 -939 -944 4.5 -900 -897 -886 -894 7.4 1.09
30 -900 -878 -872
-943 -940 -938 -940 2.5 -883 14.7 1.24
32
-941 -939 -936 -938 2.5 -885 -874 -870 -876 7.8 1.35
34
-956 -951 -945 -950 5.5 -883 -870 -864 -872 9.7 1.69

ON potential, day 7 OFF potential, day 7


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev dev. .m
0 .
-1350 -1347 -1345 -1347 2.5 -1013 -1016 -1010 -1013 3 7.25
2
-1357 -1355 -1354 -1355 1.5 -1013 -1004 -1012 -1010 4.9 7.49
4
-1343 -1339 -1337 -1340 3.1 -1004 -1002 -1007 -1004 2.51 7.29
6
-1309 -1304 -1297 -1303 6.0 -1011 -1013 -998 -1007 8.1 6.43
8
-1274 -1269 -1267 -1270 3.6 -1007 -997 -989 -998 9.0 5.91
10
-1307 -1300 -1297 -1301 5.1 -984 -980 -976 -980 4 6.97
pipe 2
12
-1044 -1039 -1038 -1015 3.5 -972 -975 -968 -972 3.5 0.93
14
-1057 -1051 -1054 -1040 3.2 -972 -973 -970 -972 1.5 1.48
16
-1064 -1063 -1063 -1055 3 -986 -980 -975 -981 5.5 1.61
18
-1064 -1063 -1064 -1063 0.6 -981 -982 -979 -981 1.5 1.78
20
-1075 -1072 -1070 -1064 0.6 -989 -985 -980 -985 4.5 1.72
22
-1044 -1039 -1038 -1072 2.5 -996 -988 -988 -991 4.6 1.76
pipe 3
24 4.2
-1045 -1044 -1045 -1045 0.6 -978 -972 -970 -973 1.56
26
-1045 -1044 -1042 -1044 1.5 -970 -968 -971 -970 1.5 1.61
28
-1041 -1039 -1038 -1039 1.5 -968 -966 -965 -966 1.5 1.59
30 -956 -949
-1035 -1033 -1032 -1033 1.5 -944 -950 6.0 1.80
32
-1038 -1038 -1037 -1038 0.6 -949 -945 -937 -944 6.1 2.04
34
-1040 -1038 -1037 -1048 1.5 -940 -932 -927 -933 6.6 2.5

288
Table 6: ON and OFF polarisation potential in mV Vs. CSE for 3 identical pipes buried
vertically at different depths in bentonite adding 45 % (w/w), 0.5 M NaCl solution.
Current density applied was 100mA.m-2, after day 1 and day 7, respectively. The ON
and OFF potential readings were taken three times for each depth (r1, r2 and r3).

ON potential, day 1 OFF potential, day 1


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev dev. .m
0 . 5.1
-1228 -1225 -1223 -1225 2.5 -1007 -1000 -997 -1001 2.21
2
-1222 -1220 -1219 -1220 1.5 -1017 -1006 -1011 -1011 5.5 2.17
4
-1212 -1211 -1208 -1210 2.1 -1010 -1005 -1002 -1006 4.0 2.39
6 -1005 -982
-1192 -1191 -1192 -1192 0.6 -989 -992 11.8 2.61
8
-1221 -1218 -1217 -1219 2.1 -1003 -1001 -994 -999 4.7 2.21
10
-1236 -1233 -1232 -1233 2.1 -991 -1000 -989 -993 5.9 2.17
pipe 2
12
-1073 -1070 -1068 -1070 2.5 -998 -984 -981 -988 9.1 0.89
14
-1072 -1072 -1070 -1071 1.2 -976 -990 -982 -982 7.0 0.97
16
-1087 -1086 -1087 -1087 0.6 -1007 -992 -996 -998 7.8 0.97
18
-1094 -1093 -1093 -1093 0.6 -993 -1002 -998 -998 4.5 1.03
20
-1107 -1104 -1103 -1104 2.1 -1006 -998 -985 -996 10.6 1.17
22
-1122 -1120 -1119 -1120 1.5 -1008 -1002 -989 -1000 9.7 1.30
pipe 3
24
-1066 -1064 -1063 -1064 1.5 -991 -977 -973 -980 9.45 0.91
26
-1059 -1045 -1041 -1048 9.5 -989 -979 -978 -982 6.1 0.72
28
-1063 -1062 -1059 -1061 2.1 -986 -980 -972 -979 7.0 0.89
30 -985 -977 -973
-1068 -1062 -1063 -1064 3.2 -978 6.1 0.93
32
-1046 -1043 -1039 -1042 3.5 -972 -964 -960 -965 6.1 0.84
34
-1061 -1056 -1053 -1057 4.0 -978 -963 -957 -966 4.0 0.99

ON potential, day 7 OFF potential, day 7


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev dev. .m
0 .
-1442 -1442 -1441 -1442 0.6 -1037 -1025 -1029 -1030 6.1 4.47
2
-1439 -1438 -1436 -1438 1.5 -1017 -1003 -996 -1005 10.7 4.70
4
-1384 -1383 -1378 -1382 3.2 -1026 -1003 -993 -1007 16.9 4.07
6
-1390 -1388 -1386 -1388 2 -1023 -1012 -978 -1004 23.5 4.17
8
-1387 -1385 -1385 -1386 1.2 -1001 -988 -981 -990 10.2 4.3
10
-1411 -1408 -1407 -1409 2.1 -980 -974 -985 -980 5.5 4.66
pipe 2
12
-1083 -1079 -1077 -1080 3.1 -991 -980 -977 -983 7.4 1.05
14
-1105 -1103 -1100 -1103 2.5 -1002 -992 -977 -990 12.6 1.23

289
16
-1116 -1113 -1113 -1114 1.7 -1021 -1009 -997 -1009 12 1.14
18
-1122 -1119 -1117 -1119 2.5 -1018 -1009 -997 -1008 10.5 1.21
20
-1120 -1119 -1116 -1118 2.2 -1013 -998 -1003 -1005 7.6 1.23
22
-1134 -1131 -1129 -1131 2.5 -1016 -1005 -986 -1002 15.2 1.40
pipe 3
24
-1074 -1072 -1071 -1072 1.5 -989 -983 -979 -984 5.0 0.96
26
-1080 -1077 -1076 -1078 2.1 -1007 -988 -975 -990 16.1 0.96
28
-1078 -1079 -1077 -1078 1 -977 -973 -961 -970 8.3 1.17
30 -970
-1078 -1077 -1075 -1077 1.5 -965 -959 -965 5.5 1.22
32 -949
-1075 -1072 -1070 -1072 2.5 -964 -938 -950 13.1 1.32
34
-1079 -1076 -1075 -1077 2.1 -955 -948 -930 -944 12.9 1.44

Table 7: ON and OFF polarisation potential in mV Vs. CSE for 3 identical pipes buried
vertically at different depths in bentonite adding 45 % (w/w), 0.5 M NaCl solution.
Current density applied was 150mA.m-2, after day 1 and day 7, respectively. The ON
and OFF potential readings were taken three times for each depth (r1, r2 and r3).

ON potential, day 1 OFF potential, day 1


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev. dev. .m
0
-1243 -1241 -1237 -1240 3.1 -1028 -1003 -994 -1008 6.3 1.68
2
-1253 -1248 -1244 -1248 4.51 -1022 -1007 -1002 -1010 10.4 1.72
4
-1221 -1220 -1217 -1219 2.1 -1017 -1004 -998 -1006 9.7 1.54
6
-1243 -1237 -1234 -1238 4.2 -1009 -1004 -994 -1002 7.6 1.71
8
-1259 -1253 -1246 -1253 6.5 -1011 -997 -991 -999 10.3 1.84
10
-1319 -1309 -1307 -1312 6.4 -1019 -1017 -994 -1010 13.9 2.19
pipe 2
12
-1151 -1144 -1138 -1144 6.5 -1021 -1013 -1003 -1012 9.0 0.96
14
-1159 -1154 -1151 -1155 4.0 -1024 -1019 -1007 -1017 8.7 1.0
16
-1164 -1158 -1154 -1159 5.0 -1034 -1029 -1023 -1029 5.5 0.94
18
-1180 -1172 -1169 -1174 5.7 -1029 -1021 -1017 -1022 6.1 1.1
20
-1192 -1185 -1182 -1186 5.1 -1030 -1024 -1011 -1022 9.7 1.19
22
-1224 -1217 -1214 -1218 5.1 -1022 -1016 -1007 -1015 7.6 1.47
pipe 3
24
-1111 -1109 -1105 -1108 3.1 -1031 -1024 -1019 -1025 6.0 0.60
26
-1134 -1133 -1130 -1132 2.1 -1029 -1025 -1023 -1026 3.1 0.77
28
-1149 -1145 -1143 -1146 3.1 -1035 -1029 -1019 -1028 8.1 0.85
30
-1157 -1147 -1144 -1149 6.8 -1032 -1025 -1018 -1025 7 0.9

290
32
-1122 -1138 -1140 -1133 9.87 -1026 -1022 -1009 -1019 8.9 0.83
34
-1127 -1131 -1136 -1131 4.51 -1020 -1014 -1008 -1014 6 0.85

ON potential, day 7 OFF potential, day 7


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev dev. .m
0 .
-1418 -1407 -1396 -1407 11 -1070 -1061 -1047 -1059 5.0 2.52
2
-1064 -1058 -1054 -1395 4.9 -1062 -1051 -1044 -1052 9.1 2.48
4
-1364 -1359 -1357 -1360 3.6 -1052 -1043 -1047 -1047 4.5 2.26
6
-1346 -1339 -1335 -1340 5.6 -1039 -1025 -1019 -1028 10.3 2.26
8
-1356 -1351 -1351 -1353 2.9 -1023 -1017 -1007 -1016 8.1 2.44
10
-1385 -1379 -1378 -1381 3.8 -1026 -1017 -1003 -1015 11.6 2.65
pipe 2
12
-1152 -1143 -1139 -1145 6.7 -1030 -1025 -1007 -1021 12.1 0.9
14
-1151 -1149 -1141 -1147 5.3 -1029 -1021 -1016 -1022 6.6 0.91
16
-1181 -1174 -1170 -1175 5.6 -1031 -1024 -1019 -1025 6.0 1.1
18
-1186 -1183 -1181 -1183 2.6 -1036 -1025 -1017 -1026 9.6 1.14
20
-1199 -1196 -1193 -1196 3 -1038 -1029 -1021 -1029 8.6 1.21
22
-1220 -1215 -1212 -1216 4.1 -1035 -1028 -1022 -1028 6.6 1.36
pipe 3
24
-1121 -1117 -1113 -1117 4 -1031 -1019 -1011 -1020 10.1 0.70
26
-1140 -1136 -1132 -1136 4 -1026 -1019 -999 -1015 14.1 0.88
28
-1158 -1149 -1144 -1150 7.1 -1029 -1019 -1009 -1019 10 0.95
30
-1163 -1153 -1149 -1155 7.2 -1027 -1018 -1011 -1019 8.1 0.98
32
-1159 -1154 -1153 -1156 3.2 -1021 -1008 -1004 -1011 8.9 1.05
34
-1168 -1164 -1163 -1165 2.7 -1022 -1013 -1007 -1014 7.6 1.09

Table 8: ON and OFF polarisation potential in mV Vs. CSE for 3 identical pipes buried
vertically at different depths in bentonite adding 45 % (w/w), 0.5 M NaCl solution.
Current density applied was 200mA.m-2, after day 1 and day 7, respectively. The ON
and OFF potential readings were taken three times for each depth (r1, r2 and r3).

ON potential, day 1 OFF potential, day 1


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev. dev. .m
0
-1123 -1120 -1118 -1120 2.5 -1137 -1131 -1122 -1030 7.6 0.49
2
-1147 -1146 -1143 -1145 2.1 -1139 -1128 -1123 -1030 8.2 0.62
4
-1146 -1145 -1142 -1144 2.1 -1124 -1117 -1111 -1017 6.5 0.69
6 -1121 -1101
-1163 -1160 -1158 -1160 2.5 -1111 -1011 10 0.81

291
8
-1157 -1155 -1153 -1155 2 -1024 -1017 -1012 -1018 6.0 0.74
10
-1181 -1177 -1175 -1178 3.1 -1041 -1035 -1029 -1035 6 0.78
pipe 2
12
-1139 -1128 -1124 -1130 7.8 -1011 -1003 -990 -1001 10.6 0.70
14
-1110 -1103 -1098 -1104 6.0 -1010 -998 -989 -999 10.5 0.57
16
-1131 -1124 -1121 -1125 5.1 -1007 -1001 -993 -1000 7.0 0.68
18
-1136 -1135 -1129 -1133 3.8 -1011 -998 -991 -1000 10.1 0.72
20
-1151 -1143 -1142 -1145 4.9 -1014 -1001 -997 -1004 8.9 0.77
22
-1126 -1122 -1119 -1122 3.5 -1017 -1011 -1002 -1010 7.6 0.61
pipe 3
24
-1191 -1186 -1184 -1187 3.6 -1050 -1041 -1038 -1043 6.2 0.78
26
-1175 -1171 -1163 -1170 6.1 -1044 -1039 -1032 -1038 6.0 0.72
28
-1224 -1204 -1210 -1213 10.3 -1057 -1048 -1039 -1048 9 0.9
30 -1084 -1079 -1068
-1215 -1198 -1209 -1209 8.6 -1077 8.2 0.71
32
-1198 -1193 -1191 -1194 3.6 -1087 -1075 -1074 -1079 7.2 0.62
34
-1201 -1197 -1196 -1198 2.7 -1074 -1064 -1058 -1065 8.1 0.72

ON potential, day 7 OFF potential, day 7


Pipe 1
Depth St. St. Rs
r1 r2 r3 av. r1 r2 r3 av.
(cm) dev. dev. .m
0
-1369 -1365 -1363 -1366 3.1 -1087 -1072 -1065 -1075 11.2 1.58
2
-1341 -1339 -1338 -1339 1.6 -1097 -1083 -1074 -1085 11.6 1.38
4
-1335 -1332 -1330 -1332 2.5 -1072 -1061 -1057 -1063 7.8 1.46
6
-1371 -1368 -1369 -1369 1.5 -1064 -1059 -1043 -1055 11 1.7
8
-1364 -1363 -1360 -1362 2.1 -1060 -1055 -1051 -1056 4.5 1.66
10
-1417 -1413 -1410 -1413 3.5 -1059 -1044 -1039 -1047 10.4 1.99
pipe 2
12
-1255 -1253 -1249 -1252 3.1 -1062 -1048 -1036 -1049 13.0 1.10
14
-1253 -1250 -1248 -1250 2.5 -1067 -1063 -1035 -1055 17.4 1.06
16
-1283 -1280 -1277 -1280 3 -1073 -1068 -1039 -1060 18.4 1.19
18
-1309 -1302 -1301 -1304 4.4 -1053 -1049 -1034 -1045 10.0 1.41
20
-1325 -1318 -1315 -1319 5.1 -1078 -1068 -1045 -1064 16.9 1.38
22
-1336 -1329 -1323 -1329 6.5 -1062 -1044 -1039 -1048 12.1 1.53
pipe 3
24
-1214 -1209 -1203 -1209 5.5 -1047 -1032 -1024 -1034 11.7 0.95
26
-1243 -1236 -1231 -1237 6.0 -1056 -1044 -1021 -1040 17.8 1.07
28
-1254 -1247 -1235 -1245 9.6 -1061 -1049 -1039 -1050 11.0 1.06
30 -1069
-1259 -1252 -1240 -1250 9.6 -1049 -1044 -1054 13.2 1.06
32 -1057
-1263 -1260 -1258 -1260 2.5 -1066 -1047 -1057 9.5 1.10
34
-1279 -1268 -1264 -1270 7.8 -1070 -1058 -1038 -1055 16.2 1.17

292
11.2 Appendix II: Bare mild steel immersed in 0.5 M NaCl Solution
free of bentonite.

(a) 1 day (b) 3 days

(c) 5 days (d) 7 days

(e) 14 days (f) Coupons in 0.5M NaCl solution

293
(g) Corrosion of steel coupons in 0.5M NaCl solution after 14 days

Figure 8: Corrosion progress of mild steel specimens after immersed in 3.0% NaCl
solution containing no bentonite for different time with no cathodic protection.

Table 9: Corrosion rates in mm/y for bare mild steel specimens immersed in 0.5M NaCl
solution free of bentonite. The data extracted from measurement of weight loss
experimentally for 1, 3, 5, 7 and 14 days respectively.

Time (day) Sample 1 Sample 2 average


1 0.253 0.288 0.270
3 0.243 0.259 0.251
5 0.161 0.189 0.175
7 0.155 0.171 0.163
14 0.197 0.220 0.206

Table 10: Corrosion potential in mV vs. SCE for mild steel after immersed in aerated
0.5 M NaCl solution for 14 days at OCP for test 1, 2 and 3.

Time (day) Sample 1 Sample 2 Sample 3 average


0 -360.7 -348.4 -340.2 -349.8
1 -694.9 -702.9 -664.4 -687.4
3 -725.5 -725.9 -708.2 -719.9
5 -703.9 -723.2 -703.5 -710.2
7 -718.2 -742.1 -701.5 -720.6
14 -692.8 -716.1 -669.9 -692.9

294
Table 11: LPR in Ohm.cm2 for test 1, 2 and 3 for mild steel after immersed in aerated
0.5 M NaCl at free corrosion potential, (at OCP).

Time day Sample 1 Sample 2 Sample 3 average


0 3048.8 2547.0 2519.8 2705.2
1 2466.8 3522.5 2238.0 2742.4
3 3581.3 2477.2 2876.1 2978.2
5 2442.6 2396.2 2427.4 2422.1
7 3710.2 3790.8 2398.3 3299.8
14 2042.2 2212.0 2272.2 2175.5

Figure 9: LPR for mild steel after immersed in aerated 0.5 M NaCl solution after one
hour, at OCP.

295
Figure 10: LPR runs for mild steel immersed in aerated 0.5 M NaCl solution after 1 day,
at OCP.

Figure 11: LPR runs for mild steel immersed in aerated 0.5 M NaCl solution after 3
days, at OCP.

296
Figure 12: LPR runs for mild steel immersed in aerated 0.5 M NaCl solution after 5
days, at OCP.

Figure 13: LPR runs for mild steel immersed in aerated 0.5 M NaCl solution after 7
days, at OCP.

297
Figure 14: LPR runs for mild steel immersed in aerated 0.5 M NaCl solution after 14
days, at OCP.

Table 12: showing Rp in .cm2, icorr in mA.cm-2 and corrosion rate (CR) in mm/y for 3
mild steel specimens after immersed in aerated 0.5 M NaCl solution free of bentonite at
natural corrosion potential.
Specimen 1
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 3049 0.009 0.099
1 2467 0.011 0.122
3 3582 0.007 0.084
5 2443 0.011 0.123
7 3710 0.007 0.081
14 2042 0.013 0.147
Specimen 2
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2547 0.010 0.111
1 3523 0.007 0.078
3 2477 0.011 0.122
5 2396 0.011 0.122
7 3791 0.007 0.078
14 2212 0.012 0.133
Specimen 3
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2520 0.010 0.111
1 2238 0.012 0.133
3 2876 0.009 0.10

298
5 2427 0.011 0.122
7 2398 0.010 0.111
14 2272 0.010 0.111

Table 13: Current density (CD) in mA.m-2 against time in days for mild steel immersed
in aerated 0.5 M NaCl solution for test 1, 2 and 3, under cathodic protection at -
0.80V/SCE.

Time Sample 1 (CD) Sample 2(CD) Sample 3(CD) Average(CD)


mA.m-2 mA.m-2 mA.m-2
(day) mA.m-2
0 247 244 249 247
1 142 149 1506 147
2 149 148 138 145
3 134 130 131.3 132
4 140 138 114.6 131
5 128 131 125.3 128
6 125 129 133.2 129.1
7 121 126 122.7 123.2

Table 14: Current density (CD) values in mA.m-2 against time in days for mild steel,
after immersed in aerated 0.5 M NaCl solution for test 1, 2 and 3 under cathodic
protection at -1.150V/SCE.
Time Sample 1 (CD) Sample 2(CD) Sample 3(CD) Average(CD)
(day) mA.m-2 mA.m-2 mA.m-2 mA.m-2
0 2253 2711.4 2788.5 2584.3
1 1382 1907 1910 1733
2 1320 1674.3 1820 1604.7
3 0.1192 0.158 0.1571 1448
4 788 970 986 915
5 760 851 893 835
6 508 842 805 718
7 477 456 531 488
14 441 406 405 417

299
Table 15: pH for 3 specimens after immersed in 0.5 NaCl solution for 14 days, under
CP at -1.150V/SCE.
Time (day) pH for sample 1 pH for sample 2 pH for sample 3 Average pH
0 8 8 8 8
1 9.4 9.5 9.3 9.4
2 9.4 9.3 9.3 9.33
3 9.5 9.4 9.4 9.43
4 9.4 9.5 9.6 9.5
5 9.3 9.4 9.5 9.4
6 9.4 9.4 9.4 9.4
7 9.1 9.1 9.1 9.1
14 9.1 9.2 9.3 9.2

Figure 15: Graph showing LPR plots for mild steel specimen 1 after immersed in
aerated 0.5 M NaCl solution. The test was run for 3 times. It was under cathodic
polarisation at -1.150V/SCE for 14 days. CP was OFF.

300
Figure 16: Graph showing LPR plots for mild steel specimen 2, after immersed in
aerated 0.5 M NaCl solution. The test was run for 3 times. It was under cathodic
polarisation at -1.150V/SCE) for 14 days. CP was OFF.

Figure 17: Graph showing LPR plots for mild steel specimen 3, after immersed in
aerated 0.5 M NaCl solution. The test was run for 3 times. It was under cathodic
polarisation, at -1.150V/SCE for 14 days. CP was OFF.

301
Element AN Series norm. wt. %
Carbon 6 K series 8.835241
Iron 26 K series 91.44937

Figure 18: SEM pattern for steel specimen 2 immersed in 0.5 M NaCl at -1.15V/SCE
polarisation potential.

302
C o unts
O ra yith_ S 2 _ 3p e rN a C l_ 1 1 5 0

40000

10000

0
10 20 30 40 50 60 70 80

P o sitio n [2 The ta ] (C o b a lt (C o ))

Steel only, no unidentified reflections


Figure 19: X-Ray pattern for steel coupon immersed in 0.5 M NaCl at -1.150V/SCE
polarisation potential (Sample 2).

11.3 Appendix III: Bare mild steel immersed in 0.5 M NaCl Solution
containing 1% (w/w) bentonite.

(a) 1 day (b) 3 days

303
(c) 5 days (d) 7 days

Figure 20: Photograph showing corrosion progresses of mild steel coupons immersed in
3.0% NaCl solution containing 1.0% bentonite for different time with no cathodic
protection.

(a) pH at the initial time (a) 1 day (b) 3 days

(c) 5 days (d) 7 days (e) 14 days

Figure 21: Photograph showing pH for mild steel specimens after immersed in 3.0%
NaCl solution containing 1.0% bentonite at natural corrosion potential at different time.

304
Table 16: Values of corrosion rates (mm/y) for bare mild steel specimens after
immersed in 0.5M NaCl solution containing 1.0% bentonite. The data was extracted
from weight loss measurements, experimentally for 1, 3, 5, 7 and 14 days respectively.

Time (day) Sample 1 Sample 2 Avg. C.R St. dev.


1 0.117 0.112 0.115 0.004
3 0.177 0.202 0.190 0.02
5 0.187 0.206 0.197 0.01
7 0.185 0.204 0.195 0.01
14 0.178 0.190 0.184 0.01

Table 17: Corrosion potential values in mV Vs. SCE for mild steel specimens, after
immersing in aerated 0.5 M NaCl solution containing 1% bentonite at OCP.

Time (day) Sample 1 (mV) Sample 2 Sample 3 average


0 -295.6 -320.5 -300.3 -305.5
1 -734.7 -736.174 -739.7 -736.9
3 -734.2 -739.3 -738.2 -737.2
5 -711.5 -735.7 -742 -729.7
7 -721.7 -721.1 -726.7 -723.2
14 -714.6 -699.4 -726.7 -713.6

Table 18: Rp values for test 1, 2 and 3 for mild steel specimens after immersed in
aerated 0.5 M NaCl containing 1.0% bentonite at OCP for 14 days.

Time (day) Sample 1 Sample 2 Sample 3 average


0 2102 1621 1679 1800.7
1 1565 1547 1562 1558
3 2118 1918 1730 1922
5 2382 1907 1694 1994.3
7 2739 2231 2446 2472
14 2554 2046 1850 2150

305
Table 19: Values for Rp in .cm2, icorr in mA.cm-2 and corrosion rate (CR) in mm/y for
3 mild steel specimens after immersed in aerated 0.5 M NaCl solution adding 1.0%
(w/w) bentonite at natural corrosion potential.
Specimen 1
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2102 0.012 0.134
1 1565 0.015 0.168
3 2118 0.012 0.134
5 2382 0.010 0.11
7 2739 0.010 0.11
14 2554 0.012 0.134
Specimen 2
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 1621 0.021 0.235
1 1547 0.017 0.190
3 1918 0.015 0.168
5 1907 0.012 0.134
7 2231 0.014 0.156
14 2046 0.011 0.123
Specimen 3
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 1679 0.019 0.212
1 1562 0.019 0.212
3 1730 0.017 0.190
5 1694 0.011 0.123
7 2446 0.014 0.156
14 1850 0.019 0.212

306
Figure 22: Graph showing LPR plots for mild steel sample 1, after immersed in aerated
0.5 M NaCl solution containing 1.0% bentonite for 1 hour (day 0).

Figure 23: Graph showing Rp plot for mild steel sample 1 after immersed in aerated 0.5
M NaCl solution containing 1.0% bentonite after1 day.

307
Figure 24: Graph showing Rp plot for mild steel sample 1 after immersed in aerated 0.5
M NaCl solution containing 1.0% bentonite after 14 days.

Table 20: Current density in mA.m-2 against time in days for mild steel immersed in
aerated 0.5 M NaCl solution containing 1.0 % bentonite for test 1, 2 and 3 under CP at -
1.150V/SC.
Time Sample 1 Sample 2 Sample 3 Average
0 2732 2978 3071 2927
1 1384 1487 1613 1494
2 601 699 611 637
3 565 633 502 567
4 497 542.6 459 500
5 418 457 380 418
6 309 351 368 342
7 231 281 324 279
14 207 219 248 225

308
11.4 Appendix IV: mild steel immersed in 0.5 M NaCl Solution
containing 10.0% (w/w) bentonite.

Table 21: corrosion rate (mm/y) for bare mild steel coupons immersed in 0.5M NaCl
solution containing 10.0% bentonite. The data extracted from weight loss measurement
experimentally for 1, 3, 5, 7 and 14 days respectively.
time Sample 1 Sample 2 Avg C.R St. dev.
1 0.117 0.113 0.115 0.003
3 0.177 0.184 0.181 0.01
5 0.187 0.195 0.191 0.01
7 0.185 0.194 0.19 0.01
14 0.184 0.190 0.187 0.004

Table 22: Corrosion potential values in mV vs. SCE for mild steel immersed in aerated
0.5 M NaCl solution containing 10% bentonite ( at OCP).
Time Sample 1 Sample 2 Sample 3 average
0 -309.6 -328.3 -302.3 -313.4
1 -742.2 -747.1 -746.7 -745.3
3 -740.5 -744.7 -742.8 -742.7
5 -737.9 -741.8 -741.3 -740.3
7 -739.9 -734.7 -736.1 -736.9
14 -728.8 -737.1 -730.1 -732

Table 23: The Rp values for mild steel tested in aerated 0.5 M NaCl containing 10%
(w/w) bentonite the specimens immersed in solution at OCP for 1, 3, 5, 7 and 14 days
respectively. .
Time Sample 1 Sample 2 Sample 3 average
0 4784 4530 3771 4362
1 4395 3405 2388 3396
3 3874 3293 2454 3207
5 3267 2327 2377 2657
7 2970 1613 2157 2247
14 3065 1956 1946 2322

309
Table 24: showing Rp in .cm2, icorr in mA.cm-2 and corrosion rate (CR) in mm/y for 3
mild steel specimens after immersed in aerated 0.5 M NaCl solution adding 10.0%
(w/w) bentonite at natural corrosion potential.
Specimen 1
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 4784 0.006 0.067
1 4395 0.006 0.067
3 3874 0.007 0.078
5 3267 0.008 0.089
7 2970 0.009 0.10
14 3065 0.009 0.10
Specimen 2
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 4530 0.006 0.067
1 3405 0.008 0.089
3 3293 0.008 0.089
5 2327 0.011 0.123
7 1613 0.016 0.178
14 1956 0.013 0.145
Specimen 3
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 3771 0.007 0.078
1 2388 0.011 0.123
3 2454 0.011 0.123
5 2377 0.011 0.123
7 2157 0.012 0.134
14 1946 0.013 0.145

Table 25: Values for current density in mA.m-2 over time in days for mild steel
immersed in aerated 0.5 M NaCl solution containing 10.0 % bentonite for test 1, 2 and 3
under CP at -0.80V/SCE.
Time Sample 1 Sample 2 Sample 3 Average
0 152.3 150.5 144.8 149.2
1 133.3 135 137.7 135.3
2 126 127.2 128.1 127.1
3 108.1 113 111.1 110.7
4 88.3 92 88.8 89.7
5 68.1 70.2 65.6 68.0
6 51.6 55.2 53.1 53.3
7 41.2 45 44.3 43.5
14 40.3 43 45 42.8

310
Table 26: Current density in mA.m-2 against time in days for mild steel specimens after
immersed in aerated 0.5 M NaCl solution containing 10.0 % bentonite for test 1, 2 and
3under CP at -1.150V/SCE (at OFF potential).
Time Sample 1 Sample 2 Sample 3 Average
0 2636.2 2600.3 2606.7 2614.4
1 1446.6 1453.1 1245 1381.6
2 690.9 703.6 664.4 686.3
3 386 373.6 367.3 375.6
4 252.1 233.1 241.6 242.3
5 211.7 203.1 198.2 204.3
6 168.7 138.4 144.2 150.5
7 124.9 113 122.8 120.2
14 119 116 120.8 118.6

Table 27: pH values for mild steel coupons immersed in 3.0% NaCl solution with
10.0% bentonite at -1.15V cathodic potential at different period.
Time (day) PH s1 PH s2 PH s3 pHAverage
0 8.7 8.8 8.7 8.7
1 10.3 10 10.3 10.2
3 10.5 10.6 10.8 10.6
5 9.7 10.3 9.9 10.0
7 9.8 9.9 10.2 10.0
14 10.3 9.4 9.5 9.7

11.5 Appendix V: mild steel buried inside the bentonite layer at the
bottom of the beaker below the solution level. The corrosive
electrolyte was 0.5 M NaCl containing 10.0% (w/w) Bentonite.

Table 28: corrosion rate (mm/y) for bare mild steel specimens buried inside the
bentonite layer below the solution level of 0.5M NaCl solution containing 10.0%
bentonite. The data extracted from weight loss experiments for 1, 3, 5, 7 and 14 days,
respectively.
Time Sample 1 Sample 2 Avg CR St. dev.
1 0.035 0.025 0.030 0.01
3 0.012 0.012 0.012 0.0
5 0.009 0.009 0.009 0.0
7 0.008 0.008 0.008 0.0
14 0.007 0.006 0.006 0.001

311
Table 29: Corrosion potential values for mild steel after buried inside the bentonite layer
below the solution level of aerated 0.5 M NaCl solution adding 10% bentonite.
Time Sample 1 Sample 2 Sample 3 average
0 -305.6 -308.5 -320.4 -311.5
1 -794.8 -795.9 -799.4 -796.7
3 -804.4 -813.8 -803.7 -807.3
5 -807.7 -811.6 -814.4 -811.2
7 -807.8 -803.5 -818.598 -809.9
14 -795.2 -812.754 -809.68 -805.9

Figure 25: Graph showing plots illustrating Rp runs for mild steel specimen 1 buried
inside the bentonite layer below the solution level of 0.5M NaCl solution containing
10.0% (w/w) bentonite for 1 day.

312
Figure 26: Graph showing plots illustrating Rp runs for mild steel specimen 1 buried
inside the bentonite layer below the solution level of 0.5M NaCl solution containing
10.0% (w/w) bentonite for 7 days.

Figure 27: Graph showing plots illustrating Rp runs for mild steel specimen 1 buried
inside the bentonite layer below the solution level of 0.5M NaCl solution containing
10.0% (w/w) bentonite for 14 days.

Table 30: showing Rp in .cm2, icorr in mA.cm-2 and corrosion rate (CR) in mm/y for
3 specimens buried in the bentonite layer below the solution level of 0.5M NaCl
solution adding 10.0% (w/w) bentonite for 14 days at natural corrosion potential.

313
Specimen 1
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 4367.5 0.0060 0.067
1 22796.2 0.0010 0.011
3 56088.3 0.0004 0.004
5 68080.5 0.0004 0.004
7 82552 0.0003 0.003
14 112292 0.0004 0.004
Specimen 2
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 5223.1 0.0050 0.056
1 23186.4 0.0010 0.011
3 73534.4 0.0004 0.004
5 88744.3 0.0003 0.003
7 84369.8 0.0003 0.003
14 105708.1 0.0003 0.003
Specimen 3
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 5421.9 0.0050 0.056
1 24222.0 0.0010 0.011
3 48469.6 0.0005 0.006
5 87529 0.0003 0.003
7 104946.7 0.0002 0.002
14 117527 0.0002 0.002

Table 31: The Rp values for mild steel specimens buried in the bentonite layer below the
solution level of 0.5 M NaCl adding 10% (w/w) bentonite for 14 days.
Time Sample 1 Sample 2 Sample 3 average
0 4367.5 5223 5421 5004
1 22796 23186 24222 23401
3 56089 73534 48470 59364
5 68081 88744 87529 81451
7 82552 84370 104947 90623
14 112293 105708 117527 111843

314
Table 32: Values of current density measurement for mild steel specimens buried inside
the bentonite layer at the bottom of the beakers below the solution level of an aerated
0.5 M NaCl Solution adding 10.0% bentonite, under CP at -0.80V vs. SCE.
Time/ day Sample 1 Sample 2 Sample 3 average
0 53.4 40 41.2 44.9
1 0.6 1.5 1.7 1.3
2 -0.3 -2.8 -2.1 -1.7
3 -2.3 -2.5 -1.87 -2.2
4 -2.5 -1.7 -1.3 -1.8
5 -2.3 -2.7 -2.3 -2.4
6 -1.9 -1.4 -2.4 -1.9
7 -1.8 -1.5 -1.7 -1.7
14 -1.5 -1.8 -1.8 -1.7

Table 33: pH for mild steel specimens buried in 3.0% NaCl solution adding 10.0%
bentonite under CP at -0.80 V vs. SCE for different period.
Time/days pH1 pH2 pH3 pH4
0 8 8 8 8
1 7.9 7.9 7.8 7.9
2 5.9 6 6 6
3 5.9 6 6 6
5 6 6.1 6.1 6.2
6 6 5.9 6 6.1
7 5.9 5.9 5.9 5.9
14 5.9 6 5.9 6

Table 34: Current density values in mA.m-2 over time in days for mild steel specimens
buried inside the bentonite layer, below the solution level of 0.5 M NaCl solution added
to 10.0% bentonite for test 1, 2 and 3 at -1.150V/SCE.

Time Sample 1 Sample 2 Sample 3 average


0 1041 1048.6 1114.4 1068
1 161.9 236.2 224 207.4
2 119.1 195.5 181.3 165.3
3 89.5 153.3 146.6 129.8
4 72.4 102.9 123.8 99.7
5 60.95 93.3 102.87 85.7
6 60.9 91.43 100.9 84.4
7 54.193 76.2 81.9 70.8
14 53.39 54.23 57.21 54.9

315
Table 35: pH for mild steel specimens buried inside the bentonite layer below the
solution level of 3.0% NaCl solution adding 10.0% bentonite under CP at -1.150 V vs.
SCE for different period.
Time Sample 1 Sample 2 Sample 3
0 8 8 8
1 8 8.2 8
2 8.1 8.1 8.3
3 8.1 8.4 8.1
4 8.3 8.4 8.4
5 8.1 7.9 8.4
6 7.9 7.9 7.9
7 7.9 7.9 7.8
14 7.8 7.9 7.7

C o unts
O ra yith_ S 2 _ 2 0 B B _ 8 5 0

160000

40000

0
10 20 30 40 50 60 70 80

P o s itio n [2 T he ta ] (C o b a lt (C o ))

Visible Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
* 00-006-0696 96 ferrite 0.000 1.006 Fe

Figure 28: XRD pattern for steel coupon buried inside the bentonite layer, below the
solution level of 0.5 M NaCl solution added to 10.0% bentonite under cathodic
protection at -0.85 V, after 14 days.

316
C o unts
O ra yith_ S 2 _ 3 0 p e rB B _ 8 5 0

22500

10000

2500

0
10 20 30 40 50 60 70 80

P o s itio n [2 T he ta ] (C o b a lt (C o ))

Visible Ref. Code Score Compound Displacement Scale Chemical


Name [2Th.] Factor Formula
* 00-006-0696 96 ferrite 0.000 0.890 Fe

Figure 29: X-Ray pattern for steel coupon buried inside the bentonite layer, below the
solution level of 0.5 M NaCl solution added to 10.0% bentonite under cathodic
protection at -0.85 V, after 14 days.

11.6 Appendix VI: Mild Steel Immersed in 0.5 M Sodium Chloride


Solution adding 10.0% (w/w) Bentonite and ZnCl2.

Table 36: Corrosion potential values for mild steel specimens after immersing in aerated
0.5 M NaCl solution added to 10% bentonite and ZnCl2 at 500ppm.
Time/day Sample 1 Sample 2 Sample 3 average
0 -650.2 -624.6 -629.2 -634.7
1 -712.3 -705.2 -704.3 -707.3
3 -714.9 -713.5 -714.3 -714.2
5 -712.5 -709.6 -708.9 -710.3
7 -710.8 -701.7 -702.1 -704.9
14 -719.3 -707.1 -708.2 -711.5

317
Table 37: Corrosion potential values for mild steel specimens after immersion in aerated
0.5 M NaCl solution added to 10% bentonite and 1000 ZnCl2.
Time Sample 1 Sample 2 Sample 3 average
0 -594.8 -607.2 -618.7 -606.9
1 -702.3 -698.5 -698.9 -699.9
3 -694.2 -683 -683.4 -686.9
5 -697.9 -698.1 -698.9 -698.3
7 -698.3 -703.0 -702.9 -701.4
14 -708.3 -709.0 -708.7 -708.66

Table 38: showing Rp in .cm2, icorr in mA.cm-2 and corrosion rate (CR) in mm/y for
3 specimens immersed in the solution level of 0.5M NaCl solution contain 10% (w/w)
bentonite and ZnCl2 at 500ppm for 14 days at natural corrosion potential.
Specimen 1
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 1590.0 0.016 0.178
1 2534.8 0.010 0.111
3 3184.7 0.008 0.089
5 2402.6 0.011 0.123
7 1855.2 0.014 0.156
14 2051.2 0.013 0.145
Specimen 2
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 3681.6 0.007 0.078
1 4616.9 0.006 0.067
3 4902.6 0.005 0.056
5 3725.8 0.007 0.078
7 3226.0 0.008 0.089
14 2801.1 0.009 0.099
Specimen 3
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2518.8 0.004 0.045
1 4475.6 0.006 0.067
3 5026.8 0.005 0.056
5 3641.2 0.007 0.078
7 3168.8 0.008 0.089
14 2829.9 0.009 0.099

318
Table 39: showing Rp in .cm2, icorr in mA.cm-2 and corrosion rate (CR) in mm/y for
3 specimens immersed in the solution level of 0.5M NaCl solution contain 10% (w/w)
bentonite and ZnCl2 at 1000ppm at natural corrosion potential, for 14 days.

Specimen 1
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2363.3 0.010 0.111
1 4361.8 0.006 0.067
3 3473.0 0.008 0.089
5 2784.2 0.009 0.099
7 3098.0 0.008 0.089
14 2892.1 0.009 0.099
Specimen 2
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2309.4 0.009 0.099
1 5293.7 0.005 0.056
3 4540.8 0.006 0.067
5 6127.0 0.004 0.045
7 6227.1 0.004 0.045
14 4386.4 0.006 0.067
Specimen 3
Time (day) LPR .cm2 icorr (mA/cm2) CR (mm/y)
0 2853.7 0.009 0.099
1 5172.6 0.005 0.056
3 4611.5 0.006 0.067
5 5825.4 0.005 0.056
7 5871.3 0.004 0.045
14 4323.5 0.006 0.067

Table 40: Rp values for mild steel specimens after immersing in aerated 0.5 M NaCl
solutions adding 10% bentonite and 500ppm ZnCl2, at OCP for 14 days.

Time Sample 1 Sample 2 Sample 3 average


0 1590 3681.6 2518.8 2596.8
1 2534.8 4616.8 4475.6 3875.7
3 3184.7 4902.6 5026.8 4371.4
5 2402.6 3725.8 3641.2 3256.5
7 1855.2 3226.02 3168.8 2750.
14 2051.2 2801.1 2829.9 2560.7

319
Table 41: Rp values for mild steel specimens after immersing in 0.5 M NaCl solution
adding 10% bentonite and 1000ppm ZnCl2 at OCP for 14 days.
Time Sample 1 Sample 2 Sample 3 average
0 2363 2309.4 2853.7 2508.7
1 4361.8 5293.7 5172.6 4942.7
3 3473 4540.8 4611.5 4208.4
5 2784.2 6127 5825.4 4912.2
7 3098 6227.1 5871.3 5065.5
14 2892.1 4386.4 4323.5 3867.3

Table 42: Current density values in mA.cm-2 for mild steel specimens after immersed in
0.5 M NaCl solution adding 10% bentonite and 500ppm ZnCl2 at -01.15V CP for 14
days.
Time Sample 1 Sample 2 average
0 508.9 589.5 549.2
1 175.5 44.5 110
3 56.3 25.2 40.75
5 29.6 14.5 22.05
7 27.7 16.5 22.1
14 31.4 23.2 27.3

Table 43: Current density in mA.cm-2 for mild steel specimens after immersed in 0.5 M
NaCl solution adding 10% bentonite and 1000ppm ZnCl2 at -01.15V CP for 14 days.
Time Sample 1 Sample 2 average
0 742.4 765.2 753.8
1 60.4 52.2 56.3
3 31.1 23.4 27.25
5 22 19.8 20.9
7 20.1 18 19.05
14 36.617 18 27.31

320
11.7 Appendix VII: Integrated weight loss, obtained from LPR and
experimental (Gravimetric) weight loss for specimens exposed to
0.5% M NaCl solution free of bentonite and adding 10.0% (w/w)
Bentonite.

Table 44: Integrated weight loss obtained from LPR methoand and average
experimental (Gravimetric) weight loss in gram for specimens were immersed in 0.5%
M NaCl solution free of bentonite for 14 days.

Specimen Specimen Specimen Gravimetric


Time/d 1 2 3 Average STDEF (g)
1 0.0026 0.0022 0.00298 0.0026 0.0003 0.006525
3 0.0074 0.00695 0.0084 0.0076 0.0008 0.0183
5 0.0121 0.0127 0.0137 0.0128 0.0008 0.02185
7 0.0168 0.0175 0.0192 0.0178 0.0012 0.0285
14 0.0352 0.034919 0.037544 0.0359 0.0015 0.0702

Table 45: Integrated weight loss from LPR and average experimental (Gravimetric)
weight loss in gram for specimens were immersed in 0.5% M NaCl solution adding10%
(w/w) bentonite for 14 days.

Specimen Specimen Specimen Gravimetric


Time/d 1 2 3 Average STDEF (g)
1 0.0016 0.0018 0.0024 0.0019 0.0004 0.0028
3 0.00498 0.0060 0.0081 0.0064 0.002 0.0134
5 0.0089 0.0110 0.0139 0.0113 0.003 0.0229
7 0.0134 0.0181 0.0199 0.017153 0.004 0.0303
14 0.0299 0.0448 0.0429 0.0392 0.008 0.0570

321

S-ar putea să vă placă și