Вы находитесь на странице: 1из 6

a-CuSn bronzes in sulphate medium: Inuence of the tin content

on corrosion processes
J. Muller, B. Lak

, I. Guillot
Institut de Chimie et des Matriaux Paris Est, ICMPE, UMR 7182 CNRS-Universit Paris Est-Crteil, 2 rue Henri Dunant, 94320 Thiais, France
a r t i c l e i n f o
Article history:
Received 24 January 2013
Accepted 18 July 2013
Available online 13 August 2013
Keywords:
A. Alloy
A. Copper
A. Tin
B. Electrochemical calculation
B. Polarization
C. Passive lms
a b s t r a c t
Corrosion of synthetic CuSn bronzes was electrochemically investigated on rotating disc electrode (RDE)
in a buffered medium (pH = 6.8, deaerated 10
-2
mol L
1
Na
2
SO
4
solution). Copper species diffusion coef-
cients through the patina for various tin-content bonzes were analyzed and provided information about
its organization. A joined mass-transport and electronic transfer process controls the corrosion. A faster
and more efcient passivation was observed with highest bronze tin-content. Also, copper species diffu-
sion through the layer is independent of the tin content and few 10
10
cm
2
s
1
. Such an order of magni-
tude indicates that corrosion products layer behaves like a gel.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
CuSn bronzes are used in various application areas including
electronics and networking sectors [15], lithium battery electrode
materials [68], art foundry or archeological artefacts [9]. Over
time, the surface becomes covered by a spontaneous degradation
layer, named patina. Preservation of the original surface character-
istics is important in connector eld to maintain conductivity
properties. Also, the patina presence on art objects or archeological
artefacts can be regarded as aesthetic, protective and tends to be
extremely stable. Signicant changes in the surrounding environ-
ment can obviously be detrimental to the patina stability. Conse-
quently, the corrosion rate can rapidly increase and could lead to
irreversible degradations [10]. In this context, it is interesting to
know the diverse corrosion products which composed the patina
and how does this protective/isolating layer begin and evolve. This
is a key point to build up protection, preservation and/or restora-
tion procedures for cultural heritage [11].
Most of papers devoted to archeological bronzes on which nat-
ural patinas grown during exposition for long time in different
environments (soil, atmosphere, aqueous media, etc.) assumed in
a rst time that the corrosion layer formation occurs similarly to
that of pure copper. Indeed, the main constituents of natural pati-
nas have been found to be essentially copper-containing com-
pounds, sometimes involving a selective tin dealloying. However,
the development of characterization techniques (EDS, XPS, etc.)
had allowed detecting the systematical presence of tin compounds
in the patina, probably in amorphous form [9,1215].
Robbiola et al. [15,16] have proposed patina formation mecha-
nisms to explain most of the natural corrosion layers formed on
CuSn alloys. Depending on the aggressiveness level of the environ-
ment, they have identied two main classes of corrosion patterns,
Type I and Type II. Type I mainly occurs in lower aggressive
environments (e.g. carbonate, sulphate media). It leads to thin
tin-rich layer directly grown under the original alloy surface and
then passivates it. Oxidized copper species diffuse through this
tin-rich layer and access to the original surface where they either
react with anions to form insoluble copper (II) products or dissolve
in the environmental medium. Type II concerns essentially the
most aggressive environments, particularly environments contain-
ing chlorides or in the presence of heterogeneities in the alloy. Cor-
rosion can be local or generalized. In this case, the original alloy
surface recedes, possibly until complete dissolution.
More recently, corrosion products layers have been electro-
chemically formed at the surface of articial a-CuSn bronzes in
0.01 mol L
1
sulphate aerated aqueous solution [17,18]. Further
characterizations (SEM, EDS, FTIR, etc.) of the patina conrm that
its structure and composition are very similar to those observed
on archeological artefacts and the involved mechanism is coherent
with Type I mechanism.
Since such a mechanism involving copper species diffusion
through the tin-rich layer, the aims of this paper is to determine
the corresponding diffusion coefcient values and the possible
0010-938X/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2013.07.025

Corresponding author. Tel.: +33 1 4978 1145; fax: +33 1 4978 1148.
E-mail address: laik@icmpe.cnrs.fr (B. Lak).
Corrosion Science 77 (2013) 4651
Contents lists available at ScienceDirect
Corrosion Science
j our nal homepage: www. el sevi er . com/ l ocat e/ cor sci
inuence of tin content. Three synthetic single-phase CuSn alloys
were prepared, namely Cu7Sn, Cu11Sn and Cu14Sn. These alloys
were also linearly anodically polarized in a potential range close
to the open-circuit voltage, with a rotating disc electrode (RDE)
in a deaerated 0.01 mol L
1
sulphate, pH-buffered and thermo-
regulated aqueous solution. Different RDE rates or different linear
sweep rates have been applied. Experimental electrochemical
curves are presented and discussed. Characteristic values (peak
current density J
p
and potential at which the peak current density
is obtained E
p
) verify conventional electrochemical laws. Their
exploitations lead to the coefcient diffusion determination.
2. Theoretical considerations
2.1. Scan rate effect studies
Scan rate can inuence the current responses of the system.
Corresponding electrochemical parameters are deduced from the
relation between scan rate of the potential sweep m and current
responses of the material. Whether the system is reversible or irre-
versible, the oxidation peak current density J
p
dependence vs. the
square root of scan rate

m
p
is linear. Depending of the case, equa-
tions are [1921]:
J
p
0:446nF
3=2

D
T
r
C

m
p
reversible process 1
J
p
0:496nF
3=2

a
p

D
RT
r
C

m
p
irreversible process 2
where J
p
(A cm
2
) is the peak current density of the J f

m
p
curve,
n, the number of transferred electrons, F, the Faradays constant (C),
a, the charge transfer coefcient, D, the diffusion coefcient (cm
2
-
s
1
), R, the molar gas constant (J mol
1
K
1
), T, the temperature (K),
C, the concentration (mol cm
3
) and m, the potential scan rate
(V s
1
).
A reversible system means that the overall reaction is under dif-
fusion control and the charge transfer reaction proceeds rapidly
enough to consider the reaction in an equilibrium state. In this
case, the equilibrium potential is given by the Nernst equation.
The potential E
p
at which the maximum of current density is ob-
tained is independent of m according to the following relation [19]:
E
p
1:109
RT
nF

E
1=2
3
where E
p
is the peak potential (V) and E

the half peak potential (V).


On the contrary, when system is irreversible, the charge transfer
rate is low, and the Nernst relation cannot be applied. There is a
linear relation between the peak potential E
p
and the logarithm
of the scan rate ln(m) (Eq. (4)) [19,20]:
E
p
E

RT
anF
0:780 ln

D
p
k

!
ln

anFm
RT
r ! !
4
where E is the standard potential (V) and k is the electronic trans-
fer rate constant (cm s
1
).
According to previous considerations for irreversible systems, J
p
and E
p
are linked by (Eq. (5)):
J
p
0:227nFCk

exp
anF
RF
E
p
E


5
This equation can also be expressed as follows:
lnJ
p

anF
RT
E
p
ln0:227nFCk


anF
RT
E


6
2.2. Rotating disc electrode (RDE) rate studies
The rotation disc electrode rate is an efcient tool to discrimi-
nate whether the electrochemical system is limited by mass trans-
port, electronic transfer or a mixed mass transport/electronic
transfer.
When the limitation is only due to mass transport, the peak cur-
rent density J
p
discussed above can be considered equivalent to the
diffusion limiting current density J
L
. Its dependence on the square
root of RDE rate

X
p
is given by the Levichs equation [19] (Eq. (7)).
J
L
J
p
0:620nFCD
2=3
m
1=6

X
p
7
where J
L
is the diffusion limiting current density (A cm
2
), J
p
, the
measured peak current density (A cm
2
), m, the kinematic viscosity
of the electrolyte (cm
2
s
1
) and X, the electrode rotation rate
(rad s
1
).
In more complicated cases, this relation, only veried at low
rotational rates, suggests that J
p
is controlled by charge transfer ki-
netic and not by mass transport. Indeed, when the rotation rate is
too high, the thickness of the Levich layer decreases and the cur-
rent magnitude is controlled by the charge transfer kinetic, charac-
terized by J
K
. In this case, the KouteckyLevich equation establishes
the current densities relation (Eq. (8)) [2224]:
1
J
p

1
J
K

1
J
L

1
J
K

1
0:620nFCD
2=3
m
1=6
1

X
p 8
where J
K
is the current density in absence of any mass-transport ef-
fect (A cm
2
).
3. Experimental
3.1. Materials
Three different bronzes of high purity were synthetized by
melting Cu (Copper rod, 12.7 mm, Goodfellow, 99.95%) and Sn
(Tin rod, 6 mm, Puratronic, 99.9985%, Alfa Aesar) in a water-cooled
carbon graphite crucible by electromagnetic high frequency (Celes
generator) under helium atmosphere. A second induction-melting
in electromagnetic levitation has been performed, with an over-
heating of more than 100 degrees above the liquidus temperature
in order to ensure a perfect chemical homogeneity of the melt. Al-
loys are homogenized under helium atmosphere at 750 C during
24 h in order to avoid tin segregation, and then annealed at
400 C during 2 h before water-quenching. The working electrodes
for electrochemical experiments are prepared by repeating two
times the following sequence: cold-rolling until 60% of deforma-
tion and recrystallisation annealing at 500 C during 1 h. Finally,
the alloys are water-quenched after the last annealing at 500 C
to obtain single-phase alloys. The samples are then machined at
4 mm diameter cylinders (S = 0.126 cm
2
) to remove the external
part that could present high temperature oxides. Then bronze rods
are pressed in an inert resin in order to obtain a rotating disc
electrode.
The three different binary alloys compositions are determined
by Inductively Coupled Plasma Optical Emission Spectroscopy after
conventional Aqua Regia digestion and correspond to 7.0, 10.9 and
13.8 0.2 weight% (wt%) tin. They are respectively named Cu7Sn,
Cu11Sn and Cu14Sn. The light micrographic examination of the
copper-tin microstructure reveals a single-phase a-CuSn solid
solution without hard and brittle d phase. As shown in Fig. 1, all
bronzes reveal an annealed microstructure characterized by ther-
mal twins. The average grain size, determined by image analysis,
is around 10 lm for Cu7Sn and Cu10Sn and around 30 lm for
Cu14Sn.
J. Muller et al. / Corrosion Science 77 (2013) 4651 47
3.2. Electrochemistry
All the electrochemical experiments are performed using an
Autolab PGSTAT 30 (Eco Chemie) Potentiostat/Galvanostat in a
classical three-electrode cell (capacity ca. 20 cm
3
). The counter
electrode is a grid of platinum. The reference electrode is a satu-
rated calomel electrode, SCE (Hach Lange-France), with an equilib-
rium potential of 0.24 V vs. SHE at working temperature used with
a separated compartment in order to avoid the diffusion of chloride
into the corrosive medium. The working electrode is the rotating
disc electrode made of bronzes (S = 0.126 cm
2
).
The 10
-2
mol L
1
sulphate electrolyte is prepared from H
2
SO
4
(SDS, purity = 95 %) and 1,4-piperazinediethanesulfonic acid buffer
reagent (disodium salt-PIPES, Aldrich, pH-buffering agent with
pK
a
= 6.8) used in ad hoc proportions to buffer the solution at
pH = 6.8. The cell temperature is regulated at 25.0 0.1 C with a
thermostat and purged with argon before and during the measure-
ments in order to limit the oxygen content.
The anodic polarization curves are all carried out from the
open-circuit potential Eoc. Firstly the inuence of a varying poten-
tial scan rate m (0.5, 2, 5, 20, 50 and 100 mV s
1
) and then, of a vary-
ing rotating rate X (20.9, 52.4, 104.7, 209.4 and 418.9 rad s
1
) are
studied.
4. Results
4.1. Surface preparation
To ensure the reproducibility of electrochemical experiments,
each experiment has been carried out with the same initial surface
state, which has been controlled by open-circuit potential Eoc mea-
surement. To obtain such conditions, the native oxide lm is re-
moved from the metallic surface before each experiment.
Different surface preparation methods were tested. Among them,
a single mechanical grinding procedure with SiC paper up to
2400 grade, a combination of mechanical grinding procedure and
a cathodically polarization step at 1.0 V/SCE or a combination
of mechanical grinding procedure and cathodically polarization
according to a pulsed method (multiple potential-steps between
1.0 V/SCE and 0.6 V/SCE) [25].
Considering the consecutive reduction curves recorded after
these different surface preparation methods (Fig. 2), it clearly ap-
pears that the third method is the most efcient since the cathodic
current is lower. Moreover, the reproducibility of this method is
guaranteed by a very small open circuit potential deviation mea-
sured at the end of the procedure.
4.2. Anodic polarization curves
The bronze oxidation is studied through the application of an
anodic potential ramp from Eoc. Fig. 3 represents, for Cu7Sn, a full
anodic polarization curve (m = 5 mV s
1
, X = 104.7 rad s
-1
) (Fig. 3a)
and zoomed curves conducted with different RDE rates X and con-
stant scan rate (m = 5 mV s
1
) (Fig. 3b).
The full anodic polarization curve exhibits, as expected for a pas-
sivable material, an activity peak at 40 mV/SCE followed by a cur-
rent density plateau at 260 lA cm
2
up to roughly 100 mV/SCE.
For higher potentials, the current rapidly increases as the alloy dis-
solution takes place. The curve shape is similar for Cu11Sn and
Cu14Sn [26]. Such polarization curve has been already presented
in literature with Cu13Sn in aerated 0.01 mol L
-1
sulphate solution
at m = 0.3 mV s
1
[17]. In our case, the current density peak is not as
well dened because of a faster potential scan rate (17 times faster).
Fig. 3b highlights that, at constant potential scan rate, both J
p
and E
p
values are inuenced by the rotation disc electrode rate.
The potential scan rate at constant RDE rate inuences J
p
and E
p
values too. Conclusions are similar with Cu11Sn and Cu14Sn
(curves no presented here). Table 1 gathers J
p
and E
p
values for
the three compositions in each condition. In the following of the
Fig. 1. Light micrographic observations of fully recrystallised bronzes (a) Cu7Sn, (b) Cu11Sn, and (c) Cu14Sn.
Fig. 2. Cathodic polarization curves of the Cu7Sn bronze from Eoc (m = 50 mV s
-1
,
X = 104.7 rad s
1
) after (a) mechanical grinding procedure, (b) mechanical grinding
procedure + cathodic polarization at E = 1.0 V/SCE, and (c) mechanical grinding
procedure + cathodic polarization at multiple potential-steps (in deaerated (argon)
0.01 mol L
1
sulphate solution).
48 J. Muller et al. / Corrosion Science 77 (2013) 4651
manuscript, used E
p
values are systematically corrected from the
ohmic drop considering an electrolyte resistance of 25 Xcm
2
.
4.3. Scan rate effect
Whatever the bronze, the increase of m has a delaying effect on
the alloy passivation as it is clearly evidenced by the E
p
raise
(Table 1). In relation with the theoretical part, since E
p
varies with
scan rate, the system is irreversible (Eq. (4)). In such conditions, J
p
must vary linearly with

m
p
according to Eq. (2) and ln(J
P
) must
increase linearly with E
p
according to Eq. (6). Graphs J
p
f

m
p

(Fig. 4) and lnJ
p
= f(E
p
) (Fig. 5) plotted for the three tin contents
exhibit straight lines with respective slopes a
1
(Eq. (9)) and a
2
(Eq. (10)).
a
1
2:99 10
5
nanD
1=2
C 9
a
2

anF
RT
10
The slope values a
1
and a
2
determined from linear regressions are
gathered in Table 2.
4.4. Rotating disc electrode rate studies
The relative contributions of mass transport and electronic
transfer are studied in this part by changing the rotating electrode
rate. Since the J
p
dependence with

X
p
is not linear (i.e. Eq. (7) is
not veried), a mixed process between charge transfer and mass
transport should operate. From this perspective the variation of
1
J
p
with
1
X
(Eq. (8)) has been plotted and reported on Fig. 6. The three
obtained curves are linear as expected. The slope values a
3
(Eq.
(11)) are also presented in Table 2.
Fig. 3. Anodic polarization curves of Cu7Sn bronze in a deaerated 0.01 mol L
1
sulphate medium (m = 5 mV s
1
) (a) full curve (X = 104.7 rad s
1
), and (b) zoomed curves with
different X rates.
Table 1
J
p
and E
p
values for the Cu7Sn, Cu11Sn and Cu14Sn bronzes determined with variable m and xed X = 104.7 rad s
1
(upper part) and with variable X and xed m = 5 mV s
-1
(lower
part).
Cu7Sn Cu11Sn Cu14Sn
J
p
(lA cm
2
) E
p
(mV/SCE) J
p
(lA cm
2
) E
p
(mV/SCE) E
p
(lA cm
2
) J
p
(mV/SCE)
m (mV s
1
) X = 104.7 rad s
1
0.5 106 24 38 20
2 190 15 72 9
20 528 47 430 80 249 56
50 889 82 790 120 380 70
100 1247 101 1102 140 570 90
X (rad s
1
) m = 5 mV s
1
20.9 57 21
52.4 220 30 150 25 64 22
104.7 260 40 170 29 78 25
209.4 290 44 82 26
418.9 320 49 190 15
Fig. 4. Plot of current density J
p
vs. square root of the scanning rate

m
p
of Cu7Sn,
Cu11Sn and Cu14Sn bronzes in sulphate medium: a
1
coefcient determination.
J. Muller et al. / Corrosion Science 77 (2013) 4651 49
a
3

1
0:620nFCD
2=3
m
1=6
11
4.5. Estimation of the D value for the different bronzes
FromEq. (10),

a
p
is expressed as a function of a
2
. It is introduced
in Eq. (9) to obtain the following relation between C, a
1
and a
2
:
C
a
1

F
p
2:99 10
5
n

DRTa
2
p 12
Substituting C in Eq. (11) enables to determine the diffusion coef-
cient D of the copper cations through the corroded layer. Its literal
expression is given by Eq. (13).
D
2:99 10
5

RTa
2
p
0:620a
1
F
3=2
m
1=6
a
3
!
6
13
where R = 8.314 J mol
1
K
1
, T = 298 K, F = 96485 C. m is the kine-
matic viscosity expressed in cm
2
s
1
. Assuming that the corroded
layer is porous and pores are lled with aqueous corrosive medium,
the value of kinematic viscosity of water at 298 K is taken (10
-
2
cm
2
s
-1
).
For each bronze, the D coefcient is calculated from a
1
, a
2
and a
3
experimental values using Eq. (13) (Table 3).
5. Discussion
5.1. Inuence of the tin content
Figs. 46 highlight that a-CuSn bronzes behaviour depends on
the tin content. Firstly, as shown on Fig. 4, whatever the potential
scan rate, the maximum of current density available in the studied
potential range is always lower for the higher tin content alloys.
Corrosion potential and current density measurements were car-
ried out also. Gathered in Table 4 for m = 5 mV s
1
and
X = 104.7 rad s
-1
, they indicate that the corrosion rate diminishes
as the tin content increases while E
corr
remains quite constant.
These comments suggest on the one hand that bronze cannot be
regarded as signicantly nobler by increasing tin content and on
the other hand that once the oxide layer is formed, the passive
behaviour is more efcient in the bronze with the higher tin con-
tent. Furthermore, the transient active corrosion step is less pro-
nounced for higher tin content, in agreement with previous
studies [14,26]. This statement is conrmed by the relative posi-
tion of the three straight lines plotted in Fig. 6. Indeed, when the
charge transfer is the limiting step, i.e. when
1

X
p
tending towards
zero, the current density decreases with the increasing tin content.
Considering that the oxidation process occurs at the alloy/oxida-
tion products interface, this decreasing current density can be ex-
plained either in term of passivation layer thickness increase or by
the lm properties changes.
This point is reinforced by the different slope values grouped in
Table 2 that are very close for Cu7Sn and Cu11Sn whereas they are
different for the Cu14Sn bronze. So, as Eq. (10) involves only con-
stant terms, F, R and T, either a, the charge transfer coefcient, or n,
the number of transferred electrons, may change with higher tin
content, in accordance with previous remark. This difference of
corrosion layer electrochemical establishment is probably respon-
sible of the relative tin enrichment in the passive layer and of the
more efcient passivity already reported in the literature for higher
tin contents [14,27,28].
5.2. Diffusion coefcient
Though tin content has an inuence on the electrochemical
behaviour of the bronzes as discussed above, it does not seem to
Fig. 5. Plot of neperian logarithm current density ln (J
P
) vs. signal potential E
p
of
Cu7Sn, Cu11Sn and Cu14Sn bronzes in sulphate medium: a
2
coefcient
determination.
Table 2
Values of the slopes a
1
, a
2
and a
3
of Cu7Sn, Cu11Sn and Cu14Sn bronzes.
Cu7Sn Cu11Sn Cu14Sn
a
1
(mA s
1/2
cm
2
V
1/2
) 3.93 3.42 1.75
a
2
(V
1
) 22.2 20.6 29.2
a
3
(cm
2
rad
1/2
s
1/2
mA
1
) 15.8 15.6 37.4
Fig. 6. Plot of reciprocal current density
1
J
P
vs. reciprocal square root of RDE
1

X
p of
Cu7Sn, Cu11Sn and Cu14Sn bronzes in sulphate medium: a
3
coefcient
determination.
Table 3
Diffusion coefcient D calculated for Cu7Sn, Cu11Sn and Cu14Sn.
Cu7Sn Cu11Sn Cu14Sn
D (cm
2
s
1
) 5 10
10
10 10
10
8 10
10
Table 4
Corrosion potential E
corr
and corrosion current density J
corr
values for Cu7Sn, Cu11Sn
and Cu14Sn (X = 104.7 rad s
1
, m = 5 mV s
-1
).
Cu7Sn Cu11Sn Cu14Sn
E
corr
(V/SCE) 0.10 0.10 0.09
J
corr
(mA cm
2
) 0.25 0.13 0.09
50 J. Muller et al. / Corrosion Science 77 (2013) 4651
have a noticeable inuence on the diffusion coefcient D since they
are in the same order of magnitude (Table 3). It demonstrates that
the diffusing species and the diffused medium are similar what-
ever the tin content, even if reaction mechanism and the corrosion
product layer structure and/or composition differ at high tin
content.
The calculated D value excludes in our case a mass transport
within a copper oxide lm. Indeed, expected values for diffusing
copper species in a solid lm of copper oxide are close to 10
14
10
15
cm
2
s
1
at roomtemperature [29,30]. Furthermore, the diffu-
sion cannot concern ionic species through an aqueous solution
because the order of magnitude of D should be in the range of
10
5
cm
2
s
1
for Sn
2+
species [31].
In the present work, the diffusion coefcient is intermediate,
about 10
9
10
10
cm
2
s
1
. In literature, such ranges of value are
reported. Indeed, some authors mentioned a diffusion coefcient
value of 5.10
9
cm
2
s
1
for species diffusion through a zirconium
hydrophosphate gel [32] and of 10
11
cm
2
s
1
for protons diffusion
through an antimony oxide layer formed by anodic polarization in
an H
2
SO
4
medium [33]. It can then be thought that the passivation
layer behaves like a gel. This conclusion is in agreement with for-
mer works presented by Robbiola and coworkers [27,34] which
suggests that CuSn corrosion layer formed in a 0.1 mol L
1
NaCl
medium could be assimilated to a xerogel. In such xerogel, Sn
(IV) compounds could aggregate to form colloidal particles and
then could act as a network former like it can be observed in sili-
cate glasses [35]. Whatever the tin content in the studied bronzes
such a corrosion layer structure is probably obtained. Through the
xerogel layer made of tin oxide compounds, copper cations migrate
outwards from the alloy. Articial layers and ancient bronzes pres-
ent the same prole [9,36] at the origin of the decuprication phe-
nomenon reported in the literature [11,15,27,28,37].
6. Conclusion
This paper focused on the use of electrochemical methods to
specify the tin inuence on the bronzes behaviour in a 10
2
mol L
1
sulphate medium. The main ndings are summarized
below:
1. The irreversible bronze/sulphate medium system is under a
combined control of mass and electronic transfers during the
bronze oxidation.
2. Addition of tin improves the bronze corrosion resistance.
Indeed, a decrease of both the mass transport and the charge
transfer contributions in the oxidation processes of high tin
content bronzes has been proved.
3. The behaviour of the Cu7Sn and Cu11Sn bronzes are similar
whereas Cu14Sn behaves differently. These differences could
be explained by a noticeable difference of the electrochemical
corrosion layer formation.
4. Diffusion coefcients of the copper cations coming from the
alloy to the electrolyte have been determined for the bronze/
sulphate medium system. The value of a few 10
10
cm
2
s
1
indi-
cates that rich tin compounds corrosion layer formed at the
bronze surface behaves like a gel.
Acknowledgments
The authors would like to thank P. Ochin and J.-L. Bonnentien
for the bronzes elaboration, L. Robbiola for fruitful discussions
and N. Emery for careful reading.
References
[1] C.K. Hu, B. Luther, F.B. Kaufuman, J. Hummel, C. Uzoh, D.J. Pearson, Thin Solid
Films 262 (1995) 8492.
[2] K.L. Lee, C.K. Hu, K.N. Tu, J. Appl. Phys. 78 (1995) 44284437.
[3] K.N. Tu, Mater. Chem. Phys. 46 (1996) 217223.
[4] D. Padhi, S. Gandikota, H.B. Nguyen, C. McGuirk, S. Ramanathan, J. Yahalom, G.
Dixit, Electrochim. Acta 48 (2003) 935943.
[5] K.N. Tu, J. Appl. Phys. 94 (2003) 54515473.
[6] W. Pu, X. He, J. Ren, C. Wan, C. Jiang, Electrochim. Acta 50 (2005) 41404145.
[7] J.S. Thorne, J.R. Dahn, M.N. Obrovac, R.A. Dunlap, J. Power Sources 216 (2012)
139144.
[8] R.M. Gnanamuthu, S. Mohan, C.W. Lee, Mater. Lett. 84 (2012) 101103.
[9] T. Beldjoudi, F. Bardet, N. Lacoudre, S. Andrieu, A. Adriens, I. Constantinides, P.
Brunella, Rev. Metall.-Paris 98 (2001) 803808.
[10] T.E. Graedel, K. Nassau, J.P. Franey, Corros. Sci. 27 (1987) 639657.
[11] L. Robbiola, L.-P. Hurtel, Mm. Et. Sci. Rev. Mt. 12 (1991) 809823.
[12] L. Soto, J.P. Franey, T.E. Graedel, G.W. Kammlott, Corros. Sci. 23 (1983) 241
250.
[13] N.D. Meeks, Archaeometry 28 (1986) 133162.
[14] L. Roobiola, PhD Thesis, Universit Paris 6-Pierre et Marie Curie, France, 1990,
<http://tel.archives-ouvertes.fr/tel-00495356/>.
[15] L. Robbiola, J.M. Blengino, C. Fiaud, Corros. Sci. 40 (1998) 20832111.
[16] L. Robbiola, C. Fiaud, Revue dArchomtrie 16 (1992) 109119.
[17] I. Mabille, A. Bertrand, E.M.M. Sutter, C. Fiaud, Corros. Sci. 45 (2003) 855866.
[18] E. Sidot, N. Souissi, L. Bousselmi, E. Triki, L. Robbiola, Corros. Sci. 48 (2006)
22412257.
[19] A.J. Bard, L.R. Faulkner, Electrochemical Methods, Fundamentals and
Applications, second ed., Wiley, New York, 2001.
[20] F. Scholz, Electroanalytical Methods: Guide to Experiments and Applications,
second revised and extended ed., Springer, Verlag, Berlin-Heidelberg, 2010.
[21] R.E. White, Modern Aspects of Electrochemistry, Number 45, Springer
Science+Business Media, LLC., 2009.
[22] S. Strbac, N.A. Anastasijevic, R.R. Adzic, Electrochim. Acta 39 (1994) 983990.
[23] H. Cheng, K. Scott, Electrochim. Acta 51 (2006) 34293433.
[24] B.G. Ateya, J.D. Fritz, H.W. Pickering, J. Electrochem. Soc. 144 (1997). 2606
2013.
[25] J. Muller, PhD Thesis, Universit Paris-Est Crteil, France, 2010. <http://
tel.archives-ouvertes.fr/tel-00492692/fr/>.
[26] J. Muller, G. Lorang, E. Leroy, B. Laik, I. Guillot, Corros. Eng., Sci. Technol. 45
(2010) 322326.
[27] L. Robbiola, C. Fiaud, A. Harch, Europ. Fed. of Corrosion, in: P. Marcus, B.
Baroux, M. Keddam, (Eds.), The Institute of Materials, vol. 12, 1994, pp. 150
154.
[28] L. Robbiola, R. Portier, J. Cult. Heritage 7 (2006) 112.
[29] M. Metikos-Hukovic, R. Babic, I. Paic, J. Appl. Electrochem. 30 (2000) 617624.
[30] J.O. Zerbino, L.M. Gassa, J. Solid State Electr. 7 (2003) 177182.
[31] T. Laitinen, K. Salmi, G. Sundholm, P. Viinikka, A. Ylipentti, Electrochim. Acta 37
(1992) 17971803.
[32] Y.S. Dzyazko, L.M. Rozhdestvenska, A.V. Palchik, F. Lapicque, Sep. Purif.
Technol. 45 (2005) 141146.
[33] M. Bojinov, D. Pavlov, J. Electroanal. Chem. Interfacial Electrochem. 315 (1991)
201216.
[34] L. Robbiola, T.T.M. Tran, P. Dubot, O. Majerus, K. Rahmouni, Corros. Sci. 50
(2008) 22052215.
[35] E.C. Ziemath, B.Z. Saggioro, J.S. Fossa, J. Non-Cryst. Solids 351 (2005) 3870
3878.
[36] I. Constantinides, A. Adriens, F. Adams, Appl. Surf. Sci. 189 (2002) 90101.
[37] L. Muresan, S. Varvara, E. Stupniek-Lisac, H. Otmai, K. Marui, S. Horvat
Kurbegovi, L. Robbiola, K. Rahmouni, H. Takenouti, Electrochim. Acta 52 (2007)
77707779.
J. Muller et al. / Corrosion Science 77 (2013) 4651 51

Вам также может понравиться