Vous êtes sur la page 1sur 12

Acta Materialia 52 (2004) 25292540

www.actamat-journals.com

A model for predicting fracture mode and toughness


in 7000 series aluminium alloys
D. Dumont
a

a,b

, A. Deschamps

a,*

, Y. Brechet

LTPCM ENSEEG, CNRS UMR 5614, Domaine Universitaire de Grenoble, INPG, BP 75, Saint Martin dHeres Cedex 38 402, France
b
Pechiney Centre de Recherches de Voreppe, 725 rue Aristide Berges, BP 27, Voreppe Cedex 38 341, France
Received 19 December 2003; received in revised form 30 January 2004; accepted 31 January 2004
Available online 5 March 2004

Abstract
A model is proposed, which predicts the toughness of 7000 series aluminium alloys in a variety of situations, including two alloy
compositions, dierent quench rates from the solution treatment temperature and ageing states from underaged to overaged. The
model is derived in three steps. The energy dissipated by transgranular fracture is rst calculated, using a simplied cohesive zone
approach. The energy dissipated by intergranular fracture is then calculated using a critical strain criterion, and the total dissipated
energy is then estimated using an averaging by the respective area fractions of the two modes, which are themselves dependent on the
respective energies of the two main fracture mechanisms. The model input parameters are the materials mechanical properties such
as yield stress and strain-hardening rate, geometrical features and related properties of the precipitate-free zones, and area fraction
of the grain boundaries covered with precipitates. The model predicts all the main features of the evolution of the toughness/yield
strength compromise with changing quench rate or ageing treatment. It allows to predict the evolution of toughness when dramatic
changes in the occurrence of fracture modes are observed. Finally, using the model it is possible to predict the eect of changes in
individual parameters on the overall fracture behaviour.
 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Toughness; Modelling; Precipitation; Aluminium alloys; 7000 series

1. Introduction
Modelling the fracture toughness in industrial aluminium alloys, in a broad range of situations (including
dierent alloys, states of ageing and quench rates for instance) is a dicult task, owing to the complexity of the
microstructure and the fact that various fracture modes
may contribute simultaneously to the nal rupture.
In situations of practical relevance, the microstructure contains a number of heterogeneities, which can
contribute both to the localisation of plastic ow, and
to the initiation and propagation of failure. In order of
decreasing size, these heterogeneities include intermetallic particles, quench-induced intergranular and in-

Corresponding author. Tel.: +33-4-76-82-66-07; fax: +33-4-76-8266-44.


E-mail address: alexis.deschamps@ltpcm.inpg.fr (A. Deschamps).

tragranular precipitates, and precipitate-free zones


situated at the grain boundaries.
The heterogeneity of the microstructure, as well as
the plastic properties associated with the dierent zones
of the material, induces a complex competition between
dierent failure modes, i.e. intermetallic decohesion,
intergranular fracture and transgranular fracture.
A large number of analytical models exist in the literature, which describe the fracture toughness of a material in the case where one fracture mode is dominant.
These dierent approaches can be reviewed as follows:
 The fracture controlled by void nucleation at second phase particles has been described by Rice and
Johnson [1], Burghard [2], and later by Hahn and Roseneld [3]. These models have yielded the dependence
of fracture toughness on the volume fraction of second
phase particles
KIC / fv1=6 ;

1359-6454/$30.00  2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2004.01.044

2530

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

which has been conrmed by several experimental


studies (see, e.g. [3] or [4]). However, the proportionality
factor predicted by these models did not correspond well
to the experimental facts. It implies a positive correlation
between toughness and yield strength, opposite to what
is generally observed experimentally with 7xxx alloys.
An other approach, by Garett and Knott [5], rened
by Chen and Knott [6] has aimed to describe the fracture
toughness in terms of a critical strain to failure at the
crack tip, associated with a critical opening displacement (COD) (related to the fracture of second phase
particles). This analysis yielded, at constant second
phase particle distribution, to a dependence of the
fracture toughness with the plastic properties of the
material (yield stress ry , hardening exponent n), and
some characteristics of the intermetallic particles (decohesion stress rc , spacing k):
p
KIC / n ry rc k:
2
This proportionality was found to be consistent with a
number of experimental observations in dierent commercial alloys [7], but again the trend predicted for the
inuence of the yield stress is opposite to experimental
ndings.
 A second type of approach describes the toughness
as controlled by the heterogeneity of plastic ow inside
the material, corresponding to the intense shear bands
observed notably in underaged states [8,9]. Such approaches have a limited applicability, since the cases
where the transgranular shear band fracture mode is
dominant are quite rare for 7xxx alloys since they are
mainly used in T6 or T7x tempers. An important group
of models has concentrated on the fracture toughness
controlled by intergranular ductile fracture, which is a
prominent fracture mode in high strength metallurgical
states (e.g., T6-treated material). The rst model taking
into account the localisation of deformation in the
precipitate-free zone (PFZ) is due to Embury and Nes
[10], and predicts a relationship between fracture
toughness and the area fraction of grain boundary
covered by precipitates fGB , through an expression for
the critical strain to failure in the PFZ:
p
ePFZ 1=fGB  1=2;
3
KIC ErR ePFZ 1=2 :

This model has been successfully applied to a number of


experimental situations, including the results of Unwin
and Smith [11] and Vasudevan and Doherty [12].
Hornbogen and Gr
af [13] have adapted the relationship developed by Hahn and Roseneld [3] for bulk
plastic properties to the plastic properties of the PFZ.
They developed an expression of fracture toughness as a
function of the strength, hardening exponent and width
of the PFZ, as well as the grain size D, valid for pure
integranular fracture, which yields a D1=2 dependence,

coherent with the experimental results of numerous


authors [1416]. The Horbogen and Graf model, like the
Embury and Nes approach, relies on an expression of
the critical strain to failure in the PFZ ePFZ . An alternative expression for this parameter has been proposed
by Kawabata and Izumi [17], which depends on several
parameters of the PFZ (precipitate size, density on the
grain boundaries, width of the PFZ).
More recently, Li and Reynolds [18] compared these
approaches and found them very close in terms of their
ability to describe experimental results.
 In the case where several fracture modes are simultaneously present (which is the general case in all
practically relevant situations), a number of authors
have proposed some mixture rules between the fracture
toughness predicted for each of the individual fracture
modes alone. Simple linear averaging has been proposed
by Hornbogen and Graf [13], and more recently by
Sugamata et al. [19], however the linear averaging of
fracture toughness lacks a clear physical meaning, as
mentioned by Kamat and Hirth [20], since the quantities
to be averaged have to be extensive ones, such as the
dissipated energies in the various fracture processes, and
thus the averaging should at least be performed on the
square of the fracture toughness. Such quadratic laws
have been used recently by Deshpande et al. [21] and
Gokhale et al. [22], in order to predict the inuence of
the fraction of recrystallisation fraction on the fracture
toughness.
Another strategy to approach the modelling of fracture toughness in ductile materials is to consider it as a
limiting case, for very high triaxiality, of a damage accumulation process. In this approach, the description of
the dissipated energy is obtained through the analysis of
void nucleation and growth, preferentially at second
phase particles. This mechanistic approach is based on
material-related void nucleation laws [2325]. A number
of models are available for the growth of voids, either
isolated in the material or interacting, the most popular
approach being the one by Gurson [26], later rened by
Tvergaard [27]. Finally, failure models based on the
coalescence of voids enable to predict the failure of the
material (see, e.g. [28]). Although these approaches have
not reached the point where they can predict quantitatively the fracture toughness as a function of microstructural parameters and plastic properties, they have
lead for instance to signicant advances on the understanding of the competition between intergranular and
transgranular fracture depending on the microstructural
features and on the state of triaxiality [29].
In conclusion, a large number of analytical expressions are available, which relate the fracture toughness
to microstructural parameters and mechanical properties of the constituent material. A special interest has
been given to the microstructural heterogeneities such as
PFZs at the boundaries. However one obvious missing

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

(from the solution treatment temperature down to


200 C) ranged from 800 K s1 (fast quench, F) to 7 K s1
(slow quench, S), representative of the quench rate
experienced at mid-thickness of a 150 mm thick plate
quenched into cold water. Following this quench, the
alloys have been subjected to a conventional ageing sequence, including natural ageing, and two step articial
ageing at 120 and 160 C.
2.2. Mechanical properties
The yield strength and strain-hardening characteristics were measured by conventional tensile tests; the tear
resistance was measured on Kahn Tear Test samples
[30], the relevant parameters being the unit initiation
energy, UIE (i.e. the energy dissipated before crack
propagation). Alloy AA7040 was tested in the TL direction, and alloy AA7050 in the LT direction, resulting in an intrinsically higher toughness for the latter.
Toughness tests (on CT specimens) on selected states
showed a good linear relationship between the square of
the critical intensity stress factor and the UIE [30]. Fig. 1
150
7040F
7040S
UA

UIE(N.mm-1)

element of the present state of the art is the description


of the fraction of the various fracture modes, from
which the appropriate averaging has to be performed.
Moreover, very little work exists, which describe the
evolution of fracture toughness of a single alloy with
variable heat treatments in a coherent framework encompassing situations ranging from dominant intergranular fracture to dominant transgranular fracture.
The aim of the present study is to develop a simple
analytical model for situations with competing fracture
modes, both ductile, either intergranular or transgranular. It has to be applicable to a wide range of situations,
from very low to very high toughness values, corresponding to dierent proportions of the two fracture
modes. This model is based on existing equations for the
individual mechanisms, leading to specic expressions
for the dissipation energies of the dierent failure
modes. This model will also propose a description of the
competition between the transgranular and the intergranular failure modes, and its relationship with the
relative dissipation energies.
The results of the model will be compared to an extensive experimental study, which details have been
published elsewhere [30,31]. This experimental study
provides the measurement of tear resistance (using the
Kahn Tear Test) in a wide range of situations (two alloys, dierent quench rates and ageing states). On the
same alloys, quantitative values for microstructural parameters, plastic properties and fractions of failure
modes have been measured, which allow a direct comparison with the present model.
In Section 2, we recall briey the main experimental
ndings. In Section 3, we present the model and its
comparison with the experimental results.

100

UA

OA

(a)

PA

400
450
500
Yield strength (MPa)

550

200

2.1. Alloys and heat treatments

7050F
7050S

UA

150
UIE(N.mm-1)

Two alloys have been studied in terms of the relationship between their microstructure, plastic properties
and tear resistance: alloys AA7050 and AA7040 of respective compositions 6.33% Zn, 2.46% Mg, 2.2% Cu,
0.11% Zr, 0.1% Fe, 0.08% Si and 6.51% Zn, 2.01% Mg,
1.64% Cu, 0.11% Zr, 0.08% Fe, 0.05% Si (all in wt%).
The latter has been specically developed for thick plate
applications, a slightly lower alloying content aiming at
reducing the quench sensitivity and thus improving the
toughness/yield strength compromise. One aim of the
present study is to determine which microstructural
parameters control this compromise, and what should
be the direction in which additional alloy development
should go in order to improve it further.
These alloys were subjected to dierent quench rates
after solution treatment at 483 C, which initial rate

PA

OA

50

0
350

2. Main experimental results

2531

100

OA
PA

50

UA

OA

0
350
(b)

PA

400
450
500
Yield strength (MPa)

550

Fig. 1. Plots of unit initiation energy (measured form Kahn Tear Tests)
vs. yield strength after a fast quench (F) or a slow quench (S) for
(a) AA7040 and (b) AA7050. UA stands for the underaged state, PA
for the peakaged state and OA for the overaged state.

2532

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

shows the main results concerning the alloy and quench


dependence on the yield strength/toughness compromise
during ageing. The classical dependence of toughness
during ageing is found: highest in underaged states,
lowest in peakaged states and intermediate in the overaged state. When the quench rate is lowered, a dramatic
decrease in fracture toughness is observed for all ageing
states, whereas the decrease in yield stress is almost
negligible. The dependence of toughness on the ageing
state is reduced as compared to the fast quench.
The strain-hardening behaviour of the two alloys is
very similar and has been shown to depend mostly on
the precipitate state inside the grain interiors (i.e. on the
state of ageing), whereas it is almost independent on
the quench rate. The strain-hardening rate is highest in
the underaged state, lowest in the peakaged state and
slightly higher in the overaged state.

cipitates coarser than the homogeneous precipitates; their


average size, measured on transmission electron microscope (TEM) images, is reported in Table 1. Fig. 2 shows
the microstructure at the grain boundaries in alloy 7040,
after a fast quench and in an overageing heat treatment.
After a slow quench, an important fraction of coarse,
heterogeneous precipitation is present in the material.
Grain boundary precipitates are present, which are
much larger than the precipitates formed during ageing
after a fast quench. As a consequence, they are not
observed to evolve during ageing. Heterogeneous precipitates are also present in the grain interiors, nucleated
on the Al3 Zr dispersoid particles (see Fig. 3). Their size
is quite similar in the two alloys (see Table 2), whereas
their volume fraction is much larger in the 7050 alloy,
owing to its higher quench sensitivity. A detailed analysis of the microstructure as a function of the process
route can be found in [31].

2.3. Microstructure
Both alloys contain intermetallic particles; however
their area fractions are quite dierent: 0.45% for alloy
7040, and 0.69% for alloy 7050. The recrystallised
fraction is similar in the two alloys, about 20%; however, due to the dierent rolling schedules, the grain
aspect ratio is not identical in the two alloys.
After a fast quench, no intergranular precipitates can
be detected. During the ageing treatment, ne precipitates
develop in the matrix. In the later states of ageing, the
grain boundaries appear to be densely covered with preTable 1
Average largest dimension (in nm) of precipitates on grain boundaries,
measured on TEM micrographs
Alloy

Quench

Underaged

Overaged

7040

Fast
Intermediate
Slow
Fast
Intermediate
Slow

50
80

60
70

40
50
110
40
50
80

7050

SD of the measurements is of the order of 15 nm.

2.4. Occurrence of fracture modes


The fracture modes observed on Kahn Tear Test
samples can be classied in four categories: fracture and
decohesion at second phase particles, transgranular
ductile fracture, transgranular shear fracture and intergranular fracture (see Fig. 4).
The area fraction of the rst fracture mode has been
determined by analysis of back-scattered electron
images of fracture surfaces in the SEM, using the high
Z-contrast of Fe with respect to Al. The result of these
measurements is shown in Table 3. It is observed that
the area fraction of second phase particles on the fracture surface is much larger than their area fraction on
polished surfaces (as expected), and that this parameter
is totally independent on the state of ageing and the
quench rate. This suggests that in the present situation,
second phase particles act essentially as fracture initiation sites, and that their own fracture or decohesion
occurs independently to all other fracture events. In this
framework, their inuence can be separated from all
other microstructural factors in terms of modelling.

Fig. 2. Dark eld TEM micrograph of precipitates on a grain boundary in the AA7040 alloy after a fast quench, and an over ageing heat treatment.

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

2533

Fig. 3. FEG-SEM image of alloy 7050 after a slow quench, showing coarse precipitation both on the grain boundaries and across several grains as
bands of precipitates nucleated on the dispersoids.
Table 2
Largest dimension (in nm) of transgranular quench-induced precipitates, determined from TEM and FEG-SEM images
Alloy

Quench

Underaged
TEM

FEG-SEM

TEM

FEG-SEM

7040
7050

Slow
Intermediate
Slow

190
205
260

200
180
270

180
200
210

250
210
230

Overaged

SD of the measurements is of the order of 30 nm.

The respective area fractions of the other fracture


modes have been quantitatively measured on a large
number of SEM images. Due to the diculty of separating them, the precision of the measurement can be
estimated to 10%. The result of these measurements is
shown in Table 4. The shear fracture mode is present
only in underaged materials, where shearable precipitates are present along with a low strain-rate sensitivity.
This failure mode dissipates a large amount of energy,
thus in terms of modelling it will be joined with the
transgranular ductile fracture mode. The fraction of
intergranular fracture mode is observed to increase
when the quench rate is reduced. For a fast quench, it is
maximum at peak strength, absent in the underaged
state and intermediate in the overaged state. For a slow
quench, it is constant as a function of ageing state. These
features correspond well to the microstructural observations cited above.

3. Model
The basic idea of the model is to calculate the total
energy dissipated per unit area during fracture using a

rule of mixture on the energies dissipated by the two


main fracture modes (intergranular and transgranular
fracture):
Etotal finter Einter ftrans Etrans ;

where finter and f trans correspond to the surface fraction


of the fracture surface which appears to be intergranular
or transgranular.
The model will be presented in three steps:
determination of the fracture energy for the transgranular mode,
determination of the fracture energy for the intergranular mode,
rationalisation of the area fractions of the two modes.
3.1. Energy dissipation during transgranular fracture
Our description of ductile transgranular fracture will
be largely based on a recent model by Zehnder and Hui
[32], which describes analytically the critical fracture
energy as a function of the plastic properties of the
material. This model is an analytical simplication of
the cohesive zone approach developed by Tvergaard
and Hutchinson [33]. In this approach, the energy

2534

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

Fig. 4. SEM micrographs of fracture surfaces showing typical occurrence of the four main fracture mechanisms: (a) fracture at intermetallic particles,
(b) ductile transgranular fracture, (c) ductile shear transgranular fracture and (d) intergranular fracture.
Table 3
Area fractions (in %) of intermetallic phases (Fe,Cu) on the fracture surfaces, as a function of quench rate and aging state for both alloys
Alloy

AA7040
AA7050

Underaged

Peakaged

Overaged

Fast

Slow

Fast

Slow

Fast

Slow

3.7

3.4
2.2

4.1
2.2

3.7
2.2

3.9
2.3

3.9
2.1

These area fractions can be compared with the average area fraction measured on a polished surface, which is 0.45% for AA7040 and 0.69% for
AA7050.
Table 4
Area fractions (in %) of the three main fracture modes measured on the fracture surfaces (I, intergranular; S, transgranular shear and D, transgranular ductile)
Alloy

Underaged
Fast

AA7040

AA7050

35%
65%
10%
15%
75%

S
D
I
S
D

Peakaged
Slow

Fast

40% I
40% S
20% D
45% I
5% S
50% D

10% I

Overaged
Slow

Fast

Slow
50% I

90% D

dissipation is separated in two contributions: one describing the energy necessary to create two free surfaces,
the other being related to the work in the plastic zone
ahead of the crack tip.
The rst contribution, which we will call Ctrans , can be
crudely estimated from the critical stress which induces
the decohesion at second phase particles and their size.

50% I

100% D
20% I

50% D
30% I

50%D

80%D

70% D

Typical values of 600 MPa for the decohesion stress and


10 lm for their size give a separation work:
Ctrans rdecoh dcrit  6000 J m2 :

In practice, this separation work will be estimated from


experimental results in situations where only transgranular fracture is observed (underaged states). The

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

value obtained will be checked for the correct order of


magnitude.
Since this separation work depends directly on the
void growth and coalescence between the second phase
particles [34], it can be expected that its value is proportional to the spacing of these phases. Thus, for a
given size of the intermetallic particle, the dependence
with volume fraction is expected to be
Ctrans / fv1=3 :

The second contribution to energy dissipation depends


on the plastic zone ahead of the crack tip. In order to
describe this contribution, the plastic behaviour of the
material will be described by a simple elasto-plastic
power-law:

Ee

n for r < gry ;
r
8
ry k e  Er
for r > gry ;
the parameter g takes into account the stress state at the
crack tip, which itself depends on the sample geometry.
The Kahn Tear Test samples thickness is 3 mm, thus we
will consider that plane stress is a good approximation
to the stress state. In this case g 1:2 [32]. In the present
model, crack propagation occurs when the local stress
equals the decohesion stress rdecoh . The total energy
dissipation per unit area during fracture is then calculated as follows:

Ctrans
for r < gry ;
Etrans
9
Ctrans Up Rp for r > gry ;
where Rp is the size of the plastic zone and Up is the
plastic work per unit volume to failure, in the stress state
experienced at the crack tip. In ductile fracture, only the
second case is generally met. In plane stress conditions,
the radius of the plastic zone has been calculated by
Irvin [35]:
ECtrans
Rp
:
pr2y

10

Combining Eqs. (8)(10), one obtains the total dissipated energy in the transgranular mode:
8
 31=n
2
rdecoh
>
<

1
gry
E
5
Etrans Ctrans 1 gry 1n=n 4
>
k
p
:
9
!>
rdecoh
n gry =
:
11

>
n1
;
For the sake of simplicity, the decohesion stress rdecoh
will be taken as the fracture stress rfracture measured in
uniaxial tensile tests. It can be observed that this expression includes all the parameters characterising the
plastic behaviour of the material: yield stress, work
hardening rate, stress to fracture. Notably, the ratio
rfracture =ry plays a key role; this parameter, called not-

2535

ched yield ratio, has been frequently shown to be representative of toughness [36,37].
3.2. Energy dissipation during intergranular fracture
In order to estimate the energy dissipated by pure
intergranular fracture, we will adapt the above-mentioned modelling approaches, which take into account
the characteristic dimensions of the relevant microstructural parameters, to take also into account the
energy dissipation in the grain interiors. The simplied
geometry describing the microstructure is shown in
Fig. 5. In the present approach, the grain boundary
environment is divided in two regions: the PFZ and the
grain interior, which have each their own plastic behaviour. In order to obtain an analytical description, the
plastic behaviour of both regions will be described by a
simple linear hardening law:
rgrain ryg hg eg ;

12

rGB ryb hb eb

13

the respective sizes of the PFZ and the grain are noted d
and D, and their ratio will be noted U d=D in the
following. The two regions are assumed to be loaded in
series, which means that rgrain rGB at all times. Following the approach of Embury and Nes [10], the crack
is assumed to propagate in the grain boundary when a
critical strain e is reached in the PFZ. To this critical
strain, we can calculate the stress to fracture rR and the
strain to fracture in the grain interior eg :
rR rb hb e ;
eg

rb  rg hb e
:
hg

14
15

The macroscopic strain to fracture is nally calculated


by an average between the strains experienced by the
PFZ and the grain interior (which is again coherent with
a loading in series):


rb  rg hb 
eR 1  U
16
e Ue :
hg
hg
This expression can be simplied in the usual case when
U  1 (the usual PFZ size is of the order of 2050 nm,
whereas the grain size 1050 lm in diameter); integrating the energy dissipated by plasticity until fracture
yields:


 2
2
2
2
r

h
e


r
b
b
rR  rg
g
Winter

:
17
2hg
2hg
This expression takes into account both the plastic behaviour of the grain boundaries and the characteristics
of the grain interiors.
The value of the critical strain to fracture in the PFZ
will be calculated from the Embury and Nes model [10],

2536

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

grain

PFZ

Fig. 5. Geometrical conguration used for the intergranular fracture


calculation.

taking into account the area fraction of precipitates on


the grain boundaries:

1
1

p  1 :
e
18
2
fGB
Finally, the work to fracture of Eq. (17) needs to be
multiplied by the size of the plastic zone, given by Eq.
(10), likewise the case of the transgranular fracture,
leading to the total dissipated energy per unit area in the
intergranular mode:
!

2
ECinter
hb
1
2
p  1
Einter
rb
 rg :
2pr20:2% hg
2
fGB
19
3.3. Description of the fracture toughness in mixed
fracture modes
In a fracture mode where both intergranular and intragranular fracture coexist, the total dissipated energy
is calculated using a rule of mixture between the energies
dissipated via each mode. For this rule of mixture to be
applied an essential ingredient is of course the respective
area fractions of the two fracture modes. We have determined them experimentally in a variety of experimental situations, which will enable to test the model
predictions. However, since no model exists, which
predicts these fractions as a function of the materials
properties, this inherently limits the applicability of the
model only to situations where the experimental determination of the fracture modes (which is cumbersome
and somewhat subjective) can be performed. Therefore,
we have sought a semi-empirical approach to relate the
area fractions of fracture modes to some other physical
parameters.

It can be expected that, given a constant grain morphology, the distribution of fracture between intergranular and transgranular modes must be governed by
the respective energies dissipated by each of the two
fracture modes. In extreme cases, when the dissipated
energies are equal, there should be no reason for a crack
to propagate in a grain boundary, and if the intergranular fracture energy is zero, there should be no transgranular fracture. It seems therefore reasonable to
consider that the fraction of each fracture mode is
controlled by the ratio of the dissipated energies.
In the expressions calculated above for the two fracture modes, all parameters have been obtained from
tensile tests, except the two cohesion energies Ctrans and
Cinter . The values of these parameters for the dierent
quench and ageing states are listed in Table 5.
The transgranular cohesion energy has been rst estimated for the alloy AA7040, fast quench, underaged,
where fracture is completely transgranular:
FQ
C7040
2915 J m2 :
trans

The corresponding value for alloy AA7050 has been


obtained by considering that the main dierence between the two alloys is the dierence in the area fraction
of the fracture surface covered with intermetallic particles (Table 3). Applying Eq. (1) then gives:
7040
1=3
7050 FQ
7040 FQ finter
Ctrans Ctrans
3500 J m2 :
7050
finter
The cohesion energy of the grain boundaries is much
more dicult to estimate, since there is no experimental
situation where fracture is purely intergranular. Thus,
the value of Cinter has been chosen, such as to provide
both a reasonable description of the relationship between Einter /Etrans and the fraction of the intergranular
fracture mode, and to provide a good description of the
data:
Cinter 500 J m2 :
This value, 1/6th of the transgranular cohesion energy, is
of the right order of magnitude. Finally, the transgranular cohesion energies for the slow quenches have been
adjusted to get a good description of the global value of
toughness. The values are:
SQ
C7040
1670 J m2 ;
trans
SQ
C7050
2000 J m2 :
trans

In order to calculate the dissipated energy for the intergranular fracture mode, a nal parameter needed is
the area fraction of precipitates on grain boundaries.
Following experimental observations, a constant value
for both alloys and all ageing states for a given quench
rate has been considered: 20% for a fast quench and 40%
for a slow quench.

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

2537

Table 5
Parameter values for the model
7040

UA
PA
OA
UA
PA
OA

Fast

Slow

7050

ry rtear =ry

n lnk

ry rtear =ry

n (lnk)

385
505
460
392
499
446

0.55
0.58
0.69
0.58
0.66
0.75

387
525
485
388
500
433

0.54
0.52
0.79
0.40
0.55
0.83

(1.75)
(1.26)
(1.29)
(1.32)
(0.77)
(1.01)

(6.32)
(5.74)
(6.31)
(6.43)
(5.94)
(6.45)

(1.88)
(1.29)
(1.44)
(1.46)
(0.92)
(1.14)

(6.35)
(5.74)
(6.63)
(6.09)
(5.89)
(6.88)

ry is the yield strength, measured on tensile tests; rtear is the tear strength, measured on Kahn Tear Tests; n and k are the parameters of the plastic
behaviour of the material measured on tensile tests according to the law r ry ken . In all cases: ryb 170 MPa; hb hg 1300 MPa.

The dependence between the experimental values of


finter and the calculated values of Einter /Etrans is shown in
Fig. 6. Although it is linear, the extreme values are not
the one which could be expected at rst. First, the
fraction of intergranular fracture becomes zero before
the two dissipated energies are equal. This means that
the energy gain for a crack to follow a grain boundary as
compared to propagate through a grain interior needs to
be larger than a critical value. This indicates that crack
deection occurring when a crack deviates into a grain
boundary, has a cost in energy. This fact has also been
evidenced in a similar work on the fracture of IF steels
[38]. Second, even in the case where the intergranular
energy dissipation is very small (less than 10% that of
transgranular fracture), the fraction of intergranular
fracture is very far from 100%. In fact, no fraction above
50% has been observed in the present experimental

Area fraction of intergranular fracture

1
7040
7050
0.8

0.6

study. Again, this show the diculty of the crack


bifurcation to follow the grain boundaries.
3.4. Model predictions: rationalisation of the inuence of
the microstructure
Using the expressions and experimental parameters
dened above, the model predictions can be compared
with the experimental data. Figs. 7(a) and (b) show
the model predictions for both alloys and both
quenches. All the general features are correctly described:
 The larger amplitude of toughness variation in the
fast quenched material as compared to the slow quenched material. This is essentially the result of the variation in fracture modes.
 The amplitude of the drop in toughness when the
quench rate is changed. This description is possible due
to the change in the transgranular cohesion energy, reecting the presence of coarse quench-induced precipitates in the grain interiors, which induce an easy
transgranular fracture.
 The minimum toughness value for peak ageing and
the lower value of toughness at given yield strength for
overaged material as compared to underaged material.
This results from the evolution of the global materials
properties.

0.4

3.5. Discussion: inuence of materials parameters on the


yield stress/toughness compromise

0.2

The model can be used to investigate separately the


inuence of some of the most important parameters on
the fracture toughness evolution along the ageing
treatment in order to guide further alloy development.
Figs. 8(a)(c) show the inuence, respectively of Cinter ,
Ctrans , and the area fraction of intergranular precipitates
fGB . The values considered for these parameters are
varied around the reference value for alloy AA7050 after
a fast quench.

0
0

0.2

0.4
0.6
E
/E
inter
trans

0.8

Fig. 6. Relationship between the measured area fraction of


intergranular fracture and the calculated ratio between the energies
dissipated by the intergranular and transgranular fracture modes.

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

Fast quench

150
4

0
350
(a)

Slow quench

100

50

400
450
500
Yield stress (MPa)

AA 7050
50
Fast quench

350

50

-2

G (10 J.m )

-2

G (10 J.m )

550

40
30

10

Influence of

inter

2
1.25
1
0.75
0.5
0.25

30

500

60

7050 F
7050 S
7050 F
7050 S

Slow quench

10

(b)

400
450
Yield stress (MPa)

(a)

40

0
350

2
1.25
1
0.75
0.5
0.25

550

60

20

trans

20

10

Influence of

10

200

F - experiment
S - experiment
F - model
S - model

-2

G (10 J.m )

30

AA 7040

-2

40

G (10 J.m )

2538

20
10

400
450
500
Yield stress (MPa)

550

350

400
450
Yield stress (MPa)

(b)

Fig. 7. Comparison between the experimental data (dashed lines) and


the model calculations (solid lines) in the Gc vs. yield stress diagram,
for both quenches and all ageing states. (a) AA7040; (b) AA7050.

500

550

60
Influence of f

GB

1.6

50

1.4
1.2
1
0.8
0.6

-2
3
c

It is rst interesting to focus on the respective inuence of Ctrans and fGB . In fact, these are the two parameters which are expected to change most when the
quench rate is changed: a slow quench favours intergranular precipitation, thus increasing fGB , and transgranular coarse precipitation on the dispersoids, which
is expected to decrease Ctrans . According to the model
predictions, these two parameters modify the overall
toughness/yield strength relationship in a very dierent
way: changing fGB results in an overall decrease of the
toughness level, whatever the ageing state, whereas
changing Ctrans results both in an overall decrease of
toughness but also in a reduced inuence of ageing on
toughness. Actually this parameter had to be changed
signicantly from the fast to the slow quench in order to
describe correctly the experimental data.
The second point which could be surprising at rst is
the eect of Cinter . Changing this parameter not only
changes the value of toughness in the states where
fracture is mainly intergranular, but also, and in the
largest extent, in the states where intergranular fracture

G (10 J.m )

40
30
20
10
0
350
(c)

400
450
Yield stress (MPa)

500

550

Fig. 8. Parametric evaluation of the inuence of various parameters on


the Gc vs. yield stress diagrams. The reference curve is for alloy
AA7050, after a fast quench. The parameter values are given normalised to this reference state (which always has the value 1). (a) Inuence
of the transgranular separation work Ctrans ; (b) inuence of the intergranular separation work Cinter ; (c) inuence of the area fraction of the
grain boundaries covered with precipitates fGB .

is initially very limited (like in the underaged state).


This illustrates the complexity of the eect of microstructure: changing the value of Cinter not only changes

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

the energy dissipated by this fracture mode, but also the


proportion of intergranular fracture, amplifying its
inuence.

4. Conclusions
A model has been proposed, which describes the
energy dissipated during ductile failure, using three
ingredients:
(i) A model for transgranular failure, based on the
simplied cohesive zone approach [32], which includes all the main materials parameters (intrinsic
cohesive energy, yield and failure strength, work
hardening exponent).
(ii) A model for intergranular failure, based on a classical failure criterion [10], which includes the principal geometrical features and materials properties
of the PFZ, as well as the plastic properties of the
grain interior.
(iii) A simple phenomenological model for estimating
the proportion of the two failure modes as a function of their respective dissipation energies, and a
law of mixture to estimate the materials dissipation
energy.
This model has been applied to an extensive data set,
including two dierent aluminium alloys, quenched
from the solution treatment with two dierent quench
rates, and after various ageing treatments. This set of
process parameters resulted in a large variety in microstructures, failure modes and consequently yield
strength and toughness. A key point in describing the
data with the model has been the quantitative determination of the respective proportions of failure modes as
a function of process parameters.
The main conclusions of the model application on the
microstructural features controlling ductile failure in the
present situation are:
 The evolution of energy dissipation during ageing is
controlled by a complex set of parameters, including the
intrinsic plasticity of the grain interiors and the competition between deformation in grain interiors and the
PFZ. This evolution is greatly reduced after a slow
quench, largely due to a decrease of the intrinsic cohesive energy of the grain interiors.
 The eect of quench rate on the energy dissipation
is dramatic. The application of the model shows that
this reduction is due in approximately equal proportions
to an increase in the area fraction of grain boundary
precipitates and a decrease in the intrinsic cohesive
energy of the grain interiors due to heterogeneous
precipitation.
 In the framework of the present experimental situation, a relationship between the materials average
properties and microstructure (Einter /Etrans ) and the area

2539

fraction of intergranular fracture (finter ) has been characterised. This relationship can be used to predict the
eect of individual process parameters, not only on energy dissipation during fracture, but also on the occurrence of the dierent fracture modes.
 A parametric analysis of the behaviour predicted
by the model can provide a guideline for further alloy
development.
One limitation of the present state of modelling is
that the above mentioned relationship Einter /Etrans vs.
(finter ) is expected to depend signicantly on a number of
microstructural features such as the grain morphology
and the proportion of recrystallised grains. In order to
further improve the prediction of fracture toughness, a
necessary step is therefore to model in a predictive way
the occurrence of failure modes in ductile materials as a
function of the relevant microstructural parameters.
This is by no means an easy task, but some step forward
has already been obtained through micromechanical
modelling [29].

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]

Rice JR, Johnson MA. Inelast Behav Solids 1970;641.


Burgard HC. Metall Trans 1974;5:2083.
Hahn GT, Roseneld AR. Metall Trans 1975;6A:653.
Ohira T, Kishi T. Mater Sci Eng 1986;78:9.
Garret GG, Knott JF. Metall Trans 1978;9A:1187.
Chen CQ, Knott JF. Met Sci 1981;15:357.
Eswarada Prasad N, Kamat SV, Prasad KS, Malakondaiah G.
Eng Fract Mech 1993;46(2):209.
Jata KV, Starke EA. Metall Trans 1986;17A:1011.
Roven HJ. Scripta Metall Mater 1992;26:1383.
Embury JD, Nes EZ. Metallkunde 1974;65:45.
Unwin P, Smith G. J Inst Met 1969;97:299.
Vasudevan AK, Doherty RD. Acta Metall 1987;35:1193.
Hornbogen E, Graf M. Acta Metall 1977;25:877.
Thompson DS, Zinkham RE. Eng Fract Mech 1975;7:
389.
Sanders TH, Starke EA. Acta Metall 1982;30(5):927.
Srinivas M, Malakondaiah G, Armstrong RW, Rao PR. Acta
Metall Mater 1991;39(5):807.
Kawabata T, Izumi O. Acta Metall 1976;24:817.
Li BQ, Reynolds AP. J Mater Sci 1998;33(24):5849.
Sugamata M, Blankenship CP, Starke EA. Mater Sci Eng A
1993;A163(1):1.
Kamat SV, Hirth JP. Acta Mater 1996;44(3):1047.
Deshpande NU, Gokhale AM, Denzer DK, Liu J. Metall Mater
Trans A 1998;29A:1191.
Gokhale AM, Deshpande NU, Denzer DK, Liu J. Metall Mater
Trans A 1998;29A:1203.
Goods SH, Brown LM. Acta Metall 1979;27:1.
Argon AS, Im J, Safoglu R. Metall Trans A 1975;6A:825.
Needleman A. J Appl Mech 1987;54:525.
Gurson ALJ. Eng Mater Tech 1977;99:2.
Tvergaard V. Int J Mech Sci 2000;42:381.
Thomason PF. Ductile fracture of metals. Oxford: Pergamon
Press; 1990.
Pardoen T, Dumont D, Brechet Y, Deschamps A. J Mech Phys
Solids 2003;51(4):637.

2540

D. Dumont et al. / Acta Materialia 52 (2004) 25292540

[30] Dumont D, Deschamps A, Brechet Y. Mater Sci Eng A


2003;356:326.
[31] Dumont D, Deschamps A, Brechet Y. Mater Sci Tech [in press].
[32] Zehnder AT, Hui CY. Scripta Mater 2000;42(10):1001.
[33] Tvergaard V, Hutchinson JW. J Mech Phys Solids 1992;40:1377.
[34] Wei Y, Hutchinson JW. Int J Fract 1999;95:1.

[35] Irwin GR. In: Proceedings of the 7th Sagamore Ordonance


Material Research Conference, Syracuse, NY, vol. 4; 1961. p. 63.
[36] Kaufman JG, Knoll AF. Mater Res Stud 1964;4:151.
[37] Senz RR, Spuhler EH. Met Prog 1975;107:64.
[38] S. Ma^itrejean, Ph.D. Thesis, Institut National Polytechnique de
Grenoble, France, 2000.

Vous aimerez peut-être aussi