You are on page 1of 40

T.

zel
1







INFLUENCE OF FRICTION MODELS ON
FINITE ELEMENT SIMULATIONS OF MACHINING




Tugrul zel
Assistant Professor
Department of Industrial and Systems Engineering
Rutgers, The State University of New Jersey, Piscataway, New Jersey 08854 USA
Tel: +1-732-445-1099; fax: +1-732-445-5467

E-mail address: ozel@rci.rutgers.edu
T. zel
2

Abstract
In the analysis of orthogonal cutting process using FE simulations, predictions are greatly
influenced by two major factors; a) flow stress characteristics of work material at cutting
regimes and b) friction characteristics mainly at the chip-tool interface. The influence of
work material flow stress upon FE simulations may be less or even none when there is a
constitutive model for work material that is obtained empirically from high-strain rate
and temperature deformation tests. However, the difficulty arises when one needs to
implement accurate friction models for cutting simulations using a particular FE
formulation. In this study, a thermo-mechanical updated Lagrangian finite element
formulation is used to simulate continuous chip formation process in orthogonal cutting
of low carbon free-cutting steel. The effects of using various chip-tool interface friction
models on the simulations are investigated. Experimentally measured stress distributions
on the tool rake face are utilized in conjunction with metal cutting theory to develop
several friction models and the evaluation of the results for the friction models is carried
out. The results depict that the use of various chip-tool interface friction models has at
most influence in predicting chip geometry, forces, stresses on the tool and temperature at
the chip-tool interface and the predicts are best when using variable shear friction model
at the chip-tool interface. Predictions presented in this work also justify that the FE
simulation technique used for orthogonal cutting process is an accurate and viable
analysis as long as flow stress behavior of the work material obtained realistically and
friction at the chip-tool interface is modeled correctly.


T. zel
3
1. INTRODUCTION
In metal cutting, severe deformations take place in the vicinity of the cutting edge of
the tool high strain-rates and temperatures are observed. Work material deformation
behavior in primary and secondary zones is highly sensitive to the cutting conditions
(Fig.1). The frictional conditions between the tool and the chip and tool and the work
piece are highly complex and also sensitive to the cutting conditions. As a result, the
stresses and temperatures at tool-chip interface and around the cutting edge can be
critically high in some cutting conditions and can cause excessive tool wear or even
premature tool failure. Therefore, accurate prediction of the distributions of the process
variables such as stresses and temperatures with Finite Element (FE) simulations are vital
to identify optimum cutting conditions, tool material, edge geometry and coating in order
to help improve productivity and quality of machining operations.
However, effects of work material flow stress constitutive models and interfacial
friction models at chip-tool interface on the accuracy of the predicted process variables in
FE simulations have not been adequately understood. Specifically, application of sound
friction models that are based on process variables such as pressure, temperature or
surface asperity etc. should be implemented and their influence on predicted process
variables should be assessed.
The use of FE simulations has become increasingly popular due to the advancement
in computers and the development of complex codes. In FE modeling, there are two types
of analysis in which a continuous medium can be described: Eulerian and Lagrangian. In
a Lagrangain analysis, the computational grid deforms with the material where as in a
Eulerian analysis it is fixed in space. The Lagrangian calculation embeds a computational
mesh in the material domain and solves for the position of the mesh at discrete points in
T. zel
4
time. In those analyses, two distinct methods, the implicit and explicit time integration
techniques can be utilized. The implicit technique is more applicable to solving linear
static problems while explicit method is more suitable for nonlinear dynamic problems. A
dynamic explicit FE formulation is very efficient for simulating highly non-linear
problems involving large localized deformations and changing contact conditions as
those experienced in machining.

2. LITERATURE REVIEW
A review of the technical literature reveals that currently FE modeling of cutting is
not fully capable of simulating practical machining operations. In other words, the
complexity and the diversity of cutting processes are such that a single process model or
simulation cannot be applied to all materials and cutting conditions. In the past decade,
continuous improvements in modeling orthogonal metal cutting have been developed by
a number of researchers. Mackerle [22,23] presented a bibliography for those efforts.
Some of the models used to model orthogonal metal cutting have used Eulerian
formulation. However, a vast majority has relied on Lagrangian formulation, which
allows the chip to be modeled from incipient to steady state. However, using Lagrangian
formulation requires a criterion for separation of the undeformed chip from the work
piece. As a result, the development of a realistic separation criterion is an important
factor in FE modeling of cutting. Black and Huang [4] presented a detailed evaluation of
the chip separation criteria.
Klamecki [16] developed one of the first FE models of cutting. This model of the
incipient chip formation was a three-dimensional model and strain-rate effect and friction
were not included. Later, this model was improved by applying thermo-elastic-plastic
T. zel
5
deformation. However, temperature softening of work material assumed overcoming the
strain-rate effects and the latter was neglected. Tay et al. [43] used FE simulations to
predict temperatures in steady state orthogonal machining. Usui and Shirakashi [45]
performed rigid-plastic FE simulations of steady state orthogonal cutting process. Their
model was able to solve the simultaneous equations of stress equilibrium, stress-strain
increment relationship, heat transfer, material flow stress and frictional stress
characteristics in the chip-tool interface. They also developed a friction model including
both sticking and sliding actions. Strenkowski and Carroll [39] developed a FE
simulation model based on the Lagrangian formulation that modeled the chip formation
from incipient to steady state cutting. Strenkowski and Carroll [40] also employed a
Eulerian formulation to model steady-state orthogonal metal cutting and compared it to
the Lagrangian model. Strenkowski and Moon [41] developed another Eulerian model.
Shih et al. [37] developed a model, which included the effects of sticking and sliding
friction, strain-rate, and heat generation. This model utilized a separation criterion based
on the distance between the tool tip and the nodes of the work piece being cut. Shih [38]
also developed an updated Lagrangian 2-D model including the effects of elasto-
viscoplasticity, material flow stress, strain hardening and friction. The separation criteria
for the elements in front of the tool tip created a numerical stability problem. That was
the drawback of this FEM model. Komvopolus and Erpenback [18] used an updated
Lagrangian FEM model and utilized a chip separation approach based on effective strain
criteria. Lin and Lin [19] also developed a thermo-elastic-plastic FEM model based on
updated Lagrangian method and used strain energy density as a chip separation criterion.
Zhang and Bagchi [47] also used Lagrangian method. Marusich and Ortiz [25] developed
T. zel
6
a dynamic explicit code to solve constitutive equation of the cutting process with
adaptive remeshing and simulated continuous and segmented chip formation.
Rakotomalala et al. [33] developed an Arbitrary Lagrangian Eulerian (ALE) formulation
to simulate orthogonal cutting to avoid frequent remeshing for chip separation. Pantale et
al. [32] also used ALE formulation to simulate 3-D cutting process. Sekhon and Chenot
[36] used an updated Lagrangian implicit formulation with automatic remeshing for chip
separation. Ueda and Manabe [44] used a numerical approach and simulated the chip
formation in the cutting process. Sasahara et al. simulated 3-D oblique cutting [34],
discontinuous chip formation, and machined layer residual stresses [35]. Ceretti et al. [6]
used updated Lagrangian implicit large plastic deformation FEM software to simulate
orthogonal cutting. Furthermore, they developed a module that is being incorporated into
this code to predict segmented chip formation by deleting the elements that reach a
"critical damage value" based on damage that cause fracture, [7, 17]. Lovell et al. [21]
introduced an explicit dynamic FEM method to simulate orthogonal cutting with sharp
edge but a restricted contact cutting tool. Olovsson et al. [29] used ALE formulation to
simulate cutting but using a sharp edge tool. Movahhedy et al. used ALE formulation
exclusively to simulate cutting with blunt edge tools [26,27].
zel and Altan [30,31] developed flow stress and friction models and simulated chip
flow in orthogonal cutting and milling with a round edge tool. Baker et al. [2] simulated
adiabatic shearing and serrated chip formation when cutting with a sharp edge cutting
tool using Lagrangian analysis with remeshing and by employing a damage criteria. Ng
and Aspinwall [28] used dynamic explicit Lagrangian analysis and Johnson-Cook
constitutive model and shear failure criteria [15] in conjunction with removal of damaged
T. zel
7
elements to simulate serrated chip formation using a sharp cutting edge tool. Chuzhoy et
al. [10] developed an explicit Lagrangian model with remeshing capability, utilized a
similar damage criteria for adiabatic shearing and simulated serrated chip formation and
deformations at microstructre-level when orthogonal cutting with a round edge tool.
Until late 1990s, vast majority of researchers used their own FEM code, however, the
use of commercially available software packages has increased dramatically over the last
fifteen years. These packages include DEFORM used by Ceretti et al. [6,7], zel and
Altan [30,31], Klocke et al. [17], ANSYS /LS-DYNA used by Lovell et al. [21],
ABAQUS/Explicit used by Liu and Guo [22]; Guo and Liu [13,14]; Ng and Aspinwall
[28]; Baker et al. [2]; Chuzhoy et al. [10,11]; Adibi-Sedeh and Madhavan, [1]; Zeren and
zel [46]. In many of these studies, the numerical codes developed were practical and
available commercially for end users.

3. FINITE ELEMENT MODELING OF CUTTING
There has been considerable amount of research devoted to develop analytical,
mechanistic and numerical models in order to simulate metal cutting processes.
Especially numerical models are highly essential in predicting chip formation, computing
forces, distributions of strain, strain rate, temperatures and stresses on the cutting edge
and the chip and the machined work surface [30]. The premise of the numerical models is
to be able to lead further predictions in machinability, tool wear, tool failure, quality and
integrity of the machined surfaces.
In all of the analysis of orthogonal cutting process using FE simulations, predictions
are greatly influenced by two major factors; a) flow stress characteristics of work
material at cutting regimes and b) friction characteristics at the chip-tool interface. The
T. zel
8
influence of work material flow stress upon FE simulations may be less or even none
when there is a constitutive model for work material that is obtained empirically from
high-strain rate and temperature deformation tests. However, the difficulty arises when
one needs to implement accurate friction and contact conditions for cutting simulations
using a particular FE formulation.
In this study, FEM software DEFORM-2D, which is an updated Lagrangian that
employs implicit integration method designed for large deformation simulations, is used
to simulate the cutting process. The workpiece is modeled as rigid-plastic due to large
elasto-plastic deformations. The strength of the FEM software is its ability to
automatically re-mesh and generate a very dense grid of nodes near the tool tip so that
large gradients of strain, strain-rate and temperature can be handled. In this approach,
there is no need for a chip separation criterion, making it highly effective in simulating
metal cutting process. Furthermore, high mesh density is defined around the cutting edge
as moving windows allowing excessively distorted mesh in the primary and secondary
zones to be automatically re-meshed without interruption. Thus, formation of the
continuous chip is simulated step by step per tool advance with minimum number of re-
meshing. In addition, this software has the capability to model the tool as elastic object so
that stress distributions in the tool can also be predicted. The flow diagram of the process
simulations is shown in Figure 2.
In this study, FE simulation for orthogonal cutting of low-carbon-free-cutting steel
(LCFCS) with P20 carbide cutting tool is designed by modeling workpiece as rigid-
plastic and tool as elastic bodies. Nodal velocities at the bottom of the workpiece model
are set to zero as boundary conditions and tool is moved from right to left due to the
T. zel
9
cutting speed. Work material flow stress behavior and contact friction conditions are
predefined and input to the simulation model accordingly. Cutting conditions, physical
and thermo-mechanical properties used for the work and tool materials are given in Table
1.
A constitutive flow stress model for LCFCS is used that is empirically determined as
a function of strain, strain-rate and temperature from the tests conducted at temperature of
700C and at strain-rate of 103 1/s in a Hopkinson bar apparatus [24]. The constitutive
flow stress model for LCFCS is given with Eq. 1.
N
M
A

|
|
.
|

\
|
=
1000

(1)
where,
2 2
-0.0011 -0.00004( -280) -0.0001( -600)
=910 e +120 e +50 e
T T T
A
=0.018 + 0.000038 M T
2
-0.0017 -0.00003( -370)
=0.16 e +0.09e
T T
N
Flow stress curves, shown in Fig. 2, are obtained using the constitutive model for
LCFCS and used in the FE simulations of orthogonal cutting.

4 FRICTION CHARACTERISTICS AT CHIP-TOOL INTERFACE

Friction in metal cutting is difficult to determine. The contact regions and the friction
parameters between the chip and the tool are influenced by factors such as cutting speed,
feed rate, rake angle, etc. [8], mainly because of the very high normal pressure at the
surface. The prevalent conditions at the tool-chip interface constrain the use of the
empirical values of the coefficient of friction found from ordinary sliding test conditions.
T. zel
10
While numerous works have been reported to quantitatively explain the variable and high
values of friction at the rake face, none has provided reliable quantitative predictive
models that are devoid of experimental testing.
In conventional machining at low cutting speeds, the friction mechanism is mostly
effective at the tool flank face. However, in high speed machining, a tremendous increase
in the chip velocity, the chip-tool friction contact and the temperatures are encountered at
the tool rake face. As a result, the increasing sliding velocity and frictional stress cause
significant wear on the tool rake face. Therefore, the rate of the tool wear heavily
depends on the frictional conditions in high speed machining.
Friction conditions at the chip-tool interface in early FEM models of metal cutting
have been largely ignored [16, 43] or assumed to be constant with a coefficient of friction
based on Coulombs law at the entire chip-tool interface [5]. Detailed friction modeling at
the chip-tool interface was pioneered by Usui and Shirakashi [45]. They modeled the
tool-chip interaction with a frictional force as a function of normal force as boundary
conditions depicted by Eq.2.
1
n
primary
k
f primary
k e

| |
|
|
\ .
(
(
=
(
(

(2)
where
f
and
n
are the frictional and normal stresses respectively, k
primary
is the shear
flow stress of the chip at the primary zone and was the coefficient of friction obtained
from experiments for different workpiece-tool material combination.
Shih et al. [37] introduced a model that consists of the sticking region for which the
friction force is constant, and the sliding region for which the friction force varies linearly
according to Coulombs law. The location of the separation of these two regions was
T. zel
11
estimated from examining the marks left on worn tool rake face. The coefficient of
friction was hard to predict in the sliding region and was thus chosen arbitrarily.
Much later, Lovell and his coworkers [3, 21,42] made significant contributions in this
field by emphasizing the use of variable chip-tool interfacial friction and exploring better
friction models based on asperity deformation and seizure.
All of the above works for friction models show that currently there is no analytical
model for friction conditions at the chip-tool interface, which is the primary heat source
for the tool temperatures. The chip usually absorbs heat generated in the primary
deformation zone. Although heated chip conducts some of the heat into the tool with
respect to the thermal conductivity property of the tool material. The major amount of
heat still remained at the chip-tool interface. Thus, the mechanical interaction between
chip and tool rake, and machined surface with tool flank creates heat sources that directly
affect tool temperatures and therefore tool wear.
In this study, influence of implementing different friction models on predictions of
FE simulation was investigated by comparing predicted process variables to experimental
results. Measured cutting forces, temperatures, stress distributions on the tool rake face,
chip-tool contact length and shear angle in orthogonal cutting process were obtained from
literature [8].
Friction in the workpiece-tool contact is modeled using shear friction, m, and
Coulomb friction, , relations via previously measured chip-tool contact conditions.
Initially, experimental data obtained under the cutting conditions of cutting speed,
V
c
=150 m/min, cut with, w=1 mm, and uncut chip thickness, t
u
=0.1 mm from orthogonal
end turning experiments, were used. For those conditions, measured cutting force,
T. zel
12
F
c
=174 N, thrust force, F
t
=83 N, and shear angle of =18.8 were reported by Childs et
al. [8]. The shear friction factor, m, as defined in Eq. 4 needs to be estimated as an input
into the FE simulations.
3
primary
m

=
or
primary
m
k

=
(4)

Maximum shear yield stress on the shear plane can be approximated from Eq. 3 as

s
=440 MPa according to DeVries [12].


s
c t
u
F F
t w
=
cos sin sin
2
(5)
This can also be considered as local shear flow stress in the primary deformation zone
k
primary
= 440 MPa. From the measured data maximum shear stress on the rake face was
given as 360 MPa [8]. Thus, from Eq. 4 m=0.82 was estimated.
By employing Zorevs stress distribution model [48] shown in Fig. 4, and using
measured normal and shear stress distributions over the rake face given for various
cutting speeds [8] as shown in Fig. 5 following friction models for chip-tool contact have
been formed:

4.1. Constant shear friction at the entire chip-tool interface (I)
A constant shear friction factor is applied by considering m=0.82 over the entire chip-tool
contact.

4.2 Constant shear friction in sticking region and Coulomb friction in sliding region
(II)
T. zel
13
A friction model with combination of Coulomb and shear friction is applied by defining
two distinct friction regions in the chip-tool contact where m=0.82 is applied over the
sticking region (0<x< l
p
=0.1mm) while =1.0 over the sliding region (l
p
=0.1<x< l
c
=0.6
mm).
4.3 Variable shear friction at the entire chip-tool interface (III)
Only a variable shear friction is considered as a function of normal surface pressure along
the entire chip-tool contact. Shear friction factor is calculated, as shown in the Fig. 6,
from following empirical friction model proposed by Childs et al. [8].

m
k
p
primary
n
p
=
|
\

|
.
|
|

(
(

1
1 7
1 1 7
exp
.
/ .
(6)
where k
primary
=440 MPa and
p
=360 MPa.
4.4 Variable Coulomb friction at the entire chip-tool interface (IV)
A variable Coulomb friction coefficient is used by considering as a function of normal
surface pressure along the entire tool-chip interface. Coulomb friction coefficient is
calculated as shown in Fig. 6 from the relation, =/
n
by using the experimental stress
distribution data for and
n
given in Fig. 5.

4.5 Variable shear friction in sticking region and variable Coulomb friction in
sliding region (V)
A friction model based on the combination of variable shear friction and variable
Coulomb friction is used by creating two distinct friction regions on tool rake face in the
process simulation, where m and were defined as functions of normal stress at the chip-
T. zel
14
tool interface (see Fig. 6) over the sticking region (0<x<=l
p
=0.1mm) and the sliding
region (l
p
=0.1<x<=l
c
=0.6 mm), respectively.
Evaluations of these friction models are carried out under the same cutting
condition and tool geometry in order to identify the best friction model for FE
simulations in order to predict process variables accurately in orthogonal cutting. Cutting
conditions of cutting speed, 150 m/min and uncut chip thickness of 0.1 mm were used.
The predicted process variables from the process simulations using five friction models
(Models I-V) are presented in Table 2. Comparison with experimental results depicts that
cutting force, F
c
and chip-tool contact length, l
c
are in fair agreements while other
predicted process variables exhibit good agreements with experimental results. It should
be noted that predicted process variables are closest to the experimental ones when using
variable friction models at the entire chip-tool interface (III, IV). Thus, it is concluded
that the variable friction models should be used in order to obtain more accurate results in
FE simulations of orthogonal cutting.

5. PREDICTION OF CUTTING FORCES, STRESSES AND TEMPERATURES
After testing the friction models, process variables for additional cutting speeds
(50 and 250 m/min) are predicted using only variable friction models. Similarly to
previous modeling done for the case with cutting speed of 150 m/min, measured normal
and frictional stress distributions are used in order to determine shear friction factor, m,
and friction coefficient, , as functions of normal stress at the chip-tool interface for
cases with cutting speed of 50 and 250 m/min.
Predicted process variables; cutting and thrust forces, maximum tool temperature,
chip-tool contact length and shear angle are compared with measured values as indicated
T. zel
15
in Fig. 7 and in Table 3. Predicted thrust forces match with the measured forces where as
predicted cutting forces slightly higher than the measured forces. It is believed that this is
due to the flow stress model of the work material. The temperature softening affect is not
sufficient enough to reflect the lower flow stress at higher temperatures generated at the
secondary zone.
Predicted normal and frictional stress distributions on tool rake face using
variable shear friction model at cutting speeds of 50, 150 and 250 m/min are compared
with measured stress distributions obtained with split-tool experiments by Childs et al.
[8]. These comparisons indicate that predicted normal and frictional stress on the tool
rake face are identically agree with the experimentally measured values for all three
cutting conditions as shown in Figs. 8.
In addition, all of the stress distributions can be predicted in the tool since the tool
is defined as elastic body. An example of effective stress distribution using variable shear
friction model at cutting speed of 150 m/min is given in Fig. 9. Temperature distribution
at the same cutting condition is given in Fig. 10. In addition, the temperature distributions
are also obtained from FE simulations for all three cases (50, 150 and 250 m/min) as
shown in Figs. 11-13.
Therefore, once a best friction model determined for the chip-tool interface,
accurate predictions of stress and temperature distributions can be obtained and
furthermore simulation experiments can be designed to identify optimum tool edge
geometry, cutting condition and tool material/coatings.

T. zel
16
6. CONCLUSIONS
In this study, a thermo-mechanical updated Lagrangian finite element formulation is
used to simulate continuous chip formation process in orthogonal cutting of low carbon
free-cutting steel. The effects of chip-tool interface friction modeling on the simulations
are investigated by means of constant and variable shear and Coulomb friction models.
Experimentally measured stress distributions on the tool rake face are utilized in
conjunction with metal cutting theory to implement five different friction models and the
evaluation of the results for the friction models is also carried out.
FE simulation of orthogonal cutting process at three different cutting is also
investigated and predicted process variables such as shear angle, forces, temperatures at
the chip-tool interface and stress distributions on the tool rake face are compared with
experimental results. Assessment of the friction models is performed using FE
simulations as reverse engineering tool in order to uncover the best-fit friction model. By
assuming that the measurements for process variables are accurate. The results depict that
the use of various chip-tool interface friction models has at most influence in predicting
chip geometry, forces, stresses on the tool and temperature at the chip-tool interface and
the predicts are best when using variable shear friction model at the chip-tool interface.
The following conclusions can be made from this study:
Comparisons indicated that predictions using variable shear friction and
variable Coulomb friction models on the tool rake face are much better than
friction models based on constant coefficient of friction or shear friction.
Thrust forces are predicted with less than 10% error although cutting force
predictions indicated error at about 30% error.
T. zel
17
Maximum tool temperature is predicted with less than 5% error for all cutting
conditions.
Chip-tool contact length is predicted from the resultant chip geometry in the
FE simulations and predictions depicted 20-30% error.
Predictions presented in this work also justify that the FE simulation technique used for
orthogonal cutting process is an accurate and viable analysis as long as flow stress
behavior of the work material obtained realistically and friction at the chip-tool interface
is modeled correctly.

ACKNOWLEDGMENTS
The author would like to thank Professor T.H.C. Childs for providing experimental
results and Y.C. Yen and Professor T. Altan for valuable discussions.


REFERENCES

[1] Adibi-Sedeh, A.H., and Madhavan, V., 2003, Understanding of finite element
analysis results under the framework of Oxleys machining model, 6th CIRP
International Workshop on Modeling of Machining Operations, McMaster
University, Hamilton, Canada.
[2] Baker, M., Rosler, J., and Siemers, C., 2002, A finite element model of high speed
metal cutting with adiabatic shearing, Computers & Structures, 80, 495-513.
[3] Bhattacharya, S., and Lovell, M. R., 2000, Characterization of friction in
machining: evaluation of asperity deformation and seizure-based models,
Transactions of NAMRI/SME, XXVII, 107-112.
T. zel
18
[4] Black, J. T., and Huang, J. M., 1996, An evaluation of chip separation criteria for
the fem simulation of machining, Journal of Manufacturing Science and
Engineering, 118, 545 553.
[5] Carroll, J.T., and Strenkowski, J., 1988, Finite element models of orthogonal
cutting with application to single point diamond turning, International Journal of
Mechanical Sciences, 30(12), 889-920.
[6] Ceretti, E., Fallbhmer, P., Wu, W.T., and Altan, T., 1996, Application of 2-D
FEM to chip formation in orthogonal cutting, Journal of Materials Processing
Technology, 59, 169-181.
[7] Ceretti E., Lucchi, M., and Altan, T., 1999, FEM simulation orthogonal cutting:
serrated chip formation Journal of Materials Processing Technology, 95, 17-26.
[8] Childs, T.H.C., Dirikolu, M.H., Sammons, M.D.S., Maekawa, K., and Kitagawa,
T., 1997, Experiments on and Finite Element Modeling of turning free-cutting
steels at cutting speeds up to 250 m/min, Proceedings of 1st French and German
Conference on High Speed Machining, pp. 325-331.
[9] Childs, T.H.C., 1998, Material property needs in modeling metal machining,
Proceedings of the CIRP International Workshop on Modeling of Machining
Operations, Atlanta, GA, 193-202.
[10] Chuzhoy, L., DeVor, R.E., Kapoor, S.G., and Bammann, D.J., 2002,
Microstructure-level modeling of ductile iron machining, ASME Journal of
Manufacturing Science and Engineering, 124, 162-169.
[11] Chuzhoy, L., DeVor, R.E., and Kapoor, S.G., 2003, Machining Simulation of
Ductile Iron and Its Constituents. Part 2: Numerical Simulation and Experimental
T. zel
19
Validation of Machining, ASME Journal of Manufacturing Science and
Engineering, 125, 192-201.
[12] DeVries, W.R., 1992, Analysis of Material Removal Processes, Springer-Verlag
Publications, New York.
[13] Guo, Y. B., and Liu, C. R., 2002, Mechanical properties of hardened AISI 52100
steel in hard machining processes, ASME Journal of Manufacturing Science and
Engineering, 124, 1-9.
[14] Guo Y.B., and Liu, C.R., 2002, 3D FEA modeling of hard turning, Journal of
Manufacturing Science and Engineering, 124, 189-199.
[15] Johnson, G.R. and W.H. Cook, 1983, A constitutive model and data for metals
subjected to large strains, high strain rates and high temperatures, Proceedings of
the 7th International Symposium on Ballistics, The Hague, The Netherlands, 541-
547.
[16] Klamecki, B.E., 1973, "Incipient chip formation in metal cutting-a three dimension
finite element analysis", Ph.D. Dissertation, University of Illinois at Urbana-
Champaign.
[17] Klocke F., Raedt, H.-W., and Hoppe, S., 2001, 2D-FEM simulation of the
orthogonal high speed cutting process, Machining Science and Technology, 5/3,
323-340.
[18] Komvopoulos K., and Erpenbeck, S.A., 1991, Finite element modeling of
orthogonal metal cutting, ASME Journal of Engineering for Industry, 113, 253-
267.
T. zel
20
[19] Lin, Z. C. and Lin, S. Y., 1992, A couple finite element model of thermo-elastic-
plastic large deformation for orthogonal cutting, ASME Journal of Engineering for
Industry, 114, 218-226.
[20] Liu, C.R. and Guo, Y.B., 2000, Finite element analysis of the effect of sequential
cuts and tool-chip friction on residual stresses in a machined layer, Int. J. Mech.
Sci., 42, 10691086.
[21] Lovell, M. R., Bhattacharya, S., and Zeng, R., 1998, Modeling orthogonal
machining process for variable tool-chip interfacial friction using explicit dynamic
finite element methods, Proceedings of the CIRP International Workshop on
Modeling of Machining Operations, 265-276.
[22] Mackerle, J., 1999, Finite element analysis and simulation of machining: a
bibliography (1976-1996), J. Mater. Process. Technol. 86, 17-44.
[23] Mackerle, J., 2003, Finite element analysis and simulation of machining: an
addendum a bibliography (1996-2002), J. Mater. Process. Technol. 43, 103-114.
[24] Maekawa, K., Kitagawa, T. and Childs, T. H. C., 1991, Effects of flow stress and
friction characteristics on the machinability of free cutting steels, Second
International Conference on the Behavior of Materials in Machining, 132-145.
[25] Marusich, T.D., and Ortiz, M., 1995, Modeling and Simulation of High-speed
Machining, International Journal for Numerical Methods in Engineering, 38, 3675-
3694.
[26] Movahhedy, M. R., Gadala, M. S., and Altintas, Y., 2000, FE Modeling of Chip
Formation in Orthogonal Metal Cutting Process: An ALE Approach, Mach. Sci.
Technol., 4, 1547.
T. zel
21
[27] Movahhedy M.R., Y. Altintas, and M. S. Gadala, 2002, Numerical analysis of
metal cutting with chamfered and blunt tools, Journal of Manufacturing Science
and Engineering, 124, 178-188.
[28] Ng E., and Aspinwall, D.K., 2002, Modeling of hard part machining, Journal of
Materials Processing Technology, 124, 1-8.
[29] Olovsson, L., Nilsson, L., and Simonsson, K., 1999, An ALE Formulation for the
Solution of Two-Dimensional Metal Cutting Problems, Computers and
Structures, 72, 497507.
[30] zel T., and Altan, T., 2000, Process simulation using finite element method-
prediction of cutting forces, tool stresses and temperatures in high-speed flat end
milling process, International Journal of Machine Tools and Manufacture, 40/5,
713-738.
[31] zel, T., and Altan, T, 2000, Determination of workpiece flow stress and friction
at the chip-tool contact for high-speed cutting, International Journal of Machine
Tools and Manufacture, 40/1, 133-152.
[32] Pantale, O., Rakotomalala, R., Touratier, M., and Hakem, N., 1996, A Three-
Dimensional Numerical Model of Orthogonal And Oblique Metal Cutting
Processes, Proceeding of Engineering Systems Design and Analysis, ASME, 3,
199-206.
[33] Rakotomalala, R., Joyot, P., and Touratier, M., 1993, Arbitrary Lagrangian-
Eulerian Thermomechanical Finite Element Model of Material Cutting,
Communications in Numerical Methods in Engineering, 9, 975-987.
T. zel
22
[34] Sasahara, H., Obikawa, T., and Shirakashi, T., 1994, "FEM Analysis on Three
Dimensional Cutting", International Journal of Japanese Society for Precision
Engineering, Vol.28, No.2, 473-478.
[35] Sasahara, H., Obikawa, T., and Shirakashi, T., 1994, "The Prediction of Effects of
Cutting Condition on Mechanical Characteristics in Machined Layer",
Advancement of Intelligent Production, Japanese Society for Precision
Engineering, Elsevier, pp. 473-478.
[36] Sekhon, G.S., and Chenot, J.L., 1992, Some Simulation Experiments in
Orthogonal Cutting, Numerical Methods in Industrial Forming Processes, 901-
906.
[37] Shih, A.J., Chandrasekar, S., Yang, H.T., 1990, Finite Element Simulation of
Metal Cutting Process with Strain-Rate and Temperatures Effects, Fundamental
Issues in Machining, ASME, PED, 43, 11-24
[38] Shih, A. J., 1995, Finite element simulation of orthogonal metal cutting, ASME
Journal of Engineering for Industry, 117, 84-93.
[39] Strenkowski, J.S., and Carroll, J.T., 1985, A finite element model of orthogonal
metal cutting, ASME Journal of Engineering for Industry, 1985, 107, 346-354.
[40] Strenkowski, J.S., and Carroll, J.T., 1986, Finite element models of orthogonal
cutting with application to single point diamond turning, International Journal of
Mechanical Science, 30, 899-920.
[41] Strenkowski, J.S., Moon, K.J., 1990, Finite element prediction of chip geometry
and tool/workpiece temperature distributions in orthogonal metal cutting, ASME
Journal of Engineering for Industry, 112, 313-318.
T. zel
23
[42] Tao, Z., and Lovell, M R., 2002, Towards an improved friction model in material
removal processes: Investigating the role of temperature Technical Paper SME,
MR02-188, 1-8.
[43] Tay, A.O., Stevenson, M.G., de Vahl Davis, G., 1974, Using the finite element
method to determine temperature distributions in orthogonal machining,
Proceedings of Institution for Mechanical Engineers, 188, 627-638.
[44] Ueda, K. and Manabe, K., 1993 Rigid-plastic FEM analysis of three-dimensional
deformations field in the chip formation process, Annals of the CIRP, 1993, 42,
35-38.
[45] Usui, E., and Shirakashi, T., 1982, Mechanics of machining -from descriptive to
predictive theory. In on the art of cutting metals-75 years later, ASME Publication
PED 7, 13-35.
[46] Zeren, E., and zel, T., 2003, Explicit dynamics finite element simulations of
orthogonal cutting with blunt tools: comparison of Lagrangian and Arbitrary
Eulerian Lagrangian methods, Report MARL-03-05, Rutgers University,
Piscataway, New Jersey, USA.
[47] Zhang, B. and Bagchi, A., 1994, Finite element formation of chip formation and
comparison with machining experiment, ASME Journal of Engineering for
Industry, Vol. 116, pp. 289-297.
[48] Zorev N.N., 1963, Inter-relationship between shear processes occurring along tool
face and shear plane in metal cutting, International Research in Production
Engineering, ASME, New York, 42-49.


T. zel
24
LIST OF TABLE CAPTIONS

Table1: List of parameters used in the FE simulations
Table 2: Comparison of predicted performance measures with experimental results
Table 3. Comparison of predicted process variables with experiments



LIST OF FIGURE CAPTIONS

Fig. 1. Deformation zones in orthogonal cutting
Fig 2. Modeling of orthogonal cutting process using FEM based process simulation
Fig. 3. Flow stress curves determined for LCFCS
Fig. 4. Curves representing normal (
n
) and frictional () stress distributions on the tool
rake face [48]
Fig. 5. Measured normal and shear stress distribution on cutting tool rake face in
orthogonal cutting of LCFCS [8]
Fig. 6. Variable shear friction and Coulomb friction on tool rake face
Fig. 7. Predicted and measured cutting forces at uncut chip thickness of 0.1 mm
Fig. 8. Comparison of predicted stress distributions with experiments when using variable
shear friction at chip-tool interface
Fig. 10. Predicted effective stress distribution at the tool and the workpiece using variable
shear friction in orthogonal cutting at cutting speed of 150 m/min
Fig. 11. Predicted temperature distribution using variable shear friction in orthogonal
cutting at cutting speed of 150 m/min
Fig. 12. Comparison of predicted temperature distributions process simulations with
experiments at cutting speed of 50 m/min
Fig. 13. Comparison of predicted temperature distributions process simulations with
experiments at cutting speed of 250 m/min

T. zel
25
Table1: List of parameters used in the FE simulations

ORTHOGONAL CUTTING PARAMETERS
Cutting speed, V
c
[m/min] 50,150,250
Undeformed chip thicknes, t
u
[mm] 0.1
Width of cut, w [mm] 1
Heat transfer coefficient between workpiece-tool
contact [W/m
2 o
C]
1.0x10
5

Fraction of deformation energy transformed into heat 0.9
Workpiece (Lcfcs) Properties
Thermal conductivity [W/(mC)] 62 -0.044 T
o
C
Heat capacity [N/mm/C]
Specific heat [J/kg/C]
3.5325+0.002983 T
o
C
450 + 0.38 T
o
C
Coefficient of thermal expansion [10
6
/C] 13.05 (at 20 to 200 C)
13.40 (at 20 to 300 C)
Density [g/cm
3
] 7.85
Poissons ratio 0.3
Youngs modulus [GPa] 200
Emissivity 0.295 (at 100 C)
0.335 (at 500 C)
0.635 (at 1000 C)
Tool (P20 Cemented Carbide) Properties
Youngs Modulus [GPa] 558
Poissons Ratio 0.22
Thermal conductivity [W/(mC)] 46
Heat capacity [N/mm/C]
Specific heat [J/kg/C]
2.7884
230.45
Coefficient of thermal expansion [m/(mC)] 5.8 at 200 C (390 F)
6.8 at 1000 C (1830 F)
Density [g/cm
3
] 12.1
Emissivity 0.75
Tool edge radius, [mm] 0.02
Rake angle [degree] 0
Clearance angle [degree] 5







T. zel
26
Table 2: Comparison of predicted performance measures with experimental results
Experimental Results, Childs et al. [8]
F
c
(N/mm) F
t
(N/mm) l
c
(mm) (degree) T
max
(
o
C)
174 83 0.6 18.8 590
Predicted Values from FE Simulations
Friction
Model
F
c
(N/mm)

F
t

(N/mm)

l
c

(mm)
(degree) T
max
(
o
C)

I 270 108 0.38 20.9 607
II 283 126 0.38 21.3 450
III 265 101 0.34 21.1 600
IV 272 115 0.47 18.4 620
V 297 140 0.51 17.8 489


T. zel
27

Table 3. Comparison of predicted process variables with experiments
F
c
(N/mm)

F
t
(N/mm)

T
max
(C)

l
c
(mm)

(degree)
V
c
=50 m/min, t
u
=0.1 mm
Measured 204 96 420 >0.60 12.6
Predicted (Model III) 267 108 399 0.38 14.4
Predicted (Model IV) 310 160 432 0.84 14.4
Vc=150 m/min, t
u
=0.1 mm
Measured 174 83 590 0.60 18.8
Predicted (Model III) 265 101 600 0.34 21.1
Predicted (Model IV) 272 115 620 0.47 18.4
Vc=250 m/min, t
u
=0.1 mm
Measured 190 125 780 >0.60 17.4
Predicted (Model III) 275 110 775 0.36 17.2
Predicted (Model IV) 268 101 830 0.36 17.3

T. zel
28
max strain
rate
34855 1/s
20000
18000
16000
14000
12000
10000
8000
6000
4000
2000
0
1/s
P
r
i
m
a
r
y

z
o
n
e
S
e
c
o
n
d
a
r
y

z
o
n
e

Fig. 1. Deformation zones in orthogonal cutting


T. zel
29
Flow Stress
of
Workpiece
Material
Tool Thermal
Conductivity
Cutting Forces
Frictional Stresses at
Tool Rake Face
Normal Stresses at
Tool Rake Face
Temperatures in
Deformation Zones

Chip-Tool
Contact Length, lc
Chip Thickness, tc
Shear Angle,
Orthogonal
Cutting Process
(Chip Flow)
Friction at
Chip-Tool
Interface
Cutting Conditions
* Cut Thickness, tu
* Cutting Speed, Vc
* Width of Cut, w
Strain-Rate
Tool Temperature
Distribution
Tool Stress
Distribution
Tool Geometry
* Rake Angle,
* Clearance Angle,
*Tool Edge Radius,
Heat Transfer
Between Chip and
Tool

Fig 2. Modeling of orthogonal cutting process using FEM based process simulation
T. zel
30
0 200 400 600 800 1000 1200
200
400
600
800
1000
1200
1400
Strain-rate=1000 1/s
Temperature (

C)
Stress
(MPa)

=10 mm/mm

=4 mm/mm

= 3 mm/mm

=2 mm/mm

=1 mm/mm

=0.1 mm/mm


Fig. 3. Flow stress curves determined for LCFCS




T. zel
31


Fig. 4. Curves representing normal (
n
) and frictional () stress distributions on the tool
rake face [48]

T. zel
32
0
200
400
600
800
1000
1200
0 0.1 0.2 0.3 0.4 0.5 0.6
distance from the cutting edge (mm)
S
t
r
e
s
s

(
M
P
a
)
nor.str. (Vc=50 m/min)
fric.str. (Vc=50 m/min)
nor.str. (Vc=150 m/min)
fric.str. (Vc=150 m/min)
nor.str. (Vc=250 m/min)
fric.str. (Vc=250 m/min)


Fig. 5. Measured normal and shear stress distribution on cutting tool rake face in
orthogonal cutting of LCFCS [8]























T. zel
33
0.00
0.20
0.40
0.60
0.80
1.00
1.20
0 200 400 600 800 1000
Normal stress on the tool rake face, MPa
S
h
e
a
r

f
r
i
c
t
i
o
n

f
a
c
t
o
r
,

m
0.00
0.20
0.40
0.60
0.80
1.00
1.20
friction coeffiecent
shear friction factor
F
r
i
c
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
,



Fig. 6. Variable shear friction and Coulomb friction on tool rake face
T. zel
34
0
50
100
150
200
250
300
350
400
50 100 150 200 250
Cutting Speed, Vc, (m/min)
F
o
r
c
e
s
,

(
N
)
Fc-predicted
Fc-measured
Ft-predicted
Ft-measured


Fig. 7. Predicted and measured cutting forces at uncut chip thickness of 0.1 mm

T. zel
35
0
500
1000
1500
2000
2500
3000
0 0.2 0.4 0.6 0.8 1
Distance from the cutting edge (mm)
(Vc=50 m/min,

=0, tu=0.1 mm)


S
t
r
e
s
s

(
M
P
a
)
predicted frictional
stress
predicted normal
stress
measured normal
stress
measured
frictional stress
0
500
1000
1500
2000
2500
3000
3500
0 0.1 0.2 0.3 0.4 0.5 0.6
Distance from the cutting edge (mm)
(Vc=150 m/min,

=0, tu=0.1 mm)


S
t
r
e
s
s

(
M
P
a
)
predicted frictional
stress
predicted normal
stress
measured normal
stress
measured
frictional stress
0
500
1000
1500
2000
2500
3000
3500
4000
0 0.1 0.2 0.3 0.4 0.5 0.6
Distance from the cutting edge (mm)
(Vc=250 m/min,

=0, tu=0.1 mm)


S
t
r
e
s
s

(
M
P
a
)
predicted frictional
stress
predicted normal
stress
measured normal
stress
measured
frictional stress

Fig. 8. Comparison of predicted stress distributions with experiments when using variable
shear friction at chip-tool interface
T. zel
36
1700
1530
1360
1190
1020
850
680
510
340
170
0
MPa
max effective
tool stress
1700 MPa


Fig. 9. Predicted effective stress distribution at the tool and the workpiece using variable
shear friction in orthogonal cutting at cutting speed of 150 m/min
T. zel
37
620
560
500
440
380
320
260
200
140
80
20
o
C
Max temp.
tool
601
o
C
workpiece
605
o
C


Fig. 10. Predicted temperature distribution using variable shear friction in orthogonal
cutting at cutting speed of 150 m/min


T. zel
38


Fig. 11. Comparison of predicted temperature distributions process simulations with
experiments at cutting speed of 50 m/min


Temp. (
0
C)

A = 0

B = 45

C = 90

D = 135

E = 180

F = 225

G = 270

H = 315

I = 360

J = 405

T. zel
39


Fig. 12. Comparison of predicted temperature distributions process simulations with
experiments at cutting speed of 150 m/min


Temp. (
0
C)

A = 0

B = 70

C = 140

D = 210

E = 280

F = 350

G = 420

H = 490

I = 560

J = 630

T. zel
40


Fig. 13. Comparison of predicted temperature distributions process simulations with
experiments at cutting speed of 250 m/min

Temp. (
0
C)

A = 0

B = 90

C = 180

D = 270

E = 360

F = 450

G = 540

H = 630

I = 720

J = 810

You might also like