You are on page 1of 15

Journal of Materials Processing Technology 105 (2000) 95109

Finite element analysis of the orthogonal metal cutting process


Chandrakanth Shet, Xiaomin Deng*
Department of Mechanical Engineering, University of South Carolina, Columbia, SC 29208, USA
Received 8 April 1999

Abstract
In this paper, the orthogonal metal cutting process is analyzed with the nite element method under plane strain conditions. Frictional
interaction along the toolchip interface is modeled with a modied Coulomb friction law, and chip separation is based on a critical stress
criterion and is simulated using a nodal release procedure. Finite element solutions of temperature, stress, strain, and strain rate elds have
been obtained for a range of tool rake angle and friction coefcient values. Results showing how the toolchip interfacial friction affects the
eld distributions are new and add to the existing knowledge base. This paper also reports the procedure and specic modeling techniques
for simulating the orthogonal metal cutting process using a general-purpose nite element computer code. The ndings of this paper
provide useful insights for understanding and for improving the orthogonal metal cutting process. # 2000 Elsevier Science S.A. All rights
reserved.
Keywords: Finite element simulation; Orthogonal metal cutting; Thermomechanical elds

1. Introduction
In the metal cutting process unwanted material is removed
from a workpiece in the form of chips for producing nished
parts of required dimensions and accuracy. Metal cutting is a
highly non-linear and coupled thermomechanical process,
where the mechanical work is converted into heat through
the plastic deformation involved during chip formation and
also due to frictional work between the tool, chip and
workpiece. During such thermomechanical work conversion, a temperature rise of up to 10008C has been reported in
the literature [1,2].
A thorough understanding of the material removal process
in metal cutting is essential in selecting the tool material and
in design, and also in assuring consistent dimensional
accuracy and surface integrity of the nished product.
The earliest analytical models explaining the mechanics
of metal cutting were proposed by Merchant [3,4], Piispanen
[5], and Lee and Shaffer [6]. These models are known as
shear angle models; they relate the chip shear angle to the
tool rake angle. Kudo [7] introduced curved shearing to
account for the controlled contact between the curved chip
and the straight tool face. These models assumed rigid
perfectly plastic material behavior.

*
Corresponding author. Tel.: 1-803-777-7144; fax: 1-803-777-0106.
E-mail address: deng@engr.sc.edu (X. Deng).

Analytical models including the effect of work hardening


and strain-rate effects were proposed by Palmer and Oxley
[8] and Oxley et al. [9]. Interfacial friction along the tool
chip interface was incorporated into these viscoplastic models by Doyle et al. [10]. The effect of heating in metal cutting
was included in an analytic model by Trigger and Chao [11].
Three-dimensional geometric conditions in metal cutting
were considered by Usui et al. [12] using an energy approach.
In recent years, the nite element method has become the
main tool for simulating metal cutting processes. Early nite
element studies included those by Usui and Shirakashi [13],
Iwata et al. [14], and Strenkowski and Carroll [15]. It seems
that Carroll and Strenkowski [16], Strenkowski and Moon
[17], and Tyan and Yang [18] were the rst to use Eulerian
formulations for steady-state metal cutting simulation. A
key component of Lagrangian metal cutting simulation
procedures is the use of a material separation criterion, such
as the ``distance tolerance'' criterion [19], the ``strain energy
density'' criterion [20], and a fracture-mechanics based
criterion [21]. The effectiveness of the various separation
criteria in orthogonal metal cutting simulation was evaluated
in Huang and Black [22]. They found that during steadystate cutting, the chip separation criterion does not greatly
affect the geometry of the chip and the distribution of stress
and strain. To deal with large element distortion in metal
cutting simulation, Shih and Yang [23] and Shih [24,25]
developed a mesh-rezoning technique to enhance computational efciency and accuracy.

0924-0136/00/$ see front matter # 2000 Elsevier Science S.A. All rights reserved.
PII: S 0 9 2 4 - 0 1 3 6 ( 0 0 ) 0 0 5 9 5 - 1

96

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109


Table 1
Temperature-dependent elastic properties
Young's
modulus E (GPa)

Poisson's
ratio v

Temperature
(8C)

207.0
200.0
190.0
105.0
70.0
50.0
30.0

0.3
0.3
0.3
0.3
0.3
0.3
0.3

20.0
100.0
150.0
200.0
250.0
300.0
350.0

Fig. 1. A schematic diagram of the orthogonal metal cutting process and


graphical descriptions of the terms used in this study.

conditions and for a range of friction coefcient and rake


angle values.

The objective of this work is to provide solutions of


temperature, stress, strain, and strain rate elds in orthogonal
metal cutting, with emphasis on the inuence of friction.
Detailed results of the effects of friction on the distributions
of the eld quantities have been examined for a range of rake
angle and friction coefcient values. These results are new
and complement existing ndings in the literature. It is noted
that such parametric evaluations and understanding are
difcult to achieve experimentally. This paper also communicates techniques for simulating orthogonal metal cutting
using the general-purpose nite element code ABAQUS. For
example, a stress-based chip separation criterion is proposed
to model the chip separation from the workpiece. According
to this criterion, chip separation occurs when a critical stress
state is achieved at a specied distance ahead of the tip of the
cutting tool. The frictional interaction between the chip and
the cutting tool is modeled with a modied Coulomb friction
law. Adiabatic heating conditions are assumed to account for
local temperature rise due to conversion of plastic work and
frictional work into heat. Temperature-dependent material
properties are employed in the analysis. Strain-rate effects
and large strains are also included in the analysis. The nite
element simulations are carried out under plane strain

2. Finite element modeling details


Because of the complex nature of metal cutting processes,
a complete nite element simulation procedure for metal
cutting simulation involves many component parts. In this
light, it is noted that while a user-developed custom nite
element code enables the user to modify the code according
to personal needs, its development often requires advanced
technical know-how in many areas, not to mention several
years of intensive programming and debugging. On the other
hand, many general-purpose commercial codes offer
advanced modeling and pre- and post-processing options
not available in custom nite element codes. Even though
commercial codes are not specialized for metal cutting
simulations, a careful integration of modeling options in
the codes and custom user subroutines will facilitate metal
cutting simulations using these codes. To this end, the
purpose of this section is to discuss a set of custom modeling
options in the commercial code ABAQUS that have been
successfully integrated by the authors to simulate the orthogonal metal cutting process. We begin the discussion of
these modeling options with a description of the geometry
and mesh of the model problem.

Fig. 2. The nite element mesh used in the orthogonal metal cutting simulations.

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109


Table 2
Temperature-dependent elasticplastic properties

97

Table 3
Temperature-dependent thermal expansion coefcient

Flow stress
s (MPa)

Plastic strain
ep (mm/mm)

Temperature
(8C)

Thermal expansion
coefficient a (mm/m K)

Temperature
(8C)

414.0
517.0
759.0
1100.0
409.0
512.0
754.0
1005.0
309.0
412.0
654.0
905.0
259.0
362.0
604.0
885.0
209.0
312.0
554.0
835.0
159.0
262.0
504.0
785.0

0.00
0.01
0.09
0.90
0.00
0.01
0.09
0.90
0.00
0.01
0.09
0.90
0.00
0.01
0.09
0.90
0.00
0.01
0.09
0.90
0.00
0.01
0.09
0.90

20.0
20.0
20.0
20.0
100.0
100.0
100.0
100.0
150.0
150.0
150.0
150.0
200.0
200.0
200.0
200.0
250.0
250.0
250.0
250.0
300.0
300.0
300.0
300.0

12.3
12.7
13.7
14.5

20.0
200.0
400.0
600.0

2.1. Problem geometry and nite element mesh


A schematic diagram of the model problem for the
orthogonal metal cutting process is shown in Fig. 1, where
a rigid cutting tool moves forward to the left with a constant velocity and cuts through a softer workpiece. To
facilitate plane strain conditions, the size of the chip layer
is taken to be much smaller than the thickness of the
workpiece in the out-of-plane direction. Three surface contact pairs have been dened in Fig. 1 to simulate chip
separation: (1) between the chip and the tool, (2) between
the workpiece and the tool, and (3) along the prospective
chip separation line (which separates the chip layer from the
rest of the workpiece). The chip separation criterion is
discussed later.

Fig. 3. Variation of the cutting force with tool-tip displacement for four rake angles and four friction coefcient values.

98

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

A typical nite element mesh design for the model


problem is shown in Fig. 2. The initial orientation of the
chip layer elements is used to alleviate numerical problems
due to the distortion of the elements as they separate from
the workpiece and interact with the tool surface. The inclination angle of the elements with the cutting direction is
about 648. An initial chip separation is adopted in order to
achieve a smooth transition of the cutting process from the
initial stage to the steady state. The extra triangle of the chip
layer at the left end is used simply to make the mesh
generation simpler and will not affect the simulation result
before the cutting tool approaches the left end. This type of
mesh design has been used by several investigators [15,25].
The chip layer has a height (called the cut depth) of
254 mm and is divided into 10 sub-layers of elements. The
rest of the workpiece has a length of 2540 mm and a height of
889 mm and has been divided into 11 layers, each having 50
elements along the cutting path. (This study shows that 50
elements are sufcient for the simulation to reach the steady
state before the cutting tool reaches the left end.) The above

discretization results in a total of 1160 four-node plane strain


elements and 1308 nodes in the chipworkpiece system. It is
noted that all the top ve rows of elements just below the
cutting path have dimensions of 50.8 mm50.8 mm.
The cutting tool is considered much harder than the
workpiece and is taken to be made of a stiff elastic material
with an articially high Young's modulus (e.g.
E2.11015 MPa). As such, the size of the tool in the
cutting direction is not of signicance. The tool in this
study is of a parallelogram shape and has a base length
of 407 mm and a height of 762 mm. It is divided into 60
uniform, four-node plane strain elements.
The boundary conditions for the chipworkpiecetool
system are given as follows. The upper boundary of the
tool moves incrementally towards the left with a constant
speed of v2.54 m/s (152.4 m/min) while it is restrained
vertically. Assuming that the workpiece is very long in the
cutting direction (ignoring any transient effects at the start
and end of cutting), the left end and right end of the
workpiece are restrained in the cutting direction but not

Fig. 4. Contour plots of the temperature rise for the case of rake angle a208 and for four values of the friction coefcient m.

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109


Table 4
Schedule of the metal cutting simulations
Rake
angle (a)

Friction
coefficient ( m)

Cutting speed
(m/s)

Cut depth
(mm)

158
208
258
308

0.0,
0.0,
0.0,
0.0,

2.54
2.54
2.54
2.54

0.254
0.254
0.254
0.254

0.2,
0.2,
0.2,
0.2,

0.3,
0.4,
0.4,
0.4,

0.4
0.6
0.6
0.6

99

D2.21105 s1, p2.87). This rate dependent power law


is highly suitable for high strain rate applications (such as
high-speed metal cutting). Tables 1 and 2 list the temperature-dependent elastic and elasticplastic properties for the
material. Table 3 gives the temperature-dependent thermal
expansion coefcient. Standard constant values are used for
other properties: specic heat c502.0 J/kg K and the mass
density r7800 kg/m3.
2.3. Chip separation criterion and simulation

vertically. Since the bottom boundary of the workpiece is


expected to undergo very little deformation during cutting, it
is assigned zero displacements in both directions.
2.2. Material model and temperature-dependent properties
The workpiece material considered is AISI 4340 steel and
is modeled with a viscoplastic relationship of the over-stress
power law type

p
s
1
for s  s0
(1)
e_ p D
s0
where e_ p is the effective plastic strain rate, s the current yield
stress, s0 the initial yield stress, and D and p are material
parameters (following Komvopoulos and Erpenbeck [19],

In this study, chip separation is achieved through a nodal


release procedure in ABAQUS. This is done by dening a
bonded interface (see contact pair 1 in Fig. 1) along the
cutting path and by applying a critical stress separation
criterion available in ABAQUS to the stress state at a xed
distance ahead of the tool tip. When the stress state at the
specied distance reaches a critical combination, the pair of
bonded nodes just ahead of the tool tip will be released,
resulting in chip separation from the workpiece. Specically,
the critical stress criterion refers to the attainment of a
critical value of 1.0 by the stress index:
s
 2  2
sn
t

f
sf
tf

Fig. 5. Contour plots of the normal stress s11 for the case of rake angle a208 and for four values of the friction coefcient m.

(2)

100

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

where sn equals the normal stress s2 when it is tensile


and is set to zero when it is compressive (i.e. snmax (s2, 0),
t the shear stress, and sf and tf are the material failure
stress under pure tensile and shear loading conditions,
respectively.
In general, a number of process simulation parameters
must be quantied by comparing and matching suitable
quantities with cutting test results. Due to the lack of actual
cutting data, the values of the process parameters must be
assumed in the simulations conducted in this study (which
has been a common practice in the literature so far). In
particular, the specied distance in the critical stress criterion is set to equal to one element length, or approximately
50.8 mm. The failure stress in tension is assumed to be
sf948 MPa
p and the failure stress in shear is taken to be
tf sf = 3 (following the von Mises equivalent stress
concept), which is about 548 MPa.

2.4. Modied Coulomb friction law


Friction along the toolchip interface plays a very important role in the metal cutting process. To this end, the
modied Coulomb friction law option in ABAQUS can
be applied to the contact pair 2 in Fig. 1. Let t be the chip
shear stress at a contact point along the toolchip interface
and p the normal pressure at the same point. This law states
that relative motion (slip) occurs at the contact point when t
is equal to or greater than the critical friction stress tc. When
t is smaller than tc there is no relative motion and the contact
point is in a state of stick. The critical friction stress is
determined by
tc minmp; tth

(3)

where m is the friction coefcient and tth the threshold


value related to material failure. Note that the conventional

Fig. 6. Contour plots of the normal stress s22 for the case of rake angle a208 and for four values of the friction coefcient m.

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

101

Coulomb friction law is recovered if tth is set to innity: then


the equation reduces to the conventional Coulomb frictional
law. For the AISI 4340 steel, tth is chosen to be 549 MPa,
which is slightly higher than the shear failure stress tf.

temperature rise DTp in the active plastic zones can be


obtained as:

2.5. Energy dissipation and local heating

where se is the effective stress, ep the effective plastic strain


(a dot over a quantity denotes the time rate of the quantity), J
the equivalent heat conversion factor, c the specic heat, r
the mass density, and Zp is the percentage of plastic work
transformed into heat (usually, 85%Zp95%). A value of
Zp90% is used in this study.
Similarly, local temperature rise DTf in a time interval Dt
caused by friction along the toolchip interface can be
determined using the equation below:
t_sDt
(5)
DTf Zf
Jcr

In metal cutting heat is generated due to plastic work done


in the primary and secondary shear zones and also due to
sliding friction work along the toolchip interface. During
high-speed machining, heat generated due to local energy
dissipation does not have sufcient time to diffuse away and
local heating will occur in the active plastic zones and along
the sliding frictional interface. Thus temperature rise in the
chip can be approximated with the adiabatic heating condition. Mathematically, during a time interval of Dt, the local

DTp Zp

se e_ p Dt
Jcr

Fig. 7. Contour plots of the shear stress s12 for the case of rake angle a208 and for four values of the friction coefcient m.

(4)

102

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

where t is the shear stress along the frictional interface, s_ the


slip velocity, and J, c, and r have the same meaning as in
Eq. (4). The coefcient Zf stands for the portion of frictional
work being converted into heat, which is taken as 1.0 in this
study. Of the total heat generated along the toolchip interface, some goes into the chip and the rest into the tool. In this
study, it is assumed that 50% of the frictional heat will go
into the chip.
2.6. Simulation schedule
A total of 16 simulation cases have been performed,
which cover four rake angles and four friction coefcient
values for each rake angle. This allows for a parametric
evaluation of the effect of friction and rake angle on the
temperature, stress, strain, and strain rate elds. The details
of the simulation schedule are listed in Table 4.
3. Finite element results and discussion
Finite element simulations of the orthogonal metal cutting
process have been carried out for all 16 cases. In each case,
the cutting tool is made to advance incrementally until a

steady state is reached. On an average 150230 displacement increments are required to reach the steady state
condition, corresponding to about 47 h of CPU time on
a PC with a 400 MHz Pentium chip and running Windows
on NT4.0 operating system.
Fig. 3 shows the variation of the horizontal cutting force
with the tool tip displacement for four rake angles and four
friction coefcient values. For each value of the friction
coefcient m, the cutting force is seen to approach a constant
value as the cutting tool advances, indicating the achievement of a stead-state condition. For each rake angle, the
cutting force is seen to increase as the value of the friction
coefcient increases, as a result of increased resistance due
to friction along the toolchip interface.
Finite element simulation results presented below are taken
after the cutting tool has moved more than 1.5 mm thus they
represent typical steady-state solutions. Because of space
limitations, only results for the case of rake angle a208 are
given. Solutions for other rake angles bear similar features.
3.1. Temperature distribution
The distribution of temperature rise induced by energy
dissipation and local heating is shown in Fig. 4. It is

Fig. 8. Contour plots of the von Mises effective stress se for the case of rake angle a208 and for four values of the friction coefcient m.

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

observed that temperature rise occurs mainly in the chip


region, with maximum values localized along the toolchip
interface. There is very little temperature rise in the workpiece and ahead of the tool tip. The magnitude of temperature is very much affected by the friction along the toolchip
interface. This is evidenced by the fact that the maximum
temperature values along the toolchip interface increase as
the friction coefcient value increases. However, the overall
temperature distribution in the chip is determined by plastic
work dissipation in the chip rather by friction along the tool
chip interface, as can be seen from the case of zero friction.
3.2. Stress distribution
The distributions of stress components s11, s22 and s12,
the von Mises effective stress se, and the mean stress sm are

103

shown, respectively, in Figs. 59, where the unit for the


stresses is MPa. In particular, Fig. 5 shows that the normal
stress s11 (S11 in the gure) is compressive in the chip and in
the workpiece head of the tool tip while it is tensile in the
workpiece at and behind the tool tip, which is consistent
with intuition. The normal stress s22 (S22 in the gure) is
given in Fig. 6. This stress is tensile in a sizable region ahead
of the tool tip, which is necessary in order for chip separation
to occur. It is observed that the magnitude of this tensile
stress decreases as the value of the friction coefcient
increases, making chip separation harder. Fig. 7 shows that
the shear stress s12 (S12 in the gure) maintains a constant
sign in a large region ahead of the tool tip. It is worth
noting that, according to the critical stress criterion for chip
separation, s22 and s12 are the two driving forces for chip
separation.

Fig. 9. Contour plots of the mean stress sm for the case of rake angle a208 and for four values of the friction coefcient m.

104

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

The plastic ow behavior can be observed from the von


Mises effective stress distribution in Fig. 8. It is seen that the
stress contours in the chip ahead of the tool tip are parallel
and aligned in a left forward direction. The peak contour is
seen to connect the tool tip and the turning point on the
chip's free boundary, forming the ``shear'' angle. The mean
stress distribution in Fig. 9 provides insight into possible
``constraint'' effects in the material separation process ahead
of the tool tip. The gure shows that the mean stress is
positive just ahead of the tool tip and is compressive farther
away.
3.3. Strain distribution
The distribution of strain components e11, e22, and e12 are
shown in Figs. 1012, respectively. In particular, Fig. 10
gives the contour plots of e11 (E11 in the gure) for four

different values of the friction coefcient m. It shows that


when m is zero and 0.2, e11 in the chip is compressive on the
side of the toolchip interface and is tensile on the side of the
free boundary. As m increases, a zone of positive e11 develops
at the toolchip interface near the tool tip. Interestingly, just
the opposite observations can be made about the corresponding contour plots for e22 (E22 in the gure), as shown in
Fig. 11. It is seen that when m is zero and 0.2, e22 is tensile on
the side of the toolchip interface and is compressive on the
side of the free boundary. As m increases, a zone of negative
e22 develops at the toolchip interface near the tool tip,
which is caused by sticking contact of the chip with the
tool's rake face near the tool tip.
The opposite behavior of e11 and e22 can be explained by
the fact that the chip's deformation is mostly plastic and thus
is incompressible, which leads to e11e22e330. Under
plane strain conditions, e330, thus e11e22 (i.e. e11 and e22

Fig. 10. Contour plots of the normal strain e11 for the case of rake angle a208 and four values of the friction coefcient m.

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

105

Fig. 11. Contour plots of the normal strain e22 for the case of rake angle a208 and four values of the friction coefcient m.

have opposite signs). It is noted that, as m increases, the


curvature of the chip decreases, which can be explained by
the differences in the strain contours along the chip's free
boundary.
Fig. 12 shows the distribution of the shear strain e12 (E12
in the gure). It is seen that most of the shear deformation
occurs in the chip and relatively small shear strains are
present in the workpiece. A large shear strain gradient is
seen to exist along the line connecting the tool tip and the
turning point on the chip's free boundary. This region of high
strain gradient is known as the primary shear zone. It is clear
that the strain gradient becomes larger as the friction coefcient increases.
3.4. Strain rate distribution
Distributions of the rate of change of strain components
e11, e22, and e12 are given, respectively, in Figs. 1315 (the
unit of the strain rates is m/m s1). In all cases, the contours
of the strain rates are localized in the primary shear zone,

with peak strain rates found at the tool tip and near the
turning point of the chip's free boundary. Strain rate contours are also found to exist in the chip along the toolchip
interface, which is often called the secondary shear zone.
However, the distribution of the contours in this zone is
strongly dependent on the friction coefcient m. For the four
friction values shown, the number of contours increases as m
increases from 0.0 to 0.2, and it decreases as m increases
further. The reason for this behavior is the frictional sliding
and sticking along the toolchip interface near the tool tip. It
seems that strain rates are enhanced in the contact region
when m equals 0.2 because of an optimized combination of
deformation and sliding motion, and strain rates are inhibited when m equals 0.6 because of sticking.
4. Summary and conclusions
Successful nite element simulations of the orthogonal
metal cutting process have been carried out using the

106

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

Fig. 12. Contour plots of the shear strain e12 for the case of rake angle a208 and four values of the friction coefcient m.

general-purpose nite element code ABQUS and under


plane strain conditions. This was achieved with a judicious combination of a set of advanced modeling options
in the code. In particular, chip separation was simulated
with a critical stress separation criterion and a nodal
release technique. Dry friction was assumed along the
toolchip interface and was modeled with a modied
Coulomb friction law. Local heating and temperature rise
were calculated based on plastic work in the chip and
frictional work along the toolchip interface and using
the adiabatic heating condition. An over-stress, rate-dependent, elasticviscoplastic constitutive law was employed,
along with temperature dependent material properties.
Large deformation was handled by an updated Lagrangian
formulation.
A total of 16 simulations have been performed, which
cover a range of tool rake angle and friction coefcient

values. Steady-state nite element solutions for the temperature rise, stress, strain, and strain rate elds have been
obtained and representative contour plots for these eld
quantities have been presented. Several conclusions can
be drawn from this study and from the nite element
solutions. First, this study demonstrates that it is possible
to carry out sophisticated nite element simulations of
metal cutting processes using advanced general-purpose
commercial codes. Second, the nite element simulations
were able to re-produce experimentally observed phenomena in orthogonal metal cutting, such as the existence of the
primary and secondary shear zones. Third, the nite element
solutions obtained in this study show that friction along
the toolchip interface strongly affects the distribution of
the thermomechanical elds. It is believed that details
afforded by nite element simulations will greatly benet
the engineer in gaining a better understanding of metal

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

107

Fig. 13. Contour plots of the rate of change of the normal strain e11 for the case of rake angle a208 and four values of the friction coefcient m.

Fig. 14. Contour plots of the rate of change of the normal strain e22 for the case of rake angle a208 and four values of the friction coefcient m.

108

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109

Fig. 15. Contour plots of the rate of change of the shear strain e12 for the case of rake angle a208 and four values of the friction coefcient m.

cutting processes and in aiding the design and application of


such processes.
Acknowledgements
The authors gratefully acknowledge the support of the
Mechanics and Materials Program of the National Science
Foundation (NSF Grant No.: CMS-9700405).
References
[1] G. Boothroyd, Photographic technique for the determination of metal
cutting temperatures, Br. J. Appl. Phys. 12 (1961) 238242.
[2] M.C. Shaw, Some observations concerning the mechanics of cutting
and grinding, Appl. Mech. Rev. 46 (1993) 7479.
[3] M.E. Merchant, Basic mechanics of the metal cutting process, J.
Appl. Mech. 11 (1944) A168A175.
[4] M.E. Mechant, Mechanics of the metal cutting process, J. Appl. Phys.
16 (1945) 267318.
[5] V. Piispanen, Theory of formation of metal chips, J. Appl. Phys. 19
(1948) 876881.

[6] E.H. Lee, B.W. Shaffer, The theory of plasticity applied to a problem
of machining, J. Appl. Mech. 18 (1951) 405413.
[7] H. Kudo, Some new slip-line solutions for two-dimensional steadystate machining, Int. J. Mech. Sci. 7 (1965) 4355.
[8] W.B. Palmer, P.L.B. Oxley, Mechanics of orthogonal machining,
Proc. Inst. Mech. Engrs. 173 (1959) 623638.
[9] P.L.B. Oxley, A.G. Humphreys, A. Larizadeh, The inuence of rate
of strain-hardening in machining, Proc. Inst. Mech. Engrs. 175
(1961) 881891.
[10] E.D. Doyle, J.G. Horne, D. Tabor, Frictional interactions between
chip and rake face in continuous chip formation, Proc. R. Soc.
London A 366 (1979) 173183.
[11] K.J. Trigger, B.T. Chao, An analytical evaluation of metal cutting
temperature, Trans. ASME 73 (1951) 5768.
[12] E. Usui, A. Hirota, M. Masuko, Analytical prediction of three
dimensional cutting process. Part 1: Basic cutting model and energy
approach, J. Eng. Ind., Trans. ASME 100 (2) (1978) 229235.
[13] E. Usui, T. Shirakashi, Mechanics of Machining from Descriptive
to Predictive Theory. On the Art of Cutting Metals 75 years Later,
Vol. 7, ASME PED, 1982, pp. 1335.
[14] K. Iwata, K. Osakada, Y. Terasaka, Process modeling of orthogonal
cutting by the rigidplastic nite element method, J. Eng. Mater.
Technol. 106 (1984) 132138.
[15] J.S. Strenkowski, J.T. Carroll III, A nite element model of
orthogonal metal cutting, J. Eng. Ind. 107 (1985) 347354.

C. Shet, X. Deng / Journal of Materials Processing Technology 105 (2000) 95109


[16] J.T. Carroll III, J.S. Strenkowski, Finite element models of
orthogonal cutting with application to single point diamond turning,
Int. J. Mech. Sci. 30 (1988) 899920.
[17] J.S. Strenkowski, K.J. Moon, Finite element prediction of chip
geometry and tool/workpiece temperature distributions in orthogonal
metal cutting, J. Eng. Ind. 112 (1990) 313318.
[18] T. Tyan, W.H. Yang, Analysis of orthogonal metal cutting processes,
Int. J. Numer. Meth. Eng. 34 (1992) 365389.
[19] K. Komvopoulos, S.A. Erpenbeck, Finite element modeling of
orthogonal metal cutting, J. Eng. Ind. 113 (1991) 253267.
[20] Z.C. Lin, S.Y. Lin, A coupled nite element model of thermo-elastic
plastic large deformation for orthogonal cutting, J. Eng. Mater.
Technol. 114 (1992) 218226.

109

[21] J. Hashemi, A.A. Tseng, P.C. Chou, Finite element simulation of


segmented chip formation in high-speed machining, J. Mater. Eng.
Perform. 3 (1994) 712721.
[22] J.M. Huang, J.T. Black, An evaluation of chip separation criteria for
the FEM simulation of machining, J. Manuf. Sci. Eng. 118 (1996)
545554.
[23] A.J. Shih, H.T.Y. Yang, Experimental and nite element predictions
of the residual stresses due to orthogonal metal cutting, Int. J. Numer.
Meth. Eng. 36 (1993) 14871507.
[24] A.J. Shih, Finite element analysis of orthogonal metal cutting
mechanics, Int. J. Mach. Tools Manuf. 36 (1996) 255273.
[25] A.J. Shih, Finite element simulation of orthogonal metal cutting, J.
Eng. Ind. 117 (1995) 8493.

You might also like