Anda di halaman 1dari 10

Engineering Fracture Mechanics 76 (2009) 641650

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Evaluation of a simple microstructural-electrochemical model


for corrosion damage accumulation in microstructurally complex
aluminum alloys
M.K. Cavanaugh, R.G. Buchheit *, N. Birbilis
Fontana Corrosion Center, Department of Materials Science and Engineering, Ohio State University, Columbus, OH 43210, United States

a r t i c l e

i n f o

Article history:
Received 17 September 2007
Received in revised form 11 November 2008
Accepted 12 November 2008
Available online 20 November 2008
Keywords:
Pitting corrosion
Corrosion fatigue
Life prediction
7075 Aluminum
Prolometry

a b s t r a c t
Stereological and electrochemical data collected in 0.1 M NaCl is presented for the intermetallic particles found in AA7075-T651. These values are used as inputs for a mechanistic-based corrosion model which yields a distribution of pit radii (assuming hemispherical
pitting) as function of immersion time. With an optical prolometer (OP), a distribution of
damage depths were collected from samples exposed to 0.1 M NaCl solution for times
ranging from 10 to 1344 h. Damage depths were compared with the model predictions
of pit depth. The model is able to predict observations of corrosion damage depths to
within a factor of two, but tends to under-predict damage depth at short times and
over-predict at long times. Under-prediction may be associated with early rapid dissolution of active particles. Over-prediction by the model at longer exposure times may be
associated with the fact that repassivation is not included as a part of this simple model.
2008 Elsevier Ltd. All rights reserved.

1. Introduction
Identifying and repairing corrosion damage to aging aircraft is a signicant part of the maintenance cost for the United
States Air Force [1]. Pitting corrosion is a form of attack that leads to the nucleation of fatigue cracks under cyclic loading [2].
Both eld returns and laboratory testing show the strong tendency for fatigue crack initiation at pits, and a detrimental effect
on component fatigue life [3].
A two-sided criterion has been developed to characterize the transition from corrosion pit to fatigue crack. According to
this criterion, a fatigue crack initiates from a pit when the pit reaches a size at which the stress intensity factor attains the
threshold for fatigue cracking (DKth) and when the rate of fatigue crack growth exceeds that of pit growth [4,5]. A simple
expression revealing the relationship between the critical pit size and the stress concentration can be given by [6]

ap-c p

2
DK th
2:2K t Dr

where ap-c is the size of the pit at the pit-to-crack transition, Kt is the stress concentration factor and Dr is the remote stress
variation.
As noted by Sankaran [7], aerospace structures nominally experience corrosion between ights and fatigue loading during
ight, suggesting the notion of prior corrosion and the role of pits as fatigue crack initiation sites. A recent study for 7075-T6
suggested that pits in the 2975 lm size range were capable of initiating fatigue cracks under a nominal maximum tensile

* Corresponding author.
E-mail address: buchheit.8@osu.edu (R.G. Buchheit).
0013-7944/$ - see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfracmech.2008.11.003

642

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

stress of 30 ksi (207 MPa) [3]. Unfortunately, detection of pits of this size is currently a signicant challenge for non-destructive inspection methods.
Mechanism-based models for life-cycle analysis are being developed to help address this inspection gap and to address
other aspects of life-cycle management [8]. Such models include the development of time-dependent corrosion damage accumulation functions that incorporate key material and environmental factors, along with their variability. Recent work within
our laboratory has established, with a degree of accuracy, the type of intermetallics that exist in AA7075-T651, together with
their physical characteristics [9]. Such intermetallics are strongly associated with pitting. Adding to this, the individual electrochemical properties of these intermetallics has also been characterized [10], indicating very clearly the heterogeneity in
electrochemical behavior that accompanies the heterogeneity in alloy microstructure. Specically, it has been found that
the electrochemical characteristics of these intermetallics vary signicantly in terms of corrosion potential and their ability
to sustain oxygen reduction or dissolution (self-corrosion) reactions. These data can serve as critical input to corrosion damage accumulation functions, promoting progress towards mechanism-based predictive damage accumulation models.
In this study, an example is presented of a simple approach, based on experiment and mechanism, for predicting the timedependent pit size distribution in AA7075-T651 in the size range that has been identied as critical for fatigue crack growth. The
predicted values were then compared with a quantitative characterization of pitting corrosion obtained by 3-D optical prolometry [11] and image reconstruction on alloy samples exposed to an aggressive chloride solution. Examples of other important aspects of corrosion damage accumulation not captured in this modeling approach are also identied and discussed.
2. Experimental procedures
Materials characterization was carried out using both transmission electron microscopy (TEM) (FEI TF-20) and scanning
electron microscopy (SEM), including backscattered electron diffraction analysis (Philips XL-30). Further information on the
materials characterization for AA7075-T651 (supplied by Alcoa) have been given fully in [9]. Details regarding electrochemical characterization of intermetallics on a phase-by-phase basis is given in [10]. Such testing was carried out using the electrochemical microcell method in NaCl solutions.
Optical prolometry was performed on AA7075-T651 blocks (62  25  10 mm), which were carefully polished to a 1 lm
nish with diamond paste (Buehler Metadi). The samples were boldly exposed (on the long-transverse face) in aerated 0.1 M
NaCl (pH  6.5) solution and removed after immersion times of 10 h, 24 h, 48 h, 96 h, 168 h, 672 h, and 1344 h. Following
removal, samples were lightly pickled in concentrated nitric acid under ultrasonic agitation to remove corrosion product.
An optical prolometer (OP) (Veeco Wyko) was used to characterize the sample surfaces. For purposes of collecting pit depth
data, the vertical scanning interferometry mode was used. Localized corrosion up to 2 mm in depth was measurable using
this technique. Two different types of data sets were collected from corroded samples. The rst type of data set was corrosion
depth measured with single pixel resolution across the corroded surface. These data are referred to as damage depth data
sets, and were used in comparisons to model predictions. The second type of data set were those where post processing of
the damage depth data was carried out using manual thresholding (Fovea Pro) to determine pit diameters, pit density, and
percent area corroded.
3. Results and discussion
3.1. Microstructural analysis
Corrosion pits that may potentially nucleate cracks arise largely as a result of constituent particle-induced pitting [12]. As
a result, it is critical to understand the type of particles that exist in AA7075-T651, along with their geometric characteristics.
Fig. 1 indicates the number density of particles in AA7075-T651 for MgZn2, Al7Cu2Fe, Al3Fe, Mg2Si, and Al2CuMg. A combination of X-ray microchemical and backscatter electron diffraction analyses were used to discriminate among chemical
types. From Fig. 1, it is apparent that g phase (MgZn2) exists in a much larger density than the other particles. MgZn2 is precipitated from solid solution during thermomechanical processing and articial aging. These particles are typically submicron in diameter but can coarsen to sizes large enough to be resolved by SEM. The other particles listed are formed
during solidication and are broken up and redistributed during thermomechanical processing. The distinction between
MgZn2 and the other particle chemical types is also seen in Fig. 2, which shows equivalent diameter of the particles by type.
There is about a two order of magnitude difference in size between MgZn2 and the other particles, the latter largely existing
in a size range of 110 lm in diameter.
Fig. 3 presents the nearest neighbor distances among all particle types, excluding MgZn2 particles. The data in Fig. 3 represents the constituent particles and the distribution of distances to any other constituent particle. On average, particles tend
to be clustered such that the 50th percentile for nearest neighbor distance is within 20 lm. At this proximity, the opportunity for chemical interaction among particles during localized corrosion is likely to exist.
3.2. Electrochemical analysis
A microelectrochemical cell was used to measure the anodic and cathodic polarization behavior of particle phases found
in this alloy [10]. Polarization responses have been collected in aqueous solutions over a range of chloride ion concentrations,

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

643

Fig. 1. Number density of particles, by type, that populates AA7075-T6. Rem refers to remainder that exist in low number density (e.g., Al6Mn, Al20Cu2Mn3,
Al2Cu).

pH and temperature relevant to corrosion in aircraft airframes, however in this work, only the electrochemical response in
0.1 M NaCl at neutral pH and 20 C has been considered.
In this environment, the particles that populate AA7075-T651 display vastly different electrochemical properties, giving
rise to localized corrosion susceptibility. At the corrosion potential of the alloy, some particles (depending on their chemical
type) will be polarized anodically and some cathodically with respect to their intrinsic corrosion potential in 0.1 M NaCl. The
corrosion potential of AA7075-T651 is relatively constant over periods of extended immersion in 0.1 M NaCl at a value in the
vicinity of 0.79 VSCE [13]. At this potential, the value of current developed by each particle type may be read from their
polarization curves measured by microcell measurements. These values of current density at 0.79 VSCE from polarization
curves are summarized in Table 1. Particles polarized above their intrinsic corrosion potential are expected to dissolve anodically. Conversely, those polarized cathodically are expected to support reduction reactions, mainly oxygen reduction, at enhanced rates. This will likely lead to local cathodic corrosion and dissolution of the alloy matrix phase around the periphery
of the particles. It is these latter particles that develop pits that trigger fatigue crack initiation.
From the microstructural characterization of AA7075-T651, Al7Cu2Fe was determined to be the dominant (>50%) constituent particle type in the overall particle population. The remaining fraction of the total population was comprised mainly of

Fig. 2. Equivalent diameter of particles, by type, that populates AA7075-T6. Equivalent diameter is used since the aspect ratio of particles is typically not
equal to unity.

644

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

Fig. 3. Nearest neighbor distance of particles, by type, that populate AA7075-T6.

Table 1
Summary of key parameters in regards to localized corrosion for particles in AA7075-T6.
Particle

Mean diameter (lm)

NA (#/mm2)

Mean Ecorr (VSCE)

Relationship to matrix

Mean current sustained by particle (ipart)


at Ecorr (AA7075-T651) in 0.1 M NaCl (A/cm2)

MgZn2
Al7Cu2Fe
Al3Fe
Mg2Si
Al2CuMg

0.052
1.7
1.69
2.13
5.1

3.5  107
1651
571
856
201

1.03
0.55
0.54
1.54
0.88

Anodic
Cathodic
Cathodic
Anodic
Cathodic

1.0  103
3.1  104
9.9  105
1.9  104
2.1  106

the Mg2Si and Al3Fe followed by a smaller fraction of Al2CuMg. Microelectrochemical characterization reveals that Al7Cu2Fe
is noble to the surrounding matrix phase with a corrosion potential of 0.55 Vsce. It is spontaneously passive, but supports
oxygen reduction at a rate of about 3050 lA/cm2 at the free corrosion potential of AA7075-T651, resulting in localized dissolution of the matrix phase at the particle periphery. In view of its large population in the alloy and it electrochemical characteristics, Al7Cu2Fe is treated as the primary contributor to the formation of large pits in AA7075-T651. On the other hand,
Mg2Si (and indeed MgZn2) particles are active compared to the matrix (1.54 Vsce). Upon exposure to solution, these particles dissolve very rapidly leaving small, apparently passivated pits that do not penetrate deeply into the alloy. Since these
ve particle types were determined to be the most abundant by size or frequency, the majority of pitting is likely caused
by them, and their inuence on the corrosion damage accumulation process captured in optical prolometry measurements
described below.
3.3. Predicted versus observed damage
A simple model relating pit depth to the microstructural and electrochemical characteristics of particles can be developed
based on corrosion cell principles and by taking advantage of several key assumptions and simplications. Under free corrosion conditions, the anodic current and cathodic current supported on a surface must be equal. Assuming that both the
reduction reaction kinetics on the matrix phase and the oxidation kinetics on the cathodic particle are negligible and neglecting reduction reaction kinetics within pits, the pit dissolution current of the matrix, IA, is coupled to oxygen reduction current
on the cathodic particles, IC. On this basis, Wei proposed a model based on Faradays law relating the pit depth distribution,
a*, to the average cathodic current density on intermetallic particles for predicting pit depths [14]

a ainit

Vm
 ipart dp  2p  r p 2  t
nF

In this expression Vm is the molar volume of Al, F is Faradays constant, n is equivalents per mole reacted, ainit is initial pit
depth (equal to rp for cathodic particles), ipart is average current density on participating particle (from polarization testing),

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

Fig. 4a. Representative polarization curve for Mg2Zn in 0.1 M NaCl solution pH 6.5 at 20 C.

Fig. 4b. Representative polarization curve for Al7Cu2Fe in 0.1 M NaCl solution pH 6.5 at 20 C.

Fig. 4c. Representative polarization curve for Al3Fe in 0.1 M NaCl solution pH 6.5 at 20 C.

645

646

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

Fig. 4d. Representative polarization curve for Mg2Si in 0.1 M NaCl solution pH 6.5 at 20 C.

Fig. 4e. Representative polarization curve for Al2CuMg in 0.1 M NaCl solution pH 6.5 at 20 C.

rp is the particle radius (the  denotes that this is treated as a distributed value for a population of particles, e.g., Fig. 2), t is
the exposure time, and dp is density of particles (number per unit area).
This model assumes a constant volumetric pit growth rate, which tends to hold for small pits, and a hemispherical pit
morphology. It links current to mass loss and, ultimately, volume loss with the appropriate relations. To develop a distribution of pit depths, a*, data determined from microstructural and electrochemical testing is used. In this simplied analysis,
ainit, ipart, dp are treated as single-valued parameters characteristic of the alloy, and r p is treated as a distributed variable with
a distribution constructed from data shown in Fig. 2.
Utilizing the current densities for the cathodic particles in Table 1, the average particle diameter values (used for ainit) and
the particle number density characterized by the data in Figs. 1 and 2, pit sizes were estimated as a function of exposure time
in 0.1 M NaCl solution using Eq. (2). Though the model does not account for the dissolution of anodic particles, like MgZn2, a
distribution of pits resulting from these particles was determined in a separate calculation using Faradays law, an assumption of hemispherical pit morphology and the anodic current densities for MgZn2 found in Table 1. Once the size of these pits
reached that size of particle, it was assumed that these sites passivated and did not propagate either as single pits or by pit
coalescence. Though these pit sites are too small to serve as effective fatigue crack initiation sites, they were included in the
analysis. The combined results of both types of pits are seen in Fig. 4, which shows cumulative distribution plots for predicted pit depth as a function of time up to 1344 h of exposure. The form of the model causes small particles to grow into
smaller sized pits and large particles to grow into large pits. A large fraction of the pits never grow to sizes beyond 5 lm in
diameter, and the maximum predicted pit depths approach the 40 lm depth range for 1344 h.

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

647

Fig. 5. Predicted pit depth distributions as a function of simulated immersion time in 0.1 M NaCl based on independent microstructural characterization
and electrochemical testing.

Fig. 6. Damage depth distributions as a function of immersion time in 0.1 M NaCl as determined by optical prolometry.

Fig. 7. Observed damage depth and predicted pit depth for the maximum, 99th percentile and median of the respective distributions as a function of
immersion time.

648

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

Fig. 8. Error in the model predictions relative to observed damage depths as a function of immersion time.

Fig. 5 shows cumulative distributions of pit depths determined from OP measurements as a function of exposure time for
AA7075-T651 exposed to aerated 0.1 M NaCl solution for a period up to 1344 h. Values of damage depth less than the 5th
percentile pit depth shown in Fig. 4 were removed from the data set. This removed data associated with damage due to surface roughening that was not associated with the pitting process. At the low end of the distributions, observation and prediction are forced to agree due to truncation of the optical prolometry data sets. The evolution in the damage depth
distribution is similar to predictions as a function of immersion time. The trend in maximum damage depths generally follows the predicted trend in pit depth. The maximum observed damage depth approaches 40 lm, which is similar to the maximum pit depth predicted by the model.
Fig. 6 shows a comparison of predicted and observed pit growth at the median and maximum of the pit depth distributions as a function of exposure time. This comparison shows the constant volumetric growth behavior imposed by the model
under-predicts damage depth observations at short times and appears to over-predict at longer times. This suggests a very
rapid initial accumulation of damage in the rst 100 h, perhaps associated with dissolution of active particles. At longer
times, it appears that the model will over-predict damage accumulation, particularly at the largest damage depth. The lack
of growth in the damage depth distribution suggests an inuence of repassivation, which is not accounted for in the model.
The assumptions and simplications underlying the model signicantly constrain the form of the pit size distribution and
how it evolves with time. As a result, there are important differences between prediction and observation. Fig. 7 is a plot of
the relative error expressed as a percentage between the predicted and observed pit depths at the median, 99th percentile
and maximum of the pit depth distributions as a function of exposure time. Overall, prediction and observation agree within
a factor of 2 across the exposure interval used in this experiment. There is no evidence to suggest that the model does a

Fig. 9. Maximum and median pit diameter measure on AA7075-T651 as a function of immersion time in 0.1 M NaCl.

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

649

Fig. 10. Percent surface area corroded as a function of immersion time for AA7075-T651 in 0.1 M NaCl.

better or worse job in predicting at the median or maximum of the damage depth distribution. However, as a function of
exposure time, the largest relative errors are observed at the shortest times where the model tends to under-predict the largest damage depths in the distribution.
The manner in which optical prolometry was carried out in these experiments allowed characterization of other aspects
of corrosion damage accumulation when the optical prolometry images were post processed using thresholding and image
analysis methods. Other characteristics measured included pit diameter, pit density, and percent area corroded. These characteristics were not used in the model described above, but their evolution as a function of exposure time illustrates noteworthy aspects of localized corrosion damage accumulation.
Fig. 8 shows the maximum and median equivalent pit diameters as a function of exposure time. Like pit depths, pit diameters increase rapidly during the rst 100 h of exposure. After this, in both the median and maximum cases, there is a period
of slow or no growth, possibly followed by a later period of slower growth. Pit diameters are 58 times greater than pit
depths at the same exposure times. This indicates that the pits are shallow, and possibly subject to repassivation associated
with the difculty in sustaining a critical environment. Given the aspect ratio of these pits, it is suspected that corrosion
product accumulation in pits blocks transport and helps sustain the aggressive environments needed for pit growth. It
should be noted that the pit aspect ratio indicated by the OP measurements suggest that hemispherical pit shape assumption
used in formulating the pit depth model in Eq. (2) should be rened.
Figs. 911 show the percentage of the surface area covered by pitting and the number density of pits as a function of
exposure time. The area of the sample surface corroded approaches 10% within the rst 100 h of exposure, then changes very
little for a period before increasing again at the longest exposure time. Not surprisingly, this trend in area coverage of pitting

Fig. 11. Pit density as a function of immersion time for AA7075-T651 in 0.1 M NaCl.

650

M.K. Cavanaugh et al. / Engineering Fracture Mechanics 76 (2009) 641650

mirrors the trend in pit diameter and reinforces the notion of an early episode of aggressive corrosion followed by much
slower growth or repassivation of some pits. The pit number density trend is interesting. Pit density increases to a maximum
approaching 4000 pits/cm2 within 100 h, then decreases. The decrease with continued immersion time is attributed to coalescence of adjacent pit sites, leading to fewer pits, which are larger in size. Although these two characteristics of pitting
damage are not explicitly included in the model represented in Eq. (2), recent work of Jones indicates that quantities such
as pit surface area and pit density can also play just as important a role in determining where a crack will form [3].
4. Conclusions
 Quantitative stereology shows that the predominant intermetallic particles in AA7075-T651 on the basis of size and number density are MgZn2, Al7Cu2Fe, Al3Fe, Mg2Si, and Al2CuMg. These particles strongly affect corrosion damage accumulation of the alloy. The physical and electrochemical characteristics were used in generating a model to predict corrosion
damage accumulation.
 A simple model has been formulated following the work of Wei based on the assumption of a constant volumetric growth
rate of hemispherical pits [14], Faradays law and inputs from microstructural and electrochemical characterization experiments. For the simple case of constant immersion in 0.1 M NaCl, pit depth distributions ranging up to 40 lm can be
calculated.
 A comparison between pit depth prediction and damage depth observations derived from optical prolometry illustrate
where improvements in the model predictions are needed. Optical prolometry suggests that damage accumulates
rapidly on the alloy surface and then slows signicantly at longer times. A constant volumetric pit growth model
under-predicts damage depth at short exposure times and appears to over-predict at longer times. It is speculated that
under-prediction of the model is associated rapid dissolution of active particles and over-prediction is due to the fact that
the model does not account for repassivation.

Acknowledgements
The authors would like to acknowledge the support of the Defense Advanced Projects Agency and the Aluminum Company of America. MKC thanks the National Defense Science and Engineering Graduate Fellowship Program.
References
[1] Sankaran K, Johnson B, Perez R. Kinetics of pitting corrosion and effects on fatigue behavior of aluminum alloy. In: Tri-service conference on
corrosion. Wrightsville Beach, NC: Department of Defense; 1997. p. 645.
[2] van der Walde K et al. Multiple fatigue crack growth in pre-corroded 2024-T3 aluminum. Int J Fatigue 2005;27.
[3] Jones K, Hoeppner DW. Pit-to-crack transition in pre-corroded 7075-T6 aluminum alloy under cyclic loading. Corros Sci 2005;47.
[4] Hoeppner DW. Model for prediction of fatigue lives based upon a pitting corrosion fatigue process, in Fatigue Mechanisms. In: Fong JT, editor.
Proceedings of ASMT-NBS-NSF Symposium, Kansas City, Mo., ASTM STP675; 1979. p. 84170.
[5] Kondo Y. Prediction of fatigue crack initiation life based on pit growth. Corrosion 1989;45:7.
[6] Wang YQ et al. Evaluation of the probability distribution of pitting corrosion fatigue life in aircraft materials. Acta Mech Sinica 2003;19(3).
[7] Sankaran KK, Perez R, Jata KV. Effects of pitting corrosion on the fatigue behavior of aluminum alloy 7075-T6: modeling and experimental studies.
Mater Sci Engng A 2001;297:223.
[8] Harlow DG, Wei RP. Life prediction the need for a mechanistically based approach. Key Engng Mater 2001;200:119.
[9] Birbilis N et al. Understanding damage accumulation upon AA7075-T651 used in airframes from a microstructural point of view. In: Symposium
applications of materials science to military systems MS&T. Pittsburg, PA: TMS; 2005.
[10] Birbilis N, Buchheit RG. Electrochemical characteristics of intermetallic phases in aluminum alloys an experimental survey and discussion. J
Electrochem Soc 2005;152:B140.
[11] Koul MG. Topographical analysis of pitting corrosion in AA7075-T6 using laser prolometry. Corrosion 2003;59.
[12] Birbilis N, Cavanaugh MK, Buchheit RG. Electrochemical behavior and localized corrosion associated with Al7Cu2Fe particles in aluminum alloy 7075T651. Corros Sci 2006;48(12):64.
[13] Birbilis N, Cavanaugh MK, Buchheit RG. Electrochemical response of AA7075-T651 following immersion in NaCl solution. In: Symposium corrosion of
advanced materials in honor of Prof. Koji Hashimoto ECS transactions. Los Angeles CA: The Electrochemical Society; 2005.
[14] Wei RP. A model for particle-induced pit growth in aluminum alloys. Scripta Mater 2001;44.

Anda mungkin juga menyukai