FULLTEXT01

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 58

2006:068 CIV

MASTER'S THESIS

Finite Element Simulation


of Punching

Magnus Söderberg

Luleå University of Technology

MSc Programmes in Engineering


Mechanical Engineering
Department of Applied Physics and Mechanical Engineering
Division of Computer Aided Design

2006:068 CIV - ISSN: 1402-1617 - ISRN: LTU-EX--06/068--SE


Abstract
This work is a study of a punching process using the ABAQUS/Explicit FE-code. A
damage model based on equivalent plastic strain has been used to describe crack initiation
and propagation in the blank.

An axisymmetric model created in ABAQUS/CAE was used throughout the study. The
model consists of a punch, blank holder, die and blank. In this work the punch, blank
holder and die are modelled as rigid bodies while the blank is considered elastic-plastic.

The materials used in simulations are DC04, Docol 350 YP, Docol 800DP and DC1400M.
This selection of steel qualities has a wide span in mechanical properties and fields of
application.

Material characterisation for simulation purposes has been done by conducting uniaxial
tensile tests. The data from these tests have been extrapolated beyond the necking strain by
double Voce extrapolation.

Punching experiments have been performed in an excenter driven press equipped with
position and force measurement devices. The experiments result in load curves for the
different steel qualities and gives information about the characteristic zones of the sheet
edges.

The results from simulations have been evaluated by comparisons with experiments
concerning load curves, characteristic zones of the sheet edge, work done during the
operation, maximum forces and stroke to failure.

A good compliance with experiments was found concerning maximum forces,


characteristic zones and work done during the operation.
Table of Contents
1. Introduction....................................................................................................................4
2. Influence of simulation parameters................................................................................5
2.1 FE-Model ...............................................................................................................5
2.1.1 Geometry........................................................................................................5
2.1.2 Element type ..................................................................................................6
2.1.3 Material data and model.................................................................................6
2.1.4 Contact and boundary conditions ..................................................................7
2.1.5 Failure criterion..............................................................................................8
2.2 Fe-Simulations .......................................................................................................9
2.2.1 Mesh density ..................................................................................................9
2.2.2 Adaptive mesh .............................................................................................14
2.2.3 Failure strain value.......................................................................................16
2.3 Comparison with results from DEFORM ............................................................21
2.3.1 Failure criterion in DEFORM......................................................................21
2.3.2 Load curves..................................................................................................21
2.3.3 Simulated edge geometry.............................................................................22
2.4 Elastic tools..........................................................................................................23
2.5 Comments ............................................................................................................27
3. Material properties .......................................................................................................28
3.1 Materials ..............................................................................................................28
3.2 Tensile tests..........................................................................................................28
3.2.1 Mechanical properties..................................................................................28
3.2.2 Extrapolation of test data .............................................................................29
4. Experiments .................................................................................................................30
4.1 Experimental setup ..............................................................................................30
4.2 Data from experiments.........................................................................................31
5. Simulation model.........................................................................................................32
5.1 Geometry and mesh density.................................................................................32
5.2 Material model .....................................................................................................33
5.2.1 Strain rate dependency.................................................................................34
5.3 Failure criterion....................................................................................................34
5.4 Contact and boundary conditions ........................................................................35
5.4.1 Contact conditions .......................................................................................35
5.4.2 Boundary conditions ....................................................................................37
6. Comparison with experiments .....................................................................................38
6.1 Method of evaluating simulations........................................................................38
6.1.1 Evaluation of simulated characteristic edge zones ......................................38
6.1.2 Evaluation of work.......................................................................................39
6.1.3 Evaluation of stroke to failure......................................................................39
6.1.4 Evaluation of maximum punch force...........................................................40
6.2 Load curves and edge profiles .............................................................................40
6.2.1 DC04............................................................................................................40
6.2.2 Docol 350 YP...............................................................................................41
6.2.3 Docol 800 DP...............................................................................................42
6.2.4 Docol 1400M ...............................................................................................43
6.3 Characteristic zones .............................................................................................44
6.3.1 Rollover .......................................................................................................44
6.3.2 Shear zone....................................................................................................45
6.3.3 Fracture zone................................................................................................46
6.4 Maximum punch force.........................................................................................47
6.5 Work ....................................................................................................................48
6.6 Stroke to failure....................................................................................................49
7. Discussion....................................................................................................................50
8. Conclusions..................................................................................................................51
9. Future work..................................................................................................................52
10. Acknowledgements..................................................................................................53
11. References................................................................................................................54
Appendix 1 Mechanical properties ......................................................................................55
Appendix 2 Data from punching experiments .....................................................................57
1. Introduction
Punching is a sheet metal forming procedure were material shearing mechanisms are used
to produce the desired geometry. The development today is focused against using
advanced high strength steel qualities to make lighter products with sustained or increased
product durability. These steel qualities puts higher demands on the tooling used to cut or
form the sheet steel, which makes the choice of process parameters even more important
than with ordinary mild steels, because of the narrow process window when using high
strength steels. Identification of the process parameters is today mostly done by large series
of experiments, which is costly and time-consuming [1]. In earlier work performed by D.
Thunvik [2] and others [3], [4], the DEFORM FE-code has been used to simulate the
punching process with results close to experimental. Simulations with the
ABAQUS/Standard FE-code and damage implemented by means of a user subroutine have
also shown results in compliance with experiments [5], [6]. This shows the possibility of
using FE-simulations to reduce the number of experiments that has to be conducted and
give a increased understanding concerning the influence of process parameters.

This work utilises the ABAQUS/Explicit FE-code and a failure criterion that is readily
implemented in the code. The objective has been to resolve the capacity of this code
concerning simulations of punching and establishment of a modelling procedure for this
type of problems. Simulations will be compared to experimental data concerning geometry
of the edges, load curves, work done by the punch during the process, maximum forces
involved and stroke to failure of the blank.

The first section is devoted to finding a suitable modelling approach in ABAQUS


concerning mesh densities, failure criterion and features in the software such as adaptive
mesh. The results from this section will be compared to simulations made in DEFORM [2].

4
2. Influence of simulation parameters
In this section a FE model is created in ABAQUS to resolve the influence of the variations
in the mesh density, value of the failure criterion and the adaptive mesh function that is
implemented in the software. A single case of punching in Docol 800 DP will be simulated
using the dimensions and material model as earlier in the DEFORM [2] simulations, to
make a comparison between the results from the two different software packages possible.
Simulations with elastic tools have been made to study if it is possible to evaluate the stress
state in the tools during punching and the elastic response in the tools using this modelling
approach.

2.1 FE-Model
All FE-models evaluated in this study have been created using the ABAQUS/CAE pre-
processor. The solution is obtained with the ABAQUS/Explict solver, which uses a direct
integration scheme on the structural dynamic equations. This approach is suitable for
highly dynamic problems with non-linear behaviour, which is the case when simulating
fracture of metals.

2.1.1 Geometry
The tools and blank have been modelled using an axisymmetrical model that can be seen in
Figure 2.1, values of the dimensions can be found in Table 2.1.

Figure 2.1: The axisymmetric FE-model in ABAQUS.

Table 2.1: Dimensions corresponding to the model in Figure 2.1


r1 [mm] r2 [mm] r3 [mm] r4 [mm] c1 [mm] c2 [mm] t [mm]
2.44 0.01 0 0.01 0.5 0.06 1

5
2.1.2 Element type
The elements in the blank consist of four node bilinear axisymmetric quadrilateral
elements with reduced integration (CAX4R) and three node axisymmetrical triangles
(CAX3). The triangles are used to coarsen the mesh for increased computational
efficiency, see Figure 2.4. Both elements belongs to the family of solid elements and are of
the first order, which means that the strain is computed as an average over the element
volume instead of the first order gauss point. The feature of reduced integration used in the
CAX4R element causes the integration order to be lower than full integration, in this case
only one integration point in the centre of the element is used. By using reduced integration
the numbers of constraints which are introduced by the elements are decreased, this
prevents “locking” in the elements causing a stiff response. The drawback of this technique
is that for certain modes of deformation no energy is registered in the element integration
point. These modes are usually referred to as “hourglass modes”. This problem is
addressed in ABAQUS using a “hourglass control” algorithm [9].

2.1.3 Material data and model


The blank material consists of Docol 800 DP in all simulations in this section. Yield data
were taken from [2] in an effort to make the simulations as comparable as possible, see
Figure 2.2. The yield data is extrapolated using the double Voce method, which is a way of
producing reliable data at high strains. This procedure is explained in more detail in section
3.2.2, which covers the topic of extrapolation of yield curves.

Double Voce extrapolated yield data

1400
1200
True stress [MPa]

1000
800 Yield stress Docol
600 800 DP
400
200
0
0 0.2 0.4 0.6 0.8 1
True plastic strain

Figure 2.2: Yield data for Docol 800DP extrapolated with the double Voce method.

The material model used is the isotropic von Mises hardening model. In this model the
material is assumed to have similar properties in all directions. As no reversed loading
occur during the simulations the hardening was described as isotropic i.e. the Bauchinger
effect is not modelled. The yield criterion for this model can be expressed by means of the
principal stresses as:

1
σy = [(σ 1 − σ 2 ) 2 + (σ 1 − σ 2 ) 2 + (σ 2 − σ 3 ) 2 ] (1)
2

As can be seen from Equation (1) the criterion is independent from the hydrostatic stress
condition i.e. only the deviatoric part of the stress tensor influences the criterion. This can

6
be interpreted by saying that yielding occurs at a certain level of deviatoric strain energy
[10].

2.1.4 Contact and boundary conditions


Contact between the tools and the blank is enforced by a kinematic contact condition,
using pure master-slave surface pairs established in the first step of the solution. The
master and slave designation must be chosen so the rigid tool forms the master surface and
the surfaces defined on the blank act as slave [11].

The formulation of this contact condition is of predictor/corrector type were the model is
first advanced to a kinematic state without consideration to the contact condition. The slave
nodes witch penetrates the master surface is then determined and the force to move the
slave nodes on to the master surface is calculated. These forces are calculated based on the
depth of penetration, nodal masses and time increment. The acceleration correction for the
master and slave surfaces is then calculated based on the forces needed to oppose
penetration and the inertia of the contacting bodies [11]. The surfaces which form the
contact pairs in this model are summarised in Figure 2.3 and Table 2.2.

Figure 2.3: The surfaces used in the contact pairs. 1) is the surface of the punch, 2) is the top nodes
of the blank, 3) is an internal nodal based surface, 4) is the surface of the blank holder, 5) is the
bottom nodes of the blank and 6) is the surface of the die

7
Table 2.2: A summarisation of the contact conditions numbered as in Figure 2.3
Master surface Slave surface/surfaces
1 2, 3
4 2
6 3, 5

Friction between the surfaces is implemented with a Coulomb model defined as:

τ f = µσ n (2)

Where τf is the friction shear stress, µ is the friction coefficient and σn is the normal
pressure. A friction coefficient of 0.1 has been used in all simulations.

The punch displacement is applied as a prescribed velocity of 40 mm/s and the holder
force is set to be 1350 N which is about 15 percent of the maximum punch force. The
nodes at the symmetry line are locked in the radial direction by a displacement boundary
condition. When an axisymmetrical definition is adapted, no material flow in the angular
direction is presumed.

2.1.5 Failure criterion


To estimate the start and propagation of fracture, a local fracture criterion is used in the
simulations. The criterion used in this set of simulations is the shear failure model in
ABAQUS. This model is based on a value of equivalent plastic strain at element
integration points and failure is assumed to occur when the damage parameter exceeds 1.
The damage parameter, ω, is defined as:

ε 0pl + ¦ ∆ε pl
ω= (3)
ε fpl

Where ε0pl is the initial value of the equivalent plastic strain, ∆εpl is an increment of plastic
strain and εfpl is the strain at failure [11].

8
2.2 Fe-Simulations
2.2.1 Mesh density
Because of the high strain gradients in the cut region, a sufficiently dense mesh must be
applied in this area. The mesh density will also affect the geometry of the cut edges as a
result of fracture being simulated with element deletion.

To resolve how the mesh density will affect the results four models with different mesh
densities have been analysed. The number of elements was 656, 1488, 3901 and 7693 and
the plastic failure strain was set to 1.5. Mesh densities were chosen by starting out with 16
through thickness (t in Figure 2.1) elements in the sheared zone and then dividing them by
2 until 128 through thickness elements were reached. The maximum number of elements
through thickness will hereafter be used when referring to a mesh density i.e. 16, 32, 64
and 128. For additional data about the different mesh sizes see Table 2.3. The different
mesh sizes are shown in Figure 2.4.

Table 2.3: Data about the different mesh densities


Maximum number of through Number of elements Smallest characteristic length
thickness (t) elements of elements [mm].
16 656 0.0625
32 1488 0.03125
64 3901 0.01563
128 7693 0.007813

9
Figure 2.4: a) 656 elements, 16 through thickness elements, b) 1488 elements, 32 through thickness
elements, c) 3901 elements, 64 through thickness elements and d) 7693 elements with 128 through
thickness elements. The number of through thickness elements stated is the maximum number in the
refined zone.

10
It can be seen from the load curves plotted in Figure 2.5 that the mesh density influences
the rupture behaviour and stroke to failure to a greater extent than the maximum force.
This is believed to be an effect of the more localised plastic deformation when the mesh
density is increased, which results in higher strains at an earlier stage of the simulation thus
triggering the failure criterion with following element deletion, see Figure 2.6.

Mesh density Docol 800 DP

10000

8000
64 elements
Force [N]

6000 32 elements
4000 16 elements
128 elements
2000

0
0 0.2 0.4 0.6 0.8
Stroke [mm]

Figure 2.5: Force versus stroke curves using the above mentioned mesh densities. The plastic
failure strain was set to 1.5 for all simulations.

Figure 2.6: Field plots of the equivalent plastic strain with a) 64 elements through thickness and b)
128 elements through thickness.

11
When coarser meshes are used the final rupture is delayed because of the cracks in the
blank that originates from the punch and die miss each other, as shown in Figure 2.7. This
behaviour is not observed when using 128 elements through thickness because of the
possibility for cracks to initiate closer to the edges of the punch and die. Another effect of
the increased mesh is that it seems to facilitate for the cracks to change columns in the
mesh as they propagate, see Figure 2.8 for the simulated edge geometry using different
mesh densities. The ability for the cracks to change column is of great importance if the
crack tips shall be able to meet each other and produce rupture behaviour similar to the one
observed in experiments.

Figure 2.7: Crack tips missing each other during fracture of the blank. The mesh density is 64
elements through thickness.

12
Figure 2.8: The simulated sheet edges with a) 16 through elements, b) 32 through thickness
elements, c) 64 elements and d) 128 elements. The number of elements refers to the maximum
number of elements through thickness as mentioned above A plastic failure stain of 1.5 was used in
all simulations.

13
2.2.2 Adaptive mesh

Adaptive meshing in ABAQUS is carried out in two steps: creating a new mesh and
remapping of the solution variables from the old mesh to the new mesh by an advection
sweep. The adaptive meshing task is triggered at user specified intervals with a default
value of 10 increments. During an adaptive mesh increment, the new mesh is created by
one or many mesh sweeps, which moves nodes to reduce element distortion. The number
of mesh sweeps can be specified by user input and has a default value of one. This value
has to be increased if the deformation rate is high because the numerical stability of the
advection sweep is maintained only if the difference between the new and old mesh is
small [11]. For the simulations in this section, the default values will be used.

In the section above the crack tips miss each other when using coarse meshes, which
causes a section of elements between the two cracks to experience large distortion before
final rupture. This results in an increased stiffness causing a longer stroke to fracture. To
investigate if the adaptive mesh function in ABAQUS is able to improve the solution, the
four mesh densities used in the previous section were simulated using this function on the
elements in the zone between the tool edges. The results of the simulations can be seen in
Figure 2.9. The location of crack initiation is improved with adaptivity i.e. the cracks is
starting nearer to the tool edges which is an effect of the mesh ability to follow the tool
geometry closer. For 16 and 32 elements the cracks propagates in the same element
column during the whole rupture process which gives a realistic load drop but fails to
represent the edge geometry by producing a very short sheared zone, see Figure 2.10. With
64 elements the simulations produces results that are improved over the ones without
adaptivity but still suffers from the fact that a element column is “trapped” between the
two crack tips which results in a deviation from the simulation with 128 elements at a
stroke of 0.16 mm. This leads to the conclusion that between 64 and 128 through thickness
elements are needed for the simulations. For capturing the edge geometry, 128 elements
will probably be required but 64 elements can be used to study the influence of parameter
values on the force stroke curve to reduce the cpu time required to perform the simulations.

Mesh density Docol 800 DP

10000

8000
16 elements
Force [N]

6000 32 elements
4000 64 elements
128 elements
2000

0
0.00 0.10 0.20 0.30 0.40
Stroke [mm]

Figure 2.9: Force versus stroke curves for different mesh densities using adaptive meshing. The
plastic failure strain was set to 1.5 for all simulations.

14
Figure 2.10: The simulated sheet edges with a) 16 elements, b) 32 elements, c) 64 elements and d)
128 elements in the thickness direction, using adaptive mesh.

15
2.2.3 Failure strain value
To investigate the influence of the failure strain value five models with the same mesh
density, 64 through thickness elements with adaptive meshing has been simulated. The
values of failure strain have been set to 1.5, 2.0, 3.0, 3.5 and 4. The results from the
simulations showed that the failure strain value mainly affected the stroke to fracture and
had a small influence on the maximum force, which can be seen in Figure 2.11.

Shear failure value Docol 800 DP

10000

8000 Shear 1.5


Shear 2.0
Force [N]

6000 Shear 3.0


4000 Shear 3.5
Shear 4.0
2000 Experiment
0
0.00 0.10 0.20 0.30 0.40 0.50
Stroke [mm]

Figure 2.11: Force versus stroke curves for different values of plastic failure strain. A mesh density
of 64 through thickness elements with adaptive mesh was used in all simulations. The simulated
curves are denoted with “Shear” and a corresponding value of the failure strain.

The influence of failure strain value on the characteristic zones have been investigated
using a results from simulations in DEFORM[2] as a benchmark for a realistic edge
geometry. This was done because no experimental data for the edge geometry could be
found for this set of process parameters.

Measurments of the characteristic zones were taken by using the distance query tool in
ABAQUS/CAE, which calculate the distance between nodes picked by the user. Defintion
of the zones have been done by taking the rollover as the distance between the top of the
blank and the first node at the transition to shear zone. The shear zone is measured between
the transition point until initiation of fracture and the remaining part down to the lower
edge of the blank is considered to be the fracture zone. Burr is measured from the lower
edge of the blank. See Figure 2.12 for details concerning how the zones are measured.

16
Figure 2.12: a) Rollover is measured from the top of the sheet edge to the first node of the shear
zone. b) The shear zone is measured from the end of rollover until the start of fracture. c) The
fracture zone is measured from the start of fracture until the lower edge of the sheet. d) Burr is
measured from the lower edge of the sheet.

The characteristic zones were measured using the procedure outlined above and the results
are presented together with results from DEFORM for comparison, see Figure 2.13. The
measured edge profiles are presented in Figure 2.14.

C haracteristic zones

1.2
Length of characteristic zones [mm]

0.8 Rollover
Shear zone
0.6
Fracture zone
0.4 Burr height

0.2

0
1 2 3 4 5 6

Figure 2.13: The characteristic zones measured with different values on the failure strain. Bar 1-5
are measurements from simulations done in ABAQUS with failure strain: 1) 1.5, 2) 2, 3) 3, 4) 3.5,
5) 4. Bar 6 is measurements from a reference case taken from DEFORM. All ABAQUS simulations
were done with 64 elements through thickness and adaptive mesh.

17
Figure 2.14: Simulated sheet edges with value of strain at failure of a) 1.5, b) 2.0, c) 3.0, d) 3.5 and
e) 4.0. The edges were simulated with 64 through thickness elements and adaptive meshing.

As the shear zone was considered to short to be realistic in the simulations made with
ABAQUS, an attempt to produce edge geometry closer to the one achieved with DEFORM
was made by increasing the mesh density to 128 elements through thickness. Failure strain
18
values of 3.5 and 4 were used. The model with failure strain of 3.5 were simulated both
with and without adaptive mesh but as can bee seen from Figure.2.15 the results with
adaptive mesh were not as close to experimental results as the ones with a standard
Lagragian mesh definition. Because of the erroneous results produced with adaptive
meshing, it was discontinued for the simulations with a failure strain of 4, which produced
a stroke to failure similar to experiments. The deterioration of the results with adaptive
mesh can probably be explained by the large number of mappings of the solutions to a new
mesh that has to be done when the failure strain value is set to high values, yielding an
iteration count of about 5 million increments. Also some kind of incompatibility with
adaptivity and failure combined is suspected. The simulated edge geometry for the models
with 128 elements through thickness can be seen in Figure 2.16. Measurements of the
characteristic zones are presented in Figure 2.17.

Docol 800 128 elements through thickness

16000
14000 Failure strain = 3.5
12000
Failure strain = 4.0
Force [N]

10000
8000
Experimental
6000
4000
Failure strain = 3.5
2000 adap
0
0 0.1 0.2 0.3 0.4
Stroke [mm]

Figure.2.15: Force versus stroke curves with 128 through thickness elements. “adap” indicates
that adaptive mesh have been used.

19
Figure 2.16: The simulated sheet edges with 128 through thickness elements with a failure strain
value of a) 3.5, b) 4 and c) 3.5 with adaptive mesh.

Characteristic zones
Length of characteristic zones [mm]

1.2

0.8 Rollover
Shear zone
0.6
Fracture zone
0.4 Burr height

0.2

0
1 2 3 4

Figure 2.17: Measurements of the characteristic zones with different values of failure strain. Bar 1-
2 are measurements from ABAQUS using a failure strain of 1) 3.5 and 2) 4. Bar 3 is simulated in
ABAQUS using a failure strain of 3.5 and adaptive mesh. Bar 4 is a reference case from DEFORM.
All simulations in ABAQUS used 128 elements through thickness.

The difference in the initial response in simulations and experiments can probably be
explained by the fact that the experimental punching setup has elastic properties when
loaded while the tools in the simulations are modelled as rigid. Another possible source of
error is that the material model used in the simulations is unable to predict the actual
behaviour of the material when subjected to this kind of deformation.

20
2.3 Comparison with results from DEFORM
DEFORM is an implicit code that is mainly used to simulate different kind of metal
forming operations such as forging, machining and rolling. To facilitate these kind of
simulations it has a more advanced mesh updating function were an adaptive refinement
procedure is implemented in the code, which allows severe case of deformation without
distorting the mesh. The code also has more models for fracture initiation and propagation
than is implemented in ABAQUS today. A drawback with such a specialised code is the
reduced capability to simulate more general problems were for example non-metal
materials is included. Another issue is that DEFORM is not as commonly used in industry
as ABAQUS. Therefore a comparison between the results produced with the two different
codes is of interest to see if the process can be simulated with no mesh updating and the
less advanced failure criterion used in ABAQUS.

2.3.1 Failure criterion in DEFORM


The Cockroft and Latham failure criterion were chosen to predict failure initiation and
propagation in the work used for comparison [2]. Other authors have also used this failure
criterion with acceptable results [4]. The formula for the Cockroft and Latham failure
criterion is given by:

εf
σ*
³ σ dε = C (4)

Where σ * is the maximum principal stress, σ is the effective stress, ε is the effective strain,
ε f is the effective strain at fracture and C is the damage value which is a material
parameter. When the damage value C reaches a critical value in an element, given as user
input, the element is deleted from the mesh. The interpretation of this criterion is that a
high degree of triaxiality invokes fracture at a lower strain level compared to the uniaxial
case.

2.3.2 Load curves


Two important results from the simulations are the maximum force and the stroke to
failure. ABAQUS and DEFORM shows good agreement in predicted maximum force and
stroke to failure compared to experiments. The noise produced in the ABAQUS curve is a
result of mass scaling techniques being used to reduce the cpu time necessary to perform
the simulations.

21
ABAQUS vs DEFORM and experiment

10000

8000
Experiment
6000
Force [N]

DEFORM C=1.5
4000

2000 ABAQUS Failure


strain=4
0
0 0.1 0.2 0.3 0.4 0.5
-2000
Stroke [mm]

Figure 2.18: Comparison between load curves from ABAQUS, DEFORM and experiments.

2.3.3 Simulated edge geometry


The dimensions of the different characteristic edge zones have been compared earlier and
can be found in Figure 2.13 and Figure 2.17. The measured zones from the two codes
differ somewhat but are in reasonably good agreement, keeping in mind that the
measurements are dependent on how the author chooses to define the limits of the different
zones. The appearance of the edges is presented in Figure 2.19. The edges simulated in
ABAQUS appear to be smoother than in DEFORM because of the higher degree of
discretization necessary to capture the geometry and fracture behaviour when a adaptive
mesh refinement method are not available.

Figure 2.19: Simulated sheet edges in DEFORM (left) and ABAQUS (right). The ABAQUS
simulation has 128 elements through thickness with a plastic failure strain of 4,Adaptive meshing
was not used.

22
2.4 Elastic tools
In order to resolve the difference between using elastic and rigid tools, a simulation model
using elastic tools were created see Figure 2.20. The boundary conditions is applied in the
same manner as the simulations using rigid tools except for the contact between the tools
and the blank, which is modelled as a pure master-slave contact pair with element-node
contact surfaces. This contact condition prescribes that nodes on slave surfaces cannot
penetrate the master surfaces but nodes from master surfaces can penetrate slave surfaces.
In this case, the blank is modelled as a nodal based slave surface and the tools are element
based master surfaces. The motivation for the choice of master and slave surfaces is that
ABAQUS requires parts that can fail to be modelled with nodal surfaces and nodal
surfaces is always slave using this contact definition [11].

Figure 2.20: The FE-model with elastic tools.

Figure 2.21 shows the difference between rigid and elastic tools using 128 through
thickness elements in the blank and a failure strain of 3.5. The elastic response in the tools
can be seen as a slight displacement of the curve to the right in the beginning of the
simulation and faster rupture of the blank. The faster rupture is assumed to be an effect of
the elastic springback in the tools during load drop. A greater elastic response would
probably be achieved by using tools that are longer in the 2 direction in Figure 2.20.

23
Elastic vs Rigid tools

11500

9500

7500
Force [N]

Shear 3.5 128 el elast


5500
Shear 3.5 128 el
3500

1500

-500
0 0.1 0.2 0.3
Stroke [mm]

Figure 2.21: Force stroke curves with rigid and elastic tools. Both simulations had a plastic failure
strain of 3.5 and 128 elements through thickness.

The stress in the tools is shown for different levels of stroke in Figure 2.22-19. The Von
Mises stress, x-stress component (S11) and y-stress component (S22) is presented in the
mentioned order, the orientation of the directions can be seen in figure 2.22 (1 is the x
direction and 2 is the y direction). It can bee seen that the blank mesh is too coarse to
capture the sharp corner radius of the punch. This leads to that only a few nodes from the
blank are in contact with the punch creating very high local stress levels (over 6500 MPa in
Figure 2.23) and sharp gradients in the stress field. Severe penetration of the elements in
the slave surface by the punch master surface is also observed.

24
Figure 2.22: Effective stress and components in the punch and blank at a stroke of 0.05 mm. The
stresses are given in MPa.

Figure 2.23: Effective stress and components in the punch and blank at a stroke of 0.15 mm. The
stresses are given in MPa.

25
To resolve these problems the elements in the refined area was divided by 2 in the x
direction. This resulted in a better interaction between the punch and blank and
smoothened the gradients in the stress field see Figure 2.24. The effective stress level was
about 1000 MPa evenly distributed around the radius instead of 2600 MPa directly under
the node contact points, which were the case in Figure 2.23. With these results in mind,
conclusions can be made that a very high level of mesh refinement is needed to resolve the
state of stress in the punch during the cutting operation when using this type of contact
condition.

Figure 2.24: Effective stress and stress components in the punch and tool at a stroke of 0.05mm.
The stresses are given in MPa.

26
Figure 2.25: Effective stress and components in the punch and blank at a stroke of 0.15mm. The
stress is given in MPa.

2.5 Comments
It has been shown in this section that simulations with ABAQUS using an axisymmetrical
model can predict the maximum punch force and stroke to failure with good accuracy,
using a plastic failure strain of 4. Calibrations against experiments have to be done to
determine the value of the plastic failure strain.

Comparisons with simulations made with the DEFORM FE-code showed good agreement
in terms of the characteristic edge zones and stroke to failure when using a value on the
failure parameter that produces a stroke as measured in experiments.

Simulations using elastic tools showed that the elastic response had a small effect on the
load curve (Figure 2.21), causing a delay in the rise of the punch force and a shorter stroke
to failure. A greater difference is expected if simulations with tool dimensions similar to
the ones used in experiments would be made, this has not been done because of the long
cpu time such a simulation would require. The tool stress could not be resolved using the
mesh density from the simulations with rigid tools because of the single sided contact
condition allowed penetration by the punch radius into the mesh of the blank. This was
improved when the mesh density in the blank was increased in the radial direction,
showing more reasonable stress levels, but further studies has to be made if estimations of
the accuracy shall be possible.

27
3. Material properties

3.1 Materials
The materials that will be used in the simulations are DC04, Docol 350 YP, Docol 800 DP
and Docol 1400M. The Docol steels are cold reduced high and ultra high strength qualities.
The corresponding numerical value to each Docol grade corresponds to the lowest yield
strength for the YP qualities and lowest tensile strength for the DP and M qualities. DC04
is a mild steel quality suitable for deep drawing.

3.2 Tensile tests


To generate the necessary input data for the simulations tensile tests have been performed
for the above mentioned materials. Specimens for each material have been created in the
rolling direction (0 degrees), 45 degrees and transverse to the rolling direction (90
degrees). The tests were made and documented by SSAB Tunnplåt. True stress versus true
strain curves generated during the tests can be found in Appendix I.

3.2.1 Mechanical properties


The mechanical properties for the different materials in the directions mentioned earlier are
given in Table 3.1. The parameter r is the plastic strain ratio that is used to assess the
anisotropy of blank materials [7]. It can be expressed in terms of strains as:

εw
r= (5)
εt

Were εw is the strain in the width direction and ε t is the strain in the thickness direction.
Another parameter that is useful for determining the properties of sheet metal is r which is
measure of the normal anisotropy i.e. the average anisotropy in the plane of the sheet. A
high r value shows that the preferred flow direction is in the plane of the sheet ( r >1)
while a low value ( r <1) indicates a preferred flow direction in the thickness direction. The
normal anisotropy parameter can be calculated from the plastic strain ratios with the
following formula [8]:

r0 + 2r45 + r90
r= (6)
4

Were the indices denote the orientation of the tensile direction given as the angle to the
rolling direction.

28
Table 3.1: Mechanical properties of the different sheet materials. r is the normal anisotropy.
Material Direction Rp0.2 [MPa] Rm [MPa] r r n
designation
Docol 350 YP 0° 363 423 0.67 0.11
45° 369 416 0.97 0.11
90° 403 432 1.05 0.92 0.11
Docol 800 DP 0° 653 873 0.70 0.11
45° 645 862 0.98 0.11
90° 659 887 0.87 0.88 0.11
Docol 1400 M 0° 1277 1486 0.75 0.11
45° 1258 1459 0.61 0.11
90° 1278 1496 0.62 0.65 0.11
DC04 0° 184 308 2.26 0.21
45° 187 315 1.55 0.20
90° 176 299 2.86 2.06 0.20

3.2.2 Extrapolation of test data


Data produced with uniaxial tensile tests is only valid until the point of necking. Necking
causes a triaxial stress state in the specimen that alters the strain hardening behaviour. The
strain developed in the blank during a cutting operation exceeds the necking strain in
uniaxial tension by far, which calls for other procedures to characterise the material
response under high strain. One way to overcome this problem is extrapolating the yield
data from the tensile test, starting with the point at the onset of necking. This has been
done using a method called double Voce extrapolation were the yield data is fitted to the
expression given by:
P P
σ = σ 0 + C1 (1 − e −C ε ) + C 3 (1 − e −C ε )
2 4
(7)

Were σ 0 is the initial yield strength of the material and ε p is the plastic strain. A method
for determining the constants in Equation (7) has been developed by Dr Mats Sigvant [13].
This method is based on using the Ag value and two points before to resolve the constants
K and n in a power law extrapolation such as:

σ = K (ε p ) n (8)

Thereafter the stress at a fourth point is computed by adding a strain of four percent to the
strain that is found at the Ag and inserting this value into Equation (8). This set of points is
then used to compute the constants in Equation (7). The resulting curve increases almost
like a power law after the Ag value but as the strain increases the exponential terms
vanishes and leaves only the initial yield stress and the constants C1 and C3 thus forming a
plateau at higher strains. See Figure 3.1 for extrapolated yield curves concerning the
different materials used in this study.

29
Double Voce extrapolated yield curves

1800.00
Yield stress [MPa] 1600.00
1400.00
1200.00 Docol 800 DP
1000.00 Docol 1400 M
800.00 Docol 350 YP
600.00 DC04
400.00
200.00
0.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Plastic strain

Figure 3.1: Double Voce extrapolated yield data for the different materials.

4. Experiments
To generate data for validation of the simulations, punching tests with the before
mentioned materials have been performed by Uddeholm Tooling AB. The experiments
give information about the characteristic zones of the cut edges, forces used to penetrate
the blank and the stroke required to complete the punching operation. For a summarisation
of experimental data see Appendix 2.

4.1 Experimental setup


The machine used to produce the experimental data is an excenter driven press with
sensors for measuring position and forces added to the main tripod. A piezoelectric load
cell is used for registering forces while an inductive position sensor is measuring the
position with a sampling frequency of 10kHz. The layout of the machine and position of
the sensors can be seen in Figure 4.1-2.

Figure 4.1: A section view of the main tripod were: 1) Position of the load cell, 2) Punch, 3) Blank
holder, 4) Die.

30
Figure 4.2: The main tripod with the position measuring sensor marked with a red arrow.

4.2 Data from experiments


In order to be able to compare the load curves from the experimental punching machine
with simulations, some postprocessing had to be done. Because of the limited space in the
machine, the load cell used to register the forces had to be placed in such a way that all
forces used in the operation is registered see Figure 4.3 for an example of an experimental
load curve.

Punch forces DC350YP

22000
20000
18000
Total force [N]

16000
14000
12000
10000
8000
6000
-0.2 0.8 1.8 2.8 3.8
Distance betw een edges [m m]

Figure 4.3: Experimental load curve

To extract the data of interest from the experimental load curve a straight line was fitted to
the part of the force curve were the punch had not yet made contact with the blank and
only the force of the blank holder is being registered, see Figure 4.4. The equation of this
line gives the distance dependency of the holder force, which then can be withdrawn from
the total force. The stroke is also corrected so the zero level is when the punch makes first
contact with the blank.

31
Blank holder force vs stroke

10000

8000
Force [N]

6000

4000

2000

0
0 0.5 1 1.5 2 2.5 3 3.5 4
Distance between edges [mm]

Figure 4.4: Straight line fitted to the blank holder force

The above corrections have been made to all experimental load curves and the result is
presented in Figure 4.5

Experimental load curves

35000
30000
25000
DC04 Experiment
Force [N]

20000
DC350YP Experiment
15000
DC800DP Experiment
10000
DC1400M Experiment
5000
0
-5000 0 0.2 0.4 0.6 0.8 1
Stroke [mm]

Figure 4.5: Corrected experimental load curves.

5. Simulation model
The experimental data presented in Section 4.2 have been produced with different tool
geometry than were used during the parameter study in Section 2. A different material
model was also to be used for DC04. Therefore a new simulation model had to be made to
simulate the punching operation.

5.1 Geometry and mesh density


An axisymmetric FE-model was used in all simulations with tool and blank dimensions
matching those that were used in the experiments, see Figure 5.1 and Table 5.1. The mesh
density was 128 elements through thickness in the sheared zone and were chosen based on
the results from Section 2.

32
Figure 5.1: The axisymmetric FE-model.

Table 5.1: Tool and blank dimensions used with the different materials.
Material r1 [mm] r2 [mm] r3 [mm] r4 [mm] c1 [mm] c2 [mm] t [mm]
DC04 4.94 0.02 0.01 0.02 0.06 0.26 0.98
DC350YP 4.94 0.02 0.01 0.02 0.06 0.26 1
DC800DP 4.90 0.02 0.01 0.02 0.10 0.3 0.96
DC1400M 4.90 0.02 0.01 0.02 0.10 0.3 0.97

5.2 Material model


The isotropic von Mises material model described in Section 2.1.3 have been used for all
materials except DC04 for which the model proposed by Hill in 1948 have been evaluated.
Usage of this model is motivated by the high values of the plastic strain ratio shown by
DC04. The yield criterion for this material model can be expressed in terms of stress
components as:

2 2 2
f (σ ) = F (σ 22 − σ 33 ) 2 + G (σ 33 − σ 11 ) 2 + H (σ 11 − σ 22 ) 2 + 2 Lσ 23 + 2 Mσ 31 + 2 Nσ 12 (9)

Were the constants F, G, H, L, M and N are obtained by tests of the material in different
directions. The definition of these constants is:

1 1 1 1
F= ( 2 + 2 − 2 ),
2 R22 R33 R11
1 1 1 1
G= ( 2 + 2 − 2 ),
2 R33 R11 R22
1 1 1 1
H= ( 2 + 2 − 2 ),
2 R11 R22 R33
(10)
3
L= 2
,
2 R23
3
M = ,
2 R132
3
N= ,
2 R122

33
The parameters R11, R22, R33, R12, R13 and R23 are anisotropic yield stress ratios defined as
follows:

σ 11 σ 22 σ 33 σ 12 σ 13 σ 23
, , , , , (11)
σ0 σ0 σ0 τ0 τ0 τ0

The type of anisotropy that will be modelled here is the normal anisotropy, which
describes the material as isotropic in the plane with increased yield strength in the
thickness direction [11]. This kind of behaviour is modelled by setting (with the directions
1 and 3 in the plane of the blank and 2 in the thickness direction):

R11=R33=1 (12)

r +1
R22 = (13)
2

Were r is the normal anisotropy parameter given by Equation (6). The remaining
parameters are set to a value of unity.

5.2.1 Strain rate dependency


The material is modelled with a strain rate dependent function to account for the increased
yield strength during high strain rate deformation. This is implemented in ABAQUS by
specifying a yield stress ratio as a tabular function of the equivalent plastic strain rate:

σ (ε pl )
R= (14)
σ0

Were σ (ε pl ) is the yield stress at a specified plastic strain rate and σ 0 is the reference
yield stress at a zero strain rate. The yield stress ratios used for the different materials and
plastic strain rates can be found in Table 5.2.

Table 5.2: Yield stress ratios for the materials used in the simulations at different levels of
equivalent plastic strain rate
Yield stress ratio Equivalent plastic strain rate
DC04 DC350YP DC800DP DC1400M
1 1 1 1 0
1.064 1.069 1.04 1.054 10
1.145 1.126 1.072 1.078 100
1.28 1.191 1.108 1.106 1000
1.28 1.191 1.145 1.127 10000
1.28 1.191 1.193 1.160 100000

5.3 Failure criterion


The shear failure model that is described in Section 2.1.5 has been used in all simulations
to predict the rupture behaviour.

34
5.4 Contact and boundary conditions
5.4.1 Contact conditions
At first attempts were made using the pure master-slave kinematic contact definition as in
Section 2. This proved to be insufficient when simulating materials that showed large
plastic deformation before rupture. When the mesh became distorted the distance between
the nodes in the blank became too great which caused penetration at the punch edge as can
bee seen in Figure 5.2.

Figure 5.2: Penetration of the punch edge into the blank. The part of the punch that has penetrated
the blank is marked with red.

This effect altered the state of deformation around the punch edge, allowing high strains to
be generated in the elements thus triggering the failure criterion at an early stage of the
simulation. To prevent this phenomenon a new approach was made by defining a balanced
penalty contact condition between the edge and side of the punch and the top elements of
the blank as shown in Figure 5.3. A similar definition was also added between the blank
and die.

Figure 5.3: The penalty contact pair consisting of a) the discrete rigid punch and b) the top
element surface of the blank.

The penalty contact is a less stringent enforcement of contact than the kinematic contact
condition but allows for more general types of contact allowing rigid surfaces to act as

35
slave surfaces. With the penalty contact condition the algorithm searches for slave node
penetrations and if penetration is found a force is applied to the slave nodes to oppose the
penetration while an force that is equal acts on the master surface on the penetration point.
This is often referred to in the literature as adding a spring between the surfaces to prevent
penetration [14]. When the balanced penalty contact algorithm is used the forces are
computed in two steps were the master and slave designation is switched between the
surfaces. An average of the two forces is then applied [11]. When using this approach of
contact modelling the penetration of the punch edge into the blank is eliminated, see

Figure 5.4: Deformation state around the punch edge with balanced penalty contact.

When the first element is removed from the top surface of the blank the penalty contact is
no longer active. The contact between the blank and punch is then enforced by a kinematic
contact condition. The motivation for two different contact algorithms to handle the a
single contact between two surfaces is that ABAQUS does not allow using the same
contact algorithm to enforce contact with both the top elements of the blank and the
internal nodal surface that becomes exposed during the rupture.

36
5.4.2 Boundary conditions
As the model is axisymmetric the only boundary condition that needs to be applied directly
to the blank is a locking of the nodes at the symmetry line in the radial direction. To hold
the blank in place during the punching operation a force of 10kN is applied in the
downward direction on the blank holder. This force is applied during the first step of the
solution with a ramp function during 0.001s and is then held at this level throughout the
simulation.

Friction between the blank and tools are of Coulomb type, which is described by Equation
(2). A friction coefficient of 0.2 has been used throughout the simulations.

The punch is controlled by a prescribed velocity boundary condition were it is given a


velocity of 70 mm/s. This velocity has been determined by a linear regression on
experimental data, see Figure 5.5. The velocity is then found by differentiating the
equation of the regression line with respect to time.

Punch stroke vs time y = 70.144x + 0.0261

1.2
1
Stroke [mm]

0.8
0.6
0.4
0.2
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016
Time [s]

Figure 5.5: A straight line fit to experimental data concerning the stroke as a function of time. The
zero level for both time and stroke is set at the position were the punch makes first contact with the
blank.

37
6. Comparison with experiments
It is necessary to compare the simulations with experimental results to evaluate the validity
of the simulations. In this work comparisons to experiments will be made concerning the
load curves, edge profiles, characteristic zones of the cut edges, maximum punch force,
work done during the punching operation and stroke to failure. All simulations have been
performed with the FE-Model described in Section 5.

6.1 Method of evaluating simulations


To be able to extract the necessary data from the simulations in a reliable manner a
framework for evaluating the simulations have been established.

6.1.1 Evaluation of simulated characteristic edge zones


The rollover zone is generated by the initial plastic deformation of the blank. This has been
taken to be between the top of the blank and the first node that is found on the edge at the
transition point to the sheared zone, see Figure 6.1.

Figure 6.1: Measurement of the rollover zone.

As the punch continues to penetrate the blank the shear zone is formed. This zone is
defined to be between the rollover zone and the start of crack initiation leading to final
rupture of the blank. When simulating materials that experiences large deformation during
forming of the shear zone the mesh becomes distorted making the crack initiation hard to
distinguish. Therefore two different measurement criteria have been evaluated: one were
the shear zone is measured to the point were the first element is deleted at the punch edge
and one where a crack of 0.05mm is to be initiated near the punch edge before the
measurement is taken. A graphical interpretation of the two criterions is shown in Figure
6.2.

38
Figure 6.2: Two different criterions for determining the lower boundary of the shear zone: a) A
crack tip of 0.05mm at the edge of the punch b) The first element deletion at the edge of the punch.

The rupture zone is then evaluated from the lower boundary of the shear zone down to the
lower edge of the blank.

6.1.2 Evaluation of work


The area under the graph of the load curve is used to evaluate the amount of work done by
the punch during the punching operation. A numerical integration method is used to
compute the area under the curve. The boundaries for integration are between the first
contact of the punch and blank and until the load has dropped to 25% of the max level
because of some curves are lacking zero crossing.

6.1.3 Evaluation of stroke to failure


The stroke to failure is measured at the final load drop. If no zero crossing is observed at
the primary part of the load drop, a line will be drawn with the same slope and the stroke
will be measured at that point see Figure 6.3.

Figure 6.3: A schematic description of how the stroke to failure are mesured when zero crossing is
not found in the primary part of the load drop.

39
6.1.4 Evaluation of maximum punch force
The maximum force is taken as the first peak value registered in the load curve.

6.2 Load curves and edge profiles


6.2.1 DC04
This steel quality has good deep drawing qualities and is highly ductile, resulting in a long
stroke to failure. At the stroke that failure occurs the mesh is heavily distorted thus
preventing the crack tips originating from the punch and die to propagate through the mesh
and meet causing a delayed final rupture. When modelling the normal anisotropy using the
Hill 48 material model the maximum force is overpredicted but the force level during
forming of the shear zone is closer to experiment. The failure strain used when simulating
this material was set to 3.5.

DC04

12000

10000

8000
Force [N]

6000 Experiment
Simulation Hill 48
4000 Simulation von Mises

2000

0
0.00 0.20 0.40 0.60 0.80 1.00
-2000
Stroke [mm]

Figure 6.4: Comparison between simulated and experimental load curve for DC04. The failure
strain value was set to 3.5 in both simulations. When using the Hill 48 material model, normal
anisotropy was assumed.

40
When comparing edge profiles the spatial orientation of the zones differs as an effect of the
larger rollover zone that was formed in experiments. The shorter rollover zone in
simulations also causes a longer fracture zone than the one being measured in experiments.
The simulated and experimental edge profiles are presented in Figure 6.5.

Figure 6.5: Experimental and simulated edge profiles. The profile to the left is simulated with Hill
48 material model while the right profile has been simulated using an isotropic von Mises model.

6.2.2 Docol 350 YP


When simulating this material the model was able to capture the stroke to failure and
maximum forces involved with reasonably good accuracy. The failure strain value was set
to 3.5 for the simulation presented in Figure 6.6.

Docol 350 YP

14000
12000
10000
Force [N]

8000 Simulation
6000 Experiment
4000
2000
0
0.00 0.20 0.40 0.60 0.80 1.00
Stroke [mm]

Figure 6.6: Comparison between simulated and experimental load curves for Docol 350YP. The
failure strain was set to 3.5 in the simulation.

41
The experimental and simulated edge profiles in Figure 6.7 shows good agreement when
compared to each other. It can bee seen from the picture of the experimental edge profile
that the shear zone has a slight variation along the edge which cannot be represented in the
simulations due to symmetry and thus adding a possible error source.

Figure 6.7: Experimental and simulated edge profiles.

6.2.3 Docol 800 DP

The simulation of Docol 800DP was able to capture the maximum force and the final load
drop with good accuracy but underpredicted the stroke to failure. This could have been
done by increasing the failure strain value but then the work done (area under the graph)
when simulating the process would be more than measured in experiments. Therefore the
failure strain value was set to 3 yielding a simulated load curve that can be seen in Figure
6.8.

Docol 800DP

20000

15000
Force [N]

10000
Simulation
5000 Experiment

0
0.00 0.20 0.40 0.60 0.80 1.00
-5000
Stroke [mm]

Figure 6.8: Comparison between simulated and experimental load curve. The failure strain was set
to 3 in the simulation.

42
When comparing experimental and simulated edge profile a good agreement was found,
see Figure 6.9.

Figure 6.9: Experimental and simulated edge profiles.

6.2.4 Docol 1400M


Simulations of Docol 1400M captured the maximum force and final load drop with good
accuracy. The stroke to failure was as in the case of Docol 800DP underpredicted because
of the difference between simulations and experiments in the initial stage of the process.

Docol 1400M

35000
30000
25000
20000
Force [N]

15000 Experiment
10000 Simulation

5000
0
0.00 0.20 0.40 0.60 0.80 1.00
-5000
-10000
Stroke [mm]

Figure 6.10: Comparisons between simulated and experimental load curve. The failure strain was
set to 3 in the simulation.

43
When comparing experimental and simulated edge profiles the rollover was overpredicted
in the simulations but the overall appearance of the edge were captured.

Figure 6.11: Experimental and simulated edge profile.

6.3 Characteristic zones


The characteristic zones from simulations have been measured by the procedure outlined
above and are here compared to experimental values in correlation diagrams. Numerical
values concerning the different zones of the edge geometry are summarised in Table 6.1-3

6.3.1 Rollover
The rollover zone is predicted with good accuracy for DC350YP and DC800DP but the
simulation fails to capture the behaviour DC04 and DC1400M.

Rollover

0.2
Simulation [mm]

0.15

0.1

0.05

0
0 0.05 0.1 0.15 0.2
Experiment [mm]

Figure 6.12: Rollover measured from simulations plotted against experimental values.

Table 6.1: Numerical data concerning the rollover height measured in experiment and simulation.
Material Experiment [mm] Simulation [mm]
DC04 Hill 48 0.16 0.096
DC04 v.Mises 0.16 0.1
DC350YP 0.1 0.09
DC800DP 0.1 0.082
DC1400M 0.01 0.04

44
6.3.2 Shear zone
The shear zone has been measured using two different criterions for evaluation. Results
closer to experimental values are achieved when the shear zones are measured after a crack
tip of 0.05 mm has initiated at the edge of the punch.

Shear zone

0.6
0.5
Simulation [mm]

0.4
0.05mm crack
0.3
Crack initiation
0.2
0.1
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Experiment [mm]

Figure 6.13: Shear zone measured from simulations plotted against experimental values, measured
with two different criterions for evaluation.

Table 6.2: Data concerning the dimensions of the sheared zones measured in experiments and
simulations. The simulated shear zone is measured with two different criterions for evaluation.
0.05 mm crack Crack initiation
Material Experiment [mm] Simulation [mm] Simulation [mm]
DC04 Hill 48 0.56 0.49 0.34
DC04 v.Mises 0.56 0.47 0.36
DC350YP 0.5 0.44 0.3
DC800DP 0.25 0.25 0.194
DC1400M 0.2 0.19 0.165

45
6.3.3 Fracture zone
The length of the fracture zone is dependent on the measurement of the other zones. The
series of data presented shows that variation of the fracture zone when the shear zone is
measured by two different criteria of evaluation. The largest difference in fracture zone is
also shown by DC04 were the simulations did not capture the rollover zone.

Fracture zone

0.8
0.7
Simulation [mm]

0.6
0.5 0.05 mm crack
0.4
Crack initiation
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8
Experiment [mm]

Figure 6.14: Fracture zone measured from simulations plotted against experimental values.

Table 6.3: Data concerning the dimensions of the fracture zones measured in experimental and
simulations.
Material Experiment [mm] Simulation [mm]
DC04 Hill 48 0.25 0.39
DC04 v.Mises 0.25 0. 414
DC350YP 0.4 0.47
DC800DP 0.61 0.638
DC1400M 0.75 0.61

46
6.4 Maximum punch force
The simulations were able to capture the maximum punch force with good accuracy for all
materials used in this study

Maximum force

35000.00
30000.00
Simulation [N]

25000.00
20000.00
15000.00
10000.00
5000.00
0.00
0 5000 10000 15000 20000 25000 30000 35000
Experiment [N]

Figure 6.15: Maximum punch force measured from simulations plotted against experimental
values.

Table 6.4: Numerical data concerning the maximum forces measured in experiment and
simulations.

Material Experiment [N] Simulation [N]


DC04 Hill 48 8773 9166
DC04 v.Mises 8773 8353
DC350YP 11731 11704
DC800DP 17990 17999
DC1400M 30431 29542

47
6.5 Work
When comparing the work performed by the punch during the operation a good match was
found between simulations and experiments. The work done during the simulations has
been used as an evaluation parameter for materials were experimental and simulated load
curves differ at the initial stage of the simulation

Work

12
10
Simulation [Nm]

8
6
4
2
0
0 2 4 6 8 10 12
Experiment [Nm]

Figure 6.16: Work calculated from simulated load curves compared to experimental values.

Table 6.5: Data concerning work measured in experiments and simulations.


Material Experiment [Nm] Simulation [Nm]
DC04 Hill 48 5.68 5.79
DC04 v.Mises 5.68 5.35
DC350YP 6.76 6.10
DC800DP 7.18 7.26
DC1400M 9.66 9.80

48
6.6 Stroke to failure
As mentioned before the stroke to failure was underpredicted for Docol 800 and 1400
while for DC04 and Docol 350 YP the mesh distortion cased a delay in fracture behaviour.

Stroke to failure

1
Simulated stroke [mm]

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
Experimental stroke [mm]

Figure 6.17: The stroke to failure measured in simulations plotted against experimental values.

Table 6.6: Data concerning the length of the stroke to failure measured in experiments and
simulations.
Material Experiment [mm] Simulation [mm]
DC04 Hill 48 0.84 0.92
DC04 v.Mises 0.84 0.94
DC350YP 0.71 0.75
DC800DP 0.56 0.52
DC1400M 0.47 0.39

49
7. Discussion
The results from this study have shown that it is possible to simulate a punching process
using ABAQUS/Explicit and an axisymmetric model definition. Materials included in the
study have yield strengths ranging from 184 to 1277 MPa, which gives a good indication
of the usefulness of the model. As the materials mechanical properties differ by far
simulation difficulties emerges at different areas. The softer material experiences a large
deformation before final rupture which causes mesh deformation resulting in a delayed
fracture, a problem that most likely could have been solved with an adaptive mesh
refinement function. For high strength qualities the initial response in the load curve shows
differences between simulations and experiments. This have been observed in earlier work
using the DEFORM FE-code (Figure 2.18) which implies that the source of error could lie
in the material definition or the elastic response recorded in the experimental punching
machine.

During the course of the work some problems were encountered when simulating materials
that has a large amount of plastic deformation before crack initiation and rupture. With
these materials penetrations of the punch radius into the blank caused a premature failure
of the material or when the failure strain parameter was increased an abortion of the
simulation due to distorted elements. This could be avoided by adding a blended penalty
contact definition between the radius of the punch and top elements of the blank thus
capturing the state of deformation in the blank in a more realistic way.

When comparing experiments and simulations a set of evaluation a set of criterions had to
be adapted for evaluating the simulations. This was done because the transition points
between the characteristic zones are diffuse when large deformations are present in the
FE-mesh. Experiments also shows variations in the characteristic zones when measuring at
different positions on the blank edge which adds an element of insecurity to the
measurements as variations cannot be captured in the simulations due to symmetry.

50
8. Conclusions
• Simulations using ABAQUS can give information regarding the maximum force and
characteristic zones with acceptable accuracy.

• Using a failure strain of 3.5 for DC04 and DC350 YP and 3 for DC800DP and
DC1400M proved to be adequate as similar amount of work are measured in
experiments and simulations with these settings.

• For materials that show a large sheared zone the transition between the characteristic
zones becomes diffuse making precise measurements hard to accomplish. Criterions
for evaluation had to be adapted.

• The mesh density used is of great importance for capturing the rupture behaviour of the
blank metal.

• Correct representation of the contact between tools and the blank is a key factor when
simulating a punching process.

51
9. Future work
• Simulations of parameter variation with experimental verification.

• Implementing improved failure criterions in ABAQUS.

• Evaluation of tool stresses by using elastic tools in the simulations.

• Simulations of metal trimming.

• Making the experiments even more accurate by eliminating elastic response and
dynamic effects.

52
10. Acknowledgements
This project has been founded by SSAB Tunnplåt, TDI AB and Volvo Cars Body
Components within the SIMR member program.

The author would like to express his gratitude to the following people for their
contributions to the project:

Berne Högman, Anders Thuvander and Ulrika Åhs, Uddeholm Tooling AB for producing
experimental data used to verify the simulations.
Daniel Eriksson, SSAB Tunnplåt, for supplying tensile test data.
Mats Sigvant, Volvo Cars Body Components, for extrapolating yield data and valuable
advice concerning FE-simulations.
Lars Gunnarson, Swedish Corrosion and Metals Research Institute, who has been my
supervisor during the course of this work and has given me great support and guidance.
Prof Arne Melander, Swedish Corrosion and Metals Research Institute, for help and good
advice during this work.

53
11. References
1. “Tooling solutions for advanced high strength steels selection guidelines”, SSAB and
Uddeholm Tooling, 2004
2. D. Thunvik, FE-simulation of trimming, Swedish Institute for Metals Research, 2004.
3. E.Taupin, J. Breitling, W-T. Wu, T. Altan, “Material fracture and burr formation in
blanking results of FEM simulations and comparison with experiments”, Journals of
Materials Processing Technology, 1996
4. G. Fang, P. Zeng, L. Lou, “Finite element simulation of the effect of clearance on the
forming quality in the blanking process”, Journals of Materials Processing Technology,
2002.
5. R. Hambli, A. Potrion, “Finite element modelling of sheet metal blanking operations
with experimental verification”. Journals of Materials Processing Technology, 2000.
6. W. Klingenberg, U.P. Singh, “Finite element simulation of the punching/blanking
process using in process characterisation of mild steel”, Journals of Materials
Processing Technology, 2002.
7. E.M. Mielnik, “Metalworking science and engineering”, McGraw-Hill, Inc, 1991.
8. A. Elsner, “Advanced hot rolling strategies for IF and TRIP steels”, Delft University
Press, 2005.
9. HKS-ABAQUS, “ABAQUS Theory manual”, Version 6.5-1.
10. N. Ottosen, M. Ristinmaa, “The Mechanics of Constitutive Modelling”, Volume 1
Classical topics, Division of Solid Mechanics, Lund University, 1999.
11. HKS-ABAQUS, “ABAQUS Analysis users manual”, Version 6.5-1.
12. “Handbok och formelsamling i hållfasthetslära”, Instutionen för hållfasthetslära KTH,
1998.
13. M. Sigvant, “The Hemming Process, A Numerical and Experimental Study, Chalmers
University of Technology, Göteborg, Sweden, 2003.
14. J.O. Hallqvist, “LS-DYNA Theoretical manual”, LSTC, 1998.

54
Appendix 1 Mechanical properties

1600
DC04 (t=0.97 mm)
1400
Docol 350YP (t=1 mm)
1200
True stress [MPa]

Docol 800DP (t=0.98 mm)


Docol 1400M (t=1 mm)
1000
800
600
400
200
0
0 0.05 0.1 0.15 0.2 0.25
True strain

Figure 1: True stress versus true strain in the rolling direction for the materials included in this
study.

1600
1400 DC04 (t=0.97 mm)
Docol 350YP (t=1 mm)
1200 Docol 800DP (t=0.98 mm)
True stress [MPa]

1000 Docol 1400M (t=0.98)

800
600
400
200
0
0 0.05 0.1 0.15 0.2 0.25
True strain

Figure II: True stress versus true strain in the 45 degree direction for the materials included in this
study.

55
1600
DC04 (t=0.97 mm)
1400
Docol 350YP (t=1 mm)
1200 Docol 800DP (t=0.98 mm)
True stress [MPa]

1000 Docol 1400M (t=0.98)

800
600
400
200
0
0 0.05 0.1 0.15 0.2 0.25
True strain

Figure III: True stress versus true strain in the transverse direction for the materials included in
this study.

56
Appendix 2 Data from punching experiments
Sheet Sheet Punch Punch Punch Roll Shear Fracture Burr Hole Hole Max
material thickness material hardness diameter over zone zone heigh diameter diameter force
[mm] [HRC] [mm] [mm] [mm] [mm] [µm] along transverse [N]
[mm] [mm]
DC04 0.97 Sverker 60-62 9.88 0.16 0.56 0.25 11.25 9.88 9.88 8773
21
Docol 1 Sleipner 60-62 9.88 0.1 0.5 0.4 15.25 9.88 9.88 11731
350YP
Docol 0.98 Not Not 9.8 0.11 0.25 0.61 9.6 9.82 9.82 17990
800 DP included included
Docol 0.98 Not Not 9.8 0.01 0.2 0.75 7 9.54 9.54 30431
1400M included included

57

You might also like