Вы находитесь на странице: 1из 21

Finite Elements in Analysis and Design 38 (2002) 863 883

www.elsevier.com/locate/nel

A nite element study of the e#ect of friction in orthogonal


metal cutting
Guoqin Shi1 , Xiaomin Deng , Chandrakanth Shet
Department of Mechanical Engineering, University of South Carolina, Columbia, SC 29208, USA

Abstract
The process of orthogonal metal cutting is studied with the nite element method under plane strain
conditions. A computational procedure has been developed for simulating orthogonal metal cutting using
a general-purpose nite element code. The focus of the results presented in this paper is on the e#ect of
friction on thermomechanical quantities in a metal cutting process. A series of nite element simulations
have been performed, in which a modied Coulomb friction law is used to model the friction along the
toolchip interface, and a nite element nodal release procedure is adopted to simulate chip separation from

the workpiece. A tool rake angle ranging from 15 to 30 and a friction coe1cient ranging from 0.0 to
0.6 have been considered in the simulations. The results of these simulations are consistent with experimental observations in the literature. In particular, it is found that shear straining is localized in the primary
shear zone while the material near the tool tip undergoes the largest plastic strain rate. However, the maximum temperature rise, which is induced by energy dissipation due to plasticity and friction, occurs along
the toolchip interface, not in the primary shear zone. Furthermore, the maximum temperature, the contact
length, the shear angle, and the cutting force are found to depend strongly on the coe1cient of friction.
? 2002 Elsevier Science B.V. All rights reserved.
Keywords: E#ect of friction; Finite element simulation; Orthogonal metal cutting

1. Introduction
Metal cutting is one of the most common manufacturing processes for producing parts of desirable dimensions. It is used to remove unwanted material from a workpiece and obtain specied

Corresponding author. Tel.: +1-803-777-7144; fax: +1-803-777-0106.


E-mail address: deng@engr.sc.edu (X. Deng).
1
Current address: Structures, Materials and Propulsion Laboratory, Institute for Aerospace Research, National Research
Council of Canada, Montreal Road, Ottawa, ON, Canada K1A 0R6.
0168-874X/02/$ - see front matter ? 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 1 6 8 - 8 7 4 X ( 0 1 ) 0 0 1 1 0 - X

864

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

geometrical dimensions and surface nish. Understanding of the material removal process in metal
cutting is important in selecting tool material and design and in assuring consistent dimensional accuracy and surface integrity of the nished product, especially in automated production and precision
parts manufacturing.
Metal cutting is a highly nonlinear and coupled thermomechanical process, where the coupling is
introduced through localized heating and temperature rise in the workpiece, which is caused by the
rapid plastic Gow in the workpiece and by the friction along the toolchip interface. For example,

temperature rise of up to 1000 C has been reported in the literature (e.g. [1,2]). As such, metal
cutting modeling has not been easy. Early analyses of metal cutting were based on simple models,
such as the shear-angle approach proposed by Merchant [3,4], Piispanen [5], and Oxley [6], and the
slip-line eld theory by Lee and Sha#er [7] and Kudo [8] based on rigid-perfectly plastic material
behavior. These models were later extended to include the e#ect of work hardening [9,10], friction
[11] and built-up edge [12].
In the last two decades, the nite element method has been applied to study and simulate metal
cutting processes and various nite element simulation techniques have been developed. An early
nite element model by Usui and Shirakashi [13] treated steady-state metal cutting based on empirical
data and assuming a rate-independent deformation behavior. A later study by Iwata et al. [14]
considered the e#ect of friction between the chip and the tool rake face but was limited to very low
cutting speeds and strain rates and assumed rigid-plastic deformation.
Perhaps the early comprehensive nite element studies of metal cutting were those by Strenkowski
and Carroll [15] and Carroll and Strenkowski [16]. They used the general-purpose nite element code
NIKE2D and employed an updated Lagrangian formulation to model the orthogonal metal cutting
process. They developed a technique for element separation in front of the tool tip and proposed an
element-separation criterion based on the magnitude of plastic strains. Their technique was used to
simulate the cutting process from the incipient stage to the steady state and to predict cutting force,
chip geometry, plastic deformation and residual stress in the workpiece. Alternatively, Strenkowski
and Moon [17] proposed a steady-state metal cutting technique based on an Eulerian formulation.
They used the technique to predict chip geometry and temperature distribution. A good correlation
between model predictions and metal cutting measurements was found.
The nite element study by Komvopoulos and Erpenbeck [18] focuses its attention on chip formation in orthogonal metal cutting and on the e#ect of such factors as plastic Gow properties of
the workpiece material, friction at the toolworkpiece interface, and wear of the tool, on the cutting
process. The simulation of chip separation was achieved by using the distance tolerance criterion.
The work by Tyan and Yang [19] analyzed the orthogonal metal cutting process based on a limit
analysis theorem. An Eulerian reference coordinate system was used to describe the steady state
motion of the workpiece relative to the tool.
More recently, Shih and Yang [20] and Shih [2123] conducted a combined experimental and
numerical investigation of the orthogonal metal cutting process. The e#ects of large strain, high
strain-rate and temperature were considered. The chip separation criterion used in this investigation
was based on the distance in the cutting direction between the tip of the cutting tool and the
nite element nodal point located immediately ahead of the tip. The studies by Hashemi et al. [24]
and Lei et al. [25], respectively, employed fracture-mechanics and crack length based algorithms to
simulate chip separation from the workpiece and the simultaneous breakage of the chip into multiple
segments.

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

865

Despite the great progress that has been achieved in modeling metal cutting processes, as indicated
by the references cited above, there are still several fundamental issues that remain open and deserve
detailed analysis. One of these issues is the frictional behavior along the toolchip interface and its
e#ect on thermomechanical quantities involved in metal cutting. From the point of view of numerical
simulations, the friction between the toolchip interface has been modeled in the literature according
to the Coulomb Friction Law. However, there are currently no established experimental techniques
that can be used to measure directly the coe1cient of friction during metal cutting, not to mention
validating this law. Another issue is the understanding of tool wear and its correlation, if any,
with thermomechanical quantities in metal cutting. This is a complicated issue and its understanding
depends on many factors. It appears that so far research e#ort has been focused on the physics
of wear and no attempt has been made to quantify wear and tool life by combining experimental
observations with numerical simulation solutions of thermomechanical quantities in metal cutting.
The purpose of this study is to address the issue of the e#ect of friction along the toolchip
interface on thermomechanical quantities in orthogonal metal cutting and to seek insight into possible ways of calibrating the frictional behavior along the toolchip interface. To achieve the stated
purpose, this study adopts the nite element method to simulate the orthogonal metal cutting process
and to obtain parametric evaluations of the e#ect of friction (such as the maximum temperature
rise and the contact length along the toolchip interface) according to a prescribed friction law.
Specically, the general-purpose nite element code ABAQUS is adopted to study orthogonal metal
cutting with continuous chip formation. To facilitate the simulation of chip separation from the
workpiece, a stress-based chip separation criterion is proposed. 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. To account for the e#ect of friction along the contact surfaces between the chip and
the cutting tool, contact surfaces are dened where a Modied Coulomb Friction Law is applied.
Local temperature rise due to heat generation caused by plastic work and friction is considered and
temperature-dependent material properties are employed. Strain-rate e#ect and large strains are also
included in the analysis. The nite element simulations are carried out under plane strain conditions
and for a range of friction coe1cient values and rake angle values. It is noted that the results of
this study have been previously presented at a professional conference [26].
The rest of the paper is arranged as follows. Section 2 describes some important details of the
nite element simulations, such as the chip separation criterion and the modied Coulomb friction
law. Section 3 presents the results of the nite element simulations for the temperature distribution,
shear angle, contact length, etc. Finally Section 4 gives a summary of the main results and concludes
with a few important observations.
2. Simulation details
A schematic of the orthogonal metal cutting process is shown in Fig. 1, where a workpiece is
fed to a cutting tool, which cuts through the workpiece and separates a chip from the workpiece.
Equivalently, it can be said that the cutting tool moves forward while the workpiece stays stationary. The chip formation is assumed to be continuous. Numerical simulations of the orthogonal metal
cutting process reported in this study were made possible through a specialized computational procedure developed by the authors for the general-purpose nite element code ABAQUS. Time- and

866

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

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

rate-dependent elasticplastic behavior of the workpiece material is considered. An iterative procedure based on the Updated Lagrangian Formulation was employed to deal with large deformations.
Discussions given below focus on some special measures adopted in the current study. They are
taken for several reasons: (a) to observe certain common practices in the literature, (b) to make
use of special options in the ABAQUS code, and (c) to simplify the simulations so that they are
computationally tractable.
2.1. Adiabatic heating
Heat generated in metal cutting can have a signicant e#ect on chip formation, tool wear and
the quality of the nished surface. The heat generation mechanisms are the plastic work done in
the primary and secondary shear zones and the sliding friction along the toolchip interface. In
high-speed metal cutting, heat generated in the workpiece and chip does not have su1cient time
to di#use away. Therefore, temperature rise in the workpiece and chip is mainly due to localized
adiabatic heating. Based on this argument and to simplify the analysis, adiabatic heating conditions
are generally assumed in metal cutting simulations, so is the case here. Consequently, each nite
element integration point is treated as if it is thermally insulated from its neighbors.
Let OTp be the change in temperature (local temperature rise) in the workpiece and chip induced
by plastic work in a time interval Ot. Under adiabatic conditions OTp can be obtained using the
equation below
OTp
e p
= p
;
Ot
Jc

(1)

where e is the e#ective stress, p is the e#ective plastic strain (a dot over a quantity denotes the
time rate of the quantity), J is the equivalent heat conversion factor, c is the specic heat,  is
the mass density, and p is the percentage of plastic work that is transformed into heat (usually,
85% 6 p 6 95%). A value of p = 90% is used in this study.

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

867

Table 1
Temperature-dependent elastic material properties

Youngs modulus E (GPa)

Poissons ratio

Temperature ( C)

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

Similarly, the local temperature rise OTf caused by friction in a time interval Ot can be determined
under adiabatic conditions as follows:
OTf
s
= f
;
Ot
Jc

(2)

where is the frictional shear stress, s is the slip velocity, and J; c, and  have the same meanings
as in Eq. (1). The coe1cient f stands for the portion of friction-induced heat along the toolchip
interface that goes into the chip (the rest of the heat goes into the tool), which is taken to be 0.5 in
this study. It is noted that in general is dependent on the contact pressure p, the friction coe1cient
 and the temperature.
2.2. Material parameters
The workpiece material is taken to be AISI 4340 steel, which is described in this study by the
following commonly accepted rate-dependent elasticplastic material model

p

p = D
1
for  0 ;
(3)
0
where p is the e#ective plastic strain rate,  is the current yield stress, 0 is the initial yield stress, D
and p are material parameters. Following [15], D and p are taken to be temperature independent and
have values of 2:21 105 s1 and 2.87, respectively. The temperature-dependent Youngs modulus
values are listed in Table 1 and the temperature-dependent elasticplastic behavior of the material
is shown Table 2, where a one-to-one relation between the Gow stress (the current yield stress) and
the accumulated plastic strain is given.
Other material parameters for the steel are as follows. The thermal parameters are considered
constants, with the specic heat c=502:0 J=kg K and the mass density =7800 kg=m3 . A temperaturedependent thermal expansion coe1cient is used, which is list in Table 3 at a range of temperature
values. Because the cutting tool is much sti#er than the workpiece material, it is modeled as a rigid
material, with an articially large Youngs modulus of E = 2:1 1015 MPa.

868

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883
Table 2
Temperature-dependent elasticplastic material behavior

Flow stress  (MPa)

Plastic strain p (mm=mm)

Temperature ( C)

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

Table 3
Temperature-dependent thermal expansion coe1cient

Thermal expansion coe1cient  (m=m K)

Temperature ( C)

12.3
12.7
13.7
14.5

20.0
200.0
400.0
600.0

2.3. Separation criterion


Several chip separation criteria have been advanced in the literature to facilitate the simulation
of metal cutting processes. The criterion by Strenkowski and Carroll [15] states that chip separation
occurs in a nite element simulation when the plastic strain at the node ahead of the tip of the
cutting tool reaches a critical value. Iwata et al. [14] introduced a separation criterion based on
stress history and Lin and Lin [27] suggested a criterion based on the strain energy density. More
recently, Shih [21] proposed a criterion with the distance ahead of the tool tip as the controlling

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

869

Fig. 2. A graphical description of some denitions used in this study.

parameter. A comparison of various chip separation criteria can be found in the paper by Huang
and Black [28].
In the current study, to make use of a special option in the ABAQUS code, we propose a
stress-based chip separation criterion. This criterion states that chip separation occurs when the
stresses along the cutting path reach a critical combination at a specied distance in front of the
tool tip. To implement this criterion in ABAQUS, the cutting path in the workpiece is dened by a
contact surface (see contact pair 1 in Fig. 2). Paired nite element nodes on the contact surface are
initially in perfect bond. When the chip separation criterion is met at the specied distance in front
of the tool tip, the pair of nite elements above and below the contact surface immediately before
the tool tip will separate. After the separation and as the tool moves forward, the element above
the contact surface will move into contact pair 2 (which refers to the chiptool interface) and the
element below the contact surface will move into contact pair 3, as illustrated in Fig. 2.
The particular combination of stresses used in the stress-based criterion is dened in terms of a
stress index parameter f given below
   
n 2
2
f=
+
; where n = max(2 ; 0):
(4)
f
f
In the above, and 2 are the shear and normal stress components at a specied distance in front
of the tool tip along the cutting path, as shown in Fig. 2, and f and f are the failure stresses of
the material under pure tensile and shear loading conditions, respectively. A critical combination is
considered reached if the stress index f attains a value of 1.0
In this study, the specied distance for the stress-based chip separation criterion is chosen to equal
one element size (approximately 50:8 m). (The element size is designed to be uniform along the
cutting path.) Due to the lack of actual material data, the failure stresses f and f have to be

870

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

assumed for the steel used in this study. Based on the yield strength of the steel, the failure stress
in tension is taken to be f =
948 MPa. Based on the von Mises yield surface, the failure stress in
shear is chosen to be f = f = 3, which is about 548 MPa.
2.4. Friction law
Friction plays a very important role in metal cutting. It not only determines the power requirement
for removing a given volume of metal, but also controls the surface quality of the nished product
and the rate of wear of the cutting tool. Friction is also di1cult to model in metal cutting. It is
widely accepted that two distinct contact regions, namely, the sticking region and the sliding region,
exist simultaneously along the toolchip interface. In the sticking region, a critical friction stress, c ,
is considered to exist, while in the sliding region, a coe1cient of friction, , is often assumed with
regard to the Coulomb friction law.
In order to explore the e#ect of friction along the toolchip interface in orthogonal metal cutting, a
modied Coulomb friction law is employed in the current study. This law is provided as a generalized
Coulomb friction model in ABAQUS and can be used with all contact options in the code. This
friction model can be described as follows. Let be the chip shear stress parallel to the toolchip
interface. During metal cutting, if at a toolchip contact point is less than the critical friction stress
c then no relative motion occurs between the chip and the rake face of the cutting tool. In this case,
the contact point is in the sticking region. On the other hand, if is equal to or greater than the
critical friction stress c , then relative motion occurs between the chip and the tool, and the contact
point is in the sliding region. The physical basis for the modied Coulomb friction law has been
discussed in detail by Kalpakjian [29].
According to the modied Coulomb friction law, the critical friction stress c is determined by
c = min(p; th );

(5)

where p is the normal pressure across the toolchip interface,  is the coe1cient of friction, and
th is a threshold value for the conventional Coulomb friction stress (which is given by the product
p). Eq. (5) states that the critical friction stress equals the conventional Coulomb friction stress
when it is smaller than the threshold value, otherwise it equals the threshold value. When th is
set to innity, Eq. (5) recovers the original Coulomb friction law. In this study, th is chosen to
be 549 MPa for the AISI 4340 steel, which is slightly higher than the shear failure stress f . The
rationale is that when shear failure occurs in the chip at the toolchip interface, the chip will also
move relative to the tool.
2.5. Mesh design and boundary conditions
The nite element simulations of the orthogonal metal cutting process are carried out under
plane strain conditions. The workpiece in Fig. 2 is taken to have a thickness of 10 mm and the
in-plane geometry of the toolworkpiece system is discretized by the nite element mesh shown in
Fig. 3. The tilted orientation of the elements in the chip layer is designed to counter the e#ect of
the expected large distortion of the elements as the chip is separated from the workpiece by the
cutting tool. This type of e#ective mesh design was originally proposed by Strenkowski and Carroll

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

871

Fig. 3. The nite element mesh used in the metal cutting simulations.

[15] and has since been adopted by other researchers (e.g. [21]). The inclination angle of the tilted

elements with the cutting direction is about 64 in this study. The extra triangular geometry of the
chip layer at the left end is kept there for the simplicity of mesh generation, but it is not expected
to a#ect the simulation solution before the cutting tool approaches the left end. Furthermore, an
initial chip separation at the right-end of the chip layer is used in this study to facilitate a smooth
transition of the cutting simulation from the initial stage to the steady state.
Including the elements for the cutting tool, the nite element mesh contains a total of 1160
four-node plane strain elements and 1308 nodes. Because the chip will undergo the largest deformation, about half of the elements are placed in the chip layer, which has a height (the cut depth)
of 254 m and is divided into 10 sub-layers of elements. Below the chip layer, the workpiece is
represented by a rectangle with a length of 2540 m and a height of 889 m. The rectangle is
divided into 50 elements along the cutting path (more than enough for the simulation to reach the
steady state before the cutting tool gets near the end of the cutting path). Consequently the ve
rows of elements right below the cutting path are square elements of dimensions 50:8 m 50:8 m.
Relatively speaking, the workpiece material below the bottom boundary of the rectangle is expected
to undergo very little deformation, which is taken to be the zero displacement boundary condition
for the bottom boundary. Neglecting any transient e#ects at the start and the end of the cutting
simulations, we can treat the workpiece to be very long in the cutting direction, so that the left and
right boundaries of the rectangle can be considered constrained in the cutting direction.
The cutting tool is usually simplied as a rigid block in the literature. Consistent with this simplication, this study models the cutting tool with an elastic material with a very high sti#ness
(Youngs modulus). The parallelogram representing the cutting tool has a base length of 407 m
and a height of 762 m and is discretized into 60 equal-sized four-node elements. During the cutting simulation, the top surface boundary of the cutting tool is constrained to zero displacement
in the y-direction and is programmed to move in the negative x-direction at a constant speed of
v = 2:54 m=s (152:4 m=min). The movement of the cutting tool is accomplished incrementally. That
is, at the beginning of every new time increment Ot, the top surface of the tool is made to move
forward in the cutting direction by an increment of Ox = vOt.

872

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883
Table 4
Schedule of the metal cutting simulations
Rake angle ()

Friction coe1cient ()

Cutting speed (m=s)

Cut depth (mm)

15

0.0
0.2
0.3
0.4

2.54

0.254

20

0.0
0.2
0.4
0.6

2.54

0.254

25

0.0
0.2
0.4
0.6

2.54

0.254

30

0.0
0.2
0.4
0.6

2.54

0.254

2.6. Simulation schedule


The focus of this study is to examine the e#ect of friction along the toolchip interface on the
metal cutting process. As such, the simulation schedule in this study includes a total of 16 cases,

with the tools rake angle ranging from 15 to 30 and the coe1cient of friction ranging from 0.0 to
0.6. The detail of the simulation schedule is listed in Table 4. The reason why di#erent rake angles
are used is that when the rake angle varies, the resultant shear and pressure force components on
the toolchip interface will have di#erent relative values at the same cutting speed, which will cause
changes in the e#ect of the friction. At each rake angle, four di#erent values for the coe1cient
of friction will be employed at the same cutting speed and cut depth (because of di1culty with

numerical convergence, the highest friction coe1cient value used is 0.4 when the rake angle is 15 ).
All simulations in this study are performed on a DEC Alpha workstation.

3. Results and discussions


The sixteen simulations in this study have generated a wealth of information about the orthogonal
metal cutting process. The results reported in this section are selected to reGect the focus of this
study, namely the e#ect of friction in orthogonal metal cutting. Similar results have not been seen
in the literature.

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

873

Fig. 4. Chip formation for the case of rake angle  = 25 and friction coe1cient  = 0:0, where the time shown is the
elapse time since the start of the cutting process.

3.1. Attainment of a steady state

Simulation results for the case of rake angle  = 25 and friction coe1cient  = 0:0 are used here
to illustrate the evolution of continuous chip formation in orthogonal metal cutting and to provide
a sense of when a stead state is achieved in the simulations. Fig. 4 shows four plots of the chip
geometry at di#erent instants of the cutting process. The last plot is for the instant at an elapse time
of 0:62 ms since cutting begins, which is reached after the tool has cut through 30 nite elements
along the cutting path. It can be seen that the contact length in the second plot, which is attained
along the toolchip interface at the time of 0:28 ms, remains the same in the last two plots. Similarly,
the shape of the chip and of the deformed elements (up to the end of the contact zone) appears to be
the same in the last three plots. The above observation suggests that a steady state is reached very
quickly after the initiation of the cutting process, at least since the time of 0:28 ms. This conclusion
is further conrmed by the variation of the cutting force with time, as shown in Fig. 5 (the cutting
force is the resultant force exerted by the cutting tool on the workpiece in the cutting direction
and is over a total workpiece plate thickness of 10 mm). Indeed, the cutting force appears to have
attained a constant value at the time of 0:13 ms, indicating the achievement of a steady state. In
this light, all subsequent steady state results (including those in this and the next subsection) will
be drawn from simulation solutions obtained at the time of 0:62 ms.
3.2. Comparison with experimental data
Before we proceed with the results of the nite element simulations discussed above, it is appropriate to provide some indications about the adequacy of the developed computational procedure.

874

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

Fig. 5. Cutting force variation for the case of rake angle  = 25 and friction coe1cient  = 0:0, where the time shown
is the elapse time since the start of the cutting process.

Fig. 6. A comparison of nite element cutting force predictions with experimental measurement [30] for various cutting
velocities.

To this end, we will rst show a comparison of predicted steady state cutting forces with measured
(mostly transient) cutting forces as a function of cutting speed.
Because of limited experimental data in the literature, the comparison in this section is based on
the cutting test results reported in [30] on AISI 4340 steel specimens, in which the cutting force
and cutting speed were measured during orthogonal cutting tests.
To compare with these test results, nite element simulations were carried out on a workpiece

model of size 76:2 mm 25:4 mm 9:53 mm. The depth of cut is 0:18 mm and the rake angle is 3 .
A friction coe1cient of 0.3 is used (which is not chosen to best t the test data). The nite element
model used 7128 four noded plane strain elements with a total of 9366 nodes. Simulations were
performed with three di#erent cutting speeds (2.0, 2.12, and 2:35 m=s) until a steady state cutting
condition is attained.
The predicted steady-state cutting forces at the three cutting speeds are compared in Fig. 6 with
those reported in [30] (note that the cutting force is distributed over the workpiece plate thickness

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

875

Fig. 7. Distribution of the shear strain rate 12 at time = 0:62 ms and for the case of rake angle  = 25 and friction
coe1cient  = 0:2.

of 9:53 mm). The comparison shows that nite element simulation can basically capture the trend
observed in the experimental data. The quantitative di#erences between the predicted and measured
data are reasonable, considering the uncertainty of many material property values and the fact that
nite element predictions are for steady states while experimental results are from transient tests. In
particular, it is noted that the di#erence between the prediction and measurement becomes smaller as
the cutting speed increases, which coincides with the fact that the cutting tests in [30] at the higher
cutting speeds can be better approximated by a steady state.
3.3. General steady-state results
To observe some common features of the thermomechanical quantities involved in orthogonal
metal cutting, general results will be presented in this section.
Typical distributions of thermomechanical quantities in steady state metal cutting are presented

next for the case of rake angle  = 25 and friction coe1cient  = 0:2. To illustrate the amount of
shear deformation being generated in the chip, the distribution of the shear strain rate component
12 is shown in Fig. 7. As expected, severe shear deformation of the chip occurs in a kink band (the
primary shear zone) which forms an angle (the shear angle) with the cutting direction. The most
severe shear strain rate occurs near the tool tip.
The distribution of the von Mises e#ective stress is given in Fig. 8, for which the following
observation can be made. As material points in the chip layer move towards the cutting tool from
the far left, the stress level experienced by the points builds up, reaches maximum in the primary
shear zone, and relaxes as the points pass through the primary shear zone. It is noted that the contour
lines are not as localized within the primary shear zone as those of the shear strain rate, even though
the largest stress levels are still found there.

876

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

Fig. 8. Distribution of the von Mises e#ective stress e at time = 0:62 ms and for the case of rake angle  = 25 and
friction coe1cient  = 0:2.

Fig. 9. Distribution of the e#ective plastic strain p at time = 0:62 ms and for the case of rake angle  = 25 and friction
coe1cient  = 0:2.

Perhaps the most interesting observations can be made for Figs. 9 and 10, where contours for the
e#ective plastic strain and the temperature eld are shown, respectively. First of all, the shape and
distribution of the contour lines are almost identical in the two gures. Secondly, the contour lines
form two bands, one at the location of the primary shear zone and the other along the toolchip

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

877

Fig. 10. Distribution of the temperature eld at time=0:62 ms and for the case of rake angle =25 and friction coe1cient
 = 0:2.

interface, which is often referred to as the secondary shear zone. Thirdly, and most interestingly,
the highest e#ective plastic strain and temperature levels are not found in the primary shear zone,
but in the secondary shear zone, especially right at the toolchip interface.
The above seemingly peculiar phenomenon can be explained as follows. In a metal cutting process,
material points can be imagined to be Gowing from one location to another as the cutting tool cuts
through the workpiece. As the material points in the chip layer Gows from the far left of the
workpiece towards the cutting tool, they undergo severe plastic straining (as indicated in Fig. 7)
in the primary shear zone and at the tool tip. Temperature is increased instantly in these areas
due to the induced local adiabatic heating according to Eq. (1), which relates the temperature rise
proportionally to the e#ective plastic strain rate. Because the highest plastic straining in the primary
shear zone occurs near the tool tip (see Fig. 7), so do the most severe local heating and temperature
rise, as suggested by the converging contour lines at the tool tip (see Figs. 9 and 10). Then, the
accumulated plastic strain and temperature rise are carried into the chip as the material points Gow
past the primary shear zone, where further plastic straining and local heating are experienced along
the toolchip interface because of severe contact and friction in the contact zone. Consequently, the
maximum e#ective plastic strain and temperature are found along the toolchip interface, at some
distance away from the tool tip.
Distributions of the normal pressure and the critical friction stress (according to the modied
Coulomb friction law, Eq. (5)) along the toolchip interface are shown in Fig. 11 as functions of
the nite element nodal position along the interface, where node number 0 refers to the tool tip node.
As can be seen, the pressure reaches its maximum at a short distance away from the tool tip, then
it decreases gradually until it becomes zero at 15 nodal points away. It is noted that the distance
between the tool tip to the point where the pressure becomes zero denes the contact length, which
will be discussed further in the next subsection.

878

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

Fig. 11. Variations of the normal pressure p and critical friction stress c along the toolchip interface, at time = 0:62 ms

and for the case of rake angle  = 25 and friction coe1cient  = 0:2, where node number 0 refers to the tool-tip node.

3.4. E3ect of friction


This section discusses the e#ect of friction on chip formation, the contact length, the shear angle,
the cutting force, and the maximum temperature. The ndings presented here are based on nite
element simulation solutions assuming the modied Coulomb friction law. Four coe1cient of friction

values (ranging from 0.0 to 0.6) and four rake angle values (15 , 20 , 25 , and 30 ) are considered

in the simulations. When the rake angle takes the small value of 15 , the coe1cient of friction varies
only between 0.0 and 0.4. At the higher coe1cient value of 0.6, the friction force is too large and
the cutting operation cannot proceed.
The typical steady state chip geometry at time = 0:62 ms is shown in Fig. 12 for the case of

rake angle  = 25 and various values for the friction coe1cient. It is clear that, as the value of
the friction coe1cient increases, the chip curvature becomes smaller and the toolchip contact zone
gets longer. The strong dependence of the contact length on the friction coe1cient as well as on
the rake angle is more clearly illustrated in Fig. 13, where a monotonic and one-to-one relationship
between the contact length and the friction coe1cient is observed for all the rake angle cases. For
each friction coe1cient value, a monotonic and one-to-one relationship also seems to exist between
the contact length and the rake angle only in this case the contact length increases as the rake
angle decreases.
The shear angle, which is the angle between the cutting direction and the kink band (the primary
shear zone) originating from the tool tip (see Figs. 12 and 7), also depends quite strongly on the
friction coe1cient and the rake angle, as shown in Fig. 14. For a xed rake angle, the shear angle
decreases as the friction coe1cient increases, and for a xed friction coe1cient, the shear angle
increases as the rake angle increases.
The dependence of the cutting force on the friction coe1cient is plotted in Fig. 15 and is very
similar (perhaps with more linearity) to that given in Fig. 13 for the contact length. As expected,
a larger friction coe1cient implies a higher frictional resistance along the toolchip interface, hence

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

879

Fig. 12. Chip formation at time=0:62 ms, for the case of rake angle =25 and various values for the friction coe1cient .

Fig. 13. E#ect of friction on the contact length along the toolchip interface, for various rake angle values.

a higher cutting force to overcome it. On the other hand, as the rake angle decreases, the normal
force across the toolchip interface becomes larger for the same amount of cutting force. The larger
normal force leads to a larger frictional resistance while at the same time the sliding force (which
must overcome the frictional resistance) becomes smaller, that is why the required cutting force
increases as the rake angle decreases, as demonstrated by Fig. 15.

880

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

Fig. 14. E#ect of friction on the shear angle, for various rake angle values.

Fig. 15. E#ect of friction on the cutting force, for various rake angle values.

The e#ect of friction on the temperature rise will be shown only with regard to the maximum
temperature, which occurs along the toolchip interface at some distance away from the tool tip (see
Fig. 10). As shown in Fig. 16, the maximum temperature rise due to local heating caused by plastic
work in the chip and by friction along the toolchip interface has an appreciable dependence on
the friction coe1cient and on the rake angle. The most pronounced dependence is seen for the case

with a small rake angle of 15 . For the rake angles of 20 and 25 , there is a clear monotonic and
one-to-one relationship between the maximum temperature and the friction coe1cient  when  is
smaller than 0.4, above which a plateau is observed. The reason for the existence of the plateau is
as follows. At higher friction coe1cient values, the normal pressure across the toolchip interface
is su1ciently high, such that the critical friction stress becomes constant according to the Modied

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

881

Fig. 16. E#ect of friction on the maximum temperature, for various rake angle values.

Coulomb friction law (see Eq. (5)) instead of increasing monotonically with the normal pressure
and the friction coe1cient. Since the temperature change due to local heating caused by friction
is proportional to the frictional shear stress (see Eq. (2)), which is limited by the critical friction
stress, the maximum temperature rise along the toolchip interface is also bounded from above. If
the threshold value in Eq. (5) is raised, the maximum temperature in Fig. 16 for the higher friction
coe1cient values will increase.
The simulation results presented above provide a clear understanding of the e#ect of friction in
orthogonal metal cutting. Even though the particular simulations reported in this paper are carried
out assuming the modied Coulomb friction law, the e#ect of friction according to other friction
models can be examined similarly. It is well known that one of the obstacles blocking the practical
application of nite element simulations in metal cutting is the verication of the friction models
being used in the simulations. It is usually di1cult to determine the actual value of the friction
coe1cient from cutting tests. As such, the point that is worth making here is that a combination of
numerical simulations and experimental measurements may provide a way of calibrating the friction
models as well as the parameters contained in the models. For example, the friction coe1cient in
the modied Coulomb friction law may be calibrated by comparing experimentally measured cutting
force with numerically predicted cutting force. The contact length, the shear angle, and the maximum
temperature provide additional verication data for such an integrated computational=experimental
approach.

4. Summary and conclusions


A computational procedure has been developed based on the general-purpose nite element code
ABAQUS, for modeling and simulating the orthogonal metal cutting process under plane strain
conditions. A new chip separation criterion based on the stress state in front of the tool tip has been
proposed and used successfully in the reported nite element simulations. A nite element nodal

882

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

release procedure is adopted to facilitate the simulation of chip separation from the workpiece. The
focus of this study is on the e#ect of friction in orthogonal metal cutting. Sixteen nite element
simulations have been performed, in which four friction coe1cient values and four rake angles are
considered. The main ndings of this study are as follows:
1. The general-purpose nite element code ABAQUS can be used to simulate the orthogonal metal
cutting process through a proper combination of user options and subroutines.
2. Plastic shear straining in the chip is localized within a band (the primary shear zone) originating
from the tool tip, with the maximum shear strain rate occurring at the tool tip.
3. The maximum temperature rise and maximum e#ective plastic strain occur somewhat away from
the tool tip along the toolchip contact interface.
4. Chip formation, the toolchip contact length, the shear angle of the primary shear zone, the required tool cutting force, and the maximum temperature rise all depend strongly on the coe1cient
of friction and on the rake angle. For a xed rake angle, as the friction coe1cient increases, the
chip curvature and the shear angle decrease, while the contact length, the cutting force and
the maximum temperature increase. For a xed friction coe1cient, as the rake angle increases,
the shear angle increases, while the contact length, the cutting force, and the maximum temperature decrease.
5. The ndings of this study suggest that it may be possible to devise a practical procedure for
establishing a friction model for the toolchip contact interface and for calibrating the friction
parameters in the model by using a combined experimental=computational approach. For example,
assuming the modied Coulomb friction Model is valid then the value of the friction coe1cient
may be determined by comparing the predicted and measured values for the cutting force, the
contact length, and the maximum temperature.
Acknowledgements
The authors gratefully acknowledge the support of the National Science Foundation (NSF Grant
No. CMS-9700405).
References
[1] G. Boothroyd, Photographic technique for the determination of metal cutting temperatures, British 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] P.L.B. Oxley, Shear angle solutions in orthogonal machining, Int. J. Mach. Tool Des. Res. 2 (1962) 219229.
[7] E.H. Lee, B.W. Sha#er, The theory of plasticity applied to a problem of machining, J. Appl. Mech. 18 (1951)
405413.
[8] H. Kudo, Some new slip-line solutions for two-dimensional steady-state machining, Int. J. Mech. Sci. 7 (1965)
4355.
[9] W.B. Palmer, P.L.B. Oxley, Mechanics of orthogonal machining, Proc. Inst. Mech. Eng. 173 (1959) 623638.

G. Shi et al. / Finite Elements in Analysis and Design 38 (2002) 863 883

883

[10] P.L.B. Oxley, A.G. Humphreys, A. Larizadeh, The inGuence of rate of strain-hardening in machining, Proc. Inst.
Mech. Eng. 175 (1961) 881891.
[11] E.D. Doyle, J.G. Horne, D. Tabor, Frictional interactions between chip and rake face in continuous chip formation,
Proc. Roy. Soc. Land. A 366 (1979) 173183.
[12] W.B. Heginbotham, S.L. Gogia, Metal cutting and the built-up nose, Proc. Inst. Mech. Eng. 175 (1961) 892917.
[13] E. Usui, T. Shirakashi, Mechanics of machining from descriptive to predictive theory, in: On the Art of
Cutting Metals 75 Years Later, ASME PED-Vol. 7, 1982, pp. 1335.
[14] K. Iwata, K. Osakada, Y. Terasaka, Process modeling of orthogonal cutting by the rigid-plastic 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.
[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] K. Komvopoulos, S.A. Erpenbeck, Finite element modeling of orthogonal metal cutting, J. Eng. Ind. 113 (1991)
253267.
[19] T. Tyan, W.H. Yang, Analysis of orthogonal metal cutting processes, Int. J. Num. Meth. Eng. 34 (1992) 365389.
[20] A.J. Shih, H.T.Y. Yang, Experimental and nite element predictions of the residual stresses due to orthogonal metal
cutting, Int. J. Num. Meth. Eng. 36 (1993) 14871507.
[21] A.J. Shih, Finite element simulation of orthogonal metal cutting, J. Eng. Ind. 117 (1995) 8493.
[22] A.J. Shih, Finite Element Analysis of orthogonal metal cutting mechanics, Int. J. Mach. Tools Manuf. 36 (1996)
255273.
[23] A.J. Shih, Finite Element Analysis of the rake angle e#ects in orthogonal metal cutting, Int. J. Mech. Sci. 38 (1996)
117.
[24] 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.
[25] S. Lei, Y.C. Shin, F.P. Incropera, Thermomechanical modeling of orthogonal machining process by nite element
analysis, Int. J. Machine Tools Manuf. 39 (1999) 731750.
[26] G. Shi, X. Deng, Finite element simulations of orthogonal metal cutting processes, Presented at the Joint
ASME=ASCE=SES Summer Mechanics Meeting, June 29 July 2, Northwestern University, Evanston, Illinois, 1997.
[27] Z.C. Lin, S.Y. Lin, A coupled nite element model of thermo-elasticplastic large deformation for orthogonal cutting,
J. Eng. Mater. Technol. 114 (1992) 218226.
[28] 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.
[29] S. Kalpakjian, Manufacturing Processes for Engineering Materials, 3rd Edition, Menlo Park, Addison-Wesley,
California, 1997.
[30] D. Lee, The e#ect of cutting speed on chip formation under orthogonal machining, J. Eng. Ind. 107 (1985) 5563.

Вам также может понравиться