You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/334254357

Thermomechanical treatment and corrosion resistance correlation in the


AA2198 Al–Cu–Li alloy

Article  in  Corrosion Engineering Science and Technology · July 2019


DOI: 10.1080/1478422X.2019.1637077

CITATIONS READS

12 222

8 authors, including:

João Victor de Sousa Araujo Aline Bugarin


University of São Paulo University of São Paulo
25 PUBLICATIONS   164 CITATIONS    7 PUBLICATIONS   63 CITATIONS   

SEE PROFILE SEE PROFILE

Uyime Donatus Caruline de Souza Carvalho Machado


Ceres Power Holdings Instituto de Pesquisas Energéticas e Nucleares
51 PUBLICATIONS   490 CITATIONS    25 PUBLICATIONS   121 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Corrosion Protection of Electrogalvanized Steel by environmentally friendly conversion coatings. View project

TSA anodizing for corrosion protection of Al alloys View project

All content following this page was uploaded by Aline Bugarin on 28 August 2019.

The user has requested enhancement of the downloaded file.


Corrosion Engineering, Science and Technology
The International Journal of Corrosion Processes and Corrosion Control

ISSN: 1478-422X (Print) 1743-2782 (Online) Journal homepage: https://www.tandfonline.com/loi/ycst20

Thermomechanical treatment and corrosion


resistance correlation in the AA2198 Al–Cu–Li alloy

João Victor de Sousa Araujo, Aline de Fátima Santos Bugarin, Uyime


Donatus, Caruline de Souza Carvalho Machado, Fernanda Martins Queiroz,
Maysa Terada, Antonello Astarita & Isolda Costa

To cite this article: João Victor de Sousa Araujo, Aline de Fátima Santos Bugarin, Uyime
Donatus, Caruline de Souza Carvalho Machado, Fernanda Martins Queiroz, Maysa Terada,
Antonello Astarita & Isolda Costa (2019): Thermomechanical treatment and corrosion resistance
correlation in the AA2198 Al–Cu–Li alloy, Corrosion Engineering, Science and Technology, DOI:
10.1080/1478422X.2019.1637077

To link to this article: https://doi.org/10.1080/1478422X.2019.1637077

Published online: 04 Jul 2019.

Submit your article to this journal

Article views: 11

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ycst20
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY
https://doi.org/10.1080/1478422X.2019.1637077

Thermomechanical treatment and corrosion resistance correlation in the AA2198


Al–Cu–Li alloy
João Victor de Sousa Araujo a, Aline de Fátima Santos Bugarina, Uyime Donatus a, Caruline de Souza
Carvalho Machado a, Fernanda Martins Queiroz a, Maysa Teradaa, Antonello Astarita b and Isolda Costa a

a
Instituto de Pesquisas Energéticas e Nucleares, IPEN/CNEN, São Paulo, Brazil; bDepartment of Chemical, Materials and Industrial Production
Engineering, University of Naples Federico II, Napoli, Italy

ABSTRACT ARTICLE HISTORY


The influence of T3, T8 and T851 thermomechanical treatments on the microstructure and corrosion Received 7 May 2019
resistance of the AA2198 was investigated. Differential scanning calorimetry and scanning electron Accepted 24 June 2019
microscopy were used for microstructural characterisation, whereas electrochemical methods were
KEYWORDS
employed to analyse the corrosion behaviour of the alloy. The morphology and composition of Aluminium–lithium alloys;
constituent particles were similar for the T3 and T8 thermomechanical treatments but varied in the thermomechanical
T851. There was an inverse relation between T1 phase density and corrosion resistance. The T3 treatments; severe localised
treatment with the highest corrosion resistance was the one with the lowest density of T1 phase. corrosion
The mechanisms of corrosion varied with the thermomechanical treatments.

Introduction
The mechanical and electrochemical behaviour of the Al– the precipitation of the T1 phase during thermomechanical
Cu–Li alloys have been recently studied by many researchers treatments in order to improve the mechanical performance
[1–11]. These alloys play an important role in the aerospace of the Al–Li alloys [18,19].
industry [12]. The reason for the constantly growing interest The T1 phase plays an important role in the corrosion
in the study of these materials is the possibility of substituting resistance of Al–Cu–Li alloys. Buchheit et al. [15] proposed
the conventional Al alloys used in aircraft structures. These that the dissolution of T1 particles preferentially located at
materials have showed improvement in mechanical proper- grain and subgrain boundaries in the AA2090 leads to inter-
ties compared with previous generations of Al–Cu–Li alloys granular attack. These authors also revealed that constituent
and promote improvements in fuel efficiency and, as a conse- particles lead to galvanic corrosion at the particle/matrix
quence, lower emission of CO2 gas. These alloys are particu- interface. Ma et al. [20] investigated the effects of microstruc-
larly suitable for any applications where the strength-to- ture on the corrosion resistance of the AA2099-T83 and
weight ratio and fatigue resistance are of major concern. observed that the corrosion attack caused by the constituent
The AA2198 alloy studied in this work is one of the most particles was superficial when compared with that caused
advanced alloys among the commercially available ones by T1 phase dissolution. They named this last type of attack
from the third generation Al–Cu–Li alloys. It has been chosen as severe localised corrosion (SLC). Li et al. [21] proposed
as the material for the fuselage skin sheet material of the A350 another corrosion mechanism associated with the T1 phase.
Airbus [13]. They suggested that upon exposure to corrosive environ-
In Al–Cu–Li alloys, the elements Cu, Mg and Li are added ments, the anodic T1 phase initially undergoes selective dis-
as alloying elements to improve the mechanical properties of solution of Li and Al and, subsequently, becomes cathodic
the alloys [14]. The microstructure of these alloys is com- to the matrix due to Cu enrichment which results in the
posed of micrometric particles, also called constituent par- attack of the surrounding matrix.
ticles (CPs), and nanometric hardening phases. In the Zou et al. [22] investigated the influence of aging time on the
AA2198 alloy, the CPs are cathodic to the aluminium matrix corrosion behaviour of the AA2198-T83 alloy through electro-
and cause localised attack around their periphery [15]. In Al– chemical tests in 3.5 wt-% NaCl environment. They revealed
Cu–Li alloys, the nanometric phases, such as the δ′ (Al3Li), Θ′ that mechanical properties and corrosion resistance gradually
(Al2Cu), S (Al2CuMg) and T1 (Al2CuLi), contribute to their decreased with aging at 175°C until 18 h (peak aged), but after
mechanical resistance [16]. These phases are formed during this stage, the corrosion resistance increased due to T1 phase
aging and are significantly affected by thermomechanical precipitation. However, as mentioned earlier, many works
treatments (tempers), which influence the mechanical and [2,10,11,18,19,23] have showed that precipitation of T1 phase
corrosion resistance of the alloys [17]. For example, the leads to low corrosion resistance in the alloys Al–Cu–Li.
AA2020 alloy shows lower anisotropy degree in the peak Ye et al. [8] studied the effect of annealing on the corrosion
aged condition (T8) compared with the underaged (T3) resistance of artificially aged AA2050 alloy (T6). They found
one. The higher amounts of T1 precipitates formed in the that annealing after cold-rolling reduced the alloy suscepti-
T8 condition compared with the T3 reduce the tensile aniso- bility to intergranular corrosion due to heterogeneous pre-
tropy significantly [16]. Thus, many researchers have studied cipitation of T1 phase.

CONTACT João Victor de Sousa Araujo joao-neutron@hotmail.com Instituto de Pesquisas Energéticas e Nucleares, IPEN/CNEN, Av. Prof. Lineu Prestes, 2242
São Paulo, Brazil
© 2019 Institute of Materials, Minerals and Mining Published by Taylor & Francis on behalf of the Institute
2 J. V. DE SOUSA ARAUJO ET AL.

Yan et al. [9] studied the naturally aged AA2050 alloy Table 2 . Heat treatments for the aluminium alloys used in this study.
(T351 designation), and the same alloy artificially aged at Heat treatment
condition Description
155 and 200°C for different times. They concluded that the
intergranular corrosion found for some of the conditions T3 (1) Solution heat treated; (2) cold worked; (3) naturally aged.
T8 (1) Solution heat treated; (2) cold worked (3) artificially aged.
tested is controlled by mechanisms other than the T1 phase T851 (1) Solution heat treated; (2) stress relieved by stretching;
precipitation and related it to solute segregation, with recent (3) artificially aged.
evidence for grain boundary Li segregation in the AA2050-
T351 alloy [7]. These works showed that thermomechanical
conditions significantly affect the corrosion resistance of the Microhardness measurements were recorded from the
new-generation of Al–Cu–Li alloys and, consequently, the samples surface using a load of 300 gf for a dwell time of
corrosion behaviour related to each of the thermomechanical 10 s, and 20 measurements were taken from each alloy. A
treatment used needs proper investigation. Buehler 1600 Series microhardness tester was employed for
The present work investigates the microstructures related these measurements.
to the AA2198 alloy in the T3, T8 and T851 temper con- Scanning Kelvin Probe Force Microscopy (SKPFM) was
ditions and proposes corrosion mechanisms that are corre- carried out using the Nanoscope IIIa MultiMode microscope
lated with the microstructure of the alloy. in tapping mode. The topographic and surface potential
images were obtained using the lift mode. The measurements
were performed at room temperature and humidity between
Experimental 30 and 55%.
The chemical compositions and the temper conditions of the Corrosion resistance was evaluated by immersion test,
AA2198 alloys with the T3, T8 and T851 designations used in open circuit potential (OCP) measurements and electroche-
this study are presented in Tables 1 and 2, respectively. mical impedance spectroscopy (EIS) as a function of immer-
Samples of the alloys were ground with silicon carbide sion time. Electrochemical measurements were performed
papers of #500, #1200 and #4000 and then polished with dia- using a three-electrode set-up configuration with the alu-
mond suspensions of 3 and 1 μm. The thicknesses of the alloy minium alloy samples as working electrodes. The reference
samples in the T3, T8 and T851 conditions were 3, 4 and electrode used was an Ag/AgCl, (KClsat), and a platinum
1.6 mm, respectively. wire was employed as counter-electrode. A surface area of
The microstructures of the AA2198 alloys were observed 1.0 cm2 of the working electrode was exposed to the electro-
after etching in a solution with 2 (v/v)% hydrofluoric and lyte. The experiments were carried out in naturally aerated
25 (v/v)% nitric acid. The evolution of the corrosion phenom- 0.01 mol L−1 NaCl solution at temperature of (23 ± 2)°C.
ena was studied as a function of immersion time by surface EIS measurements were performed on a Solartron 1287
observation using a Leica DMLM optical microscope coupled potentiostat coupled to a frequency response analyser
to a Leica EC3 camera and controlled by a LAS ES software. (FRA). The voltage perturbation range used in the EIS
In order to quantify the depths of corrosion penetration in the measurements was 20 mV (rms), and the acquisition rate
alloys, a ZESCOPE model optical profilometer was used. was 10 points per decade in a frequency range of 10 kHz to
Analysis of the microstructure and corrosion evolution on 10 mHz. Each test was performed at least three times to
the surfaces of the aluminium alloys were carried out by Scan- evaluate the reproducibility of the results.
ning Electron Microscopy (SEM) using TM3000 and JSM-
6701F microscopes. Both microscopes are equipped with
energy dispersive X-ray spectroscopy (EDS) detectors. Results
Differential scanning calorimetry (DSC) data were Microstructural characterisation
obtained in nitrogen (99.999 wt-%) atmosphere using a
DSC-50 SHIMADZU equipment coupled to a TA-60WS. Elongated grains were observed aligned in the direction of
Samples were ground with silicon carbide paper before the deformation for the alloys subjected to T3 and T8 thermo-
acquisition of DSC data. The weights of the samples mechanical treatments (Figure 1(a,b,d,e)). This is common
employed in the DSC tests were in the range of 20–30 mg. in materials that are extruded or rolled [24]. Equiaxial grains
The heating rate used in DSC measurements was were observed on the surface (LT/L directions) of the alloy
10°C min−1 and the scanning temperature ranged from subjected to the T851 treatment (Figure 1(c,f)). Furthermore,
50 to 550 °C. etching of the AA2198-T851 revealed grains with features
that are related to grains with crystallographic orientation
whose higher packing planes 〈111〉 are oriented in the same
Table 1. Chemical composition (in wt-%.) obtained by inductively coupled
plasma – atomic emission spectroscopy (ICP-AES) of the Al–Cu–Li alloys used
direction as the maximum shear stress [24,25], Figure 1(g).
in this study. This leads to the creation of slip bands which are preferred
Elements (wt-%) AA2198-T3 AA2198-T8 AA2198-T851 locations for T1 phase precipitation. Slip bands are intro-
Al Balance Balance Balance duced in the materials during stretching processes and they
Cu 3.20 3.32 3.31 cause microstructural changes [11]. Figure 1(g) shows signifi-
Li 1.02 0.96 0.97
Mg 0.30 0.31 0.31
cant differences among the grains. Grains A and B present
Ag 0.29 0.20 0.25 slip bands but not grains C and D.
Zr 0.38 0.51 0.40 The microstructure is also composed of constituent par-
Fe 0.04 0.04 0.04
Si 0.05 0.03 0.03
ticles of micrometric sizes. These particles are formed during
Zn 0.02 0.006 0.006 casting due to the low solubility of iron in aluminium [4,26].
Mn 0.003 0.003 0.003 In each of the alloys tested, more than 100 constituent par-
Cu/Li 3.1 3.5 3.4
ticles were analysed by EDS. These were preferentially aligned
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 3

Figure 1. Optical micrographs of AA2198 alloy with (a) T3, (b) T8, and (c) T851 tempers; (d), (e) and (f) are higher magnification images of the squared regions in (a),
(b), (c); and (g) is a higher magnification image of the squared region in (f).

in the direction of deformation (Figure 2(a)), and were found to the T3 and T8, and many of them were irregularly shaped.
as clusters (Figure 2(b,c)), or isolated, as shown in the circled The shape of the precipitates is determined by minimisation
region in Figure 2(b). The constituent particles were predo- of interface energy and strain energy [27]. The results showed
minantly composed of Al, Cu and Fe (Figure 2(d)). It was that the constituent particles distribution is more hetero-
found that 92% of these particles in the alloy with T851 des- geneous in the T851 designation.
ignation were mainly composed of Al, Cu, and Fe, and only The T1 phase (Al2CuLi) is the main hardening phase in
8% of them were composed of Al and Cu. The particles com- Al–Cu–Li alloys [28–30]. According to the literature [31],
posed of Al, Cu and Fe presented varied morphologies and the precipitation of this phase occurs at temperatures between
irregular shapes, Figure 3(a); whereas those with Al and Cu 138 and 260°C. The presence of this phase in the tested alloys
displayed regular morphologies, either circular or rectangular was investigated by DSC curves. These curves allow the study
Figure 3(b). of precipitation and dissolution of second phases from
The area fraction of the particles was quantified through exothermic and endothermic peaks [31–35]. The height of
image analysis in an area of 220 × 170 μm2. The percentage the peak indicates the intensity of the event [33]. The T3
of these particles in the T3 and T851 alloys is similar, but alloy showed a high peak in the range of 240–260oC related
in the T8 designation it was nearly twice that of the others with the precipitation of T1 phase (Figure 6). The peak A is
studied (Figure 4). For alloy T851, the Cu/Fe ratio was in associated with the dissolution of δ′ (AlLi3) that occurs
the range of 2–7, whereas in the other tempers (T3 and T8) between 200 and 240 oC; peak B is associated with the precipi-
this ratio was mainly in the range of 2–4, Figure 5. The par- tation of T1 phase; while peaks C and D are associated with
ticles in the T3 and T8 alloys displayed similar morphologies. T1 phase coarsening and dissolution, respectively [31]. It
On the other hand, the T851 alloy showed differences in the has been reported that natural aging causes minor precipi-
composition and morphologies of these particles in relation tation of T1 phase in Al–Li alloys [18,36]. As a result, the
4 J. V. DE SOUSA ARAUJO ET AL.

Figure 2. SEM micrographs showing constituent particles in the 2198 alloy. (a) Constituent particles in the direction of deformation. (b) Higher magnification micro-
graph of the squared region of the micrograph (a). (c) Clusters of intermetallic over the alloy surface. (d) Typical EDX profile of the constituent particle.

T3 alloy has the largest amounts of Li in solid solution among measurements clearly differentiated these two treatments
the tested tempers explaining the high peak of T1 precipi- with the latter alloy presenting higher hardness compared
tation in the T3 designation. The lower peaks related to T1 to the first.
phase precipitation in the T8 and T851 tempers (indicated
by square in Figure 6) compared to the T3 one is due to
Corrosion immersion test
the artificial aging of these alloys that favours precipitation
of this phase [36]. These results are supported by microhard- Figure 8 shows the evolution of corrosion on the surface of the
ness measurements that showed an increase in the hardness AA2198 alloy for the various treatments tested at increasing
values for the alloys subjected to the T8 and T851 temper periods of immersion in 0.01 mol L−1 NaCl solution. The loca-
treatments (Figure 7) compared with the T3. Although DSC lised corrosion in the T3 alloy was mainly associated with the
results did not show significant differences between the T8 matrix surrounding the constituent particles due to the for-
and T851 thermomechanical treatments, the microhardness mation of microgalvanic cells leading to trenching and shallow

Figure 3. (a,b) SEM micrographs of the constituent particles in the AA2198-T851 alloy with different morphologies and (c) typical EDX profile of the constituent
particle in (b).
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 5

Figure 4. Percentage of constituent particles in various tempers of the AA2198


alloy.

pits. The difference in potential between the matrix and con-


stituent particle was confirmed by SKPFM. Figure 9 shows the
topographic and volta potential maps of a coarse Al–Cu–Fe
intermetallic particle in the matrix of an AA2198 alloy. The esti-
mated mean volta potential difference between the particle and
the matrix, based on the analysis of 20 constituent particles was
in the range of 30 ± 9 mV (Figure 9(b,c)) with higher potentials
related to the constituent particle relative to the matrix. These
results show that galvanic coupling leading to localised attack
might occur between the particles and matrix.
For the T8 and T851 alloys, two forms of corrosion were
identified: trenching and shallow pits associated with the con-
stituent particles and SLC inside the grains, related to harden-
ing phases. These two types of attacks were observed from the
early hours of immersion. In the T3 alloy, however, only loca-
lised attack associated with the constituent particles was
identified as the main corrosion process (Figure 10(a,b)).
For the T8, the attack progressed in the direction of defor-
mation as indicated by the arrow in Figure 10(c) and the
SLC development led to the selective dissolution of grains
(Figure 10(d)). For the T851 alloy, attacks associated with
the selective dissolution of slip bands (Figure 10(f)) were
observed. When the alloy is corroded, slip bands are revealed
due to the activities of T1 phase (Figure 1(e–g)). In the T8 and
T851 alloys, intragranular attack predominated, and the grain
boundaries were preserved, as evidenced in Figure 10(e). Figure 5. Histogram of Cu/Fe ratio in the constituent particles in the (a) T3, (b)
The depths of corrosion attack in the various alloys tested T8 and (c) T851 tempers of the AA2198 alloy.
were evaluated by optical profilometry (OP) for 50 sites of
trenches (LC) and SCL on the alloys after 24 h of immersion
test. OP has been previously used to investigate differences in
depths of corrosion attack [37]. In this study, the mean
depth of penetration estimated by OP for the trenches (LC)
was in the range of 1–5 μm, whereas in the SLC sites, the
mean depths of penetration was dependent on the temper.
For instance, it was in the range from 8 to 15 μm, for the T8
temper, and from 8 to 35 μm, for the T851. Figure 11 shows
compares the penetration depths related to the two main
types of attack in the AA2198 alloy for the three thermomech-
anical treatments.

Electrochemical tests
OCP variation with time of exposure to test solution is pre- Figure 6. DSC heat flow curves for the AA2198 alloy with tempers T3, T8 and
sented in Figure 12. It shows that the OCP for the T3 temper T851.
6 J. V. DE SOUSA ARAUJO ET AL.

predominant with time of test for the T3 temper and the elec-
trochemical behaviour of all alloys tested was similar after
24 h of test. Only the experimental data for high to medium
frequencies, a flattened capacitive arc, were fitted using a
CPE//R pair. These are related to charge transfer processes
coupled to the charging of the double layer. A high quality
fitting with chi-squared values in the order of 10−4 for T8
and T851 tempers, and of nearly 10−3 for the T3, was
obtained. The fitting and experimental results are compared
in Figure 14(a–c) as a function of time. A constant phase
element (CPE) was used instead of an ideal capacitor to
take into consideration the heterogeneity of the surface, as
already mentioned in the literature [38,39]. The CPE values
obtained from data fitting of the CPE//R pair for the T8
and T851 conditions, Figure 14(a), varied between approxi-
Figure 7. Microhardness of AA2198 alloy with T3, T8 and T851 tempers.
mately 15 and 30 μF cm−². These values are typical of char-
ging of the double layer. The oxide layer remaining on the
surface also has a slight contribution to this time constant.
was nearly 100 mV nobler than for the T8 and T851 ones. An initial increase in R was accompanied by a decrease in
This could be due to the higher amounts of copper in solid CPE values, Figure 14(b). This indicates an initial reduction
solution in the T3 alloy. Also, the OCP related to the T3 in surface activity that could due to the partial coverage of
was more stable than for the T8 and T851 one. The initial active sites at the surface. It was subsequently followed by R
decrease in OCP after immersion must be related to oxide decrease and CPE increase showing that new sites of the sur-
film attack. The results suggest that the oxide film on the face were activated resulting in higher proportion of electro-
T3 alloy is more resistant than that on the other alloys. The chemically active areas at the surface with time. This was
greater instabilities for the T8 and T851 alloys compared to supported by surface observation with time. α values
the T3 is most likely due to the greater amounts of T1 increased with time from around 0.8 to 0.9, Figure 14(c).
phase in the first tempers. The results also showed that the temper related to the lowest
EIS results for the three thermomechanical treatments CPE and highest R values was the T3 one that is indicative of
tested (T3, T8 and T851) as a function of immersion time the highest resistance of this temper to corrosion attack.
in 0.01 mol L−1 NaCl solution are shown for the AA2198
alloy in Figure 13(a–d) as Nyquist and Bode phase angle dia-
grams. These results suggest the presence of two-time con- Discussion
stants for the T8 and T851alloys, and three-time constants The dissimilarities between the constituent particles in the
for the T3 alloy. The highest impedances were related to alloy of this study, AA2198, and that used in the work of
this last temper. The impedance decreased with time for all Ma et al. [4], AA2099-T8, are explained by the differences
tempers. Similar impedance results were obtained for the in the alloys compositions. The AA2099 presents lower Cu
T8 and T851 alloys. Evidence of diffusional controlled pro- and Ag content, but higher Li, Mn and Zr compared to the
cesses was observed at low frequencies, mainly for the T8 AA2198. Slight differences in the composition of Al–Cu–Li
and T851 tempers. Diffusion became increasingly alloys lead to significant changes in their microstructures

Figure 8. Optical micrographs of AA2198 with T3, T8 and T851 tempers after various periods of immersion in 0.01 mol L−1 NaCl solution.
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 7

Figure 9. (a) AFM image of the surface of an AA2198-T8 alloy over a Cu,Fe-rich constituent particle; (b) corresponding surface potential map; and (c) volta potential
line analysis across the Cu, Fe-rich constituent particle.

Figure 10. SEM micrographs of AA2198 with different tempers after 24 h immersion in 0.01 mol L−1 NaCl solution: (a,b)T3; (c,d) T8; and (e,f) T851.

Figure 11. OCPs of AA2198 alloy with T3, T8 and T851 tempers in 0.01 mol L−1 NaCl solution.
8 J. V. DE SOUSA ARAUJO ET AL.

least five different compositions of constituent particles con-


taining some, or all, of the following elements Al, Cu, Fe, Mn
and Zn, and some constituent particles contained Li. In the
present study, however, one type of constituent particles lar-
gely predominated (92%), that is, that composed of Al, Cu
and Fe. This shows that each Al–Cu–Li alloy presents its
own characteristics and, consequently, the corrosion resist-
ance of each of this kind of alloy deserves proper
investigation.
The microstructure differences among the Al–Cu–Li
alloys and its effect on their corrosion resistance can be
explained by their fabrication history. The T3 treatment com-
prises solution heat treatment followed by cold rolling and
natural aging, whereas the T8 and T851 treatments involve
artificial aging. Artificial aging promotes nucleation of T1
phase [2,23]. This was confirmed by the DSC curves and
Figure 12. Optical profilometry and SEM images of AA2198 alloy with different
tempers after 24 h immersion in 0.01 mol L−1 NaCl solution. microhardness measurements. SLC was not seen in the
AA2198-T3 temper during immersion test but abound in
the T8 and T851 alloys. Many authors have reported mechan-
and, consequently, in their corrosion resistance. Ma et al. [4] isms for SLC initiation and propagation that were related
revealed three groups of intermetallic compounds with differ- with the heterogeneity of plastic deformation [2,19,41]. As
ent compositions, that were classified as: (a) low, (b) inter- the plastic deformation during alloy production is not uni-
mediate and (c) high Cu-content particles. MacRae et al. form. Some grains are more deformed than others, and grains
[40] also investigated the AA2099-T8 alloy and found at with different levels of deformation, according to their

Figure 13. Comparison of EIS results for the AA2198 alloys with T3, T8 and T851 tempers at various times of immersion in 0.01 mol L−1 NaCl solution.
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 9

Figure 14. Variation of CPE, R and α as a function of time of immersion in 0.01 mol L−1 NaCl solution. These values were obtained from fitting the experimental data
to the EEC of Figure 13.

orientation, are created. T1 phase precipitates preferentially at stress relief before cold working, this stage results in a particu-
dislocations and, consequently, highly deformed grains pre- lar type of attack, that is, that related to slip bands resulting
sent high density of T1 phase and these are highly susceptible from the stretching stage.
to corrosion attack. Recently, Zhang et al. [42] in their study The variations in the depths of attack due to SLC which
on the 2A97-T6 alloy, a newly developed Al–Cu–Li alloy, were observed between the T8 and T851 tempers result
proposed this explanation. Ma et al. [2], in their study on from a heterogeneous distribution of T1 phase in both alloys,
the AA2099-T8 suggested that the SLC was related to attack and the dissimilarities in this phase content, as indicated by
of the T1 phase preferentially precipitated in deformation the microhardness measurements. On the other hand, the
bands and, therefore, to preferential dissolution of these differences in depth of corrosion penetration by trenching
bands. The corrosion attack in the AA2198-T8 alloy of this might be related to the different sizes of constituent particles.
study was also s related to the deformation bands introduced The differences in the OCP values of the three tested alloys
during cold working. This resulted in intragranular attack of show that artificial aging plays a significant role in the electro-
the grains with high strain. Since, in the T851 condition, chemical behaviour of this alloy. The noblest potentials
artificial aging is preceded by a stretching stage to induce related to the T3 alloy is due to the highest amount of copper

Figure 15. Schematic diagrams illustrating the corrosion mechanism associated with constituent particles: (a) surface after polishing; (b) constituent particle after
polishing; (c) corrosion reactions during the immersion in NaCl; (d) features corrosion after exposure to the solution; and (e) surface final after testing.
10 J. V. DE SOUSA ARAUJO ET AL.

in solid solution in this alloy. The higher kinetics of surface due to the preferential dissolution of aluminium and lithium
attack in the case of the T8 and T851 alloys is related to the and are counter balanced by oxygen reduction on the catho-
higher amount of T1 in these two alloys compared to the dic particles, Figure 15(c). These processes result in trenching
T3. The higher amounts of T1 phase in the artificially aged in the matrix surrounding the constituent particles, Figure 15
alloys (T8 and T851) also resulted in increase in the corroded (d). The attack related to the constituent particles was superfi-
areas with the removal of passive film and decrease in OCP. cial, Figure 15(e), as observed in Figure 10(a,b). Trenching
Huang et al. [6] related the effect of aging time with the also occurs in the T8 and T851 tempers due to the similar
OCP changes in the 1460 alloy. They showed that lower cathodic particles in these alloys. But, in these last two tem-
OCP values were related to longer aging times and correlated pers, SLC predominated, as illustrated by the schematic
it to the increase in the volume of T1 phase with extended images in Figures 16 and 17, respectively.
aging times. The plastic deformation of individual grains depends on
In this work, corrosion mechanisms were proposed for the grain orientation and consequently, some grains are more
localised corrosion in each of the three tempers studied. For deformed than others, Figure 16(a). Assuming that grain B
the T3 temper (Figure 15), corrosion initiated at the areas sur- is more deformed that grains A and C, the T1 phase density
rounding the constituent particles which are preferentially is higher inside grain B, Figure 16(b). In the T851 condition,
located along the deformation direction, Figure 15(a). Figure the stretching before artificial aging favours the formation of
15(b) shows the constituent particles in cross section after the slip bands, Figure 17(a) and preferential precipitation related
initiation of the corrosion process. The anodic reactions at the to them, Figure 17(b) The most active elements in T1 phase
matrix surrounding the particles result in copper enrichment (Li and Al) are preferentially dissolved resulting in

Figure 16. Schematic diagrams illustrating the corrosion mechanism for the T8 condition: (a) surface after polishing; (b) grain with higher density T1 phase (c), (d)
and (e) corrosion reactions during the immersion in NaCl; (f) corrosion features after exposure to the solution.

Figure 17. Schematic diagram illustrating the corrosion mechanism for the T8 condition: (a) surface after polishing; (b) grains with higher density T1 phase (c), (d)
and (e) corrosion reactions during the immersion in NaCl; (f) features corrosion after exposure to the solution.
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 11

crystallographic attack in the alloy, and SLC, Figures 16 and Caruline de Souza Carvalho Machado http://orcid.org/0000-0003-
17(c,d) and hydrogen evolution, (e), Figures 16 and 17(f). 4172-119X
Fernanda Martins Queiroz http://orcid.org/0000-0003-3309-1116
T1 phase precipitation inside the grains, for the T8 and Antonello Astarita http://orcid.org/0000-0003-3214-3375
T851 tempers, explains the intragranular attack observed. Isolda Costa http://orcid.org/0000-0002-4987-3334
However, the mechanism of propagation in these two tem-
pers is different. For the T8, selective dissolution of the
most deformed grains was observed, Figure 10(b,c). For the
References
T851 condition, SLC is related to the selective dissolution of
slip bands. The slip bands are anodic to their surroundings [1] Zou Y, Chen X, Chen B. Corrosion behavior of 2198 Al–Cu–Li
and due to the potential difference between the bands (with alloy in different aging stages in 3.5 wt % NaCl aqueous solution.
J Mater Res. 2018;33:1011–1022. DOI:10.1557/jmr.2018.33.
a great number of T1 phase) and the T1-free matrix, sur- [2] Ma Y, Zhou X, Liao Y, et al. Localised corrosion in AA 2099-T83
rounding the bands, propagation of the corrosion attack aluminium-lithium alloy: The role of grain orientation. Corros Sci.
occurs along the slip bands, Figure 10(e,f). 2016;107:41–48. DOI:10.1016/j.corsci.2016.02.018.
[3] Xu Y, Wang X, Yan Z, et al. Corrosion properties of light-weight
and high-strength 2195 Al-Li alloy. Chin J Aeronaut. 2011;24:681–
Conclusions 686. DOI:10.1016/S1000-9361(11)60080-0.
[4] Ma Y, Zhou X, Thompson GE, et al. Distribution of intermetallics
The results of this work showed that the microstructure and in an AA 2099-T8 aluminium alloy extrusion. Mater Chem Phys.
the mechanisms of corrosion in the AA2198 alloy are strongly 2011;126:46–53. DOI:10.1016/j.matchemphys.2010.12.014.
[5] Cavaliere P, Cabibbo M, Panella F, et al. 2198 Al-Li plates joined
affected by their thermomechanical treatments. The cor-
by Friction Stir Welding: mechanical and microstructural behav-
rosion mechanism related to the T3 thermomechanical treat- ior. Mater Des. 2009;30:3622–3631. DOI:10.1016/j.matdes.2009.
ment is due to galvanic coupling between the cathodic 02.021.
constituent particles and the anodic matrix, leading to tren- [6] Huang J-L, Li J-F, Liu D-Y, et al. Correlation of intergranular cor-
ching and shallow pits. Differently from the T3 thermomech- rosion behaviour with microstructure in Al-Cu-Li alloy. Corros
Sci. 2018;139:215–226. DOI:10.1016/j.corsci.2018.05.011.
anical treatment, the T8 and T851 presented SLC as the main
[7] Ott N, Yan Y, Ramamurthy S, et al. Auger electron spectroscopy
corrosion mechanism. This mechanism was related to intra- analysis of grain boundary microchemistry in an Al-Cu-Li alloy.
granular attack of the grains with high strain, according to Scr Mater. 2016;119:17–20. DOI:10.1016/j.scriptamat.2016.03.021.
their crystallographic directions, and with high T1 phase den- [8] Ye Z-H, Cai W-X, Li J-F, et al. Impact of annealing prior to sol-
sity. The stretching step in the T851 designation modified the ution treatment on aging precipitates and intergranular corrosion
behavior of Al-Cu-Li Alloy 2050. Metall Mater Trans A Phys
features of intragranular attack propagation relatively to the
Metall Mater Sci. 2018;49:2471–2486. DOI:10.1007/s11661-018-
T8. While in the T8 temper, corrosion propagation was 4596-1.
through the slip dislocations, in the T851, the corrosive attack [9] Yan Y, Peguet L, Gharbi O, et al. On the corrosion, electrochem-
propagated through the slip bands resulting from the stretch- istry and microstructure of Al-Cu-Li alloy AA2050 as a function
ing step. of ageing. Materialia. 2018;1:25–36. DOI:10.1016/j.mtla.2018.05.
003.
[10] de Sousa Araujo JV, Donatus U, Queiroz FM, et al. On the severe
localized corrosion susceptibility of the AA2198-T851 alloy.
Acknowledgements
Corros Sci. 2018;133:132–140. DOI:10.1016/j.corsci.2018.01.028.
We thank Olandir V. Correa from IPEN for technical support in DSC [11] Donatus U, Terada M, Ospina CR, et al. On the AA2198-T851
analysis and we are grateful to the ‘Laboratório de Filmes Finos do Insti- alloy microstructure and its correlation with localized corrosion
tuto de Física da Universidade de São Paulo’, Brazil, for the SPM facility. behaviour. Corros Sci. 2018;131:300–309. DOI:10.1016/j.corsci.
2017.12.001.
[12] Rioja RJ, Liu J. The evolution of Al-Li base products for aerospace
Disclosure statement and space applications. Metall Mater Trans A Phys Metall Mater
Sci. 2012;43:3325–3337. DOI:10.1007/s11661-012-1155-z.
No potential conflict of interest was reported by the authors. [13] Abd El-Aty A, Xu Y, Guo X, et al. Strengthening mechanisms,
deformation behavior, and anisotropic mechanical properties of
Al-Li alloys: A review. J Adv Res. 2018;10:49–67. DOI:10.1016/j.
Funding jare.2017.12.004.
[14] Li J, Zhang Z, Zhang J, et al. Review of research on corrosion
The authors are grateful to FAPESP (Proc. 2013/13235-6) for financial behavior of Al-Li alloys. J Chin Soc Corros Prot. 2003;23:319–320.
support for this research, CNPq (2017-9/169569) for the grants of [15] Buchheit RG, Moran JP, Stoner GE. Localized corrosion behavior
João Victor de Sousa Araujo, CNPq (2015-4/133557) for the grants of of alloy 2090 – the role of microstructural heterogeneity.
Aline de Fatima Santos Bugarin, FAPESP (2017/03095-3) for the grants Corrosion. 1990;46:610–617. DOI:10.5006/1.3585156.
of Uyime Donatus, FAPESP (2016/20572-7) for the grants of Fernanda [16] Prasad NE, Ramachandran TR. Phase diagrams and phase reac-
Martins Queiroz and CAPES PROEX (88882.333459/2019-01) for the tions in Al-Li Alloys. Kidlington: Elsevier Inc.; 2014.
grants of Caruline de Souza Carvalho Machado. [17] Deschamps A, De Geuser F, Decreus B, et al. Precipitation in Al-
Cu-Li Alloys: from the kinetics of T1 phase precipitation to micro-
structure development in Friction Stir Welds. 13th International
Data availability statement Conference on Aluminium Alloys; 2012. p. 1145–1154.
DOI:10.1002/9781118495292.ch172.
The raw/processed data required to reproduce these findings [18] Ma Y-L, Zhou X-R, Meng X-M, et al. Influence of thermomechani-
cannot be shared at this time as the data also forms part of an cal treatments on localized corrosion susceptibility and propa-
continuing study. gation mechanism of AA2099 Al–Li alloy. Trans Nonferrous
Met Soc China. 2016;26:1472–1481. DOI:10.1016/S1003-6326
(16)64252-8.
ORCID [19] Huang W, Ma Y, Zhou X, et al. Correlation between localized plas-
tic deformation and localized corrosion in AA2099 aluminum-
João Victor de Sousa Araujo http://orcid.org/0000-0001-6375-0480 lithium alloy. Surf Interface Anal. 2016;48:838–842. DOI:10.
Uyime Donatus http://orcid.org/0000-0001-8871-3571 1002/sia.5817.
12 J. V. DE SOUSA ARAUJO ET AL.

[20] Ma Y, Zhou X, Huang W, et al. Localized corrosion in AA2099- welded 2050-T3 alloy. J Alloys Compd. 2017;722:330–338.
T83 aluminum-lithium alloy: The role of intermetallic particles. DOI:10.1016/j.jallcom.2017.06.141.
Mater Chem Phys. 2015;161:201–210. DOI:10.1016/j.matchemphys. [33] Dorin T, Deschamps A, De Geuser F, et al. Quantitative descrip-
2015.05.037. tion of the T 1 formation kinetics in an Al–Cu–Li alloy using differ-
[21] Li JF, Li CX, Peng ZW, et al. Corrosion mechanism associated with ential scanning calorimetry, small-angle X-ray scattering and
T1 and T2 precipitates of Al-Cu-Li alloys in NaCl solution. J Alloys transmission electron microscopy. Philos Mag. 2014;94:1012–
Compd. 2008;460:688–693. DOI:10.1016/j.jallcom.2007.06.072. 1030. DOI:10.1080/14786435.2013.878047.
[22] Zou Y, Chen X, Chen B. Corrosion behavior of 2198 Al–Cu–Li [34] Qin H, Zhang H, Wu H. The evolution of precipitation and micro-
alloy in different aging stages in 3.5 wt% NaCl aqueous solution. structure in friction stir welded 2195-T8 Al-Li alloy. Mater Sci Eng
J Mater Res. 2018:33:1–12. A. 2015;626:322–329. DOI:10.1016/j.msea.2014.12.026.
[23] Proton V, Alexis J, Andrieu E, et al. Characterisation and under- [35] Wu YP, Ye LY, Jia YZ, et al. Precipitation kinetics of 2519A alumi-
standing of the corrosion behaviour of the nugget in a 2050 alu- num alloy based on aging curves and DSC analysis. Trans
minium alloy Friction Stir Welding joint. Corros Sci. Nonferrous Met Soc China (English Ed). 2014;24:3076–3083.
2013;73:130–142. DOI:10.1016/j.corsci.2013.04.001. DOI:10.1016/S1003-6326(14)63445-2.
[24] Hansen N, Jensen DJ. Development of microstructure in FCC [36] Proton V, Alexis J, Andrieu E, et al. The influence of artificial age-
metals during cold work. Philos Trans R Soc A Math Phys Eng ing on the corrosion behaviour of a 2050 aluminium-copper-
Sci. 1999;357:1447–1469. DOI:10.1098/rsta.1999.0384. lithium alloy. Corros Sci. 2014;80:494–502. DOI:10.1016/j.corsci.
[25] Gable BM, Zhu AW, Csontos AA, et al. The role of plastic defor- 2013.11.060.
mation on the competitive microstructural evolution and mechan- [37] Ghosh R, Venugopal A, Narayanan PR, et al. Environmentally
ical properties of a novel Al-Cu-Li-X alloy. J Light Met. 2001;1:1– assisted cracking resistance of Al–Cu–Li alloy AA2195 using
14. DOI:10.1016/S1471-5317(00)00002-X. slow strain rate test in 3.5% NaCl solution. Trans Nonferrous
[26] Lin Y, Zheng ZQ, Li SC. Effect of solution treatment on micro- Met Soc China (English Ed). 2017;27:241–249. DOI:10.1016/
structures and mechanical properties of 2099 Al-Li alloy. Arch S1003-6326(17)60028-1.
Civ Mech Eng. 2014;14:61–71. DOI:10.1016/j.acme.2013.07.005. [38] Schmutz P. Characterization of AA2024-T3 by Scanning Kelvin
[27] Godfrey A, Hansen N, Juul Jensen D. Microstructural-based Probe Force Microscopy. J Electrochem Soc. 2006;145:2285–
measurement of local stored energy variations in deformed metals. 2295. DOI:10.1149/1.1838633
Metall Mater Trans A Phys Metall Mater Sci. 2007;38 A:2329– [39] Blanc C, Lavelle B, Mankowski G. The role of precipitates enriched
2339. DOI:10.1007/s11661-007-9279-2. with copper on the susceptibility to pitting corrosion of the 2024
[28] Prasad KS, Prasad NE, Gokhale AA. Microstructure and precipi- aluminium alloy. Corros Sci. 1997;39:495–510. DOI:10.1016/
tate characteristics of aluminum-lithium alloys. Kidlington: S0010-938X(97)86099-4.
Elsevier Inc.; 2013. [40] MacRae CM, Hughes AE, Laird JS, et al. An Examination of the
[29] Li JF, Zheng Z, Ren W, et al. Simulation on function mechanism of composition and microstructure of coarse intermetallic particles
T1 (Al2CuLi) precipitate in localized corrosion of Al-Cu-Li alloys. in AA2099-T8, Including Li Detection. Microsc Microanal.
Trans Nonferrous Met Soc China. 2006;16:1268–1273. 2018:24:1–17.
[30] Thompson B, Noble GE. T1 (Al2CuLi) precipitation in aluminum- [41] Ma Y, Zhou X, Huang W, et al. Crystallographic defects induced
copper-lithium alloys. Met Sci. 1972;1:167–174. localised corrosion in AA2099-T8 aluminium alloy. Corros Eng
[31] Sidhar H, Mishra RS. Aging kinetics of friction stir welded Al-Cu- Sci Technol. 2015;50:420–424. DOI:10.1179/1743278214Y.
Li-Mg-Ag and Al-Cu-Li-Mg alloys. Mater Des. 2016;110:60–71. 0000000237.
DOI:10.1016/j.matdes.2016.07.126. [42] Zhang X, Zhou X, Hashimoto T, et al. Corrosion behaviour of
[32] Sidhar H, Mishra RS, Reynolds AP, et al. Impact of thermal man- 2A97-T6 Al-Cu-Li alloy: The influence of non-uniform precipi-
agement on post weld heat treatment efficacy in friction stir tation. Corros Sci. 2018;132:1–8. DOI:10.1016/j.corsci.2017.12.010.

View publication stats

You might also like