You are on page 1of 10

Materials Science & Engineering A 586 (2013) 418427

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

On the role of microstructure in governing fracture behavior


of an aluminumcopperlithium alloy
B. Decreus a,b, A. Deschamps a,n, P. Donnadieu a, J.C. Ehrstrm b
a
b

SIMAP, INP Grenoble CNRS UJF, BP 75, 38402 St Martin d'Hres Cedex, France
Constellium CRV, Voreppe Research Center, BP 27, 38341 Voreppe Cedex, France

art ic l e i nf o

a b s t r a c t

Article history:
Received 26 April 2013
Received in revised form
24 May 2013
Accepted 18 June 2013
Available online 6 July 2013

The inuence of precipitate microstructure on fracture mechanisms is studied in a recently developed


AlCuLi alloy, AA2198. The intra-granular and inter-granular microstructures are varied independently
by changing the quench rate from the solution heat treatment, the amount of pre-stretching and the heat
treatment time. Fracture toughness is evaluated by short bar chevron tear tests that make possible to
evidence clearly the mechanisms of inter-granular fracture. It is shown that intergranular ductile fracture
signicantly occurs in all conditions of heat treatment where substantial precipitation has taken place.
This mechanism is mainly controlled by the state of inter-granular precipitation and plays a major role to
determine the value of transverse fracture toughness, while the strength and ductility of the alloy are
mainly controlled by the state of intra-granular precipitation.
& 2013 Elsevier B.V. All rights reserved.

Keywords:
AlCuLi
Precipitation
Intergranular fracture

1. Introduction
Among precipitation hardening aluminium alloys, AlCuLi
alloys possess a combination of properties that has made them
attractive for structural applications, especially in the aerospace
sector. Developments conducted in the eighties and early nineties
were based on the main factor that the addition of Li decreases the
alloy density and increases the modulus [1,2]. Early alloys with
high lithium content relative to copper content have found limited
applications, due to issues related to ductility losses during long
term ageing at relatively low temperatures [3], insufcient damage
tolerance or high costs of processing. However, in the context of
the competition between Al alloys and composite materials for aircraft structures, and durably high fuel prices, new alloy development has led to alloys with optimised compositions where most of
the issues of former generation alloys have been overcome [47].
With the right combination of copper and lithium contents a more
efcient precipitation strengthening can be reached, particularly
with the T1 phase that forms very efcient obstacles to dislocation
motion because of its thin platelet shape of high aspect ratio [8
11]. These alloys nd a number of applications in new airplane
programs, for instance under the commercial name AIRWAREs
[12]. They attract currently a strong interest in the materials
science community on all aspects of primary and secondary

Corresponding author. Tel.: +33 4 76 82 66 07; fax: +33 4 76 82 66 44.


E-mail address: alexis.deschamps@grenoble-inp.fr (A. Deschamps).

0921-5093/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2013.06.075

processing, microstructure development and various properties


[1325].
The ternary AlCuLi system experiences a complex precipitation sequence, exhibiting aspects of both binary AlCu and AlLi
systems (see [26] for a more complete description). The (standard)
addition of minor alloying elements such as Mg and Ag provides
additional complexity. The binary sequence of AlCu leads to
single atomic layer GPI zones as well as GPII and precipitates,
whereas the AlLi sequence gives the (Al3Li) phase, the formation of which is known to depend strongly on the Li content. Heat
treating the ternary AlCuLi system can generate a number of
phases, the one offering the highest strength being the T1 phase
(nominally Al2CuLi). However, depending on the alloy composition
and processing conditions other minor phases like T2 and TB can
be precipitated, particularly at the grain boundaries [27]. The
presence of Mg can lead to the formation of precipitates of the S
(Al2CuMg) sequence.
One of the key issues that has been recognised since the early
development of AlLiCu alloys is their propensity to intergranular
fracture. In alloys containing of the order of 1.82.2 wt% Li (e.g.
AA8090, AA2090, AA2091), intergranular decohesion is commonly
observed [2835] in various microstructure conditions, from very
underaged to later stages of ageing. Classically a mixture between
ductile intergranular (with shallow dimples) and brittle intergranular (featureless except some trace of intergranular precipitates)
fracture is observed. In these alloys the common feature is the
presence of a high volume fraction of -Al3Li ordered precipitates
in the matrix, except in a precipitate-free zone close to the grain
boundaries. Other precipitates include S and T1 phases within the

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

grains and either phase [28] or Li and Cu containing phases


(T2 and TB) [36] at the grain boundaries. Different causes have
been invoked to explain this extensive occurrence of intergranular
fracture and have received extensive attention particularly in the
1990s, and have been reviewed in several papers ([29,30,32,34,35]
and references therein): presence of intergranular precipitates
associated to the presence of precipitate free zones (particularly
in terms of -Al3Li precipitation), planar slip (due to the collective
shearing of precipitates), embrittlement of grain boundaries by
liquid metal impurities, or a loss of grain boundary coherency due
to Li segregation. The rst mechanism was deemed more important in cases where alloys exhibit ductile intergranular fracture, on
the basis of the experimental evidence that toughness and
occurrence of intergranular fracture were well correlated to the
grain boundary precipitate microstructure [28]. Conversely, the
last hypothesis seems nowadays to be favoured in the cases where
brittle intergranular fracture is observed [34,35].
However, most recently developed alloys such as AA2198 or
AA2050 have a Li content limited to lower concentrations (of the
order of 1 wt%), in order to minimise or even suppress the
formation of the phase. These alloys are completely free of
and their intragranular microstructure in the T8 aged condition is
dominated by the T1 phase and, to a lesser extent, the phase
[26]. In such alloys, pronounced intergranuar fracture is still
observed in the aged T8 temper [23,24]. Similarly, other alloys
with relatively low Li concentration such as AA2020, which are
exempt of precipitation have been also shown to be subject to
intergranular fracture and delamination [37]. However, the relationship between the microstructure of these low-Li containing
alloys and their fracture mode has not been studied in great detail.
The limited number of studies concerning the fracture behaviour
of these recent alloys [23,24] have reported mechanical testing in
the rolling plane and no detailed microstructural analysis. Slanted
fracture is promoted by the intergranular fracture mode, and a
brous fracture surface is observed, related to the unrecrystallized
grain structure, and related very high grain aspect ratio. The shear
component of deformation makes it difcult to determine precisely the microstructural features linked with the intergranular
fracture mechanism. Although it has been proposed that reducing
the Li content may be an efcient way to inhibit Li segregation to
the grain boundaries [35], it is not clear if these alloys are still
subjected to this mechanism. In the high Li containing alloys, most
authors have used short transverse testing in order to evidence the
mechanisms of intergranular fracture [3134]. Using such testing

Fig. 1. Optical micrograph of the grain structure.

419

conditions has proven to be a very efcient way to identify clearly


the intergranular fracture mode.
Following a detailed study of the microstructure development
inside the grains and related mechanical properties during heat
treatment of the AA2198 alloy that has been published recently
[26,38], the aim of the present paper is to evaluate the evolution
of fracture mechanisms and related fracture toughness in a large
variety of heat treatments where different parameters are varied
acting on the microstructure at the grain boundaries (quench rate
from the solution treatment temperature, subsequent ageing time)
and on the intra-granular microstructure (pre-stretch and ageing
time). The mechanical properties are evaluated by simple tensile
tests and by short transverse chevron notch (short bars) samples,
giving access to short transverse fracture toughness and related
fracture surfaces. The microstructure at the grain boundaries
is characterised by electron microscopy and the intra-granular
microstructure has been characterised in the formerly published
work [26].

2. Material and experimental methods


Alloy AA2198 has the composition range (all in wt%) [2.93.5]
Cu[0.81.1]Li[0.250.8]Mg[0.10.5]Ag[0.040.18]Zr. It was
provided by Constellium Voreppe Research Centre, France, as
rolled sheet of 12 mm thickness with a fully brous grain structure. The sheets showed a strong Brass texture (112 {110}). Fig. 1
shows an optical micrograph of the grain microstructure.
The heat treatments consisted rst in a solution treatment and
a quench. Unless stated otherwise, the samples were quenched
into cold water. Some samples were also quenched into hot water
(80 1C) or air cooled, in order to vary the grain boundary microstructure prior to the articial ageing treatment. The samples were
then stretched at least 2% plastic strain and kept several weeks at
ambient temperature (T351 temper). The articial ageing treatment consisted in a heating ramp to 155 1C at 20 K h  1, followed
by an isothermal hold at 155 1C.
Tensile tests were carried out in the rolling (L) direction of the
plates using at tensile tests of section 5  3 mm2 and gauge
length 60 mm, at a constant strain rate of 1.5  104 s  1.
Short bar chevron tear tests were carried out with a loading
direction in the short transverse (ST) direction of the plates with
sample dimensions (in mm) of 19  12.5  11 corresponding to
B  W  2H in Fig. 2. The tests were carried out at a traverse speed
of 0.5 mm min  1 with a pre-load of 5 N, according to ASTM
E1304-97 [39]. The conditional plane strain toughness KQv was

Fig. 2. Geometry of the short bar chevron notched sample.

420

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

where Q stands for the conditional nature of the measured


toughness, and M states for the determination from the maximum
load. Y nm is the tabulated minimum stress intensity factor (depending on the sample geometry), and Fm is the maximum load
measured during the test. Some conditions must be met to
calculate this toughness value [39]. All the results presented here,
except for the observation of fracture surfaces, were obtained in
valid conditions; other results are omitted.
Fracture surfaces were observed using a standard scanning
electron microscope LEO Stereoscan 440 at 20 kV. High resolution
images were obtained on a Zeiss Ultra-55 FEG-SEM at 4 kV with
in-lens secondary and back scattered detectors.
Conventional transmission electron micrographs were obtained
on electropolished samples with a Jeol 3010 microscope operating
at 300 kV. Complementary observations for the grain boundary
microstructure were performed in scanning transmission electron
microscopy (STEM) mode on a FEI Titan 80300 equipped a high
angle annular dark eld (HAADF) detector.

calculated as follows:
Y nm F m

K Q vM p
B W

600
16h

100h

True stress (MPa)

500

500h

8h

4h

400
1h

300
200
100
0

3. Tensile tests

10

12

14

16

True strain (%)


Fig. 3. Tensile curves as a function of ageing time at 155 1C.

Fig. 3 shows the tensile curves obtained for samples heat


treated from 1 h to 500 h at 155 1C. The microstructure evolution
during this ageing treatment has been described in detail in [26].

Fig. 4. Fracture surfaces of tensile test samples (SEM micrographs) at low and high resolution for two different states of ageing: (a) and (b) 4 h at 155 1C; (c) and (d) 16 h at
155 1C.

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

500

600

70
16h

4h

Load (N)

1h

300
100h
500h

200

100

0.5

1.5

Displacement (mm)

Yield stress

60

500

50
Toughness

40

400
300

30

Elongation

20

100

10
0

200

Yield stress (MPa)

KqvM (MPa.m-1/2), 4*U.A. (%)

8h

400

421

10

100

0
1000

Time at 155C

Fig. 5. (a) Loaddisplacement curves for the short bar tests on samples aged for different times at 155 1C. (b) Values of toughness measured from the short bar tests along
with corresponding yield stress and uniform elongation of tensile tests in the same ageing conditions.

Fig. 6. Low resolution fracture surfaces (SEM micrographs) of the notch tip of the short bar tests for the following ageing times at 155 1C: (a) 1 h; (b) 8 h; (c) 16 h; (d) 100 h.
An increase of the fraction of at intergranular fracture is observed when the ageing time proceeds.

After 1 or 4 h at 155 1C, the microstructure consists mainly of a


homogeneous solid solution, and no signicant fraction of the
strengthening T1 precipitates is present. After 8 h at 155 1C, the
formation of the T1 phase is really signicant and the microstructure between 16 and 500 h is mostly stable, consisting on a
fraction close to equilibrium of very ne T1 platelets resulting in
an almost constant yield strength. Fig. 4 shows the fracture
surfaces of the tensile specimens tested in two ageing conditions,
namely in the absence of T1 precipitates (4 h (a, b)) and near peak

strength (16 h (c,d)). In the rst case, fracture appears to be


completely ductile trans-granular. The dimples are oriented due
to the slanted geometry of the fracture surface. In the second case,
the fracture surface appears as multiple ne steps of a few mm,
with smooth walls between them. These fractographs resemble
that obtained in similar conditions by Chen et al. [23] and Steglich
et al. [24] and are characteristic of a fracture controlled by
intergranular decohesion with a heavily elongated grain structure
in the loading direction. Due to shearing and friction during the

422

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

Fig. 7. Detail of ductile fracture area of the short bar sample aged 1 h at 155 1C: (a) in secondary electron mode showing the presence of dimples and (b) in back-scattered
electron mode showing the presence of large intermetallics associated with the dimples.

Fig. 8. Detail of ductile fracture area of the short bar sample aged 100 h at 155 1C: (a) low resolution image in back-scattered electrons mode showing no large scale
intermetallics on the intergranular surface; (b) medium resolution secondary electrons image showing the presence of a high density of small dimples on the at
intergranular fracture surface; (c) high resolution secondary electrons image showing the presence of small particles in these dimples.

last stages of the fracture process, it is impossible to describe the


microscopic mechanisms that explain this occurrence of intergranular fracture from these observations. Short bar chevron tests

presented in the next paragraph allow for loading in the direction


normal to the grain boundaries and observing the details of the
fracture mechanisms.

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

423

Fig. 9. Transmission electron micrographs showing the microstructure at the grain and sub-grain boundaries. (a) Dark-eld micrograph showing a sub-grain boundary in the
sample aged for 16 h at 155 1C; (b) dark-eld micrograph showing a grain boundary in the sample aged for 16 h at 155 1C; (c) and (d) STEM-HAADF micrographs of grain
boundaries in the sample aged for 100 h at 155 1C.

4. Chevron notch short bar tests


Fig. 5(a) shows the loaddisplacement curves for the chevron
notch short bar tests after different heat treatment times. For the
samples heat treated 1 h and 4 h at 155 1C where the material is
still very soft, generalised plasticity prevented the measurement of
fracture toughness from these tests. However, it appears clearly
that the behaviour changes dramatically between 4 and 8 h, then
between 8 h and 16 h, and again between 16 h and 100 h, after
which it remains stable. Fig. 5(b) shows the evolution with heat
treatment time of three parameters of the mechanical tests,
namely the yield strength and uniform elongation obtained from
the tensile tests, and the fracture toughness obtained from the
short bar tests. As evidenced in [26], the evolution of yield stress is
essentially representative of the evolution of the volume fraction
of T1 precipitates inside the grains. When T1 precipitates rst
appear the yield stress starts to increase, and when this fraction
saturates, the yield stress stabilises around 500 MPa. The evolution
of elongation is quite well correlated to this evolution of yield
stress, as discussed in detail in [38]. The conjunction of the
increase in yield strength and the corresponding reduction of the
strain hardening rate capability results in a reduction in uniform
elongation. Interestingly, the evolution of fracture toughness is
quite different from that of the uniform elongation. The toughness
also decreases with heat treatment time, but this decrease occurs
later. In the rst stages of the increase of yield strength, the
strength increase compensates the reduction in elongation to keep
the toughness roughly constant. It is only after 50 h at 155 1C that

the toughness decreases signicantly, while the yield strength and


elongation (albeit tested with a different loading direction) are
now constant with ageing time.
In order to understand the origin of this toughness drop during
the peak strength plateau, fractographic observation of the short
bar samples will now be presented.
Fig. 6 presents low resolution images of the fracture surfaces
close to the notch tip, for different heat treatment times (1 h, before
the increase in strength; 8 h, during the strength increase; 16 h
close to peak strength before the drop in toughness; and 100 h at
peak strength after the drop in toughness). In the rst sample the
fracture is entirely ductile transgranular. No at areas are recorded.
After 8 h and 16 h at 155 1C, a large majority of the fracture surface
is still ductile transgranular; however a few at and smooth (at this
scale) surfaces can be observed that correspond to interganular
fracture. After 100 h at 155 1C, most of the fracture surface consists
of at and smooth intergranular zones. It is therefore tempting to
associate the decrease in fracture toughness to the occurrence of a
fracture mode of low energy dissipation [40], namely inter-granular
decohesion along grains that have a large extension normal to the
short transverse loading direction.
The details of the fracture surface are shown in Fig. 7 for the
sample aged 1 h at 155 1C, and in Fig. 8 for the sample aged 100 h
at 155 1C. In the rst case, the ductile transgranular fracture is
classically characterised by large dimples, which include broken
intermetallic particles. These particles have a bright contrast
(Fig. 7b) indicating a larger average atomic number than the
matrix that is conrmed by the composition close to Al7Cu2Fe

424

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

70

60

Def 16h
60

Increasing
ageing time

50

KqvM (MPa.m-1/2)

KqvM (MPa.m-1/2)

55

Undef 16h
45
40

Fast quench

50
40

Intermediate quench

30
20

Slow quench

Undef 100h
35
30
100

10

Def 100h

Undeformed
T351
120

140

160

180

200

Microhardness

20

40

60

80

100

120

Time at 155C (h)

Fig. 10. (a) Evolution of the compromise between toughness (as measured by the short bar test) and microhardness in the sample aged at 155 1C from the T351 condition as
compared to samples aged from an undeformed condition; (b) inuence of the rate of quenching on the toughness (as measured by the short bar test) evolution with ageing
time at 155 1C.

given by EDS analysis. In the second case, the situation is entirely


different. The at areas of the fracture surface do not present any
chemical contrast at low resolution (Fig. 8(a)), which means that
no coarse intermetallic particle is associated with this fracture
mechanism. However, at higher resolution a homogeneous distribution of very ne dimples is observed (Fig. 8(b) and (c)). Their
size is between 100 and 200 nm, and very small particles (less
than 50 nm in diameter) are observed at the core of the dimples.
A quantitative chemical analysis of these particles is clearly out of
the range of EDS-SEM, but a qualitative comparison between the
EDS signal on these particles and on the surrounding matrix
clearly shows that they are enriched in Cu and not in Fe. It is not
possible using this technique to estimate a potential enrichment in
Li. Although not shown here for space reason, it should be
mentioned that similar observations were made on all at areas
of short bar samples, with a general tendency of decreasing the
dimple size with increasing the ageing time [41].
These fractographic observations provide good evidence that
the intergranular fracture mode responsible for the at areas on
the short bar test fracture surfaces is actually ductile, with a
damage initiation related to ne Cu-rich particles lying on the
grain boundaries. These Cu-rich particles, less than 100 nm in
diameter, are likely to be intergranular precipitates formed during
the articial ageing treatment. In order to conrm this hypothesis,
TEM observations are reported in Fig. 9. Fig. 9(a) and (b) shows
conventional dark eld micrograph of a sub-grain boundary and a
grain boundary after 16 h at 155 1C (at the beginning of the peak
strength plateau). In the matrix, a high density of ne T1 platelets
is observed. At sub-grain boundaries, this density is even higher,
but the precipitates have a similar shape and size. Since T1
precipitates are known to nucleate on dislocations, sub-grain
boundaries can be regarded as regions of particularly high density
of nucleation sites for the precipitates. At the grain boundaries,
however (Fig. 9(b)), hardly any precipitation can be observed,
which is consistent with the fact that fracture is still mostly
transgranular in the short bar tests. Fig. 9(c) and (d) shows
STEM-HAADF (Z-contrast) images at grain boundaries of the
sample aged 100 h at 155 1C. Inside the grains the microstructure
is qualitatively identical to that of the sample aged 16 h (it can
actually be shown that the two intragranular microstructures are
quantitatively almost identical [26]). However the grain boundaries now include a high density of incoherent particles of size
approximately 50 nm. Their bright contrast in the HAADF micrographs tells that they must be Cu-rich, in agreement with [36] and

with similar observations in [13]. These precipitates may be T2 or


TB phases; however their crystal structure was not determined
here. Their size and morphology suit particularly well with that of
the particles observed at the core of the dimples on the fracture
surfaces (Fig. 8(c)).

5. Effect of the variation of intergranular and intragranular


microstructures
The preceding experiments show that the evolution of transverse toughness is correlated with the state of intergranular
precipitation, which in turn controls the occurrence of ductile
intergranular fracture. However, during the studied articial ageing treatment, the intragranular and intergranular microstructures
evolved concurrently, so that unravelling their respective effects
remains challenging.
In order to provide additional evidence, we have sought to
change independently the intergranular and intragranular microstructures in two ways. First, we have compared samples aged in
the classical way (namely quenched, stretched and aged) to
samples where the stretching step was omitted. As shown in
[42], intragranular precipitation in the absence of pre-deformation
is very sluggish and the precipitation kinetics is decreased by a
factor of 10. However it is likely that the precipitation kinetics at
the grain boundaries is at rst order unaffected by the absence of
stretching and thus it becomes possible to compare samples with
different intra-granular precipitates and similar inter-granular
microstructure. Secondly, we have compared samples aged by
the same procedure (including the stretch), but having experienced different quenching conditions from the solution treatment
(intermediate quench, in hot water and slow quench, air cooled).
These samples can thus be expected to have similar intragranular
precipitates but different inter-granular microstructures, with a
higher fraction of intergranular precipitates in the slowly cooled
materials.
Fig. 10(a) shows a graph representing the different sets of
microhardness/toughness parameters for samples aged at 155 1C
in the deformed and undeformed materials. In both cases an
inverse correlation exists, however very different in the two
samples. As seen in the former sections, for the stretched
samples, at near peak strength (16 h) the toughness is still very
high. At this ageing time, the unstretched material shows a
comparable toughness value; however the microhardness is

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

425

Fig. 11. Fracture surfaces (SEM micrographs) of short bar samples on slowly quenched materials: (a) intermediate quench, 100 h ageing; (b)(d) slow quench, 100 h ageing.

much lower due to the sluggish intragranular precipitation


kinetics. After 100 h at 155 1C, the unstretched material has
reached the hardness value of the stretched material aged 16 h,
but has a much lower toughness. Its toughness is now similar to
that of the stretched material aged 100 h at 155 1C. Therefore all
these results are consistent with a hardness controlled by the
state of intragranular precipitation and a toughness mostly
controlled by the inter-granular precipitation (which is at rst
order controlled by the ageing time).
Fig. 10(b) shows the evolution with heat treatment time of
toughness for samples quenched with different procedures (fast,
intermediate and slow). Clearly, decreasing the quench rate
reduces very signicantly the toughness value in all ageing
conditions. The remaining evolution is more complex, with a
convergence of the intermediate and fast quench samples at long
ageing time, while the slowly quenched material keeps a lower
toughness at any ageing time. The fracture surfaces of the short
bar tests after 100 h at 155 1C for the intermediate or slow
quenched materials are shown in Fig. 11. They present a high
fraction of at intergranular areas, with a high density of small
dimples (albeit larger than that of the fast quenched material,
presumably due to a larger size of intergranular precipitates), and
some areas of facetted brittle fracture, which could correspond to
the formation at the grain boundaries of a continuous layer of a
brittle incoherent intermetallic phase after a sufciently slow
quench.

6. Discussion
As summarised in the Section 1, the fracture behaviour of Al
LiCu alloys containing of the order of 2 wt% Li has been extensively investigated in the literature. These alloys have been shown
to be prone to intergranular fracture, often of brittle nature. This
behaviour was shown to be strongly related to the presence of Li
by several mechanisms that depend on the alloy composition (and
related precipitates formed) and heat treatment, namely Li segregation, grain boundary precipitates, zones free of precipitates at
the grain boundaries, and planar slip due to the ordered phases.
In the more recently developed alloys where the Li content is
limited to about 1 wt% in order to suppress the formation of to
the benet of the T1 phase, the literature on the fracture mechanisms is still incomplete, although intergranular fracture has also
been observed in such alloys to play a prominent role. The
microstructural observations made in the present paper and in
the two companion papers published earlier on the same alloy
[26,38], as well as the fractographic observations reported here,
show that these alloys have a very different microstructure as
compared to their higher Li counterparts. This is especially
important with respect to the mechanisms that have been invoked
to explain the fracture behaviour. First, they do not show any
precipitation, nor any precipitate-free zone (PFZ). After ageing, the
T1 platelets have an extension of more than 50 nm, which helps
to explain the absence of PFZ because any precipitate nucleated

426

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

at a distance smaller than 25 nm from a grain boundary can


extend to the grain boundary provided that some Cu and Li
supersaturation still remains. Secondly, although they are prone
to planar slip in the naturally aged condition due to the presence
of solute clusters, when aged, their slip becomes quite homogeneous [38], because of the particular shearing mechanism of the T1
precipitate, which is very different from that of . Thirdly, no
evidence of brittle intergranular fracture has been found in the
present study for all the heat treatments investigated, except for
the slowly quenched material. However in the latter case, this
brittle fracture was facetted and did not follow the shape of the
grains, so that we believe that it is related to the decohesion along
the interface of large facetted precipitates. In all other cases the
macroscopically at intergranular fracture surfaces included a high
density of very small dimples (size of about 100 nm), associated
with precipitates whose size and composition were consistent
with that observed at the grain boundaries in the transmission
electron microscope, namely Cu-rich, of average atomic number
higher than the matrix as proven by the higher brightness in Zcontrast.
Therefore, it can be expected that the mechanisms inducing
intergranular fracture may be of a different nature in the high-Li
and low-Li containing alloys. Although we cannot rule out completely a loss of the grain boundary cohesion due to the presence
of Li in the alloy, a good indication that such decohesion does not
play a major role in the low-Li containing alloy investigated here is
that we have not observed brittle fracture. In addition the values of
toughness are quite high for transverse toughness tests in precipitation hardening aluminium alloys, which is a further indication that grain boundary cohesion is not affected at rst order.
Ductile grain boundary fracture is usually related to three parameters [28,43]: the presence of grain boundary precipitates, a soft
precipitate free zone, and a high stress triaxiality. In the present
case no signicant precipitate free zone is visible. However the
competition between intergranular and transgranular fracture
appears to be affected by the trixiality ratio, since grain boundary
fracture is systematically observed in the areas of highest triaxiality of the chevron notch specimen (close to the initial notch).
This effect of external triaxiality may actually help the accelerated
void growth in such conditions where no PFZ is present.
The evolution of fracture mechanism as well as the evolution of
toughness with the different microstructural states evaluated here
are consistent with the proposed mechanism controlling intergranular fracture, namely nucleation and growth of voids at the
grain boundary precipitates. The toughness drops simultaneously
to the increase in the fraction of intergranular fracture when
ageing proceeds, and this happens independently (at rst order) of
the intragranular precipitation kinetics, which was varied by
playing on the plastic deformation prior to the heat treatment.
Similarly, a slower quench rate from the solution treatment, which
favours a high fraction of grain boundary precipitates, results in a
lower fracture toughness.

of intragranular and intergranular microstructure on the strength,


the uniform elongation in tension, and transverse fracture toughness as well as related fractography.
While the strength and uniform elongation are mainly controlled by the state of precipitation within the grains, we have
strong indications that the transverse fracture toughness is largely
determined by the occurrence of intergranular fracture, which is
controlled by the presence of incoherent, Cu rich grain boundary
precipitates. These precipitates serve as damage nucleation sites
and the resulting fracture surfaces present a high density of submicrometric dimples. In practice, the choice of a heat treatment
at the very beginning of the peak strength plateau, where the
intergranular precipitation microstructure is not yet fully developed, makes possible to avoid a massive appearance of intergranular fracture and thus keep a high toughness value. Particular
care must yet be taken to ensure a sufciently fast quench rate
from the solution treatment so that the grain boundary microstructure prior to articial ageing is free of precipitates.

Acknowledgements
C. Sigli is thanked for fruitful discussions. L. Charpenay is
thanked for helping with the short bar tests. The French research
agency (ANR) is thanked for nancial support under the project
ALICANTDE. We are grateful to the Canadian Centre for Electron
Microscopy, a facility funded by the Canada Foundation for
Innovation and the Ontario Government where the HAADF STEM
work was carried out.
References
[1]
[2]
[3]
[4]
[5]
[6]

[7]

[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]

7. Conclusions
In the present work we have attempted to shed light on the
fracture mechanisms that prevail in the recently developed AlCu
Li alloys with moderate Li content where the microstructure after
articial ageing is dominated by the precipitation of the T1 phase
and where no phase is formed. In order to study the microscopic
mechanisms related to intergranular fracture, which is frequently
met in these alloys, testing has been performed in the short
transverse direction. By studying the effect of heat treatment time,
pre-stretch before ageing, and quench rate from the solution
treatment, we have been able to evaluate independently the effect

[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]

E. Starke Jr., J.T. Staley, Prog. Aerosp. Sci. 32 (1996) 131.


F.W. Gayle, F.H. Heubaum, J.R. Pickens, Scr. Metall. Mater. 24 (1990) 79.
S. Katsikis, B. Noble, S.J. Harris, Mat. Sci. Eng. A 485 (2008) 613.
T. Warner, Mat. Sci. Forum 519521 (2006) 1271.
P. Lequeu, K.P. Smith, A. Danielou, J. Mater. Eng. Perform. 19 (2010) 841.
A. Danielou, J.P. Ronxin, C. Nardin, J.C. Ehrstrm, Proceedings of the 13th
International Conference on Aluminium Alloys, TMS, Warrendale, USA,
Pittsburgh, PA, 2012, pp. 511516.
J. Boselli, G. Bray, R.J. Rioja, D. Mooy, G. Venema, G. Feyen, W. Wang,
Proceedings of the 13th International Conference on Aluminium Alloys, TMS,
Warrendale, USA, Pittsburgh, PA, 2012, pp. 581586.
J.C. Huang, A.J. Ardell, Mater. Sci. Technol. 3 (1987) 176.
J.M. Howe, J. Lee, A.K. Vasudvan, Metall. Trans. A 19A (1988) 2911.
P. Donnadieu, Y. Shao, F. De Geuser, G.A. Botton, S. Lazar, M. Cheynet, M. de
Boissieu, A. Deschamps, Acta Mater. 59 (2011) 462.
C. Dwyer, M. Weyland, L.Y. Chang, B.C. Muddle, Appl. Phys. Lett. 98 (2011)
201909.
T. Warner, J.C. Ehrstrm, B. Chenal, F. Eberl, Light Met. Age 67 (2009) 3.
N. Brodusch, M. Trudeau, P. Michaud, L. Rodrigue, J. Boselli, R. Gauvin, Microsc.
Microanal. 18 (2012) 1393.
N. Ward, A. Tran, A. Abad, E.W. Lee, M. Hahn, E. Fordan, O.S. Es-Said, J. Mater.
Eng. Perform. 20 (2011) 989.
R.J. McDonald, A.J. Beaudoin, Model. Simul. Mater. Sci. Eng. 18 (2010) 065007.
S. Kalyanam, A.J. Beaudoin, R.H. Dodds Jr., F. Barlat, Eng. Fract. Mech. 76 (2009)
2174.
I. Eberl, C. Hantrais, J.-C. Ehrtsrom, C. Nardin, Sci. Technol. Weld. Join. 15 (2010)
699.
S. Richard, C. Sarrazin-Baudoux, J. Petit, Proceedings of the TMS 2009,
San Francisco, Minerals, Metals & Materials Soc, 2009, pp. 6976.
A. Astarita, A. Squillace, A. Scala, A. Prisco, J. Mater. Eng. Perform. 21 (2012)
1763.
C. Bitondo, U. Prisco, A. Squilace, P. Buonadonna, G. Dionoro, Int. J. Adv. Manuf.
Technol. 53 (2011) 505.
P. Cavaliere, J. Light Met. 2 (2002) 247.
P. Cavaliere, M. Cabibbo, F. Panella, A. Squillace, Mater. Des. 30 (2009) 3622.
J. Chen, Y. Madi, T.F. Morgeneyer, J. Besson, Comput. Mater. Sci. 50 (2011) 1365.
D. Steglich, H. Wafai, J. Besson, Eng. Fract. Mech. 77 (2010) 3501.
Y.E. Ma, Z. Zhao, B. Liu, W. Li, Mater. Sci. Eng. A-Struct. Mater. Prop.
Microstruct. Process 569 (2013) 41.
B. Decreus, A. Deschamps, F. De Geuser, P. Donnadieu, C. Sigli, M. Weyland,
Acta Mater. 61 (2013) 2207.
S.C. Wang, M.J. Starink, Int. Mater. Rev. 50 (2005) 193.
A.K. Vasudvan, R.D. Doherty, Acta Metall. 35 (1987) 1193.
S. Lynch, Mater. Sci. Eng. A. 136 (1991) 25.

B. Decreus et al. / Materials Science & Engineering A 586 (2013) 418427

[30] S. Lynch, A. Wilson, R. Byrnes, Mater. Sci. Eng. A. 172 (1993) 79.
[31] C.G. Bennet, S.P. Lynch, R.B. Nethercott, M. Kerr, E.D. Sweet, Mat. Sci. Eng.
(1998) 32.
[32] S.P. Lynch, B.C. Muddle, T. Pasang, Acta Mater. 49 (2001) 2863.
[33] S.P. Lynch, B.C. Muddle, T. Pasang, Philos. Mag. A 82 (2002) 3361.
[34] A.C. Bregianos, A.G. Crosky, P.R. Munroe, A.K. Hellier, Int. J. Fract. 161 (2010)
141.
[35] T. Pasang, N. Symonds, S. Moutsos, R.J.H. Wanhill, S.P. Lynch, Eng. Fail. Anal. 22
(2012) 166.
[36] N. Eswara Prasad, K.S. Prasad, S.V. Kamat, G. Malakondaiah, Eng. Fract. Mech.
51 (1995) 87.

427

[37] K.V. Jata, A.K. Vasudevan, Mater. Sci. Eng. A 241 (1998) 104.
[38] A. Deschamps, B. Decreus, F. De Geuser, T. Dorin, M. Weyland, Acta Mater.
61 (2013) 4010.
[39] ASTM-International, Standard Test Method for Plain-strain (chevron-notch)
Fracture Toughness of Metallic Materials, 2002.
[40] D. Dumont, A. Deschamps, Y. Brechet, Acta Mater. 52 (2004) 2529.
[41] B. Decreus, Ph.D. thesis, Institut Polytechnique de Grenoble, France, 2010.
[42] B. Decreus, A. Deschamps, F. De Geuser, C. Sigli, Advanced Engineering
Materials http://dx.doi.org/10.1002/adem.201300098, in press.
[43] T. Pardoen, D. Dumont, A. Deschamps, Y. Brechet, J. Mech. Phys. Solids 51
(2003) 637.

You might also like