You are on page 1of 8

Available online at www.sciencedirect.

com
Available
Available online
online at at www.sciencedirect.com
www.sciencedirect.com
Available online at www.sciencedirect.com

ScienceDirect
ScienceDirect
Structural
Procedia
Structural Integrity
Structural Procedia
Integrity
Integrity Procedia 00(2016)
(2016)
2 (2016)
00 000000
25352542
000000
www.elsevier.com/locate/procedia
www.elsevier.com/locate/procedia
Structural Integrity Procedia 00 (2016) 000000
www.elsevier.com/locate/procedia

21st European Conference on Fracture, ECF21, 20-24 June 2016, Catania, Italy
21st European Conference on Fracture, ECF21, 20-24 June 2016, Catania, Italy
Unit cell simulations and porous plasticity modelling for
Unit cell
XV Portuguese simulations
recrystallization
Conference on Fracture, and porous
textures
PCF 2016, plasticity
in10-12 Februarymodelling
aluminium alloys
2016, Pao defor Arcos, Portugal
recrystallization textures in aluminium alloysa
Thermo-mechanical L.E.B Dhlia,modeling
a,
, J. Faleskogof b
b
, T.aBrvik
high a
a
pressure
, O.S. Hopperstad turbine a
blade of an
L.E.B(SIMLab),
DhliCentre , J.forFaleskog , T. Brvik (CRI),, Department
O.S. Hopperstad
University of airplane gas(NTNU),
turbineNO-7491engine
aStructural Impact Laboratory Research-based Innovation of Structural Engineering, Norwegian
a Science and Technology Trondheim, Norway
Structural Impact Laboratory (SIMLab), Centre for Research-based Innovation (CRI), Department of Structural Engineering, Norwegian
b Department of Solid Mechanics, Royal Institute of Technology, SE-100 44 Stockholm, Sweden
University of Science and Technology (NTNU), NO-7491 Trondheim, Norway
a b c
P. Brando , V. Infante , A.M. Deus *
b Department of Solid Mechanics, Royal Institute of Technology, SE-100 44 Stockholm, Sweden

a
Department of Mechanical Engineering, Instituto Superior Tcnico, Universidade de Lisboa, Av. Rovisco Pais, 1, 1049-001 Lisboa,
Abstract Portugal
b
IDMEC, Department of Mechanical Engineering, Instituto Superior Tcnico, Universidade de Lisboa, Av. Rovisco Pais, 1, 1049-001 Lisboa,
Abstract
The well-known Gurson model has been heuristically extendedPortugal to incorporate effects of matrix anisotropy on the macroscopic
yielding
c of porous
The CeFEMA,
well-known ductile
Department
Gurson of solids.
model Typical
Mechanical
has been components
Engineering, of Superior
Instituto
heuristically recrystallization
extended toTcnico, textures forde
Universidade
incorporate effects aluminium
of Lisboa,
matrixAv.alloys
Roviscowere
anisotropy on used
Pais, to calibrate
1, 1049-001
the Lisboa,
macroscopic
the Barlat Yld2004-18p yield criterion using a full-constraint Taylor homogenization
Portugal method. The resulting yield
yielding of porous ductile solids. Typical components of recrystallization textures for aluminium alloys were used to calibrate surfaces were
further employed
the Barlat Yld2004-18pin unityield
cell simulations
criterion usingusing the finite element
a full-constraint Taylormethod. Unit cell calculations
homogenization method. Theare invoked
resulting to investigate
yield surfaces werethe
evolution of the approximated microstructure under pre-defined loading conditions and to calibrate
further employed in unit cell simulations using the finite element method. Unit cell calculations are invoked to investigate the the proposed porous plasticity
Abstract
model. Numerical results obtained from the unit cell analyses demonstrate that anisotropic plastictheyielding has porous
great impact on
evolution of the approximated microstructure under pre-defined loading conditions and to calibrate proposed plasticity
the mechanical response of the approximated microstructure. Despite the simplifying assumptions
model. Numerical results obtained from the unit cell analyses demonstrate that anisotropic plastic yielding has great impact on that underlie the proposed
During their
constitutive model, operation,
it seems modern aircraft engine components are subjected unittocell.
increasingly todemanding operating conditions,
the mechanical response oftothecapture the overall
approximated macroscopic
microstructure. response
Despite ofthethesimplifying However,
assumptions further enhance
that underlie the
the numerical
proposed
especiallythe
predictions, themodel
high pressure
should turbine
be (HPT) blades.
supplemented with Such
a void conditions
evolution cause these that
expression partsaccounts
to undergo
for different
directionaltypes of time-dependent
dependency, and a
constitutive model,
degradation, one itofseems
which toiscapture
creep. the overallusing
A model macroscopic
the finite response
element of the unit
method cell. However,
(FEM) to further
was developed, enhance
in order theable
to be numerical
to predict
void coalescence
predictions, the criterion
model should in order to capture
be supplemented the last
with stages of deformation.
c the
2016 creep behaviour
The Authors. of HPT
Published byblades. Flight
Elsevier B.V. dataarecords
void evolution
(FDR) expression
for a specific that aircraft,
accounts provided
for directional
by a dependency,
commercial and a
aviation
void coalescence
Copyright 2016
company,under criterion
The
were responsibilityin
Authors. order
used to obtainofthermalto
Published capture
by the
Elsevier last
B.V. stages
and mechanicalThis isof
andeformation.
dataopen access article under the CC BY-NC-ND license
for three different flight cycles. In order to create the 3D model
Peer-review
c needed
2016 The byathe Scientific
B.V. Committee of ECF21.
 forAuthors. Published Elsevier
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
the FEM analysis, HPT blade scrap was scanned, and its chemical composition and material properties were
Peer-review
obtained.under
Peer-review
Keywords: The responsibility
under
Porous responsibility
data that was
plasticity; ofgathered
Plasticthe Scientific
ofanisotropy;
the Committee
Scientific
was Unit
fed cell
into of FEM
Committee
the ECF21.
simulations;of model
ECF21. and differentGurson
Barlat Yld2004-18p; simulations
model were run, first with a simplified 3D
rectangular block shape, in order to better establish the model, and then with the real 3D mesh obtained from the blade scrap. The
Keywords: Porous plasticity; Plastic anisotropy; Unit cell simulations; Barlat Yld2004-18p; Gurson model
overall expected behaviour in terms of displacement was observed, in particular at the trailing edge of the blade. Therefore such a
model can be useful in the goal of predicting turbine blade life, given a set of FDR data.
1. Introduction
2016 The Authors. Published by Elsevier B.V.
1. Peer-review
Introduction
underalloys
responsibility
Wrought metal used forofengineering
the Scientificapplications
Committee ofoften
PCF exhibit
2016. anisotropic behaviour under plastic deforma-
tions. Rolling
Wrought and alloys
metal subsequent annealing
used Blade;
for of aluminium alloys
engineering induces plastic anisotropy in which
underthe material axes
Keywords: High Pressure Turbine Creep; Finiteapplications often
Element Method; exhibitSimulation.
3D Model; anisotropic behaviour plastic deforma-
align with
tions. the rolling,
Rolling normal, and
and subsequent transverse
annealing directions. alloys
of aluminium The resulting
inducestextures
plastic are mainly composed
anisotropy of cube
in which the and goss
material axes
generic
align textures
with (Barlat
the rolling, and Richmond,
normal, 1987)directions.
and transverse in addition
Theto resulting
some degree of random
textures texture.
are mainly Such aluminium
composed of cube andalloys
goss
are frequently
generic employed
textures in structural
(Barlat and Richmond, applications, for instance
1987) in addition in automotive
to some degree of and marine
random industries.
texture. Hence, predictive
Such aluminium alloys
material
are models
frequently capable in
employed of structural
accounting for the microstructural
applications, for instance evolution is of and
in automotive great importance
marine for structural
industries. integrity
Hence, predictive
assessment
material and design.
models capable of accounting for the microstructural evolution is of great importance for structural integrity
assessment and design.
Corresponding author. Tel.: +47 73594677
*E-mail address: lars.e.dahli@ntnu.no
Corresponding
Corresponding author.Tel.:
author. Tel.:+47
+351 218419991.
73594677
E-mail
E-mail address:
address: amd@tecnico.ulisboa.pt
lars.e.dahli@ntnu.no

2452-3216 2016 The Authors. Published by Elsevier B.V.


2452-3216 c 2016
Peer-review under responsibility
The Authors. of the
Published Scientific
by Elsevier B.V.Committee of PCF 2016.
Copyright 2016 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
Peer-review under responsibility of the Scientific Committee of ECF21.
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
2452-3216 c 2016 The Authors. Published by Elsevier B.V.
Peer reviewunder
Peer-review under responsibility
responsibility of Scientific
of the the Scientific Committee
Committee of ECF21.
of ECF21.
10.1016/j.prostr.2016.06.317
2536 L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542
2 L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000

A commonly observed ductile fracture mechanism is by nucleation, growth and final coalescence of microscopic
voids (Hancock and Mackenzie, 1976). This mechanism is largely dependent upon the degree of triaxiality of the
imposed stress states (Rice and Tracey, 1969). Triaxial loading cases that commonly arise in practical application
facilitates enlargement of pre-existing and nucleated voids, which becomes important due to the loss of load-carrying
capacity and material softening. In addition, previous investigations (Zhang, 2001; Gao et al., 2010) have revealed a
pronounced effect of the Lode angle on ductile fracture even for isotropic matrix materials that are independent of
the third deviatoric stress invariant J3 . This effect is closely related to the void evolution, which differs substantially
between states of generalized tension, shear and compression.
Ductile fracture of anisotropic metal alloys has received considerable interest in the literature during the past
decade. Numerical unit cell studies that incorporate plastically anisotropic matrix behaviour (Keralavarma and Benz-
erga, 2010; Steglich et al., 2010; Keralavarma et al., 2011) reveal that the mechanical response is markedly dependent
upon the induced anisotropy during plastic deformations. This dependence is rather obvious in terms of stress-strain
response, but anisotropy also affects the evolution of the voids; both growth rate and shape evolution is related to
the degree of anisotropy and orientation of the material axes. In effect, this has profound influence on the aggregate
material behaviour.
Local approaches to ductile fracture include porous plasticity models that account for the evolution of damage
during plastic deformation. One such constitutive relation is the Gurson model (Gurson, 1977) which has been widely
employed in numerical studies in the literature and subjected to many modifications. Among these include extensions
to incorporate plastic anisotropy effects of the matrix material. Benzerga and Besson (2001) used upper-bound limit
analysis to derive a yield function for matrix materials governed by Hills yield criterion for anisotropic materials.
Efforts have also been made by Benzerga et al. (2004), and in this study the model was also supplemented with a
coalescence criterion. Monchiet et al. (2008) and Keralavarma and Benzerga (2010) provide a constitutive model for
plastically anisotropic porous solids in the case of non-spherical voids, however restricted to remain spheroidal. We
should also note that some studies have been devoted to extend the Gurson modelling framework for matrix materials
that are governed by a crystal plasticity formulation (see e.g. Han et al. (2013); Paux et al. (2015)) which is inherently
anisotropic.
The present study is largely inspired by the work undertaken by Steglich et al. (2010), where combined use of unit
cell simulations and a homogenized material model gave promising results for predictions of the direction-dependent
deformation and crack propagation of an Al2198 sheet metal alloy. They used the Gurson model in its original form,
however incorporating an equivalent stress measure that accounts for the plastic anisotropy of the matrix material. The
matrix material was described by the anisotropy model introduced in Bron and Besson (2004). We propose a similar
phenomenological extension of the Gurson model in order to incorporate plastic anisotropy. The linear-transformation
based yield criterion by Barlat et al. (2005) is used to describe the plastic anisotropy of the matrix material. Only an
isotropic distribution of spherical voids that undergo spherical void growth will be considered. Thus, any anisotropy
effects of initial void and particle morphology are precluded.

2. Matrix description

A hypoelastic-plastic framework is assumed for the material behaviour. The elastic deformations are approximated
by Hookes law while the plastic response is governed by orthotropic plasticity using the Barlat Yld2004-18p consti-
tutive model (Barlat et al., 2005) with the associated flow rule and isotropic work-hardening. The two generic textures
employed in this study are the main components for a recrystallization texture in aluminium alloys (Barlat and Rich-
mond, 1987). A full-constraint Taylor homogenization procedure was used to calibrate the yield surfaces for the two
textures shown in Fig. 1. We note that this method assumes that all grains are subjected to the same deformation,
thus neglecting possible effects of stress and strain gradients within the grains. Details regarding the yield surface
calibration may be found in Saai et al. (2013).
Table 1: Generic material parameters for the matrix material.

E [MPa] 0 [MPa] Q [MPa] C

70000 0.3 100 100 10


L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542 2537
L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000 3

120 120
150 90 150 90
2 2

180 60 180 60

210 30 210 30

3 1 3 1
240 0 240 0

270 330 270 330


300 300
(a) (b)

Fig. 1: Plot of the yield surface in the deviatoric plane for (a) cube and (b) goss textures. The yield surfaces are generated for the case when the
material axes coincide with the principal stress directions. The deviatoric angles = 0 , 30 , . . . , 330 correspond to those investigated in this study.

The Barlat Yld2004-18p model consists of two linear transformations of the stress deviator

s = C : s s = C : s (1)

where s and s are the transformed deviatoric stress tensors, C and C are the fourth-order transformation tensors
accounting for plastic anisotropy, and s is the stress deviator. The equivalent stress is then defined as
3 3 m1
1  
eq () = |S i S j |m (2)
4 i=1 j=1

where S i and S j are the principal values of the transformed deviatoric stress. The reader is referred to Barlat et al.
(2005) for further details about the anisotropic yield function. In the present work, the anisotropic yield criterion is
calibrated for plane stress states, reducing the number of independent coefficients in C and C and the computational
cost of the calibration.
Plastic yielding of the matrix material is governed by the yield function

(, M ) = () M 0 (3)

where the matrix flow stress is assumed to be described by a one-term Voce rule
 
M = 0 + Q 1 exp (C p) (4)

with the material parameters found in Table 1. The accumulated plastic strain p is calculated from
 t
: dp : dp
p = p= dt (5)
eq 0 eq
where the stress and plastic rate-of-deformation tensors are denoted and d p , respectively.

3. Porous plasticity model

To approximate the unit cell response in a single material element subjected to homogeneous deformation, a phe-
nomenological extension of the Gurson model is proposed
   
eq 2 3h
(, M , f ) = + 2 f q1 cosh q2 1 (q1 f )2 0 (6)
M 2M
2538 L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542
4 L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000

Here, eq is the macroscopic equivalent stress governed by Equation (2), h is the macroscopic hydrostatic stress, M
is the matrix flow stress, f is the void volume fraction and qi are the material parameters introduced by Tvergaard
(1981). The model parameters qi will be calibrated from unit cell analyses, to optimize the correspondence between
the homogenized material model and the mechanical response of the approximated microstructure for a variety of
loading cases. These results are presented in Section 5.
In this work, we adopt the void evolution law

f = (1 f ) trD p (7)

where trD p denotes the volumetric plastic strain rate. Note that this relation does not account for non-spherical void
evolution during plastic deformation. For the highly anisotropic matrix materials considered in this work, this assump-
tion is evidently violated due to the directional dependency of the matrix flow stress.
The matrix flow stress M is described by a generic work-hardening law calculated from Equation (4) with the
matrix parameters listed in Table 1. For the voided solid, the matrix accumulated plastic strain is calculated from
 t
: Dp : Dp
p = p= dt (8)
(1 f ) M 0 (1 f ) M

where the macroscopic stress and plastic rate-of-deformation tensors are denoted and D p , respectively.

4. Unit cell modelling

The unit cell simulations were carried out under the assumption that the principal stress directions are collinear to
the orthotropy axes which enable the use of symmetry conditions to reduce the discretized model to a one-eight model,
as illustrated in Fig. 2. In the numerical procedure, the material axes of the unit cell (mi in Fig. 2a) remain fixed. The
principal stress directions are then varied in different analyses by changing the deviatoric angle (), which is defined
as the angle between 1 and the current stress state in the deviatoric plane (see Fig. 1). Hence, effects of changing
the main loading directions relative to the anisotropy axes may be elucidated. Additionally, the stress triaxiality (T ) is
varied to demonstrate its effect on the aggregate mechanical response.

m2 i
k
m1
m3

(a) (b)

Fig. 2: Illustration of (a) the representative volume element used in this study, where the material axes are aligned with principal stress directions
and (b) the corresponding FE model where symmetry conditions in all three directions have been utilized.

In order to control the macroscopic stress state imposed to the unit cell, we prescribe values of the stress triaxiality
(T ) and the deviatoric angle (). To this end, the principal stress vector is written on the form

1 cos
 
1
cos   1
2 vm cos 2 vm
2
cos 2
2 = eq  3 
+ h
1
= eq



 3 
+ T 1 (9)
3 3 cos + 2
1
3
cos + 2
1
3 3
L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542 2539
L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000 5

where vmeq is the macroscopic equivalent von Mises stress, h is the macroscopic hydrostatic stress and is the angle
in the deviatoric plane, taking values on the range 0 2. The stress triaxiality T is defined on the usual form
h
T= (10)
vm
eq

The unit cells are subjected to a wide range of stress states, corresponding to = 0 , 30 , . . . , 330 and T =
0.667, 1.0, 1.667, 3.0, which results in a total of 48 simulations for each texture. A multi-point constraint (MPC) user
subroutine was implemented in the implicit FE solver Abaqus/Standard to maintain the imposed stress state through-
out the numerical simulations. Details regarding the MPC subroutine and its application to unit cell analyses may be
found elsewhere, see for instance Faleskog et al. (1998), Barsoum and Faleskog (2007), and Dhli et al. (2016).

15.10
= 0 T = 1.0
= 90
= 120
11.57 = 240

f
f0 8.05
(a) (b)

4.53

1.00
0.0 0.2 0.4 0.6 0.8
vm
Eeq
(e)
(c) (d)

Fig. 3: Deformed configurations for the Goss texture at maximum equivalent stress for loading situations corresponding to a stress triaxiality
of T = 1.0 for the deviatoric angles (a) 0 , (b) 90 , (c) 120 , and (d) 240 . Fringes of accumulated plastic strain are shown on the deformed
configurations. The corresponding void growth plots are presented in (e).

Fig. 3 shows fringes of accumulated plastic strain on deformed configurations of the unit cell with the Goss texture
for various deviatoric angles and a stress triaxiality of T = 1.0. The deformed configurations correspond to the
first frame after the macroscopic equivalent stress reaches its maximum vm eq = 0, and consequently the macroscopic
effective deformation is in general different between the shown configurations. Fig. 3e shows that the void growth for
the Goss texture is linked to the magnitude of the imposed stress, from which it is evident that the void evolution is
affected by the loading. Another interesting observation is that even though the loading states = 120 and = 240
are almost indistinguishable in terms of vm eq , seen from the radius of the yield surface in Fig. 1b, the void growth
rate is very different. Also, their void shapes evolve quite differently. This indicates that the plastic flow direction, or
correspondingly the normal to the yield surface, is important for the void evolution. However, the void shape is also
affected by the local field quantities in the proximity of the void, and it is thus difficult to separate the various effects.
In general terms, the interplay between the plastic flow and the equivalent stress seems decisive for the macroscopic
behaviour of the unit cell.

5. Calibration of porous plasticity model

The constitutive relation given by Equation (6) was calibrated to the numerical unit cell calculations by means of a
non-linear least-square error routine. Model parameters qi were varied in the optimization procedure. The stress and
void volume fraction residuals to be minimized were defined as
 Eeqmax  Eeqmax
0
|GT UC
eq eq |dE eq | f GT f UC |dEeq
e =  E max   , e f =  E0max   (11)
eq 1 GT UC eq 1 GT + f UC dE
0 2 eq + eq dE eq 0 2 f eq
2540 L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542
6 L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000

2.10 4.70
= 0
T = 1.0 = 60
= 120
= 180
1.58 3.78 = 240
= 300
Unit cell
vm
eq
f
Gurson
0 1.05 f0 2.85
= 0
= 60
= 120
= 180
0.52 1.92
= 240
= 300
Unit cell T = 1.0
Gurson
0.00 1.00
0.000 0.125 0.250 0.375 0.500 0.000 0.125 0.250 0.375 0.500
vm vm
Eeq Eeq

(a) (b)

2.10 4.50
= 30
T = 1.0 = 90
= 150
= 210
1.58 3.62 = 270
= 330
Unit cell
vm
eq
f
Gurson
0 1.05 f0 2.75
= 30
= 90
= 150
= 210
0.52 1.88
= 270
= 330
Unit cell T = 1.0
Gurson
0.00 1.00
0.000 0.150 0.300 0.450 0.600 0.000 0.150 0.300 0.450 0.600
vm vm
Eeq Eeq

(c) (d)

Fig. 4: Response of unit cell and the homogenized material model in terms of (a) and (c) equivalent von Mises stress, and (b) and (d) void volume
fraction against the equivalent strain. All curves shown are for the Cube texture.

where the superscripts GT and UC denote Gurson-Tvergaard and unit cell, respectively, and || denotes the magnitude.
Note that the limit of the definite
 integral is the equivalent strain at maximum stress from the respective unit cell
max UC
analyses Eeq = Eeq vm
eq = 0 with the equivalent strain being defined by

2  
Eeq = E E (12)
3 ij ij
where Ei j are the macroscopic strain deviator components. A weighted residual on the form

e = w e + w f e f (13)

was adopted in the optimization process. In the present work, the residuals were given equal weight w = w f because
the calibrated qi -values were only slightly affected by the residual weights.
Table 2: Parameters retrieved from the optimization procedure.

Texture q1 q2

Cube 1.912 0.791


Goss 1.282 0.842

The resulting parameters of the optimization procedure can be found in Table 2. For comparative reasons, we note
that the calibrated parameters in the study by Steglich et al. (2010) were q1 = 1.22 and q2 = 1.16. An obvious reason
L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542 2541
L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000 7

2.20 4.60
= 0
T = 1.0 = 60
= 120
= 180
1.65 3.70 = 240
= 300
Unit cell
vm
eq
f
Gurson
0 1.10 f0 2.80
= 0
= 60
= 120
= 180
0.55 1.90
= 240
= 300
Unit cell T = 1.0
Gurson
0.00 1.00
0.000 0.200 0.400 0.600 0.800 0.000 0.200 0.400 0.600 0.800
vm vm
Eeq Eeq

(a) (b)

2.70 4.10
= 30
T = 1.0 = 90
= 150
= 210
2.03 3.32 = 270
= 330
Unit cell
vm
eq
f
Gurson
0 1.35 f0 2.55
= 30
= 90
= 150
= 210
0.68 1.78
= 270
= 330
Unit cell T = 1.0
Gurson
0.00 1.00
0.000 0.225 0.450 0.675 0.900 0.000 0.225 0.450 0.675 0.900
vm vm
Eeq Eeq

(c) (d)

Fig. 5: Response of unit cell and the homogenized material model in terms of (a) and (c) equivalent von Mises stress, and (b) and (d) void volume
fraction against the equivalent strain. All curves shown are for the Goss texture.

for the difference between the values found in the present study and these referred values is that we consider widely
different materials, both in terms of texture and work-hardening properties. Also, the current work covers a wider
range of stress states which may influence the optimized solution parameters.
Figs. 4 and 5 compare the calibrated model response against the unit cell calculations for a stress triaxiality of
T = 1 and for all deviatoric angles in terms of equivalent stress and void growth. From these curves we may observe
that the model captures the general trends of the unit cell simulations in terms of equivalent stress-strain response.
With reference to Figs. 4a, 4c, 5a, and 5c, the discrepancy between the respective curves is more pronounced for the
generalized axisymmetric states than for the generalized shear states. Also, the cube texture seems somewhat better
replicated by the proposed Gurson model. This is most likely due to the less extreme anisotropy of this texture as
compared to the Goss texture (see Fig. 1), leaving it more compatible with the framework of the original Gurson
model. We may readily see from Figs. 4b, 4d, 5b, and 5d that the void evolution is not accurately predicted by this
model. This is presumably due to the assumption of spherical void growth which is employed in Equation 7. In the
case of anisotropy, the void shape evolution will depend upon the orientation of the material axes (see Section 4).

6. Concluding remarks

A heuristic extension of the Gurson model to account for plastic anisotropy of the matrix material is proposed. Unit
cell analyses were employed to investigate effects of plastic anisotropy on the mechanical response and to calibrate
the porous plasticity model. Unit cell calculations revealed the great influence of matrix anisotropy on the stress-
strain response and the microstructural evolution, which in the present work is approximated by a single parameter
accounting for the void volume fraction. The void growth rate is greatly affected by the equivalent stress magnitude.
2542 L.E.B Dhli et al. / Procedia Structural Integrity 2 (2016) 25352542
8 L.E.B. Dhli et al. / Structural Integrity Procedia 00 (2016) 000000

Numerical simulations also demonstrate the directional dependency of the void evolution, which is linked to the plastic
flow direction. The calibrated porous plasticity model shows predictive capabilities. However, the equivalent stress-
strain response is captured to a greater extent than the void growth. This is most likely related to the void evolution
expression which does not account for the void shape evolution and directional dependency due to plastic anisotropy.
It thus seems probable that the predictions would be enhanced if the void evolution is more correctly accounted for
by the model. Also, it is noted that void coalescence is not accounted for in the present work. More involved unit cell
analyses are needed to study this appropriately, which is considered for future work.

Acknowledgement

The financial support of this work from the Centres for Research-based Innovation (CRI) SIMLab and CASA at
the Norwegian University of Science and Technology (NTNU) is gratefully acknowledged.

References

Barlat, F., Aretz, H., Yoon, J.W., Karabin, M.E., Brem, J.C., Dick, R.E., 2005. Linear transfomation-based anisotropic yield functions. International
Journal of Plasticity 21, 10091039.
Barlat, F., Richmond, O., 1987. Prediction of tricomponent plane stress yield surfaces and associated flow and failure behavior of strongly textured
F.C.C. polycrystalline sheets. Materials Science and Engineering 95, 1529.
Barsoum, I., Faleskog, J., 2007. Rupture mechanisms in combined tension and shear-Micromechanics. International Journal of Solids and Structures
44, 54815498.
Benzerga, A.A., Besson, J., 2001. Plastic potentials for anisotropic porous solids. European Journal of Mechanics, A/Solids 20, 397434.
Benzerga, A.A., Besson, J., Pineau, A., 2004. Anisotropic ductile fracture: Part II: Theory. Acta Materialia 52, 46394650.
Bron, F., Besson, J., 2004. A yield function for anisotropic materials Application to aluminum alloys. International Journal of Plasticity 20,
937963.
Dhli, L.E.B., Brvik, T., Hopperstad, O.S., 2016. Influence of loading path on ductile fracture of tensile specimens made from aluminium alloys.
International Journal of Solids and Structures doi:10.1016/j.ijsolstr.2016.03.028.
Faleskog, J., Gao, X., Shih, C.F., 1998. Cell model for nonlinear fracture analysis I. Micromechanics calibration. International Journal of Fracture
89, 355373.
Gao, X., Zhang, G., Roe, C., 2010. A Study on the Effect of the Stress State on Ductile Fracture. International Journal of Damage Mechanics 19,
7594.
Gurson, A., 1977. Continuum Theory of Ductile Rupture by Void Nucelation and Growth: Part I Yield Criteria and Flow Rules for Porous Ductile
Media. Journal of Engineering Materials and Technology 99, 2-15.
Han, X., Besson, J., Forest, S., Tanguy, B., Bugat, S., 2013. A yield function for single crystals containing voids. International Journal of Solids
and Structures 50, 21152131.
Hancock, J., Mackenzie, A., 1976. On the mechanisms of ductile failure in high-strength steels subjected to multi-axial stress-states. Journal of the
Mechanics and Physics of Solids 24, 147160.
Keralavarma, S.M., Benzerga, A.A., 2010. A constitutive model for plastically anisotropic solids with non-spherical voids. Journal of the Mechanics
and Physics of Solids 58, 874901.
Keralavarma, S.M., Hoelscher, S., Benzerga, A.A., 2011. Void growth and coalescence in anisotropic plastic solids. International Journal of Solids
and Structures 48, 16961710.
Monchiet, V., Cazacu, O., Charkaluk, E., Kondo, D., 2008. Macroscopic yield criteria for plastic anisotropic materials containing spheroidal voids.
International Journal of Plasticity 24, 11581189.
Paux, J., Morin, L., Brenner, R., Kondo, D., 2015. An approximate yield criterion for porous single crystals. European Journal of Mechanics -
A/Solids 51, 110.
Rice, J., Tracey, D., 1969. On the ductile enlargement of voids in triaxial stress fields. Journal of the Mechanics and Physics of Solids 17, 201217.
Saai, A., Dumoulin, S., Hopperstad, O.S., Lademo, O.G., 2013. Simulation of yield surfaces for aluminium sheets with rolling and recrystallization
textures. Computational Materials Science 67, 424433.
Steglich, D., Wafai, H., Besson, J., 2010. Interaction between anisotropic plastic deformation and damage evolution in Al 2198 sheet metal.
Engineering Fracture Mechanics 77, 35013518.
Tvergaard, V., 1981. Influence of voids on shear band instabilities under plane strain conditions. International Journal of Fracture 17, 389407.
Zhang, K., 2001. Numerical analysis of the influence of the Lode parameter on void growth. International Journal of Solids and Structures 38,
58475856.

You might also like