You are on page 1of 7

Corrosion Science 53 (2011) 13941400

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Electrochemical aspects of exfoliation corrosion of aluminium alloys: The effects


of heat treatment
T. Marlaud a,b, B. Malki a,, A. Deschamps a, B. Baroux a
a
b

Laboratoire Science et Ingnierie des Matriaux et Procds (SIMAP), Grenoble INP, UJF, CNRS, 38402 Saint Martin d, Hres, France
Rio Tinto Alcan Centre de Recherches de Voreppe, BP 27, 38341 Voreppe Cedex, France

a r t i c l e

i n f o

Article history:
Received 10 June 2010
Accepted 8 January 2011
Available online 18 January 2011
Keywords:
A. Aluminium
B. Galvanostatic
C. Exfoliation corrosion
C. Hydrogen absorption

a b s t r a c t
Electrochemical approaches are used to investigate the exfoliation corrosion (EFC) of a 7XXX series aluminium alloy that has undergone different tempering treatments. EFC was produced under an articial
crevice at open circuit potential in neutral chloride solutions, and is found to be associated to current
and potential transients. EFC was also produced under galvanostatic control conditions. Observations
made through Scanning Electron Microscopy (SEM) suggest that these transients result from the progression of inter-granular cracks. Last, over-ageing heat treatments that are known to decrease both metal
hardness and EFC sensitivity were found to decrease the number of transients.
2011 Elsevier Ltd. All rights reserved.

1. Introduction
Aluminium alloys with an elongated grain structure parallel to
the plate surface are known [1] to be sensitive to exfoliation corrosion. High strength 2XXX and 7XXX series alloys used in aircraft
are particularly sensitive to this form of inter-granular corrosion
[2], which is considered to be a major cause of airframe degradation [3]. For instance, it has been shown that EFC above a certain
critical level could signicantly decrease fatigue life [4]. The resistance of such alloys to EFC is generally increased through over-ageing treatments, which unfortunately, signicantly decrease the
mechanical strength [1].
In the last decade, the effect of heat treatment on the corrosion
behaviour of aluminium alloys has been extensively studied as evidenced by the works on liform corrosion [58]. More recently the
effect of temper on localised corrosion kinetics of 7XXX alloys has
been studied using the foil penetration technique [9], which presents the advantage of measuring corrosion rates under various
experimental conditions. Some electrochemical transients were
observed that the authors attributed to the selective grain attack
combined with inter-granular corrosion, both controlled by the
alloying elements content [10,11]. Having this in mind, we intended in this paper to investigate the inuence of tempering on
the EFC behaviour of some new 7XXX aluminium alloys.
The mechanisms of EFC generally involve some type of mechanical wedging stresses, created by the precipitation of aluminium

Corresponding author. Fax: +33 4 76 82 67 67.


E-mail address: Brahim.Malki@simap.grenoble-inp.fr (B. Malki).
0010-938X/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2011.01.010

chloride/oxy-chloride complexes, which have higher molar volumes than that of aluminium [12]. This precipitation occurs along
grain boundaries and may cause the surface grains to lift up [13].
Other possible mechanisms, such as stresses caused by hydrogen
bubbles or hydrogen embrittlement are barely taken into consideration, despite the fact that hydrogen originating from the cathodic
reaction in an acidic electrolyte can be absorbed by the aluminium
matrix, inducing then a degradation of its mechanical properties
[1417]. Therefore, the question arises whether hydrogen generated by the cathodic reaction may also contribute to EFC through
the mechanisms described in the literature [18,19].
In terms of experimentation, various procedures have been proposed to assess the susceptibility of aluminium alloys to EFC. Alloy
and aircraft manufacturers commonly use tests such as those set
forth in ASTM G66, ASTM G85, and ASTM G34 (the so-called EXCO,
i.e. exfoliation corrosion, test) [2023]. However, most of these
tests, although widely accepted, yield only qualitative results.
The EXCO test, for instance, is based on the visual examination of
a surface corroded in an acidic medium. Thus, a method that would
provide a quantitative assessment of EFC sensitivity would be
useful to optimise alloys and tempering and to predict lifetimes
[11,24].
This study is an attempt to better understand the EFC mechanisms at work in 7XXX aluminium alloys with different heat
treatment histories, including the standard industrial tempers T6
(used to maximise metal hardness) and T76 (used to improve
EFC resistance). It examines the electrochemical transients
observed in chloride-based electrolytes at open circuit potential
(OCP), or under galvanostatic, and in some cases under potentiostatic conditions. OCP measurements are advantageous in that they

T. Marlaud et al. / Corrosion Science 53 (2011) 13941400

135C
15C/min
As received:
6h at 120C

12h

24h

48h

96h

T6

~130h
T76

Fig. 1. Ageing heat treatments used in the present study: various durations all at
135 C, following an initial 6-h ageing phase at 120 C. The T6 and T76 tempers
correspond, respectively, to 12 and 96 h at 135 C, and will be the focus of this
paper.

measure current and potential uctuations under free corroding


conditions, which are closer to in-service conditions than the standard ASTM tests. The electrochemical tests were supplemented by
extensive SEM fractographic observations.
2. Experimental
Samples were taken from a 25 mm plate of a non-commercial
aluminium alloy provided by Rio Tinto Alcan. The chemical composition of this AlZnMgCu alloy was: 10.3 wt.% Zn, 2.0 wt.% Mg
and 1.6 wt.% Cu. Small additions of Zr and Sc in this alloy ensure
that the grain structure is fully brous, i.e. that the grain structure
is highly elongated parallel to the sample surface (the rolling
plane), following a rolling and solution treatment. In accordance
with standard industrial practices, the alloy rst underwent a solution treatment and was quenched to room temperature in cold
water. It was next subjected to a small amount of plastic deformation and aged 6 h at 120 C (as received condition). Further ageing
treatments were then performed at 135 C with different holding
times, including the T6 and over-aged T76 tempers (Fig. 1).

1395

All electrochemical measurements were performed on samples


mechanically polished down to 1 lm. The corrosion attacks were
mainly performed on the rolling plane Longitudinal/Long-Transverse (L/LT, see Fig. 2a) at one-fourth the thickness of the initial
plate. Some transverse corrosion attacks were also performed,
where the electrolyte is in contact with the Short-Transverse/
Long-Transverse (ST/LT, see Fig. 2b) plane of the substrate.
The OCP measurements were performed in a 1 M NaCl, 0.25 M
NaNO3 aerated aqueous solution at pH 5.5 [25]. The experimental
setup included two working electrodes (WE) and a reference electrode (SCE) (Fig. 3a). The two working electrodes were some disks
measuring approximately 20 mm in diameter, the L/LT plane being
in contact with the electrolyte. On one of the two working electrodes an articial crevice was created by pressing a polymethyl
methacrylate (PMMA) cylinder onto the sample (see the assembly
in Fig. 3b). The presence of this crevice-forming system served to
greatly accelerate initial exfoliation corrosion under open circuit
conditions. The conned area between the PMMA cylinder and
the surface of the specimen was controlled by applying a torque
of 0.02 N m.
The two working electrodes were connected through a low
impedance ammeter, which made it possible to measure the OCP
corrosion current. The difference in potential between the two
electrodes was less than 250 lV, and their potential with respect
to the reference electrode SCE was measured using a high input
impedance voltmeter (1013 O). During the corrosion tests, the data
sampling rate was equivalent to fs = 18 Hz. An anti-aliasing lter
with a cut-off frequency set to 7.66 Hz was used to eliminate unwanted high-frequency interference.
Galvanostatic measurements were performed using a standard
three-electrode cell (a single working electrode, a platinum auxiliary electrode and the reference electrode SCE). The working electrode corresponded to the crevice-forming device. The electrolyte

Fig. 2. Diagram of the two specimens taken from the sheet of metal. The arrows indicate the surfaces tested during the galvanostatic experiments. The acronyms L/LT and ST/
LT stand for longitudinal/long-transverse and short-transverse/long-transverse.

im

(a)

(b)

ZRA

Crevice

sample
V1

PMMA support

V2

Fig. 3. (a) Diagram of the experimental setup with two working electrodes (subscripts 1 and 2) connected through a low resistance ammeter (ZRA). im is the measured
current through the ammeter and V1 = V2 = V is the measured potential; (b) a diagram of the screw/wing system of the articial crevice assembly.

1396

T. Marlaud et al. / Corrosion Science 53 (2011) 13941400

-708

-709

V(t)
-710
4

-711

incubation time

I(t)

Potential (mV/SCE)

Intensity (A)

equal to that obtained at the end of the EXCO test [2629] and also
corresponds to that of the crack tip solution measured during the
environmentally-assisted cracking of alloy AA7050 [30]. Preliminarily experiments were conducted to optimise the applied current
IG, in order to reduce experimental time and achieve noticeable
exfoliation corrosion morphologies. The IG value of 2.5 mA was
subsequently used for all experimentation.
3. Results
3.1. Open circuit testing

-712
0

-713
0

20

40
Time (h)

60

Fig. 4. Current and potential transients occurring during the OCP measurements of
the T6 specimen in the 1 M NaCl, 0.25 M NaNO3 electrolyte. Exfoliation corrosion
occurs after 20 h of incubation.

The OCP experiments involved a 64-h period of immersion for


the T6-tempered specimen. Analysis of the current and potential
signals (Fig. 4) showed that electrochemical activity occurred after
a long period of incubation (20 h). This activity was characterised
by an increase in current, associated with a drop in potential,
and followed by a gradual return to steady state values. The rst
idea which comes in mind is to suspect that these transients originate from metastable pitting [3133]. However, micrographic
observations suggest that these transients reveal something more.
Indeed, exfoliation corrosion was clearly evident on the working
electrode by way of the SEM observations performed at the end
of the test (Fig. 5). Moreover, careful observations during the early
stages of the experiment indicated that some blisters appear in the
form of protuberances (Fig. 6), in the absence of any visible corrosion product around the site, suggesting a possible effect of
hydrogen.
3.2. Galvanostatic measurements

Fig. 5. SEM image of the T6 specimen with the articial crevice assembly, after 48 h
of immersion in the 1 M NaCl, 0.25 M NaNO3 electrolyte at open circuit potential.

used was based on that of the OCP measurements and was supplemented by 0.033 M AlCl3 in order to buffer the variation of Al3+ cations and to hold the pH stable at 3.2. This value is approximately

Galvanostatic measurements were rst carried out for 24 h on


both the T6 and T76 samples. In keeping with the EXCO test results
[26], only the T6 samples developed clear exfoliation corrosion
morphology (lifted grains). No visible corrosion layers were observed on the T76 sample, which evidenced only a powder-like
corrosion over the entire surface. Fig. 7 shows the potential, recorded over time, for the two samples. The T6 temper shows several uctuations in potential, each of them consisting of a sharp
drop in potential followed by a gradual return to a steady state value. This variation corresponds, more or less, to the time dependent
function: V 0  DV m :1  expt=s where s is a time constant,
equalling approximately 100 s. Note also that the asymptotic value

Fig. 6. SEM images of the early stages of exfoliation corrosion on the T6 specimen obtained during the OCP tests. Blisters generally take the form of protuberances as shown
in the close-ups.

1397

T. Marlaud et al. / Corrosion Science 53 (2011) 13941400

Potential (mV/SCE)

(a) -600

-640
T6

-680
T76
-720
0

12

Time (h)

(b)
Potential (mV/SCE)

-620

3 min

T6

Fig. 9. Potential recorded during the galvanostatic experiments carried out for heat
treatments ranging from peak-aged temper T6 to over-aged temper T76. The details
for the heat treatments are shown in Fig. 1.

-640

-660

4000

T76

8000

8100

8200
8300
Time (s)

8400

8500

Fig. 7. (a) Evolution of the measured potential during the galvanostatic experiments for both the T6 and T76 specimens. Note the large uctuations in potential
for the T6 specimen, as illustrated in close-up (b).

of the potential stays relatively constant throughout the whole


test. For the T76 temper, no such transients were observed. This
transient behaviour was also observed in potentiostatic mode, with
applied potentials corresponding to the quasi-steady state values
achieved during the galvanostatic tests: 620 mV/SCE for T6 and
680 mV/SCE for T76. As expected, the recorded current is similar
in both cases (with a mean value close to 2.5 mA), and well-dened
current transients (1 min of lifetime) were observed only for the
T6 temper (Fig. 8).
From these initial observations, it is tempting to consider these
transients as characteristic of EFC susceptibility. In order to further

3.2

Intensity (mA)

2.8

-620 mV/SCE, T6

2.4
1.5 min

-680 mV/SCE, T76


3.4

3.42
Time (h)

3000

2000

1000

0
10

100
Overaging time (h)

Fig. 10. Impact of the ageing time on the total number of transients greater than
2 mV (threshold) recorded during the 24 h galvanostatic experiment.

investigate this possibility, galvanostatic experiments were carried


out at intermediate ageing intervals between the T6 and T76 tempers, specically after different holding times during the 135 C
heat treatment, which is known to modify the EFC sensitivity of
the alloy. The potential transients recorded during these experiments are reported in Fig. 9. As with the previous experiments, it
was observed that both the average potential and the magnitude
of the transients decrease as ageing time increases. Additionally,
a logarithmic decrease in the total number of transients, as a function of time, is also observed (Fig. 10), i.e. as ageing time increases,
the number of transients and EFC sensitivity both decrease.
3.3. Micrographic observations

3.38

Cumulative transients

-680

3.44

Fig. 8. Transient current recorded during the potentiostatic tests for the T6 and T76
samples.

After the galvanostatic treatment, the specimens were immersed for 24 h in a large volume of pure water before being observed with SEM. Fig. 11 shows the cross sections of the T6 and
T76 specimens after 90 min of galvanostatic testing. The corrosion
morphologies are signicantly different; they are quite localised
for the T6 temper specimen, and more uniform for the T76 temper
specimen. The corrosion is deeper for the T6 specimen: 100 lm
below the surface of the sample, compared to 50 lm for the T76

1398

T. Marlaud et al. / Corrosion Science 53 (2011) 13941400

Fig. 11. (a) FEG-SEM images of a cross section of the T6 sample after 3/2 h of galvanostatic testing, evidencing long inter-granular cracks with non-dissolved precipitates
(indicated by arrows) revealed using the chemical contrast imaging mode; (b) FEG-SEM images of a cross section of the T76 sample after 3/2 h of the same test, evidencing
more diffuse inter and inter-sub-granular corrosion.

(b) 3000

(a)

Mean depth (m)

a)

T6

2400

1800

1200

T76
600

10

20

30
Time (h)

40

50

Fig. 12. (a) Illustration of the corrosion depths measured on the cross sections of both the T6 and T76 samples for statistical analysis of EFC kinetics. (b) Variation of the mean
value as a function of time.

specimen. The long cracks observed in the T6 sample extend in the


rolling direction, along the grain boundaries with non-dissolved
precipitates. In these backscattered electron images, grain boundaries are dark in colour due to the presence of precipitate-free
zones with bright spots caused by grain boundary precipitates.
By contrast, the corrosion morphology of the T76 specimen consists of inter- and inter-sub-granular corrosion paths, which reveals preferential anodic dissolution of the grains (Fig. 11b). No
visible cracks can be observed. Thus, it appears that the over-aged
T76 specimen is sensitive only to inter- and inter-sub-granular corrosion. Lastly, a statistical analysis of the average corrosion depth,
as measured on the cross section of both the T6 sample and T76
sample, was performed and the mean depth was found to be much
less for the T76 specimen than for the T6 specimen (Fig. 12).
3.4. Effect of surface orientation
In order to evaluate the possible effects of lift-out stress caused
by the corrosion products formed during EFC, corrosion tests were
performed under the same conditions along the transverse orientation of the T6 specimen, on the LT-ST plane. In this scenario, grains
would, presumably, be less prone to lift up since the elongated
grains are normal to the grain surface (see Fig. 2b for the geometric
conguration of the sample). The potential transients recorded
(not reported here) still exhibit well-dened transients, even

though there were less events than in the longitudinal tests (L/LT
plane). SEM images of the cross section reveal band-like corrosion
zones with some long inter-granular cracks (200300 lm) that extend far into the unaffected metal (Fig. 13). Moreover, when these
samples were aged under air (50% humidity ratio) after the test,
cracks still propagated in the rolling direction until they crossed
the whole sample after 15 days, as shown in Fig. 14a. Fractographic
investigations (Fig. 14b, and images at a higher magnication in
Fig. 14c and d) show that the damage consists of mostly brittle, inter-granular cracks with no evident corrosion attack.
4. Discussion
Our results show that the appearance of exfoliation corrosion,
either under OCP conditions underneath an articial crevice or under galvanostatic or potentiostatic control, systematically correlates with the presence of well-dened voltage or current
transients. The magnitude of these transients, as well as their
amplitude, is shown to be a major parameter that parallels improved EFC susceptibility when prolonging heat treatment beyond
T6. For the future, this type of measurement opens the door to the
quantitative evaluation of EFC susceptibility for a wide range of alloys and corrosion media. However, it is rst necessary to understand the mechanisms giving rise to these transients and their
relation to exfoliation events.

T. Marlaud et al. / Corrosion Science 53 (2011) 13941400

Fig. 13. SEM images of a cross section (on the L/ST plane) of the T6 sample after a
brief period (6 h) of transverse galvanostatic testing. Note the band-like form of the
selective dissolution along grain boundaries.

The average rate of cracks propagation (Fig. 12) corresponds to


around 12 lm/min, not unlike what is frequently observed in
stress corrosion cracking [34]. Such cracks may propagate extre-

1399

mely fast once grain boundary strength is exceeded, creating sudden contact between the electrolyte and the fresh metallic surface,
promoting in turn the dissolution of the aluminium matrix. Such
an elementary crack advancing step is expected to be at the origin
of the transients, e.g. the intense electrochemical activity observed
during the galvanostatic measurements. The transients can be
regarded as the triggering response of the experimental setup to
a crack occurrence, and the gradual return of the potential to a
steady state value as the result of the formation of aluminium chloride/oxy-chloride complexes [35,36] on the bare surface thereby
limiting the anodic dissolution kinetics.
The question remains as to whether the stresses causing these
cracks are due to the volumic effect of solid corrosion products.
No such corrosion products were observed in the experiments
reported here, thus suggesting the possibility of a hydrogen effect.
In the acidic medium formed in a corrosion zone due to the
hydrolysis of the aluminium cations, Al dissolution is likely
accompanied by hydrogen emissions, depending on the local pH
and potential at the bottom of the corroding zone [37]. Hydrogen
bubbling was, in fact, clearly visible during and after each test.
The molecular recombination of these hydrogen atoms at or near
grain boundaries is assumed to generate enough pressure to induce crack initiation. This scenario would be consistent with the
fact that very long and extremely thin cracks are observed far
from the corrosion location (in locations where it is highly unlikely that uplifting mechanical stresses due to corrosion products
would be have much effect). However, most importantly, it would
be consistent with the observation that similar cracks are
observed when corrosion proceeds along the long direction of
the grains normal to the surface of the sample, in which case no
signicant external stress is exerted on the grain boundaries by
the corrosion products. Lastly, if a hydrogen evolution reaction
is behind the local stress producing the cracks, it is signicant that
the presence of a crevice, known to produce local acidication,
was found to favour EFC initiation in a nearly neutral electrolyte
under open circuit conditions.

Fig. 14. (a) Images of a cross section of the T6 sample after 6 h of transverse galvanostatic testing followed by 30 days under air (50% humidity ratio), showing the extension
of thin cracks throughout the entire sample; (b) shows the fractograph after fracturing the sample in (a); (c) and (d) are magnications of (b).

1400

T. Marlaud et al. / Corrosion Science 53 (2011) 13941400

From a metallurgical viewpoint, increasing the duration of the


ageing treatment beyond the T6 temper modies the size and composition of the precipitates and, consequently, the composition of
the surrounding matrix. This modies in turn the average electrode
potential [38], which may explain the marked decrease of this
potential with the ageing holding time. Moreover, the ageing treatment also modies the chemical composition in the neighbourhood
of the grain boundaries, in the so called precipitate-free zone,
thereby increasing their toughness, resulting in fewer cracks and
electrochemical transients.
5. Conclusions
To summarise, this work shows that:
(i) In quasi-neutral electrolyte and under open circuit conditions, the presence of a crevice favours the development of
exfoliation corrosion in the 7XXX series aluminium alloy in
question. This corrosion is characterised by the presence of
current and potential transients with a spectrum depending
upon the heat treatment of the alloy.
(ii) Theses transients were also observed under galvanostatic
control conditions. Increasing the ageing time beyond the
T6 temper, which is known to drastically decreases the sensitivity to EFC was found to signicantly decrease the number of transients as well as average electrode potential.
(iii) SEM observation performed after the galvanostatic tests
revealed the presence of inter-granular cracks, indicating
that the dominant EFC mechanism is discontinuous and suggesting that the propagation of exfoliation corrosion is
related to a fracture mechanism. The observed transients
are believed to result from the crack propagation. The comparison between longitudinal and transverse corrosion tests
also suggests that corrosion products have no impact on
cracks initiation, whereas there is reason to suspect that
hydrogen recombination plays a decisive role.

Acknowledgements
C. Henon (Alcan), T. Warner (Alcan) and M. Reboul (SIMAP) are
warmly thanked for stimulating discussions.
References
[1] R. Develay, Traitements thermiques des alliages daluminium in Techniques
de lingnieur, M1290, M1291.

[2] M.J. Robinson, N.C. Jackson, Corros. Sci. 41 (1999) 10131028.


[3] J.P. Chubb, T.A. Morad, B.S. Hockenhull, J.W. Bristow, Int. J. Fatigue 17 (1995)
4954.
[4] M. Liao, G. Renaud, N.C. Bellinger, Int. J. Fatigue 29 (2007) 677686.
[5] A. Afseth, J.H. Nordlien, G.M. Scamans, K. Nisancioglu, Corros. Sci. 43 (2001)
20932109.
[6] A. Afseth, J.H. Nordlien, G.M. Scamans, K. Nisancioglu, Corros. Sci. 44 (2002)
145162.
[7] A. Afseth, J.H. Nordlien, G.M. Scamans, K. Nisancioglu, Corros. Sci. 44 (2002)
24912506.
[8] J.T.B. Gundersen, A. Aytac, J.H. Nordlien, K. Nisancoglu, Corros. Sci. 46 (2004)
697714.
[9] H. Huang, G.S. Frankel, Corrosion 63 (2007) 731743.
[10] Qingjiang Meng, G.S. Frankel, J. Electrochem. Soc. 151 (2004) 271283.
[11] X. Zhao, G.S. Frankel, Corros. Sci. 49 (2007) 920938.
[12] D.J. Kelly, M.J. Robinson, Corrosion 49 (1993) 787795.
[13] M.J. Robinson, Corros. Sci. 23 (1983) 887899.
[14] P.V. Petroyiannis, Al.Th. Kermanidis, P. Papanikos, Sp.G. Pantelakis, Theor.
Appl. Fract. Mech. 41 (2004) 173183.
[15] N.D. Alexopoulos, P. Papanikos, Mater. Sci. Eng. A 498 (2008) 248257.
[16] Amjad Saleh El-Amoush, J. Alloys Compd. 443 (2007) 171177.
[17] P.V. Petroyiannis, A.T. Kermanidis, P. Papanikos, S.G. Pantelakis, Theor. Appl.
Fract. Mech. 41 (2004) 173183.
[18] H. Kamousti, G.N. Haidemenopoulos, V. Bontozoglou, V. Pantelakis, Corros. Sci.
48 (2006) 12091224.
[19] ASTM-G34 (1974). Standard Test Method for Exfoliation Corrosion
Susceptibility in 2XXX and 7XXX Series Aluminum Alloys.
[20] ASTM-G66-99 (2005). Standard Test Method for Visual Assessment of
Exfoliation Corrosion Susceptibility of 5XXX Series Aluminum Alloys (ASSET
Test).
[21] ASTM-G85 (2002). Standard Practice for Modied Salt Spray (Fog) Testing.
[22] B.W. Lifka, D.O. Sprowls, Corrosion 22 (1966) 715.
[23] D.O. Sprowls, J.D. Walshal, Simplied exfoliation testing of aluminium alloys,
localized corrosion-cause of metal failure, ASTM STP 516 (1972).
[24] E.A.G. Liddiard, J.A. Whittaker, J. Inst. Met. 89 (1960) 377384.
[25] M.C. Reboul, J. Bouvaist, Werstoffe und Korrosion 30 (1979) 700712.
[26] T. Marlaud, B. Malki, A. Deschamps, M. Reboul, B. Baroux, ECS Trans. 3 (31)
(2007) 285293.
[27] C.J. Baes, R.E. Mesmer, The Hydrolysis of Cations, second ed., R.E. Krieger
Publishing Company, 1976.
[28] A. Turnbull, Chemistry within localized corrosion cavities, Proceedings of the
Second International Conference on Localized Corrosion 9 (1990) 359373.
[29] S. Lee, B.W. Lifka, ASTM STP 1134, Am. Soc. Test. Mater. (1992) 119.
[30] K.R. Cooper, L.M. Young, R.P. Gangloff, R.G. Kelly, Mat. Sci. F. 331 (2000) 1625
1634.
[31] K. Sasaki, H.S. Isaacs, J. Electrochem. Soc. 151 (2004) 124133.
[32] S. Horl, B. Malki, B. Baroux, J. Electrochem. Soc. 153 (2006) 527532.
[33] H.S. Isaacs, C. Scheffey, R. Huang, ECS Trans. 11 (22) (2008) 112.
[34] W. Dietzel and K.-H. Schwalbe, Application of the rising displacement test to
Scc investigations, slow strain rate testing for the evaluation of
environmentally induced cracking, in: R.D. Kane (Ed.), Cracking: Research
and Engineering Applications, ASTM STP 1210, Philadelphia, 1993, pp. 134
148.
[35] G.C. Wood, W.H. Sutton, J.A. Richardson, T.N.K. Riley, A.G. Malherbe,
Corrosion 3 (1974) 526546.
[36] M.C. Reboul, T.J. Warner, H. Mayet, B. Baroux, Corros. Rev. 15 (1997) 471496.
[37] M. Boinet, J. Bernard, M. Chatenet, F. Dalard, S. Maximovitch, Electrochim. Acta
55 (2010) 34543463.
[38] E.H. Hollingsworth, H.Y. Hunsicker, Mater. Handb. 9th ed. 1987.

You might also like