Download as pdf or txt
Download as pdf or txt
You are on page 1of 116

COPPER DISTRIBUTIONS IN

ALUMINIUM ALLOYS

No part of this digital document may be reproduced, stored in a retrieval system or transmitted in any form or
by any means. The publisher has taken reasonable care in the preparation of this digital document, but makes no
expressed or implied warranty of any kind and assumes no responsibility for any errors or omissions. No
liability is assumed for incidental or consequential damages in connection with or arising out of information
contained herein. This digital document is sold with the clear understanding that the publisher is not engaged in
rendering legal, medical or any other professional services.
COPPER DISTRIBUTIONS IN
ALUMINIUM ALLOYS

T. H. MUSTER
A. E. HUGHES
AND
G. E. THOMPSON

Nova Science Publishers, Inc.


New York
Copyright © 2009 by Nova Science Publishers, Inc.

All rights reserved. No part of this book may be reproduced, stored in a retrieval system
or transmitted in any form or by any means: electronic, electrostatic, magnetic, tape,
mechanical photocopying, recording or otherwise without the written permission of the
Publisher.

For permission to use material from this book please contact us:
Telephone 631-231-7269; Fax 631-231-8175
Web Site: http://www.novapublishers.com

NOTICE TO THE READER


The Publisher has taken reasonable care in the preparation of this book, but makes no
expressed or implied warranty of any kind and assumes no responsibility for any errors or
omissions. No liability is assumed for incidental or consequential damages in connection
with or arising out of information contained in this book. The Publisher shall not be liable
for any special, consequential, or exemplary damages resulting, in whole or in part, from
the readers’ use of, or reliance upon, this material.

Independent verification should be sought for any data, advice or recommendations


contained in this book. In addition, no responsibility is assumed by the publisher for any
injury and/or damage to persons or property arising from any methods, products,
instructions, ideas or otherwise contained in this publication.

This publication is designed to provide accurate and authoritative information with regard
to the subject matter covered herein. It is sold with the clear understanding that the
Publisher is not engaged in rendering legal or any other professional services. If legal or
any other expert assistance is required, the services of a competent person should be
sought. FROM A DECLARATION OF PARTICIPANTS JOINTLY ADOPTED BY A
COMMITTEE OF THE AMERICAN BAR ASSOCIATION AND A COMMITTEE OF
PUBLISHERS.

LIBRARY OF CONGRESS CATALOGING-IN-PUBLICATION DATA

ISBN : 978-1-60741-201-4 (E-Book)

Available upon request

Published by Nova Science Publishers, Inc.  New York


CONTENTS

Preface vii
Chapter 1 Introduction 1
Chapcter 2 Alloy Manufacture 5
Chapter 3 Alloy Microstructure 9
Chapter 4 Electrochemistry 25
Chapter 5 Corrosion 31
Chapter 6 Chemically Pretreated Surfaces 49
Conclusions 83
Acknowledgements 85
References 87
Index 99
PREFACE

Aluminium alloys are used extensively throughout the world, in items such as
decorative architectural applications through fasteners to high strength structural
applications. Such a diverse range of application areas has a similarly diverse
range of requirements for materials properties and performance. The mechanical
properties are achieved through alloying aluminium with a wide range of
elements. Copper, which is one of the major alloying additions, is added in
varying amounts to many of the different aluminium alloy series, with the lowest
levels in the purest wrought aluminium alloys (AA1xxx series) and the highest
levels in the high strength AA2xxx series. The distribution of copper in aluminium
alloys varies from copper atoms dispersed in solid solution through the formation
of clusters of copper atoms and then onto to a range of intermetallic compositions
and particle sizes. The presence of copper in all these forms, particularly in the
AA2xxx series, has a significant impact on the chemistry and electrochemistry of
the surface of the alloy and, hence, on the susceptibility to corrosion and
approaches to metal finishing. This chapter examines how the copper distributions
change, as a result of corrosion reactions, and explores the influence of these
changes on continued corrosion. The influence of the distribution of copper in
aluminium alloys on metal finishing processes and the redistribution of copper as
a result of metal finishing is also examined.
Chapter 1

INTRODUCTION

Aluminium alloys were first developed for commercial use in the mid 19th
century in France [Polmear (1989)]. Since that time there has been considerable
alloy development to produce the vast range of cast and wrought alloys that are
available today.
Aluminium alloys are used extensively throughout many industries. For
example, an examination of the categories of the Aluminium Surface Science and
Technology Conference proceedings from Bonn 2003 indicates applications in
architecture, packaging , electronics, transport, lithography, capacitor foils and
heat exchangers [ASST proceedings (2004)]. Other sources indicate the extensive
usage of aluminium alloys in transport (30%), packaging (18%), building and
construction (21%), mechanical engineering (8%), electrical engineering (9%),
household articles (8%) with 8% assigned to miscellaneous uses [Riotinto
website]. The high strength to weight ratio of alloyed aluminium makes it an
excellent candidate for the transport industry where it is used in train, automotive,
shipping and the aircraft industry sectors.
A large number of alloys have been developed over the years to meet the
requirements of different application areas, and new alloys and heat treatments
continue to be developed as new needs arise. One example is in the automotive
industry, where light weight aluminium casting alloys are replacing ferrous-based
materials for engine heads and blocks. Specific alloys vary from producer to
producer, but generally heads are made from AA319 and its variants, and blocks
are beginning to be made from AA380 and its close variants. A further example is
in the aerospace industry where Aluminium-Lithium-Copper alloys were
developed to replace high strength Aluminium-Copper alloys, which have been
used for nearly eighty years in aircraft manufacture [Bovard (2006)] and possibly
2 T. H. Muster, A. E. Hughes and G. E. Thompson

as long as one hundred years if airships and aluminium alloys of engines in the
first aircraft are included.
Some of these developments are successful whereas others are not
immediately taken up. Aluminium-Lithium-Copper alloys, for example, have
been available commercially since the 1950’s [Polmear (1989)], but there has
been an ongoing reluctance to take up these alloys until there is an improvement
in corrosion performance and fracture toughness [Bovard (2006)]. However, there
are exceptions, and the alloy AA1420 (Al-5Mg-2Li-0.5Mn) has a high corrosion
resistance and has been used successfully on at least one advanced military
aircraft produced in the former Soviet Union [Polmear (1989)]
The aircraft industry uses medium to high strength AA2xxx and AA7xxx
alloys which typically have higher copper contents up to approximately 6 and 3 %
respectively. Recent trends in the aerospace industry are aiming at increased
operational lifetimes [Schmitt (1998), Brown (1992)] with corrosion issues
associated with the airframe becoming a high priority. Older AA2xxx series alloys
such as AA2024-T3 alloy and AA7xxx series alloys such as AA7075-T6 are the,
so-called, legacy alloys [Bovard (2006)]. The corrosion issues related to an
airframe increase markedly with operational lifetime beyond around twenty five
years. Table 1 lists some of the legacy alloys [Bovard (2006)] which, not too
surprisingly, are also some of the most studied alloys in terms of corrosion and
metal finishing. AA2024-T3 is one of the most corrosion prone alloys because of
the high levels of copper, and AA7075-T6 is particularly prone to intergranular
attack [Davis (1999), Hatch (1984)]. The relationship of temper to corrosion
performance should not be overlooked since a change in heat treatment can cause
a significant change in the distribution of alloying compounds and intermetallics.
For example, AA2024-T4 has a very high susceptibility to various forms of
corrosion attack, which can change from intergranular to pitting by changing
quench conditions (which influences the amount of precipitation of solute from
solid solution [Hatch (1984)]. Similarly problems with stress corrosion in
AA7xxx-series alloys can be overcome by heat treating to an over-aged condition
(T7x, where a lower value of x means a greater degree of over-aging).

Table 1. Legacy Alloys

UNS Number Introduction Aircraft


AA2024-T3 1935 Douglas DC3
AA7075-T6 1945 Boeing B29
AA7075-T73 1960 Douglas DC9
Introduction 3

New alloy design has led to the development of alloys that supercede the
more corrosion prone alloys. Some of the new alloys include AA7x5x-T77 alloy
which has been used on the Grumman A-6, AA7058 on the Boeing 777, and
AA7085 on the Airbus A380 [Bovard (2006)]. Improvements in temper have also
led to improved resistance to intergranular attack in the AA7xxx alloys, as the T77
condition suffers only a minimal reduction in mechanical properties compared to
the T6 temper. In the AA2xxx series, AA2190-T8 alloy has been introduced into
the fuselage sheet.
Such a broad range of applications comes with an equally broad range of
service conditions which, in turn, require a wide range of approaches for
aluminium finishing. Some common finishing processes include anodizing and
conversion coating for corrosion protection, adhesive bonding and painting
finishes varying from architectural facades to adhesive bonding in aircraft
applications, to coating for casings for electronic hardware. [Laevers et al. (1993),
Arai et al. (1984), Dunn et al. (1971)].
This chapter considers the influence of copper in corrosion and metal
finishing of aluminium alloys. Copper takes a special place in the role of both
corrosion of aluminium alloys and its metal finishing, because of its
electrochemical properties. Copper is one of the most noble alloying elements
used in aluminium alloy manufacture; hence, it exhibits distinctly different
electrochemical characteristics from the aluminium matrix. Copper-containing
phases on the surface tend to be cathodically protected since they exhibit a net
cathodic reduction reaction:

O2 + 2H2O + 4e- → 4OH- (neutral media) ...1

O2 + 4H+ + 4e- → 2H2O (acidic media) ...2

2H+ + 2e- → H2 ↑ (acidic media) ...3

whilst the aluminium matrix has a net anodic reaction which leads anodic
dissolution via:

Al → Al3+ + 3e- …4

The presence of Cu accelerates the rate of surface electrochemical reactions


in equations 1 to 4. Copper also tends to accumulate on the surface during both
corrosion reactions, and in metal finishing, which further complicates subsequent
surface processing and reaction with the external environment.
Chapcter 2

ALLOY MANUFACTURE

In order to address the materials requirements of the broad range of


applications where aluminium is used, it is alloyed with a range of different
elements. On the basis of the predominant alloying metal, aluminium alloys have
been divided into different series, designated AA1xxx series to the AA8xxx series.
The alloy designations for wrought alloys is defined by a four-number system
known as the International Alloy Designation System (IADS) [Starke and Staley
(1996)]. The first number designates the major alloying addition (Table 2), the
second number refers to modifications of the original alloy or to impurity levels,
and the last two numbers indicate the specific alloy. Table 2 lists the wrought
aluminium alloys according to their series and the predominant alloying elements
with an example of a common alloy for each series and the concentration range of
copper addition for the designated alloy.
While the last two numbers may specify the individual alloy, the designation
simply gives the compositional bounds for the alloy, and individual alloys made
under the same alloy designation and the same temper can have slightly different
compositions. The microstructure of the alloy is, in part, determined by the heat
treatment that the alloy undergoes. Most heat treatments involve solution treating,
where the alloy is taken to a sufficiently high temperature to generate the
equilibrium single, solid phase for the alloy concerned but below the solidus
temperature. For example, the solution treatment temperature in the Al-Zn-Mg-Cu
system varies widely with zinc content and temperature of incipient melting. The
complexity of the phase diagram as a function of zinc and magnesium content for
this system, at the solution treatment temperature, is evident in Figure 1. When
the alloy is quenched, the solute elements are retained in the single phase
aluminium as a supersaturated solid solution, with the exception of insoluble
6 T. H. Muster, A. E. Hughes and G. E. Thompson

intermetallic particles of phases such as those provided in Figure 1. The solid


solution then decomposes during the ageing treatment to produce precipitation
hardening. Large constituent phase such as η (MgZn2) can also form on grain
boundaries during the ageing process. Similarly the Al-Cu-Mg system is complex
and solution treatment needs careful control since it has to be performed a few
degrees below the solidus temperature.

Table 2. Series Designation for Wrought Aluminium Alloys

Alloying Example Heat


Series Common Usage Cu (wt%)
Metal Alloy Treatable
1xxx - AA1100 aluminium foil 0.05 – 0.201 No
food handling /
AA1200 packaging 0.052 No
containers
2xxx Cu, Mg AA2024 high strength 3.8 – 4.91 Yes
3xxx Mn AA3003 beverage cans 0.05 – 0.201 No
AA3104 beverage cans 0.152 No
structural alloys /
AA5005
5xxx Mg automotive and < 0.22 No
decorative
AA5052 automotive < 0.12
structural alloys
AA6060/ < 0.12
6xxx Si, Mg /general purpose/ Yes
automotive
AA6061 0.15 – 0.41 Yes
7xxx Zn. Mg, Cu AA7075 high strength 1.2 – 2.01 Yes
Other
8xxx Including Li and AA8081 automotive 0.7 - 1.31 No
Ni
1
Polmear (1989), 2 Kaufman (2004).

Figure 2 shows the ternary phase diagram for the Al-Cu-Mg system.
AA2024-T3 typically contains 1.2 to 1.8 wt% magnesium and 3.8 to 4.9 wt%
copper which means, if AA2024 alloy was a ternary alloy, S and θ phases should
be present in the α-aluminium matrix. Commercial alloys have additional
elements which change the composition of precipitates so that in AA2024-T3
alloy there is a range of Al-Cu-Mn-Fe containing intermetallics as well as S and θ
phases.
Alloy Manufacture 7

10
α+T
8

6 α+S+T
g
%M

α+T+M+S
4 α+S

2 α+M
α+S+M
α α+T+M

2 4 6 8 10 12
%Zn

Figure 1. Section of the Al-Zn-Mg-Cu phase diagram (1.5% Copper at 460°C): S=


Al2CuMg; T=Al6CuMg4 + Al32(MgZn)49; M= MgZn2 + AlCuMg (after Polmear, 1989).

α+θ+S
3

α+S
% Cu

α+S
2

α+ θ α+S+T
1

α+S
α

1 2 3
% Mg

Figure 2. Section of ternary Al- Mg-Cu phase diagram (1.5%Copper at 460°C) S=


Al2CuMg, θ =Al2Cu, α = solid solution (after Polmear, 1989). The shaded section of the
diagram is at 460°C whereas the other phases are for 190°C.
8 T. H. Muster, A. E. Hughes and G. E. Thompson

Table 3. Temper Designations

Alloy Basic Secondary


Description of Secondary Treatment
Type Treatment Treatment
Non Heat
H 1 only cold worked
Treatable
“ 2 cold worked and partly annealed
“ 3 cold worked and stabilised
partial solution treatment followed by
Heat Treatable T 1
natural ageing
“ 2 Annealed cast products only
“ partial solution treatment followed by
3
natural ageing
solution treatment followed by natural
“ 4
ageing
“ 5 artificially aged
solution treatment followed by artificial
“ 6
ageing
“ 7 solution treatment and stabilisation
solution treatment,
“ 8
cold work, followed by artificial ageing
solution treatment followed by artificial
“ 9
ageing and cold worked

Solution treatment is followed by quenching, usually into cold water, to


achieve the maximum supersaturation of the alloying components. Quenching
does, however, introduce residual stresses into the product (particularly sheet
product) which can be reduced by stretching or roller levelling. From this point,
wrought aluminium alloys can be divided into heat treatable alloys, which
improve their mechanical properties with subsequent heat treatments, and non-
heat treatable alloys, which develop their strength through strain hardening. This
latter group includes the Al-Mg, Al-Mn and Al-Mg-Mn alloys. The mechanical
properties of heat-treatable wrought alloys can be manipulated using artificial
aging treatments, as described above. Artificial aging can also influence the
corrosion performance of a particular alloy, and its surface response to metal
finishing treatments. The temper designation for both classes of alloys are listed in
Table 3. Slow cooling can also be used, but this depends on the individual alloy
since there may be a propensity for the formation of large undesirable
intermetallic particles.
Chapter 3

ALLOY MICROSTRUCTURE

BULK MICROSTRUCTURE
To understand how copper promotes corrosion in aluminium alloys and why
it accumulates at the surface of alloys during metal finishing processes it is
instructive to examine the microstructure of aluminium alloys, particularly those
with higher copper contents. Traditionally, the “microstructure” has, by default,
referred to the bulk microstructure, which is the focus of this section, but equally
important is the surface microstructure (see following section), since this is the
interface where reactions of relevance to metal finishing or corrosion commence.
As noted previously, there are two broad classifications for wrought
aluminium alloys; non-heat treatable and heat treatable [Hatch, (1984)]. Non-heat
treatable alloys, which obtain most of their strength through solid solution
hardening and strain hardening, contain major additions of chromium, iron,
magnesium, manganese, silicon and zinc, whilst only minor additions of copper
are permitted (i.e. 0.12 – 0.15 wt% in can stock alloys AA3003 and AA3104,
AA1100, and 1 wt% in AA8280 and AA8081 alloys). Heat treatable aluminium
alloys can contain increased levels of copper (up to 6.3 wt%) [Hatch (1984)].
Ultimately, these alloying elements are present in either solid solution in the
matrix, intermetallic particles, or both. Copper, together with magnesium, zinc
and silicon, are appreciably soluble at high temperature and considerably less
soluble at low temperature. This results in the precipitation of various phases
during solidification of the alloy [Hatch (1984)].
For the precipitate-hardened alloy, the mechanical properties are improved by
the precipitation of alloying components from solid solution. These fine
precipitates often start as clustering of alloying components called Guinier-
10 T. H. Muster, A. E. Hughes and G. E. Thompson

Preston (GP) zones that grow into a range of precipitates with increasing
temperature and time. Initially, these precipitates are coherent with the aluminium
lattice, which is desirable, but continued ageing will take the precipitates through
degrees of coherency until they are so large that they become incoherent with the
lattice. These phases are usually identified using greek letters and the degree of
coherency is identified using “primed (‘)” symbols. For example, in the case of
Al-Cu binary alloys where the θ-phase forms (Al2Cu), the stages are GP → θ” →
θ’ → θ with each step indicating an incremental loss of coherency with the
primary matrix. Table 4 indicates some of the typical hardening precipitates.
A second class of precipitates are the dispersoids that form by solid state
reaction during preheating of the ingot; they represent an important part of the
alloy microstructure since they control grain growth. They are formed through
precipitation with either chromium, manganese, titanium or zirconium and form
dispersoids particles such as Al12CrMg2, Al20Cu2Mn3, Al12Mn3Si, Al3Ti and
Al3Zr. These particles are usually a few nanometers up to 200 nm in size [Starke
and Staley (1996)].

Table 4. Compositions of Hardening Precipitates

Alloy Type Precipitate Composition Precipitate Nomenclature


Al-Cu Al2Cu θ
Al-Mg Al8Mg5 β
Al-Si Si -
Al-Cu-Mg Al2CuMg S
Al-Mg-Si Mg2Si β
Al-Zn-Mg MgZn2 η
(Al,(Cu,Zn))49Mg32 T
Al-Li-Mg Al3Li δ
Al2LiMg -
Al-Li-Cu Al3Li δ
Al2CuLi T1

A third class of particles are the constituent particles, most of which are
formed during solidification of the initial ingot. As stated above, at low levels,
copper is present in solid solution in the matrix (α-Al). The AA2xxx, AA7xxx and
AA8xxx series alloys also have copper in solid solution as well as being
incorporated into a range of intermetallic phases called constituent particles. As
noted previously, the microstructure of these alloys is complex and depends on
thermal and ageing treatments. Common constituent particles and the alloys they
appear in are listed in Table 5 [Stake and Staley (1996)].
Alloy Microstructure 11

Table 5. Some Typical Constituent Particles found in Wrought Al-Alloys

Alloy Constituent Particles


2X24 Al7Cu2Fe, Al12(Fe,Mn)3Si, Al2CuMg Al2Cu , Al6(Cu,Fe)
2X19 Al7Cu2Fe, Al12(Fe,Mn)3Si, Al2Cu)
6013 Al12(Fe,Mn)3Si
7X75 Al7Cu2Fe, Al6(Fe,Mn), Al12(Fe,Mn)3Si, Mg2Si
7X50 Al7Cu2Fe, Al2CuMg, Mg2Si
7055 Al7Cu2Fe, Mg2Si
2090 Al7Cu2Fe
2091 Al7Cu2Fe, Al3Fe, Al12Fe3Si
2095 Al7Cu2Fe , Al2CuLi, Al6CuLi3
8090 Al3Fe

A simple summary of the microstructure of aluminium alloys is not possible


since it varies considerably from series to series, alloy to alloy and even from
temper to temper. General summaries of the alloy microstructure has been given
elsewhere [Hatch (1984), Vander Voort (2004)] and are not repeated here except
for consideration of the distribution of copper in the respective alloys.

AA1xxx Series Alloys

The AA1xxx series include high and super purity aluminium and generally
only contains impurity levels up to 1% of iron and silicon as major impurities.
The types of intermetallic particles that form include Al3Fe and silicon particles.
AA1xxx series alloys are commonly used as cladding for AA2024-T3 alloy in the
aircraft industry to provide protection for the more corrosion prone AA2024-T3
alloy.

AA2xxx Series Alloys

The AA2xxx series, containing copper and magnesium, are high strength
aluminium alloys and are therefore often used in applications which require such
strength i.e., aircraft manufacture. They have a high damage tolerance and fatigue
resistance.
The mechanical properties of the AA2xxx series alloys are determined by the
ternary Al-Cu-Mg phase diagram. In Al-Cu-Mg ternary systems that fall in the α
+ S phase region (Figure 2), the precipitation of Al-Cu-Mg particles occurs in the
12 T. H. Muster, A. E. Hughes and G. E. Thompson

following sequence; Copper in solid solution with aluminium (α-Al), first


precipitates out as clusters of solute elements before GP zones form as fine rods in
the <001>α directions. Much later in the precipitation sequence, or more usually,
where cold work is applied, S-phase precipitates form as laths within the
microstructure on {012}α planes in <001>α directions. The development of these
types of phases for AA2024-T3 alloy can be followed using Cu63 NMR as shown
in Figure 2 (right), where copper in solid solution is clearly distinguished from the
precipitated forms of copper such as the S-phase [Bastow (2003, 2005, 2006),
Nairn (2006)]. The development of different forms of precipitate can also be
followed, and quantified, as shown in Figure 2 (right) for the binary Al - 4 wt%
Cu.
S

GPZ
α α

θ''

4000 3000 2000 1000 6000 4000 2000 0


ppm ppm

Figure 2. 63Cu NMR spectra. Left - of Al-4 wt% Cu solution treated, quenched and aged at
130oC for two hours showing the development of θ’-precipitates within the alloy. Right -
AA2024 solution treated and quenched and aged at 177oC for two hours showing the
development of S-phase (Courtesy of T. Bastow).

Equally, electron microscopy can be used to follow the formation of


precipitates in aluminium alloys. θ’ precipitates, in the quaternary Al-Cu-Mg-Ag
alloy aged to the T6 condition, can be seen, viewed edge-on, along the [001]α
plane in Figure 3(a), using transmission electron microscopy. The dominant Ω
phase provides the mottled appearance, since it is inclined to the plane in which
the image of the θ’ precipitates has been taken. The backscattered image in Figure
3(b) shows a triple point junction between three grains in an Al-Cu binary alloy.
Alloy Microstructure 13

The θ’ precipitates can be seen in the individual grains as well as the larger θ
precipitates in the grain boundaries. The precipitation of the θ precipitates within
the grain boundary has led to copper depletion in the vicinity of the grain
boundary. As will be shown later these small changes in copper distribution have
significance for corrosion performance.

100 nm 200 nm

(a) (b)

Figure 3. (a) Transmission micrograph of Al-Cu-Mg-Ag alloy aged to the T6 condition


showing θ’ precipitates as plates viewed edge on. A small amount of the minor phase (Σ) is
observed as cuboids, and the dominant Ω-phase is present on [111]α incline planes. Imaged
along [001]α (courtesy of Dr. R Lumley). (b) Backscattered scanning electron micrograph
of a Al-Cu binary alloy showing θ’ precipitates in the matrix and θ precipitates in Cu-
depleted grain boundaries (provided by Professor G. Thompson).

In corrosion and metal finishing, the most widely studied, copper-containing


alloy in the AA2xxx series alloys is sheet AA2024-T3, although AA2014-T3 alloy
is becoming more prominent in aircraft manufacture, and therefore a subject of
research, [Starke and Staley (1996), Henon (2006) Smith (1993)]. For AA2024-
T3 sheet, total constituent particle number densities have been reported from
300,000/cm2 [Chen et al. (1997)] to 530,000/cm2 [Juffs (2003), Hughes et al.
(2006)] for polished surfaces and as high as 11,700,000/cm2 for the rolled surface
[Juffs (2003), Hughes et al. (2006)]. However, for rolled and polished surfaces,
the surface area occupied by intermetallic particles was similar, suggesting that
rolling leads to significant breakup of intermetallic particles. This is also reflected
in the average particle size which was much smaller for the rolled surface than the
polished surfaces. Particle size distributions for the intermetallic particles in
AA2024-T3 have been reported by Jakab et al. (2005) and Hughes et al. (2006)
for polished surfaces, with slightly different distributions revealed, but with
similar volume fractions of intermetallic particles (Table 5).
14 T. H. Muster, A. E. Hughes and G. E. Thompson

Table 5. Intermetallic Particles Distributions in AA2024-T3 Alloy

Average Particle % Surface


Source
Size (μm) Area
Polished
Hughes et al. [2006), Juffs (2003) 6.6 2.89
Jakab et al. (2005) 3.1 2.18
Rolled
Hughes et al. [2006), Juffs (2003) 2.0 2.82

It is not clear whether the difference in the particle number density between
300,000 and 530,000/cm2 represents a significant variation. Certainly there will be
batch variation and probable processing effects; further the particle population
densities will depend on the resolution of the techniques used for the counting
statistics. Another possibility for the differences in the published figures is the
processing history of the alloy. Specifically, for sheet alloy, the gauge (or
thickness) reflects the number of rolling passes that the alloy undergoes. Clearly,
at each pass the potential exists for further breakup of intermetallic particles and
changes in the intermetallic size and spatial distributions and grain refinement
(see section on Surface Microstructure). Examination of cross sections of
AA2024-T3 alloy revealed that the distribution of intermetallic particle density
across the sheet can change significantly, as depicted in Figure 4 [Juffs (2003)].
The variation in particle density is accompanied by an increase in particle size
towards the centre of the sheet; this is reflected in the larger particle size on the
polished surfaces (toward the sheet centre) versus the rolled surface. The
characteristics of intermetallic particle distributions is an area which warrants
further investigation since second phase particles are often sites of corrosion
initiation and the influence of clustering of these particles is largely unknown.
Focusing on the larger intermetallic particles, Buchheit et al. (1997) reported
that roughly 60% of the constituent particles of particle diameter exceeding 0.2
μm were Al2CuMg (S-phase). The remaining 40% of intermetallics comprised a
range of Al-Cu-Fe-Mn containing phases. The composition of Al-Cu-Fe-Mn
phases has been suggested to take various forms. Gao et al. (1998) suggest
compositions based upon (Al,Cu)x(Fe,Mn)ySi such as modified forms of Al8Fe2Si
or Al10Fe2Si type intermetallics, although in low silicon–containing AA2xxx
series alloys, these compositions are different. For example, Buchheit et al. (1997)
reported that of the remaining 40% of intermetallic particles, the most notable
included Al7CuFe2, Al6MnFe2, (Al,Cu)6Mn, and a number of undetermined
compositions in the class Al6(Cu,Fe,Mn) where the Cu:Fe:Mn ratios were
Alloy Microstructure 15

approximately 2:1:1. The Al-Cu-Fe-Mn particles consistently exhibit cross-


sectional diameters in the range 10 to 50 μm, possess a high hardness, and are
generally irregular in shape [Liao and Wei (1999)]. Further Scholes et al. (2006)
reported that this class of intermetallic particles underwent fracture during
milling, whereas the S-phase particles remained largely intact.

100

90
Particle Count

80

70

60

50

2 4 6 8 10 12 14 16
Position Across Sheet

Figure 4. Intermetallic Particle Count taken on frames across a section of AA2024-T3 with
a thickness of 1.2 mm. The sample was mounted in bakelite and polished down to 1 μm
(after Juffs, 2003).

There is emerging interest in the spatial relationship of intermetallic particles


in surfaces. In an extensive study of clustering, Juffs (2003) examined several
methods of determining clustering of intermetallic particles in aluminium alloys.
One of the most sensitive methods was the pair correlation function in which the
average number of nearest neighbours is determined as a function of distance
from the average particle. Juffs (2003) observed clustering in AA2024-T3 alloy
for both polished and rolled surfaces; indeed the number of nearest neighbours
was more than double that expected on a polished surface with a random
distribution of intermetallic particles, and a little under twice as many for the
rolled surface. On the other hand Jakab et al. (2005) found no significant
clustering in AA2024-T3 alloy. The statistical sampling between the two studies
may explain the differences. In the former study, Juffs (2003) counted several
thousand particles, whereas, Jakab et al. (2005) did not indicate the number of
16 T. H. Muster, A. E. Hughes and G. E. Thompson

particles counted, although it appeared to be less than one hundred. As will be


further detailed in the section on corrosion, clustering may prove to be an
important issue for pit initiation and is a promising area for further research.
At the submicron scale of the alloy microstructure, there is an even
distribution of Al20Cu2Mn3 dispersoids. Guillaumin and Mankowski (1999) have
reported that the coarse S-phase intermetallic particles are surrounded by a
dispersoid free zone. However, Buchheit et al. (1997) suggested that only those
particles that precipitate after secondary solution heat treatment will have a
precipitate free zone surrounding them. At the finest scales there are lenticular
particles around 100 nm in length which comprise the Al2CuMg hardening
precipitates. More generally in AA2xxx alloys, the hardening precipitates of θ-
phase (Al2Cu) and S-phase (Al2CuMg), depend on the copper to magnesium ratio
[Hatch (1984)].

AA3xxx Series Alloys

The AA3xxx series is formed by alloying with manganese at levels between 1


and 1.25 % and is dispersion-hardened through the presence of Mn-containing
particles (Al6Mn). According to Hatch (1984) AA3003 alloy is the only AA3xxx
series of commercial interest. The larger constituent particles can generally be
divided into two classes which vary in their relative number according to the alloy
composition (including impurities such as iron and silicon which are typically at
levels of 0.7 and 0.3 to 0.6, respectively). These impurity levels are also higher
than that for the many of the AA1xxx series alloys. Afseth et al. (2001), for
example, reported that 60% of the constituent particles in AA3005 alloy were
Al6(Fe,Mn) and the remaining particles were α- Al12(Fe,Mn)3Si. The AA3xxx
series of alloys has gained some attention in recent years due to its enhanced
filiform corrosion susceptibility which is largely related to the near-surface
microstructure. This is discussed in the next section.

AA4xxx Series Alloys

The AA4xxx series is formed by alloying with silicon at levels up to 13 wt%


and may contain low levels of copper (typically 0.3%). These alloys are often
used in brazing allplications [Hatch, 1984]. AA4047, for example is used as a
cladding for AA3005 but according to Hatch (1984) has no copper. Because of
impirty iron in the alloy β-AlFeSi can form as well as silicon particles.
Alloy Microstructure 17

AA5xxx Series Alloys

These alloys, based on alloying with magnesium, have some of the best
corrosion resistance of all aluminium alloys as well as relatively good strength
[Polmer (1989), Vander Voort (2004)]. The alloys are often used for welding
applications, and are usually manufactured as plate [Polmear (1989)]. Because of
their corrosion resistance, they are often employed in marine manufacture such as
small craft or ship superstuctures. Magnesium contents range from as little as
0.8% and up to 5% for wrought alloys. Apart from the typical impurity phases
such as those related to silicon and iron, the main phases are Al3Mg2 and β-phase
(Al8Mg5). With significant levels of either silicon, copper or zinc, hardening
precipitates such as Mg2Si, Al2CuMg and Al2Mg3Zn3 can also form. Very small
levels of other elements are added to AA5xxx series alloys, such as chromium
which forms Al18Mg3Cr2 dispersoids. [Vander Voort (2004)].

AA6xxx Series Alloys

The 6xxx series alloys are alloyed with both magnesium and silicon which are
usually added in a ratio whereby Mg2Si form by precipitation from solid solution
or silicon is added in excess. These alloys gain their strength by heat treatment
and precipitation of the Mg2Si phase. The most common form of production of
AA6xxx series alloys is as extrusions where Si is added between 0.8 and 1.2%
[Polmear (1989)] and quenching immediately from the extrusion die means the
alloy only require subsequent low temperature ageing (e.g. 180°C) to improve
mechanical properties.
As with other alloy classes, copper can be added ; the medium strength alloys
AA6013, AA6056 and AA6111, for example, have up to about 1% copper. This is
to enhance precipitation hardening, reported to be a result of Q-phase formation.
Alloy AA6111 finds major use as an automotive bodysheet alloy, having a good
combination of formability and strength. The AA6xxx series alloys containing
copper, do, however, have inferior corrosion properties to copper-free AA6xxx
series alloys.

AA7xxx Series Alloys

The AA7xxx series alloys system is based on the ternary Al-Zn-Mg system
but has copper included to alleviate severe problems with stress corrosion
18 T. H. Muster, A. E. Hughes and G. E. Thompson

cracking. The quaternary alloys also display an improved response to age-


hardening. Some AA7xxx series alloys are considered to be weldable and are used
extensively in transport. These alloys typically have 3 to 7 % zinc and 0.8 to 3.0%
magnesium. Chromium, manganese and zirconium are also added for grain size
control. The two phases that form in wrought Al-Zn-Mg alloys, through eutectic
decomposition, are Mg2Zn and Al2Mg3Zn3 [ASM (2004)]; the dispersoid phase is
Al12CrMg2.
The AA7xxx series alloys are well known for their susceptibility to stress
corrosion, which is due to grain boundary precipitation of η-phase (Zn2Mg) and
depletion of the adjacent grains [Pickens and Langan (1987)]; it will not be dealt
with in this chapter. The addition of copper can reduce the susceptibility to stress
corrosion cracking.
The Al-Zn-Mg-Cu alloys possess some of the highest tensile strengths and
other mechanical properties of all aluminium alloys due to their age hardening
properties [Polmear, 1989]. Because of their excellent mechanical properties, the
alloys have been used extensively in aircraft manufacture and AA7075 series,
particularly in the T6 or T73 condition, have been used most extensively (Table
1). T73 is a duplex aging that creates an overaged microstructure. AA7050 alloy
is a further important structural alloy and it has increased levels of copper to assist
age hardening.
Birbilis and Buchheit (2005) include the following intermetallics for AA7075
in the T6 condition: Mg2Si, MgZn2, Al12Mn3Si, Al7Cu2Fe, Al2Cu, Al2CuMg,
Al3Fe, Al12CrMg2, Al20Cu2Mn3, Al6Mn, Al3Ti, Al6Zr, Al3Mg2, Al32Zn49,
Mg(Al,Cu). (While the most common intermetallic particles are listed in Table 5,
a more detailed analysis of the intermetallic particles in a complex alloy like
AA7075-T6 will reveal a much more extensive list of particles.) Clustering of
constituent intermetallic particles in 7075-T6 on the as-received surface was also
studied by Juffs (2003) using the radial distribution function. The highest degree
of clustering occurred in this alloy when compared to other alloys (including
AA2024-T3), with nearly 4 neighbours expected within a radius of 5 μm.

SURFACE MICROSTRUCTURE
The nature of surface microstructure has historically emerged from
tribological studies [Fishkis and Lin (1997), Schey (1983)] and, in recent years,
has increasingly been addressed as part of filiform corrosion susceptibility [Asfeth
(2001), Leth-Olsen (1997), Mol et al. (2002)]. The surface microstructure is often
more complex than the corresponding bulk microstructure. At the most
Alloy Microstructure 19

fundamental level, magnesium, lithium and silicon typically segregate to the


external surface during heat treatment of both cast and wrought alloys. Textor and
Amstutz (1994) reported that magnesium and, particularly lithium, are enriched in
the surface oxide by 2-4 orders of magnitude compared to their bulk
concentration. Internal segregation to grain boundary interfaces and the
development of internal depletion zones was considered briefly in the previous
section. Segregation to the external surface occurs via the two routes of (i) bulk
diffusion and (ii) grain boundary segregation. In general, the surface enriched
elements have a high negative free energy for oxide formation and high diffusion
coefficient in aluminium metal [Textor and Amstutz, (1994)]. However, the
tribology of forming processes such as rolling and extrusion add an extra and
complex dimension to the nature of surface layers [Fishkis and Lyn (1997)].
The study of segregation phenomena in metallic alloys goes back many years
and there have been extensive studies on the theory of segregation [Seah (1980),
Hondros and Seah (1977), Du Plessis and Tagauer (1992), Luckman (1988),
Darken (1967), Hofmann (1987), Du Plessis and van Wyk (1988) and Guttmann
(1975)] These theories deal extensively with binary alloys, with some
consideration of ternary alloys, but do not include the influence of phase
precipitation or precipitation in stacking faults. Nevertheless, they provide a basic
understanding of the kinetic and thermodynamic driving forces for segregation to
surfaces. Carney et al. (1990), for example, suggested that magnesium enrichment
in binary alloy powders is in accordance with the diffusion rate of magnesium in
molten aluminium. However, mass loss due to evaporation at high temperatures is
higher than the rate of accumulation due to segregation and magnesium depletion
can occur. Lea and Molinari (1984) and Viswanadham et al. (1980) showed that
the maximum segregation occurred at around 475°C. Even on polished surfaces,
the surface oxide on Al-Cu-Fe-Mn intermetallic particles apparently forms by
preferential oxidation of aluminium in the intermetallic phase [Roberts et al.
(2002)].
In addition to heat treatments providing a driving force for segregation,
mechanical processing can significantly change the surface microstructure of the
alloy. In the last ten years there has been renewed interest in the reasons for
increased filiform corrosion in north western Europe. The increase in filiform
susceptibility was due to the incorporation of lower etch rate processes in the
metal finishing of architectural alloys. The lower etch rates failed to remove the
surface modified layer produced by rolling and related heat treatments, which may
be rendered extremely electrochemically active.
Multiple-pass rolling causes considerable modification to the surface of
aluminium alloys creating new surface layers with a very fine grain structure and
20 T. H. Muster, A. E. Hughes and G. E. Thompson

incorporated oxide (Figure 5) [Fishkis and Lin (1997), Afseth et al. (2001)]. This
type of transformation is called a Grain Refined Surface Layer (GRSL) [Leth-
Olsen (1998)]. The surface layers are characterised by a high porosity, very fine
grain structure and large oxide content. Oxides that have been detected in the
surface include γ-Al2O3, MgO and the spinel phase MgAl2O4. The latter is only
formed above 350°C [Fishkis and Lin (1997), Scamans and Butler (1975) Pronko
et al. (1988), M. Pijolat et al., (1988), C. Lea and J. Ball (1984), S.K. Toh et al.,
(2003)], although Lumley et al. (1999) reported its formation at 275°C. The
mechanism of modification proposed by Fishkis and Lin (1997) was a three-step
process involving:

1. Formation of surface cavities by plowing, adhesive wear, delamination


wear or transverse surface cracking,
2. Filling of the cavities with wear debris, including oxide, metal fines and
lubricants,
3. Covering the cavities with thin metal layers during continuing rolling,
leading to a shingled surface appearance.

Oxide Fragments
Surface Oxide
GRSL

Intermetallic
Bulk Metal
Particles

Figure 5. Schematic diagram of the restructured layer as a result of mechanical work such
as rolling including the incorporation of oxides particulates and a recrystallised zone. The
recrystallised layer is called the Grain Refined Surface Layer (GRSL).
Alloy Microstructure 21

There are various reports on the depth of the modified surface region. Fishkis
and Lin (1997) reported that the recrystallised surface layer changed from a depth
of 8 μm after a first pass roll to 3 to 5 μm on subsequent rolling. Afseth et al.
(2001) examined a rolled sheet of AA3005 alloy and revealed a deformation layer
about 1μm thick. Both studies suggest that recrystallisation results in a very fine
surface grain structure of dimensions down to 40 nm at the outer surface and up to
400 nm elsewhere in the recrystallised zone. Similar results were reported by
Leth-Olsen et al. (1998) for AA8006, AA3005 and AA1xxx series alloy.
Subsequent heat treatment of the AA3005 can result in precipitation of
manganese-containing particles which renders the surface layer extremely surface
active [Afseth (2001), Scamans et al. (2003)]. Milling also produces changes in
surface structure as reported by Scholes et al. (2006), where crushing of Al-Cu-
Fe-Mn type intermetallics to a depth of around 4 μm from the surface of milled
AA2024-T3 alloy was observed. There was also significant folding of the matrix
alloy creating subsurface crevices up to 5μm long and a few microns depth.
In both these instances, the GRSL is likely to be in a metastable state with
respect to ageing and changes in the oxide composition. The surface oxide can
change as a result of exposure to the environment. Viswanadham et al. (1980)
studied changes to the magnesium enriched surface oxide of an Al - 5.5 wt% Zn-
2.5 wt% Mg alloy. They observed an increase in aluminium oxides on the surface
on the magnesium oxide after storage in various moist environments at 50°C.
They attributed this enrichment to aluminium diffusion from the underlying alloy
through grain boundaries in the surface magnesium oxide onto the external
surface.
A further impact of rolling on the surface microstructure relates to the
intermetallic particles phases. Lunder and Nisancioglu (1987) reported that during
mechanical processing, i.e., rolling, the large intermetallics in the surface are
covered with a layer of the matrix aluminium alloy. Rolling also has a mechanical
impact by breaking up the intermetallic particles as shown in Table 5. Particle
counting of the two main classes of intermetallic particles in AA2024-T3 (S-phase
(Al2CuMg) and Al-Cu-Fe-Mn-containing intermetallics) indicated that the
number density on the rolled surface was twice as high and the average size was
about one third (Table 5), suggesting intermetallic breakup [Hughes et al. (2006)].
Milling was also observed to result in the fracture of intermetallic particles and
covering of the intermetallics particles with the aluminium alloy matrix [Scholes
et al., (2006)]. Although in this case it was reported that the S-Phase particles
remained largely intact. From studies of a range of aluminium alloys, Lunder and
Nisancioglu (1987) have also observed that the alloy matrix can cover the
intermetallic particles in rolled surfaces.
22 T. H. Muster, A. E. Hughes and G. E. Thompson

(a) (b)

Figure 6. EELS Maps from AA7475-T6 alloy: (a) aluminium (blue), magnesium (green)
and oxygen (red); (b) copper (red), zinc (green) and iron (Blue). Color rules: R + G =
Yellow, R + B = Magneta, G + B = Cyan, R + G + B = White.

(a) (b)

Figure 7. EELS Maps from AA2024-T3: (a) oxygen/aluminium/copper and (b)


oxygen/aluminium/magnesium. Color rules: R + G = Yellow, R + B = Magneta, G + B =
Cyan, R + G + B = White. See text overlay on figure for colour assignments.

These studies indicate that, generally, while the surface microstructure and
composition may be understood, individual treatments may result in considerable
variation in the surface layers. Examples of the surface oxide are given in three-
colour maps in Figure 6 for AA7475-T6 where it is revealed that the surface layer
Alloy Microstructure 23

varies considerably in thickness and incorporates the matrix metal (bottom left)
folded into an oxide which varies in thickness from 250 to 500 nm. The three-
colour map of the copper, zinc and iron elemental maps shows that iron-
containing particles are incorporated into the surface oxide. The chattering
associated with mechanical damage due to ultramicrotomy, changes in moving
from the bulk to the surface but these changes are deeper than the oxide coating
and indicate the depth of the GRSL.
Figure 7 shows a thin continuous layer for the surface oxide for AA2024-T3
alloy with considerable incorporation of magnesium oxide. Copper oxide particles
are also present in the surface oxide layer. These particles are of similar size to the
Al20Cu2Mn3 dispersoid phase, but appear to be copper oxide since neither
aluminiumn nor manganese was detected in them; the origin of these particles is
not clear. This example demonstrates the complex nature of the surface oxide.
In summary, the surface of aluminium alloys may have a complex structure
and composition that depends on the processing history and the storage
environment. It is the objective of metal finishing processes to remove such layers
to produce a surface which has a well defined, reproducible structure and
chemistry, thus minimizing the history of the alloy on its subsequent performance
during coating processing and its performance in-service.
Chapter 4

ELECTROCHEMISTRY

This section focuses on the reactivity of aluminium alloy surfaces, and the
influence of copper with regards to electrochemical phenomena, particularly
corrosion. Aluminium is thermodynamically unstable, appearing highly negative
on the electrochemical series (standard reduction potential = -1.42 VSCE).
However, aluminium owes its, often exemplary, corrosion resistance to its ability
to form a passivating surface oxide layer. Under most conditions, this surface
oxide is able to form and, if damaged, can easily reform [Davis (1999), Hatch
(1984)]. The presence of the passivating film allows aluminium to achieve
potentials in the proximity of -0.75 VSCE in aqueous solution (Table 6). The oxide
formed at the surface of pure aluminium is composed of two layers. A thin
(usually < 5 nm) compact amorphous passivating film is formed adjacent to the
metal. This film has variously been reported as either γ-Al2O3 [Pryor, (1971)], γ’-
Al2O3 [Wilsdorf (1951)] or glassy [Fehlner and Mott, (1970)]. A thicker,
hydrated, and porous oxide forms the outer surface layer. This thicker layer is
generally designated as Al2O3.nH2O, and may be composed of aluminium
hydroxides or oxyhydroxides, depending on the formation conditions [Alwitt
(19740, Vedder and Vermilyea (1970), Davis (1999)].
When aluminium or its alloys are exposed to aqueous solutions, the protective
oxide film may breakdown, resulting in corrosive attack of the underlying metal.
At near-neutral pH the solubility of aluminium oxides is low (solubility constant
<10-32); however, the stability of aluminium oxides is highly dependent upon the
pH of the environment (Hemingway et al., 1991). Thermodynamically,
aluminium oxides are stable between pH 4 – 9 (Pourbaix, 1966) and increasingly
unstable at high and low pH. Copper oxides tend to be stable at high pH, and
become increasingly soluble at low pH (Figure 8).
26 T. H. Muster, A. E. Hughes and G. E. Thompson

5
Boehmite
Al(OH)3
4
Log (soluble species) [M]

CuO
Cu(OH)2

0
0 2 4 6 8 10 12 14
pH

Figure 8. Solubility of Al2O3. H2O (Boehmite), amorphous Al(OH)3, CuO and hydrated
CuO in distilled water as a function of pH (data from Pourbaix, (1966)).

In electrochemical systems, copper exhibits increased nobility in comparison


with aluminium (standard reduction potential of +0.24 VSCE). The nobility of a
metal is a direct reflection of its Gibbs free energy, which can be measured in
terms of the work function (the energy required to remove an electron from the
bulk of the metal into a vacuum). These are 4.17 ± 0.09 and 4.76 ± 0.23 eV for
aluminium and copper, respectively, (Lide, 2001). When copper is alloyed with
aluminium a common Fermi level is achieved, and the additional energy required
to remove an electron from copper is maintained as a difference in the surface
potential (in air) and a solution potential when immersed in an electrolyte (Muster
and Hughes, 2006). As a result, the introduction of copper into aluminium alloys
generally increases the solution potential in comparison with pure aluminium
(Table 6). However, adding active metals, such as magnesium and zinc, tends to
decrease the nobility of aluminium alloys. Figure 9 provides a summary of the
effects of major copper and magnesium additions on the solution potentials of
aluminium alloys. Up to a concentration of 4 wt%, the addition of copper to
aluminium increases the solution potential by approximately 37 mV per percent
copper. Further increasing the copper content (from 4 to 7 wt%) has limited effect
on the solution potential (Figure 9, Table 6).
Electrochemistry 27

- 0.70
Potential (V) Cu

- 0.78

- 0.86 Mg

2 4 6
Alloying element (wt %)

Figure 9. Effects of copper and magnesium alloying elements on electrolytic solution


potential of aluminium. Potentials (0.1N Calomel scale) are for high-purity binary alloys
solution heat treated and quenched, then measured in a solution of 53 g/l NaCl plus 3 g/l
H2O2, after Davis (1993).

Table 6. Solution potentials of relevant metals, alloys and secondary phases


determined according to ASTM G69 (Davis, 1999)

Metal Solution potential (VSCE) % Cu


Aluminium (99.999%) -0.75
Copper (99.999%) +0.00
Magnesium -1.64
Zinc -0.98
2024-T3 -0.60 3.8 - 4.9
6061-T6 -0.74 < 0.1
7075-T6 -0.74 1.2 - 2.0
7475-T7 -0.75 1.2 -1.9
8090-T3 -0.70 1.0 -1.6

In addition to determining electrochemical potentials, the Gibbs free energy


of alloying elements also controls the enrichment of elements at the alloy surface
and in the surface oxides during corrosion processes [Skeldon et al. (1995), Zhou
et al. (1999)]. Copper and other more noble elements (i.e. gold) have high Gibbs
free energies of oxide formation per equivalent (ΔGº/n) relative to that of alumina
28 T. H. Muster, A. E. Hughes and G. E. Thompson

and therefore show extensive enrichment at the metal/oxide interface. In contrast,


magnesium, for example, has a lower ΔGº/n value than aluminium and, therefore,
is more likely to appear in the oxide following corrosion processes [Zhou et al.,
1999]. In instances where copper ions are released from the metal surface during
electrochemical processes, the proportion of copper in the oxide, relative to its
bulk alloy concentration, is decreased, and under most conditions oxidized copper
is undetectable at the oxide/electrolyte interface [Zhou et al., 1999]. It follows that
copper ions have an increased mobility through the surface oxides in comparison
with aluminium ions, and that the Cu2+-O bond energy is lower than that of Al3+-
O. This allows copper ions to migrate more rapidly from the metal/oxide
interface, through the oxide and into the electrolyte [Zhou et al., 1999].
Where copper is present at the surface of aluminium alloys, it shows vastly
different behaviour to aluminium. The surface of copper is particularly efficient at
supporting cathodic reactions (e.g., oxygen and water reduction). Limiting
cathodic currents measured for pure copper and pure iron are reported to be in the
vicinity of 1.5 mA cm-2, whereas limiting currents on pure aluminium are three
orders of magnitude lower (0.5 – 1 μA cm-2) [Seegmiller et al., (2004)]. Note that
the reference to iron is included since its electrochemical behaviour is somewhat
complimentary to that of copper in aluminium alloys. In a stagnant and naturally
aerated aqueous solution the oxygen reduction at the surface of copper is reported
to occur at 20-30 μA cm-2 [Vukmirovic et al., 2003]. It has also been shown that
copper oxide films do not significantly impede oxygen reduction, presumably due
to their high electronic conductivity [Vukmirovic et al., 2003]. The high
conductivity is related to the semiconducting nature of copper oxides, which is
defined by the low bandgaps of 2.1 eV and 1.3 eV for Cu2O and CuO,
respectively [Richardson et al., 2001].
As discussed earlier, the oxide that forms on most commercial alloys is
predominantly a mixed aluminium and magnesium oxide (often with other
alloying elements such as silicon) and is likely to possess a complex
microstructure. The heterogeneities introduced by varying oxide coverages,
particularly over intermetallic phases, are thought to lead to the onset of corrosion
due to localised defects or poor oxide stability [Guillaumin et al., 2001;
Vukmirovic et al., 2002]. Copper-containing aluminium alloys generally become
more susceptible to corrosion with increasing copper concentration [Davis, 1999].
Copper containing alloys such as AA2xxx and AA7xxx series alloys are
susceptible to pitting and intergranular corrosion, particularly in commonly
encountered, chloride-containing environments [Leblanc and Frankel, 2002]. In
addition to the altered stability of surface oxides, the precipitation of intermetallic
phases results in a heterogeneous surface structure that exhibits areas of varying
Electrochemistry 29

solution potential in the presence of an electrolyte. This creates localized


electrochemical cells between microstructural features at the surface.
The electrochemical properties of the intermetallic phases found within
aluminium alloys have been characterized by various authors, using bulk
synthesized intermetallics [Buchheit, 1995 ; Ilevbare et al., 2004], microelectrode
studies [Suter and Alkaire, 2001] and combinations of the two [Birbilis and
Buchheit, 2005]. It is commonly accepted that all copper-containing intermetallics
except S-phase are net cathodes in comparison with the surrounding aluminium
matrix [Buchheit, 1995; Ilevbare et al., 2004]. S-phase generally possesses
potentials more negative than -0.8 VSCE in a wide range of electrolytes and it is, at
least initially, the site of preferred anodic reactions [Ilevbare et al., 2004]. Copper-
containing intermetallics have bulk electrochemical potentials that increase in
nobility in the order: Al2CuMg < Al7Cu2Fe < Al6(Fe,Mn) < Al2Cu <
Al20Cu2(Fe,Mn)3 [Birbilis and Buchheit, 2005; Ilevbare et al., 2004; Muster et al.,
2007].

Figure 10. Net cathodic currnet densities for copper and intermetallic phases at the
corrosion potential of AA7075-T651 alloy in naturally aerated 0.1 M NaCl (Data from
Birbilis and Buchheit, 2005).

An understanding of the efficiency of individual intermetallics to function as


cathodes can be obtained from the data of Birbilis and Buchheit (2005), who
reported the average currents at bulk intermetallic phases at the corrosion
potential of AA7075-T651 in 0.1 M NaCl (i.e. -965 mVSCE). Copper containing
intermetallic phases support cathodic currents exceeding 310 μA cm-2 and 470 μA
30 T. H. Muster, A. E. Hughes and G. E. Thompson

cm-2 for Al7Cu2Fe and Al2Cu phases, respectively. Intermetallic phases containing
manganese are shown in Figure 10 to support decreased cathodic reaction rates
(less than 120 μA cm-2). Asfeth et al. (2002) demonstrated that the incorporation
of increasing amounts of manganese into Al6(Fe,Mn) intermetallic structure
improved the filiform corrosion resistance of AA3xxx alloys, a result that was
attributed to solution potentials approaching that of the aluminium matrix.
Schneider et al. (2004) showed that the incorporation of iron into bulk
Al20Cu2(Fe,Mn)3 increased cathodic reaction efficiency to achieve similar levels
to Al2Cu. They also showed the magnitude of cathodic currents determined for
Al2Cu and Al20Cu2(Fe,Mn)3 to be three to ten times greater than those measured
on AA2024-T3 alloy surfaces.
Of the non-copper-containing intermetallic compounds present in aluminium
alloys, iron-containing intermetallics, Al3Fe and α-Al12(Fe,Mn)3Si, associated
with AA1xxx and AA3xxx series alloys, are known to act as cathodic sites [Asfeth
et al., 2002 ; Davoodi et al., 2005]. Bulk Al3Fe has been shown to have a noble
corrosion potential (similar to that of Al2Cu) and to support cathodic currents as
high as 200 μA cm-2 [Park et al., 1999]. However, the superior ability of copper-
containing intermetallics to host cathodic reactions was noted by Birbilis and
Buchheit (2005); they found that whilst Al3Fe was more noble than Al2Cu (and
therefore exhibits the larger driving force for cathodic reactions), Al2Cu sustained
higher cathodic rates than Al3Fe. Non-copper containing electrochemically active
phases in aluminium alloys include Al32Zn49, MgZn2, Mg2Si, Al8Mg5 [Davis,
1999] and Al-Si-Mg particles in AA6xxx series [Guillaumin and Mankowski,
2000].
The previous considerations have revealed that intermetallic phases
containing copper may be divided into two types:- (1) those with a cathodic
potential with respect to the matrix and (2) the active Al2CuMg S-phase. The
latter, as will become clear, has unique properties that influence aluminium alloy
corrosion and surface finishing performance. The following sections detail the
corrosion phenomena associated with copper present in (1) solid solution, (2) in
cathodic intermetallic phases, and (3) S-phase intermetallics.
Chapter 5

CORROSION

THE INFLUENCE OF COPPER IN SOLID SOLUTION


Since copper has a more noble formation potential than aluminium,
enrichment of copper can occur at the surface of the solid solution phase
[Habazaki et al. (1997)]. As discussed in the following sections, the majority of
the research concerning matrix enrichment has been carried out on aluminium
alloys exposed to chemical polishing, electropolishing, alkaline etching and
anodisinig. However, many of the findings appear to be systematic and are likely
to hold for corrosion processes, whilst other behaviours appear to be dependent
upon the electrochemical energy applied to the system [Jung et al. (1985)]. In a
range of corrosive environments, aluminium has been shown to preferentially
oxidise, resulting in the build-up of copper within a layer approximately 2 - 5 nm
thick at the alloy surface [Jung et al. (1985), Habazaki et al. (1997)]. Strehblow et
al. (1978) reported relatively thick enrichment zones (up to 65 nm) for anodized
Al-Cu binary alloys; however, as stated by the authors, this apparent increased
thickness of the copper-enriched zone may result from surface roughening. In
environments where aluminium alloys continually experience anodic dissolution,
then it has been suggested that even alloys with relatively little copper (0.06 to 26
at%) may display copper enrichment at the metal-oxide interface [Jung et al.,
(1985), Garcia-Vergara et al. (2004), Korovela et al. (1999), Blanc et al. (1997)].
Once a certain level of enrichment occurs, copper atoms (and other noble alloying
elements) are thought to arrange themselves into clusters through surface
diffusion and eventually protrude from the alloy surface with high curvature due
to undermining of the surrounding aluminium matrix [Sieradzki (1993), Habazaki
et al. (1997)]. The copper clusters may be released into the oxide layer by one of
32 T. H. Muster, A. E. Hughes and G. E. Thompson

two mechanisms: (1) they may be undermined and released as elemental copper-
rich nanoparticles (these particles are likely to have short lifetime due to their
increased free energy, which increases significantly for particles with a small
radius of curvature [Brinker and Scherer (1990)], or (2) copper ions may be
oxidized directly from the protruding clusters, which is also promoted through
increased surface curvature that moves the potential of copper to more negative
values [Newman and Sieradzki (1994)]. Based upon experimental evidence, the
surface of all aluminium alloys containing trace copper upwards will become
enriched in copper over sufficient periods of time. Zhou et al. (1999) have also
demonstrated that the level of copper enrichment is also influenced by grain
orientation.
As referred to previously, the process of copper enrichment at the aluminium
alloy surface can vary with applied (over)potential, which suggests that
observations from accelerated electrochemical processes may not be directly
transferable to the progression of damage occurring at the corrosion potential.
Jung et al. (1985) investigated the copper enrichment at the surface of Al-Cu
binary alloys (containing up to 2 at% copper) after polarizing samples at either -
500 mVSCE or -100 mVSCE in sulphuric acid. They found at -0.5 VSCE, where
copper was cathodically protected, that the surface anodic film contained very
little copper, but that copper particles (4 to 25 nm) were present on the surface.
Further copper enrichment was present beneath the oxide film. These particles
were thought to develop from an enriched layer beneath the oxide. Some of the
particle dimensions were larger than the film thickness and, consequently, had no
oxide covering. At a more aggressive potential of -0.1 VSCE less enrichment was
achieved at the alloy/oxide interface, which was attributed to the active
dissolution of copper. Under these more aggressive conditions there was evidence
of copper (or Cu-Al Alloy) fragments incorporated into the oxide. In terms of
general corrosion performance, the enrichment of copper at the alloy surface is
also likely to increase the number of flaws that exist in the aluminium oxide.
The level of copper and particularly its distribution at the grain boundaries
can also have a profound effect on intergranular corrosion performance. Problems
occur most commonly in high-copper wrought alloys (i.e. AA2xxx and AA7xxx),
particularly when slow quenching leads to precipitation from solid solution and
subsequent intermetallic particle growth, particularly along grain boundaries.
Slow quenching or artificial aging (at 190 ºC) allows the formation of copper-rich
precipitate phases close to the grain boundaries (these differ from coarse
intermetallics that are developed prior to quenching or heat treatment). Copper
(and other alloying elements) diffuse from the matrix into the particle, leaving a
copper-depleted region surrounding the intermetallic [Zhang and Frankel (2003)].
Corrosion 33

In addition to grain boundaries, precipitate-free zones also develop around coarse


secondary intermetallics such as Al2CuMg [Guillaumin and Mankowski (1999)].
Decreased copper in these zones leads to a more active potential (up to 120 mV
have been reported between the AA2024-T3 alloy matrix and Cu-depleted zones
[Davis (1999)] and, as detailed by Galvele and co-workers [Galvele and de
Micheli (1970) Muller and Galvele (1977)], a more active breakdown potential at
the depleted zones. The breakdown (or pitting) potential is an electrochemically
measured potential that is associated with the breakdown of the oxide (or metallic
phase), above which the oxide can no longer re-passivate the surface.
Two breakdown potentials have systematically been reported for high Cu-Al
alloys [Galvele and de Micheli (1970), Guillaumin and Mankowski (1999), Zhang
and Frankel (2003)]. Galvele and Micheli (1970) attributed the first breakdown
potential to the dissolution of copper-depleted zones, and, the second, to pitting of
the grain bodies. For AA2024-T3 alloy, Guillaumin and Mankowski (1999)
suggested that the first breakdown potential was related to the preferential
dissolution of Al2CuMg particles and the second to matrix and grain boundary
breakdown. More recent work by Zhang and Frankel (2003) on AA2024-T3 alloy
agreed that Al2CuMg dissolution leads to the first breakdown potential, but that
the second breakdown results from the initiation and growth of intergranular
corrosion. Breakdown potentials provide a good indication of whether pitting or
intergranular corrosion is likely to occur in particular environments but their
values are somewhat reliant upon the exact experimental details (i.e. scan rate),
and contributing localized corrosion processes cannot always be isolated.
Nevertheless, experimental evidence has shown that the addition of copper
initially shifts the breakdown potential of aluminium to a more noble value.
However, during ageing of the alloy, the breakdown potential gradually decreases
as the copper-containing alloys become increasingly susceptible to intergranular
corrosion. In addition, the susceptibility of aluminium alloys to intergranular
corrosion is not strictly reserved for alloys with higher copper concentrations.
Svenningsen et al. (2004) studied model AA6xxx series extrusions, finding that
alloys possessing low copper (approx. 0.02 wt%) were resistant to localized
corrosion whilst copper additions of approximately 0.17 wt% were prone to
intergranular attack. Electron microscopy showed the presence of copper-
containing precipitates (Al5Cu2Mg8Si6) at the grain boundaries of alloys
susceptible to intergranular corrosion. The Al5Cu2Mg8Si6 phase was demonstrated
to possess a noble potential in comparison to the matrix and accelerated matrix
dissolution.
34 T. H. Muster, A. E. Hughes and G. E. Thompson

THE INFLUENCE OF COPPER IN CATHODIC


INTERMETALLICS
Cathodic constituent particles are present in AA2xxx, AA3xxx, AA6xxx and
AA7xxx alloys. Copper-containing intermetallic phases (particularly phases with
higher levels of copper in the Al-Cu-Fe-Mn family, such as Al6(Cu,Fe,Mn),
Al2Cu, Al7Cu2Fe and Al20Cu2(Fe,Mn)3 phases), support high cathodic reaction
rates under most solution conditions, and assist in the initiation of localized pitting
corrosion [Scully et al., (1993), Chen et al., (1996), Schneider et al. (2004),
Birbilis and Buchheit (2005)]. It is noted that cathodic intermetallic phases that do
not contain copper (i.e. Al3Fe, α-Al(Fe,Mn,Si)) can initiate localized corrosion;
however, initiation rates are much reduced compared to alloys with copper-
containing intermetallics [Nisancioglu (1990), Blanc and Mankowski (1997),
Birbilis and Buchheit (2005)]. The ability of Al-Cu-Fe-Mn (or similar) phases to
support high cathodic reaction rates commonly results in pitting initiation where
breakdown of the oxide occurs adjacent to cathodic particles [Chen et al. (1996),
Liao et al. (1998), Alodan and Smyrl (1998), Guillaumin and Mankowski (1999),
Leclere and Newman (2002), Vukmirovic et al. (2002), Ilevbare et al. (2004)].
Figure 11 shows scanning electron micorgraphs of a section through two pits
developed on AA2024-T3 alloy which had been immersed in 0.5M NaCl for 24
hours. The top surface and sections of the alloy can be seen in tilted micrographs
(Figures 11 (a) and (b) where the top surface is rotated 30° from the normal
allowing a view of the sectioned face), and sections of the S-phase pit can be seen
in Figures 11(c) and (d). The large pit has two Al-Cu-Fe-Mn intermetallic particle
remnants and displays excessive trenching with etch patterns. It is likely that most
of the intermetallic particles have been removed, presumably by dissolution. Slow
dissolution of Al-Cu-Fe-Mn phases have been observed, pitting has also been
seen on the Al-Cu-Fe-Mn particles themselves [Ilevbare et al. (2004)], which is
likely to result from the heterogeneous structure of the intermetallics, as shown by
SEM and AFM imaging (Figure 12). The small pit to the left of the large pit in
Figure 11a is associated with an S-phase particle and has quite different
characteristics, which will be discussed in the following section.
Whilst the appearance of trenches around Al-Cu-Fe-Mn phases have been
observed by many, there is some uncertainly with regards to the mechanism of
trenching initiation and development. Even less is understood about the
propagation of pits from existing trenches. Trenching of the matrix surrounding
cathodic intermetallics may be interpreted as a galvanic corrosion between the
particle and the matrix [Buchheit et al. (1997), Ilevbare et al. (2004)], but is also
Corrosion 35

attributed to etching under the high pH conditions generated by cathodic reactions


(oxygen and/or water reduction). Several studies have reported pH values around
9.5 at the edge of cathodic intermetallics [Park et al. (1999), Vukmirovic et al.
(2002)].

(a) (b)
Cu-Fe-Mn-Al

S-Phase

(c) (d)

Figure 11. AA2024-T3 which has been corroded in 0.5 M NaCl for 24 hours. (a)
secondary and (b) backscatter images of top surface and section of two pits. The large pit
has Al-Cu-Fe-Mn intermetallic remnants and displays trenching where the small pit to the
left (white square) has Cu particulates and is assumed to result from S-phase etchout, (c)
secondary and (d) backscatter images of section of S-phase pit.

The high pH conditions are suggested to enable “cathodic” corrosion of the


aluminium adjacent to the intermetallics, assisted by chemical reaction with the
surface oxide, producing aluminate ions and hydrogen gas [Moon and Pyun
(1997, 1999)]:

+ −
Oxidation via oxygen vacancies: 2 Al + 3H 2O → Al2O3 + 6 H + 6e …5a

− −
Direct formation of hydroxide film: Al + 3OH → Al (OH )3 + 3e …5b
36 T. H. Muster, A. E. Hughes and G. E. Thompson

− −
Dissolution of alumina: Al2O3 + 2OH + 3H 2O → 2 Al (OH ) 4 …6a

− −
Dissolution of aluminium hydroxide: Al (OH )3 + OH → Al (OH ) 4 …6b

Combined oxidation (5) and dissolution (6):


− − −
Al + 4OH → Al (OH ) + 3e 4 …7

− −
Water reduction : 2 H 2O + 2e → H 2 + 2OH …8

Combined water reduction (8) and oxidation and dissolution (7):

3
Al + 3H 2O + OH − → Al (OH ) 4− + H2 …9.
2

The rate at which equation 9 occurs is increased at high pH, and has been
shown to be independent of the current density; the reaction rate is almost as high
when the alloy is under open-circuit conditions as when cathodically polarized
[Moon and Pyun, 1997 ; Buchler et al. (2000), Leclere and Newman (2002)].
Increased buffering of the solution decreases the rate of pitting around cathodic,
intermetallic particles; this has been attributed to the inability of high pH
conditions to be sustained across the surface of the surrounding matrix [Park et al.
(1999), Vukmirovic et al. (2002)]. In order to establish the high pH conditions at
cathodic intermetallics an equal number of anodic reactions are required at some
distance from the particles [Buchler et al. (2000)]. These anodic reactions are
thought to occur on the matrix aluminium, in the dispersoid-free zone surrounding
the intermetallic, or, if present, on nearby Al2CuMg particles [Ilevbare et al.
(2004)]. Leclere and Newman (2002) suggest that local galvanic cells were
unlikely to exist over the small distances involved in trenching in a conducting
electrolyte; however, the formation of colloidal alumina gels between anodic and
cathodic regions on the surface may aid in maintaining potential differences and
stabilizing pit growth [Park et al. (1999), Ilevbare et al. (2004)]. Furthermore,
Ilevbare et al. (2004) reported that, as the pH rises at the site of cathodic
intermetallics, and decreased pH conditions develop at anodic sites, increased
galvanic potential differences are created (by comparison with potential
differences at the bulk electrolyte pH). Schneider et al. (2004) suggested that a
simple pH-induced description might not be adequate to explain all trenching
Corrosion 37

phenomena. For instance, increased chloride ion concentration led to an earlier


onset of trenching but had little influence on oxygen reduction kinetics. Also,
experimentally observed trench widths of 1 – 2 μm are significantly narrower than
expected given that high pH zones extended up to 50 μm. Schneider et al. (2004)
suggested that trenching could in fact be described by a model based upon a
galvanic-couple assisted passive breakdown of the dispersoid-free zone, where the
trenching may be viewed as an unstable pitting process. Such a model could
describe the conversion of trenches to form stable pits containing acidic anolyte
solution, if sufficiently high anodic currents could be generated. In contrast, the
mechanism of a model, where trenches formed at high pH and later switching to
an acid pitting mechanism, is not clear despite some attempted explanation [Park
et al. (1999), Leclere and Newman (2002)].

Figure 12. Atomic force microscopy topography image of AA2024-T3 exposed to non-
chromate deoxidation (Turco SmutGo) for 5 minutes. The large Al-Cu-Fe-Mn particle
shows a heterogeneous microstructure. Trenching is observed around the Al-Cu-Fe-Mn
particle and smaller pits to the bottom left were the site of Al2CuMg intermetallics
(courtesy Dr T. Muster).
38 T. H. Muster, A. E. Hughes and G. E. Thompson

The mechanism of pitting in aluminium alloys at anodic sites has been the
subject of corrosion research for many years. Local anodic reactions, occurring at
the site of the oxide breakdown can result in pits development. Chloride (and
other halide) anions assist in the breakdown of oxides and pit initiation [Davis
(1993)]. Early stages in pit growth are thought to be autocatalytic due to the
generation of a low pH environment at the base of the pit, which catalyses further
pit growth [Hatch (1984)]. The presence of a cathode external to the pit, enables
the anodic production of aluminium ions (eqn 10), which react with water forming
aluminium hydroxides and releasing protons (eqn 11). Protons are reduced at
cathodic sites within the pit to release hydrogen gas (eqn 12). The presence of
various salts can alter the exact composition of corrosion products described in
equation 11, depending on complexation chemistry. For instance, chloride ions
complex with Al3+ ions at anodic sites and prevent the formation of aluminium
hydroxides.

Al → Al(3aq+ ) + 3e − …10

Al(3aq+ ) + 3H 2O → Al (OH )3( s ) + 3H (+aq ) …11

2 H + + 2e − → H 2( g ) …12

After the initiation of a pit, it may either passivate or continue with stable
growth [Szklarska-Smialowska (1999)]. Whilst the autocatalytic mechanism may
explain pit initiation and early stages of growth, it is suggested that cathodic
reactions involving oxygen reduction are still required for stable pitting in most
instances [Pride et al. (1994)]. For the continued growth of pits, the solution
potential must exceed a critical pitting (breakdown) potential. Further, it has been
shown that the ratio of current to pit radius must exceed approximately10-2A/cm
for stable pit growth [Pride et al. (1994)]. The addition of copper to aluminium
alloys increases the pitting potential; however, the corrosion potential is also
increased and, in general, the resistance against pitting usually decreases with
increasing copper content [Davis (1999)]. The surface of metallic copper is highly
efficient at reducing oxygen, and therefore, copper-rich sites allow oxygen and
proton reduction reactions to occur with an enhanced efficiency, thus increasing
the probability of stable pit growth [Davis, 1999]. The high cathodic currents
supported by copper-containing intermetallics are thought to rely upon mixed
reaction control [Buchheit et al. (1999), Jakab et al. (2005)]. Jakab et al. (2005)
Corrosion 39

investigated the oxygen reduction reaction (equation 1) occurring on the surface


of AA2024-T3 alloy. The amount of copper on the surface was varied (through
exposure to an alkaline etch) and the corresponding changes in the oxygen
reduction reaction (ORR) modeled. The charge-transfer controlled ORR was
shown to increase linearly with copper coverage whilst the mass-transfer
controlled ORR increased as a complex function of copper coverage and the
inverse of the boundary layer thickness.
The efficiency of copper containing intermetallics to act as cathodes also
enables them to support copper plating reactions whereby soluble copper ions are
reduced back to copper metal. Chen et al. (1996) studied the corrosion processes
on polished AA2024 alloy, noting that Cu-deposition as nodules was observed on
the surface of Al-Cu-Fe-Mn particles. This indicates that Cu is dissolved from
either intermetallics particles or the matrix and undergoes reduction on the
cathodic particles. Deposition of Cu is particularly a problem in cooling systems
which comprise mixed metals, and may contain brass/copper and aluminium.
Blackwood and Chong (1998) studied the deposition of Cu onto AA6063 plate
from CuSO4 solutions with and without Cl- ions. They noted that the deposition of
copper ions onto copper metal was more favoured than copper ions onto
aluminium metal; however, in the presence of chloride ions, they proposed that
the formation of CuCl+ lowered the energy barrier for deposition of copper onto
aluminium allowing more deposition of copper onto the aluminium. It is worth
noting here that the surface of aluminium alloys containing low amounts of
copper can be altered by the presence of soluble copper ions [Bjørgum et al.
(1995)]. Where heavy metals are deposited on the surface and become soluble
(i.e. at low pH), they are able to reduce back to their metallic state at cathodic
sites whilst oxidizing aluminium [Szklarska-Smialowska (1999)]. Copper is the
greatest concern in this respect, as only a low concentration (i.e. 0.02 - 0.05 ppm)
is needed to initiate pitting at neutral or acidic pH. Even alloys with greater than
99.5 % purity have cathodic regions that are active enough to plate copper onto
their surface [Bjørgum et al. (1995)], and therefore the availability of copper ions
from the environment can have severe consequences for all aluminium alloys.
Several studies have shown the importance of cathodic Cu-containing
intermetallics in controlling pit initiation and growth. Studies carried out on areas
less than 0.01 mm2 on polished AA2024-T3 alloy have demonstrated that Al-Cu-
Fe-Mn phases are critical for pit initiation in chloride-containing environments
[Leblanc and Frankel (2002)]. Small areas of the alloy, exposed to 0.5 M NaCl
without the presence of Al-Cu-Fe-Mn phases, failed to display pitting over
periods of two hours, even where active phases such as Al2CuMg phases were
present. Liao et al. (1998) identified constituent particles as initiators of pitting
40 T. H. Muster, A. E. Hughes and G. E. Thompson

and of pit growth, however, they suggested that pits were shallow unless there
was a cluster of constituent particles, which would promote deeper pitting. For
instance, Figure 12 shows the subsurface attack on AA2024 during corrosion
along grain boundaries and around intermetallic phases. The brighter contrast
between corroded intermetallics and uncorroded intermetallics suggests copper
enrichment of the remnant particles. Subsurface corrosion is thought to be
promoted by the presence and linking of copper-rich intermetallic phases.

Figure 12. Left: Backscattered scanning electron micrograph of corroded AA2024 showing
subsurface attack, Right: Conceptual model for particle-induced subsurface attack after
Liao et al. (1998).

Given that stable pit growth requires a significant exchange current to


continue growth [Pride et al. (1994)], it appears logical that the area fraction of
the various phases, and their proximity to each other, could play a controlling role
in pit formation. The work of Boag et al. (2005) and Ilevbare et al. (2004) also
support the views of Liao et al. (1998), that the proximity and distribution of
intermetallic phases may control initiation kinetics and sustained pit growth. Also,
it is presently unknown whether colloidal copper particles and copper metal plated
on the surface through adsorption and reduction of dissolved copper in solution,
which may also deposit on the walls of a pit, play any significant role in
stabilizing pit growth by providing efficient local cathodes. Further understanding
of pit growth is likely to delineate the importance of intermetallic distribution
profiles, colloidal copper particles and re-plated copper.
Corrosion 41

THE INFLUENCE OF COPPER IN S-PHASE INTERMETALLICS


Al-Cu-Mg ternary alloy systems are a special case due to the presence of
Al2CuMg (S-phase) particles. S-phase particles are of interest for two reasons;
firstly, they make up approximately 64 % of the intermetallic surface area, and
secondly, they have a corrosion potential significantly less noble than the common
AA2024-T3 and AA7075-T6 alloys, and are subject to selective dissolution under
most corrosive conditions [Buchheit et al. (1997)]. In AA2xxx series aluminium
alloys, Al2CuMg phases are known to be the most active, and pits are commonly
generated on, or around, these particles. In recent years, significant work has been
performed to detail the mechanisms of damage associated with S-phase, and
secondly, in developing methodologies to minimize its influence on corrosion and
surface finishing processes [Juffs et al. (2001), McGovern et al. (2000), Yoon and
Buchheit (2006)]. Several studies have noted that S-phase is anodic to the
aluminium matrix and therefore preferentially attacked [Alodan and Smyrl (1998),
Chen et al. (1996)]. In contrast, other studies have reported that the matrix
surrounding S-phase particles has been attacked (analogous to trenching discussed
above), which can eventually lead to undermining and ‘drop-out’ of the particle
[Guillaumin and Mankowski, 1999]. Figure 13 shows a 2D profile across an
Al2CuMg particle showed that has been etched around the side of the particle.
These two observations appear at first glance to be conflicting; however, several
detailed analyses of S-phase properties have led to consensus on most of the
mechanistic details [Buchheit et al. (2000)].
It summary, the consensus is that the S-phase is initially anodic and then,
upon the dealloying of magnesium and aluminium, a copper-rich phase is left
[Chen et al. (1996)]. The enrichment in copper results in S-phase being
transformed from being a net anode to a net cathode. For example, Figure 14
shows a plume of hydrated aluminium-magnesium oxide emanating from the
surface of AA2024-T3 alloy. In Figure 14 (a) and (b) the plume of oxide is
formed on the surface after treatment in cerium chloride solution; it demonstrates
a morphology that is typical of S-phase dissolution. Figures 14 (c) and (d) show
another plume of corrosion product from an S-phase particle. This plume also
protrudes from the surface and bright copper-particulates are incorporated into the
corrosion plume. These copper particulates are release from the copper-rich phase
following the selective dissolution of aluminium and magnesium.
Previously, in Figure 11, a small pit was identified as the site of an S-phase
remnant. The corrosion product is full of heavy atomic number particulate
material (bright spots in Figure 11 (d)), which were identified as copper. Higher
magnification images (Figure 15), revealed that the individual bright spots in
42 T. H. Muster, A. E. Hughes and G. E. Thompson

Figure 11(d) themselves comprise clusters of nanoparticles 50 to 200 nm in size.


This smaller pit shows the characteristics described by Buchheit et al. (2000) for
the mechanical redistribution of Cu redistribution via transport in the corrosion
product as will be discussed below.

a. Polished
~ 0.3 μm

b. Polarised

Figure 13. 2D profiles of Al2CuMg after (a) polishing and (b) 5 min of polarization at
above the breakdown potential of S-phase (-750 mVSCE), after Guillaumin and Mankowski
(1999).

Buchheit et al. (1999) demonstrated that potential scans, conducted on bulk


S-phase, provided indirect evidence of the dealloying of S-phase. Anodic
polarization scans, starting at 50 mV cathodic to the steady state open-circuit
potential (OCP) in 0.5 M NaCl, showed a positive shift (up to 150 mV) in the
OCP determined during the scan. These positive shifts in the OCP were attributed
to cathodic corrosion, where increased alkalinity leads to the preferential
dissolution of aluminium from the matrix. Interestingly, after the anodic scan,
when one would expect to find the enrichment of elemental copper at the surface,
the OCP remains at similar values to previously unexposed S-phase. Buchheit et
al. (2000) suggested that electrical isolation of deposited colloidal copper means
that it does not alter the OCP by any significant amount.
Corrosion 43

(a) (b)

(c) (d)

Cu particulates

Figure 14. (a) Secoondary and (b) backscattered electron images of AA2024-T3 which has
been exposed to acidified CeCl3 solution with 3% H2O2 and rinsed in ethanol (from
Gorman, 1998). Milled surface of AA2024-T3 alloy which has been exposed to neutral salt
spray for 8 days.

S-phase is an active anode with dissolution rates reported to be equivalent to


20 mA cm-2 (from visual estimation) during corrosion processes on an AA2024
alloy in 0.1 M sulphate and 0.005 M chloride solution [Ilevbare et al. (2004)].
Aldykewicz et al. (1995) reported decreased anodic currents, in the range of 30 to
60 μA cm-2, using scanning vibrating probe analysis over an AA2024 alloy in
0.012 M NaCl solution. There are a number of reasons for the large difference in
the magnitude of estimated current exchanges, including electrolyte concentration,
pH and non-Faradaic dissolution of magnesium and aluminium into solution.
Buchheit et al. (1999) showed that in addition to the S-phase being an active
anode, it supports high current densities for cathodic reactions (exceeding 1 mA
cm-2). The efficiency of S-phase to support high cathodic reaction rates has been
attributed to the formation of large surface area ‘sponges’ of copper produced by
44 T. H. Muster, A. E. Hughes and G. E. Thompson

dealloying [Buchheit et al., 1999]. (A visual example of a copper-rich sponge can


be found in Figure 21 in the section on deoxidizing/desmutting.) Numerous
studies have detailed the existence of Cu sponges [Buchheit et al. (1997),
Buchheit et al. (1999), Chen et al. (1996), Guillaumin and Mankowski (1999),
Kolics et al. (2001)]. The cathodic properties of the copper-rich phase lead to
dissolution of the matrix surrounding the original particle. Also, the matrix is
thought to be depleted in dispersoid particles in the areas surrounding S-phase
particles [Guillaumin and Mankowski (1999)], although some analysis shows that
this may not always be the case [Buchheit et al. (1997)]. Guillaumin and
Manikowski (1999) reported what appeared to be copper deposits on the matrix
directly adjacent to the copper-rich S-phase remnants. Buchheit et al. (2000) have
also observed pitting around the periphery of S-phase particles and suggested the
initiation of secondary galvanic cells.
Numerous studies have reported copper enrichment at the surface of
aluminum alloys associated with S-phase intermetallics [(Buchheit et al. (1997),
Chen et al. (1996)]. Movement of copper particulates of 10 to 100 nm in
dimension have been observed to originate from the site of Al2CuMg particles and
to migrate to the outermost surface [Buchheit et al. (1997), Chen et al. (1996),
Buchheit et al. (2000)]. Several potential mechanisms of copper enrichment have
been suggested; including the direct oxidation of copper into solution and the non-
Faradaic release of copper nanoparticles from copper-rich sponges [Buchheit et
al. (2000)].

Cu-particles

Amorphous
(a) Oxide (b)

Figure 15. (a) Secondary and (b) Backscattered electron images of Cu particles embedded
in an amorphous oxide. These Cu particles are thought to be the corrosion product form
the dissolution of S-phase depicted in Figure 10(c) and (d).

Studies have shown the presence of copper ions in solution during


electrochemical testing of S-phase and AA2024 alloys. Vukmirovic et al. (2002)
Corrosion 45

found that copper ions were released from AA2024 alloy. Buchheit et al. (2000)
used stripping voltammograms to confirm the release of both copper ions and
elemental copper into solution during the dissolution of Al2CuMg into aqueous
chloride solutions. Significant peaks were observed relating to the oxidation of Cu
→ Cu+ and Cu+ → Cu2+ for Al2CuMg when held at its open-circuit potential and,
additionally, when polarized ± 50 mV with respect to the open-circuit potential.
Therefore, evidence suggests that it is possible to remove copper from the surface
of aluminium alloys even though bulk copper is thermodynamically stable at the
corrosion potentials achieved in solution. As a consequence, it is likely that
copper redistribution cannot be completely suppressed, even using
electrochemical techniques [Buchheit et al. (2000)]. The importance of S-phase
dissolution in the generation of surface copper was recognized in the work of
Vukmirovic et al. (2002), who compared the amount of Cu released into solution
from model and commercial AA2024 specimens. Commercial AA2024 alloy
showed much higher levels of solubilised copper, which was attributed to its
ability to form copper-rich sponges. The model sample, consisting of a low-
copper alloy aluminium covered with physical vapour deposited copper islands,
produced much lower amounts of solubilised copper.
In order for copper ions to directly oxidize from the alloy in solution, the
potential of the copper phase must be decreased. Buchheit et al. (2000) suggested
two mechanisms by which the open-circuit potential of copper can approach the
more negative values achieved in Al-Cu-Mg aluminium alloys. Firstly, the
complexing of copper ions with chlorides can alter the activity of copper ions in
solution and decrease open-circuit potentials. This effect is only possible at pH
values exceeding approximately 5.5 and becomes increasingly significant at high
pH. Secondly, the radius of curvature of materials has a large effect on their free
energy and reduces the open-circuit potential, particularly where curvatures
approach the nanometer scale. This scenario is possible given that preferential
leaching of magnesium and aluminium can lead to the presence of high surface
area regions of copper. In order to reduce surface energy, the structure coarsens
through coalescence mechanisms based upon differential solubilities [Brinker and
Scherer (1990), Buchheit et al. (1997)]. Buchheit et al. (1997) suggested that
particle remnants of metallic copper become detached from the surface,
presumably through the mechanical action of pit growth and corrosion product
formation or through solution movement, as depicted in Figure 16 [Buchheit et al.
(1997)].
The argument that Cu-particles are released by mechanical action is supported
by the work of Dimitrov et al. (1999), who immersed AA2024-T3 into both
stirred (500 rpm) and unstirred 0.5 M NaCl solutions for periods up to 35 minutes.
46 T. H. Muster, A. E. Hughes and G. E. Thompson

Upon the removal of the AA2024-T3 samples, stirred samples routinely possessed
1.8 times more copper than that of the unstirred sample. Dimitrov et al. (1999)
and later Vukmirovic et al. (2002) suggested that the rate of redeposition of
copper fragments, previously removed from the surface via non-faradaic
processes, was increased due to the added convection during stirring.

Figure 16. Schematic diagram of a mechanism for redistribution of copper by dissolution


of large Al2CuMg and Al2Cu intermetallic particles after Buchheit et al. (2000).
Dealloying of the particle is caused by anodic polarization of the particle by the
surrounding matrix, resulting in the formation of a copper-rich network (which coarsens
with age) that under hydrodynamic flow or stresses arising through corrosion product
formation, releases small copper-rich clusters. The electrically isolated copper clusters
oxidize at their OCP, and may precipitate as copper-oxides or be reduced back to
elemental Cu at cathodic sites and serve as efficient local cathodes that stimulate
secondary pitting.

S-phase shows some evidence of passivation under high pH, possibly as a


result of the stability of magnesium oxide in alkaline environments [Ilevbare et
al., 2004]. However, even at high pH, any passivity is overcome by the significant
anodic polarization that exists when coupled to an aluminium matrix phase
[Ilevbare et al., 2004]. The oxide covering the S-phase varies from that of the
matrix oxide [Guillaumin et al., 2001] and is thought to have weak protective
properties [Vukmirovic et al., 2002]. Therefore, the weakened oxide and anodic
Corrosion 47

polarization experienced by the S-phase particles results in rapid attack in


corrosive environments.

Figure 17. Schematic diagram of the advanced stages of corrosion induced by dealloyed S-
phase particles. Trenching is observed around the particles, however, the nature of the
reactions within the trench is an area of controversy. It is thought that the trenching is
initiated by alkaline conditions due to the cathodic activity of the Cu-rich S-phase remnant.
However, in the latter stages of corrosion, acidity might also develop at the base of the
trench setting up a differential aeration cell. Secondary pitting is also observed due
reduction of Cu ions on the matrix which serve as local cathodic sites that stimulate anodic
reactions on the adjacent matrix.

A significant body of work has detailed the peculiar behaviour of S-phase


intermetallics, however, there appears to be some areas that could benefit from
further study. The exact structure of dealloyed S-phase is not well understood.
Studies by Sieradzki and co-workers have made some progress in this area of
noble metal enrichment, describing the dealloyed structures in terms of
percolation theory [Sieradzki (1993), Newman and Sieradzki (1994), Vukmirovic
et al. (2002)]. The dealloyed structure is predicted to be determined by a
combination of the dissolution rate and copper content. High dissolution rates do
not allow surface diffusion and relaxation processes to coarsen the structure,
leading to more open networks. Also, copper contents cannot be too high,
otherwise dealloying will not lead to an isolation of the percolation network,
which is required for the mechanical release of copper clusters. With a copper
concentration of 25 at% and an orthorhombic structure, S-phase is considered to
48 T. H. Muster, A. E. Hughes and G. E. Thompson

be borderline in its ability to release copper clusters [Vukmirovic et al. (2002)].


Hydrodynamic forces would certainly assist in fragmentation.
Finally, the influence of S-phase and its corrosion by-products on corrosion is
not well documented or understood. Figure 17 shows a schematic of the types of
corrosion processes that may occur around an S-phase after longer time periods
(compared with the dealloying mechanism presented in Figure 16). Trenching is
initiated around the dealloyed particle as it switches from behaving as a net anode
to a net cathode. Some trenched areas may be converted to acidic pits due to the
separation of anodic and cathodic regions due to the establishment of differential
aeration cells, which are created via the precipitation of colloidal alumina gels
(corrosion products) within the pit. Pitting itself may also be enhanced through
copper plating at cathodic sites on pit walls, and the deposition of elemental
copper on the aluminium matrix may result in the formation of secondary pits.
The role of oxidized copper particles is not clear in terms of advanced corrosion,
although their presence would ensure a constant supply of soluble copper, which
is expected to decrease overall corrosion performance.
Chapter 6

CHEMICALLY PRETREATED SURFACES

FINISHING PROCESSES
Metal finishing is a term that encompasses a range of chemical processes for
taking aluminum and its alloys from the as-received state to a state where the
surface has been prepared (in a reproducible way) for further processing, such as
the application of an organic coating or adhesive [Critchlow and Brevis (1996)].
Many of the metal finishing processes involve a series of treatment steps that
require alternate immersion in acid and/or alkali treatments designed to modify
the surface chemistry prior to anodizing or conversion coating. A range of
mechanical treatments, such as grit blasting with graded alumina or silica, also
exist for aluminium surface preparation [Critchlow and Brevis (1996)], but there
is little or no information on how such mechanical treatments affect the copper
distribution.
During the last twenty years, there has been considerable pressure to make
metal finishing processes more environmentally friendly and less of a health risk.
This has led to research into a range of new processes. Much of this work has
been performed in research environments where, often, a disjunct occurs between
the outcomes of the research and the requirements of the metal finishing industry
and its endusers.
Such differences in perspectives between the research community and the
metal finishers may be demonstrated by the following example. For many years,
chromate/HNO3/HF deoxidisers have been used for the treatment of copper-
containing, sheet aluminium products. This deoxidizer effectively removes S-
phase particles and essentially dissolves the Al-Cu-Fe-Mn intermetallic particles
in proportion to their composition without preferential dissolution that leads to
50 T. H. Muster, A. E. Hughes and G. E. Thompson

copper enrichment of the intermetallic remnant [Hughes et al. (2003)] as might be


observed with other deoxidizer combinations [Hughes et al. 2001]. Further, the
etch rate is not so high as to remove sufficient aluminium matrix to continually
expose new intermetallic constituent particles. Thus, it is possible to obtain a
surface which has a significantly reduced population of intermetallic particles, but
it does develop a copper-enriched layer beneath the surface oxide. Recently,
Moffitt et al. [2001] who, on the observation of this thin copper enrichment layer
beneath the surface oxide on AA2024-T3 alloy after treatment in a nitric acid
based deoxidizer, suggested: “The surface … maintains a copper enriched surface
despite the apparently long held claim, that this eliminates surface copper
enrichments”. The issue here is not whether Moffitt is correct or those he quotes
are incorrect, but the perspective from which they approach the concept of
cleanliness. In metal finishing, nitric acid solutions remove copper-containing
smut, giving a visually bright appearance which is all the processor needs to know
for quality control; however, if the surface is examined with sophisticated
techniques such as TEM, Auger or XPS, then copper will be detected in enriched
layers and intermetallic remnants. Moffitt et al. [2001] partly condede this point in
their next sentence, citing that the original claim was made in 1958 well before
surface science techniques were commonly available. With this example in mind,
one aim of this section is to highlight the differences between the requirements for
metal finishing, and the approach of surface science, both of which have a part to
play in the understanding of the current processes, and the development of new
ones.
The metrics used by each community can be used to further elaborate on the
difference between the metal finishing industry and the research community.
Metal finishers deal with etch rates, coating weights and processing bath
maintenance to evaluate the operating state of coating lines. They use visual
appearance as one of the major quality control tools, although more sophisticated
tools are commonly available and used (X-Ray fluorescence, for example). The
eveness of colour of the coating is a good example of a visual quality control tool
[Schmidt-Hansberg and Schubach (2003)]. Metal finishers also use performance
testing in standardized corrosion environments, such as Mil-C-81605 and Mil-C-
5541E in conjunction with standard operating procedures such as ASTM B117 to
estimate performance “in-service”. Research laboratories, by contrast, often use
sophisticated techniques to study small areas of samples. They often use fresh
solutions rather than “aged” or “sweetened” solutions, which have some metal
dissolved into it, or a surface condition, such as polished to 1 μm or less, as
opposed to chemically cleaned. Hence, it is important to identify these differences
when relating the surface chemistry, metallography and morphology to metal
Chemically Pretreated Surfaces 51

finishing parameters, when improving current processes and developing new


ones.
Even within the metal finishing industry, it is clear from a number of studies
that changing the processing parameters, or introducing new processing products,
can result in marked differences in the performance of the metal finishing process.
Ketcham and Brown’s (1976) work on the performance of chromate conversion
coatings (CrCC) pretreated with a variety of different commercial products, all
qualified and commercially available (alkaline cleaners and deoxidation
solutions), showed that considerable variation resulted between products in terms
of the final corrosion resistance provided by the CrCC. In another study, Leggatt
et al. (2002), found that CrCCs applied on AA2024, AA6061 and AA7075 alloys
in the field, failed to pass corrosion resistance performance requirements despite
the use of appropriate standards for application of the CrCC’s. These standards
were designed to give the required performance. These studies show that etching
and surface condition, including surface enrichment of alloying components play
an important role in performance.
From the point of view of the enrichment of alloying elements, particularly
copper, there are two main considerations:

1. etch rate of the metal finishing solution


2. solubility of the alloying element in the metal finishing solution

The etch rate determines the amount of buildup of material on the surface, but
the chemistry of the treatment solution has a bearing on the whether that material
is dissolved into solution. To this end sequestering agents are often added to
commercial products to complex the metal ions to prevent them from depositing
on the surface during the treatment step. From the perspective of etchants, a high
etch rate (NaOH or fluoride) results in fast removal of aluminium, but leaves
copper behind on the surface. For this reason, HNO3 is either used as a post
treatment in the case of alkali treatments, or added to acidic treatment steps. In the
following section the influence of individual treatment steps on aluminiuim-alloys
will be examined to illustrate how they change the surface condition.

SOLVENT CLEANING
The purpose of solvent cleaning, using either solution or vapour processes, is
to remove oils and greases that may have been applied to the surface of the alloys
for the purposes of corrosion protection during transport. Typical vapour
52 T. H. Muster, A. E. Hughes and G. E. Thompson

degreasers included organo chlorines such as trichloroethane, trichloromethane


and trichloroethylene.
Solvent cleaning can be divided into the type of cleaning used in the metal
finishing industry and those used in the research laboratory. In the former,
solvents, such as those listed in Table 7, are often used, whereas in the laboratory,
a different range of reagents such as acetone, ethanol or methanol are frequently
used for sample preparation. In the metal finishing environment, solvent cleaning
is disappearing.

Table 7. Solvents Used for Wiping Workpieces Clean

Metal Finishing Research Environment


Paraffin (kerosene) Acetone
White Spirit Ethanol
Carbon tetrachloride Methanol
Methylated Spirits Trichloro ethane

Clearly the list on the left of Table 7 is different from that on the right and has
some implications for drawing conclusions about metal finishing and corrosion
processes from laboratory research. For example, Chidambaram and Halada
(2001) explored acetone degreasing of polished AA2024-T3 alloy, and observed
carbonyl groups on the surface; this led them to suggest the formation of surface
acetates (both copper and aluminium). They further proposed the formation of
acetic acid, and suggested that in the acidity produced from acetic acid, copper
chloro-complexes are formed which assist corrosion. They also point to the role of
photo-oxidation in modifying the speciation of the complexes.
While these studies cannot be directly related to surfaces during metal
finishing, they have some relevance in terms of artifacts produced in the research
environment that are not likely to be present in the processing environment.
In the metal finishing industry, issues associated with solvent wiping have
also been identified. Trichloroethane, when used in vapour degreasing, can react
with water to produce HCl which dramatically attacks the aluminium [King
(1988)]. Commercial solvents have added stabilizers and inhibitors to prevent
such reactions [King 1988]. Hughes et al. (1996) also noted the presence of
surface chlorine, using XPS, on aluminium oxide after a solvent wipe with
trichloroethane, supporting the model that where acidity develops on the surface,
HCl may have an important role to play in subsequent corrosion events.
Chemically Pretreated Surfaces 53

DETERGENT CLEANING
Metal finishing has undergone considerable change in the last 20 years due to
environmental legislation and concerns regarding occupational health. From the
environmental point of view there has been a drive to remove volatile organic
chemicals from industrial processes. This has impacted metal finishing processing
due to the removal of vapour degreasers, and changes in the composition of paint
systems from low-solids to high-solids based as well as a general trend away from
organic-based to aqueous –based paint systems.
Solvent cleaning using the vapour phase has been replaced by detergent
cleaners, which generally have very low or non-existent etch rates. Under these
circumstances, there are only minor changes from the original structure and
chemistry of the surface, and possible adsorption of surfactants onto the oxide
surface.

ALKALINE CLEANING
Aqueous alkaline cleaners were traditionally used after vapour degreasers to
further saponify any oils or grease remaining on the surface after vapour
degreasing (or detergent cleaning) as well as give a mild etch to the surface
[Wernick et al. (1987), King (1988)]. These solutions are usually carbonate-based
and inhibited for non-etch cleaners, or caustic based with inhibitor for etch
cleaners[Wernick et al., (1987)]. Inhibitors include silicates, chromates,
phosphates, fluorides, silicofluorides or organics to reduce the rate of etching
[Wernick et al., (1987)]. Modification to the surface during this process step
usually involves the generation of basic zinc and magnesium oxides on the
surface; these must be removed during deoxidizing or desmutting steps. Ketcham
and Brown (1976) demonstrated that different formulations of these types of
cleaners can have a significant impact on corrosion resistance of chromate
conversion coatings, which is almost certainly related to residual species on the
surface after cleaning and surface morphology.
Build-up of transition metal alloying components during these processes is
dependent on the etch rate. Etching of copper-containing alloys in NaOH
(chemical milling, described below) leads to the development of a heavy copper-
rich smut on the surface. The loose copper-rich smut is readily removed in HNO3-
containing solution [Nelson et al. 2001, Dimitrov et al. (1999)]; however, it is not
clear whether a copper-enriched layer is present at the Al/Al-oxide interface.
54 T. H. Muster, A. E. Hughes and G. E. Thompson

Chemical cleaning in NaOH is moderately reactive, with reports of etch rates of


approximately 0.18 and 0.28 μm/min for 0.1M [Liu (2005)] and 0.6M NaOH
[Nelson (2001)] respectively. There have been a number of studies of the
influence of 0.1M NaOH solutions on aluminium alloys. Liu et al. (2005) have
observed copper enrichment and formation of copper nanoparticles in the surface
oxide of a sputter deposited Al-30 at% Cu alloy after treatment in 0.1M NaOH.
Based on work on anodized coatings and model binary Al-Cu alloys, they have
estimated that the maximum buildup is around 1 x 1016 atoms/cm2 [Liu et al
2003]. This level of buildup occurs with the removal of only 5 nm of the alloy
[Liu et al. 2005]. Continued etching of the aluminium matrix, resulting in the
formation of aluminium oxide, leads to incorporation of copper particles which
can be oxidized since they are no longer cathodically protected by the aluminium
matrix. Cupric oxide is also a narrow band gap semi-conductor which can display
n- or p-semicondiction [Siripala and Kumra, (1989); Di Quarto et al., (1985);
Sutter et al. (1995); Millet et al. (1995)]. The electron donor capability of this
oxide may mean that it could act as a cathode.
Studies, using positron techniques 1 , of oxide/hydroxide formation on
relatively high purity aluminium after alkaline cleaning, showed increased number
of nano-sized defects, partly related to enrichment of small defects near the
surface [(Wu et al. (1994)]. Nelson et al. (2001) examined the surfaces of
AA2024-T3, AA6061-T6 and AA7075-T6 alloys after treatment in 0.6M NaOH
with a prior treatment in HNO3/HF. They found considerable development of
basic oxides on the surface incorporating alloying additions such as magnesium
and zinc. The level of copper observed on AA6061-T6 alloy was only slightly
lower than AA7075-T6 alloy, despite it having only one quarter as much copper
in the alloy content. The level of copper enrichment for AA2024-T6 alloy was
roughly three times that of AA6061-T6 alloy, even though it had ten times the
bulk copper content, as determined by ICP analyses. However, as pointed out in
the work of Roberts et al. (2002), the enrichment underneath the surface oxide
may have been greater, but XPS probing was unable to penetrate through the
surface oxide.
Constituent particles undergo selective phase and elemental etching in
alkaline solutions. For example, Lunder and Nisancioglu (1987) found enrichment
of iron and manganese on a range of non copper-containing interemtallic

1
Positron techniques include Positron Annihilation Lifetime Spectroscopy (PALS) and Doppler
Broadened Energy Spectroscopy (DBES). In both techniques, positrons impinge on a surface and
annihilate with electrons within the material. If the positrons reside within a void in the material
then they will have a longer lifetime than the rest of the material, which provides some information
on the size of the void.
Chemically Pretreated Surfaces 55

constituent particles including Al3Fe, Al6(Mn, Fe) and Al6Mn. They also found
that the alkaline treatment increased the susceptibility to neutral salt spray for a
range of different alloys, with sheet AA5052 being the exception. Preferential
removal of aluminium during alkaline treatment should lead to the formation of
aluminium-rich oxides on the external surface of the constituent particles with
probable enrichment of zinc and magnesium for particles containing these
alloying elements.
The silicated, carbonate based cleaners [King (1988)] have a much lower etch
rate, but still modify the surface. Hughes et al. (1996) observed an increase in
magnesium, zinc and silicon with increased immersion time in a silicated alkaline
cleaner using XPS. It was suggested that the silicon was present in the form of
silicate on the basis of the Si 2p binding energy. Moffitt et al. (2001) observed
similar changes for AA2024-T3 alloy as well as AA7075-T6 alloy, although they
did not analyse silicon on the surface. Both studies suggested that the levels of
copper increased with alkaline treatment above that expected in the AA2024-T3
matrix. Conversely, Toh et al. (2004) did not observe any copper enrichment on
AA7475-T7651 alloy after treatment in a mild alkaline cleaner, but did observe
incorporation of silicon.
These differences are likely to be due to differences in the etch rates of the
alkaline cleaners, which are influenced by pH, operating temperature, solution
chemistry and the composition of the intermetallic particles. All of these can vary
with the product and even during processing leading to varied etch rates. It can be
expected that, if there is a high etch rate of the aluminium, and low solubility of
either the alloying elements in the etch solution, or compounds formed from the
alloying elements in the etch solution, then there will be a build-up of the
transition metals including iron, manganese and copper in the oxide coating the
matrix as well as transition metals in intermetallic particles.

CHEMICAL MILLING
Chemical milling is an alkaline etching process which operates at etch rates of
12 to 37 μm/min [Wernick et al., (1987)]. Chemical milling of aluminium alloys
is achieved through the addition of 75 to 150 g/L (approx. 2 M to 4 M) of NaOH.
The rate of milling can be modified via the addition of certain chemicals [Wernick
et al. (1987)] of which silicate, phosphate and chromate are some of the most
common.
Milling copper-containing alloys leave a heavy smut on the surface which is
removed using a nitric acid treatment [[Wernick et al., (1987), King (1988),
56 T. H. Muster, A. E. Hughes and G. E. Thompson

Nelson et al. (2001), Hughes et al]. Liu et al. (2005) observed the development of
copper nano-particles as a result of etching Al-30 at.% Cu alloy. Nanoparticles
were formed as a result of enrichment of copper beneath the surface oxide,
followed by clustering of copper atoms and occlusion by the oxide film. Other
studies on more pure aluminium alloys show that the intermetallics in the surface
tend to etch slower than the surrounding matrix [Lunder and Nisancioglu, (1987),
Short and Shearsby (1969)]. Indeed, Montiero et al. (1991) found severe etching
associated with the interemtallic particles for a commercial purity alloy. Even on
pure aluminium, Wu and Herbert (1996) observed buildup of iron, copper and
gallium after caustic etching 99.98 % pure Aluminium in 1M NaOH. It was
suggested that copper particles were contained in a highly defective 10 nm oxide
on the surface of the alloy.

WATER RINSING
Relatively little work has been done on the influence of water rinsing, with
the majority of the work undertaken some time ago. Clearly, for a chemically
active aluminium surface, exposure to water at room temperature and trace ions
can influence the amount of oxide and the adsorbed species. Most of the published
data is related to the Forest Product’s Laboratory etch (FPL). The nascent oxide
had a particular structure that was suitable for adhesive bonding or anodizing
[Pocius (1981), Russell and Garnis (1976), Sun et al. (1978)]. Russell and Garnis
(1976) demonstrated that the contact electrical resistance increased significantly
with the length of the rinse time, with higher resistance generated on purer alloys.
Studies of oxide growth in aqueous solution have largely been confined to
processes at elevated temperature. At temperatures above 50°C to boiling, oxide
growth is dominated by pseudo-boehmite (Al2O3.nH2O, n = 2 or 3) and boehmite
(Al2O3.H2O) formation, whereas, at 40°C, bayerite eventually forms the dominant
product. [Vedder and Vermilea (1970), Altenpohl (1966), Alwitt ((1974),
Underhill and Rider (2005)]. Clearly, surface condition is important since it has
been reported that as little as 100 ppm of Si in solution can retard oxide growth by
absorption onto the aluminum surface [Vedder and Vermilea (1970), Altenpohl
(1966)]. The adsorption of chromate ions on the surface can also significantly
impede oxide formation at elevated temperature [Gorman et al. (2003)]. At room
temperature, many species, like chromate and phosphate, adsorb irreversibly, thus
having a major impact on oxide formation [Heine and Pryor (1967), Ergun et al.
(1997), Sotoudeh et al. (1981), Böhni and Uhlig (1969), Badawy and Al-Kharafi
(1997), Németh et al. (1998), Vermilyea and Vedder (1970)]. Additionally, oxide
Chemically Pretreated Surfaces 57

growth is a thermally- activated process and the driving force for oxide formation
will be much less at room temperature [Vedder and Vermilea (1970), Altenpohl
(1966)].
Whether the presence of surface copper affects oxide growth in rinse water at
room temperature is not well understood. At 40° to 50°C, the ultimate film
thickness of oxide films developed in deionised water was similar for a range of
alloys [Gorman et al. (2003), Underhill and Rider (2005)]. However, several
studies suggest that the initiation of oxide growth may be more rapid with
increased copper content [Gorman et al. (2003), Underhill and Rider (2005)]. The
improved initiation of oxide growth may arise because the copper, which acts as a
net cathode, promotes aluminium dissolution at a greater rate during the initial
stages of immersion than on pure aluminium alloys.

DEOXIDATION/DESMUTTING
Deoxidation or desmutting is a term used to describe removal of the oxides
left after alkaline cleaning, and is intended to be a mild etching process rather than
a milling process. The term “deoxidation” is preferred in North America, whereas
“desmutting” is preferred in Europe. For the purposes of this chapter we will use
the term “deoxidation” as the process has broader chemical implicatons than the
simple removal of “smut”.
As with any process that results in etching of aluminium alloy surfaces,
preferential etching of aluminium results in a build up of alloying components
leading to surface enrichment, as well as texturing of the surface. The first part of
this section will deal with the types of deoxidisers that are available and their
impact on surface enrichment, particularly copper enrichment, and the second part
will deal with surface textures.

Surface Chemistry

The composition of the surface and the residual chemical species after
treatment in deoxidizing solutions varies according to the chemistry of the
treatment solution and the etch rate. As mentioned previously, Ketcham and
Brown (1976) observed that different deoxidisers had a significant influence on
the corrosion resistance of the CrCC. Nelson et al. (2001) also observed a
significant deterioration in performance in corrosion resistance of the CrCC on
58 T. H. Muster, A. E. Hughes and G. E. Thompson

AA2024-T3 and AA7076-T6 alloys when a simple HNO3/HF treatment was used
instead of a chromate-based deosidizer. This was attributed to a difference in the
copper levels on the surface. Liu et al. (2004, 2005) observed that the CrCC
coating weight decreased with the enrichment of copper at the surface due to
anodic dissolution during coating. Therefore there is evidence that different
surface treatments influence the performance and deposition kinetics of coatings
such as CrCC. This is likely to be due to the surface chemistry in the form of the
passivity of the surface oxides, absorbed species and surface morphology. None
of these aspects is a metric used in the metal finishing industry so one of the aims
of this section is to draw the connections between the characteristics measured
using the sophisticated techniques in the research environment to the metrics used
in the metal finishing industry.
A review of the literature reveals that there are relatively few studies on
deoxidisers in comparison with the numbers on conversion coating and anodizing.
A number of studies are related to the Forest Products Laboratory (FPL) etch,
which has been used as a preparative treatment for adhesive bonding. Since the
surface texture related to this treatment is important for bonding, it is discussed in
detail in the next section. Many of the studies relate to current, chromate based
deoxidisers [Hughes et al. (1996), Moffitt (2001)2 , Gorman et al. (2003)] and
chromate replacements for these deoxidisers which will be discussed below.
Examination of the chromate/HNO3/HF deoxidizer indicates that it leaves a
thin chromate-passivated oxide on the surface for a range of alloys [Gorman et al.
(2003)]. There is also evidence of copper, after deoxidation, for a range of Al-
alloys, including AA1100-O, AA2024-T3, AA3004-H19, AA5005-O, AA6061-
T6 and AA7075-T6 alloys, despite some of these alloys having very low bulk
copper content [Gorman et al. (2003)] and also some indication of copper
enrichment with this deoxidizer on AA2024-T3 alloy [Moffitt et al. (2001),
Gorman et al. (2003), Hughes et al. (1996)].
Figure 18 (a) shows a high magnification secondary electron image of the
surface after deoxidation in a chromate/HNO3/HF for ten minutes. It is evident
that there is a network structure across the surface which is decorated with larger
nodules; the network and the nodules are assumed to be part of the oxide covering
the surface. An XPS depth profile through this structure is presented in Figure 18
(b) where it can be seen that the copper signal increases with a corresponding
decrease in the oxygen signal suggesting copper enrichment at the aluminium-

2
The Authors assume that Moffitt et al. used a chromate based deoxidiser since the product number
they use (Parker Amchem Deoxidiser No 7) is a chromate based deoxidiser, although they do not
explicitly note this in their experimental section.
Chemically Pretreated Surfaces 59

oxide/aluminium-metal interface [Hughes et al. (1996), Moffitt et al. (2001)]. The


depth profile has been divided into three regions: Region I is the external surface
which has a high chromium concentration, Region II is the oxide, and Region III
is the interfacial region with the underlying metal. Given that the oxide ridges
observed in Figure 18 (a) are the thickest part of the coating, and, therefore, the
last part of the oxide to be removed during depth profiling, these results suggest
that the copper enrichment is greatest at the interface between the oxide ridges
and the underlying aluminium as depicted in Figure 19. The results do not
indicate whether there is also a continuous film beneath the thinner parts of the
surface oxide.
The model presented in Figure 19 is not unlike models presented by Caicedo-
Martinez et al. (2003) for the development of surface textures during chemical
polishing of aluminium alloys. However, in this case, the deoxidizer solution
removes all alloying components and impurities, effectively leaving only copper
enrichment on the surface. Species from the deoxidizer solution are also
incorporated into the surface oxide, since it contains chromium and fluorine.
Given that chromate in conversion coatings solutions react with the cathodic sites,
which Liu et al. (2003) and Brown et al. (1992) have shown to be on the ridges of
textured aluminium surfaces, then the chromate in the deoxidizer solution may be
reacting with copper-enriched texture on the aluminium surface forming a
network of mixed chromium, aluminium oxides or oxyfluorides. Aluminium
metal ridges have been included beneath the oxide ridges in a model similar to
that of Caicedo-Martinez et al. (2003) where the copper provides some cathodic
protection to the metal beneath it and promotes anodic dissolution of the adjacent
metal. These metal ridges have not been observed, as yet, experimentally.
For the chromate-free deoxidizers, there are a range of compositions.
Generally, if a high etch rate is required then the deoxidizer will contain HF.
Thus, a simple deoxidizer formulation combines a mineral acid and HF, which is
sufficient for low copper-containing aluminium alloys. High etch rate deoxidation
can also be achieved with elevated temperature as discussed below. For the
purposes of copper-removal, HNO3-based solutions are preferred.
Chromate-free deoxidisers, with the ability to remove copper, include
formulations based on HNO3/BrO3- [Bibber (1991,1993), Kloet et al. (2005), Toh
et al. (2004)], rare earths [Hughes et al. (2001,2003), Kimpton et al. (2000),
Campestrini et al. (2004)] and deoxidisers based on HNO3/HF [Montiero et al.
(1991,1988), Nelson et al. (2001)], as well as HNO3/HF with added oxidants, i.e,
oxidant/ HNO3/HF. A number of these deoxidisers have oxidants that are intended
to perform a similar function to chromate, such as Fe3+ ions [Hughes et al.
60 T. H. Muster, A. E. Hughes and G. E. Thompson

(2003,2004)]. The latter forms the basis of several, commercial, non-chromate


deoxidizers.
The presence of the chromium and fluoride containing oxide after
deoxidation, leads into some often overlooked aspects of deoxidation and the
relationship that it bears to the subsequent conversion coating step. In the last
section, the possibility of oxide growth during water rinsing was mentioned. It is
possible that the presence of this thin Cr-Al-F oxide layer, which shows some
passivating properties [Hughes et al. (1996)], may prevent oxidation during
rinsing after deoxidation, and also provides a nascent oxide structure that
promotes rapid growth of the CrCC after immersion in the conversion coating
bath.

60

(a)
(a) 50 (b) Al
O
40 Cu ( x5 )
Cr ( x5 )
Atomic %

F ( x5 )
30
I II III
20

10

0
500
500 nm
nm 0 100 200 300 400
Sputter Time (sec)

Figure 18. (a) Secondary electron image of the surface of AA2024-T3 after deoxidation in
Cr6+/HF/HNO3 and (b) aluminium, oxygen and copper (×5) XPS depth profiles through the
surface in (a).

The etch rate of Fe3+/HNO3/HF deoxidizers are lower than that of the
corresponding Cr6+/HNO3/HF deoxidizer by about two thirds, depending on the
alloy [Hughes et al. (2003)], and result in a slower etching of the surface oxides
that remain after alkaline cleaning [Hughes et al. (2004)]. Intermetallic removal is
effective; S-phase are completely removed, and the Al-Cu-Fe-Mn intermetallic
particles severely etched, although there in no data in the literature on whether
there is preferential enrichment as a result of etching. The etch pattern on the
matrix, typical of the Fe3+/HNO3/HF deoxidizer is shown in Figure 20, and is
similar to the etch patterns on the surface of AA2024-T3 alloy after treatment in
the chromate deoxidizer, but leaves iron-containing deposits on the surface
Chemically Pretreated Surfaces 61

[Bibber. 1993;, Hughes et al. (2004)] as well as copper enrichment, presumably at


the aluminium/aluminium-oxide interface [Hughes et al. (2003)].

Aluminium Oxide
Ridges

Cu enrichment

Aluminium Oxide

Al alloy
Aluminium Matrix

Figure 19. Proposed model of oxide ridges for the surface of AA2024-T3 after deoxidation
in Cr6+/HF/HNO3 deoxidizer.

Cerium-based deoxidisers have also been investigated for cerium based


conversion coatings [Hughes et al. (1999), Kimpton et al. (2000), Campestrini et
al. (2004)]. One low etch rate version of this deoxidizer results in little more than
preferential etching of Mg-oxides from the surface oxide. However, a high etch
rate composition of this deoxidizer has the oxidant/ HNO3/HF formulation with
Ce4+ being the oxidant, but it has an additional mineral acid in H2SO4. Both
Hughes et al. (1999) and Campestrini et al. (2004) observed the retention of
copper after immersion in a high etch rate Ce4+/H2SO4/HNO3/HF deoxidizer. Both
studies observed the deposition of copper particles (200 nm in dimension) from
the deoxidizer onto the matrix as well as onto the intermetallic particle remnants,
which additional oxidants (H2O2, K2S2O8) assisted in removing.
The selective dissolution of species (aluminium, magnesium, manganese, iron
and silicon) can leave a range of structures during deoxidation. In deoxidizers
containing HNO3 (and HF), there tends to be even etching of the intermetallic
particles (Hughes et al. (1996, 2001, 2003), as well as a high rate of intermetallic
removal from fallout due to undermining by etching in the presence of HF
[Campestrini et al. (2004)]. Copper enrichment occurs when either other mineral
62 T. H. Muster, A. E. Hughes and G. E. Thompson

acids are operated at elevated temperatures, or the deoxidizer composition does


not include HNO3.

(a)

(b)
Figure 20. a) Secondary electron image of the surface of AA2024-T3 alloy after
deoxidation in Fe3+/HF/HNO3 and (b) surface reconstrtuction from strereopair of another
part of the surface showing the scalloping (etching) pattern.
Chemically Pretreated Surfaces 63

(a) (b)

20 μm 10 μm

(c) (d)

2 μm 5 μm

Figure 21. Electron micrographs of Cu-Fe-Mn-Al containing intermetallic particles after


deoxidation at ambient temperature in various deoxidisers. (a) BrO3-/HNO3, (b)
Fe3+/HNO3/HF, (c) Ce4+/H2SO4/HNO3/HF and (d) Na2S2O8/H2SO4/HF.

For example, Figure 21 shows some backscattered and secondary electron


images of etched Al-Cu-Fe-Mn type intermetallic particles. In Figure 21 (a) and
(b), where strong oxidants (i.e. BrO3- or Fe3+) are used in the deoxidizer
formulation, there is even etching of the intermetallic particles, but no significant
copper enrichment of the intermetallic particle remnants. In the case of the
Ce4+/H2SO4/HNO3/HF deoxidizer, there has been selective removal of aluminium,
iron and manganese leaving a copper sponge (Figure 21(c)). This sponge appears
to have precipitated copper from solution on the surface. Where
Na2S2O8/H2SO4/HF was used (Figure 21(d)), there is also selective removal of
iron, manganese and aluminium, leading to copper-enrichment of the intermetallic
particle, but no sponge was apparent in these particles, suggesting a different
mechanism of copper enrichment. In this last case, there was no detectable copper
in the deoxidizer solution indicating limited opportunity for deposition onto the
intermetallic particles of dissolved copper from solution. In these latter two cases,
64 T. H. Muster, A. E. Hughes and G. E. Thompson

the presence of SO32- appears to have a significant influence on the mechanism of


dissolution, but the details have not yet been revealed.
Where the copper becomes electrically isolated, presumably either through
the oxidation of the underlying alloy, effectively electrically isolating the copper-
sponge from the aluminium matrix, or through undermining due to etching, the
particle can dissolve to form Cu2+, which becomes available for deposition onto
the surface. As stated previously, deposition (or plating out) has not been
observed with HNO3-based deoxidisers, however, as seen above, it can occur with
other formulations.
There are a few studies that attempt to make the connection between etch rate
and changes to the surface as a function of chemistry. Hughes et al. (2001,2003)
examined the attack on AA2024-T3 alloy using H3PO4, H2SO4 and HNO3, alone
and with additions of HF, and HF/Na2S2O8. In the mineral acids alone, at ambient
temperature, the matrix was protected by the surface oxide, and the intermetallic
particles were clearly protected by a layer of aluminium matrix, deposited during
rolling [Lunder and Nisancioglu (1987)]. Etching of the matrix can be achieved at
elevated temperatures (60°C) and a heavy copper smut can be developed on the
surface [Campestrini et al. (2004)]. Heavy copper smut can also be developed at
room temperature with the addition of an etchant, like HF, to the mineral acids in
formulations, such as H2SO4/HF, H3PO4/HF, H2SO4/HF/Na2S2O8 and
H3PO4/HF/Na2S2O8,. This occurs as a result of etching of the matrix without
dissolution of the copper. It was also noted that the addition of Na2S2O8 as an
oxidant had a minimal affect on reducing the copper levels [Hughes et al. (2001)].
In an endeavour to delineate the relationship between etch rate and the
observations of various surface techniques, Hughes and co-workers (2001,2003)
have proposed a qualitative model to describe the stages of attack by deoxidisers
on a rolled aluminium alloy surface. The process proceeds in three stages:

Stage 1: Upon immersion in the acidic deoxidiser solution, there is


preferential dissolution of components of the oxide remaining
after alkaline cleaning. These include the basic magnesium and
zinc oxides and also silicon-containing phases. Low etch rate
deoxidisers (Ce(IV), HNO3/BrO3- at 20°C or simple mineral
acids) generally do not proceed beyond this Stage unless the
operating temperature is elevated substantially.
Stage 2: Deoxidisers with any significant etch rate, proceed beyond Stage
1 into Stages 2 and 3. During Stage 2, the surface oxide left after
alkaline cleaning is completely removed and etching of the
underlying alloy begins. This is a complex process since the alloy
Chemically Pretreated Surfaces 65

contains a number of alloying additions in a variety of phases and


each of these components must reach an equilibrium between
accumulation and dissolution. Stage 2, therefore, is the
intermediate stage in which the dissolution of aluminium leaves
behind an accumulation of alloying components. Stage 2 is
reached by deoxidiser combinations containing HF and also BrO3-
/HNO3 at 40°C or 60°C and probably H2SO4 deoxidisers operated
at increased temperatures such as those used in adhesive bonding
since these give the characteristic network of oxide ridges. The
balance between surface oxide formation and dissolution moves
towards thinning of the oxide and the build-up of alloying
components, (particularly copper). It is not clear whether the rate
of oxide thinning is delayed by the Cu(I)/Cu(0) redox couple,
although data on low copper-containing alloys may help to
resolve this issue. The presence of oxidants in the system may
assist in copper dissolution (but this is likely to depend on the
details of the chemistry) so that an equilibrium concentration
might be obtained.
Stage 3: In Stage 3, alloy component dissolution reaches an equilibrium
between dissolution and accumulation. The Cr(VI) and Fe(III)
deoxidisers, described earlier, (which contained HNO3, in
combination with HF) and the BrO3-/HNO3 deoxidiser at 60°C
reach Stage 3. Similar results were observed for rare-earth based
deoxidisers when they contained fluoride.

The HNO3/BrO3- deoxidizer provides a good example of all three stages as a


function of temperature and time. Figure 22 shows that at 20°C that there is very
little attack on the matrix, but there is significant attack on the intermetallic
particles; after 10 minutes the particles were nearly completely removed. Thus,
the attack on the matrix oxide is confined to Stage 1. Attack on the matrix was
evident at 40°C where, after 10 minutes, the deoxidizer had moved through Stage
2 and commenced Stage 3 since the characteristic etch patterns were observed. At
60°C the deoxidiser moves quickly into Stage 3, where the network of oxides
ridges is apparent after deoxidation for only 5 minutes.
To further illustrate the stages of deoxidation, the etch rates for a number of
different aluminium alloys in chromate/HNO3/HF deoxidizer are shown in Figure
23. It is clear that the etch rates are higher initially for the AA7xxx series alloys
and lower for AA2024-T3 alloy, but, subsequently, both alloys reveal similar
values. The initial differences, during Stage 1, represent changes in the dissolution
66 T. H. Muster, A. E. Hughes and G. E. Thompson

of the surface oxide. The 7475-T7651 alloy remains in Stage 1 and 2 the longest
because this particular batch of alloy had a thick surface oxide [Toh et al. (2003)].
Eventually all the alloys display similar etch rates in Stage III which represent the
dissolution of the substrate aluminium, where copper enrichment and intermetallic
dissolution occur. A qualitative estimate of the three stages of dissolution is also
included in Figure 23.

20ºC

(a) (b) (c)

40ºC

(d) (e) (f)

60ºC

(g) (h) (i)

1 min 5 min 10 min

Figure 22. Scanning electron micrographs of AA2024-T3 alloy after deoxidation for
various times and temperatures in BrO3-/ HNO3- based deoxidiser. (Scale marker for (c) =
1 µm, for all other images = 500 nm.). Reproduced with permission from Hughes et al.
(2003).

A qualitative model, similar to that described, could be made more


quantitative by combination of the knowledge of copper build-up under anodic
coatings (see below), enabling prediction of the levels of copper-enrichment after
the pretreatment process. The two main issues that need to be addressed here are:
(i) the buildup of copper at the aluminium/aluminium-oxide interface, and (ii) the
removal of intermetallic particles from the surface. Hence the copper-enrichment
could be described by:

Cuenrich = f(EM, EIM,Dcu) …13


Chemically Pretreated Surfaces 67

where EM is the etch rate of the matrix, EIM, is the etch rate of the intermetallic
particles and Dcu is the amount of redeposited copper. To determine the influence
of these terms on copper enrichment, it would be necessary to know the copper
content of the matrix, as well as of individual particles. It would also be necessary
to know the influence of the solution chemistry on etching of the intermetallic
particles. If it is assumed that these three variables are independent, then equation
13 can be separated into the sum of three independent terms:

Cuenrich = f(EM) + f( EIM) + f(Dcu) ….14

19

18

17
I II III
Etch Rate (μm/h)

16

15

14

13

12

11

0 5 10 15 20
Time (mins)

Figure 23. Etch rates for chromate-based deoxidizer three different aluminium alloys. ● =
7475-T7651, ■ = 2024-T3 and ▲= 7075-T6. I= stage I, II = Stage 2 and III = Stage 3.

Clearly, terms like f(EIM) and f(Dcu) would need to incorporate the surface
area of the copper and well as the coverage of the aluminium. For a number of
deoxidisers such as the chromate/HNO3/HF and HNO3/BrO3- (60°C) deoxidizers,
where the intermetallics are virtually completely removed and there is no
significant redeposition of copper then equation 14 is simply related to the build-
up of copper under the surface oxide. For more complicated systems, specific
models would need to be developed for the enrichment of copper due to
intermetallic particle dissolution and copper redeposition. A simpler approach to
68 T. H. Muster, A. E. Hughes and G. E. Thompson

populating the terms of equation 14 is through the use of a look-up table which
gives a figure for a selected acid combination under specified conditions.

40 nm

5 nm

40 nm

5 nm Oxide

Cu-enrichment Aluminium

Figure 24. Model of surface oxide structure based on scanning electron microscopy stereo
pairs after Venables et al. (1979). (Cu-enrichment has been added by the authors and did
not appear in the original paper by Venables et al. (1979))

Figure 25. Copper enrichment levels are the electropolishing of aluminium-copper binary
alloys as a function of copper content in the alloy (after Liu et al. (2003)).
Chemically Pretreated Surfaces 69

Surface Structure

Many copper-containing aluminium alloys develop surface textures or


structures comprising networks of ridges across the surface during
electropolishing or chemical treatment (e.g. Figures 20 or 21). Liu et al. (2003)
have shown that the individual ridges consist of a metal ridge covered by an
oxide. A number of models have been put forward for the structure of the ridge
network on the surface [Brown (1949), Hunter and Robinson (1973)]. At this
point, it seems to be controlled by the distribution of alloying elements in the
surface [Thompson et al. (1987), Caicedo-Martinez et al. (2003)].
These types of structures are extremely important for adhesive bonding
applications, particularly in the aerospace industry. Work by Venables et al.
(1979) suggested that a particular type of nodular oxide structure (Figure 24)
promoted good bond strength but was susceptible to degradation during storage,
particularly if fluoride was present. In early work, it was recognized that this
structure was somehow related to copper in the processing bath or the copper
content of the alloy [Pocius, (1981), Sun et al. (1978, 1980)], but precise details
have not been determined. Based on the work of Caicedo-Martinez et al. (2003)
the authors have proposed that there is copper enrichment beneath the oxide
ridges which facilitate dissolution of the adjacent matrix; these have not, as yet,
been observed experimentally.

ELECTROPOLISHING
The electrochemical polishing of a metal is a process where a DC current is
applied to oxidize the metal and remove it from the surface. Leveling occurs due
to the field gradient across the surface, which results in higher dissolution rates at
raised areas of the surface than in troughs. According to Thompson et al. (1987),
an anodic oxide film forms on aluminium during electropolishing; the film
develops at the metal/film interface through O2- ingress and with the outwardly
mobile Al3+ ions ejected into the electrolyte under the influence of the electric
field. In addition, a through-film dissolution of aluminium ions occurs, along with
dissolution of the outer oxide regions, which are exposed to the reactive
electrolyte. For aluminium-copper alloys, preferential oxidation of aluminium
enables development of a copper enriched layer in the alloy immediately beneath
the film developed in the electropolishing bath. Copper enrichment at the alloy
surface during electropolishing has been observed to reach levels corresponding to
70 T. H. Muster, A. E. Hughes and G. E. Thompson

40 at%, which are contained in a layer approximately 2 nm thick (Habazaki et al.,


1995).
The thickness of the enriched layer is thought to be independent of bulk
copper concentration and electropolishing conditions. However, the amount of
copper contained within this thin enriched layer can vary significantly. Liu et al.
(2003) electropolished aluminium-copper binary alloys ranging in copper
concentrations from 0.0025 at.% to 1.5 at.% at 20 V in perchloric acid/ethanol
solution (20/80 v/v) at 278 K for various times. They reported metal removal rates
of 40 nm s-1. Copper enrichments at the surface of each binary alloy were
independent of time between one minute and fifteen minutes electropolishing,
indicating that all enrichment occurred within the first minute. Figure 25 shows
that the amount of copper enrichment was found to increase rapidly up to 0.4 at%
copper, whereas alloys containing greater amounts of copper showed a relatively
weak enrichment. Liu et al. (2003) introduced an enrichment factor, defined as the
ratio of copper atoms in the enriched layer (atoms cm-2) compared to that of the
bulk concentration (at%). Therefore, the enrichment factor decreases for alloys
with increasing copper concentration (Figure 25).
If significant copper enrichment is achieved, copper-rich clusters formed at
the alloy surface may be released into the oxide film, which leads to smutting of
the electropolished surface. As described previously, nitric acid etchants are
generally used to remove copper smut. For instance, Liu et al. (2003) used a 50
wt% HNO3 solution at ambient temperature for 60 s to remove smut after
electropolishing laboratory samples. It follows that decorative aluminium alloys
used to present bright surface finishes usually have a low amount of copper.
Typically, AA5xxx series alloys containing minimal copper are used, such as;
AA5005, AA5050, AA5252, and AA5657 [Hatch (1984)].

ANODISING
Anodizing is an electrochemical process that, in appropriate electrolytes,
generates a cellular-structured, porous anodic oxide film on the surface of
aluminium. Anodizing on copper-containing alloys is reported to show reduced
film growth rates due to the preferential dissolution of intermetallic compounds
[Takahashi et al. (2003)].
Unlike acid electrolytes, where porous anodic films are generally formed,
Thompson et al. (1987), examined barrier-type anodic film formation on
aluminium-copper alloys in near-neutral electrolytes. Anodic oxide growth
proceeds at both the solution-film and alloy-film interfaces under the electric field
Chemically Pretreated Surfaces 71

[Habazaki et al. (1996)]. Habazaki et al. (1996) reported that about 40% of the
film thickness was formed at the film/electrolyte interface due to the outward
diffusion of Al3+ and the remainder is formed at the metal/oxide interface through
the inward diffusion of O2-/OH- species. For binary solid solution alloys the
efficiency of film growth is only marginally influenced by the development of a
layer of copper-enrichment in the alloy, which occurs immediately adjacent to the
alloy-film interface [Habazaki et al. (1995), Liu et al. (2004)]. During anodizing
of such alloys, it has been demonstrated that the anodic film develops initially in
the absence of incorporated copper species, due to preferential oxidation of
aluminium. Such anodic oxidation allows copper to accumulate at the alloy/film
interface until it reaches a concentration of about 40 at% in a layer of thickness of
about 2 nm. Habazaki et al (1995) used RBS to demonstrate that the extent of
copper enrichment at the interface was independent of the anodizing conditions
and was similar to the enrichment developed by electropolishing of an alloy of
equivalent copper content. At the critical level of enrichment, copper oxidizes at
the alloy/film interface and enters the oxide in its alloy proportions in the
continued presence of the enriched layer. The applied voltage was also found to
influence the incorporation of copper into anodized films (Table 8), where
increased voltages up to 300 V led to a linear increase in copper enrichment at the
alloy surface. For the particular conditions used by Habazaki et al. (1996), 140 V
was calculated as being a critical voltage, which, if exceeded, would lead to the
incorporation of copper into the anodized film.

Table 8. Compositional analysis determined by RBS of enriched alloy layer


and anodized films formed under varying applied voltage at 5 mA cm-2 in 0.1
M ammonium pentaborate at 293 K (Habazaki et al., 1996)

Anodizing voltage (V) Copper in enriched alloy Atomic aluminium:copper in


layer (at%)1 anodized film (× 10-3)
15 4.8 ≤ 0.2
50 16.0 ≤ 0.2
150 41.2 ≤ 0.2
300 Not determined 4.4
1
The bulk alloy copper content was 0.9 at%.

Where a significant enrichment of copper has occurred at the metal/film


interface, oxygen gas-filled voids develop in the barrier film through the semi-
conducting nature of the Cu(II)-O bond; the gas pressure achieves very high
levels, with eventual rupture of the surrounding alumina. Interestingly, the
enrichment and consequent gas generation appears to be alloy-grain orientation
72 T. H. Muster, A. E. Hughes and G. E. Thompson

dependent [Zhou et al. (1999)]. In separate studies, where such gas generation
was eliminated by anodizing of thin layers of the copper-containing alloy, the
outward mobility of Cu2+ ions was shown to be 2-3 times that of Al3+ ions, which
gives rise to their loss at the film/solution interface and generation of an anodic
film of copper content less than that of the bulk alloy. For the bulk alloy, the
consequence of gas generation and film rupture is to generate a heavily flawed
anodic film.
Concerning porous anodic film formation, Shimizu et al. (1997) examined
anodizing of a binary Al-Cu alloy that had been artificially aged to develop a fine
distribution of θ’ precipitates. Porous anodic film formation, proceeding
exclusively at the alloy-film interface due to O2- ion ingress under the field, led to
recession of the alloy/film interface. However, when θ’ precipitates of relatively
high copper content relative to the adjacent matrix where encountered, immediate
gas generation developed to sufficiently high pressures that ruptured the anodic
film. Subsequently, film repair proceeded, removing the θ’ precipitate and
generating a porous anodic film of enhanced porosity.

(a) (b)

Figure 26. Transmission electron micrographs of (a) unclad 2014 T3 alloy and (b) clad and
after anodizing in sulphuric acid, revealing the resultant anodic films of contrasting
morphologies (courtesy of Prof. G. Thompson).

Figure 26 reveals transmission-electron micrographs of anodic films formed


on clad and unclad AA2014-T3 alloy in sulphuric acid. A distinct affect of copper
on the alloy is revealed in the anodic film which displays increased porosity
compared with that on the comparatively pure clad material. This is related to
Chemically Pretreated Surfaces 73

copper enrichment in the alloy, gas generation and possible film rupture and film
repair. Interestingly, additional mechanisms of anodic film formation are under
consideration, which question the established mechanism of film growth at the
aluminium/film interface that has been thought to proceed through dynamic
equilibrium with field assisted dissolution at the pore base [Garcia-Vergara et al.
(2006)]. That is, for anodic films formed in acid electrolytes (eg. sulphuric acid
and phosphoric acid with significant amounts of incorporated electrolyte-derived
anions) stress induced flow of the film material is being explored to explain pore
formation. In other words, plastic flow of the anodic film material is thought to
proceed from the barrier layer beneath the pore base to the adjacent cell material.
For the film formed on the AA2014 alloy shown in Figure 26, it appears that the
lateral porosity or layered film regions are associated with a cyclic oxidation of
copper from the enriched layer, in addition to the consequences of oxide plasticity
and flow under the field [Iglesias-Rubiane et al. (2006)].

CONVERSION COATING
Conversion coating is a term used to describe a coating process that
transforms (“converts”) the natural oxide on the surface of aluminium to an oxide
with more desirable properties, such as improved paint adhesion and corrosion
resistance. The most commonly used conversion coatings are chromate (and
chromate-phosphate) conversion coatings (CrCC), which are a family of closely
related, commercially-available processes. Thus, products for cleaning,
deoxidation and conversion coating are available for a range of applications, but
since this review is concerned with copper distributions, the focus is on the
processes used for treating high-strength, copper-containing alloys, which are
used extensively in the aerospace industry. Non-chromate coating formulations
are also available [Buchheit and Hughes (2003), Nylund (2000)] with the Ti/Zr
processes [Knudsen et al. (2003), Deck and Reichgott (1992), Lunder et al.
(2004), Tomlinson (1997)] and a Co-based process [Schreiver (1992), Schreiver
(1996), Roland (1998), Hughes et al. (2004)] being the most common; however,
these do not perform well on Copper-rich alloys [Chalmers (1995)]. There are
also a number of other processes such as permanganate processes [Bibber (1991),
Hughes et al. (2006)], silane treatments, self-assembled monolayers and rare earth
processes [Schmidt-Handsberg and Schubach (2003), Rivera et al. (2004)], the
last of which particularly benefits from the presence of copper at the surface
[Hughes et al. (1995,2004), Palomino et al. (2006), Campestrini et al. (2001)].
74 T. H. Muster, A. E. Hughes and G. E. Thompson

Conversion coating is a multi-step process involving a number of the cleaning


steps described previously, prior to the deposition of the conversion coating. The
cleaning steps usually involve a mild alkaline cleaning followed by a rinse and
then, for copper-containing alloys, treatment in a HNO3/HF deoxidizer usually
contains chromate 3 . As stated previously, in aerospace applications, the CrCC
process works best with a chromate-based deoxidizer prior to conversion coating
[Ketcham and Brown, 1976]. Further, the use of nitric acid in the deoxidizer
formulation for copper-containing alloys generally removes copper associated
with the smut and the vast majority of the intermetallics from the surface [Hughes
et al. (1996,2003)]. Thus, the only copper of any significance on the surface is
localised in an enriched layer between the aluminium oxide and the aluminium
metal [Moffitt et al. (2001), Hughes et al. (2001)].
Despite the fact that the Cr(VI)/HNO3/HF pretreatment removes most of the
copper from the surface, the alloying content (particularly copper) of aluminium
alloys still has a marked influence on the coating weight. As pointed out in several
texts [Ketcham and Brown (1976), Buchheit and Hughes (2003), Eppensteiner
and Jenkins (1995)], the amount of deposition of the CrCC is greater with purer
alloys, and lower for highly alloyed aluminium. Thus, copper-containing alloys of
commercial importance, such as AA7xxx and AA2xxx series, can be expected to
have some of the lightest coating weights of the aluminium alloy family.
Typical coating times are up to a few minutes for most alloys, with increased
coating times leading to some unusual behaviour. Liu et al. (2005) observed that
at longer immersion times (up to 20 minutes) the amount of deposited CrCC on
AA2014-T6 and a range of magnetron sputtered binary Al-Cu alloys, decreased as
a direct result of an increase in the surface copper levels. Trathen et al. (1993)
reported that the corrosion resistance and adhesive failure of a paint system on the
CrCC both showed unusual behaviour in the time to pitting in neutral salt fog
testing as well as adhesive failure, across a range of alloys for coating times of 10
minutes.
There have been numerous studies of the deposition of the CrCC onto
aluminium-alloys generally, but fewer on the deposition onto copper-containing
aluminium-alloys. At the most basic level, film growth proceeds through contact
of the alloy with the coating solution where the fluoride attacks and thins the
surface oxide. With sufficient thinning electron tunneling may proceed [Brown et
al. (1993), Katzman et al. (1976)]. Fluorine has been detected at the CrCC/metal
interface [Hughes et al. (1997), Vasquez et al. (2002), Abd Rabbo et al. (1978),

3
For Al alloys with low copper content, a mineral acid/HF mixture is commonly used such as
H3PO4/HF.
Chemically Pretreated Surfaces 75

Treverton and Davis, (1977)], suggesting that fluorine-containing species such as


AlF3 or AlOF have formed at the interface. Once the reaction starts then the
mechanism of coating in its most basic form is described by [Katzman et al.
(1979)]:

Cr2O72- + 2Al0 +2H+ + 6HF → 2AlF3 + 2CrOOH + 3H2O ....14

The reaction scheme can be elaborated to include the effects of thinning of


the surface oxide on the aluminium prior to the deposition reaction. As the
reaction proceeds, more chromium species are deposited, however, a limiting
thickness is eventually reached which depends on bath chemistry. It should be
noted that there are two classes of CrCC processes: unaccelerated and accelerated.
The unaccelerated processes contain a mixture of HF, chromic acid and nitric acid
and are operated at a pH below 2. Accelerated processes, in addition to the usual
components have an added accelerator, the most common of which is K3Fe(CN)6
[Xia and McCreery (1999)].
The coating weight [Arrowsmith et al. (1984)], coating thickness [Juffs et al.
(2002), Katzman et al. (1979), Campestrini et al. (2001)] or intensity of
characteristic lines from the conversion coating [Schram et al. (1998), Kendig et
al. (1993), Drozda and Maleczki (1985), Xin and McCreey (1999), Treverton and
Amor (1985)] increase with time. For unaccelerated coatings, at longer coating
times, the coating weight tends to stabilize [Katzman et al. (1979)]. For
accelerated coatings (those containing K3Fe(CN)6), the coating weight also
becomes self limiting, but, at longer immersion times, the coating becomes loose
and powdery with poor adhesion.
Typical coating thicknesses for the non-accelerated and the accelerated
formulations are shown in Figure 27. These data are derived from several sources
as listed in the figure caption, and, generally, for five minutes immersion time in
the coating solution. It is evident that the coatings from the accelerated
formulations are about an order of magnitude thicker than those from the
unaccelerated solutions. Coatings on copper-rich intermetallic particles are much
thinner than the surrounding matrix. It should be noted that there is considerable
variation in the coating thickness with the accelerated coating. For example,
Treverton and Amor (1985) estimated a coating thickness of over 1 micron after
only 3 minutes coating. Hagans and Hass (1994) found lower values, but than
others, this may be due to the low pH (pH 1) of the coating solutions that were
used. Osbourne (2001) has show that virtually no coating is developed at pH 1
whereas maximum coating weights are obtained around pH 2.
76 T. H. Muster, A. E. Hughes and G. E. Thompson

1200
accelerated
1000 no acceleration
Coating Thickness (nm)

800

600

400

200

0
a b c d e f a b c a c a b b

FeAl3
FeAl3
--

--
--
CuAl2
CuAl2
CuAl2
Cu2FeAl7

FeAl3
Matrix

Matrix

CuMgAl2
Matrix

Cu2FeAl7

3
CuAl

CuMgAl

Figure 27. Coating thicknesses for a range of surfaces in unaccelerated and accelerated FeAl
coating solutions. a = Juffs et al. (2002), b = Hagans and Haas (1994), c = Vasquez et al.
(2002), d = Katzman et al. (1979), e = Sun et al. (2001), f = Treverton and Amor (1985).
In the case of Vasques et al. (2002) the matrix value from Hughes et al. was used as the
100% value and the coating thickness over the IM phases was estimated accordingly.

The coating thickness on the intermetallic particles is similar for both the
unaccelerated and the acceleterated processes as can be seen in Figure 27. From
Figure 28, for the unaccelerated coating, it can be seen that the coating develops
over the matrix more quickly than over the intermetallic particles after five
seconds immersion, which agrees with the results of Hagans and Haas (1994) for
shorter coating times. Clearly, the coating thickens over the intermetallic particles
and develops to the same thickness as the matrix (~ 100 nm) over longer times as
seen for five minute coatings in Figure 27. In the accelerated process, the coating
over the matrix is nearly an order of magnitude thicker than the intermetallic
particles, which have a similar thickness to the coating in the unaccelerated
process. Hagans and Haas (1994) suggested that this was due to the formation of
copper-ferrocyanides on the surface of the copper-containing intermetallic
Chemically Pretreated Surfaces 77

particles. Cyano species were also observed on iron-containing species, but the
particular phase was not identified. This is discussed in more detail below.

Figure 28. Atomic force microscopy of the surface of AA2014 alloy after 5 seconds at
ambient temperature in a chromate conversion coating solution (0.8gl-1 NaF , 4.0gl-1 CrO3
and 3.5gl-1 Na2Cr2O7). Reproduced with permission from Liu et al. (2001).

The role of the ferricyanide accelerator has not yet been fully elucidated
although there are two models for its role. The first, proposed by Treverton
(1981), suggests that ferricyanide preferentially adsorbs onto the surface of
chromium oxide/hydroxide gel particles as they form at or near the surface. The
adsorbed ferricyanide then blocks the sites for chromate absorption onto the gel
particles; thus more Cr(VI) is available for reaction with the aluminium surface.
The second model, proposed by Xia and McCreery (1999), considers the
Fe(II)/Fe(III) couple to act as a catalyst (redox mediator in their terminology),
with reduction of Fe(III) to Fe(II) during aluminium oxidation and being oxidized
from Fe(II) to Fe(III) by Cr(VI) reduction. These reactions facilitate the
generation of Cr(III) hydroxyoxide on the surface. Xia and McCreery (1999) also
examined a number of other catalysts, such as IrCl62-, Fe3+ and V3+, which
accelerated the deposition of the CrCC, but none was as effective as K3Fe(CN)6.
Hagans and Haas (1994) examined the influence of ferri/ferro cyanide
78 T. H. Muster, A. E. Hughes and G. E. Thompson

accelerators and confirmed that they resulted in coating thicknesses about an order
of magnitude greater than the unaccelerated formulations on the matrix (Figure
27). However, the nature of the surface at which reactions occur are unclear in
these models.
The issue relating to the nature of the reacting surface has, to some extent,
been clarified by Osborne (2001), who discussed the nature of chromate
conversion coating in a sol gel context. He suggested that at low pH (< pH 1.2)
individual sol particles are formed adjacent to the surface, whereas, at more
moderate pH a gel may form on the surface. Thus, the reacting surface is probably
either the external surface of the sol particles at low pH, or the surface of the gel
coating at more moderate pHs. The other area of debate is the nature of the cyano
complexes on the surface; again, while there is strong evidence for certain species
e.g. Berlin green [Xia and McCreery (1999)], the isomerisation reactions between
iron and chromium cations in hexacyano complexes are complex, and there are
probably a number of species present at the surface [Basset Brown et al. (1968)].
The literature on the relationship of copper distributions on the deposition
mechanism is complicated by the differing methods of preparation of the
aluminium alloys. In an effort to understand the mechanism of deposition on well-
characterised substrates, many studies have focused on polished aluminium alloy
surfaces [Hagans and Haas (1994), Xia and McCreery (1999), Juffs et al. (2002),
Vasquez et al. (2002), Liu et al. (2000)], as well as on polished bulk intermetallics
[Juffs et al. (2001,2002), Vasquez et al. (2002), Lunder et al. (2005)]. There are
relatively few studies that deal with the commercial method of preparation,
including appropriate alkaline cleaning, deoxidation and conversion coating, as
used on aerospace alloys [Hughes et al. (1997), Lyttle et al. (1995)]. Several
studies use chemical pretreatment steps prior to deposition of a coating [Treverton
and Davies (1994,1997), Schram et al. (1998), Arrowsmith et al. (1984), Katzman
et al. (1979), Drodza and Maleczki (1985), Campestrini et al. (2001), Sun et al.
(2001), Meng and Frenkel (2004), Kloet et al. (2005), Vasquez et al. (2002)].
Some of these works use a deoxidation step designed to increase the surface levels
of copper. The difference between these various starting surfaces (polished or
deoxidized/copper-enriched) may have major implications for the application of
the findings to properly deoxidized surfaces. For example, Sun et al. (2001)
showed that HF/H2SO4 pretreatment of polished surfaces led to a level of copper-
enrichment. After conversion coating this copper-enriched surface, copper was
found on the external surface of the CrCC as well as the CrCC/metal interface. In
the same study, no copper was found on the external surface of the CrCC on
polished surfaces. Thus the copper left on the surface after HF/H2SO4
pretreatment had migrated onto the CrCC external surface during the conversion
Chemically Pretreated Surfaces 79

coating process. Copper-enrichment may occur via a mechanism similar to that


described in corrosion [Zahavi et al. (1978)] and anodizing [Habazaki et al.
(1975)] where small copper nanoparticles detach from the enriched layer and are
incorporated into the developing coating during its formation.
With regard to the work on polished surfaces, it is important to note the
differences between them and the chemically pretreated surfaces. Polished
surfaces have a much higher exposed area of intermetallic particles than the
chemically treated surface used in practical applications, typically 3 to 6% for
AA2024-T3 alloy [Hughes et al. (2006), Jakab et al. (2005), Buchheit et al.
(1997)], but probably less for other alloys. The oxide covering the matrix is a very
thin aluminium oxide [Nylund and Olefjord (1994)] and thus compositionally
different to that produced on the rolled surface or the deoxidized surface, as
described previously, with its GRSL (see section on surface microstructure).
Thus, the polished surface is likely to be more active than the deoxidized surface
where the intermetallic particles have been removed.
The drive to understand the deposition process on these polished surfaces has
led to a number of studies of deposition of chromate onto intermetallic phases and
significant advances have been made in the understanding of deposition onto
these types of phases [McGovern et al. (2000), Juffs et al. (2001,2002), Vasquez
et al. (2002)]. Figure 27 shows coating thicknesses over intermetallic phases is
generally thinner than the matrix in the accelerated case. McGovern et al. (2000)
studied the deposition of CCC’s onto an Al-Cu-Mg ingot containing various Al-
Cu-Mg phases. Raman spectroscopy showed that the intensity of the 860 cm-1
band, indicative of a mixed chromium oxide, decreased over phases with higher
copper content. They also noted larger carbon and nitrogen peaks at the sites of
copper-containing intermetallics using AES. Vasquez et al. (2002) examined
deposition onto various intermetallic phases in AA2024-T3 alloy, as well as onto
intermetallic phases manufactured by laser ablation, and observed the deposition
of similar species onto all phases.
Hagans and Haas (1994) used AES, XPS and ion-beam depth profiling to
study CCC formation on AA2024-T3 for formation times of up to 3 min. The rate
of film formation was faster over the matrix and slower over intermetallic phases.
It was suggested that ferrocyanide in the conversion coating formulation
interacted with copper-rich intermetallic phases to form compounds such as
Cu4Fe(CN)6 and Cu2Fe(CN)6 on the surface. All these studies show that the oxide
covering the intermetallic phases is much thinner. Copper appears to have a major
role in limiting the thickness of the coating on these phases, with Al2Cu having
the lowest coating thickness.
80 T. H. Muster, A. E. Hughes and G. E. Thompson

In terms of the coating on the matrix, it was noted above that copper-
containing alloys tend to have a reduced coating weight. As stated previously, Sun
et al. (2001) observed different copper distributions in the CrCC as a result of
different pretreatment conditions, but the model of the coating for AA7075-T6
alloy was similar to that proposed by Hughes et al. (1997) for AA2024-T3 alloy
and Treverton and Davies (1981) for pure aluminium. This model is slightly
different to that proposed by Vasquez et al. (2002) for AA2024-T3 alloy; they
observed some copper on the external surface of the CrCC due to the presence of
copper-containing intermetallics. A polished surface was used without
deoxidation and, therefore, the intermetallic phases had not been removed. More
recently Meng and Frankel (2004) examined a number of AA7xxx series alloys in
(i) polished and (ii) polished and acid etched conditions. They used an HF/H2SO4
acid etch to artificially increase the copper content at the surface prior to
chromating. Hughes et al. (2001) found that this type of pretreatment leads to a
loose copper-containing smut on the surface of AA2024-T3 alloy, which is not
typical of copper buildup on the alloy during commercial processing in a HNO3 -
based deoxidiser. However, Meng and Frankel (2004) found two breakdown
potentials in polarization curves of the polished AA7xxx surfaces. The first was
associated with the dissolution of hardening precipitates while the second was
combined intergranular and selective grain attack. The breakdown potential
increased with copper content of the AA7xxx series alloy for the polished surface,
indicating that the increased copper within the alloy improved the corrosion
resistance for both the polished, and the polished and conversion coated alloys.
The first breakdown was not observed in the conversion coated specimens,
indicating that either the conversion coating treatment had removed these phases
or that they had been passivated.
A detailed examination of a number of these works seems to suggest that the
CrCC process itself does not significantly change the prevailing copper
distribution on the alloy surface. The kinetics of the coating process, however, is
strongly influenced by the presence of copper. The work of Sun et al. (2001) best
illustrates this point, since no enrichment of copper was observed during chromate
conversion coating of polished surfaces of AA2024-T3 alloy, whereas copper,
already present on the surface due to deoxidation, remained enriched at the metal
surface beneath the CrCC, and some of this copper was present on the external
surface of the CrCC. These results suggest that, while the fluoride ions in the
coating solution (equation 14) are intended as an etchant for the underlying metal,
they only thin the surface oxide so that electron tunneling through the oxide can
promote deposition reactions, rather than having a strong etching effect on the
underlying aluminium. The role of fluoride as an etchant for the oxide is
Chemically Pretreated Surfaces 81

reinforced by the work of Campestrini et al. (2004). They anodized aluminium


prior to conversion coating, and demonstrated that deposition of the conversion
coating only began after the anodized layer was removed by fluoride attack.
Indeed, it is often overlooked that for the CrCC process, a chromate deoxidizer is
preferable. The composition of the deoxidizer solution is similar to the CrCC, but
more acidic, and the deoxidizer leaves a surface oxide rich in chromium
containing species; these could well be a nascent CrCC, so that the subsequent
immersion in the CrCC solution completes the coating process but at higher pH
where deposition reactions occur.
The fact that the CrCC may not change the copper distribution means that the
role of a copper-enriched layer on the deposition process needs to be examined.
One view [Vasquez et al. (2002)] is that the presence of copper on the surface
simply reduces the available aluminium surface area for the reaction with the
conversion coating solution. For this model to work, a surface coverage effect
could only be obtained if the enriched copper layer at the alumnium/aluminium-
oxide interface comprised thin islands (assuming appropriate deoxidation) on the
surface, rather than a continuous layer. (In the latter case the whole surface would
be inactive and no coating would precipitate.). There is no real evidence to date of
the overall distribution to support either model. Copper-enrichment arises from
four sources; the matrix, hardening precipitates, dispersoid particles and large
intermetallic particles. Since the large intermetallic particles are removed during
deoxidation, copper-enrichment is likely to be due to dissolution of the matrix,
hardening precipitates and the dispersoid particles which may provide some
granularity to the enriched layer thickness. However, the enrichment due to the
solid solution component should be continuous. Hence, it seems more likely that
the copper influences the reaction kinetics. It should be noted that commercial
formulations contain sequestering agents to complex copper in solution to
minimize the amount of plating out of dissolved copper.
To understand how the copper-enriched layer has an influence on the kinetics
of reaction of underlying alloy composition it is instructive to examine the
reactions of the CrCC solution with the intermetallic phases. Ferricyanide has
been observed to deposit onto the copper-containing intermetallic phases and form
an insoluble compound, as suggested by Hagans and Haas (1994). Juffs et al.
(2002) proposed, however, that, upon adsorption of the ferricyanide onto the
copper, it is reduced to ferrocyanide due to cathodic activity. Concurrently, the
chromate that is also adsorbed onto the copper-containing sites, is reduced to
Cr(III), not through the Fe(II)/Fe(III) couple but through cathodic reduction at the
intermetallic phase. Hence, the activity of the copper containing sites is lost,
because, with Cr(III) and ferrocyanide on the surface, there is no mechanism for
82 T. H. Muster, A. E. Hughes and G. E. Thompson

oxidation of the ferro/ferri couple, thus, removing the accelerating role of the
Fe(II)/Fe(III) couple. Under either of these mechanisms (copper ferri/ferro
cyanides or cathodic reduction of Fe(III) and Cr(VI)), the loss of the ferricyanide
at the reaction interface, would mean that the coating solution will behave more
like an unaccelerated solution giving thinner coatings over intermetallic phases as
seen in Figure 27.

Other Conversion Coatings

As mentioned earlier, other non-chromate coating formulations are also either


under development or available commercially [Buchheit and Hughes (2003),
Nylund (2000)]. Those that are available commercially include the Ti/Zr
processes [Knudsen et al. (2003), Deck and Reichgott (1992), Lunder et al.
(2004), Tomlinson (1997)], Co-based process [Schreiver (1992,1996), Roland
(1998), Hughes et al. (2004)], permanganate processes [Bibber (1993), Hughes et
al. (2006)], silane treatments, self-assembled monolayers [Schmidt-Handsberg
and Schubach (2003)] and rare earth processes. Most of these processes do not
perform adequately on copper-containing alloys [Chalmers (1995)]. This is
probably because many of these formulations are specifically designed to interact
with a hydrated aluminium oxide surface, rather than a copper-rich surface, to
form bridging Al-O-M bonds from which a coating “superstructure” can be built.
One obvious exception to this is the rare earth processes which benefit from
the presence of copper-enrichment [Hughes et al. (2004), Scholes et al. (2006),
Campestrini et al. (2004)], but these are currently limited to application to
architectural alloys [Schmidt-Handsberg and Schubach (2003)]. This coating
system is intrinsically different from the other coating systems in that it relies on
the cathodic activity of the surface for the deposition reaction. Unlike CrCCs the
coating weight increases with copper content of the alloy. Formulations contain
H2O2 which is cathodically reduced to generate a local pH rise in which the rare
earth (principally cerium species) is deposited onto the surface.
A second exception is the cobalt-based process [Schreiver (1992), Schreiver
(1996), Roland (1998)] which has a poor performance on copper-containing
alloys compared with alloys of reduced copper content [Chalmers (1995)].
CONCLUSIONS

The importance of microstructure and chemistry of aluminium alloys has


been demonstrated for copper-containing aluminium alloys. Alloying elements
such as copper are present in solid solution, intermetallic particles or both. The
distribution of alloying elements within aluminium alloys depends on alloy
composition and processing history.
Copper is one of the most noble alloying elements that is added to aluminium
alloys. Because of this property, the copper distribution has a major impact on
corrosion processes. Generally, copper-containing intermetallic particles will
facilitate cathodic reactions resulting in the anodic dissolution of the adjacent
aluminium matrix leading to pitting and other degradation reactions. In a special
case, S-phase (Al2CuMg) intermetallic particles, while initially behaving as a net
anode (i.e. undergoing dissolution), eventually become copper–rich and transform
into very efficient cathodes. Alloys containing these intermetallic particles such as
some of the AA2xxx series alloys, can expect to have severe corrosion issues
related to the presence of copper.
In metal finishing, proper pre-treatment, such as deoxidation, should remove
the majority of intermetallic particles including the copper-containing particles.
However, copper enrichment occurs just beneath the surface oxide wherever
etching processes are used. While nitric-acid containing processes remove copper-
rich smut from the surface, they cannot remove the copper enriched layer which
forms as a result of etching. This layer gives rise to some typical coating variation
of the different alloy series such as variation in coating weight during chromate
conversion coating.
The investigation and understanding of copper distributions in aluminium
alloys is part of a larger topic of microstructure and process control, fabrication,
forming and joining, surface modification and recycling to provide light alloys
84 T. H. Muster, A. E. Hughes and G. E. Thompson

that meet current requirements and future challenges. The future challenges are to
develop cheaper, stronger, more formable and lighter alloys to compete with
composites.
ACKNOWLEDGEMENTS

The authors will like to thank Mr Tim Harvey, Dr Scott Furman, Dr R.


Lumley and Dr R. Taylor for critical reading of the text and for feedback. Mr Tim
Harvey is also acknowledged for assistance with the references. The authors
would also like to thank past and present colleagues and students in the Light
alloys Group in Corrosion and Protection Centre at Manchester.
REFERENCES

Abd Rabbo, M. F.; Richardson, J. A.; Wood, G.C. Corros. Sci., 1978, 18, 117-
123.
Afseth, A.; Nordlien, J. H.; Scamans, G. M.; Nisancioglu., K. Corros. Sci. 2001,
43, 2093-2109.
Afseth, A.; Nordlien, J. H.; Scamans, G. M.; Nisancioglu, K. Corros. Sci. 2002,
44(11), 2529-2542.
Afseth, A., Nordlien, J. H., Scamans, G. M.Nisancioglu, K.; Corros. Sci. 2002,
44(11), 2543-2559.
Aldykewicz, A. J.; Isaacs, H. S.; Davenport, A. J. J. Electrochem. Soc., 1995,
142(10), 3342-3350.
Alodan, M. A.; Smyrl, W. H. J. Electrochem. Soc., 1998, 145, 1571-1577.
Altenpohl, D. Corros., 1966, 15, 143t.
Alwitt, R. S. J. Electrochem. Soc., 1974, 121, 1322-1328.
Arai, K.; Suzuki, T.; Atsumi, T. J. Electrochem. Soc., 1985, 132(7), 1667-1671.
Arrowsmith, D. J.; Dennis, J. K.; Sliwinski, P. R. Trans. Inst. Met. Finish., 1984,
6, 117-120.
ASST Proceedings, 2003. Proceedings of the 3rd International Symposium on
Aluminium Surface Science and Technology (ASST 2003), Bonn, Germany,
May 2003.
ASTM B117 Standard: In Annual Book of ASTM Standards; ASTM International,
West Conshohocken PA, Vol. 03.02, 1997.
Badawy, W.A.; Al-Kharafi, F.M. Corros. Sci., 1997, 39, 681-700.
Basset Brown, D.; Shriver, D.F.; Schwartz, L.H. Inorg. Chem., 1968, 7, 77-83.
Bastow, T. J.; Celotto, S. Acta Mater. 2003, 51, 4621-4630.
Bastow, T. J. Phil, Mag. 2005, 85, 1053-1066.
Bastow, T. J.; Hill, A. J. Mat. Sci. Forum, 2006, 519-521, 1355-1360.
88 T. H. Muster, A. E. Hughes and G. E. Thompson

Bastow, T. J.; Hill, A. J. Mat. Sci. Forum, 2006, 519-521, 1385-1390.


Bibber, J. W. Prod. Finish., 1991,56(2), 51-54.
Bibber, J. W. Met. Finish., 1993, 91(12) , 46-47.
Birbilis, N.; Buchheit, R. G.; J. Electrochem. Soc., 2005, 152(4), B140-B151.
Bjørgum, A.: Sigurdsson, H.: Nisancioglu, K. Corros., 1995, 51(7), 544-557.
Blackwood, D. J.; Chong, A. S. L. Br. Corros. J. 1998, 33(3), 225-229.
Blanc, C.; Mankowski G. Corros. Sci. 1997, 39(5), 949-959.
Blanc, C.; Lavelle, B.; Mankowski, G.. Corros. Sci. 1997, 39(3), 495-510.
Boag, A. P.; McCulloch, D. G.; Jamieson, D. N.; Hearne, S. M.; Hughes, A. E.;
Ryan, C. G.; Toh, S. K.; Nucl. Inst. Meth. Phys. Res. B, 2005, 231, 457-462.
Böhni, H.; Uhlig, H. H. J. Electrochem. Soc., 1969, 116, 906-910.
Bovard, F. V.; Moran, J.; Weiland, H. Proceedings of the workshop on
“Corrosion Modeling to Enable Corrosion Informed Material Selection to
Life Prediction”, Cernobbio, Lake Como, Italy, 29-31 July 2006, Opening
session.
Brinker, C. J.; Scherer, G. W. Sol-Gel Science: The physics and chemistry of sol-
gel processing. Academic Press, Inc., San Diego, CA., 1990, pp. 237-238.
Brown, A. E., Nature, 1949, 163, 961.
Brown, A. S.; Mat. Perform., 1992, 31(9), 55-58.
Brown, G. M.; Shimizu, K.; Kobayashi, K.; Thompson, G. E.; Wood, G. C.
Corros. Sci., 1992, 33, 1371-1385.
Brown, G. M.; Shimizu, K.; Kobayashi, K.; Thompson, G. E.; Wood, G. C.
Corros. Sci., 1993, 35, 253-256.
Buchheit, R. G. J. Electrochem. Soc,, 1995, 142, 3994-3996.
Buchheit, R. G.; Hughes, A. E. in Corrosion: Fundamentals, testing and
Protection. Crammer, S. D.; Covina, B. S, ASM Handbook, American
Society for Materials, Materials Pk Ohio, 2003; Vol 13A, pp. 720-736.
Buchheit, R. G.; Martinez, M. A.; Montes, L. P. J. Electrochem. Soc., 2000,
147(1), 119-124.
Buchheit, R. G.; Grant, R. P.; Hlava, P. F.; McKenzie, B.; Zender, G. L. J.
Electrochem Soc. 1997, 144(8), 2621-2628.
Buchheit, R. G.; Montes, L. P; Martinez, M. A.; Michael, J.; Hlava, P. F. J
Electrochem Soc., 1999, 146(12), 4424-4428.
Buchler, M.; Watari, T.; Smyrl, W. H. Corros. Sci., 2000, 42, 1661-1668.
Caicedo-Martinez, C. E. ; Koroleva, E. V. ; Thompson, G. E. ; Skeldon, P. ;
Shimizu, K. ; Habazaki, H.; Hoellrigl, G. ; Smith, G. ; Flukes G. ; Foord, D.T.
ATB Metallurgie, 2003, 43, 301-307.
References 89

Caicedo-Martinez, C. E.; Koroleva, E. V.; Thompson, G. E.; Skeldon, P.;


Shimizu, K.; Habazaki, H.; Hoellrigl, G. Surf. Int. Anal. 2002, 34(1), 405-
408.
Caicedo-Martinez, C. E.; Koroleva, E. V.; Thompson, G. E.; Skeldon, P.;
Shimizu, K.; Hoellrigl, G; Campbell, C.;. McAlpine, E. Corros. Sci. 2002,
44(1), 2611-2620.
Campestrini, P.; van Westing, E. P. M.; de Wit, J. H. W. Electrochim. Acta, 2001,
46(17), 2631-2647.
Campestrini, P.; van Westing E. P. M.; de Wit, J. H. Electrochim. Acta, 2001,
46(17), 2553–2571.
Campestrini, P.; Terryn, H.; Hovestad, A.; de Wit, J. H. W. Surf. Coat. Technol.
2004, 176, 365-381.
Campestrini P.; Goeminne, G.; Terryn, H.; Vereecken, J.; de Wit, J. H. W. J.
Electrochem. Soc., 2004, 151, B59-B70.
Carney. T. J.; Tsakiropoulos P.; Watts, J. F.; Castle, J. E. Int. J. Rap. Sol., 1990, 5,
189-217.
Chen, G. S.; Gao M;. Wei, R. P. Corros. 1996, 52(1), 8–15.
Chen, G. S.; Liao, C-M.; Wan, K-C.; Gao, M.; Wei, R. P. In: Effects of the
Environment on the Initiation of Crack Growth. Van Der Sluys, W. A.;
Piascik, R. S.; Zawierucha, R. Eds. ASTM STP 1298. American Society for
Testing and Materials, 1997, pp. 18-33.
Chalmers, P.D. Alternatives to Chromium for Metal Finishing, National Centre
for Manufacturing Sciences, Report No 0273RE95, 1995.
Chidambaram, D.; Halada, G. P. Surf. Int. Anal. 2001, 31, 1056-1059.
Critchlow, G. W.; Brewis, D. M. Int. J. Adhesion and Adhesives, 1996, 16(4),
255-275.
Crossland, A. C. ; Thompson, G. E.; Smith, C. J. E.; Habazaki, H.; Shimizu, K.;
Skeldon, P. Corros. Sci. 1999, 41, 2053-2069.
Darken, G. L. S. Trans. Metall. Soc., 1967, 239, 90–96.
Davis, J.R. Corrosion of Aluminum and Aluminum Alloys. ASM International,
Materials Park, OH, 1999: pp 11, 33-34.
Davoodi, A.; Pan, J.; Leygraf, C.Norgen, S.; J. Electrochem. Soc., 2005, 8(6),
B21-B24.
Deck P. D. ; Reichgott, D. W. Metal Finishing, 1992, 90(9), 29-35.
Dimitrov, N.; Mann, J. A.; Sieradzki, K. J. Electrochem. Soc., 1999, 146(1), 98-
102.
Di Quarto, F.; Piazza, S.; Sunseri, C. Electrochim. Acta, 1985, 30, 315-324.
Drozda T.; Maleczki, E. J. Radioanal. Nucl. Chem. Letts, 1985, 95, 339-349.
90 T. H. Muster, A. E. Hughes and G. E. Thompson

Dunn, C. G. ; Bolon, R. B.; Alwan, A. S.; Stirling, A. W., J. Electrochem. Soc.,


1971, 118, 381-390.
Du Plessis J. ; Taglauer, E. Surf. Sci., 1992, 260, 355–360.
Du Plessis, J.; van Wyk, G. N. 1988, 49, 1441-1450.
Eppensteiner F. W.; Jenkins, M. R. in: Chromate Conversion Coatings. Murphy,
M. Metal Finishing, Hackemack, NJ, Vol. 93(1A), 1995, 460-473.
Ergun, M.; Balaşi, M.; Tosun, A. Br. Corrosion J., 1997. 32, 117-120.
Fehlner, F. P.; Mott, N. F. Oxidation of Metals, 1970, 2, 59.
Fishkis, M.; Lin, J. C. Wear 1997, 206, 156-170.
Galvele, J. R.; de Micheli, S. M. Corros. Sci., 1970, 10, 795-807.
Gao, M.; Feng, C. R.; Wei, R. P. Metall. Mater. Trans. A, 1998, 29(4), 1145-
1151.
Garcia-Vergara, S. G.; Colin, F.; Skeldon, P.; Thompson, G. E.; Bailey, P.;
Noakes, T. C. Q.; Habazaki, H.; Shimizu, K. J. Electrochem. Soc., 2004,
151(1), B16-B21.
Garcia-Vergara, S. J.; Iglesias-Rubianes, L.; Blanco-Pinzon, C. E.; Skeldon, P.;
Thompson, G. E.; Campestrini, P. Proc. Roy. Soc. A, 2006, 462, 2345-2358.
Gorman, J. D. Characterisation of a Rare Earth Conversion Coatings on AA2024-
T3, Ph.D Thesis, Applied Physics, RMIT University, 1998.
Gorman, J. D.; Hughes, A. E.; Jamieson, D.; Paterson, P. J. K. Corros. Sci., 2003,
45, 1103-1124.
Guillaumin, V.; Mankowski, G. Corros. Sci., 1999, 41, 421-438.
Guillaumin, V.Mankowski, G.; Corros. Sci., 2000, 42, 105-125.
Guillaumin, V.; Schmutz, P.; Frankel, G. J. Electrochem.Soc., 2001, 148(5),
B163-B173.
Guttmann, M. Surf. Sci., 1975, 53, 213-227.
Habazaki, H., Shimizu, K., Paez, A., Skeldon, P., Thompson, G. E., Wood, G. C.,
Xhou, X. Surf. Int. Sci., 1995, 23, 892-898.
Habazaki, H.; Shimizu, K.; Skeldon, P.; Thompson, G. E.; Wood, G. C.; Zhou, X.
Corros. Sci., 1997, 39(4), 731-737.
Hagans, P. L.; Haas, C. M. Surf. Interface Anal., 1994, 21, 65-78.
Hatch, J. E., Aluminum: Properties and Physical Metallurgy, ASM, Metals Park,
OH, 1984, pp .58-104, 134-199, 200-241, 242-319, 351-378.
Heine M. A.; Pryor, M. J. J. Electrochem. Soc.,1967, 114, 1001-1006.
Hemingway, B. S.; Robie, R. A.; Apps, J. A. American Mineralogist, 1991, 76,
445-457.
Henon. C, Presented at the Aluminium Surface Science and Technology
Conference, Bonn, 2006: Proceedings to be published in: ATB Metallurgie.
Hofmann, S. J. Chim. Phys. 1987, 84, 141–147.
References 91

Hondros E. D.; Seah, M. P. Int. Met. Rev., 1977, Review 222, 262-301.
Hughes, A. E.; Boag, A. P.; Pedrina, L. M.; Juffs, L.; McCulloch, D. G.; Du
Plessis, J. ; Paterson, P. J. K.; Snook, I. K.; O’Malley, B. Presented at the
Aluminium Surface Science and Technology Conference, Bonn, 2006:
Proceedings to be published in: ATB Metallurgie.
Hughes, A. E.; Gorman, J. D.; Harvey, T. G.; Galassi, A.; McAdam, G. Corros.,
2006, 62, 773-780.
Hughes, A. E.; Gorman, J.; Harvey, T. G. ; McCulloch, D.; Toh, S. K. Surf. Int.
Anal., 2004, 36, 1585-1591.
Hughes, A. E.; Gorman, J. D.; Miller, P. R.; Sexton, B. A.; Paterson, P. J. K.,
Taylor, R. J. Surf. Int. Anal., 2004, 36, 290-303.
Hughes, A. E.; Hardin, S. G.; Harvey, T. G.; Nikpour, T.; Hinton, B. R. W.;
Galassi, A.; McAdam, G.; Stonham, A.; Harris, S. J.; Church, S.; Figgures,
C.; Dixon, D.; Bowden, C.; Morgan, P.; Toh, S.K.; Mcculloch, D.; Du
Plessis, J. ATB Metallurgie, 2003, 43(1-2), 459-466.
Hughes, A. E.; Harvey, T. G.; Nikpour, T.; Muster, T. H.; Hardin, S. G. Surf. Int.
Anal., 2004, 36, 15-23.
Hughes, A. E.; Nelson, K. J. H.; Miller, P. R. Mater. Sci. Technol., 1999, 15,
1124-1132.
Hughes, A. E.; Taylor, R. J.; Hinton, B. R. W.; Wilson, L. Surf. Int. Anal., 1995,
23, 540-550.
Hughes, A. E.; Taylor, R. J.; Hinton, B. R. W. Surf. Int. Anal., 1997, 25, 223-234.
Hughes, A. E.; Taylor, R. J.; Nelson, K. J. H.; Hinton B. R. W.; Wilson, L. Mat.
Sci. Technol., 1996, 12, 928-936.
Hughes, A. E., Taylor, R. J., Nelson, K. J. H.Hinton, B. R. W.; Materials Science
and Technology 1996, 12(11), 928.
Hughes, A. E.; Theodossiou, G.; Elliott, S.; Harvey, T. G.; Miller, P. R.; Gorman,
J. D.; Paterson, P. J. K. Mater. Sci. Technol., 17, 1642-1652.
Hunter, M. S., Robinson, D. L., J. Met., 1973, 717.
Iglesias-Rubianes, L.; Garcia-Vergara, S. J.; Skeldon, P.; Thompson, G. E. private
communication (2006).
Ilevbare, G. O.; Schneider, O.; Kelly, R. G.; Scully, J. R. J. Electrochem. Soc.,
2004, 151(8), B453-B464.
Jakab, M. A.; Little, D. A.; Scully, J. R.; J. Electrochem. Soc., 2005, 152(8),
B311-B320.
Jaime Vasquez, M.; Halada, G. P.; Clayton, C. R.; Longtin, J. P. Surf. Int. Anal.,
2002, 33, 606-616.
Jaime Vasquez, M.; Kearns, J. R.; Halada, G. P.; Clayton, C. R. Surf. Int. Anal.,
2002, 33, 796-808.
92 T. H. Muster, A. E. Hughes and G. E. Thompson

Juffs, L.; Investigation of Corrosion Coating Deposition on Microscopic and


Macroscopic Intermetallic Phases of Aluminium Alloys. Melbourne, RMIT,
2003.
Juffs, L.; Hughes, A. E.; Furman, S. A.; Paterson, P. J. K. Corros. Sci., 2002,
44(8), 1755-1781.
Juffs, L.; Hughes, A. E.; Paterson, P. J. K. Micron, 2001, 32, 777-787.
Jung, D.Y.; Dumler, I.; Metzger, M. J. Electrochem. Soc., 1985, 132, 2308-2312.
Katzman, H. A.; Malouf, G. M.; Bauer, R.; Stupian, G.W. Appl. Surf. Sci., 1979,
2, 416-432.
Kaufman, J. G. (2004). Aluminum Alloy Database, Knovel. Online source
http://www.
knovel.com.
Kendig, M. W.; Davenport, A. J.; Isaacs, H. S. Corros. Sci.,1993, 34, 41-49.
Ketcham, S. J.; Brown, S. R. Metal Finishing, 1976, 74(11), 37-41.
Kimpton, H. J.; Erricker, S. L.; Smith C. J. E.; Harris, S.J. Alum. Surf. Sci.
Technol. (Proc. 2nd Int. Symp.). Terryn, H.; Ed ,.ATB Metallurige; 2000, 467-
472.
King, R.G. Surface treatment and Finishing of Aluminium, Pergamon Press,
Oxford, 1988, p. 18.
Kloet, J. V.; Hassel A. W.; Stratmann, M. Z. Physik Chemie, 2005, 219, 1505-
1517.
Knudsen, O. Ø.; Bjørgum A.; Tanem, B.S. ATB Metallurgie, 2003, 43, 175-180.
Kolics, A.; Besing, A. S.; Wieckowski, A. J. Electrochem. Soc., 2001, 148(8),
B322-B331.
Koroleva, E. V. ; Thompson, G. E. ; Hollrigl, G. ; Bloeck, M. Corros. Sci., 1999,
41, 1475-1495.
Laevers, P.; Terryn, H.; Vereecken J.; Thompson, G.E. Corros. Sci., 1993, 35,
231-238.
Lea C.; Ball, J. Appl. Surf. Sci., 1984, 17, 344-362.
Lea C.; Molinari, C. J. Mater. Sci., 1984, 19, 2336-2351.
Leblanc, P.; Frankel, G.S.; J. Electrochem. Soc. 2002, 149(6), B239-B247.
Leclere, T. J. R.; Newman, R. C. J. Electrochem. Soc., 2002, 149(2), B52-B56.
Leggat, R. B.; Taylor, S. R.; Zhang, W.; Buchheit, R.G.; Corros., 2002, 58(3),
283-291.
Leth-Olsen, H.; Nordlein J. H.; Nisancioglu, K. J. Electrochem. Soc., 1997, 144,
L196-L197.
Leth-Olsen, H.; Nordlein J. H.; Nisancioglu, K. Corros. Sci., 1998, 40, 2051-
2063.
Liao, C. M.; Olive, J. M.; Gao, M.; Wei, R. P. Corros., 1998, 54(6), 451-458.
References 93

Liao, C.M., Wei, R. P., Electrochim. Acta, 1999, 45, 881-888.


Lide, D. R. CRC Handbook of Chemistry and Physics. 81st Ed., CRC
Press LLC, 2001,
p. 12-124.
Liu, Y.; Arenas, M. A.; Skeldon, P.; Thompson, G. E.; Habazaki, H.; Shimizu, K.;
Bailey P.; Noakes., T. C. Q. Corros. Sci., 2006, 48, 1874-1884.
Liu, Y.; Bailey, P.; Noakes, T. C. Q.; Thompson, G. E.; Skeldon, P.; Alexander,
M. R. Surf. Int. Anal., 2004, 36, 339-346.
Liu, Y.; Colin, F.; Skeldon, P.; Thompson, G. E.; Zhou, X.; Habazaki, H.;
Shimizu, K. Corros. Sci., 2003, 45, 1539-1544.
Liu, Y.; Colin, F.; Thompson, G. E.; Skeldon, P.; Alexander, M.; Shimizu, K.;
Habazaki, H.; Bailey, P.; Noakes, T. C. Q. ATB Metallurgie, 2003, 43, 391-
395.
Liu, Y.; Skeldon, P.; Thompson, G. E.; Habazaki, H.; Shimizu, K., Corros. Sci.,
2004, 46(2), 297-312.
Liu, Y.; Skeldon, P.; Thompson, G. E.; Habazaki H.; Shimizu, K. Corros. Sci.,
2005, 47(2), 341-354.
Liu, Y. Thompson, G. E. Skeldon, P. Smith, C. J. E. Shimizu, K. ATB
Metallurgie, 2001, 41, 479-483.
Luckman, Treatise on Materials Science and Technology, 1988, 30, 17-63.
Lumley, R. N.; Sercombe, T. B.; Schafer, G. B. Metall. Mat. Trans A, 1999, 30A,
457-463.
Lunder O.; Nisancioglu, K. Corros. Sic., 1987, 44, 414-422.
Lunder, O.; Simensen, C.; Yu, Y.; Nisancioglu, K. Surf. Coat. Technol., 2004,
184, 278-290.
Lunder, O.; Walmsley, J.C.; Mack, P.; Nisancioglu, K. Corros. Sci., 2005, 47,
1604-1624.
Lyttle, F. W.; Greegor, R. B.; Bibbins, G. L.; Blohowiak, K. Y.; Smith, R. E.;
Tuss, G. D. Corros. Sci.., 1995, 37, 349-369.
McGovern, W. R.; Schmutz, P.; Buchheit, R. G.; McCreery, R. L. J. Electrochem.
Soc., 2000, 147(12), 4494-4501.
Mil-C-81706ª: Detail Specification Chemical Conversion Materials For Coating
Aluminum and Aluminum Alloys, U.S. Department of Defence, 2002.
Mil-C-5541E: Military Specification Chemical Conversion Coatings on
Aluminum and Aluminum Alloys, U.S. Department of Defence, 1990.
Millet, B.; Fiaud, C.; Hinnen C.; Sutter, E. M. M. Corros. Sci., 1995, 37, 1903-
1918.
94 T. H. Muster, A. E. Hughes and G. E. Thompson

Moffitt, C. E.; Wieliczka D. M.; Yasuda, H. K. Surf. Coat. Technol., 2001, 137,
188-196.
Mol, J. M. C.; de Wit, J. H.; Van der Zwaag, S. J. Mater. Sci., 2002, 37, 2755-
2758.
Meng, Q.; Frenkel, G.S. Corr., 2004, 60, 897-905.
Montiero, F. J.; Barbosa, M. A.; Ross D. H.; Gabe, D. R. Surf. Int. Anal., 1991,
17, 519-528.
Montiero, F. J.; Barbosa, M. A.; Gabe, D. R.; Ross D. H. Surf. Coat. Technol.,
1988, 35, 321-331.
Moon, S.-M.Pyun, S.-I.; Corros. Sci., 1997, 39(2), 399-408.
Moon, S.-M.Pyun, S.-I.; Electrochim. Acta, 1999, 44, 2445-2454.
Muller, I. L.; Galvele, J. R. Corros. Sci., 1977, 17, 179-193.
Muster, T. H.; Hughes, A. E. J. Electrochem. Soc., 2006, 153(11), B474-B485.
Muster, T. H.; Hughes, A. E.; Harvey, T. G.; Nikpour, T.; Hardin, S. G.
Proceedings of the 9th International Conference on Aluminium Alloys,
Brisbane, Australia. Institute of Materials Engineering Australasia Ltd., 2004,
pp. 1243-1248.
Muster, T. H. 2007. Unpublished data.
Nairn, K. M.; Gable, B. M.; Stark, R.; Ciccosillo, N.; Hill, A. J.; Muddle, B. C.;
Bastow T. J. Materials Science Forum, 2006, 591, 519-521.
Nelson, K. J. H.; Hughes, A. E.; Taylor, R. J.; Hinton, B.R.W.; Wilson L.;
Henderson, M. Mat. Sci. Technol., 2001, 17, 1211-1221.
Németh, Z.; Gáncs, L.; Gémes G.; Kolics, A. Corros., 1998, 40, 2023-2027.
Newman, R. C.; Sieradzki, K. Science, 1994, 263, 1708-1709.
Nisancioglu, K. J. Electrochem. Soc., 1990, 137(1), 69-77.
Nylund, A. Aluminium Transactions, 2000, 2, 121-137.
Nylund A.; Olefjord, I. Surf. Int. Anal., 1994, 21, 283-289.
Nylund A.; Olefjord, I. Surf. Int. Anal., 1994, 21, 290-297.
Osborne, J. H. Prog. Org. Coat., 2001, 41, 280-286.
Park, J. O.; Apik, C. H.; Huang Y. H.; Alkire, R. C. J. Electrochem. Soc., 1999,
146, 517-523.
Pickens J. R.; Langan, T. J. Metal. Trans. A, 1987, 18A, 1735-1744.
Pijolat, M.; Chiavazza V.; Lalauze, R. Appl. Surf. Sci., 1988, 31, 179-188.
Pocius, A.V. Symposium on Adhesion Aspects of Polymeric Coatings.
Minneapolis, 1981. Mittal, K.L., Plenum Press, NY, 1983, p. 173.
Polmear, I. J. Light Alloys: Metallurgy of the Light Metals. 2nd Ed.; Honeycombe,
R. W. K.; Hancock, P.; Eds.; Edward Arnold, London, 1989, pp 8-12, 18-53,
54-143.
Palomino, L.E.M.; Aoki, I.V.; de Melo, H.G. Electrochim. Acta, 2006, 51, 5943.
References 95

Pourbaix, M.; Atlas of Electrochemical Equilibria in Aqueous Solutions.


Pergamon Press,1966, pp. 168-176, 384-392.
Pride, S. T.; Scully, J. R.; Hudson, J. L. J. Electrochem. Soc., 1994, 141(11),
3028-3040.
Pronko, P. P.; Bhattacharya, R. S.; Kleek, J. J.; Froes, F. H. Metall. Trans., 1988,
19, 1372-1374.
Pryor, M. J. Oxidation of Metals, 1971, 3, 253.
Richardson, T. J.; Slack, J. L.; Rubin, M. D. Electrochim. Acta, 2001, 46(13-14):
2281-2284.
Riotinto Website Link:
http://www.riotinto.com/investor/databook/downloads/commodities/
comm_alumin.pdf#search=%22aluminium%20consumption%22 Accessed
14/12/2006.
Roberts, A.; Engelberg, D.; Liu, Y.; Thompson, G. E.; Alexander, M. Surf. Int.
Anal., 2002, 33, 697-703.
Roland, W. A. ATB Metallurgie, 1998, 38, 51-57.
Russell, W. J. J. Appl. Poly. Sci:, Appl. Poly. Sym., 1977, 32, 105-117.
Russell, W. J.; Garnis, E. A. SAMPE Quart., 1976, 7, 5-12.
Scamans, G.M.; Afseth, A.; Thompson G. E.; Zhou, X. ATB Metallurige, 2003,
43, 90-94.
Scamans G.M.; Butler, P. Metal. Trans., 1975, 6A, 2055-2063.
Schey, J. A. Tribology in metalworking: Friction, Lubrication and Wear.
American Society for Metals, Metals Park, OH, 1983, p. 85.
Schmidt-Hansberg, Th.; Schubach, P. ATB Metallurgie, 2003, 43, 9-14.
Schmitt, G. F. Sampe J., 1998, 43, 32-36.
Schneider, O.; Ilevbare, G. O.; Scully, J. R.; Kelly, R. G. J. Electrochem. Soc.,
2004, 151(8), B465-B472.
Scholes, F. H.; Furman, S. A.; Hughes A. E.; Markley, T. A. Corros. Sci. 2006,
48, 1812-1826.
Schram, T.; De Laet, J.; Terryn, H. J. Electrochem. Soc., 1998, 145, 2733-2739.
Schreiver, M.P. Nonchromated Oxide Coatings For Aluminium Substrates, EU
Patent No. 488430, 1992.
Schreiver, M.P. Nonchromate Cobalt Complex Conversion Coating For
Aluminium Substrates, World Intellectual Property Organisation Patent No.
9605335, 1996.
Seah, M.P. J Phys F: Metal Phys, 1980, 10, 1043-1064.
Scully, J. R.; Knight, T. O.; Buchheit, R. G.; Peebles, D. E. Corros. Sci., 1993, 35,
185-195.
96 T. H. Muster, A. E. Hughes and G. E. Thompson

Seegmiller, J. C.; Bazito, R. C.; Buttry, D. A. J. Electrochem. Soc., 2004, 7(1),


B1-B4.
Shimizu, K.; Kobayashi, K.; Thompson, G. E.; Skeldon P.; Wood, G. C. Corros.
Sci, 1997, 39, 281-284.
Short, E.P.; Shearsby, P.G. Trans. Inst. Met. Finish., 1969, 47, 27-30.
Sieradzki, K. J. Electrochem. Soc., 1993, 140(10), 2868-2872.
Siripala, W.; Kumra, K. P. Semicond. Sci. Tech., 1989, 4, 4465-4469.
Skeldon, M.; Shimizu, K.; Skeldon, P.; Thompson G. E.; Wood G. C. Corros.
Sci., 1995, 37(9), 1473-1488.
Smith, D.J.E.; Baldwin, K.R.;, Hewins, M.A.H., Gibson, M.C. Progress in the
Understanding and Prevention of Corrosion Vol II; Barcelona, Spain; The
Institute of Materials, 1 Carlton House Tce. London, 1993, pp 1652 – 1663.
Sotoudeh, K.; Nguyen, T. H.; Foley R. T.; Brown, B. F. Corros., 1981, 37, 358-
363.
Strehblow, H. H.; Melliar-Smith, C. M.; Augustyniak, W. M. J. Electrochem.
Soc., 1978, 125, 915-919.
Starke, E. A.; Staley, J. T. Progress in Aerospace Sciences, 1996, 32, 131-172.
Strehblow, H. H.; Melliar-Smith, C. M.; Augustyniak, W. M. J. Electrochem.
Soc., 1978, 125, 915-919.
Sun, T. S.; Chen, J. M.; Venables, J. D.; Hopping, R. Appl. Surf. Sci., 1978, 1,
202.
Sun, T. S.; McNamara, D. K.; Ahern, J. S.; Chen., J. M.; Ditchek, B.; Venables, J.
D. Appl. Surf. Sci., 1980, 5, 406.
Sun, X.; Li, R.; Wong. K. C.; Mitchell K. A. R.; Foster, T. J. Mater. Sci., 2001,
36, 3215-3220.
Sutter, E. M. M.; Millet, B.; Fiaud C.; Lincot, D. J Electroanal. Chem., 1995, 386,
101-109.
Svenningsen, G.; Hurlen Larson, M.; Lein, J. E.; Nordlien, J. H.; Nisancioglu, K.
Proceedings of the 9th International Conference on Aluminium Alloys,
Brisbane, Australia. The Institute of Materials Engineering Australasia Ltd.,
2004, pp. 818-824.
Szklarska-Smialowska, Z. Corrosos. Sci., 1999, 41, 1743-1767.
Takahashi, H. In: ASM Handbook, Corrosion, Davis, J. R.; Ed.; Vol. 13A; ASM
International: Materials Park, OH, 2003; pp 736–740.
Textor, M.; Amstutz, M. Anal. Chim. Acta, 1994, 297, 15-26.
Treverton, J.A.; Amor, A. J. Microscopy, 1985, 40, 383-393.
Thompson, G. E.; Xu, Y.; Skeldon, P.; Shimizu, K.; Hau, H.S.; Wood, G. C.
Philos. Mag., 1987, B55, 651-667.
References 97

Toh, S. K.; Hughes, A. E.; McCulloch, D. G.; duPlessis J.; Stonham, A. Surf. Int.
Anal., 2004, 36, 1523-1532.
Toh, S. K.; McCulloch, D. G.; du Plessis, J.; Paterson, P. J. K.; Hughes, A. E.;
Jamieson, D.; Rout, B.; Long, J. M.; Stonham, A. Surf. Rev. Letts, 2003, 10,
365-372.
Tomlinson, C. E. Elecrochem. Soc. Proceed. (Environ. Acceptable Inhib. Coat.),
1997, 95, 159-168.
Trathen, P.; Hinton, B.; Hammon, K. A.; Hughes, A. E.; Taylor, R. Proceedings
of the Australian Corrosion Association Conference, Newcastle, November,
1993, p.1-12.
Treverton, J. A., Amor, M. P. J. Microscopy, 1983, 140, 383-393.
Treverton J. A.; Davies, N. C. Surf. Int. Anal., 1981, 3, 194-200.
Treverton J. A.; Davies, N. C. Metals Tech., Oct 1977, 480-489.
Underhill, P. R.; Rider, A. N. Surf. Coat. Technol., 2005, 192, 199-207.
Vander Voort, G. F. ASM Handbook: Metallography and Microstructure, Vander
Voort, G. F.; ASM International, Metals Park, OH, 2004, Vol. 9, pp 107-115,
711-751.
Vedder W.; Vermilyea, D. A. Trans. Far. Soc., 1969, 65, 561-589.
Vermilyea, D. A.; Vedder, W. Trans. Far. Soc., 1970, 66, 2644-2654.
Venables, J. D.; McNamara, D. K.; Chen J. M.; Sun, T. S. Appl. Surf. Sci., 1979,
3, 88-98.
Viswanadham, R. K.; Sun T. S.; Green, J. A. S. Corros., 1980, 36, 275-278.
Vukmirovic, M. B.; Dimitrov, N.; Sieradzki, K. J. Electrochem. Soc., 2002,
149(9), B428-439.
Vukmirovic, M. B.; Vasiljevic, N.; Dimitrov, N.; Sieradzki, K. J. Electrochem.
Soc., 2003, 150(1), B10-B15.
Wernick, S.; Pinner, R.; Shearsby, P. G. The Surface Treatment and Finishing of
Aluminium and its Alloys, 5th Ed. Finishing Publications Ltd., Teddington,
UK, 1996, p. 205.
Wilsdorf, H. G. F. Nature, 1951, 168.
Wu, X.; Asoka-Kumar, P.; Lynn, K. G.; Herbert, K. R. J. Electrochem. Soc.,
1994, 141, 3361-3368.
Wu X.; Herbert, K. J. Electrochem. Soc., 1996, 143, 83-91.
Xia L.; McCreey, R. L. J. Electrochem. Soc., 1999, 146, 3696-3701.
Yoon, Y.; Buchheit, R. G. J. Electrochem. Soc., 2006, 153(5), B151-B155.
Zahavi, J.; Zangvil, A.; Metzger, M. J. Electrochem. Soc., 1978, 125, 438-444.
Zhang, W.; Frankel, G.S. Electrochim. Acta, 2003, 48, 1193-1210.
98 T. H. Muster, A. E. Hughes and G. E. Thompson

Zhou, X.; Thompson, G. E.; Skeldon, P.; Habazaki, H.; Shimizu, K.; Wood, G.
C.; Paez, M. A.; Scamens, G. M. Proceedings of the 14th International
Corrosion Congress, Capetown, South Africa, 1999.
INDEX

alkaline, 31, 39, 46, 47, 51, 53, 54, 55, 57, 60,
# 64, 74, 78
alkalinity, 42
2D, 41, 42
alloys, vii, 1, 2, 3, 5, 6, 8, 9, 10, 11, 12, 13,
14, 15, 16, 17, 18, 19, 21, 23, 25, 26, 27,
A 28, 29, 30, 31, 32, 33, 34, 38, 39, 41, 44,
45, 49, 51, 53, 54, 55, 56, 57, 58, 59, 65,
absorption, 56, 77 67, 68, 69, 70, 73, 74, 78, 79, 80, 82, 83, 85
accelerator, 75, 77 aluminium, vii, 1, 3, 5, 6, 8, 9, 10, 11, 12, 15,
acetic acid, 52 17, 18, 19, 21, 22, 23, 25, 26, 27, 28, 29,
acetone, 52 30, 31, 32, 33, 35, 36, 38, 39, 41, 42, 43,
acid, 32, 37, 49, 50, 52, 55, 59, 61, 68, 70, 72, 45, 46, 48, 49, 51, 52, 54, 55, 56, 57, 58,
73, 74, 75, 80, 83 59, 60, 61, 63, 64, 65, 66, 67, 68, 69, 70,
acidic, 3, 37, 39, 48, 51, 64, 81 71, 73, 74, 75, 77, 78, 79, 80, 81, 82, 83
acidity, 47, 52 aluminium alloys, vii, 1, 2, 3, 5, 8, 9, 11, 12,
adhesion, 73, 75 15, 17, 18, 19, 21, 23, 26, 28, 29, 30, 31,
adsorption, 40, 53, 56, 81 33, 38, 39, 41, 45, 54, 55, 56, 57, 59, 65,
aerospace, 1, 2, 69, 73, 74, 78 67, 69, 70, 74, 78, 83
AFM, 34 aluminum, 44, 49, 56, 89, 90, 92, 93
Africa, 98 aluminum surface, 56
Ag, 12, 13 ammonium, 71
age, 18, 46 amorphous, 25, 26, 44
ageing, 6, 8, 10, 17, 21, 33 anions, 38, 73
agents, 51, 81 anode, 41, 43, 48, 83
aging, 2, 8, 18, 32 application, vii, 1, 49, 51, 78, 82
aid, 36 aqueous solution, 25, 28, 56
air, 26 aqueous solutions, 25
aircraft, 1, 2, 3, 11, 13, 18 argument, 45
alkali, 49, 51 artificial, 8, 32
ASTM, 27, 50, 87, 89
Atlas, 95
100 Index

atoms, vii, 31, 54, 56, 70 chemicals, 53, 55


attacks, 52, 74 chemistry, vii, 23, 38, 49, 50, 51, 53, 55, 57,
attention, 16 64, 65, 67, 75, 83, 88
Australasia, 94, 96 chloride, 28, 37, 38, 39, 41, 43, 45
Australia, 94, 96 Chloride, 38, 52
automotive, 1, 6, 17 chromium, 9, 10, 17, 18, 59, 60, 75, 77, 78,
availability, 39 79, 81, 89
cladding, 11, 16
classes, 8, 16, 17, 21, 75
B cleaning, 51, 52, 53, 54, 57, 60, 64, 73, 74, 78
clustering, 9, 14, 15, 18, 56
backscattered, 12, 43, 63
clusters, vii, 12, 31, 42, 46, 47, 70
band gap, 54
Co, 73, 82
barrier, 39, 70, 71, 73
coatings, 51, 53, 54, 58, 59, 61, 66, 73, 75, 76,
behaviours, 31
82
benefits, 73
cobalt, 82
binding, 55
commercial, 1, 16, 28, 45, 51, 56, 60, 74, 78,
binding energy, 55
80, 81
blocks, 1, 77
commodities, 95
Boeing, 2, 3
communication, 91
boiling, 56
community, 49, 50
bonding, 3, 56, 58, 65, 69
complexity, 5
bonds, 82
components, 8, 9, 51, 53, 57, 59, 64, 65, 75
borderline, 48
composites, 84
bounds, 5
composition, 6, 14, 16, 21, 22, 23, 38, 49, 53,
brass, 39
55, 57, 61, 62, 81, 83
breakdown, 25, 33, 34, 37, 38, 42, 80
compositions, vii, 5, 14, 59
by-products, 48
compounds, 2, 30, 55, 70, 79
concentration, 5, 19, 26, 28, 37, 39, 43, 47,
C 59, 65, 70, 71
conductivity, 28
carbon, 52, 79 conductor, 54
carbonyl groups, 52 Congress, iv, 98
cast, 1, 8, 19 consensus, 41
casting, 1 construction, 1
catalyst, 77 continuing, 20
catalysts, 77 control, 6, 10, 18, 38, 40, 50, 83
cathode, 38, 41, 48, 54, 57 controlled, 39, 69
cations, 78 convection, 46
cavities, 20 conversion, 3, 37, 49, 51, 53, 58, 59, 60, 61,
CCC, 79 73, 74, 75, 77, 78, 79, 80, 81, 83
cell, 47, 73 cooling, 8, 39
cerium, 41, 61, 82 copper, vii, 2, 3, 5, 6, 9, 10, 11, 12, 13, 16, 17,
Chalmers, 73, 82, 89 18, 22, 23, 25, 26, 27, 28, 29, 30, 31, 32,
chemical, 31, 35, 49, 53, 57, 59, 69, 78 33, 34, 38, 39, 40, 41, 42, 43, 44, 45, 46,
Index 101

47, 48, 49, 51, 52, 53, 54, 55, 57, 58, 59, electrochemical, 3, 25, 26, 27, 28, 29, 31, 32,
60, 61, 63, 64, 65, 66, 67, 68, 69, 70, 71, 44, 69, 70
72, 73, 74, 75, 76, 78, 79, 80, 81, 82, 83 electrochemical reaction, 3
copper oxide, 23, 28 electrochemistry, vii
correlation, 15 electrolyte, 26, 28, 29, 36, 43, 69, 71, 73
correlation function, 15 electrolytes, 29, 70, 73
corrosion, vii, 2, 3, 8, 9, 11, 13, 14, 16, 17, 18, electron, 12, 13, 26, 34, 40, 43, 44, 54, 58, 60,
19, 25, 27, 28, 29, 30, 31, 32, 33, 34, 35, 62, 63, 66, 68, 72, 74, 80
38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, electron microscopy, 12, 68
50, 51, 52, 53, 57, 73, 74, 79, 80, 83 electronic, iv, 3, 28
corrosive, 25, 31, 41, 47 electronics, 1
coverage, 39, 67, 81 electrons, 54
covering, 21, 32, 46, 58, 79 electrostatic, iv
cracking, 18, 20 energy, 19, 26, 28, 31, 39, 45, 55
CRC, 93 engineering, 1
cross-sectional, 15 engines, 2
cyanide, 77 environment, 3, 21, 23, 25, 38, 39, 52, 58
environmental, 53
equilibrium, 5, 65, 73
D etching, 31, 35, 51, 53, 54, 55, 56, 57, 60, 61,
62, 63, 64, 67, 80, 83
decomposition, 18
ethane, 52
defects, 28, 54
ethanol, 43, 52, 70
deformation, 21
Ethanol, 52
degradation, 69, 83
Europe, 19, 57
degree, 2, 10, 18
evaporation, 19
density, 14, 21, 36
evidence, 32, 33, 42, 45, 46, 58, 78, 81
deposition, 39, 48, 58, 61, 63, 64, 74, 75, 77,
expert, iv
78, 79, 80, 81, 82
exposure, 21, 39, 56
deposits, 44, 60
external environment, 3
diffusion, 19, 21, 31, 47, 71
extrusion, 17, 19
dispersion, 16
distilled water, 26
distribution, vii, 2, 11, 13, 14, 15, 16, 18, 32, F
40, 49, 69, 72, 80, 81, 83
distribution function, 18 fabrication, 83
donor, 54 failure, 74
Doppler, 54 family, 34, 73, 74
fatigue, 11
faults, 19
E feedback, 85
Fermi, 26
earth, 65, 73, 82
Fermi level, 26
electric field, 69, 70
filiform, 16, 18, 19, 30
electrical, 1, 42, 56
film, 25, 32, 35, 56, 57, 59, 69, 70, 71, 72, 73,
electrical resistance, 56
74, 79
102 Index

film formation, 70, 72, 73, 79 H2, 3, 61, 63, 64, 65, 78, 80
film thickness, 32, 57, 71 handling, 6
films, 28, 57, 70, 71, 72 hardness, 15
fines, 20 health, 49, 53
flow, 46, 73 heat, 1, 2, 5, 8, 9, 16, 17, 19, 21, 27, 32
fluorescence, 50 heavy metal, 39
fluoride, 51, 60, 65, 69, 74, 80 heavy metals, 39
fluoride ions, 80 heterogeneous, 28, 34, 37
fluorides, 53 high pressure, 72
fluorine, 59, 75 high temperature, 5, 9, 19
foils, 1 host, 30
folding, 21 House, 96
food, 6 household, 1
fracture, 2, 15, 21 hydrodynamic, 46
fragmentation, 48 hydrogen, 35, 38
France, 1 hydrogen gas, 35, 38
free energy, 19, 26, 27, 32, 45 hydroxide, 35, 36, 54, 77
hydroxides, 25, 38

G
I
gallium, 56
gas, 35, 38, 71, 72, 73 images, 35, 41, 43, 44, 63, 66
gauge, 14 imaging, 34
gel, 77, 78, 88 immersion, 49, 55, 57, 60, 61, 64, 74, 75, 76,
gels, 36, 48 81
generation, 38, 45, 53, 71, 72, 73, 77 impurities, 11, 16, 59
Germany, 87 inactive, 81
Gibbs, 26, 27 indication, 33, 58
Gibbs free energy, 26, 27 industrial, 53
gold, 27 industry, 1, 2, 11, 49, 50, 51, 52, 58, 69, 73
grain, 6, 10, 13, 14, 18, 19, 21, 32, 33, 40, 71, inhibitor, 53
80 inhibitors, 52
grain boundaries, 6, 13, 21, 32, 33, 40 initiation, 14, 16, 33, 34, 38, 39, 40, 44, 57
grain refinement, 14 initiation rates, 34
grains, 12, 18 injury, iv
greek, 10 intensity, 75, 79
groups, 52 interface, 9, 28, 31, 32, 53, 59, 61, 66, 69, 71,
growth, 10, 32, 33, 36, 38, 39, 40, 45, 56, 57, 72, 73, 74, 78, 81, 82
60, 70, 73, 74 intermetallic compounds, 30, 70
growth rate, 70 intermetallics, 2, 6, 14, 18, 21, 29, 30, 32, 34,
35, 36, 37, 38, 39, 44, 47, 56, 67, 74, 78,
79, 80
H ions, 28, 32, 35, 38, 39, 44, 45, 47, 51, 56, 59,
69, 72, 80
H19, 58
Index 103

iron, 9, 11, 16, 17, 22, 23, 28, 30, 54, 55, 56, mass loss, 19
60, 61, 63, 77, 78 matrix, 3, 6, 9, 10, 13, 21, 23, 29, 30, 31, 32,
isolation, 42, 47 33, 34, 36, 39, 41, 42, 44, 46, 47, 48, 50,
Italy, 88 54, 55, 56, 60, 61, 64, 65, 67, 69, 72, 75,
76, 78, 79, 80, 81, 83
mechanical, iv, vii, 1, 3, 8, 9, 11, 17, 18, 19,
J 20, 21, 23, 42, 45, 47, 49
mechanical properties, vii, 3, 8, 9, 11, 17, 18
Jung, 31, 32, 92
mechanical treatments, 49
media, 3
K melting, 5
metal ions, 51
kerosene, 52 metallography, 50
kinetics, 37, 40, 58, 80, 81 metals, 26, 27, 39, 55
King, 52, 53, 55, 92 methanol, 52
metric, 58
microelectrode, 29
L microscopy, 12, 33, 37, 77
microstructure, 5, 9, 10, 11, 12, 16, 18, 19, 21,
laser, 79 22, 28, 37, 79, 83
laser ablation, 79 military, 2
lattice, 10 mobility, 28, 72
leaching, 45 models, 59, 67, 69, 77
lead, 28, 41, 44, 45, 47, 55, 71 monolayers, 73, 82
legislation, 53 Moon, 35, 36, 94
lifetime, 2, 32, 54 morphology, 41, 50, 53, 58
linear, 71 movement, 45
literature, 58, 60, 78
lithium, 19
Lithium, 1, 2 N
lithography, 1
localised, 28, 74 NaCl, 27, 29, 34, 35, 39, 42, 43, 45
London, 94, 96 nanometer, 45
lubricants, 20 nanometer scale, 45
nanometers, 10
nanoparticles, 32, 42, 44, 54, 79
M natural, 8, 73
network, 46, 47, 58, 59, 65, 69
magnesium, 5, 6, 9, 11, 16, 17, 18, 19, 21, 22, New York, iii, iv
23, 26, 27, 28, 41, 43, 45, 46, 53, 54, 55, Ni, 6
61, 64 nitric acid, 50, 55, 70, 74, 75
magnetic, iv nitrogen, 79
magnetron, 74 NMR, 12
maintenance, 50 nodules, 39, 58
manganese, 9, 10, 16, 18, 21, 23, 30, 54, 55, normal, 34
61, 63 North America, 57
104 Index

phosphate, 55, 56, 73


O phosphates, 53
physics, 88
observations, 32, 41, 64
plastic, 73
occlusion, 56
plasticity, 73
occupational, 53
play, 40, 50, 51, 52
occupational health, 53
polarization, 42, 46, 80
Ohio, 88
polarized, 36, 45
oil, 6
poor, 28, 75, 82
oils, 51, 53
poor performance, 82
organic, 49, 53
population, 14, 50
organic chemicals, 53
pore, 73
orientation, 32, 71
porosity, 20, 72, 73
orthorhombic, 47
porous, 25, 70, 72
oxidants, 59, 61, 63, 65
positron, 54
oxidation, 19, 36, 44, 45, 52, 60, 64, 69, 71,
positrons, 54
73, 77, 82
powders, 19
oxide, 19, 20, 21, 22, 23, 25, 27, 28, 31, 32,
precipitation, 2, 6, 9, 10, 11, 13, 17, 18, 19,
33, 34, 35, 38, 41, 44, 46, 50, 52, 53, 54,
21, 28, 32, 48
55, 56, 57, 58, 59, 60, 61, 64, 65, 66, 67,
prediction, 66
68, 69, 70, 73, 74, 75, 77, 79, 80, 81, 82, 83
preparation, iv, 49, 52, 78
oxides, 20, 21, 25, 27, 28, 38, 46, 53, 54, 55,
pressure, 49, 71
57, 58, 59, 60, 61, 64, 65
private, 91
oxygen, 22, 28, 35, 37, 38, 58, 60, 71
probability, 38
oxyhydroxides, 25
probe, 43
process control, 83
P production, 17, 38
promote, 40, 80
packaging, 1, 6 propagation, 34
paper, 68 property, iv, 83
particle density, 14 protection, 3, 11, 51, 59
particles, 6, 8, 9, 10, 11, 13, 14, 15, 16, 18, 19, protons, 38
21, 23, 30, 32, 33, 34, 36, 39, 40, 41, 44, pseudo, 56
45, 46, 47, 48, 49, 54, 55, 56, 60, 61, 63,
64, 65, 66, 67, 75, 76, 77, 78, 79, 81, 83
Q
passivation, 46
passive, 37
quality control, 50
percolation, 47
percolation theory, 47
performance, vii, 2, 8, 13, 23, 30, 32, 48, 50, R
51, 57, 82
pH, 25, 26, 35, 36, 38, 39, 43, 45, 46, 55, 75, radial distribution, 18
78, 81, 82 radius, 18, 32, 38, 45
pH values, 35, 45 Raman, 79
phase diagram, 5, 6, 7, 11 Raman spectroscopy, 79
Index 105

random, 15 separation, 48
range, vii, 1, 3, 5, 6, 10, 14, 17, 21, 29, 31, 43, series, vii, 2, 3, 5, 10, 11, 13, 14, 16, 17, 18,
49, 52, 54, 57, 58, 59, 61, 73, 74, 76 21, 25, 28, 30, 33, 41, 49, 65, 70, 74, 80, 83
rare earth, 59, 73, 82 services, iv
rare earths, 59 shape, 15
reaction rate, 30, 34, 36, 43 shipping, 1
reactivity, 25 silane, 73, 82
reading, 85 silica, 49
reagents, 52 silicate, 55
recession, 72 silicates, 53
recycling, 83 silicon, 9, 11, 14, 16, 17, 19, 28, 55, 61, 64
redistribution, vii, 42, 45, 46 sites, 14, 30, 36, 38, 39, 46, 47, 48, 59, 77, 79,
redox, 65, 77 81
reduction, 3, 25, 26, 28, 35, 36, 37, 38, 39, 40, sol-gel, 88
47, 77, 81 solid phase, 5
reflection, 26 solid state, 10
relationship, 2, 15, 60, 64, 78 solidification, 9, 10
relaxation, 47 solubility, 25, 51, 55
relaxation process, 47 solutions, 39, 45, 50, 51, 53, 54, 57, 59, 75, 76
relaxation processes, 47 solvent, 51, 52
relevance, 9, 52 solvents, 52
repair, 72, 73 South Africa, 98
research, 13, 16, 31, 38, 49, 50, 52, 58 Soviet Union, 2
resistance, 2, 3, 11, 17, 25, 30, 38, 51, 53, 56, Spain, 96
57, 73, 74, 80 spatial, 14, 15
resolution, 14 speciation, 52
retention, 61 species, 53, 56, 57, 61, 71, 75, 77, 78, 79, 81,
risk, 49 82
rods, 12 spectra, 12
rolling, 13, 14, 19, 20, 21, 64 spectroscopy, 79
room temperature, 56, 57, 64 S-phase, 12, 14, 16, 21, 29, 30, 34, 35, 41, 42,
43, 44, 46, 47, 48, 49, 60, 83
sponges, 44, 45
S stability, 25, 28, 46
stabilize, 75
S phase, 11
stabilizers, 52
salt, 43, 55, 74
stages, 10, 38, 47, 57, 64, 65
salts, 38
standard operating procedures, 50
sample, 15, 45, 46, 52
standards, 51, 87
sampling, 15
statistics, 14
Scanning electron, 66
steady state, 42
scanning electron microscopy (SEM), 34, 68
stock, 9
science, 50
storage, 21, 23, 69
search, 95
strain, 8, 9
segregation, 19
strength, vii, 1, 2, 6, 8, 9, 11, 17, 69, 73
self limiting, 75
106 Index

stress, 2, 17, 18, 73 transmission, 12, 72


stretching, 8 transmission electron microscopy (TEM), 12,
students, 85 50
substrates, 78 transport, 1, 18, 42, 51
sulphate, 43 treatable, 8, 9
summaries, 11 trend, 53
Sun, 56, 69, 76, 78, 80, 96, 97 tribological, 18
supply, 48 tribology, 19
surface area, 13, 41, 43, 45, 67, 81 trichloroethylene, 52
surface chemistry, 49, 50, 58 tunneling, 74, 80
surface diffusion, 31, 47
surface energy, 45
surface layer, 19, 21, 22, 25 V
surface modification, 83
vacancies, 35
surface region, 21
vacuum, 26
surface structure, 21, 28
values, 32, 33, 35, 42, 45, 65, 75
surface treatment, 58
variables, 67
surfactants, 53
variation, 14, 22, 51, 75, 83
susceptibility, vii, 2, 16, 18, 19, 33, 55
visual, 43, 44, 50
switching, 37
voids, 71
symbols, 10
systematic, 31
systems, 11, 26, 39, 41, 53, 67, 82 W

warrants, 14
T
water, 8, 26, 28, 35, 36, 38, 52, 56, 57, 60
wear, 20
temperature, 5, 9, 10, 17, 55, 56, 57, 59, 63,
weight ratio, 1
64, 65, 70, 77
welding, 17
tensile, 18
workers, 33, 47, 64
tensile strength, 18
theory, 19, 47
thermal, 10 X
thermodynamic, 19
Ti, 73, 82 XPS, 50, 52, 54, 55, 58, 60, 79
time, 1, 10, 32, 48, 55, 56, 65, 70, 74, 75
time periods, 48
titanium, 10 Z
tolerance, 11
toughness, 2 zinc, 5, 9, 17, 18, 22, 23, 26, 27, 53, 54, 55, 64
transfer, 39 zinc oxide, 64
transformation, 20 zirconium, 10, 18
transition, 53, 55 Zn, 5, 6, 7, 10, 17, 18, 21
transition metal, 53, 55

You might also like