Anda di halaman 1dari 11

Understanding damage accumulation upon AA7075-T651 used in airframes from a

microstructural point of view


N. Birbilis
1
, M.K. Cavanaugh
1
, R.G. Buchheit
1
, D.G. Harlow
2
, R.P. Wei
2
1
Dept. of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, USA
2
Dept. of Mechanical Engineering and Mechanics, Lehigh University, Bethlehem, PA 18015, USA
Keywords
Localized corrosion, pitting, TEM, SEM, aluminum alloys, intermetallics particles, micro-electrochemistry
Abstract
Corrosion and corrosion fatigue of aircraft components remains a critical issue for the service readiness of aging aircraft. The
favorable mechanical properties of AA7075-T651 arise as a result of a heterogeneous microstructure that compromises the
corrosion resistance of the alloy, ultimately resulting in localized corrosion. Understanding the factors that govern extent, rate
and morphology of damage accumulation is essential in developing models to predict corrosion and corrosion fatigue
resistance. The electrochemical properties of AA7075-T651 have recently been investigated on a phase-by-phase basis. As a
result, quantitative microstructural characterization of AA7075-T651 takes on a critical importance in the understanding of
corrosion damage accumulation, based on localized chemistry (and hence electrochemistry) of the alloy.
This work presents results of a study using analytical microscopy (SEM, TEM, STEM, EDXS, EBKP) to reveal what phases,
intermetallics and particles constitute AA7075-T651; allowing characterization and discrimination on a chemical basis.
Knowledge of the type of localized chemistry that exists may be coupled with electrochemical results and used to typify
damage accumulation processes on the nano and microscale.
Introduction
Aluminum alloy 7075-T651 is commonly found in several key components of aircraft, however, the large population of
intermetallic particles within this alloy compromises corrosion resistance, ultimately resulting in localized corrosion [1-2].
Understanding the factors which govern the extent of corrosion damage accumulation is essential, since localized corrosion is
a precursor to corrosion fatigue crack initiation [3-4].
The electrochemical properties of a range of intermetallics common to high strength aluminum (Al) alloys (including
AA7075) have each recently been investigated on an individual basis [5]. Such data may serve as critical input to models
designed to predict corrosion, and also to provide a detailed physical insight to the electrochemistry of the alloy. It was found
that the electrochemical characteristics of these intermetallics vary significantly in terms of corrosion potential, breakdown
potential (if applicable) and the ability to sustain oxygen reduction or dissolution (self corrosion) reactions [5].
As a result, quantitative microstructural characterization of AA7075-T651 takes on significant importance for interpreting the
electrochemical response of the alloy. Knowledge of the type of intermetallic-matrix interaction expected and the extent of
resultant charge transfer that occurs can be used to typify damage accumulation. This knowledge may be further exploited to
develop predictive models for damage accumulation and subsequent integration into a fracture mechanics approach for
investigating corrosion fatigue via use of the appropriate analytical expressions [6].
The microstructures developed in commercial high strength aluminum alloys such as AA7075-T651 are complex and
incorporate a combination of equilibrium and non-equilibrium phases [7-8]. Typically, AA7075 has a chemical composition
incorporating up to ten alloying elements. Such elements include primarily Zn, Mg and Cu, however appreciable and specific
amounts of Fe, Si, Cr, Ti, Zr and Mn are often present (both as deliberate additions and as impurities). Over the past three
decades, the microstructure of AA7075 has been well characterized [8-14], along with the corresponding physical metallurgy
[7]. The literature quotes evidence supporting the presence of the following intermetallics in 7xxx series alloys (not all are
simultaneously present, and temper and precise composition will regulate the types and proportion): MgZn
2
, Mg
2
Si,
Al
20
Cu
2
Mn
3
, Al
12
Mn
3
Si, Al
7
Cu
2
Fe, Al
2
Cu, Al
2
CuMg, Al
3
Fe, Al
12
Mg
2
Cr, Al
20
Cu
2
Mn
3
, Al
6
Mn, Al
3
Ti, Al
6
Zr, Mg
2
Al
3
, Al
32
Zn
49
, and
Mg(AlCu) [8-9, 12-16]. The role of these intermetallics with respect to mechanical properties is beyond the scope of this
paper, since we are concerned with primarily identifying particle types and distributions for the purpose of ascribing
electrochemical parameters. Therefore, those intermetallics which may be vital in the development of mechanical properties
may not necessarily impact the corrosion response of the alloy, and vice-versa.
As noted above, the composition of commercial AA7075-T651 stimulates a large number of varied intermetallic particles.
Such particles may be precipitated phases, largely based upon the principal alloying elements (Zn and Mg). When they are
homogeneously dispersed, their effect on localized corrosion behavior has been difficult to discern. However, when they are
concentrated on grain boundaries, they may affect intergranular corrosion and stress corrosion cracking susceptibility [11, 17-
18]. Typical precipitated phases include primarily MgZn
2
(), however under certain conditions small amounts of Al
2
CuMg
(S) or Al
2
Cu () have been claimed to form [15].
Constituent particles primarily evolve from the presence of Fe, Mn and Si. They are comparatively large and irregularly
shaped, and can be formed during alloy solidification whilst not being appreciably dissolved during subsequent thermo-
mechanical processing. Rolling and extrusion tends to break-up and align constituent particles into bands within the alloy.
Often constituents are found in colonies made up of several intermetallic crystals or several different compound types.
Because these particles are rich in alloying elements, their electrochemical behavior is often significantly different than the
surrounding matrix phase [5]. In high strength Al alloys, pitting is nominally associated with some fraction of the constituent
particles present in the alloy [19-22]. Typical examples include Al
3
Fe and Al
7
Cu
2
Fe. Such particles do not represent a
significant aspect in the development of mechanical properties in high strength Al alloys, and, as a result, the size distribution
and density of such particles has not received specific attention in the literature in any great detail.
In regards to localized corrosion, the intermetallics of particular interest are those which appear in the greatest proportion, by
size or by frequency; since ultimately along with their corresponding electrochemical characteristics, will dominate the
corrosion response of the alloy.
Experimental
Commercial AA7075-T651 supplied by Alcoa was used in this study. Specimens were prepared from about 10mm below the
surface of a 76mm thick rolled plate. The face / plane incorporating the longitudinal direction (often called the LT face) were
studied here. This face represents boldly exposed surfaces such as wing skins and was chosen to correlate with a parallel
study focused on corrosion damage metrology, for which the microstructure-electrochemistry-damage relationship will
ultimately form the development of an empirically based model.
SEM, EDXS and collection of Electron Backscattered Diffraction Patterns (EBDP) was carried out using a Philips XL-30
FEG-ESEM. The specimens used were nominally 101010mm in size and metallographically mounted in conductive
bakelite. Surface preparation included successive polishing down to a 0.05m surface finish employing a vibratory polisher
and a gamma alumina suspension. TEM was carried out using a Philips CM-200 (conventional) and a Philips TF-20
(analytical), both operating at 200kV. Discs of 3mm diameter were punched from alloy sections carefully ground to a
thickness of ~150m, and twin-jet electropolished into thin foils using a 25% Nitric acid 75% Methanol electrolyte at -35C.
Results and Discussion
Identification of Intermetallics
The vast number and corresponding size distribution of intermetallics in AA7075-T651 demands techniques capable of
quantitative examination across this size distribution. As a result, both SEM and TEM based methods were employed in this
work. The ability to assess chemical composition along with structural information was considered of principle importance in
the development of accurate intermetallic identification, since positive identification of a phase is not readily possible by
EDXS alone. However, the crystal structure of most intermetallics in aluminum alloys is unique and previously defined [7-8].
Figure 1 General microstructure of AA7075-T651 when viewed with SEM (backscattered electron mode)
The general microstructure of AA7075-T651 (via SEM analysis) may be seen in Fig. 1. The image shows (constituent)
particles aligned in the direction of working. Owing to the use of imaging via backscattered electrons, we see light particles
along with darker particles, when compared to the matrix. Such particles are generally >1m in diameter, making them well
suited to spot EDXS and EBDP analysis.
An example of such analysis is seen in Fig. 2, whereby an individual (light) particle has been isolated. The slight loss of
material surrounding the particle is attributed to artifacts produced via the vibratory-polishing process required to generate a
surface flat enough for EBDP analysis.
Fig. 2 shows the EDXS spectrum and the EBDP collected for the particle, along with the EBDP for grains either side of the
particle. Based upon the chemical analyses of the EDXS, we see that the particle is comprised of Al-Cu-Fe, in the proportion
typical of Al
7
Cu
2
Fe constituent particles. From the EBDP analysis, we can confirm that the structure of the particle is
tetragonal, consistent with the theoretical structure of Al
7
Cu
2
Fe; tetragonal (a = 6.336A, c = 14.87A) space group P4/mnc.
Figure 2. SEM image showing Al
7
Cu
2
Fe particle in 7075-T651 matrix. The EDXS spectrum and associated EDBP of
the particle are shown, together with the EBDP of two grains either side of the Al
7
Cu
2
Fe particle (SEM image is tilted
and therefore particle appears slightly elongated in vertical axis).
Similarly, such an EDXS/EBDP analysis may be carried out on a dark particle, as seen in Fig. 3. In the lower of the SEM
images, corresponding to a backscattered electron image, the particle appears to be a black hole-like feature, however in the
secondary electron image above, it is seen that the particle is indeed solid and well defined. The EDXS/EBDP analysis
reveals that this particle is Mg
2
Si.
Figure 3. SEM images showing Mg
2
Si particle in 7075-T651 matrix, Secondary Electron Image (above),
Backscattered Electron Image (below). The EDXS spectrum and associated EDBP of the particle are shown, together
with the EBDP of two grains in the vicinity of the particle (SEM image is tilted and therefore particle appears slightly
elongated in vertical axis).
By repeating the characterization process shown in Figs. 2 and 3 over a number of particles, it was possible to identify a
sufficient sample, along with quantification of their physical characteristics such as size, shape, location and number density.
In contrast to SEM analysis, TEM analysis does not reveal any significant presence of constituent particles. Constituent
particles are nominally large and either destroyed by the TEM sample preparation process (grinding / polishing), or simply
fall out of the TEM foil in the thin regions under investigation.
The general microstructure of AA7075-T651 may be seen in the montage of scanning TEM (STEM) images seen in Fig. 4,
where a grain boundary triple point is seen towards the right of the montage and the grain boundary and neighboring grains
are seen to the left.
500nm

Figure 4. Montage of (dark field) STEM images showing the general structure of AA7075-T651
On the submicron level, the alloy reveals a dense population of (strengthening) particles, namely MgZn
2
(). These particles
appear as light particles in Fig. 4, the images of which were collected using a high angle annular dark field detector.
Identification of phase has been covered in [23], whereby diffraction spots arising from such particles are visible when
orienting close to the <110>

zone axis. In this work, we will not focus on distinguishing between (which is known to
populate the grain boundary) and the transition phase (which is known to populate within the grain). Furthermore, for the
purposes of corrosion, it is posited that the electrochemical response of both forms of the precipitate are similar, since the
difference between and is a slight difference in lattice parameters whilst the chemistry is nominally identical [7].
Fig. 5 shows a conventional dark field TEM image showing the grain boundary and associated phase in more detail.
Figure 5. Dark field TEM image of the microstructure of AA7075-T651 showing grain boundary precipitates and
matrix precipitation.
Intermetallic distribution
The results of (2-d) stereological investigations are summarized in Table 1, where d is the mean particle equivalent diameter
and N
A
is the particle number density. The equivalent diameter is used since intermetallic particles do not have a nominal
aspect ratio of 1. These values are presented together with the Nearest Neighbor Distance and discussed further below.
Table I. Summary of results
Particle
d
(m)
Mean N
A
(#/mm
2
)
A
A
(m
2
)
Mean
Aspect
Ratio
. .Neighb N dist
(m)
Standard deviation in
parentheses
MgZn
2
0.052 3.5 x 10
7
0.0017 1.64 0.1 (0.05)
Al
7
Cu
2
Fe 1.7 1651 3.6 1.55 11.2 (8.6)
Al
3
Fe 1.69 571 4.1 1.37 17.8 (14.0)
Mg
2
Si 2.13 856 4.6 1.56 15.5 (11.1)
Al
2
CuMg 5.1 101 11 1.91 43.9 (30.5)
It should be pointed out that intermetallics not seen in sufficient quantities to generate a statistical analysis include Al
2
Cu,
Al
6
Mn and Al
20
Cu
2
Mn
3
; such particles were however identified in TEM analysis at very low N
A
. Other particles such as the
dispersoids Al
3
Ti, Al
3
Zr, Al
12
Mn
3
Si and Al
12
Mg
2
Cr were not able to be characterized or easily discriminated in the alloy, and
are thus likely to appear at low N
A
and small d. A total of <1% of particles were indeterminate.
We note at this point, that unlike in 2xxx series alloys, Al
2
CuMg (S-phase) is not a precipitated phase in the form observed
here. In this case, S is present as a constituent particle that likely formed during alloy processing and at elevated temperatures
(above those used for age hardening). This is apparent by the low N
A
and relatively large d values for S phase.
We also note that in some cases, Al
7
Cu
2
Fe contained small amounts of manganese which substituted for iron, leading to the
Fe + Mn content forming 10wt% of the intermetallic. Mn was nominally present in such cases below a few weight percent,
and at times not present at all. Therefore, the intermetallic is referred to as Al
7
Cu
2
Fe, since it maintains the same crystal
structure all times it was tested.
If we plot the N
A
vs. particle type (Fig. 6), we observe that phase is present in the greatest population by almost four orders
of magnitude; followed by the constituent particles Al
7
Cu
2
Fe, Mg
2
Si and Al
3
Fe.
Figure 6. Number density for intermetallic particles in AA7075-T651
Conversely, if we plot d vs. particle type (Fig. 7), we observe that has the smallest d by nearly two orders of magnitude,
followed by Al
7
Cu
2
Fe, Mg
2
Si and Al
3
Fe.
Figure 7. Mean equivalent diameters for intermetallic particles in AA7075-T651
The results above correspond well with those of Poulose and co-workers [11] who carried out investigations exclusively on
phase. Further to the information in Fig. 7, the ability to generate particle size distributions for each individual particle type
allows for a more comprehensive insight into the specific particle characteristics. For example, if all particles were classed
together, then the size range of phase would dominate the distribution, since it is present in the greatest abundance. The
particle size distributions, as given according to d, the equivalent diameter of each particle, are seen in Fig. 8.
Figure 8. Cumulative distribution for the intermetallic particle size in AA7075-T651
Fig. 8 displays the size range over which these particles exist. The separation between the precipitated phases () and
constituent particles is clear. Furthermore, we observe particle sizes may approach the order of 10m.
Alloy electrochemistry
The overall electrochemistry of the alloy will be dependent on the relative proportions of the intermetallic particles, along
with their individual electrochemical properties, and the ensuing electrochemical interaction with the matrix. A useful way of
determining the relative proportions of the main intermetallic particles may be seen in Table II, whereby for a given region of
material, the area of (specific) intermetallics that populate that region is seen. This information is coupled with
electrochemical parameters as seen further below. Table II shows the electrochemical properties of the intermetallics as
previously measured in 0.1M NaCl [5].
Table II. Summary of electrochemical results collected in 0.1M NaCl
Particle
d
(m)
N
A
(#/mm
2
)
Area
occupied per
1mm
2
of
alloy (mm
2
)
E
corr
(V
SCE
)
I (current) at
E
corr
(AA7075-T651)
(A/cm
2
)
MgZn
2
0.052 3.5 x 10
7
0.09 -1.03 1.0 x 10
-3

Al
7
Cu
2
Fe 1.7 1651 0.009 -0.55 -3.1 x 10
-4

Al
3
Fe 1.69 571 0.005 -0.54 -9.9 x 10
-5

Mg
2
Si 2.13 856 0.001 -1.54 1.9 x 10
-4

Al
2
CuMg 5.1 201 0.0002 -0.88 -2.1 x 10
-6

We observe that the corrosion potential values (E
corr
) for the various intermetallics vary over a wide range (-1.54 to -0.54 V
with respect to a saturated calomel electrode, SCE). The typical E
corr
of the bulk alloy (AA7075-T651) is nominally in the
vicinity of -0.9 to -0.75V
SCE
in NaCl solutions. Therefore at the stationary E
corr
of the alloy, certain intermetallics assume the
role of being anodic (viz. less noble, with E
corr
< E
corr(AA7075-T651)
), whilst others are cathodic (E
corr
> E
corr(AA7075-T651)
). A better
sense of this is seen by the I (current) values given in Table 2. These values were determined on the premise that if we
assume the E
corr
of AA7075-T651 to be -0.8V
SCE
in 0.1M NaCl (which is based on [24]), then the current density sustained by
each intermetallic type (at a potential of -0.8V
SCE
) may be evaluated. The current values were taken from polarization curves
determined in our laboratory and published elsewhere [5, 25].
Further to determining the current each intermetallic type can support for a given potential, we can combine this with the
quantitative microscopy and determine the current sustained by each intermetallic type per unit area of alloy surface (based
on charge transfer alone). The results of this determination may be seen in Fig. 9.

Figure 9. Current density ascribed to each intermetallic type in AA7075-T651 at E
corr
(AA7075-T651) in 0.1M NaCl
Fig. 9 shows that sustains a large anodic current density, which is consistent with self-dissolution of particles that may
intersect the alloy surface and come into contact with electrolyte. These particles are expected to be dissolved, and, as a
result, not have any further discernable influence on the corrosion of the alloy [24]. Similarly, Mg
2
Si can behave in a similar
manner; however, in the case of Mg
2
Si, significantly larger cavities are expected to be observed following particle
dissolution (since d
Mg
2
Si
>> d

)
This is seen in Fig. 10, whereby following 10 h immersion in 0.1M NaCl, we see corrosion damage in the form of cavities,
associated with Mg
2
Si particles. We note that damage related to is too fine to generate obvious corrosion damage (via SEM
analysis).
Conversely, intermetallics which sustain a cathodic current appear to be surrounded by peripheral type pitting. In such
cases, the damage evolution is aligned with the morphology in which the constituent particles are found. The damage
observed takes the form of trenches which surround the intermetallic, and represent corrosion of the matrix.
(cathodic particles)
'peripheral' pitting
'terminal' pitting
(anodic particles)
Figure 10. Corrosion damage upon AA7075-T651 following 10 h immersion in 0.1M NaCl.
From Fig. 10, we also see that the sphere of influence from cathodic particles allows for connection of corrosion damage
that arises from neighboring particles. This leads to pits larger than those which may be expected from a single particle,
emphasizing the need for detailed 2-d stereology (and ultimately 3-d) information prior to accurately predicting corrosion
damage evolution.
At this point, we note that the analysis we offer is based upon alloy microstructure and charge transfer considerations alone.
We have specifically neglected the effect of local pH in solution and corrosive solutions other than NaCl; this is done in an
attempt to provide a fundamental framework for which more sophisticated analyses leading to modeling can be carried out
(as discussed further below)
General Discussion
From the results presented herein, we see that constituent particles form a significant area fraction of the alloy. These
particles possess comparatively large d values and significantly contribute to the early evolution of localized corrosion of
AA7075-T651.
The results have allowed for a comprehensive insight into the microstructure-corrosion relationship for AA7075-651. The
microstructural analysis is applicable broadly in understanding commercial AA7075-T651, whilst the electrochemical
analysis of which is useful to those working in the corrosion domain.
The utility of this data towards the development of a mechanistic based model for corrosion prediction of AA7075-T651 is
apparent.
Based on what we have (quantitatively) determined as being present in the alloy, we can present a phenomenological model
that portrays the microstructure and corrosion morphology with immersion time (viz. the corrosion development) along with
the corresponding E
corr
measured in 0.1M NaCl solution. The development of corrosion damage accumulation upon AA7075-
T651 may be phenomenologically given by the representation seen in Fig. 11.
Figure 11. Phenomenological model of the early development of corrosion damage accumulation upon AA7075-T651.
A theoretical / computational model for the corrosion damage evolution of AA7075-T651 based upon physical results
presented herein will form the basis of future work.
Summary
Quantitative microscopy of AA7075-T651 revealed that, with the exception of phase, constituent particles form a
significant area fraction of this commercial alloy. The intermetallics present in the alloy were identified by type and
by physical characteristics.
Intermetallics in AA7075-T651 nominally have markedly different electrochemical properties with respect to one
another and the matrix.
Based upon stereology and electrochemical considerations, , Mg
2
Si and Al
7
Cu
2
Fe form the watch list for
corrosion of AA7075-T651. Mg
2
Si and Al
7
Cu
2
Fe, however, owing to their respective d values, are capable of
influencing rapid early corrosion of the alloy (viz. constituent particle induced corrosion).
Al
7
Cu
2
Fe presents a particularly dangerous situation with respect to corrosion, since Al
7
Cu
2
Fe particles can generate
damage beyond their dimensions with overlapping spheres of influence, which leads to larger pits. Furthermore,
the damage ascribed to such particles is not necessarily terminal, leaving behind finite cavities such as those
expected from Mg
2
Si. Instead, they lead to matrix corrosion until they may be undercut by the corrosion becoming
so severe around the particle, by which stage the corrosion damage may have encapsulated a neighboring particle (or
in the extreme case, grown to a size capable of acting as a fatigue initiation site).
References
[1]. R.G. Buchheit, A Compilation of Corrosion Potentials Reported for Intermetallic Phases in Aluminum Alloys J.
Electrochem. Soc., 142 (1995), 3994-3996.
[2]. R.G. Buchheit, R.K. Boger, M.C. Carroll, R.M. Leard, C. Paglia, J.L. Searles, The Electrochemistry of
Intermetallic Particles and Localized Corrosion in Al Alloys, JOM, 53 (2001), 29-36.
[3]. D. G. Harlow, Applications of the Frechet distribution function, Int. J of Materials & Product Technology, 17
(2002) 482-495.
[4]. A.D.B. Gingell and J.E. King, The Effect of frequency and Microstructure on Corrosion Fatigue Crack
Propagation in High Strength Aluminum Alloys, Acta Materiala, 9 (1997), 3855-3870.
[5]. N. Birbilis and R.G. Buchheit, Electrochemical Characteristics of Intermetallic Phases in Aluminum Alloys: An
Experimental Survey and Discussion. J. Electrochem. Soc., 152 (2005), B140-151.
[6]. G.S. Chen, K.-C. Wan, M. Gao, R.P. Wei and T.H. Flourney, Transition from Pitting to Fatigue Crack Growth
Modeling of Corrosion Fatigue Crack Nucleation in a 2024-T3 Aluminum Alloy Mat. Sci. Eng. A, 219 (1996),
126-132.
[7]. L.F. Mondolfo, The aluminum-magnesium-zinc alloys (Rome, NY: Research and Development Center, Revere
Copper and Brass Inc., 1967).
[8]. I.J. Polmear, Light alloys, 3
rd
Edn. (London, UK: Arnold, 1995).
[9]. J.E. Hatch, (ed.), Aluminum: Properties and physical metallurgy (Metals Park, OH: ASM, 1984).
[10]. J.K. Park and A.J. Ardell, Effect of Retrogression and Reaging Treatments on the Microstructure of Al-7075-
T651, Metallurgical Transactions A, 15A (1984), 1531-1543.
[11]. P.K. Poulose, J.E. Morral and J. McEvily, Stress Corrosion Crack Velocity and Grain Boundary Precipitates in
an Al-Zn-Mg Alloy, Metallurgical Transactions, 5 (1974), 1393-1400.
[12]. J.A. Wert, Identification of Precipitates in 7075 Al After High-Temperature Aging, Scripta Metallurgica, 15
(1981), 445-447.
[13]. S.P. Ringer and K. Raviprasad, Developments in Age-Hardenable Aluminium Alloys and Rational Design of
Microstructure, Materials Forum, 24 (2000), 59-94.
[14]. M. Gao, C.R. Feng and R.P. Wei, An Analytical Electron Microscopy Study of Constituent Particles in
Commercial 7075-T6 and 2024-T3 Alloys, Met. Mat. Trans. A, 29A (1998), 1145-1151.
[15]. Q. Meng and G.S. Frankel, Effect of Cu Content on Corrosion Behavior of 7xxx Series Aluminum Alloys, J.
Electrochem. Soc., 151 (2004), B271-283.
[16]. S.K. Maloney, K. Hono, I.J. Polmear and S.P. Ringer, The Chemistry of precipitates in an Aged Al-2.1Zn-1.7Mg
at.% Alloy, Scripta Materialia, 41 (1999), 1031-1038.
[17]. S. Ohsaki, Stress Corrosion Cracking of Aluminum Alloys, Keikinzoku, 46 (1996), 456-466.
[18]. J.R. Davis, ed., Corrosion of aluminum and aluminum alloys (Materials Park, OH: ASM, 1999).
[19]. Z. Szklarska-Smialowska, Pitting Corrosion of Aluminum, Corros. Sci., 41 (1999), 1743.
[20]. R.P. Wei, C.-M. Liao and M.Gao, A Transmission Electron Microscopy Study of Constituent-Particle-Induced
Corrosion in 7075-T5 and 2024-T3 Aluminum Alloys. Mat. Trans. A, 29A (1998), 1153-1160.
[21]. G.O. Ilevbare, O. Schneider, R.G. Kelly and J.R. Scully, In Situ Confocal Laser Scanning Microscopy of AA
2024-T3 Corrosion Metrology: I. Localized Corrosion of Particles, J. Electrochem. Soc., 151 (2004), B453-464.
[22]. F. Andreatta, M.M. Lohrengel, H. Terryn and J.H.W. de Wit, Electrochemical characterisation of Aluminium
AA7075-T6 and Solution Heat Treated AA7075 Using a Micro-Capillary Cell, Electrochim. Acta, 48 (2003),
3239-3247.
[23]. S.K. Maloney, I.J. Polmear and S.P. Ringer, Effects of Cu on Precipitation in Al-Zn-Mg Alloys, Materials
Science Forum, 331-337 (2000), 1055-1060.
[24]. N. Birbilis and R.G. Buchheit, Corrosion Damage Accumulation on High Strength Al-Alloys: Some Advances in
Understanding the Role of Intermetallics, Corrosion and Materials, 29 (6) (2004), S4-8.
[25]. N. Birbilis and R.G. Buchheit, In preparation (2005).

Anda mungkin juga menyukai