You are on page 1of 15

International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Blanking tool wear modeling using the finite element


method
*
Ridha Hambli
ISTIA-LASQUO, 62 Avenue Notre Dame du Lac, 49000 Angers, France
Received 1 December 2000; received in revised form 22 January 2001; accepted 26 January 2001

Abstract

One of the main objectives of the numerical process design in metal forming is to develop adequate
tool design and establish process parameter in order to increase tool life and to improve part quality and
complexity while reducing manufacturing cost. The prediction of tool wear in sheet metal
blanking/punching processes is investigated in this paper using the finite element method. A wear prediction
model has been implemented in a finite element code in which the tool wear is a function of the normal
pressure and some material parameters. A damage model is used in order to describe crack initiation and
propagation into the sheet. The distribution of the tool wear on the tool profile is obtained and compared
to industrial observations. Furthermore, a numerical investigation has been carried out to study the effect
of tool wear on the burr formation.  2001 Elsevier Science Ltd. All rights reserved.

Keywords: Blanking; Finite element; Tool wear; Wear model; Wear prediction

1. Introduction

The tool profile of blanking/punching sheet metal processes is exposed to strong tribological
loads by high contact normal pressure and sliding distances. Tools often show adhesive and abras-
ive wear in the contact zone. The developpement of simulation including tool wear effects can
be used to predict tool life expressed as the number of the workpieces which can be blanked
without defects.
The real-time estimation of tool wear in blanking operations is important for scheduling tool
changing times and for adaptive process control and optimization. However, tool wear cannot be
measured directly during the process. An on-line tool wear estimation method can reduce the

* Corresponding author. Tel.: +33 (0)2-41-36-57-57; fax: +33 (0)2-41-36-57-43.


E-mail address: ridha.hambli@istia.univ-angers.fr (R. Hambli).

0890-6955/01/$ - see front matter  2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 8 9 0 - 6 9 5 5 ( 0 1 ) 0 0 0 2 4 - 4
1816 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Nomenclature
D damage variable
Dc critical damage value at fracture
sel initial yield stress
s0 isotropic non-linear hardening law
seq equivalent Von Mises stress
sH hydrostatic pressure
sTn elastic prediction of the stress tensor components at increment (n).
sn stress tensor at increment (n)
sn+1 Cauchy stress tensor at increment (n+1)
eeq equivalent plastic strain
eD threshold logarithmic strain at which Lemaitre damage initiates
eR logarithmic strain value at fracture
epl plastic part of the strain tensor
⌬epl increment of plastic part of the strain tensor
⌬e total strain increment
f yield function coupled with damage
Cel Hooke elastic operator
Un displacement field at increment (n)
B strain–displacement matrix
Jnr Jacobien tensor at increment (n) and iteration (r)
Knr stiffness matrix at increment (n) and iteration (r)

production costs. The cost of forming tools usually represents a substantial amount of the total
manufactering cost of forming parts [1,2]. In addition, unexpected tool changes due to excessive
wear are cause unaccceptable down times of the manufacturing process.
In order to contribute towards a development of a system for the on-line assessment of tool
wear during the blanking/punching processes, a finite element model has been developed and
implemented into a finite element code [3] allowing for the numerical prediction of tool wear
evolution in function of the various process parameters.
Abrasive wear most commonly occurs in blanking/punching tooling when the surface of the
working material contains hard particles, such as carbides or oxides [4]. These carbide or oxide
particles are harder than the tooling components and will scratch the tool when the working
material slides over the tool surfaces. To increase the abrasive wear resistance of blanking tool
materials, the working hardness of the tool can be increased or tool steel chemistry which forms
hard carbide particles itself can be chosen. Tooling optimization is greatly dependent upon pre-
dicting the dominant failure mechanism for a given application and adjusting tool properties
accordingly through material selection and heat treatment.
R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1817

2. Wear model

In metal blanking, the relative motion of the tool-sheet surfaces may results in a loss of tool
material through adhesive wear [2]. This wear process is initiated by the interfacial adhesive
junctions that form in contact zone [4,5]. As a normal load is applied, local pressure at the contact
area become extremely high. Therefore, the surfaces adhere together and friction between the
sliding surfaces induces the generation of wear particles. The presence of particles with high
hardness accelerates wear by abrasion.
Wear is defined as a slow degradation of the blanking tool caused by friction involved between
tool and metal sheet [6]. The rate of wear is affected by parameters such as tool material, blanked
part material, punch-die clearance, punch velocity, lubrification and material thickness.
The wear resulting from adhesive wear process has been described phenomenologically by the
Archard equation [7]:
V FN
Wad⫽ ⫽k (1)
s 3H
Wad is the worn volume per unit sliding distance, V is the volume of the material removed by
wear from surface, k is a wear coefficient depending on the contacting materials and the sliding
contact conditions. s is the sliding distance, H is the hardness of the sheet and FN is the normal
load applied on the tool. Inspection of Eq. (1) shows that the hardness H is the only material
property appearing in the model. Typical values of the wear coefficient k are given in [2,4] for
a combination of contacting materials. In the present investigation, the value k was taken in the
order of 10E-05.
A simplified expression for the volume of abrasive wear can be given by [4]:
bFNs
V⫽ tan(q) (2)
pH
where b represents that part of the asperities having the ability to cut and q the angle of the
assumed cone-shaped asperities for the hardest material. If the parameters of the wear models are
assumed to be constant through time, the above wear models can be rewritten as:
V⫽gwFNs (3)
where gw denotes a wear coefficient depending on sliding contact conditions [4,7] and varies over
the range of 10⫺2–10⫺7 mm2/N.
In the present paper, gw is taken in the order of 1.3E-04 at the sliding interface of the workpiece
and the punch. This value corresponds to a hard tool steel.
Based on the investigation presented in [2,8–11], the wear resulting from adhesive wear causes
the cutting edges to be rounded (Fig. 1). This would reduce the sharpness of the punch during
shearing, and increase the deformation of workpiece. Moreover, the burrs of the parts become
larger, the noise level in the press becomes very high and the punch penetration corresponding
to the cracks initiation into the sheet increases.
From a numerical point of view, at each node “i” of the contact elements of the tool mesh,
the above equation can be written as:
1818 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Fig. 1. Wear profile of blanking tool.

Vi⫽(gw)i(FN)isi. (4)
For F.E. analysis, instead of normal force, it is easier to use normal contact pressure. At each
node “i”, the normal force (FN)i, can be obtained from the normal stress as (Fig. 2):

(FN)i⫽ 冕 ⍀
(P)i d⍀i (5)

⍀i is the area of each contact element “i” and (P)i the normal contact pressure acting on each
contact node “i” of the tool.

3. Finite element modeling

The law describing the material behavior should allow the description of the different stages
of the process observed experimentally starting from the elastic state and ending in the final
rupture of the sheet. For this, a behavioral law including damage and failure phenomena must be
choosen. In order to predict damage evolution into the sheet metal during blanking processes, the
continuum damage mechanics approach has been applied in this work as a means to describe the
behavior of the sheet using the Lemaitre damage model [12].
The algorithms generally implemented in the finite element codes for integration of non-linear
constitutive equations are the so-called radial return algorithms, and they are used to solve the
equations in an incremental form. They are based upon the notion of an elastic predictor–plastic

Fig. 2. Normal pressure and equivalent normal force acting on contact element.
R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1819

corrector where a purely elastic trial state is followed by a plastic corrector phase [3,9]. In this
way, an implicit algorithm has been developed which allows for the integration of the constitutive
equations. The integration methods of the non-linear constitutive equations are based on the use
of a special algorithm which solves the equations in incremental form. For this purpose, during
a small time interval [tn, tn+1], it is assumed that the whole increment is purely elastic, then an
elastic prediction is defined as:
sTn+1⫽sn⫹⌬s (6)
(6) can be written as:
sTn+1⫽Cel(en+1⫺epl
n ). (7)
The superscript (.)T refers to Trial test and Cel is the elastic modulus tensor depneding on the
damage state of the material [12].
The Von Mises yield function coupled with damage is given by:
f⫽seq⫺(1⫺D)(sel⫹s0) (8)
D is the damage variable.
If this elastic predicton satisfies the yield condition: f⬍0, the prediction is true and the local
procedure is completed. Then it can be stated that:
sn+1⫽sTn+1. (9)
Otherwise, this state must be corrected by means of a plastic correction. For this purpose, the
variables at increment n+1 must satisfy the system [9]:
f⫽0 (10)

sn+1⫺Cel(en⫹⌬e⫺epl
n ⫺⌬e )⫽0
pl
(11)

⌬Ha⫽ha(⌬eij , (sij )n+1, Haij ) (12)


a a
Ha, a=1, 2, …, n, is a set of scalar state variables and h is the hardening law for H .
Within the framework of the displacement formulation of F.E.M., the global equilibrium equa-
tions to be satisfied at each instant tn+1 can be written in the general form [3]:
F(Un+1)⫽0 (13)
Un+1 is the displacement field at step (n+1)
If this non-linear problem is solved iteratively by a Newton method, at each global iteration r
the following equation can be written:
F(Urn+1)⫹(Kn+1 n+1⫺Un+1)⫽0
r
)(Ur+1 r
(14)
where

r
Kn+1 ⫽ 冉 冊 冕
∂F
∂U
r

n+1
⫽ BTJn+1

r
B d⍀ (15)
1820 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

B is the strain–displacement matrix and Jrn+1 is the Jacobien tensor obtained by:

Jrn+1⫽ 冉冊
∂s
∂e
r

n+1
(16)

The adhesive wear model (3), can be expressed in the incremental form:
dV⫽gwdFNds (17)
dV and dFN can be expressed as:
dV⫽dDwd⍀ (18)

dFN⫽P d⍀ (19)
dDw is the depth of the wear, dw is the contact area, and P the normal contact pressure acting
on the tool.
By substituing Eqs. (18) and (19) into Eq. (17), we obtain:
dDw⫽gwPds. (20)
The wear depth Dw can be expressed in the following integral form:


s

Dw⫺gw P ds. (21)


0

This equation can discritized as follows:

冘冘
n m

Dw⫽gw P⌬s (22)


i⫽1j⫽1

m is the total number of time step ⌬t, and n indicated the total number of nodes at the punch-
part contact area.
Within a time interval [tn, tn+1], the wear prediction algorithm leads to the following incremental
wear depth form:
(Dw)n+1⫽(Dw)n⫹gw[Pn+1(ss+1⫺sn)] (23)
where (.)n and (.)n+1 denote the approximation of the variable values at increment n and n+1.

4. Numerical modeling of damage and ductile fracture

The damage phenomenon, described by initiation and growth of cavities and microcracks into
material induced by large deformations in metals has been extensively studied in order to predict
the ductile damage evolution into structures subjected to plastic loading.
In this paper, the numerical simulation of the damage evolution leading to crack initiation and
R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1821

propagation has been descibed by means of continuum damage approach using the Lemaitre
damage model [12]. For isotropic evolution, the damage law valid for any loading path is written
in the incremental form:

Ḋ⫽ 冋
Dc 2
eR−eD 3
(1⫹n)⫹3(1⫺2n)
sH
seq 冉 冊册 2
(eeq)2/n dėeq. (24)

This model depends upon material constants for damage propreties, the hardening exponent n and
Poisson’s ratio n. eD is the threshold logarithmic strain at which damage initiates, eR is the logarith-
mic strain value at fracture and eeq the logarithmic plastic strain. Dc is the critical damage value
at fracture.
Fig. 3 shows the flow chart of the method used for crack propagation analysis. First, at a given
punch displacement step “n”, The behavior law coupled with damage allows for the computation
of the damage value at each finite element of the mesh. During the analysis, the initiation of
crack is supposed to occur at any point in the metal sheet where the damage reaches its critical
value Dc. The crack propagation is simulated by the propagation of a completely damaged area.
This method leads to the decrease in the stiffness of the completely damaged elements.

5. Numerical simulation of a blanking operation

The problem studied here consists of an axisymetric blanking operation of a sheet metal with
the following dimensions:
Thickness: t⫽3.5 mm, die diameter: D⫽20 mm and punch-die clearance: c⫽10%.
In this study, the tool, is modeled by adopting a rigid body hypothesis with contact elements.
The contact surface laws are defined by a Coulomb friction model with a friction coefficient value
of 0.1
The meshing of the model is carried out by means of 1500 quadrangular four node axisymmetric
elements. Fig. 4 shows the mesh used.

6. Results and discussion

Experiments using devices equipped with electrical gauges and a force tranducer, were perfor-
med by a 4000 kN hydraulic press in order to verify the validity of the proposed finite element
model.
The blanked profile of the sheet obtained by experiment is presented in Fig. 5(d), it is noted
that the final rupture is obtained at a punch displacement of about 70% of the sheet thickness.
The computation results corresponding to different displacement steps of the punch penetration
and the corresponding graphical issues are presented in Fig. 5(a)–(c). The crack propagation can
be followed in the mesh until rupture occurs. Refering to Fig. 5(d), Fig. 5(c) shows that the
predicted blanked profile is in good agreement with the experimental one.
As can be observed, the distorsion of the mesh is restricted to a small area near the die-
punch clearance.
1822 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Fig. 3. Schematic diagram of the approach used.

Fig. 4. Mesh used for the F.E. model.


R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1823

Fig. 5. Crack propagation in the mesh (relative clearance =10%). (a) Punch penetration=25% ; (b) punch pen-
etration=30%; (c) punch penetration=70%; (d) experimental profile (penetration=70%).

The different numerical computations, indicated that due to the punch penetration into the sheet,
cracks initiate at the cutting edges of the tools [Fig. 5(a)]. Secondly, the cracks propagate in the
same direction of the punch penetration.
The local contact pressure distribution and the sliding distance variation on the blanking punch
for a penetration about 15% are presented in Fig. 6.
It can be shown that the pressure distribution in the flat contact zone of the punch is constant
for a distance about 1.6 mm. The maximum local load is applied on the cutting edge of the punch
and on the element No. 167 [Fig. 6(a)] as a consequence of the sheet bending. Fig. 6(b) shows
that the relative punch/sheet sliding is localized in the cutting edge area of the punch with a
maximum sliding value applied in the rounded area.
Fig. 7 shows the normal pressure variation on the punch profile versus the the distance (x).
It can clearly observed that the contact pressure at a particular state of the process varies with
distance (x) on the punch profile. The maximum pressure value is concentrated in the rounded area.
If the wear evolution is supposed to be linear versus the number of produced blanked part, the
wear profile of the punch over time can be obtained using the wear model (3). Starting from the
1824 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Fig. 6. Normal contact pressure contour and sliding distance variation on the circular blanking punch profile. (a)
Normal contact pressure contour. (b) Sliding distance variation.

Fig. 7. Normal contact pressure variation with location on the punch.

numerical results plotted in Fig. 6, at each node “i” of the contact elements of the punch mesh,
the local wear volume of the material can be predicted using Eq. (22) as:

冘冘
n m

Dw⫽Nbcgw P⌬s (25)


i⫽1j⫽1

where Nbc is the total number of blanking cycles.


Fig. 8 shows the wear profile of the punch obtained by the numerical prediction and the experi-
mental results after 20 000 blanking cycles.
It can be observed that the wear profile predicted by the proposed wear model give good results
compared to experimental ones. A comparison between experimental and predicted wear profiles
shows that the deviation does not exceed 10%. Wear resulting from punch/sheet contact causes
the cutting edges to be rounded [13], hence it affects the quality of the parts produced and gener-
ates the burr formation [8,9,11]. The influence of tool wear on the blanking process can be mod-
eled in a simplified manner by changing the value of the punch edge radius Rusp.
Several calculations were performed in order to compare the numerical results with experi-
R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1825

Fig. 8. Wear profile of the punch after 20 000 blanking cycles. (a) F.E.M. calculation. (b) Experiment.

mental data. The influence of the tool wear on the punching force and on the evolution of the
sheared profiles was accounted for.
In Fig. 9(a), the evolution of the punching forces relative to the punch penetration for different
simulations are shown.
When the material damage is taken into account, it is possible to simulate the punch penetration
up to 80% of the sheet’s thickness, in spite of a large mesh distortion.
The experimental plots in Fig. 9(b) show that for the various states of punch wear, there is no
difference between the maximum blanking loads. Neverthless the punch penetration Udr corre-
sponding to the crack initiation in the sheet-material increases with an increase in punch wear.
This is due to relative increase in the contact area between the punch and the sheet.
The punch penetration evolution Udr corresponding to the crack initiation against the wear
parameter Rusp obtained by experiments on 1060 steel and F.E.M. simulations, is plotted in Fig.
10. The results shows that the punch penetration prior to fracture increases linearly with increases
in the value of the punch wear radius.
The two blanked profiles obtained by finite element calculation and experiment corresponding
to a new punch with a cutting edge radius Rusp =0.01 mm and a worn punch with Rusp=0.2
mm, are shown in Fig. 11(a) and (b). As can be expected, in the case of a worn punch, the profile
of the part boundary presents a bad quality due to the presence of a burr.
The numerical results compared with the experimental ones, show the reliability of the finite
element model according to the aformentioned wear algorithm prediction in describing the influ-
ence of punch wear on the burr formation.
Fig. 12 shows the evolution of the burr height with increasing punch wear radius obtained from
experiments and F.E. simulation.
It can be shown that results are in good agreement with deviations of 11% between predicted
and experimental values.
F.E. simulation indicates that punch initiates cracks at the cutting edges of the tools. Secondly,
the cracks propagate in the same direction of punch penetration as is illustrated by the curves in
Fig. 13. These curves show the relations between relative crack propagation depth (expressed as
1826 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Fig. 9. Punching force versus punch travel for four wear states. (a) Simulation. (b) Experiment.

Fig. 10. Punch penetration at rupture Udr versus the wear radius Rusp.

a percentage of sheet thickness) and the relative punch displacement obtained by experiment and
simulation, for new and worn punchs. Crack initiation and propagation can be predicted accu-
rately. From the moment of crack initiation to complete rupture of the sheet, experimental and
numerical results are in good agreement.
R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1827

Fig. 11. Predicted and experimental blanked profiles corresponding to two states of the punch wear (clearance=10%).
(a) New punch: Rusp=0.01 mm. a1, Simulation; b1, experiment. (b) Used punch: Rusp=0.2 mm. a2, Simulation; b2,
experiment.

Fig. 12. Evolution of the burr height versus the punch wear radius.
1828 R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829

Fig. 13. Crack propagation versus punch penetration.

7. Conclusions

In this investigation, a tool wear prediction algorithm has been implemented into a finite
element code in order to predict tool wear evolution during blanking/punching processes. The
comparative study between the experimental results and the numerical ones, shows the good agree-
ment.
The numerical results compared with the experimental ones, show the reliability of the model
in predicting the tool wear profile as well as burr height evolution versus the blanking cycles.
From an industrial point of view, the wear model can be used to assess the tool life for a
given application.
In order to predict tool life and the tribological loading applied on the tool, the proposed wear
model can be used in other sheet metal forming processes if the same wear mechanisms are
active. The two curves of Fig. 13 are sufficiently close to conclude that the procedure that we
propose to simulate crack initiation and propagation, gives good results.
In general, the need for regrinding of the shearing tool is determined on the basis of allowable
burr height on the final product. The proposed wear algorithm allows for the prediction for the
need of tool regrinding in two steps. During the first step, one can compute the tool radius variation
versus the number of blanking cycles. During the second step, it possible to predict the burr
height corresponding to the number of blanking cycles. This wear wear analysis is very helpful
to improve the reliability of the shearing tool and to determine the tool repair or change.

Acknowledgements

The author wishs to thank Professor Alain Potiron for his assitance during this investigation
and Deville S.A. Company for its support.

References

[1] D. Hortig, D. Schmoeckel, Analysis of local loads on the draw die profile with regard to wear using the FEM
and experimental investigations, in: The 7th Int. Conf. on Sheet Metal, Erlangen, Germany, 1999, p. 193–202.
R. Hambli / International Journal of Machine Tools & Manufacture 41 (2001) 1815–1829 1829

[2] L. Kurt, Handbook of metal forming, McGraw-Hill, New York, 1985.


[3] ABAQUS/Standard-HKS, theory manual—Version 5.8.
[4] S. Kalpakjian, Manufacturing processes for engineering materials, 2nd ed., Addison-Wesley, Reading, MA, 1991.
[5] M.R. Jensen, F.F. Damborg, K.B. Nielsen, J. Danckert, Applying the finite element method for determination of
tool wear in conventional deep-drawing, J. Mater. Process. Technol. 83 (1998) 98–105.
[6] W.T. Carter, A model for friction in metal forming, J. Eng. Mater. Tech. 113 (1994) 8–13.
[7] J.F. Archard, Contact and rubbing of flat surfaces, J. Appl. Phys. 24 (1953) 981–988.
[8] C.M. Choy, R. Balendra, Experimental analysis of parameters influencing sheared-edge profiles, in: The 5th Int.
Conf. on Sheet Metal, University of Twente, Netherland, 1996, p. 101–10.
[9] R. Hambli, A. Potiron , Finite element modeling of sheet-metal blanking operations with experimental verification,
J Mater Process Technol (2000) 257–265.
[10] R. Hambli, Numerical fracture prediction during sheet-metal blanking processes, Eng. Fract. Mech. 68 (2000)
365–378.
[11] R. Hambli, A. Potiron, Finite element analysis of sheet-metal punching processes, in: The 7th Int. Conf. on Sheet
Metal, Erlangen, Germany, 1999, p. 153–9.
[12] J. Lemaitre, A continuous damage mechanics model for ductile fracture, J. Eng. Mater. Technol. 107 (1985) 83–89.
[13] Y. Kasuga, S. Tsutsumi, T. Mori, Investigation into shearing process of ductile sheet metals, Mem Fac Eng
Nagoya Univ Japan (1979) 1–46.

You might also like