Sie sind auf Seite 1von 29

Machining Science and Technology, 13:488–515

Copyright © 2009 Taylor & Francis Group, LLC


ISSN: 1091-0344 print/1532-2483 online
DOI: 10.1080/10910340903451506

EXPERIMENTAL STUDIES AND MODELING OF HEAT


GENERATION IN METAL MACHINING

A. Liljerehn, V. Kalhori, and M. Lundblad


AB Sandvik Coromant, Sandviken, Sweden

 Heat generation in the cutting zones due to plastic deformation and friction in the cutting
region governs insert wear, tensile residual stresses on the machined component surface and may
give rise to undesired tolerances and short component life. Therefore, it is crucial that the heat
generation is kept under control during metal cutting. In this study an analytical model for
prediction of heat generation in the primary and secondary deformation zones is compared with
results from finite element simulations and temperature measurements using IR-CCD camera.
The used cutting data are altered to study the temperature influence from tool geometry and
feed when machining stainless steel SANMAC316L and low carbon steel AISI 1045.

Keywords analytical temperature prediction, finite element modeling, infrared


charge coupled device IR-CCD, material modeling, metal machining

INTRODUCTION

Metal cutting is one of the most common manufacturing processes


used to achieve the desired shape and dimension of components. Standing
demands for higher productivity and better life for processed components
and cutting tools have historically pushed further development of
existing tooling concepts. This is to achieve better production economics
and manufacturability, and thereby higher competitive force in the
market. However, increased productivity target must not jeopardize the
component’s quality in terms of dimension accuracy, surface integrity and
longevity.
It is widely known that heat has a major impact on the component’s
characteristics and life of cutting tools. It may give rise to undesired
tolerance deviation, tensile residual stresses at the workpiece surface and
governs insert wear rate. Therefore, it is of great interest to understand the
mechanisms that cause heat development in the metal cutting, and thus

Address correspondence to A. Liljerehn, AB Sandvik Coromant, Sandviken SE-811 81, Sweden.


E-mail: anders.c.liljerehn@sandvik.com
Modeling Heat Generation in Metal Machining 489

would, if possible, reduce or control the heat development in cutting zones.


The generated heat at the cutting zone is mainly governed by elastic-plastic
strain energy dissipated at the primary, secondary and tertiary deformation
zones and friction between chip and insert on the secondary deformation
zone. The heat developments in cutting are strongly dependent on the
cutting process parameters and insert geometry. In this study, the thermo-
mechanical behavior of the insert-chip interface and the shear plane is
investigated. An analytical model is applied and the results are compared
with numerical finite element simulations and experimental measurements
using infrared CCD method.

EXPERIMENTAL SETUP

Insert Geometry and Cutting Parameters


Several different cutting geometries are selected to study its effect on
heat generation within the cutting zones. These are the cutting inserts
TCMT 16T 308-MR for roughing and TCMT 16T 308-MM for medium
roughing and TPNG 160 308 with no chip breaking geometry and a rake
angle of 3◦ , manufactured at Sandvik Coromant (Table 1). The tool grade
used was a PVD coated carbide (GC1025), where the coating consists
of a thin outside layer of TiN and a thicker inner layer of TiAlN, with
a total coating thickness of 4 m. The cutting speed was chosen to be
200 m/min for both workpiece materials, SANMAC 316L and AISI 1045.
SANMAC 316L is an improved AISI 316L stainless steel with respect to
manufacturability, produced by Sandvik Material Technology.

Experimental Setup Using Infrared CCD


The infrared charge coupled device, IR-CCD, technique was used
to measure the temperature at the cutting region. The experimental
setup for temperature measurement in turning, performed at the Swedish
Institute for Metal Research (SIMR), is illustrated in Figure 1. The CCD

TABLE 1 Workpiece Material and Cutting Insert Combinations for the


Machining Tests

Test Workpiece material Cutting insert

1 SANMAC 316L TCMT16T308-MR


2 SANMAC 316L TCMT16T308-MM
3 SANMAC 316L TPNG160308
4 AISI 1045 TCMT16T308-MR
5 AISI 1045 TCMT16T308-MM
6 AISI 1045 TPNG160308
490 A. Liljerehn et al.

FIGURE 1 Experimental IR-CCD camera setup for tool temperature measurement in orthogonal
cutting.

components used in the present camera model are made of silicon-based


semiconductors that act like condensers. During the exposure phase to
the light flux, the condenser accumulates a charge, which depends on
the number of photons collected by the semiconductor. During the image
reading sequence, the charge of each condenser is transferred to the
neighboring one under the application of an external electrical field.
The charges are then transferred from one element to the other. Finally
a diode located at the end of the measuring chain permits to read
the charge number and transfer the information as a form of a video
signal. The signal is then directly digitized and transferred to the image
memory of a computer controlled with suitable software. The number of
effective picture elements 752 × 582 (H × V) ensure the possibility to have
a 64 × 48 mm sensing area. The normally occurring thermal noise of the
CCD sensors is here eliminated by a Peltier cooling system, ensuring a
repeatable image quality.

Application of the System to the Detection


of Infrared Radiation
The Si-based CCD sensors classically used in black and white image
analysis are sensitive both to visible (400–800 nm) and near infrared
(800–1100 nm) radiation. The present type of sensors is optimized in the
near infrared range.
The adjustment of an object can be done in the visible range, using
an IR-block filter. The fine adjustment of the image can be done in
the near-IR range through replacing the IR block filter with an IR filter
Modeling Heat Generation in Metal Machining 491

(800 or 1000 nm) that will cut the visible radiation. As the area of interest
for temperature measurement in machining is located close to the tool tip,
it is necessary to enlarge the zone corresponding to the sensing area of the
CCD-sensors, which is 64 × 48 mm. To fulfill this requirement a special
fixed magnification lens (Navitar Precise Eye) with a working distance of
175 mm was used. This type of magnification lens, generally optimized
for visible radiation (80–85% in radiation transmission), can also be used
in the near infrared range, however with a slight loss in transmission
(75–80%).

Calibration of the CCD System


To calibrate the CCD system, two different methods may be applied:

• the calibration in front of a black body for which absolute temperature


is known,
• the calibration directly against a reference source made of the same
material (example, cutting insert) as the one that is supposed to be
observed during the turning operation.

In the black body calibration method (BBM), the real temperature of the
object, i.e., cutting insert, is related to the one given for a black body
derived from Planck’s law, as shown in Equation (1)

1 1 k
= + ln (Tr , ) (1)
Tr Tbb hc

where Tbb is the black body temperature, h = 66255 × 10−34 J.s, k =


13805 × 10−23 J/K, c = 29979 × 108 m/s,  is the emissivity, and  is the
wavelength corresponding to the NIR filter (850 nm).
This relation is directly applicable in the case of a monochromatic
radiation. Owing to the low emissivity variation of the cutting tools in
the near-IR region investigated, Equation (1) could be employed as a
first approximation (M’Saoubi et al., 2002). Furthermore, it shows that
the knowledge of the real temperature requires the determination of the
object’s emissivity.
In the present study, the emissivity values at different temperature were
determined for temperatures up to 1000◦ C, on the inserts. The measured
emissivities are presented in Figure 2. It shows that for temperatures
between 500−1000◦ C, the variation in emissivity values is very small.
Therefore it was decided to fix this value to  = 0386 (±0006748) when
calculating the insert temperature.
A calculation of the error propagated on temperature when emissivity
variation is neglected has been done and permits to determine the shift
492 A. Liljerehn et al.

FIGURE 2 Results from emissivity measurement for cutting insert TPNG160308.

between the real temperature and the one of a black body, as expressed in
Equation (2)

Tr − Tbb k
Error = = −Tr ln (T ) (2)
Tbb hc

Table 2 and Equation (2) indicate that the error in temperature is


quite sensitive to emissivity variations. The maximal error is ∼11% for
an emissivity of 0.2, below 8.5% for an emissivity higher than 0.3 and
lower than 5% for an emissivity value higher than 0.5. In earlier studies,
experimental emissivity measurements were carried out for different
cutting tools, such as tungsten carbide and TiN-coated tools. Results
showed experimental emissivity values to vary between 0.5 and 0.85. In the
present investigation the emissivity was found to be 0.38–0.4. Considering
the temperature range in this study (<500−1000◦ C), the maximal error
committed on the real temperature is less than 6.5% if the emissivity
variation is neglected.

TABLE 2 Error (%) in Temperatures Calculated from IR-CCD Measurements When Emissivity
Variation is Neglected

T (◦ C)\ 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

900 160 112 8.3 6.3 4.8 3.5 2.5 1.5 0.7
800 146 102 7.6 5.8 4.4 3.2 2.3 1.4 0.7
700 132 93 6.9 5.3 4.0 2.9 2.1 1.3 0.6
600 119 83 6.2 4.7 3.6 2.6 1.8 1.2 0.5
500 105 73 5.5 4.2 3.2 2.3 1.6 1.0 0.5
400 92 64 4.8 3.6 2.8 2.0 1.4 0.9 0.4
300 78 54 4.1 3.1 2.3 1.7 1.2 0.8 0.4
Modeling Heat Generation in Metal Machining 493

The calibration of the CCD with a reference cutting insert heated at


different temperatures was done in the chamber of a special furnace at
SIMR. The temperature is controlled by a thermocouple introduced in the
furnace chamber that is filled with Argon gas in order to minimize the
oxidation effect due to high temperatures. A sapphire window embedded
in the chamber enables the observation of the insert. With this method no
correction of temperature with emissivity is needed and the temperature
of the insert is assimilated to the temperature of furnace.

COMPARISON OF TWO METHODS

Based on the two methods described in the previous sections a set of


calibration curves displaying either the temperature of the black body or
the temperature of the furnace as a function of grey level were obtained
(Figures 3 and 4). Each calibration curve (grey level, n = f (T0 )) was
fitted with a mathematical function (sigmoidal with 4 parameters) of the
following form:
a
N = n0 + (T −T0 )
(3)
1+e b

Once the parameters n0 , a, b and T0 were obtained, this relation was


inverted to obtain a function describing the temperature as a function of
grey level, namely:
 
a
T = T0 + b ln −1 (4)
N − n0

FIGURE 3 Black body calibration curves (IR filter 850 nm).


494 A. Liljerehn et al.

FIGURE 4 SIMR furnace calibration curves (IR filter 850 nm).

TABLE 3 Parameters Determined for the IR-CCD Calibration


Curve at t = 20 ms

Method a b T0 n0 Corr.

Black body 334 63 767 −57 0.99


Furnace 396 75 815 −187 0.99

FIGURE 5 Comparison between black-body and SIMR furnace methods (for t = 20 < ms).
Modeling Heat Generation in Metal Machining 495

As an example Table 3 shows the parameters determined for the


calibration curve obtained within an integration time of 20 ms. Based on
the emissivity results and black calibration it was possible to calculate the
temperature of the insert using Equation (2) and compare it with the
insert temperature in the SIMR furnace. Results are shown in Figure 5,
displaying a comparison between black body and SIMR calibration curve
for t = 20 ms. It indicates a maximal deviation of 40◦ C between T(insert)
and T(insert via black body) . Possible reasons that could account for this variation
could be an underestimation of the insert surface temperature in the SIMR
furnace. The reading of the gray scale is problematic since both the insert
and the thermocouple are placed in the furnace and the thermocouple
temperature has to be read through a small window. Another reason could
be due to an underestimation of the insert emissivity. Typical treatment to

FIGURE 6 Typical treatment to obtain the tool isotherms (M’Saoubi et al., 2002).
496 A. Liljerehn et al.

obtain tool temperature distribution in a tungsten carbide cutting insert is


shown in Figure 6.

ANALYTICAL MODELLING OF CUTTING TEMPERATURE

The analytical solution handles heat generation due to the combined


effects of shear plane heat source and the tool-chip interface frictional
heat source and is based on the analytical model suggested by Komanduri
and Hou (2001a). The analytical model has however been corrected
to follow the stated boundary conditions regarding the location of the
imaged secondary heat source. The correction enables the solution to
be applied for cutting tools with a rake and relief ratio different from
90◦ . This change is important since it makes a significant difference in
tool temperature for tools with large rake and/or relief angles, which is
common in cutting tool design.

TEMPERATURE RISE IN THE CHIP FROM THE PRIMARY


HEAT SOURCE

The primary heat source, in the solution proposed by Komanduri


and Hou (2001a), is viewed as an oblique band heat source moving
in the direction of the cut in an infinite medium. This solution had
already been derived by Hahn (1951), but Komanduri and Hou (2001a)
adopted Hahn’s solution and made two important adjustments. The first
contribution was a modification of the model to yield for a semi-infinite
medium. This was accomplished by adding an adiabatic boundary on the
top of the chip surface and splitting the solution for the temperature rise
due to the primary heat source in the chip and the rest of the workpiece.
The image heat source is a reflection of the heat source in the adiabatic
plane. Adopting this method for calculating the temperature rise in the
chip due to the primary heat source (Komanduri and Hou, 2001a) and
the secondary heat source (Komanduri and Hou, 2001b) came up with the
two schematic solutions presented in Figures 7 and 8.
The temperature rise at any point in the chip, including the tool-
chip interface, due to the primary heat source, relative to the coordinate
system shown in Figure 9 and considering the previously stated boundary
conditions, can be calculated with Equation (5):
 AB     
qpls −(x−(lc −li sin()))vch /2a vch vch 
c−s = e K0 Ri + K 0 R dli (5)
2wp li =0 2a 2a i
Modeling Heat Generation in Metal Machining 497

FIGURE 7 Oblique band heat source with infinite depth, traveling in a semi-infinite solid; after
Komanduri and Hou (2001a).

The distance, Ri , between the point in the chip, M (x, z), and the finite
increment of the primary heat source, dli , is given by Equation (6):

Ri = (x − (lc − li sin()))2 + (z − li cos())2 (6)

and the distance, Ri , between the point in the chip, M (x, z), and the finite
increment of the imaged primary heat source, dli , is given by Equation (7):

Ri = (x − (lc − li sin()))2 (2tch − z − li cos())2 (7)

The heat source is moving with the velocity of the chip, vch , and having an
oblique angle  , that is given by Equation (8):

= −
(8)

FIGURE 8 Komanduri and Hou’s (2001b) proposed model for the temperature distribution due
to the secondary heat source in a continuous chip formation process in metal cutting.
498 A. Liljerehn et al.

FIGURE 9 Schematic description of Komanduri and Hou’s (2001c) model for calculating the
temperature rise in any point in a continuous chip due to both the primary and secondary heat
sources.

The coordinate system is positioned with the x-axis aligned with, and
the z-axis normal to, the tool rake face. Komanduri and Hou (2001c)
positioned the coordinate system for both the primary and the frictional
heat source to have the same position and orientation. The length of the
shear plane AB, is defined by Equation (9):
tch
AB = (9)
cos()
The thermal diffusivity, a, is a function of thermal conductivity, , specific
heat, cp , and density, , as shown in Equation (10):


a= (10)
cp ·

TEMPERATURE RISE IN THE CHIP FROM THE SECONDARY


HEAT SOURCE
The secondary deformation zone is referred to as the secondary heat
source and is defined by the contact area between the sliding chip and the
Modeling Heat Generation in Metal Machining 499

insert. The mathematical solution for this type of moving heat source has
been a subject of research for Carlsaw and Jaeger (1959), Blok (1938) and
Komanduri and Hou (2001b). Komanduri and Hou (2001b) started with
Jaeger’s solution for a source with a defined length and an infinite width,
moving along the length of the source in an infinite medium. The same
imaged heat source approach that was used for the solution of the primary
heat source was used to derive the solution of the frictional heat source
(Figure 8). The heat liberation over the contact length is regarded as
uniform, which is the same assumption that Komanduri and Hou (2001b)
as well as Hahn (1951), Trigger and Chao (1951), and Li and Liang (2005)
have made in their calculation models. The solution for the temperature
rise at any point in the chip due to the frictional heat source is presented
in Equation (11):
 lc     
qpl −(x−xj )vch /2a vch vch 
c−f = e K0 Rj + K 0 R dxj (11)
2wp xj =0 2a 2a j

where the distance, Rj , between the point in the chip, M (x, z), and
the finite increment of the secondary heat source, dxj , is given by
Equation (12):

Rj = (x − xj )2 + (z)2 (12)

and the distance, Rj , between the point in the chip, M (x, z), and the
finite increment of the imaged secondary heat source, dxj , is given by

FIGURE 10 Typical solution of the temperature rise in the chip due to the primary and secondary
heat sources.
500 A. Liljerehn et al.

Equation (13):

Rj = (x − xj )2 + (2tch − z)2 (13)

A typical output of the combined solution of Equations (5) and (11) is


presented in Figure 10, where it can be noted that only the temperature
rise in the chip is used in the final solution.

TEMPERATURE RISE IN THE TOOL FROM THE SECONDARY


HEAT SOURCE
The secondary heat source is regarded as a stationary rectangular heat
source for the temperature rise solution in the tool, in accordance with
the solution presented by Komanduri and Hou (2001b). The temperature
rise in any given point in the tool, by using the same imaged heat source
approach, can be calculated using Equation (14):
 lc  b/2  
qpl 1 1
t −f = +  dlk dyk (14)
2tool li =0 yi =−b/2 Rk Rk

The distance, Rk , between the point in the tool, M (x, y, z), and the finite
increment of the secondary heat source, dlk , is given by Equation (15):

Rk = (x − lk )2 + (y − yk )2 + (z)2 (15)

The calculation of Rk has been re-derived since the equation for this
distance presented by Komanduri and Hou (2001b) did not use the

FIGURE 11 Schematic comparison between Komanduri and Hou’s (2001c) model (on the left)
and the re-derived model (on the right) for calculating the temperature rise in any point in the
tool due to rectangular heat source at the tool-chip interface.
Modeling Heat Generation in Metal Machining 501

FIGURE 12 Typical solution of the temperature rise in the tool due to the secondary heat source.

correct mirror plane. The mirror plane in which the image heat source is
reflected is located at the tool clearance (Figure 11). This is mentioned
in Komanduri and Hou (2001b) but the equation has unfortunately not
been derived from this standpoint. The re-derived equation is presented
in Equation (16):

Rk = (lc (1 + cos(
+ )) − x − xk )2 + (y − yk )2 + (lc sin(
+ ) − z − zk )2
(16)

where

xk = lk cos(
+ ) (17)

and

zk = lk sin(
+ ) (18)

Figure 12 shows a typical solution of Equation (14), and it can be noted


that there is only a part of this solution that moves on to the final solution,
just like in the solution of the chip temperature.

TEMPERATURE RISE IN THE WORKPIECE FROM THE PRIMARY


HEAT SOURCE

The temperature rise at any point in the workpiece due to the primary
heat source relative to the coordinate system in Figure 13 can be calculated
502 A. Liljerehn et al.

FIGURE 13 Schematic description of the analytical model for calculating the temperature rise at
any point in the workpiece due to the primary heat source.

using Equation (19), and a typical solution is shown in Figure 14.


 AB     
qpls −(x−ll sin())vc /2a vc vc 
wp−s = e K0 Rl + K0 R dll (19)
2wp ll =0 2a 2a l

FIGURE 14 Typical solution of the temperature rise in the workpiece due to the primary heat
source.
Modeling Heat Generation in Metal Machining 503

FIGURE 15 Combination of the valid chip, tool and workpiece temperature solutions.

The finite increment of the primary heat source, dll , is given by


Equation (20):

Rl = (x − ll sin())2 + (z − ll cos())2 (20)

and the distance, Ri  , between the point in the chip, M (x, z), and the
finite increment of the imaged primary heat source, dli , is given by
Equation (21):

Rl = (x − ll sin())2 + (z + ll cos())2 (21)

There is only a part of the solution of Equation (19) that moves on to


the final solution, just like in the previous examples. The final solution
(Figure 15) of the temperature distribution in the chip, tool and work
pieces is then assembled from the valid parts of the solutions obtained with
Equations (5), (11), (14) and (19).

SOLUTION OF THE TEMPERATURE DISTRIBUTION


To be able to calculate the temperature rise and distribution at any
given point in the chip and tool it is essential to know the fraction of
energy that goes into the tool or follows the chip. This problem is solved
by calculating a heat partition ratio, B(x), which satisfies temperature
equilibrium along the tool-chip interface, as shown in Equation (22):

c−s + c−f B(x) = t −f (1 − B(x)) + t −i Btool _ind (x) (22)


504 A. Liljerehn et al.

where t −i and Btool _ind (x) is introduced to match the heat on the rake
face originating from the shear plane. A numerical solution that only
satisfies the equilibrium at discrete points over the tool-chip interface
cannot be used to solve the temperature distribution in the chip or the
tool. The solution of the heat partition ratio B(x) needs to be described
by a continuous function of xj along the contact length. The suggested
function presented in Komanduri and Hou (2001c) was used for solving
the heat partition ratio B(x), t −i and Btool _ind (x). The fraction of heat
conducted into the workpiece material is not constant or linear. The power
law Equations (23) and (24) are used to solve the temperature equilibrium
between tool-chip interfaces.
 m  k
x x
B(x) = (Bchip − B) + 2 B + C B (23)
lc lc

The fraction of heat conducted into the tool is expressed with


Equation (24):
 m  k
x x
1 − B(x) = (Btool + B) − 2 B − C B (24)
lc lc

where Btool is defined using Equation (25):

Btool = (1 − Bchip ) (25)

The temperature rise in the tool due to the induced heat source is
expressed by Equation (26):
 lc  b/2  
qpli 1 1
t −i = +  dlk dyk (26)
2tool li =0 yi =−b/2 Rk Rk

and the heat partition ratio for the induced heat source is found in
Equation (27):
 m i  ki
x x
Btool _ind (x) = (Binduced + Bi ) − 2 Bi − Ci Bi (27)
lc lc

The coefficients in Equations (23), (24) and (27) need to be solved to


satisfy Equation (22).

INPUT PARAMETERS

The analytical temperature model has been evaluated for six different
experimental setups. The cutting conditions and output parameters from
Modeling Heat Generation in Metal Machining 505

TABLE 4 Cutting Conditions

Test 1 2 3 4 5 6

tc [mm] 0112 0112 0224 0112 0112 0224


tch [mm] 016 019 028 018 017 032
vch [m/s] 140 118 157 128 132 140
vs [m/s] 244 225 236 237 232 236
b [mm] 30 30 30 30 30 30
[deg] 350 305 387 326 334 350
lc [mm] 023 028 038 023 027 042

TABLE 5 Cutting Forces

Test 1 2 3 4 5 6

Fc [N] 747 773 1296 738 743 1247


Ff [N] 564 601 613 513 580 663
Ffr [N] 564 641 680 587 618 727
Fs [N] 288 332 629 255 273 641

TABLE 6 Thermal Material Properties Used in the Analytical Model

Test 1 2 3 4 5 6

Tcm [◦ C] 485 520 380 477 493 420


wp
cp [J/kg K] 720 745 658 685 740 646
wp [J/s m K] 216 221 202 379 376 387
tool [J/s m K] 74 74 74 74 74 74
awp [mm2 /s] 39e−6 39e−6 40e−6 65e−6 64e−6 78e−6
wp [kg/m3 ] 7626 7599 7690 7790 7788 7822

TABLE 7 Material Properties Used in the Finite Element Model (MSC. Marc)

SANMAC 316L AISI 1045 Cutting insert TiAlN

Rp02 [MPa] 240


E [GPa] 199 206 580 439
 03 03 022 018
[mm2 /s] 165e−6 16e−6 74e−6 74e−6
 [J/s m K] 14 38 120 24
[kg/m3 ] 7822 7675 14500 4900
cp [J/kg K] 445 675 220 551
506 A. Liljerehn et al.

the trials are shown in Table 4. Cutting forces were measured during the
tests and are presented in Table 5, while the thermal material properties
used in the analytical model are listed in Table 6 and the properties used
for the MSC. Marc FE simulations are provided in Table 7.

Loewen–Shaw Model
This model is also based on an idealization of the heat sources (in
primary and secondary deformation zones) that are both regarded as
planes (Loewen and Shaw, 1954). The tool is also assumed to be perfectly
sharp and with no chip breaking geometry, as in the previous model.
This model does not take into account deformation of the material in the
secondary deformation zone but considers all work done due to friction
at the interface. On the other hand, the model does include the heat
exchange of the area of contact when sliding on a conducting surface.

Work in the Primary Deformation Zone


The estimated mean temperature of the material passing through the
primary deformation zone according to the Loewen and Shaw (1954)
model becomes:
(1 − )Fs vs
s = (28)
cp tc bvc

where (1 − ) · Fs · vs is the heat per unit time per unit area that flows into
the workpiece and (1 − ) can be calculated by Equation (29):

1
1− = (29)
1 + 1328 · a·cos(
)
vc ·tc ·cos( −
)

Temperature in the Secondary Deformation Zone


The average temperature rise in the chip due to the secondary heat
source can be expressed according to Boothroyd and Knight (1989) by
Equation (30):

Ff vch
f = (30)
cp tc bvch

and the mean chip temperature can be found through Equation (31):

Tcm = T0 + s + f (31)
Modeling Heat Generation in Metal Machining 507

The thermal dependant material properties—thermal diffusivity, a,


thermal conductivity, , specific heat, cp , and density, —have been
selected based on an iterative solution of the average chip temperature.

NUMERICAL MODELING OF HEAT GENERATION

Finite element program MSC. Marc was used to predict cutting forces,
chip formation and heat generation within the cutting zones. This is
based on an implicit updated Lagrangian formulation using continuous
re-meshing technique to model the chip formation (Kalhori, 2001).
Jaumann rate formulation is employed for the rate formulation (Crisfield,
1997). The orthogonal cutting condition has been modeled using four-
node plane strain elements. The volumetric strain is under-integrated
in order to avoid locking caused through large, incompressible plastic
strains. To predict the chip forming process for workpiece of steel AISI
1045 and stainless steel SANMAC 316L, a thermal-elastic-plastic material
model with temperature dependent properties and variable hardening was
implemented in MSC. Marc. The yield stress is updated based on strain
rate and thermal softening of workpiece material, formulated using the
Equation (32):

˙ T ) = g (p ) · ()
(p , , ˙ · (T ) (32)

where,
 1/n
p
g ( ) = 0 1 +
p
p (33)
0
p
and 0 is the initial yield stress, p is the plastic strain and 0 is the
reference plastic strain. The strain rate contribution to yield stress is
considered using the Equation (34):
 1/m1

˙ = 1+
() , if ˙ ≤ ˙ t
˙ 0
 1/m2  (1/m1 −1/m2 ) (34)
˙ ˙ t
˙ = 1+
() 1+ , if ˙ > ˙ t
˙ 0 ˙ 0

where, ˙ is strain rate, ˙ 0 is reference plastic strain rate, ˙ t is strain rate


where the transition between low and high strain rate sensitivity occurs,
m1 is the low strain rate sensitivity coefficient and m2 is the high strain
rate sensitivity coefficient. Thermal softening is accounted for using a
polynomial function of order up to five. To predict the material failure
at the primary deformation zone a damage model based on hydrostatic
508 A. Liljerehn et al.

pressure and strain to failure was implemented in MSC. Marc. The damage
value for each integration point is calculated using the Equation (35):


pi
D= p (35)
i
fi
p
where D is the dimensionless cumulative damage, i is the instantaneous
p
increment of strain, and fi is the instantaneous strain to failure. The
instantaneous strain to failure is determined considering the hydrostatic
pressure given by Equation (36):
 
p −15 · Pc · J p
f = n · exp  f0 (36)

where Pc is a pressure-dependent coefficient, defined by Equation (37):


 p
1 c
Pc = √ ln p (37)
3 t
and
√ 
3 · Pc
n = exp (38)
2

RESULTS AND DISCUSSION


The predictions from numerical and analytical models of cutting
process are compared with experimental results. The cutting forces were
measured using a Kistler dynamometer in a manual lathe (Köping),
as described by Kalhori et al. (1997), and the IR-CCD temperature
measurements were made during cutting in a CNC lathe. These are
compared with results from numerical studies in Table 8. It is observed that
the agreement of simulated cutting forces compared to the experimental

TABLE 8 Comparison of Measured (Exp) and Predicted (FEM) Machining Forces

Test 1 2 3 4 5 6

Fc [N] Exp. 747 773 1247 738 743 1296


Fc [N] FEM 750 775 1150 800 750 1230
Ff [N] Exp. 564 601 654 513 580 614
Ff [N] FEM 550 450 400 510 450 450
Modeling Heat Generation in Metal Machining 509

TABLE 9 Comparison of Measured (Exp) and Predicted (FEM) Chip Morphology Parameters

Test 1 2 3 4 5 6

tch [mm] Exp. 016 019 028 018 017 032


tch [mm] FEM 016 018 034 018 018 032
[deg] Exp. 350 305 387 320 334 350
[deg] FEM 350 327 333 320 363 350
lc [mm] Exp. 023 028 038 023 027 042
lc [mm] FEM 018 026 048 025 026 039

values is very good when it comes to the forces in the cutting direction.
However, this is not the case for feed forces, especially for TCMT16T308-
MM and TPNG160308 inserts. Several different phenomena, described
below, could be the reason for this deviation.
More accurate results may be achieved through further development
of the material model considering strain localization. The damage model
implemented in the current FEM simulation model should be further
developed in order to better capture shear localization at the primary
deformation zone. Furthermore, the friction model used in the numerical
analysis might be further improved. In the current study an average
friction coefficient for the insert-chip contact area is assumed. Since, the
friction behavior of contact surfaces depends on both the contact bodies’
temperatures and contact pressure; it should be included in the FE-model.
It is believed that this would improve the numerically predicted result.
The simulated chip morphologies are compared with experimental
results in Table 9. It is shown that there is very good agreement
between the simulation and measurements when it comes to the average
chip thickness, contact length and shear plane angle. The measured
peak temperature over the tool-chip interface for different geometries is
compared to the analytical and finite element simulations in Table 10.

TABLE 10 Comparison of Experimentally Measured (IR-CCD)


and Predicted (Analytical, FEM) Peak Temperatures at the
Tool-Chip Interface

Test Analytical [◦ C] IR-CCD [◦ C] FEM [◦ C]

1 802 850
2 890 852 850
3 1004 943 960
4 657 750
5 785 785 740
6 835 823 760
510 A. Liljerehn et al.

FIGURE 16 Comparison of analytically predicted, experimentally measured (IR-CCD), and


numerically simulated (MSC. Marc) temperatures for Test 2 (TCMT 16T308-MM tool; SANMAC
316L workpiece).

FIGURE 17 Comparison of analytically predicted, experimentally measured (IR-CCD), and


numerically simulated (MSC. Marc) temperatures for Test 3 (TPNG160308 tool; SANMAC 316L
workpiece).

FIGURE 18 Comparison of analytically predicted, experimentally measured (IR-CCD), and


numerically simulated (MSC. Marc) temperatures for Test 5 (TCMT 16T308-MM tool; AISI 1045
workpiece).
Modeling Heat Generation in Metal Machining 511

FIGURE 19 Comparison of analytically predicted, experimentally measured (IR-CCD), and


numerically simulated (MSC. Marc) temperatures for Test 6 (TPNG160308 tool; AISI 1045
workpiece).

The temperature gradients in the cutting zones for all tests, except for the
TCMT 16T308-MR insert where there are no IR-CCD results, are shown in
Figures 16–19.
The analytical model of heat generation has a tendency to
underestimate the temperature near the tool tip and overestimate the
tool peak temperature further up along the rake face, which has its
natural explanation from the neglected friction forces around the cutting
edge radius. The mean temperature is also lower along the tool-chip
interface than the IR-CCD measured temperature but the model is most
definitely suitable for over-all temperature estimations. The reason for the
underestimation, by the analytical model, near the tool tip might have a
four folded explanation:

A) Since the cutting forces along the insert-chip interface are not evenly
distributed, the assumption that the heat liberation over the rake face
would be evenly distributed may not be correct.
B) The contact between the tool and the chip needs to be modeled with
a stick-slip approach that considers the shear stress variation over the
contact length properly. This would result in higher heat liberation
intensity closer to the tool tip and a lower intensity further up where
the contact state changes from sticking to slipping. This approach has
been proven by both Huang and Liang (2003) as well as Karpat and
Özel (2006a,b) and would also result in a higher mean temperature
but a lower peak temperature.
C) In the current studies the temperature rise from friction between the
tool and the workpiece in the tertiary deformation zone is not taken
to account.
D) The heat liberation intensity in the primary deformation zone is
determined based on the calculated average plastic deformation over
512 A. Liljerehn et al.

the shear plane. This is not the case during the cutting process.
The primary deformation zone is known to involve more plastic
deformation near the cutting edge radius then on the back face of
the chip. This will further increase the temperature rise close to the
edge radius and increase the mean temperature over the contact
area.

The results from the analytical models are based on information


from measurements; i.e., cutting forces, contact length and thickness of
deformation that is caused by friction in the contact with the insert.
Contact length is hard to predict or measure but a very important
parameter since the results from the analytical model are very sensitive
to the contact area over which the heat liberation is calculated. A small
difference in contact length can result in a large difference in predicted
temperature.

CONCLUSIONS

From the result comparison between measured cutting forces and the
corresponding values predicted with the FEM simulation model it can
be concluded that the forces in the cutting direction match well in all
experiments. The prediction of the feed force on the other hand needs
to be further developed, preferably with a material model that considers
strain localization better as well as an improved friction model. The
damage model implemented in the current FEM simulation model would
also result in a better agreement regarding the feed force if it were further
developed in order to better capture the shear localization at the primary
deformation zone. It can also be concluded that the FEM model made
good temperature predictions for the SANMAC 316L workpiece material
but slightly underestimated the temperature in the AISI 1045 case, which
might have to do with the underestimation of the feed force in these
simulations.
The analytical model of heat generation made generally good
temperature predictions for the given experiments but the model is
very sensitive to input parameters. The problem with underestimation
of the temperature near the tool tip, and overestimation of the tool
peak temperature further up along the rake face, could to some extent
be worked around by considering the rubbing heat source on the
clearance side of the tool. This approach demands more in-depth analysis
of the cutting force distribution around the cutting edge of the tool.
The information that is obtained from the experimental setup in this
experiment is not enough but might be of interest in future work.
Modeling Heat Generation in Metal Machining 513

NOMENCLATURE
1−B Fraction of heat conducted into the tool
a Thermal diffusivity [mm2 /s]
AB Shear plane length [mm]
B Fraction of heat conducted into the chip
b Depth of cut, ap [mm]
cp Specific heat capacity [J/kg K]
D Dimensionless cumulative damage
E Young’s modulus [GPa]
Fc Cutting force [N]
Ff Feed force [N]
Ffr Frictional force [N]
Fs Shear force [N]
J Hydrostatic pressure [MPa]
K0 Bessel function of second kind order zero
lc Contact length [mm]
m1 low strain rate sensitivity coefficient
m2 high strain rate sensitivity coefficient
Pc Pressure coefficient
qpl Heat liberation intensity of the frictional heat source
[J/mm2 s]
qpls Heat liberation intensity of the moving shear plane
heat source [J/mm2 s]
R Distance between the moving-line heat source and point M ,
where the temperature rise is calculated [mm]
T0 Ambient temperature of workpiece [◦ C]
Tcm Average chip temperature [◦ C]
tc Uncut chip thickness, feed [mm]
tch Cut chip thickness [mm]
vc Cutting speed [m/s]
vs Shear velocity [m/s]
x, y, z Coordinates of point M where the temperature is
calculated [mm]
1− Fraction of energy conducted into the chip
Clearance angle [◦ C]
p
˙ i Instantaneous increment of strain
 Ratio of chip thickness deformed in secondary
deformation zone
p Plastic strain [mm/mm]
p
0 Reference plastic strain [mm/mm]
p
fi Instantaneous strain to failure [mm/mm]
˙ Strain rate [1/s]
˙ 0 Reference plastic strain rate [1/s]
514 A. Liljerehn et al.

˙ t Strain rate where the transition between low


and high strain rate sensitivity occurs [1/s]
 Temperature rise [◦ C]
tool Thermal conductivity of tool [J/s m K]
wp Thermal conductivity of workpiece [J/s m K]
Density [kg/m3 ]
 Initial yield stress [MPa]
¯ Von Mises stress [MPa]
 Poisson’s ratio
Shear angle [◦ C]
 Oblique angle [◦ C]

REFERENCES
Blok, H. (1938) Theoretical study of temperature rise at surfaces of actual contact under oiliness
lubricating conditions. Proceedings of General Discussion on Lubrication and Lubricants,
Institute of Mechanical Engineers London, pp. 222–235.
Boothroyd, G.; Knight, W.A. (1989) Fundamentals of Machining and Machine Tool, 2nd edition, Marcel
Dekker Inc., New York.
Carlsaw, H.S.; Jaeger, J.C. (1959) Conduction of Heat in Solids, 2nd edition, Oxford University Press,
Oxford, UK.
Crisfield, M.A. (1997) Non-linear Finite Element Analysis of Solids and Structures, Vol. 2, Advanced
Topics. John Wiley, Chichester.
Hahn, R.S. (1951) On the temperature developed at the shear plane in the metal cutting process.
Proceedings of the First U.S. National Congress of Applied Mechanics, pp. 661–666.
Huang, Y.; Liang, S.Y. (2003) Cutting temperature modeling based on non-uniform heat intensity
and partition ratio. Machining Science and Technology, 9(3): 301–323.
Kalhori, V. (2001) Modelling and Simulation of Mechanical Cutting, Luleå Technical University, Doctoral
thesis: 28, ISSN: 1402–1544.
Kalhori, V.; Lundblad, M.; Lindgren, L.-E. (1997) Numerical and experimental analysis of
orthogonal metal cutting. Transactions of the ASME, International Mechanical Engineering Congress
& Exposition, MED Vol. 6–2, Manufacturing Science and Engineering, Dallas, Texas 16–21
Nov. 1997.
Karpat, Y.; Özel, T. (2006a) Predictive analytical and thermal modeling of orthogonal cutting
process. Part I: Predictions of tool forces, stresses and temperature distributions. ASME Journal
of Manufacturing Science and Engineering, 128(2): 435–444.
Karpat, Y.; Özel, T. (2006b) Predictive analytical and thermal modeling of orthogonal cutting
process. Part II: Effect of tool flank wear on tool forces, stresses and temperature distributions.
ASME Journal of Manufacturing Science and Engineering, 128(2): 445–453.
Komanduri, R.; Hou, Z.B. (2001a) Thermal modeling of the metal cutting process, Part 1:
Temperature rise distribution due to shear plane heat source. International Journal of Mechanical
Science, 42: 1715–1752.
Komanduri, R.; Hou, Z.B. (2001b) Thermal modeling of the metal cutting process, Part 2:
Temperature rise distribution due to frictional heat source at the tool-chip interface.
International Journal of Mechanical Science, 43: 57–88.
Komanduri, R.; Hou, Z.B. (2001c) Thermal modeling of the metal cutting process, Part 3:
Temperature rise distribution due to combined effects of shear plane heat source and
the tool-chip interface frictional heat source. International Journal of Mechanical Science,
43: 89–107.
Li, K.M.; Liang, S.Y. (2005) Modeling of cutting temperature in near dry machining. Journal of
Manufacturing Science and Engineering, Transactions of the ASME, 128(2): 1–9.
Modeling Heat Generation in Metal Machining 515

Loewen, E.G.; Shaw, M.C. (1954) On the analysis of cutting tool temperatures. Transactions of the
ASME, 71: 217–231.
M’Saoubi, R.; Eggertsson, C.; Chandrasekaran, H. (2002) Application of IR-CCD technique to
map tool temperature distribution in single point turning of quenched and tempered steel.
IM Report No IM-2002-563, Swedish Institute for Metals Research.
Trigger, K.J.; Chao, B.T. (1951) An analytical evaluation of metal cutting temperatures. Transactions
of the ASME, 73: 57–68.
Copyright of Machining Science & Technology is the property of Taylor & Francis Ltd and its content may not
be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written
permission. However, users may print, download, or email articles for individual use.

Das könnte Ihnen auch gefallen