Sie sind auf Seite 1von 16

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/256699352

Failure Prediction in the Hole-Flanging Process


of Aluminium Alloys

Article in Engineering Fracture Mechanics · February 2013


DOI: 10.1016/j.engfracmech.2012.12.018

CITATIONS READS

19 403

5 authors, including:

Abdelkader Krichen Pierre-yves Manach


University of Sfax Université de Bretagne Sud
29 PUBLICATIONS 145 CITATIONS 113 PUBLICATIONS 868 CITATIONS

SEE PROFILE SEE PROFILE

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

Micro-forming of Ultra-thin Stainless Steel Sheets View project

Benchmark 3 Numisheet 2016 View project

All content following this page was uploaded by Pierre-yves Manach on 21 December 2013.

The user has requested enhancement of the downloaded file.


Engineering Fracture Mechanics 99 (2013) 251–265

Contents lists available at SciVerse ScienceDirect

Engineering Fracture Mechanics


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

Failure prediction in the hole-flanging process of aluminium


alloys
A. Kacem a, A. Krichen a, P.Y. Manach b,⇑, S. Thuillier b, J.W. Yoon c
a
LGPMM, National Engineering School of Sfax, University of Sfax, BP 1173, 3038 Sfax, Tunisia
b
LIMATB, Université Européenne de Bretagne, Université de Bretagne-Sud, BP 92116, 56321 Lorient, France
c
Faculty of Engineering and Industrial Sciences, Swinburne University of Technology, P.O. Box 218, Hawthorn, Vic 3122, Australia

a r t i c l e i n f o a b s t r a c t

Article history: The objective of this paper is to characterize and predict numerically the limits of the hole-
Received 24 April 2012 flanging process arising from material failure for two different aluminium sheets. Firstly,
Received in revised form 13 November 2012 an analysis of the types of failure, i.e. necking and cracks, that appear under different pro-
Accepted 20 December 2012
cess conditions has been performed. Then, a fracture criterion based on local strain mea-
sures in tension has been identified for both materials. A particular attention has been
paid to the modelling and identification of the constitutive law in a large strain range,
Keywords:
and then numerical predictions of the strain limits obtained from successful parts were
Hole-flanging
Ironing
compared to experimental results. The main conclusions are focused on the occurrence
Fracture of damage in the conventional hole-flanging process from numerical simulations and
Finite element model experiments, depending on several process parameters.
Damage ! 2013 Elsevier Ltd. All rights reserved.

1. Introduction

In various industrial applications, the forming of metal sheets is performed by stretching that induces tensile stresses and
causes thinning of the sheet, leading sometimes to the failure of the part. Among forming processes, hole-flanging consists in
expanding a hole in a sheet blank with a conical, cylindrical or hemispherical punch into a die, in order to produce a more or
less long hollow cylinder, as presented in Fig. 1. In this process, the highest strains are located at the periphery of the ex-
panded hole, because there is an increase of the diameter of the initial hole while the thickness of the flanged edge is re-
duced. This deformation mode corresponds to expansion, in which several defects may occur such as necking or cracks
leading to fracture [1]. The appearance of such defects is closely related to the magnitude of the expansion, mainly due to
the tensile stress component in the circumferential direction at the edge of a flanged hole.
Several investigations have already been performed for the hole-flanging process of various materials, including efforts to
determine the relationship between ductility in tension and damage in the process [2]. Johnson et al. [3] performed an exper-
imental study on the deformation of circular plates leading to the fracture of the flanged edge. The influence of the material
properties as well as the process geometry on this type of failure was also investigated: diameter of the initial hole [1,4],
corner radius of a cylindrical punch [4], surface roughness of the initial hole [5], punch shape [6]. The results showed that
strain path is independent of punch shape during forming process, conversely to the maximum punch load that depends on
it. Takuda et al. [7] suggested that the forming limits in deep drawing and hole expansion processes decrease with the punch
profile radius, according to numerical simulation and a ductile fracture criterion. Most of these studies are based on the cri-
terion of a critical thickness, which can be determined experimentally by considering that it equals the critical thinning at

⇑ Corresponding author. Tel.: +33 678717542.


E-mail address: pierre-yves.manach@univ-ubs.fr (P.Y. Manach).

0013-7944/$ - see front matter ! 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engfracmech.2012.12.018
252 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

Nomenclature

A initial yield stress in tension


b growth rate of the yield surface
B!A isotropic hardening range in tension
c clearance between the punch and the die
C rate of saturation of isotropic hardening
C1,C2 material parameters
E Young’s modulus
L characteristic length of a finite element
n the hardening exponent
Q maximum change in the size of the yield surface
r0, r90, r45 plastic anisotropy coefficients
!r normal anisotropy coefficient
Rc clearance-thickness ratio
Rd die radius
t workpiece thickness
ud equivalent plastic displacement at the complete damage
uf equivalent plastic displacement at the onset of fracture
up equivalent plastic displacement
UTS ultimate tensile stress
V measure of damage of a finite element after the onset of the fracture
Vc critical value of V
YTS yield tensile stress
Dr planar anisotropy coefficient
! logarithmic strain
!!f equivalent plastic strain at onset of fracture
!!p equivalent plastic strain
g stress triaxiality
g0 mean value of the stress triaxiality in tension
x accumulated damage
r Cauchy stress in tension
r Cauchy stress tensor
r! equivalent stress

fracture of a tensile sample [1,4,6,8–11]. Such criterion was firstly used in [12,13] to predict the limits of the deep drawing
process, but later, the ability of this criterion to deal also with the hole-flanging process was checked for several punch
shapes [6]. However, these studies allow only to predict necking. Other types of defects can be generated in the flanged edge
during the process that are invisible from the surface. Moreover, no previous studies focus on damage occurrence in the
flanged edge whether the process involves ironing or not. Finite element simulation seems to be an appropriate tool for
the prediction of such defects.

Fig. 1. Tool geometry parameters of the conventional hole-flanging process.


A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 253

As the hole is expanded during the process, a crack tends to occur at the flanged tip. The prediction of the forming limit by
hole expansion may require the use of micromechanisms inspired fracture models [14,15], but phenomenological models
have also been developed to predict ductile fracture without modelling void nucleation and growth. It is assumed that frac-
ture occurs at a specimen location where a weighted measure of the accumulated plastic strain reaches a critical value. A
comparative study of various weighting functions [16] showed that none of them can accurately describe the fracture behav-
iour of a given material over a large range of stress triaxialities. Attempts to define a more general fracture criterion led to the
introduction of the third invariant of the stress tensor in the weighting function (e.g. [17]), or to the transformation of stress
based fracture criteria into the space of stress triaxiality, Lode angle and equivalent plastic strain [18]. However, Chung et al.
[19] recently showed that a rather simple triaxiality-dependent fracture criterion could also lead to valuable results in the
prediction of damage in the hole-flanging process.
In this study, the hole-flanging process was experimentally and numerically investigated in order to determine the limits
of the process, by using a ductile fracture criterion as proposed in [19]. To account for crack formation, an inverse calibration
method based on a damage model using a stress triaxiality-dependent fracture criterion and hardening behaviour with stiff-
ness deterioration was used. The damage model was inversely calibrated by performing numerical simulations and exper-
iments for tension of a straight sample up to fracture. Then, the damage model was applied to the conventional hole-flanging
process. In addition, scanning electron microscope (SEM) observations of fractured surfaces were performed. In order to ver-
ify the proposed criterion, numerical simulations of the hole-flanging process were performed using Abaqus and compared
with experimental data for two aluminium alloys. Practically, it was observed that under certain conditions, a satisfying
flanged edge can be produced with the conventional hole-flanging process without requiring the use of a fineblanked
hole-flanging process. The first part highlights several cases of failure in the hole-flanging process, then identification of
the damage model is introduced and finally, finite element predictions are compared with experiments.

2. Limits of the hole-flanging process

In the conventional hole-flanging process, damage in the flanged edge often occurs, as described in [5], showing particular
problems such as failure at the flanged tip and cracks. In the next section, such problems are highlighted by means of exper-
imental data.

2.1. Hole-flanging process

Hole-flanging process was carried out with a dedicated conventional hole-flanging device for sheet metal settled on a
Zwick-Roell BUP200 testing machine, presented in Fig. 1 and described in details in [20]. All tests were performed without
lubricant, under displacement control and at a constant speed of the punch equal to 5 mm/min. The punch displacement and
the load were continuously recorded with a data acquisition system. Both a 2 mm thick A1050-H14 pure aluminium and a
AA6061-O Mg–Si were considered. Samples were shear cut in squares of 30 mm side and a 6 mm diameter hole was drilled
in the centre. During hole-flanging, the sample was clamped between the die and the blank-holder by applying a force of
5 kN [21]. The diameter of the punch is 12 mm and the half-angle at the top of the cone is 30". Some parameters determine
whether the hole-flanging process is performed with or without ironing, typically by varying the clearance (c) between the
punch and the die while keeping the other parameters unchanged. In practice, to control the occurrence of ironing it is con-
venient to consider the clearance-thickness ratio Rc = c/t which is defined as the ratio of the clearance (c) to the workpiece
thickness (t) [20]. Two die radii Rd of 7 and 8 mm leading to clearance-thickness ratios Rc of respectively 0.5 and 1 have been
used. Two other die radii of Rd equal to 6.50 and 6.82 mm (i.e. Rc equal to 0.25 and 0.41, respectively) have also been tested in
order to study the influence of higher ironing levels of [20].

2.2. Materials and mechanical properties

Monotonic tensile tests were carried out on rectangular samples of gauge dimensions 150 mm " 20 mm under displace-
ment control at a constant rate of 5 mm/min, according to a procedure defined in [22]. The load was measured by a load cell
of 50 kN capacity. The samples were initially shear cut, then machined to eliminate the hardened area due to cutting. The
edges were then locally polished with sandpaper, to reduce the width by some tenths of a millimeter, in order to obtain
necking in the centre of the sample. Cauchy stress r was calculated by the ratio of the load over the current section of
the sample before necking. The current section was determined from the initial section of the sample by assuming both iso-
choric plastic deformation and a homogeneous distribution of strain over the gauge length of the sample. The logarithmic
strain ! was measured by a field measuring system by 3D digital image correlation (DIC Aramis) as defined in [22]. The
use of the DIC technique allowed to obtain also the nominal stress–strain curves by defining a virtual extensometer on
the sample, which gauge length is 28.86 mm. In the range of homogeneous strains, the strain calculated from the virtual
extensometer displacement was similar to the one measured by the DIC system. The obtained stress–strain curves in the
rolling direction are shown in Fig. 2 for both materials.
The mechanical properties of both materials are presented in Table 1. To determine the material anisotropy, tensile tests
were also performed at 45" and 90" to the rolling direction (RD). The measurement of both transverse and longitudinal
254 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

Fig. 2. Nominal and Cauchy tensile stress–strain curves for (a) the A1050-H14 alloy and (b) the A6061-O alloy.

Table 1
Mechanical properties of both materials.

Material E (GPa) YTS (MPa) UTS (MPa) r0 r45 r90 !r Dr


A1050-H14 68 107 115 0.59 0.61 0.75 0.64 0.06
A6061-O 70 52 136 0.54 0.74 0.63 0.66 !0.15

strains leads to the plastic anisotropy coefficients given in Table 1. The normal anisotropy coefficient !r ¼ ðr0 þ r 90 þ 2r45 Þ=4 is
significant for both materials but the planar anisotropy coefficient Dr = (r0 + r90 ! 2r45)/2 is rather weak. Young’s modulus
was determined from the experimental stress–strain curves and Poisson’s ratio was taken equal to 0.3.

2.3. SEM observations

As in most of the forming processes, the hole-flanging process is limited by the apparition of several defects. Observations
of the surface have been performed using a Scanning Electron Microscope for both materials on both sides of the flanged
edge and typical defects are gathered in Fig. 3.
For the A1050-H14 alloy, an orange peel aspect has been observed. For hole-flanging with and without ironing, this defect
is localized in the junction between the flanged edge and the non-deformed part of the workpiece (Fig. 3a). In the case of
hole-flanging without ironing, this defect is also observed in the outer surface of the flange (Fig. 3b). Conversely, for the
A6061-O alloy, no orange peel aspect has been observed on the samples from naked eye observations. At the flanged tip,
neckings appear for both materials only when hole-flanging tests are performed without ironing (Fig. 3b). For both materials,
hole-flanging tests lead to an inner surface of the obtained hollow cylinder without apparent defect.

Fig. 3. Several types of defects in hole-flanging: (a) orange peel aspect (A1050-H14), (b) orange peel aspect and necking (A1050-H14), (c) radial fracture
(A6061-T4) and (d) axial crack (A6061-T4).
A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 255

To be compared, hole-flanging tests without ironing were also performed on a material presenting a weak ductility
(A6061-T4). In this case, radial fractures are observed near the junction zone between the flanged edge and the non-
deformed part of the workpiece (Fig. 3c), while cracks start from the flanged tip (Fig. 3d). All observed defects such as orange
peel aspect, neckings, axial and radial cracks are consistent with previous studies [1,2].

3. Identification of a damage model in tension

3.1. Macroscopic fracture criterion

A phenomenological fracture criterion taking into account the triaxiality state of the material has been used to determine
the onset of the macroscopic fracture. This criterion characterizes the damage state in the material and when a critical value
is reached, it leads to the onset of a macroscopical fracture. This criterion is expressed according to [19,23]:
Z !!f
d!p
x¼ ð1Þ
0 !f ðgÞ

where x is a state variable that indicates the onset of fracture in a given element when it reaches the unity. It should be
noted that x is also used as a damage indicator to appreciate the damage state. In the approach, work hardening and damage
are uncoupled, that means that until the criterion does not reach its critical value, mechanical properties are not affected by
the damage. In Eq. (1), !!p is the equivalent plastic strain and g is the stress triaxiality defined by g ¼ trðrÞ=3r! , where r is the
Cauchy stress tensor and r ! the equivalent stress. !f is the equivalent plastic strain at onset of fracture. It should be noted that
!!f may also depend on the Lode angle as mentioned in [18,24], but this dependency has not been considered here.
According to [18], the evolution of !f as a function of g can be determined experimentally using tensile tests with different
sample shapes leading to several stress triaxialities [25,26]. When such tests are not available, its evolution can also be
approximated analytically by the relation used by [27]:
8
>
>
1 g 6 !1=3
>
< C 1 =ð1 þ 3gÞ !1=3 < g 6 0
!f ðgÞ ¼ ð2Þ
> C 1 þ ðC 2 ! C 1 Þðg=g0 Þ2
> for 0 6 g 6 g0
>
:
C 2 g0 =g g P g0
pffiffiffi
with g0 is the mean value of the stress triaxiality in tension and C 1 ¼ C 2 ð 3=2Þ1=n with n the hardening exponent. The iden-
tification of these parameters has been performed using the hybrid method proposed in [25], involving an experimental
load–displacement curve and a finite element model of the tensile test. C2 is equal to the maximum equivalent plastic strain
reached for the fracture displacement uf which corresponds to the sharp drop-off of the load. The mean stress triaxiality g0 is
then calculated according to [25], and the evolution of ! !f with g can be plotted.

3.2. Softening behaviour after fracture initiation

The onset of fracture represents the first stage of the complete failure of the tensile sample experimentally, and of a finite
element of the mesh numerically. Another variable V is introduced to measure the damage of a finite element after the onset
of the fracture [28]. A linear evolution of V with respect to the equivalent plastic strain is chosen such as:

dup d!p
dV ¼ ¼L ð3Þ
ud ud
where L is the characteristic length of the element, up is the equivalent plastic displacement of the element and ud is the
equivalent plastic displacement at the complete damage. The element looses gradually its rigidity due to the evolution of
V to a complete damage V = 1 until for up = ud, it is deleted from the mesh. However, for high values of V, the stiffness of
the element decreases leading to its excessive distortion. To solve this problem, the element was not deleted for V = 1 but
rather for a critical value Vc < 1, such as in numerical techniques using cohesive elements. Both parameters ud and Vc have
been determined by an inverse method that consists in the numerical simulation of a tensile test with the parameters iden-
tified in Eq. (2). The values of ud and Vc leading to the first element deletion corresponding to the experimental displacement
of the complete fracture of the specimen, are kept. It has to be noted that the element deletion requires a reasonable fine
mesh, particularly in the necking zone. In this case, the deletion of an element leads to the deletion of all the damaged ele-
ments which gives a more realistic description of the fracture of the specimen. However, the use of a very fine mesh gener-
ates other difficulties that are addressed later in the paper.

3.3. Numerical simulation of the tensile test

Numerical simulations of the tensile test were performed in 3D with the explicit finite element code Abaqus v6.10, as for
dealing with necking problems, the loss of uniaxiality of the tensile test requires the use of a 3D model. The finite element
256 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

mesh as well as the boundary conditions of the numerical tensile test are presented in Fig. 4. Similarly to the experimental
test, a small notch (0.1 mm) has been done in the mesh, the size of which was adjusted in order to obtain a rupture onset
located on the edge and in the middle of the sample. The nodes located on the left side are clamped while those located on
the right side are moved with a smooth increasing velocity up to 15 mm/s. In order to reduce the calculation time, the mass-
scaling option was used with parameters leading to a kinetic energy lower than 5% of the internal energy.
According to [29], shell elements are not suitable for the numerical simulation of a tensile test after necking. Then, solid
elements C3D8R (Continuum 3D 8-nodes Reduced integration) were used. Dunand and Mohr [29] have also shown that the
element size modifies significantly the value of !
!f . To determine accurately the value of !!f , a finer mesh of about 0.2 mm was
used. The same element size is kept for the numerical simulation of the hole-flanging process to avoid the problem of the
sensitivity of !f to the mesh size. Since the element size does not have a large influence on the force–displacement curve
[29], the size of the elements was fixed to 0.4 mm for the identification of the work-hardening relationship to speed up
the computation.

3.4. Work-hardening

Isotropic hardening with von Mises criterion was chosen to model the mechanical behaviour. As previously mentioned,
the homogeneous strain range is limited in tension by the occurrence of necking where deformation is localized and can
reach high values. Necking is followed by a macroscopical fracture leading to the rupture of the specimen. After necking,
a classical approach consists in extrapolating the hardening curve up to a large strain range. This leads generally to a poor
modelling by presenting a prematurate drop-off of the load. On the basis of these successive phenomena as a function of
strain, the constitutive behaviour was established as follows: the homogeneous strain range was modelled using a classical
work-hardening law, the localized strain range was described by a linear piecewise relation identified by an inverse method,
the occurrence of the macroscopic fracture was determined by identifying a specific criterion, and finally the damage evo-
lution of the elements after the onset of fracture was modelled. Such a method was adopted by several authors to study the
limits of forming process (e.g. [30–32]) and has been recently improved in [19].
For the isotropic hardening before necking, Voce relationship was retained for the A6061-O alloy:

r ¼ r0 þ Q ð1 ! expð!b!p ÞÞ ð4Þ

where Q represents the maximum change in the size of the yield surface, and b defines the growth rate of the yield surface
(or hardening). For the A1050-H14 alloy, the Hockett–Sherby relationship [33] was used:
" #
r ¼ B ! ðB ! AÞ exp !C !!np ð5Þ

where A, (B-A) and C are the initial yield stress in the RD, the isotropic hardening range and the rate at which saturation is
achieved, respectively. These parameters were identified in order to obtain the best fit between the experimental and
numerical stress–strain curve on a tensile test in the RD.
Considère criterion was used to distinguish the different strain ranges. It was assumed that the diffuse necking begins when
the tensile force reaches its maximum, i.e.: dF = 0 or dr/d! = r for an isotropic material. Fig. 2 shows that it corresponds to a
tensile strain of 0.019 and 0.16 respectively for the A1050-H14 and the A6061-O alloy leading to a displacement of 0.52 mm
and 5.2 mm, respectively. The work-hardening relationships, fitted within these strain ranges, are presented in Fig. 5.
Fig. 6 presents an extrapolation of the work-hardening laws corrected with a linear piecewise hardening and without cor-
rection up to necking. It is observed that the hardening cannot be described satisfactorily without correction due to the sat-
uration of the Voce law (see Fig. 6a), but also with non-saturating hardening laws [32] (see Fig. 6b). However for both
materials, the curve corrected by the piecewise linear hardening in the localized strain range is in a good agreement with
experimental results.

Fig. 4. Finite element model of the tensile test.


A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 257

Fig. 5. Identification of the true tensile curves for (a) A1050-H14 and (b) A6061-O alloys in the homogeneous strain range.

Fig. 6. Force–displacement curves for (a) A1050-H14 alloy and (b) A6061-O alloy.

To analyze the effect of the element size of the mesh on the identification results, the force–displacement curves obtained
with a finer mesh size of 0.2 mm are superimposed in Fig. 6. It is observed that the mesh has no influence on the curve before
necking. However, a slight difference is observed in the necking zone, particularly after the onset of localized necking, which
does not affect the good agreement between experiments and simulation. The material parameters are reported in Tables 2
and 3 for the A1050-H14 alloy and the A6061-O alloy, respectively.

3.5. Onset of fracture

Based on the DIC measurements, the instant of onset of fracture (not the location) was defined by the first detectable dis-
continuity in the measured displacement field of the tensile test. Subsequently, a finite element simulation was performed
which post-processing gave access to the evolution of the stress triaxiality and the equivalent plastic strain. Due to the main
tensile stress state in the specimen, it was assumed that the location of the onset of fracture coincides with the location of
the highest equivalent plastic strain within the specimen at the instant of onset of fracture. The corresponding equivalent
plastic strain is referred to as fracture strain !
!f . The displacements uf corresponding to the onset of fracture are presented
in Fig. 2, uf = 3.05 mm for the A1050-H14 alloy and uf = 9.2 mm for the A6061-O alloy, respectively. In both cases, the element
presenting the maximum equivalent plastic strain for the displacement uf has been located in the centre of the geometrical
default (see Fig. 7). The value of the maximum strain C2 is 1.32 for the A1050-H14 alloy and 1.26 for the A6061-O alloy,
respectively.

Table 2
Material parameters of the Hockett–Sherby hardening law for the A1050-H14 alloy corrected by a linear piecewise relation after necking.

Before necking ð0 6 !
!p < 0:0082Þ After necking ð!
!p P 0:0082Þ
A (MPa) B (MPa) C n !p 0.07 0.2 0.4

33 116 52 0.5 r (MPa) 123 126 130


258 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

Table 3
Material parameters of the Voce hardening law for the A6016-O alloy corrected by a linear piecewise relation after necking.

Before necking ð0 6 !p < 0:095Þ After necking ð!p P 0:095Þ

r0 (MPa) Q (MPa) b !p 0.1 0.2 0.3 0.4 0.5 0.6

48 106 25 r (MPa) 146 166 180 192 202 210

Fig. 7. Distribution of the longitudinal strain for several displacement values for the A1050-H14 alloy (upper part) and for the A6061-O alloy (lower part).

Fig. 7 displays the distribution of the longitudinal strain for both alloys, measured by the DIC system and obtained by
numerical simulations for different longitudinal displacements (1 mm and 2.6 mm for the A1050-H14 alloy, 3.4 mm and
6.8 mm for the A6061-O alloy). It can be observed that in both cases, the numerical simulations are close to experimental
distributions, for the strain values as well as for the location of necking zones.
Table 4 gives the values of C1,C2, and g0. These values lead to the evolution of the fracture strain !
!f as a function of the
stress triaxiality g presented in Fig. 8.

3.6. Element deletion

The evolution of V requires only the identification of ud and Vc. This was done by an inverse method with the procedure
defined previously. It is observed in Fig. 2 that the complete fracture of the tensile sample occurs for a displacement of

Table 4
Material parameters of the onset of fracture.

Material C1 C2 g0
A1050-H14 0.64 1.32 0.49
A6061-O 0.74 1.26 0.39
A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 259

Fig. 8. Evolution of the fracture strain as a function of the stress triaxiality for both aluminium alloys.

Fig. 9. Comparison of the numerical and experimental fracture profiles for (a) the A1050-H14 alloy and (b) the A6061-O alloy.

4.2 mm for the A1050-H14 alloy and 9.8 mm for the A6061-O alloy, respectively. The values of ud and Vc that generate the
complete damage of an element for these displacement values are ud = 0.5 mm with Vc = 0.55 for the A1050-H14 alloy and
ud = 0.3 mm with Vc = 0.45 for the A6061-O alloy, respectively. Fig. 9 presents the shape of the fracture section obtained
numerically with previous values of ud and Vc and experimentally for both materials. It has to be noticed that a more accurate
value of this parameter could be obtained by using a very fine mesh in the necking zone, in order to obtain the damage of all
elements of the necking zone when this first one is damaged. Due to the dependency of ! !f to the mesh, this requires to
recalculate !f ¼ f ðgÞ for this very fine mesh which is highly time consuming.

4. Numerical prediction of the limits in hole-flanging

3D Finite element (FE) simulations were performed to highlight the damaged zones in the hole-flanging process and to
define the limits of this process. The blank was discretized with finite elements which formulation (C3D8R) and size
(0.2 mm) are similar to the one used to identify !f . To minimize the influence of mesh size in the results, adaptive meshing
or remeshing techniques were not used. The number of elements in the blank was 21,623. Tools in contact with the sheet
(punch, die and blank-holder) were considered as rigid bodies. The contact between the tools and the blank was assumed
with friction governed by the Coulomb’s law with constant friction coefficient of 0.17. During the simulation, the die and
the blank-holder were fixed. The punch was moved at a constant velocity of 15 mm/s similar to that used for the identifi-
cation in tension. The FE model is presented in Fig. 10.
To model potential fracture of the flanged edge, the element reaching the critical damage value Vc was deleted from the
mesh. To ensure that the inertial effects due to the mass scaling option and the hourglassing control of reduced integration
elements do not affect the results, both the kinetic and artificial strain energy remained negligible compared to the internal
energy.
Two cases were considered, the hole-flanging with and without ironing, in order to analyze the damage distribution in
each condition. Ironing is performed by setting a low value of the clearance (c) between the punch and the die. In this case,
the metal is squeezed between the punch and the die leading to a longer flange. Ironing is thus mainly a kind of plane strain
deformation in which the reduction of sheet thickness contributes to the increase of flange height. When the process is per-
formed without ironing, the flange is formed by edge stretching leading to a thinner flange. The study of both conditions has
260 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

Fig. 10. 3D finite element model of the conventional hole-flanging process (for sake of clarity, the blank-holder is not represented).

Fig. 11. Distribution of the stress triaxiality in the conventional hole-flanging process (A1050-H14).

an increasingly engineering interest, as in many industrial applications, the process is performed with ironing to obtain a
long flange that can be used for example to increase bearing surface or to increase the number of threads that will fit in
a tapped hole. To characterize ironing, the clearance-thickness ratio Rc is used: when Rc = 1 the process is performed without
ironing while when Rc = 0.5 it is performed with ironing.

4.1. Stress triaxiality in hole-flanging

The analysis of the distribution of the stress triaxiality in the flanged edge during hole-flanging is relevant to assess the
strain state, to determine the fracture modes and also to compare the range with the evolution of ! !f . Fig. 11 shows the dis-
tribution of the stress triaxiality in the flanged edge for different punch displacements of hole-flanging processes performed
with and without ironing.
Fig. 11 shows that the stress triaxiality is mainly between 0 and 0.49, which is coherent with the use of the tensile test for
identification (initial stress triaxiality of 0.33). Values above this range occur at the junction between the flanged edge and
the non-deformed part of the workpiece, on the inner profile, indicating a favourable region for the nucleation of ductile frac-
ture. The contact zones between the tools and the blank present mainly negative stress triaxiality values lower than !0.33,
preventing the flanged edge from fracture. For the hole-flanging without ironing, the lowest values of stress triaxiality are
observed near the die radius indicating a more safety area, as values of stress triaxiality lower than !0.33 do not lead to
fracture [34]. However, for the hole-flanging process performed with ironing, increasing values of the stress triaxiality in
the vicinity of the die radius are recorded as the punch displacement increases.
A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 261

Fig. 12. Distribution of damage x in the flanged edge for (a) the A1050-H14 alloy and (b) the A6061-O alloy and (c) localized necking at the flanged tip for
the A6061-O alloy.

4.2. Damage in hole-flanging

The process is firstly performed without ironing (Rc = 1) and the parameter x is used as an indicator to appreciate the
damage state in the flanged edge. The closer the value of x is to the unity, the higher is the damage.
Figs. 12a and 12b present the distribution of damage in the flanged edge calibrated according to 5 different levels of x for
both aluminium alloys. It can be observed that there are no zones of severe damage for the A6061-O alloy, the maximum
value reaches 0.77 in this case. Conversely, for the A1050-H14 alloy, the maximum value of x reaches 0.99. The most severe
damage zone is located at the flanged tip. Experimentally, localized neckings have also been recorded in this zone, as ob-
served in Figs. 3b and 12c. For both grades, a lower damage is observed in the junction between the flanged edge and the
non-deformed part of the workpiece and on the outer profile (die side). SEM observations with high magnification show that
these zones present experimentally an orange peel aspect, as shown in Fig. 13, that is more pronounced for the 1050-H14
alloy than for the A6061-O alloy.
Finally, the inner surface of the flanged edge is characterized by a weak value of damage that is coherent with the SEM
observations presented in Fig. 14 which do not exhibit any defect. In sheet metal forming processes, the area located in the
vicinity of the die radius is generally favourable for crack initiation. Therefore, the die radius is one of the most sensitive
parameter for all these processes. However, it is observed that it is not the case for the hole-flanging process performed with-

Fig. 13. Orange peel aspect in the junction zone between the flanged edge and the non-deformed part of the workpiece for (a) the A1050-H14 alloy
and (b) the A6061-O alloy.
262 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

Fig. 14. Aspect of the inner surface of the wall of the flanged edge for (a) the A1050-H14 alloy and (b) the A6061-O alloy.

out ironing. According to Fig. 12, this area is characterized by a weak damage that can be explained by the negative stress
triaxiality in this zone. This observation is also confirmed experimentally: the use of a very small die radius, even no die ra-
dius at all, has no influence on the success of the hole-flanging process without ironing.
Fig. 15 presents the distribution of damage in the flanged section obtained in a test performed with Rc = 0.5. Conversely to
the test performed without ironing, severe damage is observed in large zones for both materials. These areas are mainly lo-
cated in the centre of the flanged edge, not directly on the surface, which explains that no defects are clearly observed exper-
imentally with the naked eye on the surface of the samples. However, SEM observations with high magnification confirmed
the presence of scratches on the inner surface of the flanged edge (punch side).
According to these results, it appears that the external surfaces remain weakly damaged while the inner core of the
flanged edge is highly damaged. Then, in the hole-flanging process with ironing, subsequent forming operations that involve

Fig. 15. Distribution of damage x in the flanged section for (a) the A1050-H14 alloy and (b) A6061-O alloy and SEM observations of scratches on the inner
surface of the flanged edge for (c) the A1050-H14 alloy and (d) A6061-O alloy.
A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 263

Table 5
Limits of ironing in the hole-flanging process in the experimental and numerical process for the
1050-H14 alloy, ': successful, ": unsuccessful.

Clearance-thickness ratio (Rc) Rc = 0.5 Rc = 0.41 Rc = 0.25


Experimental ' ' "
Numerical ' ' "
V maximum 0.33 < Vc 0.5 < Vc 0.55 = Vc

Fig. 16. (a) safe product obtained with high level of ironing (Rc = 0.41), (b) failed product obtained with very high level of ironing (Rc = 0.25)
and (c) numerical simulation of the rupture of flanged edge (Rc = 0.25).

the deformation of the surface can be successfully performed (e.g. threading [35]) while those involving a plastic deforma-
tion of the core of the flanged edge can be unsuccessful (e.g. hemming [30]).
By comparing the damage state between both materials, it can be seen that the flanged edge of the A1050-H14 alloy is more ex-
posed to fracture than the A6061-O alloy. The A1050-H14 alloy was thus retained to study the limits of ironing in the next section.

4.3. Limits of ironing in hole-flanging

To determine the rupture of the flanged edge numerically, the criterion of deletion of finite elements is used. This criterion
suggests that an element reaching a damage value of V = Vc is deleted from the mesh. Two low values of Rc were considered
in order to perform the hole-flanging process with high level of ironing that may lead to the rupture of the flanged edge.
Table 5 presents the numerical and the experimental results for hole-flanging operations performed with the considered
clearance-thickness ratios. A good agreement between numerical simulation and experimental results is observed. Experi-
mentally and numerically, the limits of ironing in the hole-flanging process are lower than 0.41 for the A1050-H14 alloy. Figs.
16a and 16.b present safe and failed products obtained for Rc equal to 0.41 and 0.25, respectively. Fig. 16c shows the result
obtained from a very high level of ironing (Rc = 0.25) leading to the fracture of the flanged edge. A good agreement between
numerical simulation and experimental results is also observed. However, a large distortion of the finite element is recorded
before its deletion, probably due to the coarse mesh used in the identification of the parameter controlling the element
deletion and the fracture simulation. It appears that an adaptive mesh refinement in the necking zones, or a cohesive ele-
ment formulation, is required to improve the prediction of fracture in the hole-flanging process.

Fig. 17. Distribution of damage x in the flanged section for the A1050-H14 alloy using (a) the von Mises yield criterion and (b) the Hill 1948 yield criterion.
264 A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265

4.4. Influence of material anisotropy

To analyze the effect of material anisotropy on the damage prediction in the hole-flanging process, a numerical simulation
of the hole-flanging without ironing has also been performed with the Hill 1948 yield criterion. The material parameters of
the anisotropy behaviour are listed in Table 1. Fig. 17 shows a comparison of the distribution of damage x obtained using
von Mises criterion and Hill’s (1948) criterion, both in the rolling and in the transverse directions for the A1050-H14 alloy. It
is observed that the influence of anisotropy on the distribution of damage is weak for this material. The comparison between
the two orientations shows that damage is higher in the rolling direction, that is coherent with the lower anisotropy ratio in
this direction (r0 = 0.59 < r90 = 0.75), thus indicating a lower resistance to thinning and necking.
It has to be noticed that !f was considered similar whatever the orientation of the sample to the anisotropy directions.
This is generally not the case experimentally but requires a more complex identification of ! !f for tensile tests performed
at several orientations to the RD.

5. Conclusion

This study deals with the characterization of the evolution of damage and the definition of the limits in the hole-flanging
process. In this process, the stress state leads to a local thinning of the sheet which excess can generate necking or even
cracks. The evolution of damage in the hole-flanging process is predicted by finite element analysis using a ductile fracture
criterion and is compared with experimental tests and SEM observations. By a careful identification of the ductile fracture
criterion on two aluminium alloys (A1050-H14 and A6061-O), based on the numerical simulation of the fracture of a tensile
test, it is shown that the distribution of damage is different whether or not the process is performed with ironing. The main
results obtained are:

( For the hole-flanging process without ironing, the damage is mainly localized in the junction between the flanged edge
and the non-deformed part of the workpiece (peel orange aspect) and at the flanged tip (localized necking).
( For the hole-flanging process with ironing, the damage is localized in the inner core of the flanged edge, but not in the
surface of the sample. That explains why subsequent processes that involve deformation of the surface (i.e. threading)
can be successfully performed.
( Conversely to most of the sheet forming processes, the die radius has a weak influence on the damage of the flanged edge,
due to low negative stress triaxiality values.
( In the numerical simulation, the influence of anisotropy on the damage distribution is weak.

References

[1] Huang YM, Chien KH. The formability limitation of the hole-flanging process. J Mater Process Technol 2001;117:43–51.
[2] Hyun DI, Oak SM, Kang SS, Moon YH. Estimation of hole flangeability for high strength steel plates. J Mater Process Technol 2002:9–13.
[3] Johnson W, Chitkara N, Minh H. Deformation mode and lip fracture during hole-flanging of circular plates of anisotropic materials. Trans ASME J Engng
Ind 1977;99:738–48.
[4] Huang YM, Chien KH. Influence of the punch profile on the limitation of formability in the hole-flanging process. J Mater Process Technol
2001;113:720–4.
[5] Thipprakmas S, Jin M, Murakawa M. Study on flanged shapes in fineblanked-hole flanging process (FB-hole flanging process) using finite element
method (FEM). J Mater Process Technol 2007:128–33.
[6] Leu DK, Chen TC, Huang YM. Influence of punch shapes on the collar-drawing process of sheet steel. J Mater Process Technol 1999;88:134–46.
[7] Takuda H, Hatta N. Numerical analysis of formability of a commercially pure zirconium sheet in some sheet forming processes. Mater Sci Engng A
1998;242:15–21.
[8] Huang YM, Chien KH. Influence of cone semi-angle on the formability limitation of the hole-flanging process. Int J Adv Manuf Technol
2002;19:597–606.
[9] Chen TC. An analysis of forming limit in the elliptic hole-flanging process of sheet metal. J Mater Process Technol 2007:373–80.
[10] Huang YM, Tsai YW, Li CL. Analysis of forming limits in metal forming processes. J Mater Process Technol 2008;201:385–9.
[11] Li CL, Huang YM, Tsai YW. The analysis of forming limit in re-penetration process of the hole-flanging of sheet metal. J Mater Process Technol
2008;201:256–60.
[12] Huang YM, Chen JW. Influence of the tool clearance in the cylindrical cup-drawing process. J Mater Process Technol 1996;57:4–13.
[13] Huang YM, Chen JW. Influence of lubricant on limitation of formability of cylindrical cup-drawing. J Mater Process Technol 1997;63:77–82.
[14] Uthaisangsuk V, Prahl U, Bleck W. Stretch-flangeability characterisation of multiphase steel using a microstructure based failure modelling. Comput
Mater Sci 2009;45:617–23.
[15] Uthaisangsuk V, Prahl U, Bleck W. Modelling of damage and failure in multiphase high strength DP and TRIP steels. Engng Fract Mech 2011;78:469–86.
[16] Bao YB, Wierzbicki T. A comparative study on various ductile crack formation criteria. J Engng Mater Technol Trans ASME 2004;126:314–24.
[17] Xue L, Wierzbicki T. Ductile fracture initiation and propagation modeling using damage plasticity theory. Engng Fract Mech 2008;75:3276–93.
[18] Bai Y, Wierzbicki T. A new model of metal plasticity and fracture with pressure and lode dependence. Int J Plasticity 2008;24:1071–96.
[19] Chung K, Ma N, Park T, Kim D, Yoo D, Kim C. A modified damage model for advanced high strength steel sheets. Int J Plasticity 2011;27:1485–511.
[20] Kacem A, Krichen A, Manach PY. Occurrence and effect of ironing in the hole-flanging process. J Mater Process Technol 2011;211:1606–13.
[21] Krichen A, Kacem A, Hbaieb M. Blank-holding effect on the hole-flanging process of sheet aluminium alloy. J Mater Process Technol 2011;211:619–26.
[22] Zang SL, Thuillier S, Le Port A, Manach PY. Prediction of anisotropy and hardening for metallic sheets in tension, simple shear and biaxial tension. Int J
Mech Sci 2011;53:338–47.
[23] Johnson GR, Cook WH. Fracture characteristics of three metals subjected to various strains, strain rates, temperatures and pressures. Engng Fract Mech
1985;21:31–48.
[24] Bai Y, Wierzbicki T. Application of extended Mohr–Coulomb criterion to ductile fracture. Int J Fracture 2010;161:1–20.
A. Kacem et al. / Engineering Fracture Mechanics 99 (2013) 251–265 265

[25] Bao Y, Wierzbicki T. On fracture locus in the equivalent strain and stress triaxiality space. Int J Mech Sci 2004;46:81–98.
[26] Giglio M, Manes A, Vigano F. Ductile fracture locus of Ti–6Al–4V titanium alloy. Int J Mech Sci 2012;54:121–35.
[27] Yu HL, Jeong DY. Application of a stress triaxiality dependent fracture criterion in the finite element analysis of unnotched Charpy specimens. Theor
Appl Fract Mech 2010;54:54–62.
[28] Yaning L, Wierzbicki T. Prediction of plane strain fracture of AHSS sheets with post-initiation softening. Int J Solids Struct 2010;47:2316–27.
[29] Dunand M, Mohr D. Hybrid experimental–numerical analysis of basic ductile fracture experiments for sheet metals. Int J Solids Struct
2010;47:1130–43.
[30] Le Maoût N, Thuillier S, Manach PY. Aluminum alloy damage evolution for different strain paths – application to hemming process. Engng Fract Mech
2009;76:1202–14.
[31] Mohr D, Ebnoether F. Plasticity and fracture of martensitic boron steel under plane stress conditions. Int J Solids Struct 2009;46:3535–47.
[32] Luo M, Wierzbicki T. Numerical failure analysis of a stretch-bending test on dual-phase steel sheets using a phenomenological fracture model. Int J
Solids Struct 2010;47:3084–102.
[33] Hockett JE, Sherby OD. Large strain deformation of polycrystalline metals at low homologous temperatures. J Mech Phys Solids 1975;23:87–98.
[34] Bao Y, Wierzbicki T. On the cut-off value of negative triaxiality for fracture. Engng Fract Mech 2005;72:1049–69.
[35] Mathurin F, Stéphan P, Daidié A, Guillot J. 3D finite elements modeling of an assembly process with thread forming screw. J Manuf Sci Engng Trans
ASME 2009;131:151–8.

View publication stats

Das könnte Ihnen auch gefallen