Sie sind auf Seite 1von 11

Extension of Oxley’s Analysis

of Machining to Use Different


Material Models
The aim of the present work is to extend the applicability of Oxley’s analysis of machining
to a broader class of materials beyond the carbon steels used by Oxley and co-workers.
The Johnson-Cook material model, history dependent power law material model and the
Mechanical Threshold Stress (MTS) model are used to represent the mechanical proper-
Amir H. Adibi-Sedeh ties of the material being machined as a function of strain, strain rate and temperature. A
few changes are introduced into Oxley’s analysis to improve the consistency between the
Vis Madhavan various assumptions. A new approach has been introduced to calculate the pressure
e-mail: vis.madhavan@Wichita.edu variation along the alpha slip lines in the primary shear zone including the effects of both
the strain gradient and the thermal gradient along the beta lines. This approach also has
Behnam Bahr the added advantage of ensuring force equilibrium of the primary shear zone in a mac-
roscopic sense. The temperature at the middle of the primary shear zone is calculated by
College of Engineering, Wichita State University, integrating the plastic work thereby eliminating the unknown constant ␩. Rather than
Wichita, KS 67260-0035 calculating the shear force from the material properties corresponding to the strain, strain
rate and temperature of the material at the middle of the shear zone, the shear force is
calculated in a consistent manner using the energy dissipated in the primary shear zone.
The thickness of the primary and secondary shear zones, the heat partition at the primary
shear zone, the temperature distribution along the tool-chip interface and the shear plane
angle are all calculated using Oxley’s original approach. The only constant used to fine
tune the model is the ratio of the average temperature to the maximum temperature at the
tool-chip interface (␺). The performance of the model has been studied by comparing its
predictions with experimental data for 1020 and 1045 steels, for aluminum alloys 2024-
T3, 6061-T6 and 6082-T6, and for copper. It is found that the model accurately repro-
duces the dependence of the cutting forces and chip thickness as a function of undeformed
chip thickness and cutting speed and accurately estimates the temperature in the primary
and secondary shear zones. 关DOI: 10.1115/1.1617287兴

1 Introduction Merchant 关4兴, using energy minimization, found that this angle
could be determined if the average friction angle at the tool-chip
The physics of metal cutting involves complex interactions
interface were known. Various other researchers have developed
among a multitude of phenomena such as plasticity, friction, heat
2D models of machining by the use of slip line field theory 关5–7兴
generation, heat flow, material damage and phase changes that are used in conjunction with Hill’s theory of permissible intensity of
dealt with in separate disciplines such as mechanics, heat transfer, stress singularities as applicable to machining 关8兴. The most sig-
tribology, materials science and mathematics. Since the time of nificant contribution towards modeling of practical machining has
Tresca’s studies on the planing of metals, there have been signifi- been the parallel-sided shear zone theory of Oxley and co-workers
cant improvements in our understanding of machining resulting 关9–12兴. The unique feature of Oxley’s machining theory 关12兴 is
from advances in these disciplines. However, machining still that the dependence of material flow stress upon strain, strain rate
poses ample challenges to researchers. and temperature is considered to obtain the shear angle ␾ and
Machining process development involves determination of in- other outputs of interest. This theory has been modified to analyze
puts such as machine tool operating parameters and the cutting a range of machining operations 关13–15兴 and extensive experi-
tool geometry, material and coating to be adopted in order to mental validation of the theory has been carried out 关16兴. One
obtain desired values of outputs such as cutting force, chip mor- significant limitation of the model is the fact that it has been
phology, machined surface characteristics and tool life. Design almost exclusively applied to carbon steels. There is a need to
and development of machining processes and cutting tools is still extend the applicability of this theory to other materials com-
primarily based on empirical knowledge, with additional experi- monly machined.
mentation carried out as needed. This approach is costly, time
consuming and often leads to non-optimal manufacturing 关1兴.
Continuous enhancements to the accuracy of machining models 2 Description of Oxley’s Machining Theory
and determination of accurate values for critical inputs needed for
these models are prerequisites for scientific design of machining Oxley’s model of machining was developed based on experi-
processes. mental observations 关17兴 of the material deformation. As shown in
Over the past century and a half, many researchers have tried to Fig. 1, the deformation in metal cutting is concentrated in two
understand the machining process under the framework of plastic- zones—a primary shear zone centered about AB 共the nominal
ity theory. Piispanen 关2兴 and Ernst 关3兴 proposed the card model of shear ‘‘plane’’ of length L兲 and a secondary shear zone along the
chip formation in which the shear plane angle ␾ is unknown. tool-chip interface. Though the actual shapes of the two zones are
approximately as depicted in Fig. 1, the primary shear zone is
Contributed by the Manufacturing Engineering Division for publication in the
assumed to be parallel-sided and the secondary shear zone is as-
JOURNAL OF MANUFACTURING SCIENCE AND ENGINEERING. Manuscript received sumed to be of constant thickness, to simplify analysis. The pa-
March 2002; Revised June 2003. Associate Editor: Y. Shin. rameter c is used to represent the relative length of the primary

656 Õ Vol. 125, NOVEMBER 2003 Copyright © 2003 by ASME Transactions of the ASME
minimizes the overall cutting force 关19兴. This is taken to be the
correct value of ␦ and the corresponding shear plane angle, forces,
strain rates and temperatures are taken to be the actual values of
these output variables.
Oxley 关12兴 and co-workers developed a computer program to
carry out the above analysis. It is found that the analysis yields
results in good agreement with experiments when parameters ␩
and ␺ are used to tune the model. Typical values of these param-
eters used by Oxley and co-workers range from 0.5 to 1.0 for
carbon steels.

3 Extension of Oxley’s Machining Analysis to Use


Other Material Models
We have extended Oxley’s model to use any of a range of
material models, namely, the Oxley’s material model for carbon
steels, the Johnson-Cook material model, the history-dependent
power law model used by Maekawa and co-workers, and the MTS
model, so that machining of a wide range of materials can be
analyzed. In the following description of the extension it should
Fig. 1 Model of chip formation used in Oxley’s analysis for be noted that all assumptions and techniques used, other than
orthogonal machining †12‡ those explicitly discussed are the same as in Oxley’s model.
3.1 Description of the Material Models Used. Oxley and
co-workers used the velocity modified temperature concept to de-
shear zone with respect to the thickness of the primary shear zone scribe material properties as a function of strain rate and tempera-
and ␦ is the ratio of the thickness of the secondary shear zone to ture. The velocity modified temperature, defined as
the chip thickness. T mod ⫽ 共 1⫺ ␯ log10共 ␧˙ /␧˙ 0 兲兲 T (1)
In the primary shear zone the slip line along the direction AB
共Fig. 1兲 is an alpha slip line and in the secondary shear zone the increases as the temperature increases and decreases as the strain
slip line along the chip face is a beta slip line. Assuming that the rate increases. ␯ and ␧˙ 0 are constant for a given material. The flow
strain at AB is uniform, equal to one half the strain in the primary stress is related to the strain through the power law ␴
shear zone, and further assuming that the temperature and strain ⫽ ␴ 1 (T mod )␧ n ( T mod ) , where both the strength coefficient and the
rate are uniform along AB, the shear stress along AB is calculated strain hardening exponent are functions of the velocity modified
for a given c. A parameter ␩ is introduced to obtain the tempera- temperature. This leads to a complex relationship between the
ture of the middle of the shear zone as a fraction of the tempera- deformation and the accompanying increase in temperature. In
ture rise through the primary shear zone. The gradient of shear Oxley’s analysis, calculation of the total deformation work is
stress along the beta slip lines is obtained from constitutive equa- avoided and pointwise calculations of the stresses, depending
tion 关12兴 assuming the strain rate is a maximum along AB 关10兴 upon the local strain, strain rate and temperature are used to evalu-
and that the component due to the temperature gradient is negli- ate forces. Though the velocity modified temperature concept was
gible. Using the force equilibrium of an element on the free sur- originally restricted to steels, it has been extended to aluminum
face of work material close to point A the hydrostatic pressure at alloys also 关20,21兴.
point A, p A , is obtained. Applying force equilibrium along the In recent years, an extensive amount of characterization of ma-
shear plane direction and using the gradient of strength perpen- terial properties at high strain rate and temperature has been car-
dicular to the nominal shear plane the normal pressure variation ried out for use in simulation of high velocity impacts. The
along AB can be determined. Knowing the pressure and the shear Johnson-Cook model and the MTS model are widely used consti-
stress along AB, the resultant force, its direction of action and its tutive models for which the coefficients are available for a variety
moment about B 共the tool tip兲 are calculated. of materials 关22,23兴. Another model used widely by Maekawa and
Assuming uniform distribution of the normal stress along the co-workers 关24兴 is a history dependent power law material model.
rake face, the tool-chip contact length is obtained so that the mo- The equation below shows the structure of the Johnson-Cook
ment of the normal force about the point B equals the moment of material model, and lists the coefficients to be used for represent-
the resultant force along the shear plane. The normal stress ␴ n and ing the mechanical properties of materials 关22兴.

冉 冉 冊冊冉 冉 冊冊
the shear stress ␶ int on the rake face are obtained from the corre-
sponding forces and the contact length. The normal stress on the ␧˙ T⫺T r m
␴ ⫽ 共 A⫹B␧ n 兲 1⫹C ln 1⫺ (2)
rake face can be obtained in a different manner by assuming that ␧˙ 0 T m ⫺T r
the alpha slip line in the primary shear zone turns to meet the tool In the above equation, ␴ is the true stress, ␧ is the true strain, ␧˙ is
perpendicular to the rake face. The parameter c is fixed at the the strain rate, ␧˙ 0 is the reference strain rate, T is the temperature
value that causes ␴ n calculated by both ways to be the same 关11兴. of the material, T r is the reference temperature and T m is the
Assuming that there is sticking friction at the tool-chip inter- melting temperature. Brief descriptions of the MTS material
face, the shear strength of the chip (k chip ) at the average tempera- model and the power law material model used by Maekawa and
ture of the tool-chip interface should be equal to ␶ int . The maxi- co-workers are given in the Appendix. It is apparent that Oxley’s
mum temperature at the tool-chip interface is calculated using model can be extended to handle other materials by incorporating
equations derived from Boothroyd’s 关18兴 results. A parameter ␺ is these material models, especially since it is already complex
used to obtain the average temperature at the tool-chip interface, enough that a computer is required to carry out the calculations.
T a v e , from the maximum temperature rise, ⌬ ␪ m . For each as-
sumed shear plane angle, all of these calculations are carried out 3.2 Calculation of the Shear Plane Temperature. In Ox-
to find ␶ int , and the highest value of ␾ for which ␶ int ⫽k chi p is ley’s analysis the heat partition coefficient at the primary shear
taken to be the shear plane angle 关9兴. zone, ␤, is calculated using the nondimensional number R T tan ␾
The calculations above are repeated for various values of ␦, the evaluated at the average temperature along the shear plane, which
relative thickness of the secondary shear zone, to obtain the ␦ that depends on an arbitrary parameter ␩. The overall temperature rise

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 657
of the material in moving through the primary shear zone is found
using the total plastic work in the primary shear zone, accounting
for the heat conducted to the workpiece through the parameter ␤.
It should be noted that the total power is calculated as F s V s , with
F s obtained from the material model for the temperature, strain
and strain rate corresponding to the midplane, not by the integra-
tion of the power required for deformation. The midplane tem-
perature is obtained by scaling the total temperature increase in
the primary shear zone with a parameter ␩. Iteration through the
calculations is used to converge to the correct temperature at the
middle of the shear zone.
Using a slightly different approach, we calculate the tempera-
ture of the midpoint of the shear plane by equating the fraction ␤
of the total plastic work done up to the midplane of the primary
shear zone to the energy expended to increase the temperature of
the material. Point-wise evaluation of the properties of each of the
layers in the primary shear zone is used for the calculation. For
the Johnson-Cook material model the temperature of the shear
plane, T AB , is found from Fig. 2 Temperature variation along the thickness of the pri-
mary shear zone predicted by the modified model for Al 2024-

冕 T AB ␳ C p共 T 兲
冉 B
冊 T3. Cutting conditions: ␣Ä8°, t 1 Ä160 ␮ m, V Ä1.31 mÕs and b 1

冉 冊 m dT⫽ 共 1⫺ ␤ 兲 A␧ AB ⫹ ␧ n⫹1 Ä4.7 mm. ␤Ä0.38, ␧ AB Ä0.63. Average temperatureÄ110.7°C,


TW T⫺T r n⫹1 AB temperature of the middle of the shear zoneÄ117.7°C and
1⫺ ␩Ä0.54.
T m ⫺T r


⫻ 1⫹C ln
␧˙ s
␧˙ 0 冊 (3)
tively. The average shear stress along the shear plane ␶ s is calcu-
where T w , the temperature of the lower part of the shear zone, is lated as the shear force divided by the shear plane area. In Oxley’s
assumed equal to the incoming workpiece material temperature, analysis, the shear strength of the material at the strain and tem-
the partition coefficient ␤ is assumed to be constant and evaluated perature of the midpoint of the shear zone is taken to be the shear
at T AB , ␧ AB ⫽cos ␣/2冑3 cos(␾⫺␣)sin ␾ is the strain in the de- stress at the shear plane, from which the shear force is then cal-
formed material in the middle of the shear plane and ␧˙ s ⫽cV s /L is culated, without any guarantee that energy balance 共Eq. 共6兲兲 will
the strain rate in the primary shear zone assumed to be constant be satisfied. Figure 3 shows the variation of the shear strength
throughout the primary shear zone. The above equation can be through the thickness of the primary shear zone for Al 2024-T3.
solved iteratively or explicitly for T AB . The temperature of the Because of the competition between the temperature and strain
upper part of the shear zone, T EF , can be obtained in a similar there is a maximum in the shear strength along the shear zone.
manner as Also, the value of shear strength at the midplane is larger than the

冕 冉 冊
value of ␶ s obtained from the above energy balance.
T EF ␳ C p共 T 兲 B

冉 冊 m dT⫽ 共 1⫺ ␤ 兲 A␧ EF ⫹ ␧ n⫹1
TW T⫺T r n⫹1 EF
1⫺ 3.4 Calculation of the Hydrostatic Pressure at the Cutting
T m ⫺T r Edge and the Normal Stress on the Tool-Chip Interface.


⫻ 1⫹C ln
␧˙ s
␧˙ 0 冊 (4)
From the equilibrium of the free surface near point A the hydro-
static pressure, p A , can be obtained as

in which ␧ EF ⫽2␧ AB and ␤ is evaluated at T AB . Figure 2 shows


the distribution of the temperature in the primary shear zone as a
function of the distance through the thickness of the zone for Al
2024-T3 obtained from the above relations. Although the average
strain is considered to be at the middle of the shear zone, the
average temperature of the primary shear zone is different from
the temperature of the midplane AB due to the nonlinear depen-
dence of the strength on the strain and temperature of the material.

3.3 Calculation of the Shear Force Along the Primary


Shear Zone. Since the deformation work per unit volume 共w兲
can be obtained as


T
␧ EF 兰 T EF ␳ C p 共 T 兲 dT
w
w⫽ ␴ d␧⫽ (5)
0 共 1⫺ ␤ 兲
we can obtain the shear force, F s , in a consistent manner by
setting
F s V s ⫽wVt 1 b 1
(6) Fig. 3 Variation of shear strength along the thickness of the
F s wVt 1 primary shear zone predicted by the model for Al 2024-T3. Cut-
␶ s⫽ ⫽ ting conditions: ␣Ä8°, t 1 Ä160 ␮ m, V Ä1.31 mÕs and b 1
As V sL
Ä4.7 mm. Note that the shear strength at the midplane of the
where V s is the shear velocity and V, t 1 and b 1 are the cutting shear zone, k AB , is different from the energy equivalent shear
velocity, undeformed chip thickness and workpiece width, respec- strength ␶ s .

658 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME


冉 冉 冊冊
p A ⫽ ␶ s 1⫹2

4
⫺␾ (7)
temperature increase at the tool-chip interface (⌬ ␪ m ). For a rect-
angular secondary shear zone, this ratio is expressed as a function
of the relative thickness of the secondary shear zone 共␦兲, chip
This is similar to Oxley’s expression for p A , but rather than using thickness (t 2 ), length of contact 共H兲 and thermal number (R T ) as
the shear strength at the midplane of shear zone, energy equivalent
shear strength has been used. The requirement for equilibrium of
the shear zone dictates that the hydrostatic pressure at point B in
Fig. 1 be related to the hydrostatic stress at point A by
⌬␪m
⌬␪c
⫽0.06⫺0.195␦
H 冉 冊
R Tt 2 0.5
⫹0.5 log10 冉 冊
R Tt 2
H
(15)

The same approach is adopted here to determine the maximum


p A ⫺p B ⫽ 共 k u ⫺k l 兲 c (8) temperature increase along the tool-chip interface after obtaining
⌬ ␪ c from the friction force at the tool-chip interface and the chip
where k u and k l are the shear strengths at the upper and lower velocity. A constant ␺ is used to relate the mean temperature along
boundaries of the primary shear zone. k u and k l can be obtained the tool-chip interface to the maximum temperature increase along
knowing the temperature, strain and strain rate corresponding to the tool-chip interface as T a v e ⫽T EF ⫹ ␺ ⌬ ␪ m .
the upper and lower boundaries of the primary shear zone. Calcu- Knowing the average temperature, strain and strain rate along
lating p B from the above equation automatically takes into con- the tool-chip interface, the shear strength of the material (k chi p )
sideration the effects of the gradient in strain as well as the gra- can be evaluated. The largest value of ␾ that satisfies the equality
dient in temperature along the alpha slip line, the latter being of the shear strength obtained from the friction force ( ␶ int
neglected in Oxley’s analysis. The angle ␪ between the resultant ⫽F/Hb 1 ) and the one from the constitutive equation (k chi p ) is
force and the direction of the primary shear zone can be obtained chosen to be the correct value of ␾.
using the known pressure distribution and shear stress along the
shear plane.

冋 冉 冊
␪ ⫽tan⫺1 1⫹2

4
⫺␾ ⫺
c 共 k u ⫺k l 兲
2␶s 册 (9)
4 Results and Discussion
4.1 Comparison of Extended Model With Oxley’s Original
Model for 1020 Carbon Steel. We have developed a program
Knowing F s obtained from Eq. 共6兲, the normal and tangential using the Maple symbolic mathematics package to carry out this
components, N and F, of the resultant cutting force on the tool analysis. To verify the implementation of the model and to study
rake face can be obtained following Oxley as the effect of introducing the above changes into Oxley’s original
model for plain carbon steels we also implemented Oxley’s origi-
Fs nal model following the algorithm given by Oxley 关12兴 in all
N⫽ cos共 ␪ ⫺ ␾ ⫹ ␣ 兲
cos ␪ details. Figure 4 shows a comparison of the results of this imple-
(10)
mentation with the results published by Hastings et al. 关16兴 for
Fs 0.2% carbon steel under one particular set of cutting conditions. In
F⫽ sin共 ␪ ⫺ ␾ ⫹ ␣ 兲
cos ␪ all the graphs presented here experimental data points are marked
by symbols and lines represent the model predictions. It is ob-
The slip line field equation describing the relation between the
served that the correlation is very good, but not exact, especially
pressure at point B and the normal stress on the tool-chip interface
at low values of cutting speed. This may be attributed to uncer-
can be written in terms of ␶ s as
tainties in the exact values of particular parameters such as ␺, ␩
␴ n ⫽p B ⫹2 ␶ s 共 ␾ ⫺ ␣ 兲 (11) and c used by Hastings et al. to tune the model.
Figure 4 also compares the performance of the extended model
3.5 Calculations Pertaining to the Secondary Shear Zone. with that of Oxley’s original model. It can be seen that even
The length of contact 共H兲 along the tool-chip interface can be though we have eliminated the degree of freedom provided by the
obtained from moment equilibrium of the chip, using the same parameter ␩, the predictions are slightly improved for cutting
approach adopted by Oxley 关12兴, as force and chip thickness. This is one of the main contributions of

冉 冊
the present work and may be attributed to the self-consistency of
b 1 L 2 2p A ⫹p B
H⫽ (12) the model and the imposition of energy balance in the primary
N 3 shear zone for calculation of the shear force.
where N is the normal force on the tool rake face.
Assuming uniform distribution of the normal force on the tool 4.2 Comparison of Extended Model With Experimental
rake face, the normal stress can be obtained as N/Hb 1 . Equating Data for Other Materials. We have used the extension of Ox-
this and ␴ n calculated from Eq. 共10兲, the parameter c can be ley’s analysis with the modifications described above to analyze
obtained as the machining of aluminum alloys 2024-T3, 6061-T6 and 6082-

c⫽
1
冋 N ␲

⫺ ␶ 1⫹ ⫺2 ␣
k l ⫺k u Hb 1 s 2 冊册 (13)
T6, 1045 steel and copper. The constants of the Johnson-Cook
material model for 1045 steel and Al 6082-T6 were obtained from
Jaspers 关25兴, for Al 2024-T3 from Johnson and Cook 关22兴 and for
Assuming sticking friction over the tool-chip interface, the av- 6061-T6 from Rule 关26兴. The coefficients A, B and n in the
erage value of the strain in the chip material can be considered to Johnson-Cook model were modified such that the model better
be represents the data obtained from compression of those materials
to high strains 关27,25兴. The modified constants of the Johnson-
␥˙ int H Cook material model for each of the materials used in this analy-
␧ int ⫽␧ EF ⫹ (14)
冑3V c sis are listed in Table 1. The constants for the Maekawa’s material
model for 1045 steel and the MTS material model for copper were
in which ␥˙ int ⫽V c / ␦ t 2 is the average strain rate in the secondary obtained from Childs 关24兴 and Follansbee and Kocks 关28兴, respec-
shear zone and V c is the chip velocity. tively. These constants are listed in the appendix.
Stevenson and Oxley 关10兴 obtained Eq. 共15兲 below to represent The experimental data used in the comparison for Al 2024-T3
the results of Boothroyd’s numerical calculations for the ratio of and 6061-T6 were obtained by tube turning 关29兴. The Al 2024-T3
the mean temperature increase of the chip (⌬ ␪ c ) to the maximum tubes had a diameter of 70 mm and a wall thickness of 4.7 mm.

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 659
Table 1 Constants for the Johnson-Cook material model for
the materials used in this study after adjusting to fit stress
strain curve obtained from compression test

Material T m (°C) A (MPa) B (MPa) C n m


Al 2024-T3* 502 325 414 0.015 0.2 1
Al 6061-T6** 582 293.4 121.26 0.002 0.23 1.34
Al 6082-T6*** 582 250 243.6 0.00747 0.17 1.31
AISI 1045**** 1460 553.1 600.8 0.0134 0.234 1

*Constants C, n and m adapted from 关22兴


**Constants C, n and m adapted from 关26兴
***Constants C, n and m adapted from 关25兴
****Adapted from Jaspers 关25兴

fied by the manufacturer. The experimental data for 6082-T6 was


obtained from Jaspers 关25兴.
In all the analyses carried out for obtaining the results reported
below, ␺, the ratio of the average temperature increase along the
tool-chip interface to the maximum temperature increase along
this interface, is set to 0.6. Figure 5 shows the results for Al
2024-T3 as a function of the cutting speed for two different un-
deformed chip thicknesses of 80 ␮m and 160 ␮m and rake angle
of 0°. The predictions show good agreement with the experimen-
tal data. As expected, increasing the cutting speed decreases the
cutting forces and the chip thickness. It is found that increasing
the undeformed chip thickness increases the thickness of the pri-
mary shear zone and decreases the strain rate in the primary shear
zone. For instance, for the undeformed chip thickness and cutting
speed equal to 80 ␮m and 1.31 m/s respectively, doubling the
undeformed chip thickness increases the thickness of the primary
shear zone by 66% and reduces the strain rate in the primary shear
zone by 37%. It is also found that increasing the cutting speed
reduces the thickness of the primary shear zone and increases the
strain rate in the primary shear zone. For instance, for the unde-
formed chip thickness and cutting speed equal to 80 ␮m and 1.31
m/s respectively, doubling the cutting speed decreases the thick-
ness of the primary shear zone by 19% and increases the strain
rate in the primary shear zone by 1.6 times. It can also be seen that
c, the ratio of the length of the shear zone to its thickness, in-
creases with increase in the undeformed chip thickness and cut-
ting speed.
Figure 6 shows that increase in the rake angle decreases the
cutting forces. The model overpredicts the cutting forces for nega-
tive rake angles. It is found that increasing the rake angle in-
creases the thickness of the primary shear zone and decreases the
strain rate in the primary shear zone. For instance, beginning with
rake angle and undeformed chip thickness equal to 0° and 80 ␮m
respectively, increasing the rake angle by 10° increases the thick-
ness of the primary shear zone by 17% and decreases the strain
rate in the primary shear zone by 43%. This effect of the rake
angle seems to be more pronounced for higher undeformed chip
thicknesses. For instance for 160 ␮m undeformed chip thickness
and 1.31 m/s cutting speed, the strain rate is reduced in half when
the rake angle is changed from 0° to 8°. It can also be noted that
increasing the rake angle greatly decreases the c value. Figure 7
shows the capability of the model to capture the size effect. As
shown in Fig. 7共a兲, the specific cutting energy increases as the
Fig. 4 Effect of the new modifications to Oxley’s model on „a… undeformed chip thickness decreases. Figure 7共b兲 shows the
Cutting force, „b… Thrust force, and „c… Chip thickness. Cutting variation in k chi p and ␶ int as a function of the shear plane angle ␾.
conditions: ␣Ä5°, t 1 Ä0.5 mm, b 1 Ä4 mm and 0.2% carbon It can be seen that the shear plane angle, which is the angle where
steel. It can be seen that even though we have eliminated the k chi p ⫽ ␶ int , increases as the undeformed chip thickness increases.
degree of freedom provided by the parameter ␩, the predictions It can also be seen that while the shear stress required for force
are slightly improved for cutting force and chip thickness. balance ( ␶ int ) at any given ␾ is just a little lower for the higher
undeformed chip thickness 共due to variation in ␥˙ int ), the shear
strength (k chi p ) of the chip material corresponding to the strain,
The diameter and the wall thickness for 6061-T6 tubes were 57 strain rate and temperature at the secondary shear zone is much
mm and 3.3 mm respectively. The inserts used were Kennametal smaller at any given ␾ for larger undeformed chip thicknesses due
grade K68 with an edge radius of approximately 10 ␮m as speci- to the much higher temperature at the secondary shear zone.

660 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME


Fig. 6 Comparison of predicted „a… Cutting force, „b… Thrust
force with experimental data for Al 2024-T3 †29‡. Cutting condi-
tions: V Ä1.31 mÕs and b 1 Ä4.1 mm, different rake angles and
undeformed chip thicknesses.

of c obtained for Al 6082-T6 are between those for Al 6061-T6


and Al 2024-T3. For instance for ␣⫽8°, undeformed chip thick-
ness equal to 160 ␮m and cutting speed of 1.31 m/s, the values of
c obtained from the model are 9.22, 3.68 and 2.98 for Al 6061-T6,
Al 6082-T6 and Al 2024-T3 respectively. For Al 6061-T6 and Al
6082-T6, increase in the cutting speed and undeformed chip thick-
ness both lead to decrease in the c value, while a similar change
causes an increase in the c value for Al 2024-T3. With respect to
the thickness of the primary shear zone the situation is little dif-
ferent; the greatest and the least values are found for Al 6082-T6
and Al 6061-T6 respectively. The decrease in the thickness of the
primary shear zone (L/c) and increase in the strain rate with
increase in the cutting speed and decrease in the undeformed chip
Fig. 5 Comparison of predicted „a… Cutting force, „b… Thrust thickness obtained in the case of Al 2024-T3 are also found to be
force, and „c… Chip thickness with experimental data for Al the case for Al 6061-T6 and Al 6082-T6. While the ratio of the
2024-T3 †29‡. Cutting conditions: ␣Ä0° and b 1 Ä4.7 mm, differ- length to the thickness of the primary shear zone 共c兲 tends to
ent cutting speeds and undeformed chip thicknesses. highlight changes in material properties, the trends in the thick-
ness of the primary shear zone and the average strain rate through
this zone are dependent only upon the cutting conditions.
Figures 8 and 9 show the variation of the cutting forces and Figure 10 shows the comparison between the shear plane tem-
chip thickness with cutting speed while cutting Al 6061-T6 and Al perature predicted using the model and the experimental values
6082-T6 respectively. The model predictions are in good agree- 关25兴 obtained using an infrared camera to measure the temperature
ment with experimental data. It was found that under the same of point A in Fig. 1. The experimental value is found to be be-
cutting conditions the values of c obtained for Al 6061-T6 and Al tween T AB and T EF predicted by the model, but closer to T EF .
2024-T3 are the greatest and the least respectively, and the values Figure 11 shows predictions of the specific cutting energy and

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 661
Fig. 7 „a… Comparison of the size effect predicted by the
model with the experimental data for Al 2024-T3 †29‡, and „b…
The decrease in specific cutting energy with increase in unde-
formed chip thickness can be attributed to the increase in
shear angle, caused by the increase in temperature at the tool-
chip interface and the consequent decrease in k chip . Cutting
conditions: ␣Ä8°, V Ä2.62 mÕs and b 1 Ä4.1 mm.

shear plane angle for cutting 1045 steel with different undeformed
chip thicknesses using different material models. The size effect
can again be seen in Fig. 11共a兲. It can also be seen that the
Johnson-Cook material model predicts the variation of the specific
cutting force more accurately compared to the Oxley material
model and the Maekawa material model, especially for the larger
undeformed chip thicknesses. It can be seen from Fig. 11共b兲 that
while all the models tend to underpredict the shear plane angle,
the Johnson-Cook material model leads to predictions closest to Fig. 8 Comparison of predicted „a… Cutting force, „b… Thrust
the experimental data. Note that in this case Oxley’s original force, and „c… Chip thickness with experimental data for Al
model predicts specific cutting force and the shear plane angle 6061-T6 †29‡. Cutting conditions: ␣Ä8°, b 1 Ä3.3 mm, different
better compared to the modified model with different material cutting speeds and undeformed chip thicknesses.
models but the difference is of the same order of magnitude as the
variation in the experimental data.
Figure 12 shows predictions of the cutting force and tempera- speed. Figure 12共b兲 shows the effect of cutting speed on the tool-
ture at the tool-chip interface for cutting 1045 steel with different chip interface temperature. The temperature was measured experi-
cutting speeds using different material models. As can be seen in mentally 关30兴 using the tool-work thermocouple technique. The
Fig. 12共a兲, the Johnson-Cook material model predicts the cutting measured values of temperature are between the maximum tem-
force most accurately. It can be noted that the cutting force pre- perature and the average temperature at the tool-chip interface.
dicted by the Maekawa material model is almost constant for dif- The experimental data is closer to the average temperature for the
ferent cutting speeds. Also, Oxley’s material model gives good Oxley and Maekawa material models, whereas it is closer to the
agreement with experimental data only for low values of cutting maximum temperature for the Johnson-Cook material model for

662 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME


Fig. 10 Comparison of the temperature of the shear plane with
experimental data †25‡ for Al 6082-T6 for different undeformed
chip thicknesses. Cutting conditions: ␣Ä6°, V Ä6 mÕs and dif-
ferent undeformed chip thicknesses. Note that the measured
temperature seems to be close to the temperature at the end of
the primary shear zone.

measured using the tool-work thermocouple is not necessarily the


average temperature at the tool-chip interface and it depends on
the thermo-electrical properties of the tool. Note that if Oxley’s
original model were to match the experimentally measured tem-
perature closely, the cutting force would be underestimated. Also
note that the data shown in Fig. 11 are for a cutting speed of 7
m/s, which is beyond the range presented in Fig. 12. Thus it is
expected that if a similar comparison to that shown in Fig. 11
were made at any of the cutting speeds presented in Fig. 12, the
comparisons would be more in favor of the modified
models.
Figure 13 shows the variation of the cutting forces with cutting
speed for commercially pure copper. Since experimental values of
the forces do not reach steady state, representative averages of the
recorded forces are presented. Also the average strain at the tool-
chip interface was assumed to be a constant factor of the strain of
the deformed material leaving the primary shear zone. This con-
stant value was calibrated for each undeformed chip thickness for
a single cutting speed and the obtained value was used in calcu-
lations for the same undeformed chip thickness and different cut-
ting speeds. The average strain at the tool-chip interface for 100
␮m and 200 ␮m with different cutting speeds obtained this way
were 2.3␧ EF and 3.1␧ EF respectively. The predictions of the ex-
tended model using MTS material model show good agreement
with the experimental data for different cutting speeds and unde-
formed chip thicknesses and as expected, increasing the cutting
velocity decreases the cutting forces.

5 Conclusions
Oxley’s machining theory has been extended to use any of a
range of high strain rate constitutive models commonly used by
Fig. 9 Comparison of predicted „a… Cutting force, „b… Thrust the impact physics community and for which the constants are
force, and „c… Chip thickness with experimental data for Al available for a wide range of materials. In the process, changes
6082-T6 †25‡. Cutting conditions: ␣Ä8°, different cutting speeds have been made to the theory to improve its self-consistency. The
and undeformed chip thicknesses. extended model has been used to analyze the machining of steel,
aluminum and copper. The model predicts all experimentally ob-
served trends well, which is quite remarkable considering the fact
higher cutting velocities. It can also be seen that temperatures that the model parameters are typically obtained from experiments
predicted by Maekawa’s material model are more sensitive with at much lower values of strain, strain rate and temperature than
respect to the cutting speed and follow the trend in experimental that encountered in machining. Among different material model
data more closely compared to the other material models. It used the Johnson-Cook and Maekawa’s material models perform
should be noted that Stephenson 关32兴 found that the temperature best in terms of prediction of cutting forces and temperature re-

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 663
Fig. 12 Comparison predicted „a… Cutting force, and „b… Tem-
Fig. 11 Comparison of predicted „a… Specific cutting force, perature of the tool-chip interface using different material mod-
and „b… Shear plane angle using different material models with els with Oxley’s original model †31‡ and experimental data †30‡
Oxley’s original model †12‡ and experimental data †12‡ for for 0.45% carbon steel. Cutting conditions: ␣Ä5°, t 1 Ä0.2 mm,
0.45% carbon steel. Cutting conditions: ␣ÄÀ5°, V Ä7 mÕs and b 1 Ä3 mm and different cutting speeds. Oxley’s original model
different undeformed chip thicknesses. Oxley’s original model underpredicts both cutting force and temperature along the
performs better compared to modified model with different ma- tool-chip interface. The Johnson-Cook and Maekawa material
terial models. Among different material models used the models perform better in prediction of the cutting force and
Johnson-Cook material model shows the most sensitivity with temperature respectively.
respect to changes in undeformed chip thickness.

Though the modified Oxley’s model has been shown to suc-


spectively. This provides a means of support to the suggestion cessfully predict a wide range of experimental observations for a
made by some researchers that machining itself could be used as a range of materials, it is not the case that the model is complete
very high strain, strain rate and temperature material test to gen- from a point of view of understanding machining. Questions re-
erate coefficients for material models. main as to the role of energy minimization and the validity of
The predictions of the model show that increasing the cutting different assumptions used. A systematic series of studies is cur-
speed and decreasing the undeformed chip thickness both de- rently underway to further refine Oxley’s machining theory. This
crease the thickness of the primary shear zone and increase the includes a study of the sensitivity of the model to each of the
strain rate. Also, increasing the rake angle increases the thickness assumptions used 关35兴. Another study is aimed at refining the
of the primary shear zone and decreases the strain rate. It is assumptions based upon results of finite element analysis of ma-
also found that the ratio of the length to the thickness of the chining using the same material and friction models used in the
primary shear zone 共c兲 increases for Al 2024-T3 with increase in theory 关36兴. As mentioned by Oxley, better models of the geom-
cutting speed and the undeformed chip thickness while a similar etry of the secondary shear zone and the normal and frictional
change causes a decrease in the c value for Al 6061-T6 and stress distribution at the tool-chip interface will help improve the
Al 6082-T6. accuracy of the analysis and lead to improved understanding of
We believe that this model of orthogonal cutting, coupled with machining.
recent progress in the analysis of 3D machining operations with
arbitrary cutting edge geometry 关34兴 may allow the development Acknowledgment
of a software package that can provide a quick estimate of the The authors would like to thank Thomas Samuel and Sivaku-
cutting forces, chip thickness and tool temperature for a variety of mar Balasubramanian for providing us the results of experiments
machining processes. conducted as part of their M.S. theses.

664 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME


m ⫽ temperature exponent in the Johnson-Cook material
model
N ⫽ normal force at the tool-chip interface
n ⫽ strain hardening exponent in Oxley’s and the
Johnson-Cook material models
p A ⫽ hydrostatic pressure at the free surface of the nominal
shear plane
p B ⫽ hydrostatic pressure at the tool tip
R T ⫽ thermal number of work material
T ⫽ temperature
T a v e ⫽ average temperature at the tool-chip interface
T mod ⫽ velocity modified temperature
T M ⫽ melting temperature
T r ⫽ reference temperature
T w ⫽ temperature of the uncut work material
t 1 ⫽ undeformed chip thickness
t 2 ⫽ chip thickness
V ⫽ cutting velocity
V s ⫽ shear velocity
V c ⫽ chip velocity
w ⫽ deformation work per unit volume in the primary
shear zone
␣ ⫽ normal rake angle
␤ ⫽ heat partition coefficient
␦ ⫽ the ratio of the secondary shear zone thickness to the
chip thickness
␥˙ int ⫽ shear strain rate at the tool-chip interface
␧ ⫽ strain
␧ AB ⫽ strain at the midplane of the primary shear zone
␧ EF ⫽ strain at the upper boundary of the primary shear
zone
␧˙ ⫽ strain rate
␧˙ 0 ⫽ reference strain rate
␧˙ s ⫽ strain rate in the primary shear zone
␾ ⫽ shear plane angle
␩ ⫽ a constant calibration factor in Oxley’s model equal
to the ratio of the work completed till the midplane
of the primary shear zone to the total plastic work
done in the primary shear zone
␯ ⫽ the strain rate exponent in the velocity modified tem-
Fig. 13 Comparison of predicted „a… Cutting force, and „b… perature
Thrust force with experimental data †33‡ for copper. Cutting ␪ ⫽ the angle of the resultant force on the shear plane
conditions: ␣Ä8°, b 1 Ä1.17 mm, different cutting speeds and with respect to the shear plane direction
undeformed chip thicknesses. ⌬ ␪ c ⫽ mean temperature rise of the chip due to the second-
ary shear
⌬ ␪ m ⫽ maximum temperature rise along the tool-chip inter-
Nomenclature face
␳ ⫽ density
A ⫽ yield strength in the Johnson-Cook material model ␴ ⫽ flow stress
A s ⫽ area of the primary shear zone ␶ int ⫽ shear strength at the tool-chip interface obtained from
B ⫽ strength coefficient in the Johnson-Cook material equilibrium
model ␶ s ⫽ energy based shear strength of the primary shear
b 1 ⫽ width of the workpiece zone
C ⫽ strain rate constant in the Johnson-Cook material ␺ ⫽ a constant calibration factor equal to the ratio of the
model average temperature increase to the maximum tem-
c ⫽ ratio of the length to the thickness of the primary perature rise at the tool-chip interface
shear zone
C p ⫽ specific heat of work material
F ⫽ friction force at the tool-chip interface
F s ⫽ shear force along the nominal shear plane Appendix
H ⫽ length of contact between tool and chip
k AB ⫽ shear strength at the middle of the primary shear Mechanical Threshold Stress „MTS… Model. In this mate-
zone rial model the flow stress ␴ is expressed as a function of a refer-
k l ⫽ shear strength at the lower boundary of the primary ence stress ␴ˆ , the mechanical threshold stress, or flow stress at the
shear zone absolute zero temperature. ␴ˆ can be resolved into two components
k u ⫽ shear strength at the upper boundary of the primary such that ␴ˆ ⫽ ␴ˆ a ⫹ ␴ˆ t . The athermal stress ␴ˆ a characterizes the
shear zone rate independent interactions of dislocations with long-range bar-
k chip ⫽ shear strength at the tool-chip interface obtained from riers such as grain boundaries and ␴ˆ t characterizes the rate depen-
the material model dent interactions with short-range obstacles. The general form of
L ⫽ length of the shear plane the flow stress ␴ can be expressed as ␴ ⫽ ␴ˆ a ⫹s(␧˙ ,T) ␴ˆ t , in which

Journal of Manufacturing Science and Engineering NOVEMBER 2003, Vol. 125 Õ 665
s is a function of strain rate and temperature. The specific form of 关7兴 Fang, N., Jawahir, I. S., and Oxley, P. L. B., 2001, ‘‘A Universal Slip-Line
the mechanical threshold model used in this study can be written Model With Non-Unique Solutions for Machining With Curled Chip Forma-
as 关28兴 tion and a Restricted Contact Tools,’’ Int. J. Mech. Sci., 43, pp. 557–580.

再 冋 册冎
关8兴 Hill, R., 1954, ‘‘The Mechanics of Machining: A New Approach,’’ J. Mech.
1/q 1/p Phys. Solids, 3, pp. 47–53.
kT ln共 ␧˙ 0 /␧˙ 兲 关9兴 Fenton, R. G., and Oxley, P. L. B., 1968 – 69, ‘‘Mechanics of Orthogonal
␴ ⫽ ␴ˆ a ⫹ 共 ␴ˆ ⫺ ␴ˆ a 兲 1⫺ (16)
g 0␮ b 3 Machining: Allowing for the Effects of Strain-Rate and Temperature on Tool-
Chip Friction,’’ Proc. Inst. Mech. Eng., 183, pp. 417– 438.
where the flow stress ␴ is expressed as a function of the current 关10兴 Stevenson, M. G., and Oxley, P. L. B., 1969–1970, ‘‘An Experimental Inves-
threshold stress, ␴ˆ , temperature, T, and strain rate ␧˙ . k is the tigation of the Influence of Speed and Scale on the Strain-Rate in a Zone of
Intense Plastic Deformation,’’ Proc. Inst. Mech. Eng., 184共31兲, pp. 561–576.
Boltzmann’s constant, ␧˙ 0 is a constant, g 0 is the activation energy, 关11兴 Oxley, P. L. B., and Hastings, W. F., 1977, ‘‘Predicting the Strain-Rate in the
␮ is the shear modulus, b is the magnitude of the Burgers vector, Zone of Intense Shear in Which the Chip is Formed in Machining From the
and p and q are constants that represent the energy profile of Dynamic Flow Stress Properties of the Work Material and the Cutting Condi-
obstacles. The strain hardening rate can be obtained as tions,’’ Proc. R. Soc. London, Ser. A, 356, pp. 395– 410.

冋 冊冊 册
关12兴 Oxley, P. L. B., 1989, The Mechanics of Machining: An Analytical Approach

d ␴ˆ
⫽ ␪ 0 1⫺
tanh 2 冉冉 ␴ˆ ⫺ ␴ˆ a
␴ˆ s ⫺ ␴ˆ a
(17)
to Assessing Machinability, E. Horwood, Chichester, England.
关13兴 Lin, G. C. I., and Oxley, P. L. B., 1972, ‘‘Mechanics of Oblique Machining:
Predicting Chip Geometry and Cutting Forces From Work Material Properties
d␧ tanh共 2 兲 and Cutting Conditions,’’ Proc. Inst. Mech. Eng., 186, pp. 813– 820.
关14兴 Arsecularatne, J. A., Mathew, P., and Oxley, P. L. B., 1995, ‘‘Prediction of
In the case of OFE copper the corresponding values of parameters Chip Flow Direction and Cutting Forces in Oblique Machining With Nose
used are as follows 关28兴 Radius Tools,’’ Proc. Inst. Mech. Eng., B209, pp. 305–315.
关15兴 Young, H. T., Mathew, P., and Oxley, P. L. B., 1994, ‘‘Predicting Cutting
␪ 0 ⫽2390⫹12 ln ␧˙ ⫹0.034␧˙ (18) Forces in Face Milling,’’ Int. J. Mach. Tools Manuf., 34共6兲, pp. 771–783.
关16兴 Hastings, W. F., Mathew, P., and Oxley, P. L. B., 1980, ‘‘Machining Theory for

冉 冊 kT/ ␮ b 3 A Predicting Chip Geometry, Cutting Forces, etc., From Work Material Proper-
␧˙ ties and Cutting Conditions,’’ Proc. R. Soc. London, Ser. A, 371, pp. 569–587.
␴ˆ s ⫽ ␴ˆ so (19) 关17兴 Palmer, W. B., and Oxley, P. L. B., 1959, ‘‘Mechanics of Metal Cutting,’’ Proc.
␧˙ so
Inst. Mech. Eng., 173, pp. 623– 654.
p⫽2/3, q⫽1, ␴ˆ a ⫽40 MPa, ␧˙ 0 ⫽107 sec⫺1 , 关18兴 Boothroyd, G., 1963, ‘‘Temperatures in Orthogonal Metal Cutting,’’ Proc. Inst.
Mech. Eng., 177, pp. 789– 802.
关19兴 Hastings, W. F., and Oxley, P. L. B., 1976, ‘‘Minimum Work as a Possible
␧˙ so ⫽6.2⫻1010 sec⫺1 , ␮ ⫽42 GPa, Criterion for Determining the Frictional Conditions at the Tool/Chip Interface
in Machining,’’ Philos. Trans. R. Soc. London, 282, pp. 565–584.
k/b ⫽0.823 MPa/K,
3
A 1 ⫽0.31, g 0 ⫽1.6 and 关20兴 Bao, H., and Stevenson, M. G., 1976, ‘‘A Basic Mechanism for Built-up Edge
Formation in Machining,’’ CIRP Ann., 25共1兲, pp. 53–57.
␴ˆ so ⫽900 MPa 关21兴 Kristyanto, B., Mathew, P., and Arsecularatne, J. A., 2000, ‘‘Determination of
Using the above values Eqs. 共16兲 and 共17兲 can be solved simulta- Material Properties of Aluminum From Machining Tests,’’ ICME 2000—
Eighth Int. Conf. On Manuf. Eng., Sydney, Australia, August 27–30.
neously to obtain the flow stress ␴. 关22兴 Johnson, G. J., and Cook, W. H., 1983, ‘‘A Constitutive Model and Data for
Metals Subjected to Large Strains, High Strain Rates and High Temperatures,’’
Maekawa’s Material Model for AISI 1045. Maekawa and Proceedings of the 7th International Symposium on Ballistics, pp. 541–547.
co-workers 关24兴 proposed a history dependent material model for 关23兴 Tanner, A. B., McGinty, R. D., and McDowell, D. L., 1999, ‘‘Modeling Tem-
0.45% carbon steel as follows for which the strength ␴ is ex- perature and Strain Rate History Effects in OFHC Cu,’’ Int. J. Plast., 15, pp.
pressed in MPa. 575– 603.
关24兴 Private communication between Maekawa and Childs as appeared in Childs, T.

␴ ⫽A 2 e aT 冉 冊 冉冕
␧˙
1000
M ⫹m 1

strain path
e ⫺aT/N 1 冉 冊␧˙
1000
⫺m 1 /N 1
d␧ 冊 N1 H. C, ‘‘Material Property Requirements for Modeling Metal Machining,
1997,’’ Colloque C3, Journal de physique, III: XXI–XXXIV.
关25兴 Jaspers, S. P. F. C., 1999, ‘‘Metal Cutting Mechanics and Material Behavior,’’
PhD thesis, Technische Universiteit Eindhoven.
(20) 关26兴 Rule, W. K., 1997, ‘‘Numerical Scheme for Extracting Strength Model Coef-
where: ficients From Taylor Test Data,’’ Int. J. Impact Eng., 19, pp. 797– 810.
关27兴 Kobayashi, S., and Thomsen, E. G., 1959, ‘‘Some Observations on the Shear-
2
A 2 ⫽1350e ⫺0.0011T ⫹167e ⫺0.00006共 T⫺275兲 (21) ing Process in Metal Cutting,’’ ASME J. Ind., pp. 251–261.
关28兴 Follansbee, P. S., and Kocks, U. F., 1988, ‘‘A Constitutive Description of the
⫺0.000015共 T⫺340兲 2 Deformation of Copper Based on the Use of the Mechanical Threshold Stress
N 1 ⫽0.17e ⫺0.001T ⫹0.09e (22) as an Internal State Variable,’’ Acta Metall., 36, pp. 81–93.
关29兴 Samuel, T., 2000, ‘‘Investigation of the Effectiveness of Tool Coatings in Dry
M ⫽0.036 Turning of Aluminum Alloys,’’ MS thesis, Wichita State University.
关30兴 OECD-CIRP, 1966, Proceedings of the Seminar on Metal Cutting.
a⫽0.00014 关31兴 Hastings, W. F., Oxley, P. L. B., and Stevenson, M., 1974, ‘‘Predicting Cutting
Forces, Tool Life etc., Using Work Material Flow Stress Properties Obtained
m 1 ⫽0.0024. From High-Speed Compression Tests,’’ Proc. Int. Conf. on Production Engi-
neering, Tokyo, p. 528.
关32兴 Stephenson, D. A., 1991, ‘‘Tool-Work Thermocouple Temperature Measure-
References ments: Theory and Implementation Issues,’’ PED-Vol 55, Sensors and Signal
关1兴 Armarego, E. J. A., 1998, ‘‘A Generic Mechanics of Cutting Approach to Processing for Manufacturing, ASME 1992, pp. 81–95.
Predictive Technological Performance Modeling of the Wide Spectrum of Ma- 关33兴 Balasubramanian, S., 2001, ‘‘The Investigation of Tool-Chip Interface Condi-
chining Operations,’’ Mach. Sci. Technol., 2共2兲, pp. 191–211. tions Using Carbide and Sapphire Tools,’’ MS thesis, Wichita State University.
关2兴 Piispaanen, V., 1948, ‘‘Theory of Formation of Metal Chips,’’ J. Appl. Phys., 关34兴 Adibi-Sedeh, A. H., Madhavan, V., and Bahr, B., 2002, ‘‘Upper Bound Analy-
19, pp. 876 – 881. sis of Oblique Cutting With Nose Radius Tools,’’ Int. J. Mach. Tools Manuf.,
关3兴 Ernst, H., 1938, ‘‘Physics of Metal Cutting, in Machining of Metals,’’ Ameri- 42共9兲, pp. 1081–1094.
can Society for Metals, Cleveland, Ohio., pp. 1–34. 关35兴 Adibi-Sedeh, A. H., and Madhavan, V., 2002, ‘‘Effect of Some Modifications
关4兴 Merchant, M. E., 1945, ‘‘Mechanics of the Metal Cutting Processes,’’ ASME J. to Oxley’s Machining Theory and the Applicability of Different Material Mod-
Appl. Mech., 11, pp. A168 –A175. els,’’ Mach. Sci. Technol., 6共3兲, pp. 377–393.
关5兴 Lee, E. H., and Shaffer, B. W., 1951, ‘‘The Theory of Plasticity Applied to a 关36兴 Adibi-Sedeh, A. H., and Madhavan, V., 2003, ‘‘Understanding of Finite Ele-
Problem of Machining,’’ ASME J. Appl. Mech., 73, pp. 405– 413. ment Analysis Results Under the Framework of Oxley’s Machining Model,’’
关6兴 Dewhurst, P., 1978, ‘‘On the Non-Uniqueness of the Machining Process,’’ Proceedings of the 6th CIRP International Workshop on Modeling of Machin-
Proc. R. Soc. London, Ser. A, 360, pp. 587– 610. ing Operations, Hamilton, Canada, May 20, pp. 1–15.

666 Õ Vol. 125, NOVEMBER 2003 Transactions of the ASME

Das könnte Ihnen auch gefallen