You are on page 1of 27

Corrosion Science 46 (2004) 109135

www.elsevier.com/locate/corsci

Electrochemical corrosion of unalloyed copper


in chloride mediaa critical review
G. Kear

a,*

, B.D. Barker b, F.C. Walsh

c,*

Erosion-Corrosion Group, Division of Mechanical Engineering, University of Queensland,


Brisbane QLD 4072, Australia
b
Applied Electrochemistry Group, Centre for Chemistry, University of Portsmouth PO1 2DT, UK
c
Electrochemical Engineering Group, Department of Chemical Engineering, University of Bath,
Claverton Down, Bath BA2 7AY, UK
Received 28 August 2002; accepted 5 December 2002

Abstract
The literature dealing with the electrochemical corrosion characteristics of unalloyed copper
in aqueous chloride media is examined. The enormous quantity of polarisation and mixed/
corrosion potential data that has been made available in the literature over the last 50 years
has been compiled and discussed in a comprehensive review. For a wide range of electrode
geometries, the importance of the chloride ion and the mass transport of anodic corrosion
products on the corrosion behaviour of copper are made clear for both freshly polished and
lmed surfaces.
 2003 Elsevier Ltd. All rights reserved.
Keywords: A. Copper; B. Polarisation; RDE/RCE; C. Oxygen reduction

1. Introduction
Pure copper is rarely used as a corrosion resistant material within the marine
environment. Copper-based alloys, however, such as the copper nickels, aluminium
bronzes and nickel aluminium bronzes, have an extensive range of marine applications [13]. Since the turn of the century, the corrosion of copper and its alloys has
been widely studied in chloride media where it has been observed that the chloride

*
Corresponding authors. Tel.: +61-7-3365-3668; fax: +61-7-3365-4799 (G. Kear), tel.: +44-1225-383488; fax: +44-1225-826-894 (F.C. Walsh).
E-mail addresses: garethkear@uq.edu.au (G. Kear), f.c.walsh@bath.ac.uk (F.C. Walsh).

0010-938X/$ - see front matter  2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0010-938X(02)00257-3

110

G. Kear et al. / Corrosion Science 46 (2004) 109135

ion has a strong inuence on the copper corrosion mechanism. The main focus of
this review is the literature dealing with the electrochemical characteristics of the
copper metal/chloride corrosion system. An enormous quantity of research is
available on this subject and although most authors provide a necessarily limited
background to their own work, a comprehensive review has not been made available.
The discussion here, however, has been strictly limited to pure copper. The main
aspects of the work considered here include:
1. Measurement and characterisation of the anodic and cathodic reactions.
2. The inuence of solid surface species (produced via pre-treatment or in situ) on
reaction rates and mechanisms.
3. The derivation and replication of corrosion potential characteristics.
The inuence of mass transport conditions on the rate of the copper electrodissolution reaction is of particular interest to corrosion scientists and applied
electrochemists. This is particularly so for chloride ion concentrations roughly
equivalent to that of seawater ([Cl ] @ 0.50.6 mol dm3 or 3.03.6% w/v NaCl). The
anodic reaction has generally been taken to be reversible, mainly due to the rapid,
highly thermodynamically favourable complexation of the cuprous ion by the
chloride ion. The cathodic reaction, however, is dominated by oxygen reduction,
which is usually assumed to be relatively irreversible (hydrogen evolution usually
only becomes signicant at potentials more negative than )1.0 V vs. SCE). Thus,
with particular attention to the anodic reaction, the corrosion system has been
analysed with a number of ow geometries including the rotating disc and cylinder
electrodes, pipe and channel loops and the impinging jet. The inuence of both
natural and articial lm formation on the anodic and cathodic reactions is also
important as copper and copper alloys are generally only viable as corrosion resistant materials when in the passive state. At such lmed or blocked surfaces
transport processes occurring through poorly conducting solid corrosion product
layers limit both the overall anodic and cathodic reaction rates.

2. General corrosion
Early investigations of Bengough et al., in 1920 [4] and Bengough and May, in
1924 [5] studied the corrosion of copper in seawater as part of a series of reports to
the Institute of Metals. They determined that mature, protective scales formed at
the surface of copper consist largely, but not entirely, of corrosion products. They
assumed that in neutral chloride solutions, the main, initial corrosion product of
copper is cuprous chloride, CuCl formed via reaction (1).
Cu Cl ! CuCl

It was proposed that the cuprous chloride, which is only slightly soluble in dilute sodium chloride, reacts to produce cuprous oxide (cuprite) which was the
main constituent of thick scales [5]. The cuprous oxide was generally oxidised over

G. Kear et al. / Corrosion Science 46 (2004) 109135

111

Chloride solution
Atacamite (Cu 2(OH)3 Cl) or malachite (CuCO3 .Cu(OH)2 )
Cupric hydroxide Cu(OH)2 or oxide (CuO)
Cuprous oxide (Cu 2O)
Cuprous chloride (CuCl)
Copper metal

Fig. 1. General stratication scheme of species in a mature copper corrosion product lm in seawater,
after Bengough et al. [4].

time to cupric hydroxide (Cu(OH)2 ), atacamite (Cu2 (OH)3 Cl) or malachite


(CuCO3 Cu(OH)2 ) in the presence of seawater. It was predicted that if the chemical
nature of scale formation is considered only as a function of oxygen supply, a general
stratication of corrosion products for a mature lm, as shown in Fig. 1, should be
observed.
Later, Lee and Noble [6] found that cuprous chloride complexes CuCl2
and
3

CuCl3
will
be
produced
sequentially
from
CuCl
(reaction
(2))
as
chloride
ion
4
2
concentrations become successively greater than 1.0 mol dm3 (or >0.7 mol dm3
after Braun and Nobe [7]).

2

3
Cu 2Cl () CuCl
2 Cl () CuCl3 Cl () CuCl4

K1

CuCl
2

 2

Cu Cl 

105 mol2 dm6

2
3

Bianchi et al. [8] incorporated K1 (shown in Eq. (3) for 35& seawater at 25 C), to
place CuCl
2 as the main cuprous chloride complex in seawater and NaCl electrolytes
with [Cl ] concentrations approximating 0.55 mol dm3 . This is in agreement with
most authors [623].
The method of cuprous oxide production in the presence of the chloride ion is
usually taken as a precipitation reaction [8,17] rather than a direct electrochemical or
chemical formation from the base metal or cuprous chloride. The equilibrium in
reaction (4) is shifted to the right as the local concentration of, for example, the
CuCl
2 complex (produced directly from the dissolution of copper metal or CuCl)
increases and cuprous oxide is deposited in response.


2CuCl
2 2OH () Cu2 O H2 O 4Cl

112

G. Kear et al. / Corrosion Science 46 (2004) 109135

The equilibrium constant, K2 , for this reaction at 25 C is as follows,


K2

Cl 

2
 2
CuCl
2  OH 

1020

The stability of Cu2 O is inversely dependent on the concentration of chloride ions [8].
Thus, from the general literature dealing with copper corrosion, the rate of redissolution of the protective Cu2 O (as a soluble cuprous chloride complex) has been
shown to be much higher than that observed in chloride free, neutral/alkaline solution [5,8,24,25]. In a paper devoted to the thermodynamic stability of copper in
35& salinity seawater, Bianchi and Longhi [16], produced a number of equilibrium
potentialpH (Pourbaix) diagrams for the Cu/H2 O/Cl system which included the
inuence of carbonate and bicarbonate ions. An example has been reproduced in
Fig. 2, where the solid phases taken into consideration are Cu, Cu2 O, CuCl and
Cu2 (OH)3 Cl.
The majority of work in the literature is initiated either at polished surfaces or
pre-reduced surfaces. The composition and thickness of thermal and air formed
corrosion product lms, which form by oxidation of the metal during preparation/
storage should, therefore, be carefully considered. As will be seen in the later section
dealing with oxygen reduction, the presence of such lms can have considerable
inuence on reaction mechanisms. Preston and Bircumshaw [26] observed that, at
room temperature, the oxide lm on freshly polished copper consists only of a 2.5
nm thick cuprous oxide. No denite indication of the presence of cupric oxide, CuO,
was found. Other workers have discovered, however, that when a copper surface is
heated in air, oxygen or water, a duplex lm forms where the primary component is
cuprous oxide with a lesser concentration of cupric oxide and/or cupric hydroxide,
Cu(OH)2 [25,2734].
1.2

CuO 22-

O2 /OH

CuO

Cu2(OH)3Cl

E (V/SHE)

0.8

2+

Cu

0.4

HCuO2-

CuCl
0.0

H /H2

Cu 2O

Cu
-0.4

10

12

14

pH
Fig. 2. EpH diagram for copper in seawater of salinity 35& at 25 C (assuming an activity of 104
cuprous and cupric ions in solution), after Bianchi and Longhi [16].

G. Kear et al. / Corrosion Science 46 (2004) 109135

113

The choice of electrolyte has also been shown to have a critical inuence on the
electrochemistry of the corrosion system [20,35]. Chernov et al. [35] compared the
corrosion behaviour of copper in 3% w/v NaCl solution and natural seawater. They
revealed that, owing to the buering properties of seawater, the rate of formation of
cuprous oxide in seawater solution is lower, and the rate of generation of cuprous
complexes (and thus the general dissolution rate) is higher than in the NaCl solution.
From reaction (4), an increasing pH resulting from the oxygen reduction reaction
will favour the consumption of cuprous chloride complex to form cuprous oxide.
From Fig. 2, the greater thermodynamic stability of cuprous oxide over cuprous
chloride at neutral and alkaline pH values is clear. The buering properties of seawater (due to the presence of carbonate/bicarbonate and borate ions [16,36]), however, will act to limit any change in pH of the electrolyte adjacent to the active
surface.

3. Cuprous oxide as a p-type semiconductor


Pryor, with both North in 1970 [37] and Blundy in 1972 [38] connected the effectiveness of protective lms formed on copper and copper nickels with the semiconducting nature of cuprous oxide. The linear nature of the resistance response
across dierent copper alloy corrosion product lms, indicated various magnitudes
of ohmic dependence on potential dierence. It was inferred that the long-term
corrosion rates of copper-based alloys were dependent on the combined electronic
and ionic resistances of the corrosion product lms. After some modication, this
approach has been adopted by a number of authors to explain observed reduction of
corrosion rate with exposure time and lm thickness [8,39,40].
The crystal structure of Cu2 O is cubic with the Cu and O2 ions contained in two
interpenetrating cubic lattices. It is a non-stoichiometric, highly defective p-type (or
positive carrier type) semiconductor where there is a deciency in Cu ions
[24,34,37,38,4143]. The resulting positive holes allow electron transport (contributing towards the electronic conductivity) outwards through the structure while the
cation vacancies move inwards. Cuprous ion diusion (ionic conductivity) is driven
by a cation hole gradient where a larger number of vacancies exist at the electrolyte/
oxide interface than at the metal/oxide interface [44].
In a model proposed by Bianchi et al. [8], oxygen reduction on cuprous oxide
lmed copper occurs at both the oxide/electrolyte interface and at the metal/electrolyte interface at the base of pores. With the former, oxygen is reduced at the
surface of the cuprous oxide to O2 which is then incorporated into the oxide lattice.
The electrons required for the reduction process are supplied by adjacent cuprous
ions that are consequently oxidised to cupric ions. The excess of electrons at the
metal/oxide interface, the result of prior cuprous ion migration via the positive hole
mechanism, travel outwards through the lattice (the equivalent of movement of
positive charge inwards) in order to reduce the surface cupric ions back to cuprous
ions. Eiselstein et al. [42] described the solid state reaction scheme via reactions (6)
and (7).

114

G. Kear et al. / Corrosion Science 46 (2004) 109135

Cathode
Anode

Oads 2CuCu2 O () 2VCu2 O 2pCu2 O OCu2 O


2CuM 2VCu2 O 2pCu2 O () 2VM 2CuCu2 O

6
7

The nomenclature was given as:


Oads
CuCu2 O
VCu2 O
pCu2 O
OCu2 O
CuM
VM

oxygen adsorbed on the Cu2 O surface


cation in Cu2 O
cation vacancy in Cu2 O
electron hole in Cu2 O
anion in Cu2 O
Cu in base metal
Cu vacancy in base metal

Accepting a dissolution mechanism at potentials where the Cu(II) is the stable


oxidation state, the Cu2 O in the corrosion product lm will be oxidised and dissolution as Cu2 and OH will occur,
Cathode
Anode

OCu2 O H2 O () 2OH  2pCu2 O


2Cu2 O 2pCu2 O () 2Cu2
sol: 2VCu2 O

8
9

If a pore free, oxide lm is assumed, both ionic and electronic transport through the
corrosion product will be essential for the corrosion process to proceed. In terms of
doping [41], the introduction of lower-valent cations into a p-type semiconductor
will increase the electron hole concentration and decrease the cation vacancy concentration. Higher-valent cations will, of course produce the opposite eect, increasing the number of cation vacancies. Ionic transport is usually rate limiting in
terms of overall conductivity since electron and hole mobilities are generally much
greater than ionic mobility. Thus, the introduction of cations of increasing valency
such as Ni2 and Fe2 will act to proportionally increase the overall conductivity of
cuprous oxide.
4. Electro-dissolution
As previously noted, the anodic polarisation behaviour of pure copper in chloride
media has received considerable attention in the literature. The main dierence between models is the identity of the initial, electro-dissolution reaction/s of bare
copper. The reactions are generally thought to be reversible and all are universally
assumed to be under mixed (charge transfer and mass transport controlled) kinetics
close to the corrosion potential. The three possible cases are given in reactions (10)
(14) where ki is the rate constant for the reaction of interest.
Case 1
k1


Cu 2Cl () CuCl
2 1e
k1

10

G. Kear et al. / Corrosion Science 46 (2004) 109135

115

Case 2
k2

Cu () Cu 1e

11

k2

k3

Cu 2Cl () CuCl
2

12

k3

Case 3
k4

Cu Cl () CuCl 1e

13

k4

k5

CuCl Cl () CuCl


2

14

k5

Cases 1 and 3 represent the direct formation of a cuprous chloride species from the
metal, while Case 2 involves the dissolution of copper as cuprous ion in the rst
instance. Table 1 gives a summary of authors who have worked in this eld and their
proposed mechanisms.
Anodic polarisation of copper at large overpotentials in chloride electrolytes results in E vs. log i curves typied by the schematic shown in Fig. 3 [6,7,17]. With
reference to this gure, typical anodic polarisation curves can be split into three main
regions of potential:
Section I: a potential region of apparent Tafel behaviour where mixed charge
transfer and mass transport controlling kinetics are usually assumed.
Section II: a potential window of lm formation leading to a maximum peak current density and subsequent lm or metal dissolution giving a limiting current
density.
Section III: a potential above which any increase in current density is due to the
formation of Cu(II) species.

Table 1
Review of authors showing choice of reaction scheme for the initial electro-dissolution reaction of unalloyed copper in chloride media
Case 1

Case 2

Case 3

Bacarella and Griess [45]


Faita et al. [17]
Brossard [10]
Tribollet and Newman [46]
Bjorndahl and Nobe [47]
Dhar et al. [20]
de Sanchez and Schirin [12]
Wood et al. [14]
Georgiadou and Alkire [48]
Mathiyarasu et al. [19]

Taylor [49]
Bianchi et al. [8]
Lush and Carr [50]
Braun and Nobe [7]
de Sanchez and Schirin [51]
Smyrl [52]
Wood and Fry [14]
Mansfeld et al. [15]
Wagner et al. [53]
Kear et al. [54]

Flatt and Brook [55]


Bongilio et al. [9]
Walton and Brook [23]
Lee et al. [6,22]
Deslouis et al. [56]
King et al. [57]

116

G. Kear et al. / Corrosion Science 46 (2004) 109135

0.2

E (V/SCE)

0.1
Section III.

0.0

-0.1

Section II.

-0.2

-0.3

Section I.

0.5

1.0

1.5

-2

log (i / mA cm )
Fig. 3. Typical anodic polarisation characteristics of copper in aqueous chloride media.

In the remainder of this discussion on electro-dissolution, each of the above


sections will be examined in turn.
An early galvanostatic investigation into the anodic polarisation behaviour of
copper in neutral and alkaline sodium chloride solutions in the presence and in the
absence of oxygen was published by Lal and Thirsk in 1953 [58]. They were the rst
to demonstrate that the rate of anodic dissolution of copper in neutral chloride
solution close to the corrosion potential was under the inuence of both charge
transfer and mass transport. As shown in Fig. 4, a linear Tafel region (g 100 to
+200 mV) was observed where the origin (Ecorr ) of the measured E vs. i response
became more negative with increasing chloride concentration and temperature.
From 17 to 60 C, the apparent Tafel slope (E vs. log i) was equal to 2:3RT =F
(0.059 V decade1 at 25 C). R, T and F are the molar gas constant (J K1 mol1 ), the
absolute temperature (K) and the Faraday constant (A s mol1 ), respectively. The
linear relationship between overpotential and applied logarithmic current density
obeyed the relationship shown in Eq. (15) where g is the applied overpotential, i0 is
the exchange current density and z, the number of electrons exchanged (in this case,
equal to unity).


 
RT
i
g
 log
15
zF
i0
Table 3 provides a review of directly measured anodic Tafel slopes for pure
copper in aerated 3.03.5% w/w NaCl and related electrolytes.
Since the early work of Lal and Thirsk [58], most models have tried to reproduce
at least one of the following, universally observed characteristics of the copper/
chloride electro-dissolution system within this apparent Tafel region [10]:

G. Kear et al. / Corrosion Science 46 (2004) 109135

117

0.2

E (V/SCE)

0.1
0.0
-0.1

(5)
(4)
(3)
(2)
(1)
(6)

-0.2
-0.3
-0.4
5

-1

-2

log (i / A cm )
Fig. 4. Plots of E vs. log i for copper at aqueous NaCl concentrations of (1) 4.0, (2) 2.0, (3) 1.0, (4) 0.5, (5)
0.1 and (6) 4.0 mol dm3 , after Lal and Thirsk [58]. Curves (1)(5) were performed at 18 C and curve (6) at
60 C.

(a) The relation E vs. log i is linear with a slope close to 60 mV decade1 at a xed
rate of mass transport (uid velocity) and chloride concentration:


dE
59 mV decade1
16
d logi Cl ;298 K
(b) The i vs. x0:5 relation is linear with an origin close to zero at a xed applied
potential and chloride concentration:


di
Constant
17
dx0:5 Cl ;E
(c) The relation E vs. logi=x0:5 is linear with a slope equal or close to 60 mV
decade1 :
2
3
dE

7
6
59 mV decade1
18
4
5
di
d log
dx0:5

Cl ;298 K
The examples given in (b) and (c) above are in terms of angular velocity, x (rad s1 ),
of the rotating disc electrode (RDE) [6,11]. The extensive range of mathematical
descriptions of the complete electro-dissolution behaviour of copper in the apparent
Tafel region have been summarised in Table 2. The symbols D, dN , m, c and iL refer to
diusion coecient (m2 s1 ), Nernst diusion layer thickness (m), kinematic viscosity
(m2 s1 ), concentration (mol m3 ) and limiting current density (A m2 ), respectively.
aA and aC are the charge transfer coecients for the anodic and cathodic reactions.

118

G. Kear et al. / Corrosion Science 46 (2004) 109135

Table 2
Review of expressions derived to reproduce the electro-dissolution behaviour of copper in chloride media
at low positive overpotentials
Author/s

Notes

Equation describing electro-dissolution in apparent Tafel region

Bacarella and
Griess [45]

Case 1,
reversible
reaction

iF

Faita et al. [17]

Case 1,
reversible
reaction

 


DCuCl2
FE
Cl 2 exp
RT
dN



F Ee  Ee0
zF x0:5 0:667
DCuCl Cl 3 exp
0:667
2
RT
1:61m

Moreau [11]

Case 3,
quasi-reversible reaction



1
aC FE
1

exp
;
k2
RT
BU n
iL
where B
expFE=RT
!
 ,


k5 k4
FE
dN
 2
1
Cl  exp
i
k5
RT
k4
DCuCl2

de Sanchez and
Schirin [12,51]

Case 2,
reversible
reaction






1:61D0:67
m0:167 x0:5
aA F
aC F
Cu
g

k
g

exp
i zFcv0
k2 exp
2
v0
Cu
RT
RT
cCu

Brossard [10]

Case 1,
reversible
reaction

i 0:62zF m0:167 DCuCl2 x0:50 Cl 2 exp

F E  E0
RT

Case 1,
reversible
reaction


,
a1 FE
1
i Fk1 Cl  exp
RT

dN

Case 1,
reversible
reaction

i Fk1 exp

Case 2,
reversible
reaction

i Fk2 exp

Lush and Carr [50]

Tribollet and
Newman [46]
Dhar et al. [20]

Smyrl [52]

Lee and Nobe [6]

Deslouis et al. [56]

Case 2,
reversible
reaction

Case 3,
quasi-reversible reaction
Case 3,
reversible
reaction

FE
RT

i exp



 2

DCuCl2

k1 exp



a1 FE
RT

"
#

,


a1 F
dN
a1 F
g
1 k1
g
exp
RT
RT
DCuCl K2 Cl 2
2


,
a1 FE
1
RT

k2 dN

 2

k2 Cl  DCuCl2


exp

a1 FE
RT

 


k5 k4
FE
Cl  exp
2
RT
k4
DCuCl2 0:667 m0:167 x0:5 1:62k5

0:167 0:5
FD0:667
x
CuCl m


9
8
FE >
>
>k4 exp


 >
<
aA FE
RT =

i F k4 Cl  exp
2

>
>
RT
>
>
;
: k4 k5 Cl 


9
8
FE >
>
>
>
k
k
d
exp
< 4 5 N
RT =
>
>
:

k4 k5 Cl 2 DCuCl2

>
>
;

Case identities relate to Eqs. (10)(14). See the text for symbol meanings.

G. Kear et al. / Corrosion Science 46 (2004) 109135

119

Although systematic reproduction of the derivations of each relationship is beyond the scope of this work, the main assumptions of each study can be summarised.
From Table 2, it is clear that the choice of anodic reaction (see case number) is
variable but it is usually taken to be reversible. With the exception of de Sanchez and
Schirin, who assume the direct dissolution to the free cupric ion (1982 and 1988),
the electro-dissolution current is always dependent on chloride ion concentration. If
only pure charge transfer controlled current density is considered (see again later in
this section), the rate dependency of the anodic reaction on the chloride ion can be
described via a classical kinetic equation:
m

i k zF Cl 

19

where k is the potential dependent rate constant for the anodic reaction and m is the
reaction order with respect to the concentration of the chloride ion. Venkakachai
and Kannan postulated a value of 1.0 [59] for m while Faita et al. [17] have produced
a reaction order with respect to the chloride ion equal to 3. The most frequently
quoted value of m at NaCl concentrations less than 1.0 mol dm3 (7% w/v NaCl),
however, is 2.0 [6,11,18,45,46,52,56].
In all of the cases shown in Table 2, the anodic current is clearly dependent on
both charge transfer and mass transport. The diusion of CuCl
2 away from the
electrode surface is usually taken to control the mass transport controlled component of the current and H and OH ions are not directly involved in the dissolution
process. Consequently, the pH of the solution should exert no inuence over the rate
of electro-dissolution within the apparent Tafel region. This has been veried experimentally by Brossard [10]. It has also been noted that within the region of apparent Tafel behaviour, the polarisation behaviour of copper in acid chloride
solution is very similar to that in neutral chloride solution [6,7,9].
Dhar et al. [20] assumed that copper corrodes directly via the anodic dissolution
mechanism in Case 1, reaction (10), which was taken to be reversible. At high anodic
currents, when the overall rate of the anodic reaction is much faster than the diffusion rate of CuCl
2 (when mass transport will inuence the measured current), the
following expression was derived to describe the apparent Tafel behaviour:
dg
2:303RT

d logi a1 a1 F

20

where a1 and a1 are the charge transfer coecients for the forward and reverse
directions of reaction (10) in Case 1. Eq. (20) at 25 C predicts the 60 mV decade1 E
vs. log i response within the apparent Tafel region. Similar relationships have also
been produced by de Sanchez and Schirin [51], Eqs. (21) and (22).


d log di1 =dkm1
zF
1=59 mV decade1

2:3RT
dE

21

d logdi1 =dx0:5
zF

1=59 mV decade1
dE
2:3RT

22

120

G. Kear et al. / Corrosion Science 46 (2004) 109135

Slopes of g vs. log i for copper [13] and copper alloy [51] dissolution in seawater was
found to be close to 1/60 decade mV1 (or 60 mV decade1 ). Note that the value of
z 1 and the charge transfer coecient has been omitted. For lmed surfaces, Wood
et al. [13] found that the directly measured anodic Tafel slope for copper changed to
values corresponding to an activated electron transfer rate determining step. That is,
where the transfer coecient, aA , is approximately equal to 0.5 (rather than zero for
a freshly polished copper surface) which produced a Tafel slope of 110 mV decade1 .
For this case, the kinetics of charge transfer had decreased to a level where mass
transport no longer inuenced the rate of electro-dissolution and, therefore, the
value of the Tafel slope. Lush and Carr [50] noted that an increase in the apparent
anodic Tafel slope from approximately 60 mV decade1 (at low ow velocity and low
overpotential) to 90 mV decade1 (high ow velocity and high overpotential) was
produced by a change in the balance between charge transfer and mass transport
control. This eect was said to be not due to a change in the mechanism of the
copper dissolution reaction itself. It was proposed that at high ow velocity, the
diusion of ions can be enhanced to such an extent that the charge transfer controlled component of the anodic reaction scheme becomes signicant and inuences
the measured current response. It appears that at high rates of mass transport the
charge transfer coecient will become signicant and inuence the apparent Tafel
slope value. That is, the observed change in the anodic Tafel slope from 60 towards
120 mV decade1 corresponded to a change in the value of the charge transfer coecient (for a single electron transfer) from 0 to 0.5.
Some evidence for the above has been noted in the literature for pure charge
transfer controlled Tafel slopes measured for copper in chloride media. The extraction of pure charge transfer controlled current from mixed controlled current
may be extracted using the KouteckyLevich equation approach [13,54,60,61]. For
cases where aA 0:5 has been assumed, Tribollet and Newman [46] and Smyrl [52]
have quoted pure charge transfer controlled Tafel slopes of approximately 0.120
V decade1 . Kear et al. [54] measured slopes of 0.062 and 0.097 V decade1 in ltered
and articial seawaters, respectively.
Section II of the anodic polarisation response will now be considered, where a
region of peak current density and limiting current density is usually observed.
Returning to Fig. 4, Lal and Thirsk [58] assumed that for neutral solutions, a surface
adsorbed precipitate of CuCl was formed at critical current densities, the values of
which were dependent on chloride concentration and temperature (later conrmed
by Venkatachari and Kannan [59], and Wagner et al. [53]). Once the deposition of
this precipitate reached a critical rate, the current attained a xed value i.e., a region
of limiting current density. Nobe and coworkers [6,7] produced a model explaining
the entire anodic polarisation response of copper in acid chloride solution. Again, it
is apparent that the model can equally be applied to neutral and alkaline chloride
media where Cu2 O precipitation may be favoured over the deposition of CuCl [17].
The peak maximum, minimum and limiting current density responses observed in
Section II are produced as the formation of CuCl becomes faster than either its
complexion by the chloride ion or mass transport of the cuprous dichloride complex
to the bulk solution. The apex peak current, therefore, is followed by a current

G. Kear et al. / Corrosion Science 46 (2004) 109135

121

minimum as surface CuCl coverage reaches its maximum [22,23,55]. Further reaction
of the copper with chloride ions to form CuCl is reduced to a rate which Braun and
Nobe proposed as equal to the rate of Cl diusion from the bulk of the electrolyte
to the electrode surface [7]. The peak current maximum and the current minimum
were not observed for [Cl ] P 2.0 mol dm3 due to the high dissolution rate of the
cuprous chloride where a lm free surface can be assumed.
As predicted by the Levich equation [62] which describes current at a rotating disc
electrode under full mass transport control, Braun and Nobe [7] found that the
magnitude of the limiting current density was dependent on the angular velocity of
the disc according to x0:5 . Using the rotating ring disc electrode [6] it was found that
the formation of CuCl
2 was constant with increasing potential for a given chloride
ion concentration. This indicated that the maximum rate of CuCl dissolution had
been reached which was independent of potential. The limiting current density was
given by:
iL

0:167 0:5
0:62zFD0:667
x Cl 
Cl m
1  tCl

23

where tCl is the transference number of the chloride ion. This is in agreement with
the general case of passivating lm formation in conjunction with lm dissolution
[67]. It is probable, however, that at these potentials, the limiting current density is
controlled by the diusion of the cuprous complex away from the surface of the
cuprous chloride to the bulk of the electrolyte [68]. For this model, Eq. (24) will hold.
iL

0:167 0:5
0:62zFD0:667
x CuCl
CuCl m
2
2

1  tCuCl2

24

Data produced by Kear [68], which considered the anodic polarisation of copper,
strongly indicate that limiting current densities, taken from plots such as those
shown in Fig. 5, are the result of cuprous dichloride ion mass transport. Literature

values for the diusion coecients of CuCl
2 and Cl ions are shown in Table 4,
where it is apparent that low diusion coecient values and relatively low concentrations of the cuprous complex compared to the chloride ion will tend to give rise to
a mass transport limiting step in the lm dissolution mechanism.
At potentials more positive than the limiting current region, P0.05 V vs. SCE
(Section III, Fig. 3), both the cuprous chloride complex and free cupric ions are
produced from surface cuprous chloride via competing reactions [6]. In this region, a
combination of the diusion of both Cu(I) and Cu(II) species from the electrode/
electrolyte interface into the bulk solution were said to be rate determining.
5. Cathodic polarisation
A simplied relationship for the complete reduction of oxygen, reaction (25),
involves an overall exchange of four electrons resulting in the production of hydroxyl
ions or, at low pH, water molecules.
O2 2H2 O 4e ! 4OH

25

Flow geometry

Temperature (C)

Surface condition

bA (V decade1 )

Reference

NaCl

RDE

28 1

Freshly polished

[59]

NaCl
NaCl

Static
Static

Freshly polished
Freshly polished

NaCl
NaCl

Static
Pipe ow
0.16 m s1 y
1.60 m s1 z
RDE

18
30100y
>100z
30175
20

25

Freshly polished

0.045 (0 rpm)
0.040 (500 rpm)
0.035 (1000 rpm)
0.040 (2500 rpm)
0.070 (5000 rpm)
0.055 0.005
@ 0.060y
0.1400.160z
@ 0.060
0.060y low g
0.070y high g
0.090z high g
0.064| (500 rpm)
0.053| (1000 rpm)
0.054| (1600 rpm)
0.067| (2500 rpm)
All slopes 0.005
@ 0.060 (03000 rpm)
Low g high g
0.063 0.063y
0.060 0.120z (8 m s1 )
0.061 0.130z (14.7 m s1 )
0.120 50y at low g
0.115 58z (2 m s1 )
0.038 0.047 (0 285 rpm)
0.062 0.065 (0 285 rpm)
@ 0.060 (slope independent of
exposure time)
0.067
0.087y and 0.098z
0.063 0.001y and
0.066 0.002z

NaCl| and
articial seawaterr

Freshly polished
Freshly polished

NaCl
Seawater

RDE
Staticy
Enclosed
channel owz

2025
25

Freshly polished
Freshly polished
Filmed

Seawater

RCE

Natural and
articial seawaters
Seawater
Seawater
Naturaly and
articialz seawaters

RCE

Filmed (7 days)
Filmed (21 days)
Freshly polished

Static

20
30y and 50z
25 0.2

Freshly polished
Freshly polished
Freshly polished

RDE

[58]
[63]
[45]
[50]

0.061r (500 rpm)


0.060r (1000 rpm)
0.060r (1600 rpm)
0.058r (2500 rpm)
All slopes 0.005

[20]

[56]
[13]

[64]
[15]
[65]
[66]
[54]

G. Kear et al. / Corrosion Science 46 (2004) 109135

Electrolyte

122

Table 3
A review of directly measured anodic apparent Tafel slopes for pure copper in fully aerated neutral 3.03.5% w/v NaCl and natural and articial seawaters

G. Kear et al. / Corrosion Science 46 (2004) 109135

123

106.2
6500 rpm

3500 to 6500 rpm

i (mA cm-2)

200 to 1800 rpm

70.8

35.4
200 rpm

0.0
-0.4

Cu - 1.0 mV s-1

-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

E (V/SCE)
Fig. 5. Anodic polarisation curves for the electro-dissolution of freshly polished copper in ltered seawater as a function of rotating disc electrode rotation rate, after Kear [68].
Table 4
Diusion coecients for the cuprous dichloride complex ion and the chloride ion in chloride media
DCuCl2 (cm2 s1 )
5

1.0 10
5.4 106
5.0 106
5.5 106

DCl (cm2 s1 )

2.0  105

Electrolyte conditions
3

0.53.0 mol dm HCl at 25 C


0.12.0 mol dm3 HCl at 25 C
Seawater at 25 C
0.11.0 mol dm3 NaCl at 23 1 C
Seawater at 25 C

Reference
[11]
[46]
[13]
[18]
[36]

The generalised scheme describing the possible reduction mechanisms involving an


intermediate peroxide is shown in reaction (26) [61,69,70] where, in alkaline media
the peroxide species may be HO
2.

26

124

G. Kear et al. / Corrosion Science 46 (2004) 109135

The indices indicate that:


(a)
(b)


species is adsorbed at the electrode surface,


species is considered to be within the bulk of the electrolyte, and
species is within the immediate vicinity of the electrode surface.

It is clear that the complete, four-electron reduction of oxygen may occur through
a direct or indirect route. Hydroxyl ions or water molecules can be products of a
single four-electron step or the result of cumulative two-electron reduction steps
where oxygen is reduced to peroxide which in turn is reduced to hydroxyl ions.
Hence, if k6 , k10 , k9 are small and k8 , k9 large, oxygen reduction may only involve an
overall two-electron change. Clearly, this explanation is simplied. The kinetics of
the reversible electrochemical and chemical reactions along with the rates of the
adsorption/desorption processes may be equivalent to the active reduction processes.
This will result in a complication of the overall reduction mechanism. Because of
these considerations, the kinetics of oxygen reduction are expected to be very specic
to the system under study. This is found to be the case in practice [69]. The character
of the substrate, surface condition, temperature and electrolyte conditions all have
an inuence over each step in the reduction mechanism.
The earliest study dealing with the reduction of dissolved oxygen at copper in
chloride media was performed by Delahay in 1950 [71]. Over the whole range of
negative overpotentials studied, it was determined from polarisation curves and
oxygen consumption data that the number of electrons consumed was predominantly four. Although hydrogen peroxide was always formed, it was proposed that a
form of catalytic decomposition prevented the build up of this intermediate reduction product. Balakrishnan and Venkatesan [72] incorporated the rotating ring disc
electrode (RRDE) in a full kinetic study of oxygen reduction at copper in sodium
chloride solution. The reaction was studied for both freshly polished and pre-reduced surfaces. Cathodic polarisation from Ecorr to )1.4 V vs. SCE produced a single
mass transport limited wave which was taken to indicate the absence of large
quantities of peroxide. These plots are similar to those measured by Kear et al. [54]
as a function of rotating disc electrode rotation rate (see Fig. 6). Here, (a) denotes
oxygen reduction under charge transfer/kinetic control, (b) oxygen reduction under
mixed mass and charge transfer control, (c) oxygen reduction under mass-transfer
control and (d) hydrogen evolution. Reverse scans measured by Balakrishnan
and Venkatesan, however, showed two distinct waves or peaks. The second peak
at @ )0.4 V only became noticeable only on a partially or completely oxide free
surface and was identied with the reduction of oxygen to hydrogen peroxide. This
has also been veried by Venkatachari and Kannan [59].
Deslouis et al. [56] found that the morphological characteristics of cathodic polarisation curves measured at copper are strongly dependent on the time of immersion prior to polarisation. Indeed, it is clear from the literature that at potential
sweep rates P1 mV s1 non-steady-state conditions can result at surfaces supporting
corrosion products. After an equilibration time of 5 min at 0 rpm, Deslouis et al.
found that LSV curves produced a peak superimposed upon the single wave for

G. Kear et al. / Corrosion Science 46 (2004) 109135

125

0.0
(a)

i (mA cm-2)

-0.7

200 rpm

-1.3
(b)

-2.0
-2.7
(c)

-3.3
-4.0
-1.5

9500 rpm
(d)

-1.2

-0.9

-0.6

-0.3

0.0

E (V/SCE)
Fig. 6. Cathodic linear sweep voltammetry at the copper RDE as a function of angular velocity in aerated
ltered seawater at 25 0.2 C and a scan rate of 0.5 mV s1 , after Kear et al. [54].

oxygen reduction at a somewhat irreproducible potential of )0.390 0.07 V vs. SCE.


This peak was attributed to the reduction of CuCl formed during the 5 min electrode
equilibrium period. Equilibration of the electrode for 16 h at 0 rpm prior to polarisation, resulted in the disappearance of the initial peak which was replaced by a
second single peak at )0.9 V vs. SCE superimposed centrally upon the oxygen reduction limiting current plateau. The second peak was attributed to the reduction
of Cu2 O formed by hydrolysis of the initial lm of cuprous chloride over the 16 h
immersion period.
The same eect has also been noted by Bjorndahl and Nobe [47]. Under stirred
conditions, a separation of the more negative peak occurred; this was attributed to
the additional reduction of cuprous or cupric species present in solution. Halliday
[73] and Shams El Din and Abd El Wahab [34] reported the same peak separation
but attributed the peaks to the reduction of solid surface lm layers. It was proposed
that, during the cathodic polarisation of copper (which supported a duplex corrosion
product layering) an underlying layer of cuprous species will be reduced rst followed by the immediate reduction of overlying cupric species at a larger than expected overpotential. The articial formation of anodic lms and their subsequent
reduction has also been studied with the use of cyclic voltammetry as a function of
potential sweep rate [19,47,59].
In a 1983 review of oxygen reduction, Schirin [69] described the complications
arising in the reduction mechanism where anodic reaction products of copper are
present. Cuprous ions are able to react with hydrogen peroxide according to reactions (27) and (28).
Cu H2 O2 ! Cu2 OH OH

27

Cu OH ! Cu2 OH

28

126

G. Kear et al. / Corrosion Science 46 (2004) 109135

Thus, the features of the reduction of oxygen at copper electrodes should be inuenced by a combination of the copper corrosion reaction, the Cu /H2 O2 reaction
and the normal kinetics of reduction at a bare copper surface. A study of the reduction of hydrogen peroxide at pre-reduced, polished and pre-oxidised copper in
the absence of oxygen by Vazquez et al. [24] conrmed that peroxide can be electrochemically reduced on articially oxidised surfaces containing Cu(I) species. In
the absence of chloride, the reduction process was catalysed by a redox cycle involving the Cu(I)/Cu(II) couple:
Cu2 O H2 O2 ! 2CuO H2 O 2e

29

2CuO H2 O 2e ! Cu2 O 2OH

30

For a surface pre-oxidised at potentials where a Cu/Cu2 O/CuO duplex, corrosion


product lm is produced, the Cu(II) species was actually found to reduce the rate of
the hydrogen peroxide reduction process. Similarly, at a pre-reduced copper surface,
the redox catalytic mechanism was also slowed and the current vs. potential response
within the charge and mixed control region decreased. Following on from this study,
Vazquez et al. [74] considered the complete electrochemical reduction of oxygen at
pre-reduced and pre-oxidised copper electrodes. Cu, Cu/Cu2 O and Cu/Cu2 O/CuO
surface systems were investigated. The authors found that oxygen reduction is also
inhibited by the prior formation of a duplex layer of corrosion products (with CuO
present as a layer adjacent to the electrolyte). At non-duplex lm layering consisting
of only cuprous oxide, a single linear sweep voltammetric polarisation wave corresponding to a four-electron change was observed indicating the absence of signicant
intermediate peroxide production. Under conditions where cuprous oxide is stable, it
was proposed that oxygen reduction proceeds through a sequence of steps involving
the oxidation of Cu2 O:
Cu2 O O2 H2 O ! 2CuO H2 O2

31

Cu2 O H2 O2 ! 2CuO H2 O

32

Again, the cuprous oxide is regenerated electrochemically via the reduction of the
cupric oxide (reaction [30]). Oxygen reduction at corrosion product free copper was
examined using NH4 Cl as a Cu(I) complexing agent. Again, the eect was to produce
a more irreversible reduction process and to separate the linear sweep voltammetry
reduction waves for oxygen and peroxide.
King et al. [57] studied the reduction of oxygen at polished copper in neutral,
unbuered 1.0 mol dm3 NaCl using the RRDE at various oxygen concentrations.
They found that the reduction process was rst order with respect to oxygen concentration and as shown in reaction (33), following the adsorption of oxygen at the
electrode surface, the rst electron transfer is rate determining.
O2 1e ! O
2

33

The increase in interfacial pH due to the net production of hydroxyl ions in unbuered NaCl was said to favour the formation of the catalytic Cu(I) surface lm

G. Kear et al. / Corrosion Science 46 (2004) 109135

127

and the overall reaction rate through catalytic Cu(OH)ads or Cu2 O species. They
found, however, that in the presence of Cu(I) surface lms, peroxide was detected as
a stable intermediate at the RRDE. The ring current was only signicant, however,
at low negative overpotentials and at a bubbling gas composition of 100% oxygen
where hydroxide production would be signicant. From this work, King et al. [57]
proposed a reduction mechanism similar to that of Schirin and Vazquez et al. involving oxygen reduction on two types of surface site having diering degrees of
reactivity. The most reactive site towards oxygen reduction involves a Cu(I) species.
The least reactive site is found at the metal itself, Cu(0). An overall expression for the
RDE geometry was derived where the catalytic Cu(I) species was taken to be
Cu(OH)ads , Eq. (34):
"

(
0:5

Ic:t: zFAhA kCA O2  1 kCB Zx

O2  exp

 #)


a0A  aC F
aC F
g
g
 exp
RT
RT
34

where Ic:t: is the cathodic current due to charge transfer controlled oxygen reduction,
A is the surface area of the electrode, hA is the fractional surface coverage of bare
copper, a0A is the anodic transfer coecient for the formation of Cu(OH)ads . aC is the
cathodic transfer coecient and kCA and kCB are the electrochemical rate constants
for oxygen reduction at bare copper and Cu(I) lmed copper respectively. The term
Z is a constant in the Levich equation consisting of:
Z 0:64m0:167 D0:667

35

At low oxygen concentrations and/or at relatively large cathodic overpotentials,


bare copper surface sites dominate (hA ! 1) and Eq. (34) is reduced to a simple
expression describing charge transfer controlled cathodic current:


Ic:t:

aC F
g
zFAkCA O2  exp
RT


36

The majority of the literature on the electrochemistry of copper appears to indicate


that for copper the presence of solid, cuprous-based species will act to catalyse the
oxygen reduction reaction. The presence of solid cupric species or, even the complete
absence of surface species will tend to inhibit the reduction process. A review of the
published Tafel slopes for oxygen reduction at pure copper in chloride media is given
in Table 5. It can be assumed that the wide range of values shown in this table ()46
to )300 mV decade1 ) certainly has contributions from deviation in surface conditions. In addition, a review of diusion coecients for oxygen is also given in Table 6
where throughout, the values corrected to 25 C have been determined using an
approximate activation energy for diusion of 83 J mol1 [68]. In this case, deviation
is lower with the diusion coecient ranging between 1.5 105 and 2.3 105
cm2 s1 when corrected to 25 C.

128

Table 5
Review of cathodic Tafel slopes for pure copper in aerated chloride media
Temperature (C)

Surface condition

bC (V decade1 )

E vs. SCE (V)

Reference

0.5 mol dm3 NaCl

30 0.1

[72]

25

[20]

Natural seawater

25

)0.5 to )0.3
)0.5 to )0.3

[13]

Natural seawater

0.11.0 mol dm3 NaCl


0.5 mol dm3 NaCl in
0.1 mol dm3 borax
Articial seawater
Natural articial
seawaters
1.0 mol dm3 NaCl

28 1
20 1

)0.238y
)0.155z
)136| (500 rpm)
)194| (1000 rpm)
)195| (1600 rpm)
)173| (2500 rpm)
)192r (500 rpm)
)190r (1000 rpm)
)167r (1600 rpm)
)167r (2500 rpm)
All slopes 0.005
)0.150y
)0.210#
)0.035 to )0.075
)0.100 to )0.325#
(RDE: 0285 rpm)
)0.130 10
)0.30

0.5 mol dm3 NaCl|


and articial seawaterr

Freshly polishedy
and pre-reducedz
Freshly polished

[59]
[69]

)0.046
@ )0.100

[65]
[15]

)0.13 (2.0% O2 + N2 )
)0.14 (5.0% O2 + N2 )
)0.15 (10.0% O2 + N2 )
)0.17 (20.9% O2 + N2 )
)0.17 (50% O2 + N2 )
)0.16 (100% O2 )
)0.134y
)0.148z

)0.70 to )0.35
)0.75 to )0.35
)0.75 to )0.35
)0.75 to )0.35
)0.90 to )0.50
)0.80 to )0.55
@ )0.50 to )0.65y
@ )0.50 to )0.65z

[57]

Naturaly and articialz


seawaters

20

23 2

25 0.2

Freshly polishedy and


aged# 3 months
Aged seven days and
aged 21 days#
Freshly polished
Freshly polished and
pre-oxidised
Freshly polished
Freshly polished and
aged 21 days
Freshly polished and
pre-reduced

Freshly polished and


pre-reduced

[64]

[54]

G. Kear et al. / Corrosion Science 46 (2004) 109135

Electrolyte

Electrolyte

Electrode material

Temperature (C)

D  105 (cm2 s1 )

D25 C 105
(cm2 s1 )

Reference

0.68 mol dm3 NaCl


Natural seawater
0.5 mol dm3 NaCl
Natural seawater
0.5 mol dm3 NaCl
Natural seawater
0.5 mol dm3 NaCl in 0.1 mol dm3
Borax
1.0 mol dm3 NaCl
1.0 mol dm3 NaCl
0.5 mol dm3 NaCl

Monel
Aluminium brass
Carbon steel
Range of CuNi
Cu
Cu and 7030 CuNi
Cu

25
20
20
30
2025
19
20

2.3
1.7
1.7
1.9
2.0
1.4
1.9

2.3
1.9
1.9
1.8
2.0
1.6
2.1

[76]
[51]
[77]
[78]
[56]
[12]
[74]

Cu
Cu
Cu
Marinel CuNi#
Cu
9010 CuNi#
Nickel aluminium bronze

23 2
23 1
20

1.7
1.8
1.4
1.5#
2.0y 2.2z
2.2y#
1.7y , 1.8z

1.8
1.9
1.5
1.7#
2.0y 2.2z
2.2y#
1.7y , 1.8z

[57]
[18]
[79]

Naturaly and articialz seawaters

25 0.2

[68]

G. Kear et al. / Corrosion Science 46 (2004) 109135

Table 6
Literature diusion coecient values for oxygen at atmospheric pressure in chloride media

129

130

G. Kear et al. / Corrosion Science 46 (2004) 109135

6. The mixed potential


With the use of rotating disc and rotating cylinder electrodes, Venzcel et al. [75]
and subsequent workers [13,35,50,51,57] established that the corrosion rate of polished copper in NaCl solution is limited by the diusion of anodic corrosion products from the metal/electrolyte interface to the bulk solution. This was found to be
true provided that a constant concentration of dissolved oxygen was large and not
rate determining. Using a RDE Venzcel et al. [75] also noted a marked increase in
corrosion rate at the critical Reynolds number, Recrit , corresponding to transition
from laminar to turbulent ow. At very high Reynolds numbers, however, the
corrosion rate of copper was found to be virtually independent of the ow velocity
where the mass-transfer process was so rapid that the charge transfer process of the
anodic reaction was rate determining.
Faita et al. [17] showed that anodic and cathodic polarisation curves obtained via
linear sweep voltammetry as a function of rotation rate, can be used together with
corresponding values of the corrosion potential to describe the corrosion mechanism
of copper in chloride media using mixed potential theory [80]. They were able to
calculate corrosion rate values for polished copper ranging between 0.2 and 0.3
mm y1 (@ 9 lA cm2 ). Mixed charge transfer and mass transport control over the
oxygen reduction reaction was only signicant at potentials more negative than
)0.450 V vs. SCE (a potential far more negative than the open circuit potential). It
was predicted that only at very low dissolved oxygen concentrations and low ow
rates would the mixed potential exhibit a mass transport control over the cathodic
half-cell reaction. An Evans diagram, such as the schematic shown in Fig. 7, was

2> 1

Ecorr1
Ecorr2

A1
A2
C1
log icorr1 log icorr2

C2

log i

Fig. 7. Schematic polarisation curves for copper corrosion in aerated chloride media for two diering
mass-transfer conditions, x1 and x2 , after Faita et al. [17]. Anodic and cathodic polarisation curves are
represented by Ax and Cx respectively.

G. Kear et al. / Corrosion Science 46 (2004) 109135

131

-0.240

E (V/SCE)

-0.260
200 rpm

-0.280
-0.300
9500 rpm

-0.320
Filtered seawater

-0.340

400

800

1200

1600

2000

t (s)
Fig. 8. Typical mean open circuit (corrosion) potential transients for freshly polished copper in aerated,
ltered seawater as a function of RDE rotation rate, after Kear et al. [54].

used to describe the corrosion process occurring at the copper surface. At the mixed
potential, the anodic dissolution of copper is under mixed charge transfer and mass
transport control within the apparent Tafel region and the cathodic reaction (oxygen
reduction) is under full charge transfer control. The increasingly negative corrosion
potential (Ecorr1 to Ecorr2 ) and derived corrosion rate (icorr1 to icorr2 ) is solely due to the
greater anodic current within the apparent Tafel region. Evidence for this increase in
negativity of the corrosion potential with uid velocity is given in Fig. 8 for copper in
ltered seawater [54]. From this data, it is clear that immediately after immersion of
a freshly polished electrode the magnitude of change in potential is dependent on
rotating disc electrode rotation rate.
As seen in the earlier section dealing with copper electro-dissolution, the greater
currents are due to increasing rates of cuprous dichloride complex mass transport at
increasingly greater RDE angular velocities (from x1 to x2 ). With varying degrees of
acknowledgement to the work of Faita et al., this model has been adopted by a
number of other workers [18,50,51,54]. Of note, both Lush and Carr [50] and
Schirin and coworkers [13,51] predicted that the corrosion potential would become
independent of the rate of mass transport at high uid velocities and the latter authors derived expressions that are able to reproduce Ecorr [51] and icorr [13]. For a fair
estimation of the corrosion rate, de Sanchez and Schirin [51] suggested employment
of the activated Tafel oxygen reduction polarisation curve combined with measurement of the open circuit at various rotation rates. This direct method was suggested to avoid the laborious calculation of the electrochemical constants and errors
associated with changes in surface condition during anodic linear sweep voltammetry. The latter problem, however, can be eliminated with the use of a potential
step, current transient technique [54,81].

132

G. Kear et al. / Corrosion Science 46 (2004) 109135

In a 1995 study of the corrosion of copper in 1.0 mol dm3 NaCl, King et al. [18]
used the RDE as a function of oxygen concentration and mass-transfer coecient.
They demonstrated that only with a signicant decrease in the bulk oxygen concentration, did the cathodic reaction exhibit mixed charge and mass transfer control at
the corrosion potential. But, complete charge transfer control of the anodic reaction
was only predicted at very high RDE rotation frequencies, f , in excess of @ 1 kHz
(60,000 rpm). This extremely high value of rotation rate indicates the importance of
mass transport in their model. The eect of a covering layer of porous nylon membrane or compacted clay (to reduce the mass-transfer coecients of O2 and CuCl
2)
was to produce total charge transfer control of both the anodic and cathodic reactions
at the mixed corrosion potential. This was taken as a simulation of the eect porous
and micro-porous surface lms would have on the corrosion mechanism at the mixed
potential. The dependence of the steady-state corrosion potential as a function of
RDE rotation rate and oxygen concentration is shown in Table 7 [18].
From this table, it can be seen that the corrosion potential becomes less dependent
on rotation rate with decreasing oxygen concentration and consequently, the negative slope of Ecorr vs. logarithmic (electrode frequency) becomes more positive. At
very low oxygen concentrations, both the anodic and the cathodic processes are
inuenced by changes in the conditions of mass transport which will have a tendency
of cancel the rotation rate eect. Full mass transport control of the oxygen reduction
reaction was not achieved, however, even for the fully de-aerated electrolyte, due to
the irreversible nature of oxygen reduction at copper.
Several studies have also been made using linear polarisation resistance, Rp in
conjunction with the SternGeary relationship, Eq. (37) [84], where the mixed
controlled, apparent Tafel slopes for the anodic reaction are usually taken for the
term bA .
icorr

bA bC
B

2:3Rp bA bC Rp

37

A review of literature proportionality constants, B, ranging from 5 to 31 mV, is given


in Table 8 for the corrosion of copper in chloride media. Again, the deviation in
these values is probably not due only to experimental characteristics peculiar to each
worker but also to varying states of surface condition.
Table 7
Experimental data describing the dependence of copper Ecorr in NaCl solution on RDE rotation frequency
(2 s1 @ 120 rpm) and bulk oxygen concentration, after King et al. [18]
[NaCl] (mol dm3 )

[O2 ] (mol dm3 )

Ecorr vs. SCE at


f 2 s1 (V)

dEcorr
vs. SCE
d logf
(V)

dEcorr
vs.
d logO2 
SCE (V)

1.0
1.0
1.0
1.0
0.1

2.0 104 (air)


1.9 105 (2% O2 /N2 )
1.9 105 (0.2% O2 /N2 )
7 107 (de-aerated)
2.6 105 (2% O2 /N2 )

)0.306 0.008
)0.356 0.011
)0.413 0.011
)0.426 0.014
)0.246 0.007

)0.019
)0.015
)0.014
)0.011
)0.016

0.049
0.057
0.032

G. Kear et al. / Corrosion Science 46 (2004) 109135

133

Table 8
Values of the proportionality constant, B, for unalloyed copper in chloride media
Electrolyte

Value of B (V)

Reference

Natural seawater
3% NaCl
Natural seawater
33.5% NaCl and natural seawater
Naturaly and articialz seawaters
Naturaly and articialz seawaters

0.006
0.031
0.005
0.019 0.001
0.014y and 0.020z
0.018 0.001y and 0.023 0.003z

[82]
[82]
[83]
[20]
[15]
[68]

7. Conclusions
Four general conclusions can be drawn from the literature dealing with the
electrochemical corrosion of lm free copper in chloride media:
1. At potentials close to the corrosion potential the anodic reaction is under mixed
charge transfer and mass transport controlling kinetics where the mass transport
limiting step is the rate of movement of a cuprous chloride complex away from the
electrode surface to the bulk electrolyte.
2. The cathodic reaction is dominated by the comparatively irreversible reduction of
oxygen via either a two- or four-electron exchange which remains under complete
charge transfer control at potentials close to the corrosion potential.
3. At the corrosion (mixed) potential the anodic reaction is under mixed control
while oxygen reduction is under pure charge transfer control and, as such, the corrosion rate is dependent on mass transport conditions.
4. The corrosion of copper in aqueous chloride media is made complicated by the
formation of surface lms.
5. The presence of surface lms generally act to reduce the rate of both anodic and
cathodic charge transfer process and, as a result, the inuence of mass transport at
the corrosion potential reduces in importance.
Acknowledgements
The authors are grateful to Dr. Carlos Ponce de Leon for editorial assistance
during the preparation of this review.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]

A.H. Tuthill, Materials Performance 26 (1987) 1222.


E.C. Mantle, Marine Engineers Review, July (1986) 1921.
I. Ogilvie, Corrosion and Coatings South Africa 614 (13) (1986) 2, 56, 10.
G.D. Bengough, R.M. Jones, R. Pirret, Journal of the Institute of Metals 23 (1920) 65158.
G.D. Bengough, R. May, Journal of the Institute of Metals 32 (1924) 81269.
H.P. Lee, K. Nobe, Journal of the Electrochemical Society 133 (1986) 20352043.
M. Braun, K. Nobe, Journal of the Electrochemical Society 126 (1979) 16661671.
G. Bianchi, G. Fiori, P. Longhi, F. Mazza, Corrosion-NACE 34 (1978) 396406.

134
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]

G. Kear et al. / Corrosion Science 46 (2004) 109135


C.H. Bongilio, H.C. Albaya, O.A. Cobo, Corrosion Science 13 (1973) 717724.
L. Brossard, Journal of the Electrochemical Society 130 (1983) 403405.
A. Moreau, Electrochimica Acta 26 (1981) 497504.
S.R. de Sanchez, D.J. Schirin, Corrosion Science 28 (1988) 141151.
R.J.K. Wood, S.P. Hutton, D.J. Schirin, Corrosion Science 30 (1990) 11771201.
R.J.K. Wood, S.A. Fry, Journal of Fluids EngineeringTransactions 112 (1990) 218224.
F.B. Mansfeld, G. Liu, H. Xiao, C.H. Tsai, B.J. Little, Corrosion Science 36 (1994) 20632095.
G. Bianchi, P. Longhi, Corrosion Science 13 (1973) 853864.
G. Faita, G. Fiori, D. Salvadore, Corrosion Science 15 (1975) 383392.
F. King, C.D. Litke, M.J. Quin, D.M. LeNeveu, Corrosion Science 37 (1995) 833851.
J. Mathiyarasu, N. Palaniswamy, V.S. Muralidharan, Proceedings of the Indian Academy of
SciencesChemical Sciences 111 (1999) 377386.
H.P. Dhar, R.E. White, G. Burnell, L.R. Cornwell, R.B. Grin, R. Darby, Corrosion-NACE 41
(1985) 317323.
K. Kinoshita, D. Landolt, R.H. Muller, C.W. Tobias, Journal of the Electrochemical Society 117
(1970) 12461251.
H.P. Lee, K. Nobe, A.J. Pearlstein, Journal of the Electrochemical Society 132 (1985) 10311037.
M.E. Walton, P.A. Brook, Corrosion Science 17 (1977) 317328.
M.V. Vazquez, S.R. de Sanchez, E.J. Calvo, D.J. Schirin, Journal of Electroanalytical Chemistry 374
(1994) 179187.
D.D. Macdonald, Journal of the Electrochemical Society 121 (1974) 651656.
G.D. Preston, L.L. Bircumshaw, Philosophical Magazine 20 (1935) 706720.
U.R. Evans, Journal of the Chemical Society 642 (127) (1925) 24842491.
W.H.J. Vernon, Journal of the Chemical Society 643 (129) (1926) 22732282.
N.B. Pilling, R.E. Bedworth, Journal of the Institute of Metals 29 (1923) 529591.
H.A. Miley, Journal of the American Chemical Society 59 (1937) 26262629.
A. Hickling, D. Taylor, Transactions of the Faraday Society 44 (1948) 262268.
J. Kruger, Journal of the Electrochemical Society 106 (1959) 847853.
B. Miller, Journal of the Electrochemical Society 116 (1969) 16751680.
A.M. Shams El Din, F.M. Abd El Wahab, Electrochimica Acta 9 (1964) 113121.
B.B. Chernov, K.T. Kuzovleva, A.A. Ovsyannikova, Protection of Metals 21 (1985) 4246.
M. Whiteld, D. Jagner, Marine Electrochemistry: A Practical Introduction, John Wiley & Sons,
Chichester, UK, 1981.
R.F. North, M.J. Pryor, Corrosion Science 10 (1970) 297311.
R.G. Blundy, M.J. Pryor, Corrosion Science 12 (1972) 6575.
T.D. Burleigh, D.H. Waldeck, Corrosion-NACE 55 (1999) 800804.
J.M. Popplewell, R.J. Hart, J.A. Ford, Corrosion Science 13 (1973) 295309.
B.D. Craig, Fundamental Aspects of Corrosion Films in Corrosion Science, Plenum Press, London,
UK, 1991.
L.E. Eiselstein, B.C. Syrett, S.S. Wing, R.D. Caligiuri, Corrosion Science 23 (1983) 223239.
M. Yamashita, K. Omura, D. Hirayama, Surface Science 96 (1980) 443460.
K.R. Trethewey, J. Chamberlain, Corrosion for Science and Engineering, Longman, Harlow,
1995.
A.L. Bacarella, J.C. Griess Jr., Journal of the Electrochemical Society 120 (1973) 459465.
B. Tribollet, J. Newman, Journal of the Electrochemical Society 131 (1984) 27802785.
W.D. Bjorndahl, K. Nobe, Corrosion-NACE 40 (1984) 8287.
M. Georgiadou, R. Alkire, Journal of Applied Electrochemistry 28 (1998) 127134.
A.H. Taylor, Journal of the Electrochemical Society 118 (1971) 854859.
P.A. Lush, M.J. Carr, Corrosion Science 19 (1979) 10791088.
S.R. de Sanchez, D.J. Schirin, Corrosion Science 22 (1982) 585607.
W.H. Smyrl, Journal of the Electrochemical Society 132 (1985) 15561563.
D. Wagner, H. Peinemann, H. Siedlarek, The inuence of chloride ions on the corrosion performance
of DHP-copper and 90/10 coppernickeliron, in: S.A. Campbell, N. Campbell, F.C. Walsh (Eds.),
Developments in Marine Corrosion, The Royal Society of Chemistry, Cambridge, 1998, pp. 103118.

G. Kear et al. / Corrosion Science 46 (2004) 109135

135

[54] G. Kear, F.C. Walsh, D.B. Barker, K.S. Stokes, Electrochemical corrosion characteristics of copper
in ltered and articial seawater as a function of mass transfer conditions, in: EuroCorr 2000,
Institute of Corrosion, Leighton Buzzard, UK, 2000.
[55] R.K. Flatt, P.A. Brook, Corrosion Science 11 (1971) 185196.
[56] C. Deslouis, B. Tribollet, G. Mengoli, M. Musiani, Journal of Applied Electrochemistry 18 (1988)
374383.
[57] F. King, M.J. Quin, C.D. Litke, Journal of Electroanalytical Chemistry 385 (1995) 45.
[58] H. Lal, H.R. Thirsk, Journal of the Chemical Society (1953) 26382644.
[59] G. Venkatachari, K. Kannan, Bulletin of Electrochemistry 9 (1993) 400402.
[60] R. Greef, R. Peat, L.M. Peter, D. Pletcher, M.J. Robinson, Instrumental Methods in Electrochemistry, Ellis Horwood, Chichester, UK, 1985.
[61] V.N. Vesovic, N. Anastasijevic, R.R. Adzic, Journal of Electroanalytical Chemistry 218 (1987) 5363.
[62] V.G. Levich, Physicochemical Hydrodynamics, Prentice-Hall, Englewood Clis, NJ, USA, 1962.
[63] P.J. Boden, Corrosion Science 11 (1971) 353362.
[64] B.J. Little, F.B. Mansfeld, Werkstoe und Korrosion 42 (1991) 331340.
[65] J.N. Alhajji, M.R. Reda, Journal of the Electrochemical Society 141 (1994) 14321439.
[66] K.K. Satpathy, M.P. Srinivasan, S. Rangarajan, S.V. Narashiman, P.K. Mathur, Bulletin of
Electrochemistry 12 (1996) 6467.
[67] G.P. Power, I.M. Ritchie, Electrochimica Acta 26 (1981) 10731078.
[68] G. Kear, Electrochemical Corrosion of Marine Alloys Under Flowing Conditions, Thesis, University
of Portsmouth, UK, 2001.
[69] D.J. Schirin, The Electrochemistry of oxygen, in: D. Pletcher (Ed.), Specialist Periodical Reports:
Electrochemistry, The Royal Society of Chemistry, Cambridge, 1983, pp. 126170.
[70] K.-L. Hsueh, D.-T. Chin, S. Strinivasan, Journal of Electroanalytical and Interfacial Chemistry 153
(1983) 7995.
[71] P. Delahay, Journal of the Electrochemical Society 97 (1950) 205212.
[72] K. Balakrishnan, V.K. Venkatesan, Electrochimica Acta 24 (1979) 131138.
[73] J.S. Halliday, Transactions of the Faraday Society 50 (1954) 171178.
[74] M.V. Vazquez, S.R. de Sanchez, E.J. Calvo, D.J. Schirin, Journal of Electroanalytical Chemistry 374
(1994) 189197.
[75] J. Venzcel, L. Knutsson, G. Wranglen, Corrosion Science 4 (1964) 115.
[76] R. Kappesser, I. Cornet, R. Greif, Journal of the Electrochemical Society 118 (1971) 19571959.
[77] A. Bonnel, F. Dabosi, C. Deslouis, M. Duprat, M. Keddam, B. Tribollet, Journal of the
Electrochemical Society 130 (1983) 753761.
[78] J.R. Scully, H.P. Hack, D.G. Tipton, Corrosion-NACE 42 (1986) 462469.
[79] G.J.W. Radford, F.C. Walsh, J.R. Smith, C.D.S. Tuck, S.A. Campbell, Electrochemical and atomic
force microscopy studies of a copper nickel alloy in sulphide-contaminated sodium chloride solutions,
in: S.A. Campbell, N. Campbell, F.C. Walsh (Eds.), Developments in Marine Corrosion, The Royal
Society of Chemistry, Cambridge, 1998, pp. 4063.
[80] C. Wagner, W. Traud, Zeitschrift Fur Elektrochemie 44 (1938) 391402.
[81] G. Kear, D.B. Barker, F.C. Walsh, K.S. Stokes, The rotating disc electrode as a tool for the
determination of nickel aluminium bronze and copper corrosion rates in chloride media, in: Corrosion
Odyssey2001, Institute of Corrosion, Leighton Buzzard, UK, 2001, session 8/paper no. 6.
[82] L.M. Callow, J.A. Richardson, J.L. Dawson, British Corrosion Journal 11 (1976) 123131.
[83] J.C. Rowlands, M.N. Bently, British Corrosion Journal 7 (1972) 4246.
[84] M. Stern, A.L. Geary, Journal of the Electrochemical Society 104 (1957) 56.

You might also like