You are on page 1of 10

Journal of Materials Processing Technology 211 (2011) 1684–1693

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


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

Experimental and numerical study of an AlMgSc sheet formed by an incremental


process
C. Bouffioux a,∗ , C. Lequesne a , H. Vanhove b , J.R. Duflou b , P. Pouteau c , L. Duchêne a , A.M. Habraken a
a
University of Liège, Dept. ArGEnCo, Chemin des Chevreuils 1, B-4000 Liège, Belgium
b
University of Leuven, Dept. of Mechanical Engineering, Celestijnenlaan 300b, B-3001 Heverlee, Belgium
c
CRM Gent, Technologiepark 903c, B-9052 Zwijnaarde, Belgium

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

Article history: A recently developed AlMgSc alloy is studied since this material, which is well adapted to the aeronautic
Received 22 February 2011 domain, is poorly known.
Received in revised form 10 May 2011 The first objective is to reach a better knowledge of this alloy to provide the missing useful information
Accepted 15 May 2011
to the aeronautic industry and to help research institutes who want to simulate sheet forming processes
Available online 20 May 2011
by Finite Element (FE) simulations. A set of experimental tests has been performed on the as-received
sheets, material laws have been chosen and the corresponding material parameters have been adjusted
Keywords:
to correctly describe the material behaviour.
Aluminium alloy AlMgSc
Material parameter identification
The second objective is to study the applicability of the Single Point Incremental Forming process (SPIF)
Finite element simulation on this material. Truncated cones with different geometries were formed and the maximum forming
Single Point Incremental Forming angle was determined. A numerical model was developed and proved to be able to predict both the force
evolution during the process and the final geometrical shape. Moreover, the model helps reaching a better
understanding of the process.
The characterisation method described in this research and applied on the AlMgSc alloy can be extended
to other alloys. In addition, the numerical simplified model, able to accurately describe the SPIF process
with a reduced computation time, can be used to study more complex geometries.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction The Single Point Incremental Forming (SPIF) process is a flexi-


ble sheet metal forming method adapted to form various complex
The aeronautic industry is constantly looking for ways to reduce shapes using a milling machine, a multi-axis robot or a dedicated
the weight of aeronautic parts and structures. Therefore, the avail- machine without the need of specific and costly tools, such as a
ability of a new material of the AlMgSc alloy family, combining a punch and die, as described by Allwood and Shouler (2009). The
low specific weight, a high corrosion resistance, a high toughness sheet, of which the edges are clamped, is locally deformed by a
and an excellent weldability is of major interest in this domain. spherical tipped tool following a succession of contours leading to
This recently developed alloy, with compositions ranging from the final shape of the sheet.
3.1 to 5.1 at.% for Mg, 0.10 to 0.40 at.% for Sc and 0.05 to 0.20 at.% for Depending on the length of the tool path and speed of the tool,
Zr, is not commonly used. The first investigations of the material the forming process can take up to a few hours for large parts. It is, in
provider showed promising results. However, a detailed study was consequence, adapted to small batch production and rapid proto-
required to improve the knowledge of the material behaviour, to typing. Considering the maximum part size is machine dependent,
extend its applications and to study more complex processes both SPIF can be used to form small or large pieces, such as aeronau-
by experiments and by numerical investigations. tical components. Furthermore, the SPIF process, which is known
to shift Forming Limit Diagrams (FLD) to higher formability (Filice
et al., 2002) compared to conventional forming processes can be
adapted to materials such as the AlMgSc alloy.
Because SPIF has a high industrial interest, the applicability of
∗ Corresponding author. Tel.: +32 4 366 9219; fax: +32 4 366 9192. this technique on the new material studied must be experimen-
E-mail addresses: chantal.bouffioux@ulg.ac.be (C. Bouffioux), tally verified. The problem of shape inaccuracy which is usually
Cedric.Lequesne@ulg.ac.be (C. Lequesne), Hans.Vanhove@mech.kuleuven.be (H. an important limiting factor for SPIF applications, as underlined by
Vanhove), joost.duflou@mech.kuleuven.be (J.R. Duflou), pouteau@agt0.ugent.be (P.
Micari et al. (2007), must be solved. Moreover, the forming force
Pouteau), L.Duchene@ulg.ac.be (L. Duchêne), anne.habraken@ulg.ac.be
(A.M. Habraken). prediction provides crucial data to choose the forming tool, to verify

0924-0136/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2011.05.010
C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693 1685

the stiffness of the set-up and to avoid deformation and risk for the Simple tensile tests
equipment. Aerens et al. (2009) established a useful regression for- 500
mula to predict an approximate value of the axial force for any

True stress (MPa).


400
material when forming a cone. It is a function of the shape geom-
etry, the work conditions and the material tensile strength only. 300
However, even if the force is usually well predicted by Aerens’ for-
mula, its inaccuracy can exceed 20% when a new material is used. 200
Furthermore, its application is limited to cones.
100
The SPIF process can be optimized by experimental trials and
errors and/or by a numerical analysis. The numerical approach 0
predicts the forming forces and the final shape according to the 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
tool path. However such information is accurate only if the simu- True strain
lation has been validated by experiments. Previous researches by Tens_0° Tens_45° Tens_90° Hooke 0.2%
Bouffioux et al. (2008, 2010) demonstrated the effect of the material
parameter identification method, the material model itself (Flores Fig. 1. Effect of the anisotropy on tensile tests.
et al., 2007), the Finite Element Method (FEM) strategy such as the
explicit or implicit choice (Henrard, 2009), the finite element type:
solid or shell (Eyckens et al., 2010a; Henrard et al., 2010), the mesh Additional mechanical tests (tensile tests, simple and reverse
density and the contact model (Henrard, 2009). Eyckens (2010b) shear tests) and microstructure verifications showed that the cold
incorporated through-thickness shear (TTS), also known as out-of- rolled and annealed sheets had the same behaviour as the initial
plane shear, into a Marciniak–Kuczynski model and observed that ones.
the through-thickness shear has a significant effect on the increas- A set of mechanical tests was performed to describe the mechan-
ing forming limit curves of the process. Moreover, Emmens and ical behaviour and to fit the parameters of the hardening laws.
van den Boogaard (2009) presented several mechanisms that can
explain the enhanced formability of the SPIF process. 2.1. Classical tests
The numerical inaccuracy of finite element models is one of the
main problems often encountered by scientists. Furthermore, the To investigate the anisotropy of the material, tensile tests were
simulations usually require a very high computation time because performed at 0◦ , 45◦ and 90◦ from the rolling direction. Fig. 1 shows
the tool path is long and can be complex. The combination of differ- a lower yield stress and tensile strength at 45◦ from the rolling
ent strategies described by Bouffioux et al. (2008, 2010), Lequesne direction without any significant change on the strain hardening
et al. (2008) to improve the material and numerical models and coefficient n. The average Young’s modulus E = 70,500 MPa was
applied on this study aimes to be able to accurately simulate the computed from these tests.
process with a short computation time. Tensile tests, performed along the rolling direction (RD, 0◦ ) with
After this introduction showing the interest of this study and 3 different strain rates (0.0075s−1 , 0.075 s−1 and 0.1 s−1 ) did not
the complexity of reaching a good numerical model of the SPIF pro- show any significant change in the mechanical response of the
cess, Section 2 presents all the experimental tests performed on the material properties.
AlMgSc sheets. The constitutive laws used to describe the material The Lankford coefficients: r0 , r45 and r90 with r␣ = εp22 /εp33 are
behaviour, the data fitted by the inverse method and the validation classically computed from tensile tests at 0◦ , 45◦ and 90◦ from the
on a line test are presented in Section 3. Then, the experimental RD. ˛ is the angle between the longitudinal direction of the test
cone tests, which were performed on this material to evaluate its specimen and the rolling direction. εp22 and εp33 are the plastic
forming limits, are described in Section 4. A numerical model of a strain respectively in the transversal and thickness directions of
truncated cone was chosen and simulated. The results were com- the sample. As explained in Section 3.1, the Lankford coefficients
pared with the experimental forces and shape to validate the model are used to determine the parameters of the Hill yield locus.
in Section 5. Finally, Section 6 presents the wall angle sensitivity Due to the Portevin–Le Chatelier effect (serrated stress-strain
which was compared with the experimental results and analytical curves as seen in Fig. 1, a kind of instability classically observed in
predictions. The last section summarizes and discusses the results. AlMg alloys), it was neither possible to accurately define the Poisson
ratio, nor the Lankford coefficient at 45◦ (r45 ) and at 90◦ (r90 ) from
RD. However, the average r0 coefficient could be evaluated at 0.8
for a strain between 6 and 12%.
2. Experimental material characterisation The Poisson ratio was estimated equal to 0.33 as for classical Al
alloys.
As underlined by Aerens et al. (2009), before performing any Additional tests, as used to determine the yield locus shape and
forming process, it is useful to ensure the preservation of the tool- to fit the material data of the isotropic and the kinematic hardening
ing and the machinery and to verify that the platform used is stiff laws, were performed: a monotonic shear test (Figs. 2a and 3, left), a
enough to avoid significant deviation resulting in errors in the large tensile test (Figs. 2b and 3, right), two orthogonal tests (a large
geometry of the achieved parts. tensile test followed by a shear test) at two levels of pre-strain: 4%
A first evaluation of the required force to form the sheet by the and 8% (Figs. 2c and 4, left) and two Bauschinger shear tests (a shear
available SPIF equipment indicated that the as-received sheets from test followed by a reverse shear test) at three different levels of
the manufacturer (3.2 mm in thickness) were too thick and had to pre-strain: Gamma = d/b = 8.8%, 13% and 29% (Figs. 2d and 4, right).
be rolled down to 0.5 mm. A conventional symmetrical cold rolling Each test was performed three times. As the results showed a
was applied and followed by an additional annealing to recover the good reproducibility, an average curve was used and drawn for each
initial material properties. loading state.
Different annealing temperatures between 100 and 475 ◦ C and The tensile, shear and large tensile tests were performed in three
two holding times of 30 and 60 min were investigated to recover directions: at 0◦ (RD), 45◦ and 90◦ (TD) from the rolling direction to
the initial hardness of 110 HV2 and the mechanical properties. It see the effect of the anisotropy. It was observed that the behaviour
turned out that an annealing at 450 ◦ C during 30 min was adequate. was different at 45◦ than in RD or TD.
1686 C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693

a d d
b c d

Fig. 2. Description of the additional classical tests: (a) shear test, (b) large tensile test, (c) orthogonal test and (d) Bauschinger test, definition of b and d values.

Shear tests Large tensile tests


300 500
450
Shear stress (MPa).

250 400

True stress (MPa).


350
200
300
150 250
200
100 150
100
50 50
0 0
0.0 0.2 0.4 0.6 0.8 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
Gamma Strain

Sh_RD Sh_45 Sh_TD TL_RD TL_45 TL_TD

Fig. 3. Experimental shear tests (left) and large tensile tests (right) in 3 directions (0◦ , 45◦ and 90◦ from RD).

Orthogonal tests Bauschinger tests


True stress / Shear Stress (MPa)

450 250
Abs(shear stress) (MPa).

400
350 200
300
250 150
200
.

100
150
100 50
50
0 0
0.00 0.05 0.10 0.15 0.20 0.0 0.2 0.4 0.6 0.8 1.0
Major strain / gamma Sum (Gamma)
OR_4% OR_8% BA_8.8% BA_13% BA_29%

Fig. 4. Experimental orthogonal tests (4 and 8%) (left) and Bauschinger tests (8.8, 13 and 29%) (right).

All these tests were used to describe the yield locus shape and 2.2. Indent and line tests
its evolution. The decrease of the stress in the second part of the
Bauschinger tests indicated that a complex hardening law combin- The line test was performed on a square sheet with a thickness
ing both an isotropic and a kinematic part was required. of 0.5 mm, clamped along its edges (Fig. 5, left).

Y
Z
Y
100 mm

X
tool 1
X 2 3
tool 4
20 mm 60 mm 20 mm
100 mm

Fig. 5. Description of the line test: geometry (left) and tool displacements (right).
C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693 1687

The displacement of the spherical tool, having a radius of 5 mm, Table 1


Material data of the AlMgSc alloy (units: N, mm).
was composed of five steps with an initial position tangent to the
surface of the sheet (Fig. 5, right): a first indent of 3 mm (step 1), a Tests used Yield surface Voce Back stress
line movement at the same depth along the X axis (step 2), then a Coefficients (Hill’48) parameters data (Ziegler)
second indent up to the depth of 6 mm (step 3) followed by a line F = 1.11 K = 275 CA = 930
at the same depth along the X axis in the opposite direction (step Classical and indent G = 1.11  0 = 8.5 GA = 18.1
4) and the unloading. The first step corresponds to what is called tests H = 0.89 n = 439
N=3
hereafter the indent test, while all of the five steps define the line
test. The whole line test was performed three times (with the SPIF
set-up) and the bolts of the frame were tightened using the same However, the in-plane isotropic assumption above-mentioned
torque to ensure the reproducibility of the results. was used.
As explained in Section 3.1, the total tool force during the indent
yield yield
test was combined with all of the classical tests to fit the material TD = RD
data by the inverse identification method. The line test, presenting
The usual additional condition for Hill (1948) was applied (Eq.
deformation mechanism close to the SPIF, was used as validation
(3)):
of the material parameter fitting (see Section 3.2).
H+G =2 (3)
3. Material behaviour modelling
Finally, the Lankford coefficient formulation allowed to charac-
The elastic coefficients and most of the parameters related to terize the out-of-plane anisotropy (Eq. (4)):
the yield locus shape could be defined by classical tensile tests, but H
the inverse method was required to fit the hardening behaviour r0 = (4)
G
and the shear Hill coefficients. This inverse method was coupled
with the Finite Element code: “Lagamine” described by Duchêne where r0 = 0.8, as computed from the tensile tests.
and Habraken (2005). The principle of this method is to choose a The isotropic hardening was described by Voce’s formulation
set of tests, the results of which are sensitive to the material data (Eq. (5)) because this law, which predicts a saturation (Bouffioux
to adjust. These tests are simulated using an initial set of data, cho- et al., 2010; Habraken et al., 2010), was better adapted to describe
sen arbitrarily. Then, the numerical results are compared with the the material behaviour for the large deformation encountered dur-
experimental measurements. The Levenberg Marquardt minimiza- ing the SPIF process.
tion algorithm described by Dennis and Schnabel (1983) is used F = 0 + K(1 − exp(−n• εpl )) (5)
to iteratively adjust the material data until a sufficient accuracy is
reached. where  0 , K and n are the material parameters and is the equiv-εpl
All the classical tests, described in Section 2.1, and the indent test alent plastic strain.
(Fig. 5) were simultaneously used in the inverse method. Indeed, It was noticed that a simple isotropic hardening model (Eq.
Bouffioux et al. (2008, 2010) observed that simple classical tests (5)) was not sufficient to provide an accurate tool force predic-
inducing in-plane stresses were not sufficient to describe the mate- tion (Flores et al., 2007 and Henrard et al., 2010). Therefore an
rial behaviour during the SPIF process. The indent test gave the elasto-plastic law with a mixed isotropic-kinematic hardening was
missing information about the out-of-plane material behaviour used. In this law, the stress tensor  (in Eq. (1)) is replaced by
since it induces through thickness shear. In addition, the same shell ( − X), where X is the back-stress. The evolution of the back-stress
elements were used for these simulations and for the SPIF process. is described by Ziegler’s hardening equation (Eq. (6)):
It was decided to impose isotropic in-plane behaviour because 1
neither the experimental force evolution during a whole contour Ẋ = CA ( − X) ε̇pl − GA • X • ε̇pl (6)
F
of a cone, nor the cone wall thickness comparison in the different
sections (at 0◦ , 45◦ and 90◦ from RD), shows any effect of anisotropy. where CA is the initial kinematic hardening modulus and GA is the
This simplification induced the replacement, in the inverse method, rate at which the kinematic hardening modulus decreases with an
of classical test results in three directions by their average curve. increasing plastic deformation. These models use the Green strains
The elasto plastic constitutive law was to be used on a simplified and the second formulation of Piola–Kirchhoff for the stresses.
numerical model (a 90◦ pie) where the force evolution of a contour As already specified in Section 2.1, the Young’s modulus was
was replaced by an average force when the tool was in the central defined by the tensile tests (E = 70,500 MPa) and the Poisson ratio
third of the pie. was fixed equal to 0.33 as for other aluminium alloys.
The shear parameter N of the Hill’s law, which could not be
3.1. Material model defined accurately by the classical tests was combined to the Voce’s
law data and the back stress parameters to be fitted by the inverse
The elastic range was described by Hooke’s law and the plastic method (Table 1).
part by the Hill 48 law (Eq. (1)) where  ij are the stress tensor  The final data were fitted using the classical tests and the indent
components and where  F is the yield stress. test. All these tests as well as the SPIF process were simulated with
1 quadrilateral shell elements with four nodes.
FHILL () = H(xx − yy )2 + G(xx − zz )2
2 As shown in Figs. 6–8, this set of data was able to accurately
  predict all the experimental curves.
+F(yy − zz )2 + 2N xy
2 2
+ xz 2
+ yz − F2 = 0 (1)

The parameters F, G, H of Hill’s law (Table 1) could be identified 3.2. Validation of the material model parameters
from the relation between the yield stress limit in RD and TD (Eq.
(2)): The line test, described in Fig. 5, was used to verify the accu-
 racy of the material data on a complex procedure similar to the
yield 2 yield incremental forming process, performed with the SPIF set-up and
TD =  (2)
H + F RD inducing bending, and an out-of-plane stress field.
1688 C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693

Simple tensile and large tensile tests shear test


500 300
450
250
True stress (MPa).

400

Shear stress (MPa).


350 200
300
250 150
200
150 100
100
50
50
0 0
0.00 0.02 0.04 0.06 0.08 0.10 0.00 0.10 0.20 0.30
strain Gamma
SimpleTens exp SimpleTens Num.
LargeTens exp LargeTens Num. Shear exp Shear Num.

Fig. 6. Comparison between experimental and numerical tensile and large tensile tests (left) and simple shear tests (right).

Orthogonal tests Bauschinger tests


True stress / Shear stress (MPa).

450 250
400 200
350 150

Shear stress (MPa).


300 100
250 50
200 0
150 -0.2 -0.1 -50 0.0 0.1 0.2 0.3
100 -100
50 -150
0 -200
0.00 0.05 0.10 0.15 0.20 0.25 -250
Major strain / gamma Gamma
Ortho_4% exp Ortho_4% Num. Ortho_8% exp Ortho_8% Num. Bau_8.8 exp Bau_8.8 Num. Bau_29 exp Bau_29 Num.

Fig. 7. Comparison between experimental and numerical orthogonal tests (left) and Bauschinger tests (right).

Indent test
400
350
300
Force (N).

250
200
150
100
50
0
0 0.1 0.2 0.3 0.4 0.5
Time (sec)
Indent test exp Indent test Num.

Fig. 8. Comparison between experimental and numerical tool force during the indent test.

In the FEM simulation, the nodes along the edges were fixed. 4. Experimental incremental forming applicability on a
The tool force was computed by the static implicit strategy. The cone
Coulomb’s friction coefficient was unknown. The same coefficient
of 0.05 as Henrard (2009) for the alloy 3013 was applied between As illustrated in Fig. 10, a firmly clamped sheet was formed by
the tool and the sheet. It was observed that the impact of this coef- a contouring operation of the forming tool. The tool moved along
ficient on the total tool force is small (the force computed with a predetermined contour in the horizontal plane, after which the
a friction of 15% is 1% higher than the force without friction). The tool incrementally descended and started a contour in the next
mesh was adjusted to limit the number of elements (576 elements) horizontal plane, building up the workpiece layer by layer.
while keeping accuracy (Fig. 9, left), taking into account the sym- The two platforms used in the framework of this project were
metry. The same shell elements as in the inverse method were a 6-axis robot and a 3-axis rigid milling machine (Fig. 11), both
used. equipped with a dedicated force measuring device.
Fig. 9, right, shows that the model gives a good evaluation of the The forming speed (approx. 2000 mm/min), in combination with
evolution of the experimental tool total force. the long toolpath, resulted in a slow forming process.
C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693 1689

Y
Line test

1400
1200
1000

Force (N).
X
100 mm

800
600
400
200
tool
0
0 5 10 15 20 25
Time (sec)

1 00 mm Line test exp Line test Num.

Fig. 9. Line test: geometry and meshing (left) and comparison between experiment and numerical tool force evolution (right).

Fig. 10. Description of the SPIF process.

Fig. 11. Platforms description.

Thinning of the workpiece is the dominant failure mode in SPIF Determining the geometrical forming limits defined by the max-
and is related to the workpiece drawing angle ␣ (Fig. 12). For a given imum drawing angle ˛ (Fig. 12) in Single Point Incremental Forming
material and initial thickness, the maximum drawing angle repre- was done by means of well-established cone tests. Forming limit
sents the limits of the conventional incremental forming process. tests typically started with the formation of a cup with a 10◦ wall
angle. Then, successive cups with an increasing wall angle were
made. The forming limit is the angle at which the sheet fractured.
The maximum drawing angle is mainly influenced by the initial
thickness of the blank sheet. To a lesser degree, the diameter of the
spherically tipped tool and the scallop height affect the drawing
angle in a negative way. The scallop height is specified as the theo-
retic height of the ripple between two passes of the forming tool as
calculated for a milling process. The maximum value of ˛, for the
studied AlMgSc alloy, was found equal to 46◦ with an accuracy of
Fig. 12. Workpiece geometry. 1◦ (Table 2).
1690 C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693

Table 2 Two meshes were tested: a very fine mesh (2692 elements)
Cone parameters and maximum forming angle.
without the remeshing method and the coarse mesh of Fig. 14 (248
Material Thickness Ø tool Scallop height Step down ˛ elements) combined with the remeshing method. The level of the
AlMgSc 0.5 mm 10 mm 0.005 mm 0.322 mm 46◦ force was almost the same for both simulations, but the compu-
tation time was divided by a factor of 49 for the second case. The
Comparing different materials reveals a fairly low forming limit Coulomb friction coefficient , which was not accurately known,
for AlMgSc. This is a strong limiting factor for the part geometries was taken equal to 0.10.
which could be made in this material. Duflou et al. (2007) showed
that performing incremental forming at elevated temperatures can 5.1. Tool force comparison
significantly improve forming limits. However, this property was
not verified on the AlmgSc. The radial FR , tangential FT and axial FZ force components,
respectively in R, T and Z directions (Fig. 14, left) and the norm of
5. Numerical models of a cone processed by SPIF the force Ftot were compared with the experimental ones. To sim-
plify the comparison, the force evolution during each contour was
A truncated cone test was chosen to validate the material model. replaced by the average of the force. The numerical test values were
A circular AlMgSc sheet with a thickness of 0.5 mm was clamped. 30 computed when the tool was in the central third of each contour
contours were performed with a tool radius of 5 mm, a depth incre- to avoid small inaccuracies due to the boundary conditions.
ment Z of 0.5 mm between successive contours, corresponding to Fig. 15 shows that similar results were obtained with a fine mesh
a scallop height of 0.015 mm, and a wall angle of 40◦ (Fig. 13). without remeshing and a coarse mesh combined with the remesh-
Strategies had to be found to reduce the computation time. As ing method. Similar small differences could be observed for all the
introduced in Section 3, the use of shell elements instead of e.g. force components.
3D solid elements aimed to considerably reduce the computation
time. Indeed, solid elements would require minimum three layers 550
of elements to model bending. It would lead to a significantly larger 500
number of degrees of freedom. 450
400
Total force (N).
The mesh used to simulate the process is presented in Fig. 14.
Since the tool always moved in the same tangential direction, it 350
300
was possible to reduce the computation time by modelling only a
250
quarter of sheet. Rotational boundary conditions (Henrard, 2009;
200
Henrard et al., 2010) were imposed by a link between the displace-
150
ments of the edges, as schematically presented in Fig. 14, left: the
100
rotation, the tangential and the radial displacements were forced 50
to have the same values on both the horizontal and vertical bound- 0
aries of the mesh. This link is related to the six degrees of freedom (3 0 5 10 15 20 25 30
translations and 3 rotations) of each node of the edges along the X Contour
and Y axes. The remeshing method (Lequesne et al., 2008) was used
Experimental Fine mesh Coarse mesh + remesh
to refine the mesh by division of the elements close to the tool into
nine elements. The refinement was automatically removed when
Fig. 15. Comparison of the experimental total tool force with the simulation results
the tool went further, except when the elements distortion was for two cases: the fine mesh without remeshing and the coarse mesh with remesh-
high (Fig. 14, centre and right). ing.

Y
Z

X
40°
tool
a b c b a
92 mm
∅: 90 - 92 mm

Fig. 13. Description of the cone test (a = unsupported area = 1 mm, b = 17.88 mm and c = 54.25 mm).

Fig. 14. Cone test: Initial mesh, axes, radial (R) and tangential (T) directions, boundary conditions (left) and mesh evolution with the remeshing method (centre and right).
C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693 1691

FZ_S FZ Ftot
550 550
500 500
450 450
400 400

FR & Ftot (N)


350
FT & FZ (N)

350
300 300
250 250
FR
200
FT 200
150 150
100 100
50 50
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Contour Contour

Experimental Numerical Analytic Experimental Numerical

Fig. 16. Comparison of the numerical (coarse mesh + remeshing), analytical (FZ S ) and experimental total, axial, tangential and radial forces.

Fig. 16, left and right show that the numerical model with a Experimental and numerical shapes
coarse mesh combined with the remeshing method can predict all 0
the force components: FR , FT , FZ and Ftot with accuracy. In addition, -50 -30 -10 10 30 50
it would be possible to improve the level of the FT component by -4
adapting the friction coefficient. Henrard (2009) observed that fric-
-8
tion has almost no impact on FR and FZ while FT is linearly affected by
the friction coefficient. The ratio FT /FZ reached by two simulations -12
with  = 0.05 and  = 0.10 was respectively: 0.157 and 0.212. By
extrapolation, the value:  = 0.14 should lead to the experimentally -16
measured ratio FT /FZ of 0.256.
-20
The axial force when a steady state is reached: FZ S could also be
predicted by an analytic generalized formula (Aerens et al., 2009)
based on the tensile strength of the material Rm , the sheet thickness Numerical shape Robot mid-thickness, X=0 Robot mid-thickness, Y=0
t, the tool diameter dt , the scallop height h and the wall angle ˛
(Eq. (7)): Fig. 17. Comparison of the mid-thickness shapes.

0.41 •
FZ S = 0.0716• Rm • t 1.57• dt h0.09• ˛ cos ˛ (7)

where FZ S is expressed in N, Rm in N/mm2 ,


t, dt and h in mm and Only a small difference was observed between the two sections:
˛ in degrees. X = 0 and Y = 0 and between the cones formed by the robot and the
As before-mentioned, the scallop height h is the theoretic milling machine used for the experiments. The accuracy measure-
height of the ripple between two passes of the forming tool. It is ments were obtained on the tool side; it is the inner shape. The local
linked to the depth increment Z, the tool diameter and the wall thickness was found by comparing the scan data from the outside
angle by the relation (Eq. (8)): of the part with respect to the inside. The mid-thickness layer was
  computed (Fig. 17) point by point, taking into account the local wall
Z = 2 sin ˛ h(dt − h) ≈ 2 sin ˛ h• dt (8) angle and the thickness. This correction had a small impact on the
shape due to the low sheet thickness.
The tensile strength determined by the tensile tests was
The numerical test shape was extracted out of the section: X = Y
352 MPa, the scallop height was 0.0151 mm from Eq. (8) and thus
to avoid small imprecision due to the boundary conditions. It was
the force prediction FZ S was equal to 458.5 N. This level, repre-
the mid-thickness layer obtained by the nodal coordinates. Fig. 17
sented in Fig. 16, left, gave a good prediction for this alloy and
shows that the final shape was well predicted by the model. The
geometry.
higher discrepancy was observed at the centre of the cone with an
The radial force FR could be linked to the axial force FZ S (Aerens
error of 0.9 mm.
et al., 2009) by the following relation (Eq. (9)):
 −c
˛ + ˇ − 17.2• dt /10
FR = FZ • tan (9) 6. Wall angle sensitivity
s
2
where ˇ = arccos (1 − 2·h/dt ) (in degrees) and c = 2.54 for alu- Earlier research on forces in SPIF revealed the important and
minium alloys. complex non-linear influence of the draw angle on the forming
This equation provided a first approximation of FR = 111.2 N forces. Because of this irregular behaviour a more in-depth study
which was too low in the current case. However, as mentioned by was performed towards this influence. A set of cones similar to
Aerens et al. (2009), a correction term should be used to improve the cones used for the forming limit analysis was experimentally
this equation, but its determination would require additional formed to see the angle sensitivity. Once again, a circular AlMgSc
experiments that were not performed on the AlMgSc material. metal sheet with a thickness of 0.5 mm was clamped on a back-
ing plate with a diameter of 90 mm. The contours were performed
5.2. Shape comparison with a tool diameter of 10 mm and the same scallop height of
0.005 mm for all the cones. While forming the cones with draw
The experimental shapes were measured by means of laser line angles between 10◦ and 46◦ and the step down indicated in Table 3,
scanning at the end of the process when the tool was removed. forces were measured when a steady state was reached.
1692 C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693

Fig. 18. Effect of the wall angle on the force components and comparison between experimental and numerically simulated forces (left) and comparison between experi-
mental, analytical and simulated axial (FZ ) and radial (FR ) forces (right).

Table 3 tion time. It was possible to simulate the 30 contours of a cone test
Step down with respect to wall angles for a tool diameter of 10 mm and a scallop
with a simple personal computer within 5 h and last but not least,
height of 0.005 mm.
with a good precision.
Wall angle (◦ ) 10 20 30 40 45 46 The effect of the wall angle on the forming forces was also stud-
Step down (mm) 0.078 0.153 0.224 0.287 0.316 0.322 ied experimentally, analytically and by numerical simulation. Once
again, the numerical simulation model was validated.
In future work, the results could be further improved by using
Three cones with the wall angles of 20, 30 and 40◦ were also more complex elements, such as solid-shell type elements (Alves
simulated using a numerical model with a coarse mesh and the de Sousa et al., 2007) which would take into account the through-
remeshing method, the same tool path as in the experiment and a thickness shear. However, more complex elements will affect the
friction coefficient of 0.14. 30 contours were performed as it was computation time.
observed that the forces reached a steady state from the 12th con- As conclusions, it is clear that this study gives a new contribution
tour. Then, the average values of the tool forces were computed in AlMgSc characterisation and in incremental forming application
when the tool was in the central third of each contour. As a result, and simulations.
these force components with respect to the wall angle were com-
pared with the experimental forces (Fig. 18, left and right).
It is clearly visible in Fig. 18, right, that the axial force: FZ was Acknowledgments
the dominating force component in the SPIF process. Also, the non-
linearity between the draw angle and the force components is The authors would like to thank the Belgian Federal Science Pol-
visible. The numerical model was able to give a good approximation icy Office (IAP: project P6/24 and ALECASPIF: Pat2 project P2/00/01)
of the force components. for its financial support and Aleris for the material supply.
The radial FR and axial FZ tool forces were also computed by A.M. Habraken and L. Duchêne would like to thank the Fund for
Eqs. (7) and (9) (Aerens et al., 2009). These equations which were Scientific Research (FNRS, Belgium) for its support.
established for five different materials gave also a good estimation J. Duflou and H. Vanhove recognize the support by FWO-
of these two components (Fig. 18, right) even if the AlMgSc alloy vlaanderen for the project.
was not considered in that study.
References
7. Conclusions
Aerens, R., Eyckens, P., Van Bael, A., Duflou, J.R., 2009. Force prediction for single
The aims of this research were to improve the knowledge of the point incremental forming deduced from experimental and FEM observa-
tions. International Journal of Advanced Manufacturing Technology, 969–982,
AlMgSc sheets, to study the applicability of the SPIF process and to doi:10.1007/s00170-009-2160-2.
simulate it on this material. Allwood, J.M., Shouler, D.R., 2009. Generalised forming limit diagrams showing
Classical tests gave information about in-plane material increased forming limits with non-planar stress states. International Journal of
Plasticity 25 (7), 1207–1230, doi:10.1016/j.ijplas.2008.11.001.
behaviour. An indent test performed with the SPIF setup, induc- Alves de Sousa, R.J., Yoon, J.W., Cardoso, R.P.R., Fontes Valente, R.A., Grácio, J.J., 2007.
ing heterogeneous stress and strain fields similar to those present On the use of a reduced enhanced solid-shell (RESS) element for sheet forming
in the SPIF process, provided information for both in-plane and simulations. International Journal of Plasticity 23 (3), 490–515.
Bouffioux, C., Henrard, C., Eyckens, P., Aerens, R., Van Bael, A., Sol, H., Duflou, J.R.,
out-of-plane strain states. Habraken, A.M., 2008. Comparison of the tests chosen for material parameter
A mixed isotropic-kinematic hardening, described by both a identification to predict single point incremental forming forces. In: Proceedings
saturating hardening law such as Voce’s law and the kinematic of IDDRG 2008 Conference , Olofström, Sweden, pp. 133–144.
Bouffioux, C., Pouteau, P., Duchêne, L., Vanhove, H., Duflou, J.R., Habraken, A.M.,
Ziegler’s law, accurately models the material behaviour as observed
2010. Material data identification to model the single point incremental form-
during all the simulated tests. ing process. International Journal of Material Forming 3 (Suppl. 1), 979–982,
To quantify the formability of the AlMgSc alloy in the SPIF pro- doi:10.1007/s12289-010-0933-7.
Dennis, J.E., Schnabel, R.B., 1983. Numerical Methods for Unconstrained Optimiza-
cess, experimental investigations on cup tests showed that for a
tion and Nonlinear Equations. Prentice, Hall, Englewood Cli s, NJ, USA.
spherically tipped tool of Ø10 mm and a scallop height of 0.005 mm, Duchêne, L., Habraken, A.M., 2005. Analysis of the sensitivity of FEM predictions
a maximum wall angle of 46◦ was found. to numerical parameters in deep drawing simulations. European Journal of
To simulate the SPIF process, the use of shell elements, a 90◦ pie Mechanics - A/Solids 24, 614–629.
Duflou, J.R., Callebaut, B., Verbert, J., Debaerdemaeker, H., 2007. Laser assisted
coarse mesh combined with the remeshing method and rotational incremental forming: formability and accuracy improvement. CIRP
boundary conditions allowed to considerably reduce the computa- Annals—Manufacturing Technology 56 (1), 273–276.
C. Bouffioux et al. / Journal of Materials Processing Technology 211 (2011) 1684–1693 1693

Emmens, W.C., van den Boogaard, A.H., 2009. An overview of stabilizing deforma- P., 2010. An Aluminium alloy improved by ECAP and formed by SPIF process,
tion mechanisms in incremental sheet forming. Journal of Materials Processing Alecaspif, Final Report, Belgian Science Policy (Programme to stimulate knowl-
Technology 209, 3688–3695. edge transfer in areas of strategic importance–TAP2), 112 pages, Available from:
Eyckens, P., Belkassem, B., Henrard, C., Gu, J., Sol, H., Habraken, A.M., Duflou, J.R., Van http://www.belspo.be/belspo/home/publ/rappP2 en.stm.
Bael, A., Van Houtte, P., 2010a. Strain evolution in the single point incremen- Henrard, C., 2009. Numerical simulations of the Single Point Incremental
tal forming process: digital image correlation measurement and finite element Forming Process. PhD thesis. Available from: http://bictel.ulg.ac.be/ETD-
prediction. International Journal of Material Forming, doi:10.1007/s12289-010- db/collection/available/ULgetd-04182009-184505/.
0995-6. Henrard, C., Bouffioux, C., Eyckens, P., Sol, H., Duflou, J.R., Van Houtte, P.,
Eyckens, P., 2010. Formability in Incremental Sheet Forming: Generalization of the Van Bael, A., Duchêne, L., Habraken, A.M., 2010. Forming forces in single
Marciniak–Kuczynski model. PhD thesis. ISBN 978-94-6018-210-5. point incremental forming. Prediction by finite element simulations, vali-
Filice, L., Fratini, L., Micari, F., 2002. Analysis of material formability in incremental dation and sensitivity. Computational Mechanics, doi:10.1007/s00466-010-
forming. CIRP annals—Manufacturing Technology 51 (1), 199–202. 0563-4.
Flores, P., Duchêne, L., Bouffioux, C., Lelotte, T., Henrard, C., Pernin, N., Van Bael, A., Lequesne, C., Henrard, C., Bouffioux, C., Duflou, J.R., Habraken, A.M., 2008. Adaptive
He, S., Duflou, J., Habraken, A.M., 2007. Model identification and FE simulations: remeshing for incremental forming simulation. In: Proceedings of Numisheet
effect of different yield loci and hardening laws in sheet forming. International 2008 Conference , Interlaken, Switzerland, pp. 399–403.
Journal of Plasticity 23 (3), 420–449. Micari, F., Ambrogio, G., Filice, L., 2007. Shape and dimensional accuracy in single
Habraken, A.M., Bouffioux, C., Lequesne, C., Rosochowski, A., Rosochowska, M., Ver- point incremental forming: state of the art and future trends. Journal of Materials
linden, B., Vidal, V., Ailinca, G., Duflou, J.R., Vanhove, H., Schmitz, A., Pouteau, Processing Technology 191, 305–390.

You might also like